Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Circular Bioeconomy Perspectives in Sustainable Bioenergy Production

Download as pdf or txt
Download as pdf or txt
You are on page 1of 469

Energy, Environment, and Sustainability

Series Editor: Avinash Kumar Agarwal

Gurunathan Baskar
Veeramuthu Ashokkumar
Samuel Lalthazuala Rokhum
Vijayanand Suryakant Moholkar Editors

Circular
Bioeconomy
Perspectives
in Sustainable
Bioenergy
Production
Energy, Environment, and Sustainability

Series Editor
Avinash Kumar Agarwal, Department of Mechanical Engineering, Indian Institute
of Technology Kanpur, Kanpur, Uttar Pradesh, India
AIMS AND SCOPE
This books series publishes cutting edge monographs and professional books focused
on all aspects of energy and environmental sustainability, especially as it relates to
energy concerns. The Series is published in partnership with the International Society
for Energy, Environment, and Sustainability. The books in these series are edited
or authored by top researchers and professional across the globe. The series aims
at publishing state-of-the-art research and development in areas including, but not
limited to:
• Renewable Energy
• Alternative Fuels
• Engines and Locomotives
• Combustion and Propulsion
• Fossil Fuels
• Carbon Capture
• Control and Automation for Energy
• Environmental Pollution
• Waste Management
• Transportation Sustainability

Review Process
The proposal for each volume is reviewed by the main editor and/or the advisory
board. The chapters in each volume are individually reviewed single blind by expert
reviewers (at least four reviews per chapter) and the main editor.
Ethics Statement for this series can be found in the Springer standard guidelines
here https://www.springer.com/us/authors-editors/journal-author/journal-author-hel
pdesk/before-you-start/before-you-start/1330#c14214
Gurunathan Baskar · Veeramuthu Ashokkumar ·
Samuel Lalthazuala Rokhum ·
Vijayanand Suryakant Moholkar
Editors

Circular Bioeconomy
Perspectives in Sustainable
Bioenergy Production
Editors
Gurunathan Baskar Veeramuthu Ashokkumar
Department of Biotechnology Saveetha Institute of Medical and Technical
St. Joseph’s College of Engineering Sciences (SIMATS)
Chennai, Tamil Nadu, India Saveetha University
Chennai, Tamil Nadu, India
Samuel Lalthazuala Rokhum
Department of Chemistry Vijayanand Suryakant Moholkar
National Institute of Technology Silchar Department of Chemical Engineering
Silchar, Assam, India Indian Institute of Technology Guwahati
Guwahati, Assam, India

ISSN 2522-8366 ISSN 2522-8374 (electronic)


Energy, Environment, and Sustainability
ISBN 978-981-97-2522-9 ISBN 978-981-97-2523-6 (eBook)
https://doi.org/10.1007/978-981-97-2523-6

© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature
Singapore Pte Ltd. 2024

This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether
the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse
of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and
transmission or information storage and retrieval, electronic adaptation, computer software, or by similar
or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.

This Springer imprint is published by the registered company Springer Nature Singapore Pte Ltd.
The registered company address is: 152 Beach Road, #21-01/04 Gateway East, Singapore 189721,
Singapore

If disposing of this product, please recycle the paper.


Preface

The developing economies in the world, throughout their progress and diversification
in the last one century, have followed a fundamental feature established in the nine-
teenth century during industrial revolution: the linear model of using the resources
viz. the strategy of “take-make-dispose”. This model followed the pattern of obtaining
raw materials from nature, converting them into useful products (through physical/
chemical/biological processes) to be used by the consumers and finally discarding
them as waste back into the nature. This one-way flow of materials or resources
produced high adverse long-term environmental and social impacts. With fast growth
of industrial sector, coupled with rising population and modernisation of infrastruc-
ture, there is an urgent need to pursue and realise the Sustainable Development Goals
(derived from 2015 United Nations agreement) that aim to decouple the resource util-
isation (or material input) from revenue (or economic growth). A solution to meet this
goal is transition from “take-make-dispose” model to “reduce-repair-reuse-recycle”
model—which forms the basis of circular economy.
The International Society for Energy, Environment and Sustainability (ISEES)
was founded at the Indian Institute of Technology Kanpur (IIT Kanpur), India, in
January 2014 to spread knowledge and awareness and to catalyse research activities in
the fields of energy, environment, sustainability, and combustion. Society’s goal is to
contribute to the development of clean, affordable, and secure energy resources and a
sustainable environment for society, spread knowledge in the areas mentioned above,
and create awareness about the environmental challenges the world is facing today.
ISEES aimed to break the conventional silos of specialisations (engineering, science,
environment, agriculture, biotechnology, materials, fuels, etc.) to tackle the problems
related to energy, environment, and sustainability holistically. This is evident by the
participation of experts from all fields to resolve these issues. The ISEES is involved
in various activities, such as conducting workshops, seminars, conferences, etc., in
the domains of its interests. The society also recognises the outstanding works of
young scientists, professionals, and engineers for their contributions by conferring
awards in various categories.
The Seventh International Conference on “Sustainable Energy and Environmental
Challenges” (VII-SEEC) was organised under the auspices of ISEES from 16–18

v
vi Preface

December 2022 at the Indian Institute of Technology, Banaras Hindu University,


Varanasi (IIT BHU), India. This conference provided a platform for discussions
between eminent scientists and engineers from several countries, including India,
the USA, Spain, Poland, Austria, the Czech Republic, and Korea. At this confer-
ence, eminent international speakers presented their views on energy, combustion,
emissions, and alternative energy resources for sustainable development and a cleaner
environment. The conference presented two high-voltage plenary talks by Dr. Ajay
Kumar, Former Union Defense Secretary, Government of India and Dr. S. S. V.
Ramakumar, Director (R&D), Indian Oil.
The conference included 29 technical sessions on energy and environmental
sustainability topics, including two plenary talks, 15 keynote talks, 80 invited talks
from prominent scientists, and 118 contributed talks by students and researchers. The
conference included technical sessions on advanced engine technologies, air pollu-
tion monitoring, anaerobic digestion, combustion and flames, air pollution control,
biodegradation of toxic chemicals, energy and exergy, desalination and wastew-
ater treatment, environmental bioengineering, pollution and climate change: chal-
lenges and priorities, alternative transportation fuels and materials, emerging envi-
ronmental contaminants, sustainable processing of biomass, human health and envi-
ronmental sustainability, sprays and atomisation, solid waste: challenges and mitiga-
tion, solid waste management, sustainable food and agri-biotechnology, modelling
and simulations, renewable energy technologies, bioremediation, biofuels and biore-
fineries, engine emissions and control, cleaner technologies for pollution mitigation,
microbial processes, energy and environment, coal and biomass gasification, and
environmental challenges mitigation.
About 250+ participants and speakers attended the conference, where 14 ISEES
books published by Springer, Singapore, under a dedicated series, Energy, Envi-
ronment and Sustainability, were released. This conference laid the roadmap for
technology development, opportunities, and challenges in the energy, environment,
and sustainability domains. These topics are relevant for the country and the world
in the present context. We acknowledge the support from various funding agen-
cies and organisations for the successful conduct of the VII-SEEC, where the idea
of these books germinated. We would therefore gratefully like to acknowledge IIT
BHU (Special thanks to Prof. Akhilendra Pratap Singh); SERB, Government of India;
Department of Scientific and Industrial Research (DSIR) (Special thanks to Dr. Vipin
Shukla) and our publishing partner Springer (Special thanks to Swati Mehershi).
The editors would like to express their sincere gratitude to a large number of
authors from all over the world for submitting their high-quality work on time and
revising it appropriately at short notice. We want to express our special gratitude
to our prolific reviewers: Dr. Lepakshi Barbora, Dr. Pankaj Kalita, Dr. Debarshi
Mallick, Dr. Ritesh Malani, Dr. Rakesh Kumar Sharma, Dr. Anita R. Warrier, Dr.
Venkatesa Prabhu Sundramurthy, Dr. Sivaprakash Sundaresan, Dr. Zainul Akmar
Zakaria, Dr. P. Radha, Dr. Selvakumar P., Dr. M. Anil Kumar, Dr. M. Jayakumar,
Dr. Vinoth Kumar V., Dr. R. Rajesh Kannan, Dr. Vidhya C. V., Dr. S. Raja, Dr. J.
Iyyappan, Dr. Shijie Liu, Dr. Anh Tuan Hoang, Dr. Irfan Turhan, Dr. Josiel Martins
Costa, Dr. Rodrigo S. Vieira, Dr. Steven Lim, Dr. Shangeetha Ganesan, Dr. Mobolaji
Preface vii

Shemfe, Dr. Chuan Chen, Dr. Ramesh Kumar, Dr. Debora Cynamon Kligerman, Dr.
Jun Yang, Dr. Bahram Barati, Dr. Bhupendra Singh Chauhan, and Dr. Zibin Yin,
who reviewed chapters of this monograph and provided their valuable suggestions
to improve them.
This monograph has addressed the subject of circular bioeconomy for bioen-
ergy production, which application of circular economy principles for sustainable
energy production from renewable or bio-resources. Eighteen chapters in this book
have touched upon multiple and diverse aspects of production of renewable and
sustainable energy. These aspects include: utilisation of lignocellulosic biomass, use
of industrial organic waste and byproducts, resources from domestic sector, such
as waste cooking oil for biodiesel production, use of hospital-based plastic waste,
bioethanol production from biomass feedstocks, use of urban solid waste through
biorefinery, thermochemical (pyrolytic) conversion of waste biomass, bio-jet fuel
generation from waste resources, and supply chain analysis of waste biomass for
bioenergy production. The book also contains case studies on biogas generation from
organic wastes. Chapters include recent results and are focussed on current trends
in sustainable bioenergy production from waste resources and the related technical
issues. We expect that this book will serve as a source of state-of-the-art informa-
tion and knowledge about the sustainable energy production—in the form of both
gaseous and liquid biofuels—from waste resources. The case studies presented in the
book are likely to be practical examples and first-hand information for entrepreneurs
that demonstrate feasibility and viability of sustainable bioenergy production. We
hope that this book will be a handy resource for academicians, researchers, indus-
trial professionals, and postgraduate students working in the area of bioenergy and
sustainable development.

Chennai, Tamil Nadu, India Gurunathan Baskar


Chennai, Tamil Nadu, India Veeramuthu Ashokkumar
Silchar, Assam, India Samuel Lalthazuala Rokhum
Guwahati, Assam, India Vijayanand Suryakant Moholkar
Contents

1 Introduction to Circular Bioeconomy Perspectives


in Sustainable Bioenergy Production . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
Gurunathan Baskar, Veeramuthu Ashokkumar,
Samuel Lalthazuala Rokhum, and Vijayanand Suryakant Moholkar
2 Bioenergy-Based Sustainable Bioeconomy—Perspectives
and Challenges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
C. V. Vidhya, S. Nandhini, J. Mary Sheela, and M. Reenaa
3 Lignocellulosic Biomass for Sustainable Production
of Renewable Fuels: Embracing Natural Resources . . . . . . . . . . . . . . . 37
Medha Maitra, S. Sruthi, Pavada Madhusudan Rao, V. S. Avanthi,
and P. Radha
4 Industrial Organic Waste and Byproducts as Sustainable
Feedstock for Bioenergy Production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
Desta Getachew Gizaw, Selvakumar Periyasamy,
Zinnabu Tassew Redda, and Gurunathan Baskar
5 Transesterification of Waste Cooking Oil Through Microwave
Technology: Recent Advances and Challenges . . . . . . . . . . . . . . . . . . . 117
Bunushree Behera, Kolli Venkata Supraja, S. Mari Selvam,
Snehi Kinger, and Prangya Ranjan Rout
6 Green Catalysts Synthesized from Biomass for Biodiesel
Production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
Amirthavalli Velmurugan, Anita R. Warrier, and Gurunathan Baskar
7 Sustainable Solutions for Bioenergy Production
from Hospital-Based Plastic Waste—Thinking Beyond
Landfills . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
Patitapaban Dash, Chirasmita Mohanty, Pratyush Kumar Das,
Joseph M Anto Simon, Debasish Sahoo, and Gurunathan Baskar

ix
x Contents

8 Functionalized Biochar for Green and Sustainable Production


of Biodiesel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193
Hlawncheu Zohmingliana, Joseph V. L. Ruatpuia,
and Samuel Lalthazuala Rokhum
9 Versatile Pretreatment Approaches to Improve the Bioethanol
Production from Various Biomass Feedstocks . . . . . . . . . . . . . . . . . . . . 219
R. Rajesh Kannan, V. Saravanan, M. Rajasimman,
Panchamoorthy Saravanan, and Gurunathan Baskar
10 Biorefinery Avenues for Processing Urban Solid Waste:
Potential for Value-Added Chemicals and Energy . . . . . . . . . . . . . . . . 239
Swapna Gade, Yuvraj Patil, and Bhalchandra Bhanage
11 Pyrolytic Conversion of Heterogenic Natural Waste Biomass
from Rural Communities with Concomitant Valorization . . . . . . . . . 259
M. Anil Kumar, Pareshkumar G. Moradeeya,
K. Manikanda Bharath, P. Jakulin Divya Mary, and K. S. Giridharan
12 Improving the Biogas Generation Potential from Organic
Wastes Using Hydrochar as an Additive Lab-Scale Case
Study from Central Sweden: Part 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 289
Maria Kristoffersson, Maria Sandberg, and G. Venkatesh
13 Improving the Biogas Generation Potential from Organic
Wastes Using Hydrochar as an Additive Lab-Scale Case
Study from Central Sweden: Part 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 307
Annette Liisa Kariis, Maria Sandberg, and G. Venkatesh
14 Heterogeneous Hydrochar-Based Catalysts for Biodiesel
Production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 323
Muhammad Aliyu, Umer Rashid,
Wan Azlina Wan Ab Karim Ghani,
Muhamad Amran bin Mohd Salleh, Balkis Hazmi,
Ibrahim Garba Shitu, and Ali Salisu
15 Circular Bioeconomy Approaches for Valorizing Waste
Streams into Bio-jet Fuel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 361
Louella Concepta Goveas, S. M. Vidya, Ramesh Vinayagam,
and Raja Selvaraj
16 Sustainable Bioethanol Production from the Pretreated Waste
Lignocellulosic Feedstocks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 377
Belete Tessema Asfaw, Meroda Tesfaye Gari, Mani Jayakumar,
and Gurunathan Baskar
Contents xi

17 Waste Biomass Supply Chain for Sustainable Bioenergy


Production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 395
C. Nirmala, M. Sridevi, P. Loganathan, Mani Jayakumar,
and Gurunathan Baskar
18 Manifesting Sustainability Toward Food Waste into Bioenergy:
Biorefinery in a Circular Economic Approach . . . . . . . . . . . . . . . . . . . 431
Devi Sri Rajendran, Swethaa Venkatraman, R. Rahul, M. Afrrin,
P. Karthik, and Vinoth Kumar Vaidyanathan
Editors and Contributors

About the Editors

Dr. Gurunathan Baskar has 24 years of teaching and


research experience, published more than 200 research
and review articles in reputed National and Interna-
tional Journals, 37 book chapters and edited 6 books
in various fields of Biotechnology. He is Fellow of
the Institution of Engineers (India) and International
Society for Energy Environment and Sustainability;
Active Life Member of various national and interna-
tional professional bodies; listed in top 2% Scientist in
the world consecutively in 2020, 2021 and 2022 by Else-
vier BV and Stanford University, USA; and received
ISTE-Periyar Best Engineering College Teacher Award
2020 from Indian Society for Technical Education-
Tamil Nadu Section, Prof. S. B. Chincholkar Memo-
rial Award 2019 from Biotech Research Society, India,
for the outstanding work in the area of Biofuels and
Food Biotechnology, ISTE-Syed Sajid Ali National
Award 2016 for his outstanding research on renew-
able energy from Indian Society for Technical Educa-
tion, New Delhi, and Young Scientist Award 2015 from
International Bioprocessing Association, France. He has
been actively researching for creating sustainable bioen-
ergy solutions while contributing toward a greener and
circular bioeconomy environment.

xiii
xiv Editors and Contributors

Dr. Veeramuthu Ashokkumar received his Ph.D. from


the University of Madras, Chennai, India, in 2014.
During his Ph.D. he was awarded an Indo-Australian
fellowship and visited Adelaide University, Australia, as
a visiting scientist. He received his M.Phil. and M.Sc. in
Plant Biotechnology from Indian universities. Currently,
he is working as a professor at Saveetha Institute of
Medical and Technical Sciences (SIMATS), Saveetha
University, Chennai, India. He was the visiting professor
at Chulalongkorn University, Thailand. He worked as
a postdoctoral scientist in Microalgal Biorefinery at
the Department of Mechanical Engineering, Univer-
sity Technology Malaysia (UTM), Malaysia. Also, he
worked as a postdoctoral scientist at the Department
of Aeronautics and Astronautics, National Cheng Kung
University, Taiwan. He was also a visiting scientist
and visited European Universities under the Marie
Skłodowska-Curie Actions Fellowship scheme

Dr. Samuel Lalthazuala Rokhum is Assistant


Professor in the Department of Chemistry, National
Institute of Technology Silchar, Silchar. His research
interests encompass synthetic chemistry, heterogeneous
catalysis, green chemistry, and biofuel synthesis. Dr.
Rokhum has completed two SERB-funded research
projects and currently running projects funded by CSIR
and DST. He has so far published 99 papers in jour-
nals of international repute and has authored and edited
two books. In addition, he has delivered a lecture at
several national and international conferences and has
been granted a German patent.

Dr. Vijayanand Suryakant Moholkar is Professor


(HAG) of Chemical Engineering and Adjunct Faculty
at School of Energy Science and Engineering (erstwhile
Centre for Energy) at Indian Institute of Technology
Guwahati. He has been Head of the Chemical Engi-
neering Department, IIT Guwahati between 2012 and
2015, and Head of Centre for Energy, IIT Guwahati
between 2017 and 2020. His main research interests are
sonochemistry, cavitation assisted physical, chemical,
and biological processing, and thermo and biochemical
routes to biofuels. As of January 2023, he has published
more than 200 papers in renowned international jour-
nals that have received more than 11000 citations (with
Editors and Contributors xv

h-index of 62 and i-10 index of 182). He is Co-inventor


of 3 US patents (issued to CTI Nanotech, CA, USA)
on application of hydrodynamic cavitation reactors for
biomass pretreatment and bioalcohol synthesis.

Contributors

M. Afrrin Department of Food Technology, Faculty of Engineering, Karpagam


Academy of Higher Education, Coimbatore, India
Muhammad Aliyu Institute of Nanoscience and Nanotechnology, Universiti
Putra Malaysia (UPM), Serdang, Selangor, Malaysia;
Department of Material Engineering, Graduate School of Engineering, Kyushu
Institute of Technology, Tobata-ku, Kitakyushu, Fukuoka, Japan
M. Anil Kumar Centre for Rural and Entrepreneurship Development, National
Institute of Technical Teachers Training and Research, Chennai, Tamil Nadu, India
Belete Tessema Asfaw Department of Chemical Engineering, Haramaya
University, Haramaya Institute of Technology, Haramaya, Ethiopia;
College of Biological and Chemical Engineering, Addis Ababa Science and
Technology University, Addis Ababa, Ethiopia
Veeramuthu Ashokkumar Saveetha Institute of Medical and Technical Sciences
(SIMATS), Saveetha University, Chennai, Tamil Nadu, India
V. S. Avanthi Bioprocess and Bioseparation Laboratory, Department of
Biotechnology, School of Bioengineering, College of Engineering and Technology,
SRM Institute of Science and Technology, Kattankulathur, Chengalpattu District,
Tamil Nadu, India
Gurunathan Baskar Department of Biotechnology, St. Joseph’s College of
Engineering, Chennai, Tamil Nadu, India;
School of Engineering, Lebanese American University, Byblos, Lebanon
Bunushree Behera Bioprocess Laboratory, Department of Biotechnology, Thapar
Institute of Engineering and Technology, Patiala, Punjab, India
Bhalchandra Bhanage Department of Chemistry, Institute of Chemical
Technology, Mumbai, India
Pratyush Kumar Das Department of Phytopharmaceuticals, School of
Agricultural Engineering and Bioprocessing (SoABE), Centurion University of
Technology and Management, Paralakhemundi, Odisha, India
Patitapaban Dash Centre for Biotechnology, Siksha ‘O’ Anusandhan (Deemed
to be University), Bhubaneswar, Odisha, India
xvi Editors and Contributors

Swapna Gade Chemical Engineering and Process Development Division,


National Chemical Laboratory, Pashan, Pune, India
Meroda Tesfaye Gari College of Biological and Chemical Engineering, Addis
Ababa Science and Technology University, Addis Ababa, Ethiopia
Wan Azlina Wan Ab Karim Ghani Department of Chemical and Environmental
Engineering, Faculty of Engineering, Universiti Putra Malaysia, Serdang, Selangor,
Malaysia
K. S. Giridharan Centre for Rural and Entrepreneurship Development, National
Institute of Technical Teachers Training and Research, Chennai, Tamil Nadu, India
Desta Getachew Gizaw Department of Chemical Engineering, School of
Mechanical, Chemical and Materials Engineering, Adama Science and Technology
University, Adama, Ethiopia
Louella Concepta Goveas Department of Biotechnology Engineering, NMAM
Institute of Technology, Nitte (Deemed to be University), Mangalore, Karnataka,
India
Balkis Hazmi Institute of Nanoscience and Nanotechnology, Universiti Putra
Malaysia (UPM), Serdang, Selangor, Malaysia
P. Jakulin Divya Mary Centre for Rural and Entrepreneurship Development,
National Institute of Technical Teachers Training and Research, Chennai, Tamil
Nadu, India
Mani Jayakumar Department of Chemical Engineering, Haramaya University,
Haramaya Institute of Technology, Haramaya, Dire Dawa, Ethiopia;
Department of Biotechnology, Faculty of Engineering, Karpagam Academy of
Higher Education, Coimbatore, Tamilnadu, India;
Centre for Natural Products and Functional Foods, Karpagam Academy of Higher
Education, Coimbatore, Tamilnadu, India
Annette Liisa Kariis Energy and Environmental Systems Group, Department of
Engineering and Chemical Sciences, Karlstad University, Karlstad, Sweden
P. Karthik Department of Food Technology, Faculty of Engineering, Karpagam
Academy of Higher Education, Coimbatore, India
Snehi Kinger Bioprocess Laboratory, Department of Biotechnology, Thapar
Institute of Engineering and Technology, Patiala, Punjab, India
Maria Kristoffersson Energy and Environmental Systems Group, Department of
Engineering and Chemical Sciences, Karlstad University, Karlstad, Sweden
P. Loganathan Department of Electrical and Electronics Engineering, Vinayaka
Missions Kirupananda Variyar Engineering College, Vinayaka Missions Research
Foundation (Deemed to be University), Salem, Tamil Nadu, India
Editors and Contributors xvii

Medha Maitra Bioprocess and Bioseparation Laboratory, Department of


Biotechnology, School of Bioengineering, College of Engineering and Technology,
SRM Institute of Science and Technology, Kattankulathur, Chengalpattu District,
Tamil Nadu, India
K. Manikanda Bharath Centre for Rural and Entrepreneurship Development,
National Institute of Technical Teachers Training and Research, Chennai, Tamil
Nadu, India
S. Mari Selvam Department of Biotechnology and Medical Engineering,
National Institute of Technology Rourkela, Rourkela, Odisha, India
Chirasmita Mohanty Centre for Biotechnology, Siksha ‘O’ Anusandhan
(Deemed to be University), Bhubaneswar, Odisha, India;
Department of Biotechnology, School of Bio Sciences and Technology, Vellore
Institute of Technology, Vellore, Tamil Nadu, India
Vijayanand Suryakant Moholkar Department of Chemical Engineering, Indian
Institute of Technology Guwahati, Guwahati, Assam, India
Pareshkumar G. Moradeeya Department of Environmental Science and
Engineering, Marwadi University, Rajkot, Gujarat, India
S. Nandhini Department of Microbiology, Ethiraj College for Women, Chennai,
India
C. Nirmala Department of Biotechnology, Paavai Engineering College, Paavai
Institutions, Namakkal, Tamil Nadu, India
Yuvraj Patil School of Health Sciences and Technology, Dr. Vishwanath Karad
MIT World Peace University, Pune, India
Selvakumar Periyasamy Department of Chemical Engineering, School of
Mechanical, Chemical and Materials Engineering, Adama Science and Technology
University, Adama, Ethiopia;
Department of Biomaterials, Saveetha Dental College and Hospitals, SIMATS,
Saveetha University, Chennai, India
P. Radha Bioprocess and Bioseparation Laboratory, Department of
Biotechnology, School of Bioengineering, College of Engineering and Technology,
SRM Institute of Science and Technology, Kattankulathur, Chengalpattu District,
Tamil Nadu, India
R. Rahul Department of Food Technology, Faculty of Engineering, Karpagam
Academy of Higher Education, Coimbatore, India
M. Rajasimman Department of Chemical Engineering, Annamalai University,
Chidambaram, Tamil Nadu, India
Devi Sri Rajendran Integrated Bioprocessing Laboratory, Department of
Biotechnology, Faculty of Engineering and Technology, School of Bioengineering,
xviii Editors and Contributors

SRM Institute of Science and Technology (SRMIST), Tamil Nadu, Kattankulathur,


Chengalpattu District, India
R. Rajesh Kannan Department of Chemical Engineering, Annamalai University,
Chidambaram, Tamil Nadu, India
Pavada Madhusudan Rao Bioprocess and Bioseparation Laboratory,
Department of Biotechnology, School of Bioengineering, College of Engineering
and Technology, SRM Institute of Science and Technology, Kattankulathur,
Chengalpattu District, Tamil Nadu, India
Umer Rashid Institute of Nanoscience and Nanotechnology, Universiti Putra
Malaysia (UPM), Serdang, Selangor, Malaysia;
Center of Excellence in Catalysis for Bioenergy and Renewable Chemical (CBRC),
Faculty of Science, Chulalongkorn University, Pathumwan, Bangkok, Thailand
Zinnabu Tassew Redda University of Applied Sciences (HTW) Berlin, Berlin,
Germany;
School of Chemical and Bio Engineering, Addis Ababa Institute of Technology,
Addis Ababa University, Addis Ababa, Ethiopia
M. Reenaa Department of Microbiology, Ethiraj College for Women, Chennai,
India
Samuel Lalthazuala Rokhum Department of Chemistry, National Institute of
Technology Silchar, Silchar, Assam, India
Prangya Ranjan Rout Department of Biotechnology, National Institute of
Technology Jalandhar, Jalandhar, Punjab, India
Joseph V. L. Ruatpuia Department of Chemistry, National Institute of
Technology Silchar, Silchar, India
Debasish Sahoo BioInnovale Lifescience (P) Ltd, Bengaluru, Karnataka, India
Ali Salisu Department of Chemistry, Faculty of Natural and Applied Science,
Sule Lamido University, Kafin Hausa, Jigawa State, Nigeria
Muhamad Amran bin Mohd Salleh Department of Chemical and
Environmental Engineering, Faculty of Engineering, Universiti Putra Malaysia,
Serdang, Selangor, Malaysia
Maria Sandberg Energy and Environmental Systems Group, Department of
Engineering and Chemical Sciences, Karlstad University, Karlstad, Sweden
Panchamoorthy Saravanan Department of Petrochemical Technology,
UCE—BIT Campus, Anna University, Tiruchirappalli, Tamil Nadu, India
V. Saravanan Department of Chemical Engineering, Annamalai University,
Chidambaram, Tamil Nadu, India
Editors and Contributors xix

Raja Selvaraj Department of Chemical Engineering, Manipal Institute of


Technology, Manipal Academy of Higher Education, Manipal, Karnataka, India
J. Mary Sheela Department of Microbiology, Ethiraj College for Women,
Chennai, India
Ibrahim Garba Shitu Department of Physics, Faculty of Natural and Applied
Science, Sule Lamido University, Kafin Hausa, Jigawa State, Nigeria
Joseph M Anto Simon Department of Biotechnology, Sri Krishna Arts and
Science College, Bharathiar University, Coimbatore, Tamil Nadu, India
M. Sridevi Department of Biotechnology, Vinayaka Missions Kirupananda
Variyar Engineering College, Vinayaka Missions Research Foundation (Deemed to
be University), Salem, Tamil Nadu, India
S. Sruthi Bioprocess and Bioseparation Laboratory, Department of
Biotechnology, School of Bioengineering, College of Engineering and Technology,
SRM Institute of Science and Technology, Kattankulathur, Chengalpattu District,
Tamil Nadu, India
Kolli Venkata Supraja Waste Treatment Laboratory, Department of Biochemical
Engineering and Biotechnology, Indian Institute of Technology Delhi, Hauz Khas,
New Delhi, India
Vinoth Kumar Vaidyanathan Integrated Bioprocessing Laboratory, Department
of Biotechnology, Faculty of Engineering and Technology, School of
Bioengineering, SRM Institute of Science and Technology (SRMIST), Tamil Nadu,
Kattankulathur, Chengalpattu District, India
Amirthavalli Velmurugan Department of Petroleum Engineering, Academy of
Maritime Education and Training, Chennai, Tamil Nadu, India
G. Venkatesh Energy and Environmental Systems Group, Department of
Engineering and Chemical Sciences, Karlstad University, Karlstad, Sweden
Swethaa Venkatraman Integrated Bioprocessing Laboratory, Department of
Biotechnology, Faculty of Engineering and Technology, School of Bioengineering,
SRM Institute of Science and Technology (SRMIST), Tamil Nadu, Kattankulathur,
Chengalpattu District, India
C. V. Vidhya Department of Microbiology, Ethiraj College for Women, Chennai,
India
S. M. Vidya Department of Biotechnology Engineering, NMAM Institute of
Technology, Nitte (Deemed to be University), Mangalore, Karnataka, India
Ramesh Vinayagam Department of Chemical Engineering, Manipal Institute of
Technology, Manipal Academy of Higher Education, Manipal, Karnataka, India
xx Editors and Contributors

Anita R. Warrier Nanophotonics Research Laboratory, Department of Physics,


Academy of Maritime Education and Training, Chennai, Tamil Nadu, India
Hlawncheu Zohmingliana Department of Chemistry, National Institute of
Technology Silchar, Silchar, India
Chapter 1
Introduction to Circular Bioeconomy
Perspectives in Sustainable Bioenergy
Production

Gurunathan Baskar, Veeramuthu Ashokkumar,


Samuel Lalthazuala Rokhum, and Vijayanand Suryakant Moholkar

Abstract Energy security and climate change risk are the two daunting issues faced
by developing economies across the globe. The industrial, agriculture and trans-
port sectors in developing nations follow a linear model of economy based on a
“take-make-dispose” pattern. The circular economy model that follows the pattern
of “reduce-repair-recycle-reuse” offers a potential solution. This book has dealt with
the application of circular economy principles to the energy sector. Basically, this
book has covered diverse aspects and issues related to the generation of energy from
waste and sustainable resources. The topics covered by the book chapters include
energy generation (in the form of liquid and gaseous fuels) from agro- and forest
residues, municipal solid wastes, industrial organic wastes and food wastes. Tech-
nical and economic viabilities of the thermochemical and biochemical technologies
for sustainable energy generation as well as issues related to supply chain manage-
ment of waste resources have been assessed. On a whole, this book has attempted
to identify both opportunities and challenges for application of circular bioeconomy
models for bioenergy generation, which could be useful for various stakeholders
from academia and industry.

G. Baskar
Department of Biotechnology, St. Joseph’s College of Engineering, Chennai 600119, Tamil Nadu,
India
V. Ashokkumar
Saveetha Institute of Medical and Technical Sciences (SIMATS), Saveetha University,
Chennai 602105, Tamil Nadu, India
S. L. Rokhum
Department of Chemistry, National Institute of Technology Silchar, Silchar 788101, Assam, India
V. S. Moholkar (B)
Department of Chemical Engineering, Indian Institute of Technology Guwahati,
Guwahati 781039, Assam, India
e-mail: vmoholkar@iitg.ac.in

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2024 1
G. Baskar et al. (eds.), Circular Bioeconomy Perspectives in Sustainable Bioenergy
Production, Energy, Environment, and Sustainability,
https://doi.org/10.1007/978-981-97-2523-6_1
2 G. Baskar et al.

Keywords Advanced engine techniques · Alternative fuels · Methanol · Emission


control

The twenty-first century has witnessed globalization and tremendous development


in industrial, transportation, agriculture and domestic sectors across all developing
economies in world. Present annual global energy consumption is estimated at 580
million terajoules, which is expected to grow to about 740 million terajoules (~ 30%
rise) by 2040. Fossil fuels—in the form of oil, coal and gas—currently supply about
80% of the total global energy demand. The use of fossil fuels for energy genera-
tion results in huge greenhouse gas (GHG) emissions, and in 2022, these emissions
were estimated at 34 billion tonnes of CO2 equivalent. Like all other industries, the
power sector has been following the linear economy model of “take-make-dispose”
pattern. The circular economy model respects the environment and follows the pattern
of “reduce-repair-reuse-recycle” to achieve the sustainable development. Circular
economy model aims to maximize resource efficiency, and minimize waste gener-
ation by the promoting reuse, recycling, and regeneration of materials and prod-
ucts. Circular bioeconomy essentially emphasizes upon the carbon cycle. Circular
bioeconomy aims to close the carbon cycle and creates an additional carbon sink
in the realm of technology by utilizing biogenic carbon for making products and
other anthropogenic activities. The circular bioeconomy can be deemed as a circular
carbon economy that stresses upon capturing atmospheric carbon via photosynthesis
to produce renewable bioresources, devising sustainable supply chains and devel-
oping new efficient technologies for sustainable transformation of the bioresources to
value-added products and fuels for electricity generation, transportation and domestic
use such as cooking. The circular bioeconomy model in the context of energy offers
simultaneous solutions to the two daunting issues, viz. energy security due to fast
depletion of fossil fuel reserves and climate change risk due to emissions from use
of fossil fuels. For a developing economy like India, which is highly dependent on
oil imports to meet its energy needs, the implementation of a circular bioeconomy
is crucial for achieving sustainable economic growth. During the G20 presidency
of India, a new and first-of-its-kind initiative was launched entitled “Resource Effi-
ciency and Circular Economy Industry Coalition (RECEIC)” with participation of
39 global companies.
This book has addressed multiple and diverse aspects related to generation of
energy from waste, renewable and sustainable sources. Eighteen chapters of this
book present comprehensive compilation and analysis of state-of-the-art informa-
tion on technologies for sustainable energy production from renewable and waste
resources and also discuss the opportunities and challenges for commercialization
of these technologies. This chapter gives a critical SWOT analysis of the role of
bioenergy in circular economy. Chapter 2 has focussed on the opportunities and
technical/economic challenges related to production of liquid biofuels from ligno-
cellulosic biomass. Basically, the technical issues related to biomass pretreatment
and fermentation for alcoholic biofuels (bioethanol and biobutanol) production are
1 Introduction to Circular Bioeconomy Perspectives in Sustainable … 3

discussed. Chapter 3 presents an overview of bioenergy production from indus-


trial organic wastes and byproducts as sustainable feedstocks. The industrial organic
wastes considered in this chapter include sugarcane bagasse, wood chips, potato/
orange peels, oil cake, sawdust and molasses. These wastes can be converted to
liquid and gaseous fuels through biochemical/thermochemical conversions.
Biodiesel has emerged as a potential renewable liquid fuel to be blended with
petroleum-derived diesel. Last two decades have seen intense research on synthesis
of biodiesel from different feedstocks using variety of processes. Waste cooking oil
produced in domestic kitchens and restaurants is also a potential waste feedstock for
biodiesel production. Chapter 4 discusses the potential and challenges of biodiesel
synthesis using waste cooking oil. Large-scale biodiesel production requires green
and efficient catalysts. Several chapters of this book have addressed this important
issue. Chapter 5 presents a review of green and non-toxic catalysts for biodiesel
production that are derived from biomass. These carbon- or calcium-based catalysts
have large surface area, improved porosity and stability, which contributes towards
improving biodiesel yield. Biochar produced during biomass pyrolysis forms a poten-
tial green base for catalysts due to high surface area and porosity. Physical and chem-
ical activation of biochar for further increase of surface area and enlargement of the
pores can enhance its catalytic activity. In addition, functionalization of the catalyst
surface by addition of specific functional groups for a specific chemical reaction
can make these catalysts suitable for particular applications. Chapter 7 extensively
discusses the functionalization of biochar to synthesize green and sustainable cata-
lysts for biodiesel production. Heterogeneous hydrochar-based catalysts have poten-
tial applications in biodiesel production. Chapter 14 has described the synthesis of
hydrochar from sustainable feedstocks like agricultural/forestry/food wastes. The
chapter also discusses improving the catalytic performance of the hydrochar through
addition of metal nanoparticles or acid/base functional groups to the surface of the
catalyst. These catalysts have been utilized for improving yield and selectivity during
biodiesel production.
Chapter 6 addresses the matter of valorization of hospital-generated plastic waste
for bioenergy production, as an alternative to conventional disposal through landfills.
Pyrolysis of this plastic waste in fixed or fluidized bed reactors can produce char and
hydrocarbons, which can be used for bioenergy production. Bioethanol has found
applications as a green and sustainable alcoholic transportation fuel. India has already
achieved the target of 10% ethanol blending in petrol. Chapters 8 and 16 discuss
production of 2G ethanol from lignocellulosic biomass. Especially, the pretreatment
techniques for improving the bioethanol yields have been analysed. The solid waste
produced in the urban sector is a potential resource for the production of biofuels
and other value-added chemicals. Chapter 9 presents an overview of biorefinery
avenues for processing of urban solid waste into value-added products. Natural waste
biomass produced from rural sector can also be potential feedstock for bioenergy
production. Chapter 10 discusses opportunities for the valorization of this sustainable
resource for bioenergy production through pyrolysis. Chapters 11 and 12 present an
important case study on biogas generation from organic waste (like food waste,
manure and silage) at Biogasbolaget AB in Karlskoga, Sweden. This case study also
4 G. Baskar et al.

describes enhancement in biogas production by addition of hydrochar produced from


forestry sector wastes and municipal organic wastes. An economic evaluation of the
profitability of this process (after integration with hydrothermal carbonization) has
also been presented. Chapter 14 discusses the potential of heterogeneous hydrochar-
based catalysts biodiesel. Various feedstocks for hydrochar production, production
technologies and factors affecting the properties of hydrochar have been discussed.
The aviation industry has flourished remarkably in recent years and so has the
demand for aviation fuel. To achieve sustainable growth in the aviation industry, there
is a pressing need for manufacture of aviation fuel through renewable resources.
Chapter 15 presents a review of the research in the synthesis of bio-jet fuel from
sustainable waste resources. Various technologies for conversion of waste resources
into jet fuel have been discussed from the perspective of a circular bioeconomy.
Economically feasible bioenergy production also requires a continuous and unin-
terrupted waste biomass supply chain. Different social/technical aspects and issues
related to this supply chain have been discussed in Chap. 17. Lastly, Chap. 18 presents
a review and technical/economic assessment of a food waste-based biorefinery for
bioenergy production. This chapter concludes that economic viability of a food waste-
based biorefinery relies on the utilization of the organic waste for the synthesis of
a diverse range of green products within the circular economy framework.
We, the editors of this monograph, believe that the contents of this book form a
compendium of social, technical and economic facets of bioenergy from the view-
point of circular bioeconomy. We hope that academic and industrial fraterni-
ties working in the areas of sustainable development, policy planning, renewable
energy and waste-to-wealth will find this book a valuable source of state-of-the-art
information and critical analysis.
Chapter 2
Bioenergy-Based Sustainable
Bioeconomy—Perspectives
and Challenges

C. V. Vidhya, S. Nandhini, J. Mary Sheela, and M. Reenaa

Abstract Bioenergy is a renewable energy that can be obtained, from biological


resources in abundance. The resources include microorganisms, plants, animals and
recently, biomass. The applications of bioenergy are wide. It can be used for cooking
purpose, power generation, etc. It can also be used as a substitute for petroleum
products in both liquid and gaseous forms. Bioenergy is been considered as Nextgen
fuel, replacing the other fossil fuels, either partially or completely. Bioeconomy
refers to a prudential usage of bioenergy obtained from biological resources. It also
deals with conservation of bioenergy, thereby, aiming at sustainable economy for
the human welfare. The precise meaning of biological resources not only includes
plants, animals, microorganisms or biomass but also the relatable scientific and tech-
nological knowledge and innovative ideas based on the same. In the early twenty-first
century, the European Union (EU) and Organization of Economic Cooperation and
Development (OECD) provided a spotlight on bioeconomy concept. The strategies
of bioeconomy, in a holistic approach, aim at sustainable development with circular
economy. Sustainable bioeconomy strongly focuses in mitigating global challenges
like depletion of fossil fuels, endangered biodiversity, unpredictable inconsistent
climatic changes, environmental pollution, etc. It paves a new approach by promoting
green-based employment. Thereby, our country’s economic status will escalate. Its
paramount focus is on the manufacture and consumption of world’s finite resources
the bioenergy, in an equitable proportion with circularity. This chapter deals with the
bioenergy, bioeconomy, circular economy, perspectives, and challenges of the same.

Keywords Bioenergy · Sustainable bioeconomy · Circular economy

C. V. Vidhya (B) · S. Nandhini · J. M. Sheela · M. Reenaa


Department of Microbiology, Ethiraj College for Women, Chennai, India
e-mail: vidhya_cv@ethirajcollege.edu.in

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2024 5
G. Baskar et al. (eds.), Circular Bioeconomy Perspectives in Sustainable Bioenergy
Production, Energy, Environment, and Sustainability,
https://doi.org/10.1007/978-981-97-2523-6_2
6 C. V. Vidhya et al.

2.1 Bioenergy and Bioeconomy—Introduction

Bioenergy is a renewable energy. It can be procured, in abundance, from biological


resources that includes microorganisms, plants, and animals. The Intergovernmental
Panel on Climate Change 2019 states that bioenergy with carbon capture and storage,
in short BECCS, has the ability to control the emission of greenhouse gases. Bioen-
ergy is one among the important decarbonization source in the energy transition. It
can be used as a near zero-emission fuel. Besides being responsible for the release
of carbon into the atmosphere, bioenergy is still considered to be less harmful when
compared with other fossil fuels on being burnt. This is because it utilizes and releases
carbon currently, whereas fossil fuels release carbon that has been stored for millions
of years. Biofuels are the most common form of bioenergy. Bioethanol and biodiesel
are usually the common types of biofuels—green diesel, biogasoline, biobutanol,
biogas, and jet biofuel [1].
Bioenergy is been considered as Nextgen fuel, replacing the other fossil fuels,
either partially or completely. There is a great demand in the production of bioenergy
as there is a great flexibility in its usages. For example, bioenergy in its solid form
can be used in domestic combustion purpose and industrial as well. The liquid form
of biofuels can be used in motor cars, airway, and seaway transport. Natural gas can
be replaced by biomethane. Bioenergy and carbon capture and storage (BECCS) is
a novel upcoming idea as it can be an economic strategy to gain negative emission
that may reduce atmospheric carbon [2].
Bioeconomy refers to a prudential usage of bioenergy from biological resources,
and their conservation, thereby, aiming at sustainable economy for the human
welfare. In the early twenty-first century, it was European Union (EU) and Orga-
nization of Economic Cooperation and Development (OECD), who initially adopted
the concept of bioeconomy, in their framework, that promoted the wider applica-
tions of biotechnology. The strategies of bioeconomy, collectively, aim at sustainable
development with circular economy. Sustainable bioeconomy intensely focuses in
mitigating global challenges like exhaustion of fossil fuels, endangered biodiversity,
erratic climatic changes, environmental pollution, etc. It also supports a country’s
economy by promoting green-based employment. Its ultimate focus is on the produc-
tion and consumption of world’s finite resources in an equitable proportion with
circularity. This chapter deals with the bioenergy, bioeconomy based on bioenergy,
bioeconomy with circular economy, perspectives of the same, the challenges to build
it, and the challenges dealt by the same.

2.2 Sources of Bioenergy

Bioenergy plays a crucial role in the global pursuit of sustainable development and
the transition to a low-carbon economy. Bioenergy is derived from organic materials,
which can be continuously replenished through natural processes. Unlike fossil fuels,
2 Bioenergy-Based Sustainable Bioeconomy—Perspectives and Challenges 7

which are finite and contribute to climate change, bioenergy sources have the potential
to be sustainable and help reduce greenhouse gas emissions. It offers an opportunity
to diversify energy sources, reduce dependence on imported fossil fuels, and enhance
energy security.
Many countries can produce biomass locally, reducing their vulnerability to inter-
national energy price fluctuations and geopolitical tensions. Bioenergy has the poten-
tial to significantly reduce carbon dioxide emissions when compared to conventional
fossil fuels. It can be easily integrated into existing energy infrastructure, such as
power plants, heating systems, and vehicles. This adaptability makes it a feasible
option for reducing carbon emissions in various sectors.
It is derived from biological material known as biomass. Biomass has the potential
to be an increasingly cost-competitive renewable energy source and be used to make
a valuable contribution to the overall energy supply system, mainly because of its
very low cost and the fact that it is renewable. There is still considerable scope for
making better use of the existing biomass energy supplies and for developing new
supplies.
Biofuels like bioethanol and biodiesel can be used as alternatives to fossil fuels
in the transportation sector, reducing emissions from the transportation industry,
which is a significant source of greenhouse gases. Some bioenergy technologies,
such as biogas and solid biomass power plants, can provide a stable base load of
electricity and contribute to grid stability, particularly when paired with intermittent
renewable energy sources like solar and wind. The development and implementation
of bioenergy technologies drive research and innovation in the renewable energy
sector, leading to advancements in efficiency, cost-effectiveness, and sustainability.

2.2.1 Biomass

Biomass is a versatile and sustainable source of renewable energy that plays a crucial
role in meeting our needs. Biomass refers to any organic material derived from
plants or animals that can be used as an energy source. This can include wood,
crop residues, agricultural and forestry byproducts and organic waste. Biomass feed-
stocks are diverse and can be tailored to suit regional and environmental conditions.
This flexibility allows for the utilization of different types of biomasses depending
on local availability and economic considerations.
Ongoing research and development in the field of bioenergy are leading to tech-
nological advancements that enhance the efficiency and environmental performance
of biomass conversion. The economic viability of bioenergy production depends
on factors such as feed-stock costs, energy market prices, and government support.
Biomass is a valuable resource for bioenergy production with the potential to reduce
carbon emissions, enhance energy security, and support rural economies. However, its
sustainable requires careful consideration of feed-stock selection, technological inno-
vations, and supportive policies to ensure its long-time environmental and economic
benefits.
8 C. V. Vidhya et al.

2.2.1.1 Types of Biomass

Feed-stock signifies raw substances or inputs used in various processes, particularly


in the context of bioenergy production. Feed-stocks are broadly classified into three
generations based on the kind of feed-stock which is being used for its production
(Fig. 2.1).

First-Generation Feed-Stock

First-generation feed-stock refers to those sources that have been traditionally used
for biofuel production. These feed-stocks are typically derived from food crops or
crops that are primarily grown for human consumption. The production of biofuels
from first-generation feed-stock has been the starting point for the development of
bioenergy technologies. The types of first-generation feed-stock are as follows.
Corn (Maize) is a widely utilized feed-stock for ethanol production due to its high
starch content. In this process, the starch found in corn kernels is transformed into
sugars, which are subsequently fermented to produce ethanol.
Sugarcane, on the other hand, is another prominent source of bioenergy, partic-
ularly ethanol. Sugars derived from sugarcane juice or molasses undergo fermenta-
tion to yield ethanol. This feed-stock is prized for its high sugar content, efficient
conversion rate, and the presence of well-established production systems.

Fig. 2.1 Classification of biomass feed-stocks


2 Bioenergy-Based Sustainable Bioeconomy—Perspectives and Challenges 9

Palm oil, derived from the fruit of the oil palm tree, serves a dual role as a versatile
and widely consumed vegetable oil and as a prominent feed-stock for biodiesel
production. The oil undergoes intricate processing to yield biodiesel, which has
found extensive use in transportation and industrial applications.
Wheat, a staple cereal grain known for its versatility and widespread cultivation,
serves as a crucial first-generation feed-stock in the production of bioethanol. It is
rich in starch, which can be converted into ethanol through a process that mirrors the
production of bioethanol from corn.
Canola is a versatile oilseed crop cultivated primarily for its oil-rich seeds. It is
known for its high oil content, particularly in the form of oleic acid, linoleic acid,
and linolenic acid. Canola oil is not only used in the culinary industry but also serves
as a valuable feed-stock for the production of biodiesel.
Barley is a cool-season cereal grain crop that has been cultivated for thousands
of years, primarily for use in human and animal nutrition. However, in recent years,
barley has gained recognition as a valuable first-generation feed-stock for bioethanol
production. This is due to its high starch content, which can be enzymatically
converted into ethanol, making it an ideal for sustainable biofuel production.
Soybean plays a vital role in biodiesel production through a transesterification
process. Soybean oil is the primary component converted into biodiesel, with glyc-
erol as a valuable byproduct of this transformation. This diversification of feed-
stocks underscores the versatility and sustainability of bioenergy production, offering
multiple pathways to reduce our reliance on traditional fossil fuels.

Production Process
Ethanol Production The production of ethanol from various feed-stocks involves
a series of essential steps. First, there is the pretreatment phase, which encompasses
feed-stock preparation and milling to make the starch or sugars readily accessible
for subsequent processes. Following pretreatment, enzymatic hydrolysis comes into
play, where complex carbohydrates are broken down into simpler sugars, a crucial
step in the ethanol production chain. Once the sugars are available, fermentation
takes center stage. This phase employs yeast or bacteria to transform these sugars
into ethanol, marking a pivotal point in the conversion process. After fermentation,
the distillation step comes into action, serving the purpose of separating ethanol
from the fermentation mixture. The final step in ethanol production is dehydration,
aimed at removing excess water to achieve anhydrous ethanol. These meticulously
orchestrated processes together contributed to the successful production of ethanol,
a versatile biofuel with the potential to reduce our dependence on conventional fossil
fuels.
Biodiesel Production Biodiesel production follows a well-defined sequence of
steps, with each phase contributing to the creation of this sustainable biofuel.
Biodiesel, a renewable and eco-friendly alternative to conventional diesel fuel, is
primarily produced through a chemical process known as transesterification. This
process involves converting triglycerides, which are typically found in vegetable oils
10 C. V. Vidhya et al.

or animal fats, into biodiesel (fatty acid methyl esters) and glycerin. Transesterifi-
cation is a crucial step in the biodiesel production process, as it alters the chemical
structure of the feed-stock to create a fuel suitable for use in diesel engines.
Transesterification Reaction: The transesterification reaction is catalyzed and
typically involves the following steps:
Feed-stock Preparation: Vegetable oils or animal fats are first heated to reduce their
viscosity and enhance their reactivity.
Alcohol Addition: Methanol or ethanol is added to the heated feed-stock. These
alcohols serve as the reactants and work to replace the glycerol component in the
triglycerides.
Catalyst Introduction: A catalyst is introduced to facilitate the reaction. Catalysts
play a pivotal role in speeding up the process and enhancing its efficiency.
Stirring and Heating: The reaction mixture is stirred and heated to promote the
complete conversion of triglycerides into biodiesel and glycerin.
Product Separation: Product separation in biodiesel production is a crucial stage,
as it enables the isolation of the desired biodiesel product from the less valuable
byproducts, especially glycerin. It is during this phase that the biodiesel can be further
refined to meet quality standards, ensuring it is free from contaminants and impurities
that could hinder its performance in diesel engines. This step also emphasizes the
sustainable nature of biodiesel production, as both the biodiesel and glycerine can find
applications in various industries, reducing waste and maximizing the value of the
feed-stock used in the process. Proper separation, followed by thorough purification,
is essential to produce high-quality biodiesel that meets regulatory requirements and
maintains the integrity of diesel engines while simultaneously contributing to a more
environmentally friendly and renewable energy source.
Role of Catalyst in Transesterification
Catalyst Function: Catalysts are fundamental to the transesterification reaction
because they significantly accelerate the reaction rate, reduce the reaction time, and
improve the overall efficiency of biodiesel production. They achieve this by lowering
the activation energy required for the reaction, allowing it to occur more readily.
Types of Catalysts: Historically, sodium hydroxide (NaOH) and potassium
hydroxide (KOH) have been widely used as catalysts in transesterification reac-
tions. These alkali catalysts are effective in promoting biodiesel formation, but they
have some drawbacks, including the potential for soap formation and the need for
water removal. Acid catalysts, such as sulfuric acid, are also used but require careful
handling due to their corrosive nature.
2 Bioenergy-Based Sustainable Bioeconomy—Perspectives and Challenges 11

Current Innovations in Catalysts for Transesterification


Enzymatic Catalysts: One of the exciting innovations in biodiesel production is the
use of enzymatic catalysts. Enzymes, like lipases, can catalyze the transesterifica-
tion reaction under milder conditions, reducing energy consumption and the need
for excessive stirring. They also produce fewer impurities, making the purification
process easier.
Solid Catalysts: Heterogeneous solid catalysts, such as metal oxides and zeolites,
are gaining attention due to their reusability and the potential for reducing waste in
the biodiesel production process. Solid catalysts can facilitate the transesterification
reaction effectively while minimizing the formation of soaps and glycerin.
Nano-Catalysts: Nano-catalysts are on the horizon as a promising avenue for further
enhancing the efficiency of transesterification reactions. These tiny particles offer a
high surface area, which can lead to faster reaction rates and greater control over the
biodiesel production process.
Ionic Liquids: Ionic liquids are another innovative approach. They can serve as both
solvents and catalysts, eliminating the need for additional chemicals and simplifying
the reaction process. Ionic liquids are also highly tunable, allowing for precise control
over reaction conditions.
These advancements in catalyst technology are improving the sustainability
and cost-effectiveness of biodiesel production, making it a more viable option
for reducing carbon emissions and promoting a cleaner, renewable energy source.
Researchers and industry experts continue to explore and refine these catalyst
innovations to enhance the efficiency and environmental friendliness of biodiesel
production.

First-Generation Environmental and Social Impacts


The production and utilization of bioenergy have both positive and negative environ-
mental and social implications. On the positive side, bioenergy offers several advan-
tages. It contributes to the reduction of greenhouse gas emissions when compared
to conventional fossil fuels, making it a valuable tool in combatting climate change.
Moreover, when managed sustainably, bioenergy sources can be renewed over time,
providing a potentially long-term and environmentally friendly energy solution.
Additionally, bioenergy can enhance energy security by reducing the dependency
on imported fossil fuels, thus increasing a region’s energy self-reliance.
However, bioenergy also comes with its share of disadvantages and challenges.
One of the main concerns is land use competition, which arises when land is allocated
for bioenergy production at the expense of food production or natural habitats. This
competition can lead to conflicts and issues related to food security and biodiversity
conservation. Additionally, bioenergy production often requires substantial water
resources, which can result in increased water consumption and potential pollution
if not managed properly. Lastly, social concerns related to land rights, displacement
of local communities, and fluctuations in food prices are important considerations
12 C. V. Vidhya et al.

when implementing bioenergy projects. Balancing these advantages and disadvan-


tages is crucial in ensuring that bioenergy contributes positively to our sustainable
energy future while mitigating potential adverse impacts. First-generation feed-stock
has been essential in the early development of bioenergy technologies. However,
concerns over sustainability and competition with food production have spurred
research into more advanced feed-stock options.

Second-Generation Feed-Stock
Second-generation feed-stock marks a substantial progression in the realm of bioen-
ergy production. In contrast to first-generation feed-stock, which originates from food
crops, second-generation feed-stock predominantly centers on non-food biomass
sources. Bioenergy production relies on a diverse range of feed-stocks, and two
prominent categories are agricultural residues and woody biomass. Agricultural
residues, including materials like wheat straw, corn stover, and rice husks, are a
valuable resource for bioenergy due to their availability as byproducts of agricultural
activities. The utilization of these residues reduces competition with food production,
making it a sustainable option.
On the other hand, woody biomass comprises forestry residues, sawdust, and the
cultivation of dedicated energy crops like willow and poplar. This category provides
another substantial source of biomass for bioenergy production, tapping into the
potential of wood-related materials while promoting sustainable forestry practices.
Both agricultural residues and woody biomass offer environmentally friendly alter-
natives to traditional fossil fuels, contributing to the diversification of our energy
sources and the reduction of greenhouse gas emissions.

Second-Generation Production Process


The conversion of biomass into valuable biofuels involves a series of intricate
processes, each contributing to the transformation of biomass feed-stock. Initially,
pretreatment methods are employed, encompassing physical, chemical, or biolog-
ical processes. The primary objective here is to break down the complex structure of
biomass, increasing the accessibility of cellulose and hemicellulose for subsequent
processing. Following pretreatment, enzymatic hydrolysis comes into play, where
specialized enzymes play a pivotal role in breaking down cellulose and hemicellu-
lose into fermentable sugars. This step requires enzymes capable of targeting the
intricate biomass structures effectively.
The journey continues with fermentation, where microorganisms like bacteria
or yeast are utilized to convert these sugars into biofuels, such as ethanol. This
biological process is a cornerstone in the production of biofuels from biomass.
Additionally, thermochemical conversion methods, such as pyrolysis, gasification, or
hydrothermal liquefaction, can also be employed to extract biooil, syngas, or biocrude
from biomass. These sophisticated processes underscore the versatility and potential
of biomass as a renewable and environmentally friendly source of bioenergy.
2 Bioenergy-Based Sustainable Bioeconomy—Perspectives and Challenges 13

Second-Generation—Environmental and Social Impacts


The utilization of non-food biomass in bioenergy production offers several advan-
tages worth considering. Firstly, it reduces competition with food crops, alleviating
concerns related to food security and resource allocation. Additionally, this approach
has the potential for higher energy yields and lower greenhouse gas emissions,
contributing to a more environmentally sustainable energy landscape. Furthermore,
bioenergy production from non-food biomass provides an opportunity for waste
valorization, effectively reducing waste disposal and promoting a circular economy.
However, it is essential to acknowledge and address the challenges associated
with this approach. One primary challenge lies in the technological complexity and
higher processing costs often associated with non-food biomass. These complexities
can impact the economic viability of bioenergy projects. Sustainable sourcing of
non-food biomass and responsible land use considerations for energy crops are also
key challenges that must be carefully managed to prevent adverse environmental and
social impacts. Additionally, there is a need to mitigate the potential effects on biodi-
versity and ecosystems when converting non-food biomass into bioenergy, ensuring
a balanced approach to sustainable energy production. Second-generation feed-stock
represents a promising avenue for sustainable bioenergy production. Understanding
the potential of these non-food biomass sources and the technological advancements
in their conversion processes can contribute to the advancement of renewable energy
solutions and address the challenges posed by conventional fossil fuels. The future
of bioenergy lies in these innovative and environmentally responsible feed-stock
options.

Third-Generation Feed-Stocks

Third-generation feed-stocks represent a remarkable advancement in the realm of


bioenergy production, focusing on unconventional and eco-friendly sources that
bring unique advantages. Diverging from the paths of first and second-generation
feed-stocks, third-generation feed-stocks are charting alternative courses that hold
promise for overcoming various challenges inherent to traditional biofuels. A thor-
ough understanding of these cutting-edge feed-stocks and their potential applications
is pivotal for the advancement of sustainable bioenergy technologies.
Among the diverse array of third-generation feed-stocks, algae take center stage.
These aquatic organisms can be cultivated in various settings, including ponds, tanks,
or specialized photo-bioreactors. Algae stand out due to their rapid growth rates and
the potential to produce lipids suitable for biodiesel and other valuable compounds.
Within the realm of algae, micro-algae, microscopic aquatic organisms proficient
in photosynthesis, come into play. These microorganisms are prized for their high
oil content, rendering them suitable candidates for biodiesel production. They can
be cultivated in diverse environments, ranging from open ponds to closed photo-
bioreactors and even wastewater treatment systems.
14 C. V. Vidhya et al.

Macro-algae, commonly known as seaweeds, form another branch of third-


generation feed-stocks. These large aquatic plants flourish in marine environ-
ments, characterized by their rapid growth rates and significant carbon capture
capacity. This group exhibits potential for various bioenergy applications, including
bioethanol, biogas, and the extraction of other valuable compounds. The exploration
of these third-generation feed-stocks represents a dynamic shift in the bioenergy
landscape, with the potential to redefine sustainable energy production in a more
environmentally conscious and efficient manner.

Third-Generation—Production Process
The production process for third-generation feed-stocks, such as micro-algae
and macro-algae, involves distinct cultivation methods and harvesting techniques.
Microalgae cultivation necessitates essential resources like light, water, and nutrients,
typically nitrogen and phosphorus. Large-scale production is achieved through photo-
bioreactors or open ponds, depending on the scale of the operation. The harvesting
of micro-algae employs various methods, including centrifugation, filtration, and
flocculation, to separate the biomass efficiently.
In contrast, macro-algae cultivation primarily occurs in coastal areas or specialized
offshore farms. The harvesting of macro-algae can be conducted manually or using
mechanical methods, which depends on the specific context and scale of the oper-
ation. One notable advantage of macro-algae cultivation is its minimal demand for
freshwater and land resources, making it a sustainable choice for bioenergy produc-
tion that aligns with environmental conservation efforts. This distinctive cultivation
and harvesting processes highlight the adaptability and potential of third-generation
feed-stocks in bioenergy production, offering a glimpse into a more sustainable and
efficient future for renewable energy sources.

Third-Generation—Environmental and Social Impacts‘


The production of bioenergy from third-generation feed-stocks like micro-algae and
macro-algae presents a range of environmental and social implications. On the posi-
tive side, there are several advantages to consider. These feed-stocks can thrive in
non-arable land, reducing conflicts over land use for food production. Furthermore,
depending on the cultivation and conversion methods employed, micro-algae and
macro-algae have the potential to be carbon–neutral or even carbon-negative, which
can significantly contribute to mitigating greenhouse gas emissions. Additionally,
they offer the opportunity for integration with wastewater treatment processes, aiding
in reducing pollution and nutrient runoff, thereby benefiting aquatic ecosystems.
However, it is essential to recognize and address the challenges associated with
third-generation feed-stock production. High cultivation and processing costs can
limit the commercial viability of these bioenergy sources, necessitating further
research and development to optimize cost-effectiveness. Technical challenges, such
as achieving high yields and maintaining continuous cultivation, require innovative
solutions to maximize the potential of these feed-stocks. Additionally, if not properly
2 Bioenergy-Based Sustainable Bioeconomy—Perspectives and Challenges 15

managed, the cultivation and harvesting of micro-algae and macro-algae can have
adverse impacts on aquatic ecosystems, emphasizing the need for sustainable and
responsible practices in their production. Balancing these advantages and challenges
is vital in harnessing the potential of third-generation feed-stocks for sustainable and
environmentally friendly bioenergy production.

2.2.2 Miscellaneous Sources of Bioenergy (Fig. 2.2)

Energy Crops: Dedicated crops grown specifically for bioenergy production, such
as switchgrass and miscanthus.
Municipal Solid Waste: Waste materials from households, businesses, and indus-
tries can be processed to extract biogas through anaerobic digestion or incineration
for energy recovery.
Landfill Gas: Organic waste in landfills generates methane and other gases, which
can be collected and used as a source of renewable energy.

Fig. 2.2 Different sources of bioenergy


16 C. V. Vidhya et al.

Wastewater Treatment: Organic matter in sewage and wastewater can be treated


anaerobically to produce biogas, which is a sustainable energy source.
Animal Waste: Livestock and poultry waste can be processed through anaerobic
digestion to produce biogas, reducing greenhouse gas emissions and providing
energy.
Energy from Crop Residues: Agricultural residues, such as sugarcane bagasse and
rice husks, can be utilized for energy generation.
Bioenergy Derived from Food Processing Residues: Energy can be generated from
the byproducts and waste produced from food processing industries.

2.2.3 Technological and Policy Developments

Technological advancements and evolving policy frameworks are driving significant


developments in the field of bioenergy production. One notable innovation is the
emergence of biorefineries, which are integrated systems designed to maximize the
extraction of value from various feed-stocks. These biorefineries facilitate the produc-
tion of not only biofuels but also a range of bioproducts and chemicals, promoting a
more diversified and sustainable bioeconomy.
In parallel, genetic engineering research is making strides in enhancing feed-
stock properties through genetic modification. This involves improving factors such
as biomass yield, composition, and stress resistance, thereby enhancing the efficiency
and viability of bioenergy production.
Furthermore, third-generation feed-stocks, particularly micro-algae and macro-
algae, have emerged as a captivating frontier in bioenergy research and development.
These feed-stocks possess substantial potential for sustainable biofuel production and
other applications, contributing to the diversification of renewable energy sources and
environmental sustainability.
Exploring these state-of-the-art technologies and understanding their implica-
tions for bioenergy production will play a pivotal role in our ongoing efforts to
reduce our dependence on fossil fuels and address the pressing challenges posed by
climate change. These developments mark crucial steps toward a more sustainable
and environmentally friendly energy landscape.

2.2.4 Sustainability and Carbon Balance

Sustainability and achieving a balanced carbon footprint are paramount considera-


tions in the realm of bioenergy production. Sustainable bioenergy practices prioritize
environmentally responsible sourcing and production, aiming to minimize adverse
impacts on ecosystems. This entails practices that avoid deforestation and other
2 Bioenergy-Based Sustainable Bioeconomy—Perspectives and Challenges 17

harmful effects while ensuring that the entire bioenergy life cycle is managed with a
commitment to preserving our environment.
In the context of carbon neutrality, bioenergy systems strive to maintain equilib-
rium in carbon emissions. Carbon-neutral bioenergy systems are designed to offset
the carbon dioxide released during biomass combustion by reabsorbing an equiva-
lent amount of carbon dioxide through the growth of new biomass. This approach
not only reduces the net emissions of greenhouse gases but also helps mitigate the
effects of climate change. By embracing sustainable and carbon-neutral bioenergy
practices, we move closer to a future where renewable energy sources play a pivotal
role in mitigating environmental challenges and advancing a more sustainable energy
landscape.

2.2.5 Classification of Bioenergy

Bioenergy categorization can vary depending on multiple factors, such as the type of
biomass employed, the conversion technique, and the ultimate purpose. Figure 2.3
explains the classification of bioenergy.

Fig. 2.3 Classification of bioenergy


18 C. V. Vidhya et al.

2.3 Applications of Bioenergy Products

The bioenergy can contribute for more sustainable, safe, secure and economically
powerful applications as it can supply domestic clean energy sources, revitalize rural
economies, reduce the burden of countries to depend on foreign oil and petroleum
products, generate green-based jobs, etc.
The applications are wider starting from simple stove cooking to rocket fuels. The
applications are as follows:
• Applications of biomass energy for residential purpose especially, cooking with
small stoves.
• It can also be used in small scale to large-scale power plants with centralized
utilities for the production of electricity.
• It can be used for the production of day-to-day products
• It is used to produce transportation fuels
• It can also be used for production of aviation fuels
• It increases bioenergy-based job opportunities.
For the abovementioned vast applications, the bioenergy products are the sole
resources. The prudential uses of bioenergy products will lead to sustainable clean
energy supply. The bioenergy products with their application are as follows:

2.3.1 Thermal Conversion

Biomass can be burned by thermal conversion and used as energy source. This thermal
conversion includes a very high temperature starting from 180° (carbonization) to
1000 °C (gasification). Raw materials such as municipal solid waste (MSW) and
scraps from paper or lumber mills are the most familiar biomass feed-stocks for
thermal conversion. Energy is created through direct firing, co-firing, pyrolysis, gasi-
fication, and anaerobic decomposition. Compressed briquettes have high-energy
density and are generally easy to burn during direct or co-firing.

2.3.2 Direct Firing and Co-firing

Steam produced during the firing process of briquettes is used as power for turbine
which in turning on a generator and in electricity production. Co-firing eliminates
the need for new factories for biomass processing. Co-firing also eases the demand
for coal. This also reduces the amount of carbon dioxide and other greenhouse
gases released by burning fossil fuels.
2 Bioenergy-Based Sustainable Bioeconomy—Perspectives and Challenges 19

2.3.3 Syngas

Syngas is a combination of hydrogen and carbon monoxide majorly. It also has


fewer percentage of methane and hydrocarbons. This can be purified. It is cleaned
of sulfur, particulates, mercury, and other pollutants. The clean syngas is processed
into transportation biofuels.

2.3.4 Biochar

Biochar is a type of charcoal and a carbon-rich solid which is specifically useful


in agriculture. Biochar enriches soil and prevents it from leaching. Carbon sinks are
reservoirs for carbon-containing chemicals, including greenhouse gases, and biochar
acts as an excellent carbon sink. Biochar also helps in enriching the soil.

2.3.5 Anaerobic Decomposition

Anaerobic decomposition is a process where microorganisms, usually bacteria, break


down the available material in the absence of oxygen. Anaerobic decomposition is
an important process in landfills, where biomass is crushed and compressed, creating
an anaerobic (or oxygen-poor) environment. Manure and other animal waste can be
converted to sustainably meet the energy needs.

2.3.6 Biofuel

Biomass when converted into liquid biofuels such as ethanol and biodiesel acts as
a suitable renewable energy source. Biomass that is high in carbohydrates, such
as sugarcane, wheat, or corn when fermented gives ethanol. Biodiesel is made by
combining ethanol with animal fat, recycled cooking fat, or vegetable oil. Though
biofuels help in operating vehicles, they do not operate as efficiently as gasoline.
However, they can be blended with gasoline to make use of it efficiently to power
vehicles and machinery, and in turn do not release the emissions associated with fossil
fuels. Ethanol has become a popular substitute for wood in residential fireplaces.
When burned, it gives off heat in the form of flames, and water vapor only and not
smoke [3].
20 C. V. Vidhya et al.

2.3.7 Black Liquor

Processing of wood into paper produces a high-energy, toxic substance called black
liquor. With the invention of the recovery boiler in the 1930s, black liquor could be
recycled. In the USA, paper mills use nearly all of their black liquor formed to run
their mills, and because of this forest industry is one of the most energy-efficient
in this nation. Sweden has experimented very recently in gasifying black liquor to
produce syngas, which is further used to generate electricity.

2.4 Bioeconomy

The terms bioeconomy means the economy that is developed by the production and
usage of bioenergy, developed through biomass resources, from earth, like microbes,
plants, and animals. This bioenergy, thus produced, is utilized for the production in
industries and health sector. In simple terms, the economy based on bioenergy is
defined as bioeconomy. The perspectives of bioeconomy completely rely on the
properly structured protocols and mechanisms, in manipulating the genomic and
molecular level, so as to improve the industrial production to produce, improved and
novel products.
In the report “The Knowledge Based Bio-Economy (KBBE) in Europe: Achieve-
ments and Challenges” by Albrecht et al. [4], the term bioeconomy was defined as
follows: The bioeconomy is the sustainable production and conversion of biomass,
for a range of food, health, fiber and industrial products and energy, where renew-
able biomass encompasses any biological material to be used as raw material.” [4].
According to a 2013 study, “the bioeconomy can be defined as an economy where the
basic building blocks for materials, chemicals and energy are derived from renewable
biological resources.” [5]
In Berlin (November 2015), the first global Bioeconomy Summit was orga-
nized. It defined bioeconomy as “knowledge-based production and utilization of
biological resources, biological processes and principles to sustainably provide
goods and services across all economic sectors.” According to the summit, bioe-
conomy comprises of three important elements: renewable biomass, enabling and
converging technologies, and integration across applications concerning primary
production (living natural resources), health (medical and pharma-related devices),
and industries [6].
The ideology of bioeconomy is to reduce and finally stop the dependence on
exhausting renewable resources like coal, petroleum, and other fossil fuels. More
than 60 countries have bioeconomy-based strategies, out of which twenty of them
have published dedicated bioeconomy strategies in Asia, America, Africa, Oceania,
and Europe [7].
A suitable incident was the implementation of the Energy Independence and
Security Act (EISA) by the USA, in 2007. The only goal of EISA was to escalate the
2 Bioenergy-Based Sustainable Bioeconomy—Perspectives and Challenges 21

availability of renewable energy. And it was planned to be through the production of


biofuel [8]. Fifth Fuel Policy was introduced by the Government of Malaysia, in the
8th Malaysia Plan structured for years 2001–2005.
The aim was to motivate the industrial field for the production of bioenergy,
worldwide. An overall share of 20% renewable energy has been set as a target in
EU’s transport consumption by the European Commission in the year 2020 [9].

2.4.1 Bioeconomy: Concept, Policies, and Future


Perspectives

The bioeconomy/bio-based economy refers to the actions that are driven by research
and innovations related to the production, application, and conservation of biolog-
ical resources with the ultimate aim of a sustainable environment and economy in
the future. It includes all commercial and industrial sectors that rely on consuming,
creating, or processing biological resources derived from forestry, agriculture, aqua-
culture, and organic waste. The bioeconomy seeks to operate according to the circular
economy’s guiding principles of reuse, repair, and recycling, as opposed to the
traditional linear or fossil-based economy (Fig. 2.4).
In the first decade of the twenty-first century, the term “bioeconomy”—a
“leapfrogging” strategy became well known as a result of its adoption by the European

Fig. 2.4 Concept of bioeconomy


22 C. V. Vidhya et al.

Union (EU-2012) and the Organization for Economic Cooperation and Development
(OECD-2006) as a framework to promote new goods, procedures, and markets by
accelerating biotechnology innovation. The multi-sector perspective of EU bioe-
conomy policies is to safeguard the environment, prevent overexploitation of natural
resources, and increase biodiversity by adopting strategies dependent on biological
resources (plants, animals, microorganisms, and their products, or biomass).
Numerous nations support and implement comprehensive policies and initiatives
to advance their bioeconomies as a means of addressing environmental concerns like
climate change, food security, and energy independence.
The United Nations (UN) adopted the Bioeconomy 2030 Agenda in 2015, which
included 17 Sustainable Development Goals (SDGs). According to the FAO studies,
the following four connected aspects of the bioeconomy are crucial:
1. The elimination of poverty, hunger, and the reduction of inequality.
2. Sanitation and clean water mission.
3. Encourage environmentally friendly business and infrastructure.
4. Benefit for climate action and the health of land and aquatic ecosystems.
Although not mutually exclusive, the National Academies of Science, Engi-
neering, and Medicine (NASEM) classifies the bioeconomy into three main visions
as shown in Fig. 2.5.
According to NASEM, the activities of the bioeconomy revolve around improving
scientific knowledge by comprehending mechanisms and processes at the genetic
and molecular levels, as well as synthesizing and manufacturing those commercial
items, such as vaccines, biopesticides, etc. The vision for bioresources entails the
transformation of biomass or biological materials into energy sources or new goods
like bioplastics, biofuels, etc. The bioecology vision takes into account how important
ecological processes are for maximizing the use of energy and nutrients, preserving
and promoting biodiversity, and preventing soil deterioration.

Fig. 2.5 Visions of bioeconomy as proposed by NASEM


2 Bioenergy-Based Sustainable Bioeconomy—Perspectives and Challenges 23

2.4.2 Sectors Involved in Bioeconomy

In accordance with the European Union Bioeconomy Strategy, which was updated
in 2018, the bioeconomy encompasses all industries that rely on biological resources
(including organic waste, animals, plants, microbes, and derived biomass), as well
as its principles and functions. This multidisciplinary topic encompasses the subsec-
tors of biohealth, bioenergy, bioagriculture (agri-bio), and bioindustrial (chemicals
and materials). All subsectors of the bioeconomy share the objective of commercial-
izing the resulting bioproducts, processes, and intellectual property. A few important
sectors are discussed below.

2.4.2.1 Agriculture and Food-Based Bioeconomy

Technology advancements in agriculture have enhanced crop production and farming


efficiency while reducing the environmental impact. Examples of these advancements
include climate monitoring, nutrition management, vertical farming, pest and disease
control, precision and regenerative agriculture, and digital market information.
The notion of “cellular agriculture” is utilized to create agricultural goods from
cell cultures utilizing genetic engineering and synthetic biology methods. Animal
products such as meat, milk, and eggs are produced using cell culture as a step
toward animal welfare, food security, and human health. Spider silk similar plant
protein has been used as an alternative to plastics. Crops have been genetically
engineered to protect against pests and increase the yield. Genomic-assisted breeding
improves bioenergy crops, e.g., in shrub willow, a crop that is gaining popularity as
a source of bioenergy, researchers at Cornell University have discovered genes that
confer resistance to leaf rust. Fungi-based industries are considered a major part of a
growing bioeconomy, e.g., mushroom farming, mycoprotein, etc. Algae cultivation
has potential uses in the production of pharmaceuticals, bioplastics, nutraceuticals
(algal oil), food colorants and dyes, protein-rich animal feed, and algal fuels.

2.4.2.2 Blue/Ocean Economy

The term “blue/ocean economy” refers to the wise use of ocean resources to support
economic viability and the health of the marine environment. According to estimates,
it is currently valued more than US$ 1.5 trillion annually and might double in size by
2030. The ocean offers a wide range of possible blue energy sources, including wave,
thermal, and biomass (like marine algae). Because they provide enormous financial
opportunities and food security, fishing and aquaculture are essential to society.
Through interstate collaboration and trade, the ports and maritime transportation can
boost economic growth. Diverse sorts of minerals are found in the sea bed, and it
is also a source of many innovative substances that could be used to create pharma-
ceuticals, nutraceuticals, and personal care items that would boost the economy. The
24 C. V. Vidhya et al.

production of enzymes, biopolymers, and biomaterials could be accomplished using


items of marine ancestry. Economic growth may benefit from sustainable seaside
tourism. In order to improve livelihood, lower risk from natural catastrophes, and
mitigate climate change, the Indian Ocean Blue Carbon Hub (IORA) aims to develop
knowledge to maintain and restore blue carbon ecosystem throughout the Indian
Ocean.

2.4.2.3 Forest Bioeconomy

The forest bioeconomy is based on natural resources from forests and encompasses
a number of industries and industrial methods that put the idea of the circular
economy into practice. The modern forest industry uses every branch of science
and technology in its effective industrial value chains, including oil management,
drones, satellite imaging, sophisticated harvesting equipment, robust mechanical
engineering, innovative process automation, nanotechnology, and biochemical appli-
cations. The use of woodchips or wood pellets in automated heating system is based
on robust technology. In order to further advance the circular bioeconomy, techniques
utilized for industrial conversion of wood into value-added goods can be refined
and applied to agricultural leftovers like straw. Recreation and nature tourism are
essential components of the bioeconomy, in addition to carbon sequestration and
ecosystem services. Many nations, including Finland, are converting their forest-
based businesses into biorefineries. Forest biomass is utilized in many processes to
make textiles, chemicals, cosmetics, fuel, coatings, food, and animal feed [10].

2.4.2.4 Bioenergy-Based Bioeconomy

The ever-rising global energy crisis, environmental degradation, and climate change
issues due to extensive usage of fossil fuels (FFs) are pushing the world to find an
alternative green energy source to attain Sustainable Development Goals (SDGs) via
decarbonization of the global energy system (GES). Due to the expanding energy
needs of contemporary economies and the physical interactions between energy-
and non-energy-based processes and products, energy can be seen as the “glue” in
the bioeconomy. Bioenergy is described as clean energy generated from renewable
sources that can displace fossil fuels. Wide range of natural resources as shown in
Fig. 2.3 is used to produce bioenergy. Utilizing processing methods like anaerobic
digestion, fermentation, pyrolysis, or torrefaction, the bio-based economy utilizes
first-generation biomass (crops), second-generation biomass (cellulosic crop), third-
generation biomass (algae based), and fourth-generation (genetically modified algae)
to get the most out of biomass.
Regardless of the source, bioenergy is considered as a sustainable source of power
providing heat, gas and fuel. The International Energy Agency (IEA) has projected
that by 2060, bioenergy will supply roughly 17% of global energy demand and can
cut greenhouse gas emissions by 20% overall [11, 12]. The biofuel capacity in India
2 Bioenergy-Based Sustainable Bioeconomy—Perspectives and Challenges 25

is expected to grow from 5.2 billion liters in 2021 to 10.1 billion liters in 2025 and
may generate US$ 20 billion bioeconomy.

2.4.3 Benefits of Bioeconomy

Making the switch to a bioeconomy has several advantages. The bioeconomy can
boost crop and livestock productivity while ensuring food security. It prevents overex-
ploitation as it can offer sustainable management of natural resources and encourages
the use of waste materials and products to create value-added products by reducing
waste and promoting a circular economy. It can result in improved efficiency in the
use of agricultural wastes and biomass. For the production of energy, chemicals, and
other materials, it is conceivable to substitute renewable biomass or bio-based raw
materials for fossil fuels. Adopting bioeconomy strategies can enhance public health
while reducing greenhouse gas emissions. A few bioeconomy techniques, such as
afforestation and deforestation, can help reduce climate change by removing carbon
dioxide from the atmosphere. Rural development can be improved by diversifying
income sources and creating jobs in agriculture, forestry, biotechnology, and related
services. Innovation in areas like genetic engineering, synthetic biology, and biopro-
cessing is sparked by the bioeconomy and results in new goods with diverse uses. The
creation of novel medications, treatments, and medical innovations like personalized
medicine may be facilitated byproducts obtained from the bioeconomy.

2.4.4 Ways to Strengthen the Bioeconomy

The benefits of bioeconomy will not become a reality without the support of local,
national and supranational organizations across the world. Guidelines for the effective
creation and marketing of bioeconomy-related goods and services include promoting
alignment and coherence in policy across a variety of involved sectors; assuring that
all people have access to goods and services relevant to the bioeconomy. Improving
the rigidity of current production systems and stimulating customer demand as well
as acceptance will pave the way to strengthening the bioeconomy.

2.5 Bioeconomy—A Support System of Circular Economy

Bioeconomy and circular economy go hand in hand (Fig. 2.6). Circular economy is
a concept of commercial production and proper utilization of the same which
involves the 6 Rs concept of sustainability that includes, reduce, rethink, refuse,
reuse, repair, and recycle. Circular economy aims to tackle the global challenges
like endangered biodiversity, climatic change, pollution, etc., by the implementation
26 C. V. Vidhya et al.

of the three basic principles—concept of reuse, control pollution, recycle or regen-


erate the natural sources [13]. In a circular economy, the resources are either applied
or utilized in circular fashion rather than linear. Many business models promoting
circularity are those that share platforms, provide products as service, optimize the
utilization of resources, etc. The ultimate aim of these models is to promote circular
economy [14].
The historical approaches of circular economy convey the ideas that would link
biomimicry, cradle-to-cradle design principles and industrial ecology. Biomimicry
mimics time-tested designs that formulates human system pattern. Industrial ecology
acts as the base of circular economy as it is the study of flow of energy and material
in the industrial systems. Cradle-to-cradle design principles are those that device
a holistic approach to create a system that deals with procuring raw material to
waste disposal. The ideology is to decrease waste and pollution and to increase the
efficiency of resources. These intertwined concepts are stepping stones for developing
the circular economy and implementing the same.
William McDonough is named as the father of circular economy as he enlight-
ened the society about circular economy through his popular book namely Cradle
to Cradle: Remaking the Way We Make Things. He also received Fortune Award

Fig. 2.6 Design of bioenergy-based bioeconomy with circular economy


2 Bioenergy-Based Sustainable Bioeconomy—Perspectives and Challenges 27

for Circular Economy Leadership in the 2017, at Davos during the World Economic
Forum.
The Centre for Bioenergy Sustainability (CBES) deals with the environmental
impacts and the sustainability of bioenergy. This is a leading organization under the
Department of Energy Oak Ridge National Laboratory, USA. The ultimate goal of
the organization is to analyze the methods for enhancing the sustainability of current
and potential bioenergy production and distribution (environmental, economic, and
social) so as to serve and support the stakeholders and decision makers with
independent source of data and analysis for bioenergy production [15].
The CBES goals aim at considering bioenergy as an important option from a
system’s perspective; socioeconomic and environmental concerns leading to sustain-
able bioenergy system; appropriate scales for addressing the same; bartering between
land use and land management. It also lays emphasis on quantification of socioe-
conomic and environmental ideas of selecting appropriate bioenergy source (ex:
lignocellulosic feed-stock) and to implement the same at regional to global scale.
As far as the application of bioeconomy in circular economy is concerned, the
achievement of “Net Zero” to “True Zero” carbon emission is on progress. “Net
Zero” carbon emissions achievement is because of consumption of fossil fuel thereby
completely removing the carbon from the atmosphere. Related to this context, Steve
Hoy, a global smart grid technologist, says that he believes the concept of creating a
circular economy, True Zero concept has to be adapted rather than Net Zero concept
wherein fossil fuel consumption is eliminated, entirely, energy will be produced
only from renewable sources [16]. In the current scenario, to maintain these renew-
able systems, the renewable energy industry sector is intending to manufacture a
significant amount of energy and raw.
Due to the emissions attributed to fossil fuel electricity generation, the overall
carbon footprint of renewable energy technologies is significantly lower than for
fossil fuel generation over the respective systems lifespan [16]. The other applica-
tions include agriculture, textile industry, construction industry, automotive industry,
furniture and logistics industry, plastic waste management.

2.6 Indian Bioeconomy Scenario

India has a fantastic chance to rebuild its economy since it uses a variety of feed-
stock to manufacture various bioproducts. The bioeconomy of India surpassed US
$1 billion in 2021, and the country has set an ambitious goal to reach US$ 150 billion
by 2025. Bio-agri, which includes genetically modified (GMO) crops, insecticides,
marine biotechnology, and animal biotechnology, has the potential to treble from
US$ 10.5 billion to US$ 20 billion by 2025. The Indian government has decided
to hasten the launch of ethanol-blenched petrol, enhance biofuel output, and adopt
modifications to the national biofuel policy. The modifications include enabling more
feed-stocks for biofuel production, exporting in certain circumstances, promoting
research for indigenous techniques, and creating more jobs. In the National Policy
28 C. V. Vidhya et al.

on Biofuels-2018 (NPB), a target of 20% ethanol in gasoline and 5% biodiesel in


diesel has been suggested by 2030 [17].
India currently supports specialized bioenergy incubators, infrastructure funding
in the fields of synthetic biology, algal biofuels, biohydrogen, bioenzymes, CO2
capturing and utilization, and biovalorization, to develop value-added products in an
efficient manner. India and the USA are sharing leadership of the Mission Innova-
tion Collaborative platform initiative on “Innovation for Sustainable Aviation Fuel”
(ISAF) as part of Mission Innovation 2.0 [18]. Through the production of biogas
from the processing of solid agricultural waste and livestock manure, the Galva-
nizing Organic Bioagro Resources (GOBAR)-DHAN [19] project aims to enhance
sanitation in Indian village communities. Biomanures from biogas plants have been
used as organic manure under fertilizer category for retail sale. Hydrogen-blended
HCNG has been used as a clean fuel in the mobility sector to reduce pollution. Along
with flagship initiatives like “Make in India,” “Swachh Bharat,” “Doubling Farmers
Income,” etc., the concept of the bioeconomy also needs to have a solid political
foundation.
Bioeconomy policies should be aligned with international initiatives such as Paris
Agreement on climate change, United Nations 2030 Sustainable Agenda or the Aichi
Agreement on biodiversity to protect nature and life systems. It is crucial to empha-
size that while the bioeconomy offers many advantages, it also brings challenges in
terms of preserving biodiversity, effectively exploiting land, competing with food
production, and ensuring that bio-based industries are socially just and sustainable.
A balance between economic expansion and environmental protection is necessary
for the bioeconomy concept to be successfully implemented.

2.7 Challenges of Bioenergy-Based Bioeconomy

Bioenergy, in spite of being Nextgen fuel, still faces challenges in its production and
bioenergy-based bioeconomy. The major concern areas with respect to environmental
impacts of bioenergy production include the water quality and quantity ranking first
(16%) in the year of 2017, emission of greenhouse gases (6%), SOC and biodiversity
(5%), and soil erosion (0.8%). This indicates that the bioenergy production has a
direct and great impact on water resources and its pollution [20].

2.7.1 Water Quality and Quantity

Bioenergy production has its great impact on water quality and its quantity as well. It
is due to the reason that high water consumption is required for such crop plantation.
For an instance, USA with the support of EISA (2007) started a large corn ethanol
production unit to produce potential water stress both local and regional levels [21–
23]. This is due to the reason that when compared with soyabean and wheat, the corn
2 Bioenergy-Based Sustainable Bioeconomy—Perspectives and Challenges 29

plantation requires higher quantities of water and its uses a water level which equals
the amount used by a community of 5000 people. (Service, 2009). Other disadvan-
tages in bioenergy crop cultivation include nutrient pollution due to groundwater
pollution, surface runoff, and usage of more fertilizers.
The bioenergy crop plantation in erodible and marginal areas reduces the stream-
flow and nutrients loss. This was identified by scenario analysis with Soil and Water
Assessment Tool (SWAT) [24]. But these problems can be solved by following certain
concepts like the selection of proper crop species and proper management (irrigation,
appropriate fertilization, harvest rate, etc.) which will build possible balance between
bioenergy production and water resource protection [25, 26].

2.7.2 Greenhouse Gas Emission

The decrease in greenhouse gas (GHG) emissions is also an equally important


concern in the bioenergy production. Carbon dioxide and nitrous oxide are the two
primary components as they are in high quantity. Quantity of net CO2 emissions is
comparatively less when biofuels are used than in the usage of fossil fuels [25, 27].
In a pilot study, Liu et al. quantified that the maximum production of switchgrass on
marginal land would decrease emissions by 29 million tons CO2-eq /year. This was
done by replacing fossil fuel. The end results suggested that for transportation use in
the USA, 40–85% of GHG emissions can be decreased when is replaced by ethanol
on a per megajoule (MJ) energy basis. The efficiency of GHGs reduction always
varies with usage of different feed-stocks [28].
Harris et al. from their study concluded that the conversion of arable land to
the bioenergy crop plantation (second generation) will definitely lead to decreasing
CO2 emissions. Whereas, when native grasslands were converted to plantations of
first-generation bioenergy crops and short rotation coppice (SRC), as well, showed
a remarkable hike in CO2 emissions. Hence, it is indispensable to select an appro-
priate bioenergy crop types as well as its management practices, to mitigate carbon
dioxide emissions [29].
Emission of N2 O is also equally important as it is the second important greenhouse
gas and second prime reason for global warming (298 times that of CO2 ). One major
source of N2 O is agriculture [30]. For an instance, corn cultivation requires high
nitrogen-rich fertilizer, as it is the substrate for denitrification process. This directly
increases the emission of N2 O. Hence, wise selection of the type of bioenergy plant
and planting directly influences in controlling the N2 O emission.

2.7.3 Biodiversity

Bioenergy production indirectly abrades the biodiversity—Landscape choices, land-


scape configuration, land use conversion, and bioenergy production system types.
30 C. V. Vidhya et al.

For example, conversion of normal plantation may interfere with local ecosystem
but conversion of grasslands to bioenergy production plantation will increase local
productiviwty and help maintain the ecosystem functions [31, 32].
Additionally, many research studies laid emphasis on growing Miscanthus,
(comparing with other annual crops) which had reduced negative impact on biodi-
versity. This is due to the reason that cultivation of perennials provides relatively
stable habitats for wildlife [33, 34].

2.7.4 Soil Organic Carbon

Soil organic carbon (SOC) is the index of soil quality. High soil organic carbon
content improves retention of soil water and increases crop productivity and thereby
soil biodiversity. The production of bioenergy influences SOC directly. This is
through three major pathways. They are removal of residues, tillage, and conversion
of land usage.
The decrease in SOC could be controlled, through a proper residue management
like removal of little amount of residues and increased manure applications [26,
35, 36]. The biochar application improves the carbon sink of agricultural sector. It
increases the SOC but it also absorbs the inorganic carbon in the atmosphere thereby
improving the quality of air [37, 38].
Secondly, the loss of SOC is also due to soil disturbances and management prac-
tices. For example, in his study, Drewniak et al. [39] observed that tillage would
always cause SOC loss. Wise choice of most suitable land area, plant types, and the
management practices can prevent SOC [39].

2.7.5 High Cost

The expense that is involved in building up and maintenance of bioenergy plant is


relatively high as other renewable technologies do not need to accounted for. Certain
of bioenergy types of plantations or its management practices are sometimes on par
or still cheaper, than fossil fuels, as the former does not require expensive drilling
and mining processes which involve large capital.

2.7.6 Land Requirement

Socioeconomical and environmental pressure may be created in biomass production


unit as there is increased demand for biomass. The reason is due to low surface power
density of biomass as larger land areas are needed to produce the same amount of
energy, compared to fossil fuels [40].
2 Bioenergy-Based Sustainable Bioeconomy—Perspectives and Challenges 31

Bioenergy plantation requires land just like any other plantation. But if the same is
built exclusively for an industry or company, then it has to be built in close proximity
to cut off storage and transportation costs. If this condition is not possible, high
capital and investment are required for the same: as it is easily possible only for
those industries that are newly being built. Additional land requirement is essential,
and if plantation is for the sole purpose of electricity purpose, then this adds on larger
land footprint/unit of production power supply.

2.7.7 Transportation

Distant transportation of biomass is considered as prodigal and unsustainable. It is


advantageous if a plantation and resource-consuming industry are located nearby as
there will be considerably less expenditures on transport.

2.7.8 Bioenergy-Based Bioeconomy—Nextgen Challenge:


Carbon Management

The bioenergy and bioeconomy has a fundamental challenge. Is it possible for bioen-
ergy and bioenergy-based bioeconomy to feed the demand of the entire carbon-based
value chains? Recently, it was found that the annual consumption of fossil carbon for
oil, polymer industries, and textiles was more than 1 billion tons. To replace aviation
fuel alone may require a many million tons for bioenergy fuel. A part of bioenergy
has to be modified by removing oxygen and moisture so that it can be widely used
in chemical and textile industries. All these will lead not only the hike in demand
for natural resources but also reinstate the sources of bioenergy as primary asset for
international trading.
Like fossil fuel, biomass is also not equally distributed on the earth, and there may
be national discord. If the bioenergy-based bioeconomy turns out to be inefficient
in supplying adequate amount of biomass or only with inappropriate bartering for
other sustainability goals like biodiversity and food security, then it requires add-ons
like artificial photosynthesis, direct air capture, etc. But still science and technology
and industrial sector are working with full efficiency in bringing about the successful
replacement of fossil fuels with bioenergy and hold the nation’s economy status high
with bioenergy-based bioeconomy.
32 C. V. Vidhya et al.

2.8 Practical Approach

The bioenergy is not a new topic neither it is only a topic of theoretical view but
much applicable for practical applications. In Estonia, Finland, and Denmark, appli-
cation of bioenergy is well represented by the production of electricity (> 15%). The
production was mainly by the combination of heat and power—CHP. Next to these
three countries stand UK, Germany, Brazil and Sweden, Germany. Whereas in a few
other countries, it is only 2–5% [41]. The practical approach is evident from the
following companies that either produce bioenergy or use it for their production.
In India, the government of India supports bioenergy and this is evident from the
following statement. Direct supply of biodiesel (B-100) blended along with diesel
was to high-end consumers like Railways and State Road Transport Corporations
and was permitted by the Government of India on August 10, 2015 [42].
One of the leading manufacturers of biodiesel (B-100) in India is Emami Agrotech
Limited (EAL). They manufacture and supply biodiesel (B-100) to many industry
sectors and applications-oriented sectors, in our country. Emami Group has its state-
of-the-art biodiesel plant in Haldia, West Bengal. Its production capacity is 350 tons/
day. It has a state-of-the-art NABL-accredited laboratory at Haldia. This biodiesel
plant has ISO 9001-2008, ISO 14001, OHSAS 18001 and International Sustain-
ability and Carbon Certification (ISCC) Certifications. Biodiesel produced by EAL
conforms to Indian and European Quality Standards IS 15607-2016 and EN 14214
[43].
In the USA, one of the leading biodiesel producers is renewable energy group
(REG). The annual production capacity is 432 million gallons as per January 2023
data. There are eight biodiesel plants under REG. The largest plant is in Washington.
These are the ten leading bioenergy producers in the world by 2021. The compa-
nies were enlisted in Fortune Business Insights, 2021 [44]. ADM is an MNC. It was
established in Chicago in 1902. Amersco Inc. is one of the leading bioenergy assets
developers. It was established in Massachusetts, USA, in 2000. EnviTec Company
is leading biogas producer in Germany. It is a multiservice company that provides
a range of services. Enviva in the USA is the largest producer of coal’s alternate—
viable wood pellets. Drax group in the USA is a pellet manufacturing mill that is
trying out to have sustainable zero carbon level. STARBAG is a pioneer in innovation
and field strength with respect to designing and constructing biogas plant, through its
European technology. Pinnacle Renewable Energy Inc. is one of the leading wood
pellet manufacturers. The production capacity is 2.83 million metric tons/year.
A Canadian Company Enerkem ranks first in the commercial production of
renewable ethanol and methanol from non-recyclable municipal solid waste. Green
Plains Inc. (established in 2007) is a US-based company pioneering in biorefining.
The company has successfully brought down the carbon emission by over 34.3 MMT,
since its establishment. POET, a Scotland-based company, is also one among the
largest biorefinery. The company apart from being a biorefinery is also aiming at
producing alternatives for petrochemicals.
2 Bioenergy-Based Sustainable Bioeconomy—Perspectives and Challenges 33

These are few biomass energy producing and utilizing Companies Premium. Harp
Renewables Limited, California; Granutech-Saturn Systems, USA;, EnviTec Biogas
AG, Germany; Talbotts Biomass Energy Limited, UK; Taylor Biomass Energy, New
York, Green Vinci Biomass Energy Co., Ltd. China; Vattenfall AB, Sweden.
Suzuki Motor Corporation, Toyota Motor Corporation ENEOS Corporation,
Subaru Corporation, Toyota Tsusho Corporation and Daihatsu Motor Co. Ltd.,
collaboratively established the Research Association of Biomass Innovation for Next
Generation Automobile Fuels (Research Association). This association was created
to explore and analyze various possible ways to successfully optimize the production
of best quality biofuel [45].

2.9 Conclusions

In this chapter, the production and applications of bioenergy, the application of bioen-
ergy in raising bioeconomy, the association of bioeconomy and circular economy, the
interrelationship among bioenergy, its resources, bioeconomy and circular economy
and the collaborative effect of the same, on solving the mitigating environmental
changes are explained. In general, bioenergy expansion fell by 50% in 2020. Only
Europe and China reported significant expansion in 2020, adding 2 GW and 1.2 GW
of bioenergy capacity, respectively. Under Net Zero Scenario, total global bioenergy
usage, in 2030, is only going to be about 20% higher than in 2021. In 2021, about
35% of the bioenergy was from biomass, which was utilized for traditional prac-
tices that were inefficient, polluting and unsustainable. In the Net Zero Scenario, the
usage of this traditional biomass may plummet to zero by 2030 in order to achieve the
UN Sustainable Development Goal 7 on Affordable and Clean Energy. The usage
of modern bioenergy, which does not include traditional uses of biomass, nearly
doubles up from about 42 EJ in 2021 to 80 EJ in 2030. There will be escalation in the
production of bioenergy so as to cater the estimation by Net Zero Scenario in 2030.
Following scenarios are expected by the year 2030: In generation of electricity, the
use of bioenergy will nearly double up. 750 TWh of electricity (about 2.5% of total
demand) in 2021 will nearly reach about 1350 TWh (about 3.5% of total demand)
in 2030. The usage of biojet kerosene as aviation fuel was nearly zero in 2021 and
will increase in 2030 accounting for over 7% of all aviation fuel demand. In road
transport, the consumption of liquid biofuel will quadruple from 2.1 mboe/d in 2021
to over 8 mboe/d in 2030.
The usage of bioenergy in industrial sectors (mainly paper, cement, and light)
was moderately above 11 EJ of energy in 2021 and may increase considerably to
more than 17 EJ in 2030. Bioenergy with carbon capture and storage (BECCS)
plays a critical role in bioenergy management. It will capture and store bioenergy
emissions that are already carbon–neutral and create negative emissions. In 2021,
BECCS captured and stored 2 Mt of CO2 . There will be a hike to around 250 Mt of
CO2 in 2030, counterbalancing the emissions from the sectors where abatement will
not be an easy condition.
34 C. V. Vidhya et al.

Hence to head toward Sustainable Development Goals (SDGs), bioenergy produc-


tion and usage has to profoundly increase to meet up with a wide variety of its imple-
mentations, by 2030, so as to get on track with the Net Zero Scenario, according to
UN Sustainable Development Goal 7.

References

1. Vidhya, C. V. (2022). Microalgae—The ideal source of biofuel. In Biofuels and bioenergy


opportunities and challenges. Elsevier Publications. 978-0-323-85269-2.
2. Babin, A., Vaneeckhaute, C., & Iliuta, M. C. (2021). Potential and challenges of bioenergy
with carbon capture and storage as a carbon-negative energy source: A review. Biomass and
Bioenergy, 146, 105968.
3. Kumar, R., & Kumar, P. (2017). Future microbial applications for bioenergy production; a
perspective. Frontiers in Microbiology. https://doi.org/10.3389/fmicb.2017.00450
4. Albrecht, J., Carrez, D., Cunningham, P., Daroda, L., Mancia, R., Máthé, L., Raschka, A.,
Carus, M., & Piotrowski, S. (2010). The knowledge based bio-economy (KBBE) in Europe:
Achievements and challenges.
5. McCormick, K., & Kautto, N. (2013). The bioeconomy in Europe: An overview. Sustainability.,
5(6), 2589–2608.
6. Food and Agriculture Organization of the United Nations Rome. (2016). How sustainability is
addressed in official bioeconomy strategies at international, national and regional levels? An
overview.
7. Food and Agriculture Organization of the United Nations. (2022). Document card. www.
fao.org. Retrieved 16 September 2022.
8. United states Environmental Protection Agency. (2007). Summary of the Energy Independence
and Security Act.
9. Van Dam, J., & Junginger, M. (2011). Striving to further harmonization of sustainability criteria
for bioenergy in Europe: Recommendations from a stakeholder questionnaire. Author links
open overlay panel Energy Policy, 39(7), 4051–4066.
10. Arasto, A., Koljonen, T., & Similä, L. (2018). Growth by integrating bioeconomy and low-
carbon economy: scenarios for Finland until 2050. [Espoo]. ISBN 978-951-38-8699-8. OCLC
1035157127.
11. Organization for Economic Cooperation and Development. (2009). The Bioeconomy to 2030:
Designing a policy agenda (p. 3). OECD Publishing.
12. Scarlet, N., Dallemand, J. F., Monforti-Ferrario, F., & Nita, V. (2015). The role of biomass and
bioenergy in a future bioeconomy: Policies and facts. Environment and Behaviour, 15, 3–34.
13. European Parliament News. (2021).
14. Urbinati, A., Chiaroni, D., & Chiesa, V. (2017). Towards a new taxonomy of circular economy
business models. Journal of Cleaner Production., 168, 487–498.
15. Chung, J. N. (2013). Grand challenges in bioenergy and biofuel research: engineering and
technology development, environmental impact, and sustainability. Front Energy Research:
Bioenergy and Biofuels, 1.
16. Gallagher, J., Basu, B., Browne, M., Kenna, A., McCormack, S., Pilla, F., & Styles, D. (2019).
Adapting stand-alone renewable energy technologies for the circular economy through eco-
design and recycling. Journal of Industrial Ecology., 23(1), 133–140.
17. Das, S. (2020). The National policy of biofuels of India—A perspective. Energy Policy, 143.
18. Fuelling India’s future with bioenergy. (2023). https://www.pwc.in/research-and-insights-hub/
fuelling-indias-future-with-bioenergy.html
19. Galvanizing Organic Bio-Agro Resources Dhan (GOBARdhan). (2023). https://www.mys
cheme.gov.in/schemes/gobardhan
2 Bioenergy-Based Sustainable Bioeconomy—Perspectives and Challenges 35

20. Wu, Y., Zhao, F., Liu, S., Wang, L., Qiu, L., Alexandrov, G., & Jothiprakash, V. (2015). Bioen-
ergy production and environmental impacts. Official Journal of the Asia Oceania Geosciences
Society, 5 (Article number: 14).
21. Gasparatos, A., Stromberg, P., & Takeuchi, K. (2011). Biofuels, ecosystem services and human
wellbeing: Putting biofuels in the ecosystem services narrative. Agriculture, Ecosystems &
Environment, 142(3–4), 111–128.
22. Hoekman, S. K., Broch, A., & Liu, X. (2018). Environmental implications of higher ethanol
production and use in the US: A literature review. Part I—Impacts on water, soil, and air quality.
Renewable and Sustainable Energy Reviews, 81, 3140–3158.
23. Zhou, X., Clark, C. D., Nair, S. S., Hawkins, S. A., & Lambert, D. M. (2015). Environmental
and economic analysis of using SWAT to simulate the effects of switchgrass production on
water quality in an impaired watershed. Agricultural Water Management, 160, 1–13.
24. Guo, T., Cibin, R., Chaubey, I., Gitau, M., Arnold, J. G., Srinivasan, R., Kiniry, J. R., & Engel,
B. A. (2018). Evaluation of bioenergy crop growth and the impacts of bioenergy crops on
streamflow, tile drain flow and nutrient losses in an extensively tile-drained watershed using
SWAT. Science of the Total Environment, 613–614, 724–735.
25. Qin, Z., Zhuang, Q., Cai, X., He, Y., Huang, Y., Jiang, D., Lin, E., Liu, Y., Tang, Y., & Wang,
M. Q. (2018). Biomass and biofuels in China: Toward bioenergy resource potentials and their
impacts on the environment. Renewable and Sustainable Energy Reviews, 82, 2387–2400.
26. Wu, Y., Liu, S., Young, C. J., Dahal, D., Sohl, T. L., & Davis, B. (2015). Projection of corn
production and Stover-harvesting impacts on soil organic carbon dynamics in the US temperate
prairies. Science and Reports, 5, 10830.
27. Dunn, J. B., Mueller, S., Kwon, H.-y, & Wang, M. Q. (2013). Land-use change and greenhouse
gas emissions from corn and cellulosic ethanol. Biotechnology for Biofuels, 6, 51.
28. Liu, T., Huffman, T., Kulshreshtha, S., McConkey, B., Du, Y., Green, M., Liu, J., Shang,
J., & Geng, X. (2017). Bioenergy production on marginal land in Canada: Potential, economic
feasibility, and greenhouse gas emissions impacts. Applied Energy, 205, 477–485.
29. Harris, Z. M., Spake, R., & Taylor, G. (2015). Land use change to bioenergy: A meta-analysis
of soil carbon and GHG emissions. Biomass and Bioenergy, 82, 27–39.
30. Williams, A. G., Audsley, E., & Sandars, D. L. (2010). Environmental burdens of producing
bread wheat, oilseed rape and potatoes in England and Wales using simulation and system
modelling. International Journal of Life Cycle Assessment, 15(8), 855–868.
31. Correa, D. F., Beyer, H. L., Possingham, H. P., Thomas-Hall, S. R., & Schenk, P. M.
(2017). Biodiversity impacts of bioenergy production: Microalgae vs. first generation biofuels.
Renewable and Sustainable Energy Reviews, 74, 1131–1146.
32. Sang, T., & Zhu, W. (2011). China’s bioenergy potential. GCB Bioenergy., 3(2), 79–90.
33. Rowe, R. L., Street, N. R., & Taylor, G. (2009). Identifying potential environmental impacts of
large-scale deployment of dedicated bioenergy crops in the UK. Renewable and Sustainable
Energy Reviews, 13(1), 271–290.
34. Werling, B. P., Dickson, T. L., Rufus, I., Hannah, G., Claudio, G., Gross, K. L., Heidi, L.,
Malmstrom, C. M., Meehan, T. D., Ruan, L., Roberston, B. A., Roberston, G. P., Schmidt, T.
M., Schrotenboer, A. C., Teal, T. K., Wilson, J. K., & Landis, D. A. (2013). Perennial grasslands
enhance biodiversity and multiple ecosystem services in bioenergy landscapes. PNAS, 111(4),
1652–1657.
35. Sheehan, J. J., Adler, P. R., Del Grosso, S. J., Easter, M., Parton, W., Paustian, K., & Williams,
S. (2014). CO2 emissions from crop residue-derived biofuels. Nature Climate Change, 4(11),
932–933.
36. Robertson, G. P., Grace, P. R., Izaurralde, R. C., Parton, W. P., & Zhang, X. (2014). CO2
emissions from crop residue-derived biofuels. Nature Climate Change, 4(11), 933–934.
37. Li, W., Dang, Q., Brown, R. C., Laird, D., & Wright, M. M. (2017). The impacts of biomass
properties on pyrolysis yields, economic and environmental performance of the pyrolysis-
bioenergy-biochar platform to carbon negative energy. Bioresource Technology, 241, 959–968.
38. Pourhashem, G., Rasool, Q. Z., Zhang, R., Medlock, K. B., Cohan, D. S., & Masiello,
C. A. (2017). Valuing the air quality effects of biochar reductions on soil NO emissions.
Environmental Science and Technology, 51(17), 9856–9863.
36 C. V. Vidhya et al.

39. Drewniak, B. A., Mishra, U., Song, J., Prell, J., & Kotamarthi, V. R. (2015). Modelling the
impact of agricultural land use and management on US carbon budgets. Biogeosciences, 12(7),
2119–2129.
40. Daley, J. (2018). The EPA declared that burning wood is carbon neutral. It’s actually a lot more
complicated. Smithsonian Magazine. Archived from the original on 30 June 2021. Retrieved
14 September 2021.
41. International Energy Agency. (2021). World energy outlook.
42. Ministry of Petroleum and Natural Gas, Government of India. (2015). Annual report.
43. Emami Agrotech—Emami Group. https://www.emamigroup.com/group-of-companies/agr
otech/biodiesel
44. Fortune Business Insights. (2021). https://www.fortunebusinessinsights.com/blog/10-leading-
bioenergy-producers-in-the-world-10598
45. Green Car Congress—Energy Technologies Issues and Policies for Sustainable Mobility.
(2022). Six Japan-based companies established research association of biomass innovation
for next generation automobile fuels.
Chapter 3
Lignocellulosic Biomass for Sustainable
Production of Renewable Fuels:
Embracing Natural Resources

Medha Maitra, S. Sruthi, Pavada Madhusudan Rao, V. S. Avanthi,


and P. Radha

Abstract The green lignocellulosic biomass industry is now gaining recognition


as an appealing substitute for fossil resources, seeking to balance the world’s esca-
lating dependency on petroleum while being economical. The resurgence of the
most abundant lignocellulosic resources into effective goods is an intriguing option
for diminishing greenhouse gas emissions. Aromatic lignocellulose is a complex
carbohydrate polymer composed of polysaccharides derived from xylose, glucose,
and lignin. The interior framework of lignocellulosic matter is responsible for plant
cell walls’ hydrolytic resilience and rigidity, rendering them immune to microbial
destruction. Recycling of residues leads to less dumping of waste. The chapter mainly
concentrates on commercializing lignocellulosic waste biomass in the contemporary
period to develop alternative biofuels, which encompasses diverse processes such
as prior treatment, saccharification, and oleaginous microbe-mediated fermentation.
Different means of prior treatment are implemented to conquer the inflexibility of
LCB and advance its breakdown into hemicellulose, lignin, and cellulose. Compo-
nents such as bioethanol, biodiesel, and biobutanol from the lignocellulosic feedstock
act as inhibitors of detoxification and play major roles in strain enhancement, process
integration, and optimization, which are the main fundamentals for boosting biofuel
synthesis. The re-engineering of lignocellulosic material to extract fuel alternatives
has tremendous potential because of its enticing features, comprising low density,
affordability, green credentials, and biodegradability.

Keywords Lignocellulosic biomass · Biofuels · Pretreatment · Bioethanol ·


Biobutanol · Biodiesel

M. Maitra · S. Sruthi · P. M. Rao · V. S. Avanthi · P. Radha (B)


Bioprocess and Bioseparation Laboratory, Department of Biotechnology, School of
Bioengineering, College of Engineering and Technology, SRM Institute of Science and
Technology, Kattankulathur, Chengalpattu District, Tamil Nadu 603203, India
e-mail: radhap@srmist.edu.in

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2024 37
G. Baskar et al. (eds.), Circular Bioeconomy Perspectives in Sustainable Bioenergy
Production, Energy, Environment, and Sustainability,
https://doi.org/10.1007/978-981-97-2523-6_3
38 M. Maitra et al.

Abbreviations

LCB Lignocellulosic biomass


LC Lignocellulose
LCS Lignocellulosic substrate
PHB Poly hydroxybutyrate
CBP Consolidated bioprocessing
DESs Deep eutectic solvents
IL Ionic liquid
LHW Liquid hot water
ABE Acetone-butanol-ethanol synthesis
BTL Biomass to liquid
OMs Oleaginous microorganisms
DMC Direct microbial conversion
PCR Polymerase chain reaction
SHF Separate hydrolysis and fermentation
SSF Separate saccharification and fermentation
CRISPR-Cas Clustered regularly interspaced short palindromic repeats
HBA Hydrogen bond acceptor
HBD Hydrogen bond donor
DNA Deoxyribose nucleic acid
HMF Hydroxymethylfurfural
EG Endoglucanase
AFEX Ammonia fiber expansion
PSA Pressure swing adsorption
CHP Combined heat and pressure
MESP Minimum selling price
LCEE Life cycle energy efficiency

3.1 Introduction

The lignocellulosic substance is vital for feedstock utilized by several technological


advances and a source of eco-friendly green energy in the upcoming days. Although
the exploitation of fossil fuels has been extremely beneficial to the advancement
of humankind, it has also ended up with several complications, including adverse
ecological effects and a fuel shortage. This decade has seen an unsurpassed boom in
the global economy and industrial revolution, an accomplishment that is primarily
defined by the implementation of various energy assets. In addition, due to extraordi-
nary human population expansion, the reliance on fossil fuel resources has risen [1].
According to the United Nations, the worldwide inhabitants will continue to evolve
rapidly, exceeding 8.5 billion in 2030, especially 9.7 billion by the year 2050. Because
of these reasons, the severe exploitation of petroleum-based non-renewable resources
3 Lignocellulosic Biomass for Sustainable Production of Renewable … 39

has resulted in acute scarcity, unanticipated price increases, and global greenhouse
gas emissions. As a result, international governments and academics have set the
objective of discovering and producing sustainable and renewable fuels [1]. As a
result, administrations across the nation are quickening initiatives toward improving
the task and commercialization of lignocellulose as a renewable resource, intending
to address concerns such as energy shortages, environmental impact, and equitable
growth. On average, the sheer quantity of LCB manufactured through photosynthesis
on Earth surpasses 100 billion metric tons. However, only a tiny percentage has a
job for human endeavors, leading to a substantial wastage of assets. The quantity of
LCB capable of being harvested is anticipated in the USA each year, with the utilized
amount being over 1.3 billion metric tons. However, China has roughly 800 million
metric tons.
The consumption of lignin as a material for processing in the manufacturing of
bioethanol as a replacement bioenergy for fossil fuels is an indispensable aspect of
averting the hazards that accompany the combustion of fossil fuels [2]. Lignocellu-
losic resources are feedstock that could potentially be exploited to get regenerative
biofuels. They are composed of vegetative biomass and are ubiquitous in nature. They
consist mostly of the following substances: cellulose, lignin, tanning agents, saponin,
and hemicellulose from agricultural, forestry, or wood debris. Cellulose comprises
the entirety of the polysaccharide framework of the outermost layer of plant cells and
accounts for between 30 and 50% of the total weight in the dry form of lignocellu-
lose and is composed of an ordered sequence of 14 corresponding D-glucose units.
A digestive enzyme, cellulose, may digest to yield glucose. Hemicellulose, unlike
cellulose, incorporates an extensive list of heteropolymers, notably glucomannan,
glucuronoxylan, xyloglucan, xylan, and arabinoxylan [2].
The key elements of the feedstock for lignocellulosic plants are hemicelluloses
and cellulose (about most of a plant’s dry volume), which break down into glucose
via thermochemical and biological processes. There are three kinds of lignocellu-
losic feedstocks: herbaceous and woody energy crops, waste from agriculture, and
forest leftovers. Cellulose in the cell barrier should dissipate and release glucose,
resulting in can later be devoured into ethanol or another substance like biodiesel
or butanol. It should be emphasized, however, that the plant’s cell exteriors hinder
cellulosic ethanol generation due to their structural characteristics. The conversion of
feedstock is a vital stage. It predominantly comprises two treatment approaches for
synthesizing biofuel via lignocellulosic components: the biochemical and thermo-
chemical paths [3]. The enzymatic breakdown of lignocellulosic materials is often
executed via the following phases: lignin elimination, which provides availability of
both hemicellulose and cellulose; degradation of polymers composed of carbohy-
drates to collect unconstrained sugar; and brewing of the liberated sugars to ethanol
from biomass. Every year, an enormous quantity of lignocellulosic residue created
by harvesting and food service businesses can be employed as a source for the next
generation [2].
Bioenergy cultivars ought to be cultivated on the outskirts, adopting advances
in genomics along with reproductive technology breakthroughs to offer broadened
nutrients, especially starches (which are subsequently converted into energy). These
40 M. Maitra et al.

cultivars developed on vacant land will ultimately result in the manufacturing of


environmentally conscious fuels. An extensive selection of plant-based feed fluc-
tuates in structural and substance composition, leading to the industrial production
of a vast range of biofuels [3]. After considering where they originate, biofuels
can be distinguished as first-generation and fourth-generation biofuels. Second-
generation biofuel is an increasingly proficient biofuel created from diversified non-
food biomass of plant or animal provenance. Lignocellulosic biological fuel is made
by using unpalatable vegetation or plant components. Non-edible lignin-rich biomass
(such as vegetable weeds, forestry wastes, crop residue) appears widespread among
living things and might be exploited to execute just like an unprocessed component
during the synthesis of bioenergy. Lignocellulosic ethanol, butanol, amalgamated
alcohols, and additional biofuels of the second kind are the latest possibilities to over-
come the scarcity [4]. Basis substances for second-generation lignocellulose-based
energy tend to comprise biomass unsuitable for human dietary use, so it has nothing
to do with agronomy. The biomass industry from leftovers from farming, not alimen-
tary agricultural products, jatropha, timber scraps, and even microorganisms are all
viable components for biofuels of a subsequent generation. One of the most prevalent
methodologies for offering sustainable aviation energy derived from natural sources
is the conversion of LCB to ethanol. Due to the built-in susceptibility of lignin deriva-
tives to decay by bacteria alongside their substantial thermal needs, distilling the
combined production of ethanol and electricity corresponds to a modest general effi-
ciency of 50–55% [5]. The efficient utilization of lignocellulose as a bioenergy source
could encounter various setbacks, encompassing the difficulty of evolving an inex-
pensive method enabling cellulosic breakdown and sugar recovery. Depending on
the biological component of the lignocellulose substrate, this needs certain pretreat-
ment workflows. This type of pretreatment can boost production costs. Aside from
that, the employment of different chemicals in lignocellulose pretreatment produces
waste products containing chemicals or substances with inhibitory properties, which
could diminish the efficacy of fermentation and hydrolysis methodologies [2].
On the contrary, advancing lignocellulosic biomass (LCB) transformation to
refined chemicals and polymers remains an urgent hurdle. Lignocellulose has evolved
to be less susceptible to depletion. Because of this intrinsic trait, lignocellulosic
materials are resilient to enzymatic and chemical synthesis. The preliminary treat-
ment of LCB, which is an exorbitant approach when considering both expense and
vitality, is required to change the chemical and physical attributes of the lignocellu-
losic framework [4]. The fundamental variations in plant cell walls and deviations in
ultrastructure extensively affect the biomass’s preprocessing and hydrolysis (dissolu-
tion) rate. To circumvent the problems associated with chemical pretreatment, enzy-
matic pretreatment of lignocellulose is thought to be more efficient and produce less
chemical waste. Conventional hydrolysis approaches are costly, labor-intensive, and
not so eco-friendly. More treatments (pretreatment, neutralization, etc.) are recom-
mended, and liberated carbs break down under difficult hydrolysis conditions [6].
Although enzymatic pretreatment is selective, accurate, and environmentally benign,
it is costly to sustain, posing difficulty in impoverished developing nations that lack
the financial means to support or maintain the technology. Because the enzymes are
3 Lignocellulosic Biomass for Sustainable Production of Renewable … 41

temperature-sensitive, the process parameters must be carefully monitored. Various


microorganism strains have been examined for their ability and uniqueness to operate
as viable sources of lignocellulolytic enzymes [2]. These biomass are diverse and
convoluted. Its solid structures render it difficult to transform them into an assort-
ment of valuable goods. Since biomass has a lower energy density than typical
hydrocarbon fuels, it needs more biomass feedstocks to produce the same quantity
of energy. Ample moisture might make methods of conversion unsuccessful. Because
of their high humidity, the fuels are challenging to combust. In accordance with the
biomass species, these biopolymers’ topologies and/or degree of polymerization can
change [6].
So, LCB sources have been regarded as exciting fuel reserves not solely because
they are sustainable but also because they generate a net reduction in green-
house gases released into the surroundings. Lignocellulose, among the most preva-
lent green energy organic substances, constitutes almost half of the plant matter
generated by photosynthesis. This type of power resource can be summarized
as using discarded feed as an environmentally friendly power source. Bioenergy,
frequently called biomass energy, is any viable power supply deriving from non-
fossil substances. However, bioenergy can alleviate the worldwide financial system’s
concerns regarding energy safety due to its insistence on minimizing fossil fuel
availability [7]. Different biomass feedstock sources differ in content, homogeneous
shape, size, etc., which enormously affects the conversion processes. As it turns out,
the standardization of biomass feedstocks has become necessary for biorefineries.
Biomass feedstocks must be standardized before being brought to the facility for
a biorefinery to efficiently convert a given product. The technical aspects of the
chosen distinct lignin biomass conversion approaches, which comprise biochem-
ical, chemical, and biological pathways and numerous pretreatments to create liquid
fuels, have been highlighted in the present investigation. Densified organic matter
and LCB are the most beneficial choices for addressing handling, storage, and mois-
ture issues. The biological transformation treatments of different biofuels, which
comprise bioethanol, biobutanol, PHB esters, biomethanol fermentation, and acces-
sible biomass pretreatment, have been analyzed, as have the techno-economic assess-
ments and different hurdles in utilizing LCB. In broad terms, the overarching objec-
tive of this endeavor is to provide divergent techniques to generate liquid fuels using
lignocellulosic waste and to discuss the potential constraints and alternatives to fossil
fuels.

3.2 Exploring Woody and Non-woody Biomass


as a Prominent Lignocellulosic Resource

Lignocellulosic materials, such as horticulture waste, microalgae, forestry, and


metropolitan solid wastes, are significant, inexpensive, and widely accessible. These
are some of the nutrients from biomass used to generate fuels and chemical-based
42 M. Maitra et al.

goods. In the case of agro-residue, three general categories of lignocellulosic biomass


are virgin biomass, discarded biomass, and renewable energy crops in India. The
plants are included in the virgin feedstock. Waste biomass is created as a poor-
quality byproduct of several distinct sectors of industry, especially forestry (sawmill
and paper mill discards) and agricultural (corn stover, sugarcane bagasse, straw,
etc.). Biomass comprises forest residues like pine bagasse, agricultural wastes like
cereal straw, and short rotation coppices with specialized cultivars like switchgrass,
miscanthus, and poplar. Significant quantities of LCB can be obtained by energy
crops, such as switchgrass (Panicum virgatum) and elephant grass, which are used as
a foundational component for developing second-generation biofuel. These sources
of sustainable energy help reduce reliance on fuels such as coal, petroleum, green-
house gas emissions, and, therefore, environmental pollution. Pretreatment, hydrol-
ysis, fermentation, and product recovery are numerous steps required to convert LCB
into biofuels and bioproducts.
Several obstacles must be addressed to successfully convert into valuable commer-
cial goods while minimizing the generation of byproducts [8]. Lucerne, Miscanthus,
Napier grass, giant reed, and canary grass are some herbaceous plants considered
important sources of LCB, of which common grasses are regarded as the major
source. They are thought to be economically feasible and environmentally friendly
over food crops for ethanol production in addition to avoiding the “food versus fuel”
dilemma. The primary source of cellulosic biomass is the forest, where plants grow
quickly and rotate quickly. Despite their uneven distribution, forests are essential
for preventing landslides, limiting CO2 entering our atmosphere, and maintaining
the balance between people and wildlife. In the meantime, the USA produces over
370 million tons of LCB from its forests each year; other nations with abundant
forests include Canada, Russia, China, and Brazil. As mentioned earlier, half of the
world’s total LCB has been gathered in the nations, including sources such as timber,
branches, wood shavings, and wastes from dead trees like sawdust, chips, etc. Non-
woody biomass refers to biomasses with lower lignin content, like rice husks, grass
straw (wastes from fields), and sugarcane oil wastes (processing wastes). They are
large and have a low energy content [9].

3.3 Structural Composition of Lignocellulose

A typical LCB structure, predominantly composed of cellulose, hemicellulose, and


lignin, is represented in Fig. 3.1. The nature and kind of plant feed, like agricultural,
processed, or forestry, determine efficient feedstock used to produce nanocellulose
by selecting appropriate compositions [10]. The polymers cellulose, hemicellulose,
and lignin combine to form microfibrils, giving stability to the cell wall, which is
further compacted to form microfibrils [11]. Cellulose is the main element, which
consists of 1,4-linked glucose. The hydrogen bonds prevent degradation between
the polysaccharide chains. Hemicellulose is the second main element, comprising
five and six-carbon polymers like mannose, glucose, arabinose, xylose, etc. Lignin is
3 Lignocellulosic Biomass for Sustainable Production of Renewable … 43

made up of chemicals composed of sinapyl alcohol (S), p-coumaric alcohol (H), and
coniferyl alcohol (G) [12]. These polymers are polymerized in different compositions
and ratios based on plant and tissue type, cell wall, wood quality, and layers. Cellulose
has a polymerization degree of about 10,000, consisting of glucose molecules joined
by water-insoluble 1,4-glycosidic linkage. Each glucose unit contains adequate
hydroxyl groups to generate both intra- and inter-chain hydrogen bonds, which
stabilize the molecular structure and combine, forming microfibrils. The structure
of cellulose is made up of several hydrogen bonds, which are inter- and intramolec-
ular, forming a complex three-dimensional arrangement and linking crystalline and
non-crystalline areas. Crystallinity denotes the percentage of crystalline regions in
the polymer cellulose, whose structure is broken down into an amorphous form to
boost its enzymatic hydrolysis conversion.
Hemicellulose is mostly made up of two or more glycans, like xylose, arabinose,
mannose, and galactose, which are non-homogeneous. Hemicellulose is bonded by
H bonds, whereas Van der Waals forces are used to link cellulose and lignin. Hemi-
cellulose can have a variety of structures depending on the plant: typically, hexosan
and pentosan predominate in coniferous and broad-leaved wood, respectively. Hemi-
cellulose varies greatly in its concentration and composition, even within the same
plant, especially in the bark, stems, roots, and branches. Hemicellulose is more
soluble and vulnerable to chemical assault than cellulose because it has a lower
molecular weight and a larger amorphous area. Lignin is the only natural, renew-
able, and aromatic polymer found in plants’ secondary cell walls, which are made up

Fig. 3.1 Structure of lignocellulosic feedstock


44 M. Maitra et al.

of hydroxycinnamic alcohol, cetearyl alcohol, and mustard alcohol. Syringyl struc-


ture (S), p-hydroxyphenyl structure (H), and mustard alcohol are the three lignin
units. These three structural units constitute the majority of lignin polymers and are
joined together by ether and carbon–carbon bonds, such as β-O-4 β-β, β-5, as well as
α-O-4. The quantity of lignin and its structural makeup differs substantially between
different types of lignocelluloses; coniferous wood only has G structural units and
contains 25–31% lignin. Broadleaf wood, containing G and S units, on the other
hand, has 16–24% lignin. The lignin content of herbaceous species ranges from 16
to 21%; it contains all three structural units. Thus, lignin is divided into three types:
G-type (coniferous lignin), GSH-type (graminaceous lignin), and GS-type (broadleaf
lignin), based on structural unit contents [13].

3.4 Physicochemical Properties of Lignocellulose

LCB has some physical properties, including its ability to absorb moisture, compact-
ness, flow behavior, heat properties, and particle size. It contains the fixed carbon of
biomass samples, ash, moisture, and volatile matter [14]. The amount of moisture
present in biomass is referred to as moisture content. It has a significant impact on
the harvest and preparation, in addition to the transportation, storage, processing, and
final goods. The biomass specimen, when entirely burned, results in the formation
of solid residue. The amount of solid residue left over is measured by the amount of
ash. The essential constituents of biomass residue contain calcium, iron, potassium,
magnesium, silica, titanium, and aluminum. The condensable vapor and stable gases
that biomass releases during heating are considered volatile. It is influenced by the
heating limitations, such as residence time, temperature, and heating rate. Following
the heating of the biomass, the labile stuff is liberated, and the fixed carbon is the
solid combustible residue left behind. Optimum analysis determines the concentra-
tion of nitrogen, carbon, oxygen, hydrogen, and sulfur. Elemental analyzers are used
to burn a weighted specimen of biomass within a managed atmosphere and analyze
the gaseous byproducts. Energy content is often measured as the heat of combustion,
which is the total energy released as heat when it goes through an absolute burning
in the presence of oxygen within controlled conditions. It contains the quantity of
energy present inside the particular specimen of biomass. The lignocellulosic feed-
stocks can be taken based on their different physical–chemical properties: structural
composition, moisture content, electrical conductivity, pH, nutrient content, specific
size, and various shapes of the raw materials.
3 Lignocellulosic Biomass for Sustainable Production of Renewable … 45

3.5 Different Pretreatment Approaches

The crucial first stage in the breakdown of LCB’s three components and their conse-
quent high-efficiency conversion to bioenergy is pretreatment. Factors influencing
pretreatment include energy use and the country’s economy. The pretreatment step is
another important factor for determining the processes’ industrial feasibility. Further-
more, the pretreatment process must not alter the components of LCB’s natural struc-
ture. For example, the natural structure of lignin should be altered during delignifica-
tion, rendering it susceptible to valorization. Additionally, when the crystallinity of
cellulose reduces during pretreatment, it is less accessible to the enzymes. It is crucial
to maintain the various components of a successful LCB deconstruction to ensure the
efficacy of a pretreatment procedure. The efficiency of pretreatment is determined
by its capacity to (1) delignify the LCB without significantly changing the native
structure of lignin; (2) less energy utilization; (3) operate at a low cost; (4) lower the
cellulose’s crystallinity index; (5) reduce the LCB’s particle size in order to increase
its surface area to improve enzymatic hydrolysis; (6) avoid the production of enzyme
inhibitors (7) pretreat several LCB feedstocks; and these pretreatment methods can
be of three categories: physical, chemical, and physicochemical (Fig. 3.2) [15].

Fig. 3.2 Methods of pretreatment of lignocellulose


46 M. Maitra et al.

3.5.1 Physical Pretreatment Methods

Prior to using additional pretreatment techniques, LCB must first undergo phys-
ical pretreatment. It is primarily done to reduce the size of particles by increasing
surface area and decreasing polymerization and crystallinity levels. As a result, future
processes are improved and made simpler. These processes are environmentally bene-
ficial and rarely result in the production of harmful materials. The significant energy
consumption of physical pretreatment is one of its main drawbacks. The type of LCB
being used typically affects energy consumption [16].

3.5.1.1 Milling

A common pretreatment approach for diminishing biomass particle size is milling.


Milling can enhance enzymatic degradation since it reduces the amount of crys-
tallinity and the size of the material. By milling and grinding, particles can be reduced
in size by up to 0.2 mm. To an extent, biomass particle size can also be reduced;
beyond that point, the pretreatment process is unaffected by the reduction in particle
size. A stove for corn, with a little maize stover with small particles ranging from 53
to 75 m, is more effective than a maize stover with broad particles of size between
475 and 710 m. This variation in particle size demonstrates how productivity greatly
impacts pretreatment. Ball milling significantly reduces the crystallinity index from
4.9 to 74.2%, making it suitable for saccharifying straws in mild hydrolytic condi-
tions to produce highly fermentable sugars. Enzymatic hydrolysis can be combined
with milling to improve hydrolysis results, kinetic energy, and mass transfer. Unlike
biomass pretreatment, which is done without milling, the ball beads number in the
reactor plays a key role in cellulose hydrolysis due to its less enzyme loading require-
ment and ability to obtain a complete 100% hydrolysis rate. When a fluidized bed,
as opposed to a jet mill, is loaded with rice straw for grinding finely, after steam
explosion and pulverization, the maximum hydrolysis rate with the highest output
of reducing sugar is produced. Ball milling is an expensive alternative for biomass
preparation, consuming high energy, which is a major limitation. Furthermore, lignin
reduces enzyme availability to the substrate, and grinding is less desirable because it
cannot remove lignin. Crystallinity decline, biomass type, the process’s duration, the
method and type of pretreatment used, etc., can all affect polymerization and surface
area growth. A method used to bring down and reduce crystallinity and digestibility
is vibratory ball milling. Wet disk milling generates fibers that promote cellulose
hydrolysis and uses less energy than hammer milling, which results in finer bundles
of material. Because of this, wet disk milling is favored when milling is the option
[17]. The size of 0.4 mm and lesser biomass particles have no effect or impact on
the hydrolysis rate. Based on the kind of powered machinery, vibratory milling,
two-roll milling, hammer milling, ball milling, colloid milling, etc., are some of
the numerous milling techniques used. The type of milling technique utilized, the
amount of time processed, and biomass type all affect the decrease in the size of
3 Lignocellulosic Biomass for Sustainable Production of Renewable … 47

particles and crystallinity [15]. This technique has several significant benefits: (1)
The crystallinity of cellulose decreases; (2) cellulose polymerization decreases to an
extent; (3) an increase in the surface that is available for enzymatic hydrolysis; and
(4) advancement in mass transfer, as a result of particle size reduction [14].

3.5.1.2 Microwave

Microwaves are electromagnetic radiation of wavelengths between 1 mm and 1 m,


which consist of nonionizing radiation with a frequency range of 300–300,000 MHz
that selectively transmits energy to various substances. Energy-efficient microwave
heating increases hydrolysis, oxidation, alkylation, and esterification processes, and
using microwaves has gained new interest. Over the past 30 years, studies have
reported effective lignocellulosic pretreatment utilizing microwave radiation and
have steadily scaled up from lab to pilot scale. There are now two primary cate-
gories for LCB in microwave-assisted pretreatment technologies: (1) The pyrolysis
of lignin is carried out at high temperatures, greater than 400 °C, in the absence
of oxygen. Microwaves are used to depolymerize biomass and convert it to bio-
oil or gases. (2) Microwave-assisted solvolysis produces high-value compounds on
depolymerization of biomass.
When compared to conventional heating, microwave radiation offers many bene-
fits, including (a) quick heat transfer, (b) fewer reaction times, (c) simple process,
and (d) little degradation or creation of harmful compounds and extraneous goods.
Additionally, microwave hydrothermal pretreatment eliminates most acetyl groups
that are elevated by the hot spot effect of microwave irradiation from hemicellulose.
During traditional heating, energy is transported through conduction from the mate-
rial’s exterior surface to its inside. So, while the interior remains cold, overheating
on the exterior surface is still possible. By directly converting the electromagnetic
energy in microwaves into heat, on the other hand, they produce heat at the molecular
level. As a result, energy is dispersed evenly throughout the material [18]. Microwave
irradiation causes explosions within the particles of a substance, facilitating the disin-
tegration of resistant structures. Great interest has been sparked in biorefinery and
its uses by introducing microwaves as a thermal technique for processing LCB. The
existence of crystalline, ionic, and nonionic amorphous areas in a typical LCB makes
it function like a biological conductor. Hot spots are produced in the biomass due to
polar groups present in the cellulose [15].

3.5.1.3 Extrusion

One of the most popular physical preparation techniques used for LCB is extrusion.
This method’s foundation is a tight barrel with temperature control and one or two
screws that spin into it. The refractory structure of lignocellulose is disrupted by the
screw blades of the barrel, which are rotated under high temperature and pressure
[16]. Materials that fit through a die with a specific cross section come out with
48 M. Maitra et al.

a rigid, definite profile. This extrusion method is well-known for recovering sugar
from biomass. Advantages of this technique include flexibility to modifications, zero
requirements of degradation products, a controlled environment, and high output.
Extrusion machines are divided into twin-screw extruders and single-screw
extruders, composed of small cylindrical parts known as screw elements [16]. The
screw design has a significant impact on the decomposition of LCB. To boost the
generation of biogas from LCB, the impact of various screw configurations on wheat
straw and deep litter during the extrusion pretreatment process was studied [19].
They examined five screw configurations for pretreatment: mild, long, reverse, and
kneading with reverse. This resulted in efficient sugar yield for each arrangement,
causing a sizable methane output. However, the key issue in this study is the high
energy consumption, which calls for additional research to upgrade the pretreatment
process and make it highly feasible. Several variables, including screw design, speed,
and barrel temperature, influence the extrusion pretreatment. The pretreatment was
explored in 2017, employing a twin-screw extruder of sugarcane bagasse and straw,
and optimized the factors, including the type of additives, the biomass, additive ratio,
the extrusion passes no., the barrel temp, the screw speed, shape, size, and its config-
urations. Different additives, such as water, ethylene glycol, glycerol, and Tween
R80, were used for pretreatment in varied amounts. When pretreatment is carried
out using glycerol, bagasse, and straw, with varying glycerol ratios of 1:0.75 and
1:0.5, experiments revealed that glycerol was considered the optimal addition factor
[16].

3.5.1.4 Pyrolysis

Pyrolysis is regarded as another efficient method for pretreating biomass. Char and
certain gases are produced on breaking down cellulose, which occurs slowly at low
temperatures. This can be improved by supplying oxygen and mild acid hydrolysis
[20]. LCB is thermally degraded during pyrolysis at extremely high temperatures
without an oxidizing agent. Pyrolysis was carried out at temperatures ranging from
500 to 800 °C. The rapid breakdown of cellulose led to charcoal and pyrolysis oil
production. Several variables, including biomass type, pyrolysis type, and reaction
parameters, influence end products. Nowadays, thermal industries are adjusting to
this method because it produces products of high value and energy, which makes
retrofitting transport management simple, allows for combustion and storage, and
allows for flexibility in utilization and marketing. This mechanism is more effective
when oxygen is present, and the temperature is lower.
Gasification, pyrolysis, and direct combustion are the three thermochemical
processes that can convert biomass to biofuels. The various pyrolysis modes are
the cause of the various product yields. Water and polar organics are the major
components of bio-oil. Where bio-oil generation is necessary, pyrolysis is used.
The generation of liquid products (fuels) is caused by rapid pyrolysis in a controlled
atmosphere. Emerging methods are now termed mild pyrolysis or torrefaction. Other
3 Lignocellulosic Biomass for Sustainable Production of Renewable … 49

thermochemical processes that occur at 200–300 °C temperatures differ from pyrol-


ysis. In this process, biomass partially decomposes, resulting in terrified biomass.
Pyrolysis produces vapor, aerosols, and char, breaking down plant biomass. Torrefac-
tion has been divided into two types: wet and dry torrefaction. An inert atmosphere,
thoroughly dried biomass, and normal atmospheric pressure are requirements for dry
torrefaction, whose main byproduct is biochar. In wet torrefaction, the pretreatment is
performed in a pressurized water vessel. Wet torrefaction uses biomass that contains
moisture, which is later dried. Other names comprise hydrothermal carbonization/
torrefaction. Wet torrefaction requires a pressure between 1 and 250 MPa to be carried
out. Hydro-char is the principal end product of the wet torrefaction processing of
biomass [18].

3.5.1.5 Ultrasonication

Pretreatment using ultrasonication is based on the cavitation theory and uses ultra-
sonic energy. The compact structure of LCB is cleaved by the shear forces produced
by cavitation, which also facilitates the extraction of desirable constituents like
cellulose, lignin hemicellulose, etc. [17]. Chemical processing uses ultrasound at
frequencies between 20 kHz and 1 MHz, whereas medical and diagnostic processes
use higher frequencies. Ultrasound processing of biomass alters the exterior struc-
ture and generates oxidizing radicals, damaging the lignocellulose matrix’s chem-
ical structure. Additionally, lignin’s connections can be broken with ultrasound by
producing small bubbles, enabling the splitting of polysaccharides and lignin. After
reaching an assertive critical size, the generated bubbles become unstable, violently
disintegrating at pressures of 1800 atm and temperatures between 2000 and 5000 °K.
As a result, ultrasonic disruption is a useful environmentally friendly approach for
pretreating LCB [19]. The ether bonds between lignin and hemicelluloses are broken
down by ultrasonic waves, affecting cell wall stability, which makes the hemicellu-
loses more accessible and extractable. This method enhances the lignin degradation
rate and saccharification of enzymes by severing hydrogen bonds between ligno-
cellulosic molecules and decreasing the crystallinity, but it cannot produce enough
energy to alter the conformation of the surface of biomass. The effectiveness of soni-
cation is determined by various variables, including ultrasonic frequency, residence
duration, biomass type, solvent utilized, reaction kinetics, and reactor configuration/
geometry [19].

3.5.1.6 Pulsed Electric Field

In this process, the cell membrane is punched with holes using a high voltage of 5–
20 kV/cm for a few nano-milli seconds to permit invading substances to enter the cell
to activate disintegration. The sample is placed between 2 parallel plate electrodes,
where the electric field strength, E, is expressed. Mass permeability and tissue rapture
dramatically increase when the electric field is shown. Then, electric pulses are
50 M. Maitra et al.

applied in the mode of exponential decay or square waves. This method comprises
components such as a control system, material handling equipment, control system
[17]. As a result of the pulsed electric field’s disruption of the biological membrane
and local structural changes that cause a loss of semi-permeability, intracellular
chemicals can pass through the biological membrane and solve in the area. As a
result, hydrolytic enzymes can enter the treated plant cell membrane more easily
through its holes. To improve the hydrolysis of cellulose, lignocellulosic materials
(wood chips and switchgrass) are pretreated with 2000 pulses with a field strength
of about 10 kV/cm [18].
The physical approaches to pretreatment are frequently utilized to decrease
the biomass’s particle dimension. Mechanical instruments and methods including
millers, grinders, screws, and UV or microwave radiations are typically used during
physical pretreatments. Furthermore, the entire procedure is aided by using secondary
or tertiary sources such as chemicals, solvents, or enzymes. An evaluation of several
pretreatment methods indicates that the optimal procedure should overcome the
LCB’s stubborn character, boost cellulosic crystallization, and guarantee the highest
possible yield of sugars and other utilitarian value bioproducts.

3.5.2 Chemical Pretreatment Approaches

Highly complex substrates can be broken down efficiently by chemically treating


them with acids, organic solvents, and alkalis. This process also increases the
bioavailability of carbohydrates by eliminating some of the inhibitors—lignin and
reducing cellulose crystallinity or polymerization levels [21].

3.5.2.1 Alkali Pretreatment

One of the earliest attempts at biomass preparation involved the use of alkaline.
While most of the cellulose remains intact after the alkaline pretreatment, a signif-
icant quantity of lignin and hemicellulose are solubilized. Enzymes with a high
hydrolysis rate and potential yield could further hydrolyze the cellulose stored in the
residue solid. The prohibitive factor for alkaline application in industry is its high cost
[20]. The hydroxides of sodium, potassium, calcium, and ammonium are alkaline
reagents generally utilized for alkali pretreatment. The most efficient of these was
discovered to be sodium hydroxide [17]. During the alkali pretreatment procedure,
a saponification reaction breaks the intermolecular ester bonds between hemicellu-
loses and lignin. This causes the lignin and hemicellulose fragments to solubilize in
the alkali solution and introduces the cellulose into the enzyme interaction, causing it
to swell. Hence, the degree of polymerization and crystallinity is brought down due
to the modification in the structure of lignocellulose, which also broadens the internal
surface area. In this method, uronic acid and acetyl groups are eliminated in hemicel-
lulose, improving the accessibility of carbohydrates during enzyme hydrolysis [16].
3 Lignocellulosic Biomass for Sustainable Production of Renewable … 51

A few bases are also utilized, in addition to acids, while biomass is pretreated. The
amount of lignin significantly impacted the alkaline treatment’s result.
Alkali treatment requires less pressure and temperature than other pretreatment
techniques. However, it also requires several hours to days of work. Alkali treat-
ment results in less sugar degradation than acid treatment, and it is also practical and
simple to remove and recover caustic salt. High efficiency is demonstrated by alka-
line pretreatment, particularly in the delignification procedure. Alkaline pretreatment
causes the cell wall to swell, increasing the interior surface area while lowering the
cellulose’s polymerization and crystallinity. Crucial process parameters, including
alkaline loading, reaction temperature, and the amount, must be optimized to obtain
alkaline effects entirely. Alkaline reagent at high concentrations causes polysac-
charides to degrade and decompose. Hence, lower concentrations of atmospheric
pressure and temperatures are advised. Additionally, it does not produce harmful
substances like furfurals and hydroxymethylfurfural (HMF), leading to improved
efficiency in the biomethanation process [21].

3.5.2.2 Acid Pretreatment

Acids are employed in this pretreatment to pretreat LCB. Acid pretreatment loses
appeal as a pretreatment method due to the production of inhibitory compounds.
After acid pretreatment, the inhibitory chemicals produced in large quantities include
furfurals, aldehydes, 5-hydroxymethylfurfural, and phenolic acids. Concentrated
acids are also used to treat LCB. Acids like sulfuric acid and hydrochloric acid
are most commonly preferred. Fermented sugars are liberated from the LCB by this
method, as it enhances hydrolysis. Sulfuric acid pretreatment for spruce, switch-
grass, poplar, and corn stover is typical. After Bermuda grass and rye were subjected
to acid pretreatment, 19.71% and 22.93% reducing sugars were generated, respec-
tively. According to studies, diluted sulfuric acid is most frequently used to treat
LCBs before usage. The impact of diluted acid and alkali pretreatment on wild rice
grass (Zizania latifolia) for enzymatic hydrolysis was compared [17].

3.5.2.3 Organosolv Procedure

In order to isolate each component of the LCB and produce relatively pure lignin that
may be sold as a byproduct or transformed into higher-value products in a biorefinery
concept, organosolv pretreatment rises as the only available method. With or without
adding a catalytic reagent, organic solvents such as methanol, ethanol, acetone, acetic
acid, peracetic acid are utilized. Oxalic, acetylsalicylic, salicylic, and mineral acids—
hydrochloric, sulfuric, and phosphoric acid—are catalysts. Although using a catalyst
can increase pretreatment efficiency, it can also adversely affect the environment.
For example, using a chemical catalyst as acid can cause lignin condensation reac-
tions, with the formation of reactive aldehydes and furfural and 5-hydroxymethyl
furfural due to the acid-catalyzed degradation of monosaccharides [21]. The features
52 M. Maitra et al.

of pretreatment biomass, such as crystallinity, fiber length, and degree of polymer-


ization, are influenced by many variable parameters, including catalyst type, temper-
ature, solvent concentration, and reaction time. Long reactions, high temperatures,
and acid concentrations all lead to the production of inhibitors. The effectiveness of
various catalysts was examined for the production of ethanol. Sulfuric acid, sodium
hydroxide, and magnesium sulfate were found to be the most effective catalysts, but
sodium hydroxide is potent for improving digestibility. Although sulfuric acid is a
useful catalyst, it is less frequently used due to its toxicity and inhibitory properties.
This method is not cost-effective since the catalysts are of high cost. The costs,
however, can be reduced by recovering and recycling solvents. Because the presence
of solvents affects fermentation, microbial growth, and enzymatic hydrolysis, their
removal is crucial. Handling such abrasive chemical solvents comes with an addi-
tional risk. Acid aids in the depolymerization and hydrolysis of lignin. When lignin
cools, it dissolves in phenol, and sugars are present in the aqueous phase. Lignin is
soluble in Formasolv, an organosolv that uses formic acid, water, and hydrochloric
acid, allowing low-temperature processing [17]. Due to its inherent benefits, such as
the simplicity with which the solvents can be recovered by distillation, the ability to
recycle the solvents back into the pretreatment process, and the utilization of high-
quality lignin extracted from this process act as value-added byproducts for industrial
applications. Hence, this pretreatment method is flourishing in the present times.

3.5.2.4 Deep Eutectic Solvents (DESs)

The adoption of eco-friendly solvents is a critical factor for determining the overall
renewability of a process, according to the fundamentals of green chemistry. The
traditional solvents, now utilized in the industry, are hazardous to the environment
and constitute a major risk. As a result, the search for new eco-friendly solvents has
intensified. By linking a hydrogen bond donor (HBD) and hydrogen bond acceptor
(HBA) in the right proportions, DESs are formulated. DESs are less expensive,
hazardous, and more biodegradable than traditional solvents and ionic liquids (ILs).
Moreover, the synthesis of DESs is straightforward and does not require elaborate
purifying procedures. One of their key benefits is the extensive tuning ability of
the physicochemical characteristics of eutectic solvents, which are similar to ionic
liquids (ILs). This method has some unique features, such as high polarity, large
miscibility with water, and the ability to mix with other co-solvents. Hence, it is used
widely for the effective delignification of various biomass feedstocks. A new and
eco-friendly approach for the pretreatment of LCB under mild reaction conditions
has been created by the DESs’ intriguing ability to selectively eliminate lignin from
the LCB complex. DESs with acidic hydrogen donors are superior to ILs and can
effectively remove lignin and hemicellulose. Zhang provides more evidence for the
findings, and his collaborators looked into how corn cobs may be pretreated using
three DESs: monocarboxylic acid (ChCl), dicarboxylic acid (ChCl), and polyal-
cohol (ChCl). It was discovered that the strength and quantity of acid and the type of
3 Lignocellulosic Biomass for Sustainable Production of Renewable … 53

hydrogen bond acceptors significantly impacted delignification and cellulose acces-


sibility. The pretreatment of polyalcohol: ChCl mixture can be enhanced when the
free hydroxyl groups of polyalcohol and the free and etherified hydroxyl groups of
lignin interact [16].

3.5.2.5 Ionic Liquids (ILs)

ILs have long piqued the curiosity of scientists who want to pretreat lignocellulose. A
novel family of solvents, known as ILs, containing cations or anions, has low melting
points, negligible vapor pressure, and good thermal stability and polarity. Ionic
liquids typically amount to big organic cations and small inorganic anions. The degree
of anion charge delocalization and cation structure greatly impact ionic liquids’ phys-
ical, chemical, and biological properties. Pretreatment time, temperature, cations,
and anions influence the ionic liquids and biomass interaction. Compared to conven-
tional solvents, ionic liquids offer special characteristics, such as low vapor pressure,
that prevent atmospheric emissions by preventing solvent losses during evaporation.
Also, they have reasonably high thermal stability, low flammability, and recycling
potential, resulting in less waste segregation [15]. Certain factors must be considered
in this pretreatment method, such as temperature, ionic liquid and its physicochem-
ical properties, biomass type, reaction duration, and amount of water in the sample.
Because extremely viscous solutions are created during pretreatment and have a
considerable negative impact on the commercial application of ionic liquids, phys-
iological properties like viscosity are the most detrimental feature of the process.
Despite large changes in biomass structure, the content of the biomass underwent
only minor alterations after ionic liquid pretreatments. Because of its great thermal
and chemical stability, less hazardous processing conditions, low solvent vapor pres-
sure, and ability to persist in the liquid state even under the widely varied range of
temperatures, this technique is not likely to be favored over alternative techniques.
These ionic liquids rarely degrade and can be recycled with ease. The disadvantage
of utilizing ionic liquid pretreatment is that cellulase will unfurl and become inactive
as a result of the incompatibility between the two substances [21].
Chemical pretreatment has drawn more attention since it is typically less expen-
sive, has faster breakdown rates for complex organic compounds, and has superior
efficiency overall. A succinct synopsis of the multiple biochemical approaches to
pretreatment and how they relate to the LCB components is provided by Pandey
et al. 2021; nonetheless, it is still incomplete. Dedicated studies employing acidic,
alkaline, and organosolv techniques to comprehend the aforementioned procedures
can expand the potential of DESs for use in upcoming bio-based refineries.
54 M. Maitra et al.

3.5.3 Physicochemical Pretreatments

Certain physicochemical approaches can be used to solubilize the LCB components


based on temperature and moisture content and to make the lignocellulose readily
accessible during the hydrolysis stage, preventing inhibitors’ production. Despite
being more difficult to apply, pretreatment of lignocellulose feedstock using these
techniques promises good yield in the following bioprocesses. Physical pretreatment
is typically more expensive and requires more energy, making it high-priced on an
industrial scale. The amount of time and energy consumed in this technique can be
lowered by process optimization [21].

3.5.3.1 Steam Explosion

Biomass is exposed to extremely hot steam in a reactor during the steam explo-
sion pretreatment procedure. The biomass is penetrated by the high-pressure steam,
which starts an autohydrolysis reaction. Hemicellulose is hydrolyzed to soluble
sugars by utilizing organic acids as catalysts, produced initially from acetyl groups
in the biomass. Highly fermented sugars are produced in this technique, compared
to biomass hydrolysis with cellulases. Reaction time, temperature, particle size, and
moisture content are crucial parameters to consider in this process. Good perfor-
mance may result from either high temperature and a quick reaction time (270 °C,
1 min) or low temperature and a lengthy reaction time [20].
Softwoods cannot be effectively pretreated with the steam explosion, but adding
an acid catalyst throughout the process is necessary to make the substrate avail-
able to hydrolytic enzymes. Steam is utilized to maintain the biomass at a specific
temperature, eliminating the requirement for a specific dilution. After the process,
the temperature is lowered by a sudden pressure release. Rapid heat expansion causes
the biomass’s particle structure to break down, which is engaged to stop the reac-
tion. Certain parameters, such as moisture content, residence duration, chip size, and
temperature, impact the steam explosion. Some advantages of this pretreatment tech-
nique include less energy requirement since 70% less energy is required compared
to the mechanical techniques to obtain the same particle size. This method, which
involves a catalyst, has been studied and reported that it could be scaled up to large
commercial levels because of its cost-effectivity [17].

3.5.3.2 Liquid Hot Water (LHW)

Analogous to a steam explosion, LHW pretreatment employs water at high pres-


sure (up to 5 MPa) and temperature conditions (170–230 °C), replacing steam. In
contrast to the steam explosion, quick pressure release is not necessary for LHW,
and pressure is merely applied to stop water from evaporating. By freeing hemicellu-
lose’s acetyl groups and removing lignin, LHW hydrolysis hemicellulose increases
3 Lignocellulosic Biomass for Sustainable Production of Renewable … 55

the explosibility of the cellulose fibers. The processed slurry’s liquid portion still
contains detached hemicellulose, and very little monomeric sugar is formed after the
treatment. However, the LHW pretreatment is carried out at a regulated pH between
4 and 7 to impede the production of inhibitors and sugar degradation [16]. The
disadvantage of LHW is that it requires a large amount of energy for the subsequent
procedure because there is a large quantity of water involved. The advantage of this
method is that no inhibitor is produced, and no chemicals or catalysts are needed [22].
Since no chemicals or catalysts are needed for LHW pretreatment, harmful mate-
rial production is nearly nonexistent, and solvent costs are minimal for large-scale
applications. Furthermore, while the particles are broken down during the treatment,
biomass size has no effect because the process is more striking when done on a large
scale. However, the process requires considerable energy because a large amount of
water is used [16].

3.5.4 Biological Pretreatment

Prior to enzymatic saccharification, LCB can be treated using a low-cost, envi-


ronmentally acceptable method known as biological pretreatment. This method is
promising since it uses less energy, produces fewer inhibitors, and is environmentally
benign. By using entire cells or enzymes from lignin-degrading bacteria or fungi,
LCB is pretreated. The enzymes lignin peroxidase, laccases, manganese peroxidase,
and versatile peroxidase are employed to break down lignin. The greatest candidates
for these uses are fungi since they can break down lignin, hemicelluloses, and cellu-
lose. Biological pretreatment is utilized to eliminate certain components, such as
antibacterial compounds, in addition to lignin. With relatively little impact on cellu-
lose, white-rot, soft-rot, and brown fungi are renowned for removing lignin and hemi-
cellulose since these organisms contain lignin-degrading enzymes like peroxidases
and laccases, which help in the breakdown of lignin. The control of these enzymes
that degrade substances involves carbon and nitrogen supplies. While white and
soft-rot specifically target plant biomass’s lignin and cellulose components, brown
rot more frequently attacks cellulose [16].
The inclusion of microbes like mold and bacteria in a naturally occurring pretreat-
ment is becoming increasingly prevalent throughout lignocellulose pretreatments
due to its positive effects on both the natural world and the economy. Arguably, the
most important step is carefully choosing the appropriate microbial consortia. Due
to their high enzyme activity, co-cultivating bacteria and/or fungus during consoli-
dated bioprocessing (CBP) is advantageous in breaking down complicated biological
polymers [23].
56 M. Maitra et al.

3.5.5 Combined Pretreatment Approaches

Gasification, pyrolysis, and direct combustion are the three thermochemical


processes that can convert biomass into biofuels. The various pyrolysis modes are
the cause of the various product yields. Pyrolysis is mostly used when bio-oil is
required, whose major components are water and polar organics. The generation
of liquid products (fuels) is caused by rapid pyrolysis in a controlled atmosphere.
Emerging methods are termed mild pyrolysis or torrefaction. The thermochemical
process that occurs at temperatures between 200 and 300 °C differs from pyrolysis.
In this process, biomass partially decomposes, resulting in terrified biomass. An inert
atmosphere, thoroughly dried biomass, and normal atmospheric pressure are require-
ments for dry torrefaction, whose byproduct is biochar. Wet torrefaction is done in a
pressurized vessel containing moisture and is dried after the torrefaction process. Wet
torrefaction requires a pressure between 1 and 250 MPa to be carried out. Hydro-char
is the principal end product of the wet torrefaction processing of biomass [18]. Due
to the shortcomings of individual pretreatment strategies, experts have been trying
to fix these issues and boost effectiveness by combining various techniques. Plenty
of studies have been conducted using an assortment of pretreatment techniques from
a few decades back. For instance, the primary drawback of fungal pretreatment is
its extended duration of action. This limitation can be addressed by integrating the
fungal treatments with other chemical as well as physical methods. For example,
compared to the acid pretreatment alone, the fungal and diluted acid pretreatment of
the olive tree elevated the amount of sugar recuperation to 51% and expanded the
hydrolysis by enzyme output by 34%. Microbiological combination diluted acid and
microbiological combined diluted alkaline pretreatments were recently employed by
Zhang et al. [24].
In the case of physical pretreatment measures, colloid mills and extruders are
advised for pretreating LCBs with moisture contents of more than 15–20%, whereas
hammer and knife mills are advised for LCBs with moisture contents of 10–15%. As
the target particle size lowers and the moisture content rises, more energy is needed
in normal circumstances. Grinding hemp to improve methane output by 15% needed
more energy than steam pretreatment procedures [25]. In the field of pretreatment,
acid pretreatment, which is not completely eco-friendly, becomes more significant.
In acid-catalyzed pretreatment, temperature and acid concentration are two of the
foremost processing aspects. Although the C5 sugars are similarly degraded by high
temperatures and an excess of acid in the organosolv treatment, it is an effective
pretreatment approach. Alkaline prior treatments are the most feasible treatment
approach under all these facts. Ammonia-assisted alkaline treatment is an original
method; however, it could not tackle the underlying cause of hemicellulose decom-
position. This procedure also requires a high temperature. Limestone and sodium
hydroxide are fairly affordable. The primary benefit of this procedure is that it does
not turn the C5 sugars into inhibitors produced from furans. This alkaline approach
is an efficient and successful way of attaining delignification. When all the benefits
3 Lignocellulosic Biomass for Sustainable Production of Renewable … 57

and drawbacks are considered, lime-assisted or sodium hydroxide treatment can be


regarded as the most environmentally secure and feasible business approach [26].

3.6 Various Applications of Biomass for Bioenergy


Production

3.6.1 Production of Second-Generation Bioethanol Using


LCB

Bioethanol can be obtained from carbohydrates as raw materials such as fruit juices,
leftovers from breweries, soya, peanuts, wheat-rice, and particularly waste, corn
waste, trees, rice wastes which contain lignocellulose that can be used as feedstocks
in traditional methods (Fig. 3.3).
The pretreatment of the lignocellulose is an important stage, followed by hydrol-
ysis and fermentation [27]. Physical, thermochemical, physicochemical, and biolog-
ical techniques are the most frequently employed pretreatment procedures for LCB
conversion. These approaches have been thoroughly studied but have several draw-
backs, like limited yield, high processing costs, and detrimental environmental
effects. As a result, more effective green technologies are always being investi-
gated to address these issues [28]. In the case of the physical pretreatment method,
IL pretreatment and DESs pretreatment have been used. According to numerous
researches, LCB, which has been pretreated using ionic liquid treatment, showed
surface area expansion with decreased cellulose crystallinity and lignin content.

Fig. 3.3 Bioethanol production


58 M. Maitra et al.

Although ILs pretreatment methods are cost-effective and environmentally benefi-


cial, their implementation in large-scale biorefineries may be constrained. DESs, a
new class of environmentally friendly and designer solvents, were brought to the
scientific community’s attention by [29] as an alternative to ILs. For the production
of bioethanol, lignocellulose is subjected to pretreatment procedures before being
hydrolyzed and fermented because of the following:
• Improves the cellulose amorphous regions, which can be dissolved more easily
than the crystalline regions.
• Enhances the porosity of fibers by perforating chemicals and enzymes into the
structure.
• Free the cellulose from its lignin and hemicellulose contents [27].
Another appealing method that promises affordable equipment and operation costs
is using microorganisms to pretreat lignocellulose, which is the biological approach.
Combining two or three pretreatment methods is common in producing bioethanol
derived from lignin cellulose. The biomass must be mechanically milled, sliced, and
crushed to reduce the size. Subsequently, it must undergo a steam explosion, and
finally, it must be immersed in an acidic or alkaline solution. Then, the pretreated
material must be washed and neutralized [27].

3.6.1.1 Bioethanol from Pretreated Lignocellulose

In addition to fermentation, direct microbial conversion (DMC), hydrolysis, and


the biomass that has been pretreated can all be utilized for producing bioethanol.
The DMC procedure is time-consuming, the conversion yields are only moder-
ately high, and there is a considerable risk of contamination. Enzymatic hydrolysis,
combined with microbial fermentation, is a method with higher preference and has
been demonstrated to work significantly better [27].
Lignocellulosic saccharification: After being processed with lignocellulose, the
polysaccharide material is degraded by enzymes into monosaccharides. A combina-
tion of many enzymes isolated from microorganisms makes up the cellulose, which
is commercially accessible and used to hydrolyze cellulose and hemicellulose. To
break down cellulose enzymatically, the following three enzymes are used: Exo-1,4-
β-D-glucanases, including 1,4-β-D-glucan glucohydrolase, to free D-glucose from
1,4-β-D-glucan and slowly hydrolyze D-cellobiose, and enzyme 1,4-β-D-glucan
cellobiohydrolase, to free cellobiose from 1,4-β-glucan. Also, Endo-1,4-glucanases
(EG) or 1,4-D-glucan 4-glucanohydrolases which cleave 1,4-glucan linkages and β-
D-Glucosidase or β-D-glucoside glucohydrolase forms D-glucose from cellobiose
[27].
LCB saccharification using enzymes: LCB is converted to sugar monomers, but
their industrial yields are one of the primary obstacles to produce second-generation
biofuels, as it is challenging to obtain high process efficiencies. Hydrolyzing the
polysaccharide is done using acid to obtain monomers because it completely trans-
forms the crystalline form of the polysaccharide into an amorphous state by dissolving
3 Lignocellulosic Biomass for Sustainable Production of Renewable … 59

the hydrogen bonds between the cellulose chains. Cellulase, hemicellulases, and
additional microbial lignocellulolytic enzymes have been discovered in various
biological settings, including woodlands, composting soils, compost piles, sewage
sludge, and rumens [30]. Aerobic bacteria and filament-containing fungi, along with
some enzyme complexes, form the cellulosomes, which are abundant in fungal
strains, and anaerobic bacteria belong to common lignocellulolytic systems.

3.6.1.2 Strain Development for Improved Bioethanol Fermentation

Lignocellulolytic enzymes are very costly; this is a major bottleneck in biofuel-


producing industries to decrease the enzymes’ price and enhance the enzymes’ yields
with the required productivity. Recently, several techniques have been employed for
enhancing the expression of enzymes using microbial strains, including genome
and metabolic engineering, CRISPR-Cas9, protoplast fusion, protein engineering,
cell surface engineering, genetic engineering, and recombinant DNA technology.
Random mutagenesis is widely used to alter microorganisms’ natural makeup and
functionality. Numerous papers have described cyclic mutagenesis that combined
physical (UV radiations) and chemical (ethanomethane, acriflavine, and N-methyl-N
Nitro-N-nitrosoguanidine) factors. Site-directed mutagenesis is another widely used
method that causes precise changes in the recognized DNA sequences. Site-directed
mutagenesis and error-prone PCR (EP-PCR) have been demonstrated to increase
EGIII’s alkaline tolerance in Trichoderma reesei and increase enzyme activity in
Bacillus amyloliquefaciens DL-3 (by up to 7.93 times) [31]. To produce alcoholic
fuel, the microbial genes are altered through adaptation, selection, mutation, and
protoplast fusion to improve fermentation and saccharification efficiency. Choosing
genetically modified ethanol-resistant bacteria involves transposon-mediated muta-
tion collection and deletion mutant library screening [32, 33].

3.6.1.3 Fermentation Procedures for Producing Bioethanol

The simple sugars produced during enzymatic hydrolysis can be converted to


bioethanol by microorganisms. There are two processes, including separate as well
as simultaneous fermentation and hydrolysis.
Ninety-five percent of soluble carbohydrates undergo fermentation in order to
yield CO2 and ethanol. Pentose and hexose sugars are mixed in the hydrolysate
substrate of LCB, in contrast to the processing of starch hydrolysate, which solely
includes glucose. Numerous microbes that occur naturally can easily ferment
hexose carbohydrates, including galactose, glucose, and sugar mannose. Typically,
Zymomonas mobilis and Saccharomyces cerevisiae are the foremost ones. Never-
theless, pentose sugars like arabinose as well as xylose cannot be processed by
natural microbes. Conversely, several yeast strains, such as Kluyveromyces marxi-
anus, Pichia stiipitis, Pachysolen tannophilus, and Candida shehatae ferment xylose
[34].
60 M. Maitra et al.

Separate hydrolysis and fermentation (SHF): After hydrolyzing the sugars,


bacteria are added to the mixture to ferment them. Contamination, the introduc-
tion of inhibitors, and the need for more time and equipment are a few of its inherent
weaknesses [27].
Simultaneous saccharification and fermentation (SSF): Hydrolysis using enzymes
and fermentation using microbes occur simultaneously in individual machinery. This
mixture contains microorganisms and enzymes. This method has been demonstrated
to be substantially more effective than the SHF mentioned above, with less equip-
ment, less time, and a lesser risk of contamination [27]. This step is very impor-
tant to produce chemicals and fuels that are industrially relevant and involves the
conversion of sugar monomers produced by hydrolyzing the raw materials into prod-
ucts by the activity of microorganisms. Calonectria brassicae, Zymomonas mobilis,
Saccharomyces cerevisiae, Pichia stipitis, Pachysolen tannophilus, Mucorindicus,
Escherichia coli, and Candida shehatae are the wild-type microorganisms inves-
tigated in ethanol fermentation [35]. Some variables that may impact the fermen-
tation process for producing bioethanol include salt concentration, pH, carbohy-
drate content, aeration rate, ethanol concentration, and temperature. The three basic
fermentation processes used to make ethanol are batch, fed-batch, and continuous
fermentation. These methods have their own merits and demerits [35].
Batch fermentation: The high-concentration substrate is converted into a desirable
product in this fermentation. It is a very old and popular fermentation operation.
Repeated batch fermentation is a variation of this operation, which increases the
system’s effectiveness by using immobilized microbial cells rather than free ones.
Fed-batch fermentation: Unremoved products are intermittently added while using
the fed-batch method, which combines batch and continuous modes of fermentation.
This fermentation has some merits over batch fermentation, which involves high
ethanol productivity, lower toxicity of media components, short fermentation period,
and higher dissolved oxygen in the media.
Continuous fermentation: The most prevalent kind of fermentation, known as
continuous fermentation, has been utilized to produce industrial quantities of
bioethanol since it is simple to regulate the process, does not require as much time
to clean, and is less expensive to invest [36, 37].
Ethanol generation using multiple LCBs is mostly governed by the extent to which
glucan hydrolyzes into soluble sucrose. The order of glucan digestibility among the
residues from agriculture was stated to be as follows: bagasse (47%) > crop maize
stover (23%) > hemp for industrial use hurd (14%) > wheat residue (11%) > rice stalks
(10%). Conversely, the ability to be digested by bioenergy crops was found to be lesser
than that derived from agricultural wastes, with common reed ranking highest (19%),
followed by switchgrass (17%), Miscanthus (8%), and bamboo (3%). Poplar, one type
of untreated biomass from trees, has an estimated glucan digestibility of about 7%. A
comparison analysis was conducted employing sugarcane, maize, and sugar beetroot
as the feedstocks implemented in the industrialization of ethanol in Brazil, the USA,
and Europe [38]. This study shows that sugarcane is the most affordable fodder
when viewed alongside the other two since it reduces carbon dioxide emissions by
30 and 40% when compared to maize and sugar beetroot. According to estimates,
3 Lignocellulosic Biomass for Sustainable Production of Renewable … 61

sugarcane has more possibilities for ethanol output than sugar beet (5500 L/ha) and
maize (4180 L/ha) [34].

3.6.2 Production of Second-Generation Biobutanol Using


LCB

Due to the depletion of petroleum reserves and the fluctuating price of gasoline,
replacing fossil fuels with alternative energy sources is a critical concern. One of the
problematic liquid fuels is bioethanol, but its use is restricted by its higher volatility
and corrosiveness compared to petrol and its lower energy content. Biobutanol is
a biofuel with a significant potential that may be utilized directly in car engines
without modification due to its similarities to petrol characteristics [39]. By using
lignocellulosic feedstock, we can obtain biobutanol (Fig. 3.4).

3.6.2.1 Renewable and Carbohydrate-Rich Lignocellulosic Feed


for Producing Biobutanol

Organic municipal wastes, which include fibrous and wood waste, agricultural
wastes, and waste generated from industries, are all included in the feedstock, which
contains lignocellulose; also, plants are comprised of cellulose, lignin, and hemi-
cellulose [40]. Typically, LCB has 15–25% lignin, 38–50% cellulose, and 23–32%
hemicellulose. The proportions of lignin, cellulose, and hemicellulose in different

Fig. 3.4 Acetone-butanol-ethanol synthesis by Clostridium acetobutylicum


62 M. Maitra et al.

lignocellulosic materials vary. In comparison to maize cob, which has 14–15% lignin,
42–45% cellulose, and 35–39% hemicellulose, rice straw has 12–14% lignin, 28–
36% cellulose and 23–28% hemicellulose [41–43]. Because of its crystallinity and
the resultant intense inter- and intramolecular hydrogen bonding, it greatly resists
enzymatic hydrolysis. Xylans, mannans, and glucomannans are only a few polysac-
charides found in hemicellulose, another polymer in plant cell walls. By forming
hydrogen bonds with lignin or cellulose, it contributes to the sturdiness of the cell
wall. P-hydro-xycinnamoyl, coniferyl, and sinapyl alcohols make up the three alcohol
groups that make up lignin, along with a variety of interunit connections such as
biphenyl and other bonds [43, 44]. It contributes to the stability of the structure while
protecting it from oxidative stress and microbial attack [45]. The most potential feed-
stocks are LCB and other abundant and sustainable raw resources. However, most
solventogenic bacteria, including Clostridia, could not use lignocellulose because
they lacked the expression of enzymes that break down lignocellulose [46].

3.6.2.2 The Traditional Method of Producing Biobutanol


from Lignocellulose

The most popular methods for releasing sugars are thermochemical and enzymatic
hydrolysis procedures [47]. Although acid hydrolysis is a less expensive and quicker
technique, it produces inhibitors that are hazardous to cell growth and are not ecolog-
ically friendly. However, the complex structure of LCB necessitates using many
enzymes that work in concert to degrade lignocellulose. While hemicellulose is a
complex polymer of carbohydrates, which contains significant amounts of hydrolyz-
able subunits, it also requires a large number of enzymes for degradation, such
as glucuronidase, exo-xylanase, acetyxylan esterase, endo-xylanase, xylosidase [48,
49]. The biobutanol manufacturing process is less efficient due to the costly enzymes,
which contribute around 50% of the production expenses. In general, enzymatic
hydrolysis and chemical methods were used to pretreat lignocellulosic feedstock
and modify the LCB’s structure and chemical makeup to accelerate the conversion of
biomass feedstock to obtain biobutanol [50]. However, the inhibitors formed during
the lignocellulose hydrolysis and enzyme cost restrict the utilization.
Acid/alkaline hydrolysis and ozonolysis are currently used pretreatment tech-
niques involving large amounts of chemicals, including powerful oxidizing agents,
acids, and alkalis, that could hinder hydrolysis. Additionally, because these
compounds are poisonous and dangerous, corroding pretreatment is more expen-
sive because anti-corrosion equipment is required. Similarly, physical–chemical
processes involving high temperature and pressure, such as steam explosion,
ammonia fiber expansion, and moist air oxidation, render the process too expen-
sive and energy-intensive to be practical. Thus, the main pretreatment problem is to
prevent carbohydrate degradation or loss while reducing treatment costs to make the
procedure economically viable [39]. In the case of acetone-butanol-ethanol fermen-
tation, Clostridium Saccharoperbutylacetonicum, Clostridium saccahrobutylicum,
Clostridium beijerinckii, Clostridium Acetobutylicum have been used. The two
3 Lignocellulosic Biomass for Sustainable Production of Renewable … 63

stages of acetone-butanol-ethanol synthesis involve acidogenesis and solventogen-


esis. Oxidizing agents, diluted acid, alkaline processing, and liquid hot water are used
during the chemical pretreatment of lignocellulose to manufacture biobutanol. The
commonly solventogenic C. beijerinckii is used to produce butanol, which is basically
used in chemical pretreatment, specifically in alkali pretreatment techniques [51]. To
reduce lignocellulose’s high resistance to biological transformation, the cross-linked
matrices are dissolved before enzymatic hydrolysis, which is required for generating
biobutanol [52]. Before fermentation, the enzymes hydrolyze the lignocellulose using
substances like crude enzyme extract or commercial cellulase.

3.6.2.3 Fermentation Procedures for Manufacturing Biobutanol

Enzymatic hydrolysis is necessary after pretreatment to obtain butanol by trans-


forming cellulose into glucose and later into butanol. The process of turning LCB into
butanol occurs via four major processes that are based on the production of cellulase,
cellulose hydrolysis, and butanol fermentation: consolidated biotreatment (CBP),
simultaneous saccharification, and co-fermentation (SSCF) and separate hydrolysis
and fermentation (SHF) [39].
CBP: This is one of the most significant barriers to SHF and SSF. Therefore,
creating affordable, efficient, and effective fermentation and saccharification methods
is essential for biobutanol production. It, also known as direct microbial conversion,
is a substitute method to minimize this significant cost increase. Cellulase synthesis,
fermentation, and cellulose hydrolysis are accomplished during CBP in one phase.
This process is cost-effective for low-cost fermentation and biomass hydrolysis.
Because enzymatic hydrolysis and cellulase synthesis come at operational and capital
costs that are higher than those of SSF and SHF processes, it is anticipated that CBP
has the potential to lower expenses by more than 50% [39].
SSCF: During this procedure, cellulose fermentation and hydrolysis occur simul-
taneously. Also, a process named simultaneous saccharification and co-fermentation
operates to yield butanol by hydrolyzing both C5 and C6 sugars. It operates at lower
temperatures, usually 35–37 °C, to promote microbial development and butanol
production [39].
SHF: Cellulases hydrolyze cellulose to glucose under ideal circumstances at a
temperature of 50 °C. The glucose obtained is then transported into a reactor where
fermentation occurs with the help of microorganisms that produce butanol at a
temperature of 35 °C. It is currently used most frequently in the butanol manu-
facturing process of LCB. Cellulose hydrolysis and fermentation can occur in this
process under their respective ideal circumstances. Cellulases’ activity during enzy-
matic hydrolysis may be inhibited, nevertheless, by the buildup of glucose during
the process [39].
64 M. Maitra et al.

3.6.2.4 Techniques for Increasing Butanol Production


from Lignocellulose

Detoxification of inhibitors: During hydrolysis and pretreatment, sugar monomers


are formed along with many unwanted degradation products that can be produced,
including aliphatic acids, furan derivatives, and phenolic compounds. These
inhibitors were found to affect biomass conversion efficiency significantly. Therefore,
the key to effective biofuel production is the removal of these inhibiting substances
prior to fermentation. Adsorption, extraction, precipitation, and combinations are
the primary techniques used in traditional inhibitor removal processes. An alternate
approach to removing inhibitory substances from biomass hydrolysates is biological
control [39].
Strain improvement: The poor butanol yield caused by the production of acetone
and ethanol byproducts and the microorganism’s low butanol tolerance are the key
obstacles to the industrialization of butanol fermentation. To address this weakness,
butanol fermentation necessitates the creation of strong strains with increased butanol
yield and tolerance. Mutagenesis has been used to modify the strains; for example,
both C. beijerinckii BA101 and C. beijerinckii NCIMB 8052 produced 19 g/L of
butanol when treated with Methylnitronitrosoguanidine [39].
Clostridium sp. favors glucose as its major energy source. In batch fermentation,
various amounts of glucose (20, 40, 60, and 80 g/L) were employed as a carbon
substrate for C. acetobutylicum to synthesize biobutanol, and an elevated biobu-
tanol titer of 11.10 g/L was observed. In conformity with other recently reported
studies, the ideal level of glucose for batch fermentation employing suspended C.
acetobutylicum cells was 60 g/L. After glucose, xylose is the second-largest sugar
found in lignocellulosic biomass hydrolysates. ABE fermentation has been carried
out in a batch approach, with C. acetobutylicum implementing xylose as the sole
carbon source. At 60 g/L xylose, the maximum biobutanol titer was 8.52 g/L. Since
Clostridium groups are also amylolytic, they may formulate biobutanol as an alter-
native to carbon. As a result, 9.10 g/L of reduced concentration xylose is produced,
and the biobutanol titer yields 0.42 mol biobutanol/mol glucose. Alkaline pretreated
rice straw outperformed sugarcane bagasse as the substrate of choice. In the absence
of yeast extract, the use of immobilized cells during the phase of fermentation leads
to the high biobutanol generation, efficiency, and concentration of 13.80 g/L, 0.90 g/
L/h, and 0.58 mol biobutanol/mol glucose, respectively [47].

3.6.3 Production of Second-Generation Biodiesel Using LCB

Biodiesel is sustainable, eco-friendly, harmless, sulfate-free, biodegradable,


combustible, and free of aromatic chemicals. Compared to normal diesel, it has
a greater flash point and cetane number. The intrinsic lubricate nature improves
engine performance [53], negating the need for additional engine lubrication. To cut
the cost of production, it is a good idea to use cooking oil, oils from non-edible crops,
3 Lignocellulosic Biomass for Sustainable Production of Renewable … 65

and animal fats as raw materials. Additionally, this approach is unable to meet the
current energy demands for renewable fuels. Therefore, in order to generate biodiesel,
experts are seeking new oil sources. Due to its similarity in fatty acid composition
to vegetable oils, single-cell oil, a lipid generated by microorganisms like yeast,
bacteria, molds, and algae, is one of many resources that are thought to be a suitable
feedstock for the generation of biodiesel. These microorganisms can synthesize oils
in their cellular compartments using organic carbon (Fig. 3.5) [33, 54].

3.6.3.1 Traditional Feedstocks Used to Produce Biodiesel

Lignocellulose-containing feedstocks are the greatest sustainable resource for


biodiesel production. It should be emphasized that not all available wastes are often
used for producing biodiesel; some are used for producing fodder, manure, and paper
or are directly burned as fuels [47]. Edible plants are a source of vegetable oils that
serve as a starting material for biodiesel production. The objective of employing
vegetable and non-edible oils as feedstocks for biodiesel manufacturing has been
defeated by their exorbitant price, acreage need, and low oil output [48]. Oleaginous
microorganisms (OMs) can use low-cost feedstocks, such as agricultural wastes,
lignocellulosic substrates (LCSs), and waste substrates to accumulate more lipids
[55]. LCB is converted to enriched bioproducts that help generate electricity through
alternative methods in place of direct burning. Most commonly, waste cooking
oils, animal fats, and vegetable oils, either edible or inedible, are used in a trans-
esterification procedure to create biodiesel [56]. Due to global food security, oil
sourced from food sources is insufficient for large-scale biodiesel manufacturing.
Thus, looking for sustainable feedstocks for manufacturing biodiesel, including
undigestible lignocellulose-containing raw materials, is essential [57].
Agricultural residues: Numerous agricultural waste products are a source of
lignocellulose-containing raw materials that help in lipid production by generating
small amounts of sugars. Crop remnants and agro-industrial residues are two general
groups into which these residues fall [22]. Crop remnants are parts of edible products
like grains crops that are obtained after harvesting, whereas agri-industrial leftovers
are byproducts after post-harvesting processes.
Non-edible crop residues: Food crops are grown on the world’s most fertile lands.
However, due to the population’s rapid growth, we will soon be confronted with the
issues of a food shortage and an energy catastrophe. Forest-derived woody biomass
is still a significant source of raw materials for biofuel production [58–61]. Most
non-edible LCB comes from agricultural residues, followed by these byproducts
left over after superfluous elements are eliminated to create trees of proper size and
shape. Other notable sources of LCBs include sawmill rejects and sawdust, pallets,
cardboard boxes, wooden packing crates, and recovered wood obtained from the
demolition of buildings [57, 62]. Many LCBs are made from non-edible energy
crops, [63] agricultural waste, herbaceous and municipal solid waste, and forest
woody feedstocks. The most promising plant leftovers, which help generate energy,
are present in the USA, Europe, and Asia [22, 64, 65], including corn stover, rice
66 M. Maitra et al.

straws and wheat, sugarcane husk, and bagasse. Lignin, hemicellulose, and cellu-
lose are the three primary polymers that comprise LCBs. These polymers are tightly
entangled and linked together by covalent and non-covalent connections. The ligno-
cellulose that exists in plants contains cellulose, which is the primary structural
component of cell walls. Hemicellulose and lignin surround the microfibrils that make
up celluloses. The repeating unit of cellulose, known as the “clubhouse,” contains D-
glucose subunits that are connected by α-(1,4)-glycosidic linkages. Proper pretreat-
ment is needed to get cellulose units from LCBs, and then the material must be
further processed with sulfuric acid or an enzyme cocktail to produce sugars. The
hydrolysates include pentoses and hexoses released from the treated cellulose.
Industrial residues: If industrial waste or byproducts are not properly disposed
of, they can significantly negatively influence the environment in many ways. In
many situations, the expense of treating these pollutants exceeds the cost of produc-
tion, and they are often handled in ways that are harmful to the environment [30].
Anaerobic digestion of organic wastes demands expertise to assure safety due to the
buildup of dangerous gases. Oleaginous bacteria can process the waste materials as
an alternative to anaerobic digestion, metabolizing them as lipid droplets inside their
cells before further processing them into biodiesel [59]. The biological treatment
using microorganisms is more effective at lowering the amount of hazardous organic
compounds in treated effluents.
Oleaginous microbes: Under stressful circumstances, such as those involving low
nitrogen and high carbon sources, these bacteria can store fat up to 70% or more
of their dry weight. These bacteria utilize a variety of renewable resources to create
microbial oil, which is then used in the transesterification process to create biodiesel
[55]. Production of biodiesel with OM has a number of benefits, including a short
incubation period in contrast to animal and plant resources, and the production of
lipids is not affected by seasonal, meteorological, or geographic fluctuations. Micro-
bial oil has a low viscosity and is similar in caloric value and composition to oils
derived from plant and animal resources. OM are microalgae, yeasts, filamentous
fungi or molds, bacteria, and other microbial groups. Because it can supply enough oil
as a feedstock for worldwide consumption, it yields far more biodiesel per acre than
plant-based feedstock. Biodiesel obtained from microalgae is currently the subject
of research interest.

3.6.3.2 The Traditional Method of Producing Biodiesel


from Lignocellulose

Some pretreatment methods are undertaken to enhance the digestibility of ligno-


cellulose materials, including oxidative, acid, thermal, alkaline, and mechanical
pretreatment. Additionally, CO2 explosion, acid-triggered steam explosion, liquid
hot water, and ammonia fiber explosion are currently used as pretreatment methods.
The approach selected is determined by the initial makeup of the materials used [66].
The inhibitors that are generated during the process and present in the medium are
3 Lignocellulosic Biomass for Sustainable Production of Renewable … 67

what the oleaginous microorganisms use to ferment materials. Mechanical pretreat-


ment is the best method for converting LCB to produce biodiesel. Milling is a process
that reduces biomass particle size and eliminates crystallinity [67]. When biomass is
sheared, the surface area increases, decreasing polymerization levels and accelerating
total hydrolysis [68]. The LCBs are heated over 150 °C during thermal pretreatment
to saturate the biomass’s hemicellulose portion, followed by saturating lignin [69].
The two primary components of hemicellulose are glucomannan and xylan, being
the least thermally stable. When LCB is heated to a high temperature (4160 °C),
inhibitors prevent bacteria and yeasts’ growth [70, 71]. Alkaline pretreatment
involves converting dissolved polysaccharides into xylan while maintaining a low
temperature to prevent end groups from peeling off and losing carbohydrates. The
biomass itself consumes some of the alkalies used for alkaline pretreatment, with
the remaining alkali being utilized in subsequent reactions. Industries frequently
use various procedures [70]. The solubilization of the components is the main
distinction between these two procedures; while lignin fraction solubilization is seen
during alkali pretreatment, hemicellulose solubilization occurs more slowly during
acid pretreatment. Adopting an alkali pretreatment method removes the lignin frac-
tion without causing the other biomass constituents to deteriorate [66]. For specific
biodiesel production, ozonolysis, liquid hot water, steam explosion, carbon dioxide
(CO2 ) pretreatment, microwave treatment as well as some biological approaches
have been applied. The main disadvantage of utilizing this pretreatment technique,
which alters the composition of the hydrolysates, is salt production.

3.6.3.3 Bioconversion of LCB to Biodiesel Production

Complex unit processes, such as depolymerization of lignin, saccharification of


biomass, cultivation of OMs for producing lipids, and transesterification, are used in
biodiesel production using LCB [55].
Lignin depolymerization: Lignin is removed during pretreatment to enhance the
hemicellulose and cellulose hydrolysis components and create fermentable sugars.
Through this method, cellulose crystallinity is decreased, porosity is increased, and
hydrolytic enzymes and chemicals for efficient hydrolysis of biomass better access
hemicellulose. To remove lignin, several pretreatment processes such as physico-
chemical, physical, biological, and chemical, pretreatments are used, such as pulsed
electric field irradiation, extrusion, and pyrolysis to minimize the particle size,
crystallization, and degree of polymerization. Pretreatment via physical method
is economically infeasible on a larger industrial scale since it requires consider-
able energy. There are many different chemical pretreatment methods, such as basic
(potassium hydroxide, calcium oxide, and magnesium oxide), acidic (hydrochloric
acid and sulfuric acid), ionic liquid, and organosol pretreatment. The term “physic-
ochemical pretreatment” refers to the mix of chemical and physical processes, and
it encompasses some techniques such as CO2 explosion, ammonia fiber, explo-
sion, steam explosion, and hydrothermal methods. Additionally, microorganisms
or enzymes are used in the biological processing of biomass [55].
68 M. Maitra et al.

LCB saccharification: Prior to saccharification, lignocellulosic materials are


pretreated to transform carbohydrate polymers into fermentable sugars effectively.
Enzymatic and chemical techniques are used to hydrolyze the biomass cellulose
portion. At high temperatures and the presence of acid, the biomass feedstock
polysaccharides can hydrolyze chemically. The creation of furfurals, the poor sugar
yield, and the method’s lower economic feasibility are drawbacks of this approach.
Enzymatic hydrolysis ensues under low pressure, weak pH, and low temperatures
(45–50 °C) [55].
Transesterification and lipid extraction: Extraction of lipids from the microbe is
a primary requirement for cell wall degradation. A variety of processes such as bead
mill, subcritical water hydrolysis, pulsed electric field, ultrasound, high-pressure
homogenization, osmotic shock, microwave, chemical hydrolysis, enzymatic hydrol-
ysis, autolysis are commonly utilized to disrupt cells. In addition to pressing and inte-
grated solvent extraction, elevated temperature, supercritical extraction, and pressure
solvent extraction, ultrasonic extraction is employed to extract lipids from the oleagi-
nous microbes. Dried biomass residues of algae are pyrolyzed into oils that are used
for transporting fuel. Biomass containing moisture is treated by anaerobic digestion
and hydrothermal means to produce diesel fuel, gasoline, and methane. Once the
extraction has been done, lipids are transformed into biodiesel fuel through transes-
terification; this helps reduce lipids’ viscous nature. Glycerol and esters are formed
when lipid reacts with alcohol during transesterification (Eq. 3.1). The reactions
are chemically catalyzed and divided into homogeneous and heterogeneous reac-
tions. A solid acid or alkaline catalyst is required for the feasibility of heterogeneous
reactions.

(3.1)

Moreover, a homogeneous reaction takes place using an acid or alkaline catalyst.


Because of soap formation, due to higher free fatty acids, alkali-based transester-
ification is not feasible. The drawbacks of acid-based transesterification include a
longer incubation period, corrosive character, higher energy usage, higher glycerin
purification expenses, etc. [55].
3 Lignocellulosic Biomass for Sustainable Production of Renewable … 69

Fig. 3.5 Biodiesel production from LCB

3.6.4 LCB for Second-Generation Polyhydroxybutyrate


(PHB) Ester Production Production

The biodegradability of PHB esters makes them environmentally friendly materials.


They can be degraded by microorganisms present in soil, water, and composting envi-
ronments, ultimately breaking down into natural components such as water, carbon
dioxide, and biomass. This biodegradability makes PHB esters suitable for various
applications where sustainability and environmental compatibility are essential,
including packaging materials, agricultural films, disposable products, and biomed-
ical applications. PHB esters are a class of biodegradable polymers derived from
polyhydroxybutyrate, which is a naturally occurring polyester. PHB esters exhibit
improved thermal and mechanical properties compared to unmodified PHB. They
have increased flexibility, toughness, and elongation at break. Additionally, the intro-
duction of ester groups enhances the processability of PHB, allowing for easier melt
processing and shaping into various forms such as films, fibers, and molded prod-
ucts [38]. Overall, PHB esters combine the biodegradability of PHB with improved
properties, making them versatile and eco-friendly materials for a wide array of uses.

3.6.4.1 Bioconversion of Lignocellulosic Feed to PHB Esters

PHB esters can be obtained from LCB using a two-stage procedure. Initially, ligno-
cellulose is transformed into sugars, and subsequently, these sugars are fermented to
produce PHB esters. To manufacture PHB esters, the usual pretreatment approach is
significant.
70 M. Maitra et al.

3.6.4.2 Prior Treatments to Enzymatic Hydrolysis

For producing second-generation biofuel, this method aims to disrupt the lignin
structure by reducing the crystallinity of cellulose. Lignin serves as a protective
blockade, impeding the accessibility of hemicellulose and cellulose. Pretreatment
techniques are employed to degrade or alter lignin, enhancing the exposure of cellu-
lose and hemicellulose. This makes them more amenable to enzymatic or chemical
hydrolysis. Cellulose, in its native form, has a highly ordered crystalline structure,
which limits enzyme accessibility. Pretreatment methods disrupt the crystalline struc-
ture of cellulose, increasing its surface area and susceptibility to enzymatic hydrol-
ysis. Removing the hemicellulose is another crucial point to improve its enzymatic
digestibility. Eventually, increasing the porosity and surface area enhances the contact
between the biomass and the catalysts or enzymes used in subsequent conversion
processes [38]. Several chemical, biological, and physical pretreatment methods,
including milling, grinding, and size reduction techniques, are available to mechan-
ically break down the biomass structure. In the case of physicochemical methods,
they integrate physical as well as chemical workflows, for instance, ammonia fiber
expansion (AFEX), organosolv processes, and steam explosion. The choice of the
pretreatment method is contingent upon many factors, including biomass feedstock,
target products, process economics, and environmental considerations. The choice
of pretreatment method can significantly impact the efficiency and cost-effectiveness
of the overall biomass conversion process.
Enzymatic hydrolysis: Enzymatic treatment offers several advantages, including
specificity, mild operating conditions, and environmental friendliness. Here are some
key aspects of enzymatic treatment for LCB [72].
Enzymes used: Enzymes used in the enzymatic treatment of LCB include cellu-
lases, hemicellulases, and lignin-modifying enzymes. Cellulases are responsible for
the hydrolysis of cellulose into glucose, while hemicellulases break down hemicel-
lulose into various sugars. Lignin-modifying enzymes, such as lignin peroxidase and
laccase, target lignin degradation.
Enzyme production: Enzymes for biomass conversion can be obtained from
various sources. Fungi, such as Trichoderma species, are commonly utilized
for enzyme production as they have high cellulase and hemicellulase activities.
Aspergillus species can also be used for the same reason.
Enzyme cocktail: A mixture of different enzymes is often used as an enzyme
cocktail due to the intricate nature of LCB. The enzyme cocktail contains multiple
enzymes that act synergistically to break down different components of the biomass.
This approach improves the overall hydrolysis efficiency and reduces the enzyme
dosage required.
Process conditions: Enzymatic treatment generally necessitates gentle process
conditions, typically involving moderate temperatures, ranging from 40 to 60 °C and
a pH level close to neutral. The specific process conditions may vary based on the
enzymes employed and the characteristics of the biomass feedstock.
Inhibitors and enzyme stability: LCB can contain inhibitory compounds, such
as lignin-derived phenolics and sugar degradation products, which can inhibit
3 Lignocellulosic Biomass for Sustainable Production of Renewable … 71

enzyme activity. Strategies to mitigate enzyme inhibition include detoxification of


the biomass or modification of enzymes for improved stability and tolerance to
inhibitors.
Enzyme recycling: Enzymatic treatment processes can be designed to enable
enzyme recycling, where the enzymes are recovered and reused in subsequent hydrol-
ysis steps. This reduces the overall expenses associated with enzyme production,
rendering the process more economically feasible.
Enzymatic treatment of LCB significantly contributes to the generation of
biofuels, biochemicals, and bioproducts. Advances in enzyme technology, enzyme
engineering, and process optimization continue to upgrade the efficiency and cost-
effectiveness of this pretreatment method, making it a promising approach for
sustainable biomass conversion.

3.6.4.3 Fermentation Procedure for Production of PHB Ester

The obtained sugars from enzymatic hydrolysis are then used as a feedstock for
fermentation. Here is an overview of the fermentation process (Fig. 3.6):
Microorganism selection: The choice of microorganisms depends on the desired
product and process conditions. Bacterial strains capable of producing PHB
are selected for fermentation. Examples include Cupriavidus necator (formerly
Ralstonia eutropha) and other PHB-producing bacteria. These bacteria can naturally
convert glucose or other fermentable sugars into PHB as a storage material.
Fermentation medium: The sugars derived from LCB are typically present
in a hydrolysate solution, which contains various compounds, including sugars,
inhibitors, and nutrients. The hydrolysate may need to be treated to remove or
detoxify inhibitors and adjust the nutrient composition to support microbial growth
and product formation. The fermentation medium for PHB ester production typi-
cally contains the necessary nutrients and carbon sources to support the growth of
the PHB-producing bacteria and the accumulation of PHB within the cells. The
specific composition of the fermentation medium may vary based on the bacterial
strain used and the availability of LCB-derived sugars.
Fermentation process: The fermentation process involves the inoculation of the
selected microorganism into the fermentation medium containing the sugars from
LCB. Fermentation is typically carried out in bioreactors equipped with monitoring
and control systems.

3.6.4.4 Procedures for PHB Recovery

Once the fermentation is complete, the PHB must be separated from the broth [63].
The recovery of PHB from the fermentation broth involves several steps to separate
and purify the PHB from the bacterial cells and other impurities. Here is a general
outline of the PHB recovery process:
72 M. Maitra et al.

Cell harvesting: The first step is to separate the bacterial cells containing PHB
from the fermentation broth. This can be achieved through techniques such as
centrifugation, filtration, or both. The harvested cells are typically referred to as
wet biomass.
Cell disruption: Once the cells are harvested, they must be disrupted to release
the intracellular PHB granules. Cell disruption methods encompass mechanical
methods like high-pressure homogenization and bead milling, chemical methods
like treatment with enzymes or detergents, or a combination of different techniques.
PHB extraction: After cell disruption, the PHB granules are extracted from the
disrupted cells. Various extraction methods can be employed, including solvent
extraction, where a suitable organic solvent (e.g., chloroform, chloroform–methanol
mixture) dissolves the PHB while leaving other cellular components behind. The
extraction can be performed using soxhlet extraction, shaking or stirring with the
solvent, or continuous extraction methods.
PHB purification: Once the PHB is extracted, it needs to be purified to remove
impurities, such as residual cell debris, proteins, and other contaminants. Purifica-
tion methods can include precipitation, filtration, centrifugation, or chromatography
techniques. For example, the extracted PHB can be dissolved in a non-solvent (e.g.,
ethanol, methanol) to precipitate the PHB, followed by washing and drying steps to
obtain purified PHB [72].
Drying: The purified PHB is typically in the form of a wet cake or solid. It
needs to be dried to remove residual moisture and obtain dry PHB. Techniques like
freeze-drying, vacuum drying, or hot air drying can achieve drying.
PHB characterization: The recovered and dried PHB can be characterized using
various techniques to determine its molecular weight, thermal properties, crys-
tallinity, and other physical and chemical properties. This characterization helps
ensure the quality and suitability of the PHB for specific applications.

3.6.4.5 Transesterification for PHB Esters Production

A transesterification process can be employed to form a PHB ester from PHB. Trans-
esterification entails the substitution of one ester group with a different alcohol,
leading to the creation of a novel ester compound [73].
Preparation of reactants: The PHB polymer is depolymerized into monomeric
units, typically 3-hydroxybutyrate (3HB) monomers. The alcohol used for esterifi-
cation is usually a low molecular weight, like methanol or ethanol.
Reaction setup: The reaction is performed in a suitable reaction vessel, which may
be equipped with a condenser to prevent the loss of volatile components. The vessel
is often heated or maintained at a specific temperature to facilitate the reaction.
Catalyst addition: An acid or base catalyst is added to the reaction mixture
depending on the transesterification method. Common acid catalysts include sulfuric
acid (H2 SO4 ) or hydrochloric acid (HCl), while common base catalysts include
sodium hydroxide (NaOH) or potassium hydroxide (KOH). The catalyst helps to
accelerate the reaction and improve the yield of PHB esters.
3 Lignocellulosic Biomass for Sustainable Production of Renewable … 73

Fig. 3.6 Flow scheme for the production of PHB ester from LCB

Reaction initiation: The reaction mixture is heated or maintained at a specific


temperature, typically around 60–80 °C, to promote the transesterification reac-
tion. The reaction time can vary depending on the specific conditions and desired
conversion.
Monitoring the reaction: The progress of the transesterification reaction can be
monitored by sampling the reaction mixture at regular intervals and analyzing it
using strategies such as gas chromatography (GC) or nuclear magnetic resonance
(NMR) spectroscopy. This helps determine the conversion and optimize the reaction
conditions.
Purification: At the end of the reaction, the reaction mixture is cooled, and the PHB
ester product is separated from the remaining reactants, catalysts, and impurities.
In the presence of a 2% concentration of glucose and an available carbon source,
Alcaligenes eutrophus MTCC 1285 produces bioplastics. The synthesis of PHB has
been demonstrated employing molasses (pH 6.9–8) and sago effluent as feedstock. In
the manufacture of media based on molasses, higher output was achieved at pH 8.0.
The concentrations of PHB generated in sago, molasses, glucose, and thippi substrate-
based medium at pH 6.9 were 0.5, 0.73, 0.8, and 0.4 g/ml, respectively. The expenses
of producing biopolymers might be considerably minimized by harvesting leftovers
such as starch, damaged food grains, date seeds, dairy waste, and grain crops. The
efficient utilization of food and agricultural waste sources for the value-added manu-
facturing of biopolymers has been the focus of current research. Processing industry
byproducts are frequently disposed of as waste because their economic worth is lower
than their recovery cost and potential for reuse. These waste fluxes can be efficiently
valorized by producing and processing biopolymers [74].
74 M. Maitra et al.

3.6.5 Production of Second-Generation Biomethanol Using


LCB

Biomethane production from LCB typically involves a two-step process: pretreat-


ment and anaerobic digestion [75]. Figure 3.7 provides the overview of the steps
involved in biomethanol production:
Pretreatment approaches: LCB is often complex and resistant to enzymatic break-
down. Therefore, pretreatment is necessary to make biomass more accessible to
microbial digestion. Pretreatment methods can include physical, chemical, or biolog-
ical processes. The aim is to break down the complex lignin and hemicellulose
structures, making cellulose accessible for succeeding enzymatic hydrolysis.
Enzymatic hydrolysis procedure: After pretreatment, the LCB is subjected to enzy-
matic hydrolysis. Enzymes, such as cellulases and hemicellulases, are added to the
biomass to break down the cellulose and hemicellulose into simpler sugars, such as
glucose and xylose.
Anaerobic digestion: The hydrolyzed LCB is then made to undergo anaerobic
digestion. This microbial process occurs in anaerobic processes. The biomass is
placed in an anaerobic digester, a sealed vessel containing a consortium of microor-
ganisms called methanogens. These microorganisms break down the sugars into
biogas, primarily methane (CH4 ) and carbon dioxide (CO2 ).
Biogas upgrading: The biogas produced during anaerobic digestion contains
impurities, including CO2 , hydrogen sulfide (H2 S), and trace contaminants. In order
to transform the biogas into biomethane, which can be injected into the natural
gas grid or utilized as fuel for vehicles, biogas upgrading is performed. Common
upgrading techniques encompass water scrubbing, amine scrubbing, pressure swing
adsorption (PSA), and membrane separation. These techniques remove impurities
and increase the methane content in the biogas.
Gas compression and purification: The biomethane is then compressed to the
desired pressure for transportation or injection into a natural gas network. Further
purification may be carried out to meet the specific quality necessities of the gas grid.
The overall procedure followed to produce biomethane from LCB is known as
anaerobic digestion or biomethanation. It is a renewable and sustainable process
that produces biomethane and yields a nutrient-rich digestate that can be used as an
organic fertilizer. Biomethane can be utilized in combined heat and power (CHP)
systems or natural gas-fired power plants to create electricity and heat. It can also be
utilized as a fuel source in various industrial processes, including boilers, furnaces,
kilns, and cogeneration systems. Industries such as food and beverage, pulp and
paper, chemical, and pharmaceutical sectors can benefit from the use of biomethane,
reducing their carbon footprint and dependence on fossil fuels.
3 Lignocellulosic Biomass for Sustainable Production of Renewable … 75

Fig. 3.7 Biomethanol production

3.6.6 Lignocellulose-Derived High-Value Bioproducts

LCB acts as a valuable resource for the production of various high-value bioprod-
ucts [4]. Here are some examples of high-value bioproducts that can be derived from
LCB (Fig. 3.8):
Bio-based chemicals: LCB can be modified into a range of bio-based chemicals
through biochemical and thermochemical processes. For example, furan derivatives
like furfural and levulinic acid can be produced from hemicellulose, which has appli-
cations in the production of resins, plastics, and solvents. Other chemicals that can
be derived include organic acids, platform chemicals, and specialty chemicals.
Bio-based polymers: LCB can be utilized as raw materials for manufacturing bio-
based polymers. Polymers like cellulose acetate, cellulose esters, and lignin-based
polymers can be obtained from cellulose and lignin components of biomass. These
bio-based polymers have applications in packaging, textiles, coatings, and biomedical
materials, offering more sustainable alternatives to fossil fuel-based polymers.
Bio-based fuels and energy: Besides biofuels like bioethanol and biobutanol, LCB
can be used to produce advanced biofuels such as bio-jet fuel and biohydrogen.
These biofuels offer cleaner energy options, reducing greenhouse gas emissions and
dependence on fossil fuels. Additionally, LCB can be used for biomass gasification
or pyrolysis to produce syngas, bio-oil, or biochar, which can be utilized for the
creation of heat and electricity.
Specialty chemicals and natural products: LCB acts as a resource for the produc-
tion of various specialty chemicals and natural products. For example, lignin, a
complex polymer in biomass, can be chemically modified to produce high-value
chemicals like vanillin (a flavoring agent), phenolic compounds (used in adhesives
76 M. Maitra et al.

Fig. 3.8 Various bioproducts obtained from lignocellulose

and antioxidants), and aromatic compounds for fragrances. Extractives in biomass,


such as essential oils and plant-based extracts, can also be used in cosmetics, personal
care products, and flavors.
Nutraceuticals and functional ingredients: LCB can provide a source of bioac-
tive compounds and functional ingredients in the nutraceutical and pharmaceutical
industries. For instance, LCB can be processed to obtain dietary fibers, antioxidants,
prebiotics, and other bioactive compounds that offer health benefits. These ingre-
dients can be used in functional foods, dietary supplements, and pharmaceutical
formulations.

3.7 LCB Conversion: Impact on Techno-Economic


Viability

The technological-economic viability of manufacturing bioenergy from substrates


made from lignocellulosic materials is the foremost important element determining
how viable they are for business. The cost of biomass pretreatment is anticipated to
be between 19 and 22% of the total cost of a biofuel manufacturing chain. Several
authors have created techno-economic analyses of various pretreatments under the
notion of a biorefinery to clarify the profitable circumstances. The dwindling supply
of petroleum and gas and the adverse environmental implications of swift urbaniza-
tion and technological progression are two substantial challenges. Because renew-
able energy sources can be purchased easily, it is one of the biggest energy sources,
encompassing over 70% of the energy used worldwide. Nevertheless, rising elec-
tricity production and utilization coincide with elevated GHG emissions and particle
pollution levels. Employing thermochemical synthesis using technological advances,
the LCB feed can potentially be effectively altered into substantial-quality outcomes
such as bioenergy and biofuel [8]. Damartzis and Zabaniotou analyzed the extrac-
tion and production of biofuels of the subsequent generation concerning biomass
3 Lignocellulosic Biomass for Sustainable Production of Renewable … 77

leveraging thermal energy and indicated constraints as well as prospects in process


consolidation and execution [9].
The most important life cycle assessment is a widely enacted approach to eval-
uate the prospective ecological effects of specific goods by integrating quantita-
tive, qualitative, confirmable, and measurable statistics. The life cycle assessment
(LCA) inquiry has been split into four sections. The initial stage identifies objec-
tives and dimensions, shortcomings, context, and repercussions for the environment.
The following step is to undertake a life cycle supply, encompassing input materials
and energy, outcomes including products and byproducts, and release pollutants
such as wastewater, gas, and any harmful particle emissions. The next stage, also
depicted as life cycle inventory estimation, is an administrative approach that quan-
tifies the possible consequences of different aspects of an outcome. The culminating
phase, elucidation, is assessing the product-associated data, leading to conclusions
and suggestions for reducing environmental impact. The variety of lignocellulosic
substances being used, conversion technique effectivity, and maximum efficiency
must be explored to maximize the product yield and profit.
Several factors contribute to uncertainty in the case of using lignocellulosic feed-
stock, including various viable designs, a lack of understanding about how the feed-
stock market will behave in the long run in the situation when there is a highly
increased demand, ignorance of ethanol market prices, particularly the answer to
the big query, regarding how ethanol price and gasoline price interact, when ethanol
blend ratio in transportation fuel is high; and most importantly investors’ readiness
to commit to the manufacturing of subsequent generation bioethanol. Due to a lack
of access to trustworthy data and economic evaluations based on numerous process
assumptions, techno-economic analyses of the manufacture of cellulosic ethanol
have encountered challenges. Cellulosic bioenergy generation facilities could be
able to help with these problems. Nevertheless, before the state-of-the-art is broadly
disseminated to the community, it can take some time as industrial data and expe-
rience are proprietary and confidential. In an effort to reduce the cost of biofuel
production and hence investigate the minimum selling price (MESP), numerous
studies about the different bioenergy production have been conducted. Likewise,
numerous biorefinery facility approaches seek supplementary technological break-
throughs, financing, promotional activities, and lignocellulosic material valoriza-
tion, which turns off their respective commercial and economic viability positions.
Consequently, a comprehensive setup for the reputable and statistical assessment of
ecological sustainability must be constructed [76] (Table 3.1).
Table 3.1 Comparison between biobutanol, bioethanol, biodiesel, biomethane
78

Comparison Biobutanol Bioethanol Biodiesel Biomethane Citations


factor
Chemical C4 H10 O C2 H5 OH Various mono-alkyl esters CH4 [77]
formula
Energy content Higher energy content Lower energy content Comparable energy content to High energy content, [77]
compared to ethanol compared to gasoline diesel fuel similar to natural gas
Octane rating Higher octane rating, similar A lower octane rating may A lower cetane rating may No octane rating, not [78]
to gasoline require blending with require blending with petroleum relevant for gaseous fuel
gasoline for higher diesel for better combustion
performance
Vapor pressure Lower vapor pressure, less Higher vapor pressure, more Varies depending on feedstock Varies depending on [79]
volatile volatile and blend composition and blend
Density Higher density Lower density Higher density Lower density [77]
Combustion Combusts like gasoline Combusts like gasoline Combusts in a manner similar to Combusts like natural gas [78]
characteristics diesel fuel
Water content Low water content Higher water content requires Low water content Low water content [80]
dehydration before use
Raw materials LCB (crop residues, wood Sugar crops (corn, sugarcane, Vegetable oils (soybean, palm, Organic waste materials [72–74]
required chips, etc.) sugar beets) rapeseed, etc.) (agricultural waste, food
waste, etc.)
(continued)
M. Maitra et al.
Table 3.1 (continued)
Comparison Biobutanol Bioethanol Biodiesel Biomethane Citations
factor
Production 1. Pretreatment: Breakdown 1. Pretreatment: Extraction 1. Transesterification: Chemical 1. Anaerobic Digestion: [81–83]
process of LCB into fermentable of sugars from sugar crops reaction between vegetable Microbial
sugars 2. Fermentation: Conversion oils and alcohol (usually decomposition of
2. Enzymatic Hydrolysis: of sugars into ethanol methanol) to produce organic waste materials
Conversion of complex through microbial biodiesel and glycerol in an oxygen-free
sugars into simple sugars fermentation 2. Separation and Purification: environment
3. Fermentation: Conversion Removal of impurities from 2. Purification: Removal
of sugars into butanol biodiesel of impurities such as
through microbial 3. Glycerol Purification: moisture and hydrogen
fermentation Refining and purification of sulfide
glycerol byproduct 3. Upgrading: Remove
carbon dioxide and
other trace elements to
increase methane
concentration
(continued)
3 Lignocellulosic Biomass for Sustainable Production of Renewable …
79
Table 3.1 (continued)
80

Comparison Biobutanol Bioethanol Biodiesel Biomethane Citations


factor
Typical yield 8.4–12 g/L (Under the The yearly yield of LCB The production volume of India has 125 billion cubic [84, 85]
influence of different waste worldwide was around biodiesel fuel in India was 185 meters of yearly trash that
microorganisms. Differ 220 billion tons. The million liters in 2022 might be converted into
concerning various substrates following shows the yearly energy by producing
and sugar content) output of bioethanol using biomethane. This can
non-wood LCB (million produce 748.59 TWh of
tons): The following straws electric power and 4.49 EJ
are used to produce of heat energy
bioethanol: sugarcane bagasse
(1044.8), rice straw (657.5),
wheat straw (472.2), corn
straw (376.4), sorghum straw
(12.0), and oat straw (10.1).
Of these, sugarcane bagasse
appears to be an appealing
entity
Compatibility It can be blended with It can be blended with It can be used in diesel engines It can be used in natural gas [86]
gasoline in various gasoline in various with a few to no modifications vehicles (NGVs) or injected
proportions proportions into the natural gas grid
Fuel storage Less volatile and less prone to More volatile and requires Similar storage characteristics to The gaseous form allows [87]
evaporation compared to careful handling to minimize petroleum diesel for easier storage and
bioethanol evaporation distribution
Fuel efficiency Higher energy density can Lower energy density may Similar energy density to High energy content [87]
lead to improved fuel result in decreased fuel petroleum diesel, maintaining enables efficient
efficiency compared to efficiency compared to fuel efficiency combustion
bioethanol gasoline
(continued)
M. Maitra et al.
Table 3.1 (continued)
Comparison Biobutanol Bioethanol Biodiesel Biomethane Citations
factor
Co-product Byproduct glycerol can be Byproduct distillers’ grains Glycerol byproduct can be used Digestate byproduct can be [88]
utilization used in various industrial can be used as animal feed or in various applications used as a fertilizer or soil
applications in other applications amendment
Greenhouse Can contribute to less GHG Can contribute to less GHG Can contribute to less GHG Can contribute to less GHG [86, 89]
gas emissions emission in comparison to emission in comparison to emission in comparison to fossil emission in comparison to
fossil fuels fossil fuels fuels fossil fuels
Applications Fuel additive, gasoline blend Fuel additive, gasoline blend Diesel fuel substitute, heating oil Electricity generation, [90, 91]
substitute heating, transportation fuel
Advantages Higher energy density, lower Renewable, reduced GHG Renewable, reduced GHG Utilizes organic waste, [86, 89]
vapor pressure, compatibility emissions emissions reduces GHG emissions
with existing infrastructure
Challenges The complex production Competition with food Competitiveness with petroleum Feedstock availability, [92, 93]
process, cost competitiveness, production, limited feedstock diesel, feedstock availability purification process,
scalability availability infrastructure compatibility
3 Lignocellulosic Biomass for Sustainable Production of Renewable …
81
82 M. Maitra et al.

3.8 Challenges and Future Directions in Utilization


of Lignocellulose-Derived Feedstock

LCB means plant-derived biomass. Due to its abundant availability and potential as a
renewable resource, it is a promising feedstock for generating biofuels, biochemicals,
and bioproducts. Nevertheless, several challenges are linked with the usage of LCB,
and ongoing research is focused on addressing these challenges and exploring future
directions. Here are some key challenges and future directions in LCB [94].

3.8.1 Valorization of Lignin

Lignin, a complex aromatic polymer, is a major constituent of LCB. Currently, lignin


is underutilized and often treated as a low-value byproduct. Future research aims to
develop efficient lignin extraction and valorization processes, such as its conver-
sion into high-value chemicals, materials, and energy products. Finding innova-
tive applications for lignin can significantly enhance the overall economic viability
of lignocellulosic biorefineries. Lignin-infused polyurethanes develop cost-effective
composites and foams favorable to the environment, advancing efforts in drainage,
building, and soil stabilization.

3.8.2 Sustainability and Environmental Impact

As LCB is obtained from agricultural residues, energy crops, and forestry residues,
it is crucial to ensure sustainable sourcing practices and minimize the environ-
mental impact of biomass production and conversion. Future directions include
optimizing land use, minimizing water and energy requirements, reducing green-
house gas emissions, and implementing integrated biorefinery concepts to maximize
resource efficiency.
Techno-economic feasibility: For LCB to become commercially viable, it is essen-
tial to achieve cost competitiveness with fossil-based alternatives. Future efforts focus
on reducing the overall production costs through process optimization, scale-up, and
developing advanced biorefinery technologies. Improved efficiency and cost reduc-
tion in each step of the biomass conversion process are keys to achieving economic
feasibility.
3 Lignocellulosic Biomass for Sustainable Production of Renewable … 83

3.8.3 Integration of Biorefineries

Integrating biorefineries involves manufacturing multiple valuable products from


LCB by efficiently utilizing various components and byproducts. Future directions
emphasize the development of integrated biorefinery platforms that can simultane-
ously produce biofuels, biochemicals, materials, and energy, maximizing the overall
value derived from LCB. Also, genetic engineering and breeding techniques can
be employed to enhance the properties of lignocellulosic feedstocks, making them
more amenable to conversion processes. Future research aims to improve biomass
feedstocks’ yield, quality, and sustainability [95].

3.9 Conclusions

The findings outline several phases for processing lignocellulosic substrates into
feasible biofuels, including biobutanol, biomethane, PHB esters, etc. All the hydrol-
ysis, thermal, and catalytic routes will be crucial in potential biorefineries, where it
will ultimately be necessary to combine these methods to refine the entire biomass
to have the best possible yield. Resembling a petroleum refinery, a profitable biore-
finery may be formed by unifying a number of extraction processes. The econom-
ically feasible biorefinery will be able to derive a variety of different substances
as well as biofuels in this manner. The overarching objective of LC biomass, the
most abundant feedstock, intends to yield a substantial quantity of liquefied fuels to
meet the world’s energy requirements and produce high-value byproducts, utilizing
an assortment of technological innovations. With the amalgamation of plenty of
steps into an all-encompassing affordable, along with convenient interpretation,
consolidated bioprocessing illustrated tremendous possibilities in lignocellulosic
feedstock biorefinery. Scientists must concentrate on enhancing and streamlining
catalytic transformation procedures to ensure a firm basis for producing high-priced
bi-products such as HMF in tandem with ethanol. Regarding enforceable massive-
scale industries, extensive study has been focused on breakthroughs and convergence,
the implementation of possible sources, viable environment-friendly approaches, and
the formulation of techno-economic and financial frameworks to produce different
byproducts and lignocellulose-derived biofuels. Compared to the transesterification
process for biodiesel and fermentation for bioalcohol, more sophisticated manufac-
turing methods such as biomass to liquid (BTL) and Fischer–Tropsch synthesis tech-
nology for advanced-generation fuels are emerging at an alarming rate. Furthermore,
some properties have been significantly enhanced, which are oxidation durability and
thermal efficiency. It is possible to improve the production rates of second or third-
generation biofuels to a minimum of two to three times that of first-generation biofuel.
Specifically, advanced-generation biofuels offer a far greater potential to replace
fossil fuels in transportation than first-generation biofuels. An adequate combina-
tion of several feedstock sources is advised to formulate a competitive biofuel and
84 M. Maitra et al.

strike a reasonable balance among high fuel quality and cheap manufacturing prices.
Improved fuel economy and a higher chance of survival in the global fuel market
have been demonstrated by biofuels with higher life cycle energy efficiency (LCEE)
values. In order to foster the commercial potential of biofuel production process
byproducts like glycerol, more sophisticated conversion and purification methods
are also evolving quickly. From the extensive studies, it is inferred that supercrit-
ical alcohol technology and microwave radiation are fairly capable of speeding up
the rates at which feedstocks from biomass undergo conversion into fuel. Every
phase of producing different biofuels has certain drawbacks, either due to inhibitory
synthesis or whether the procedure is financially or insufficiently competitive. As a
result, constraints need to be resolved through optimized performance, and various
proceedings ought to be worked together in order to develop an overall profitable
biofuel manufacturing framework. For either the emergence of inhibitors or the fact
that it is not viable in the marketplace, almost every phase of the extraction process
has some negative aspects. Accordingly, optimization ought to be employed to fix
these issues, and these processes need to be incorporated to establish a system that
is typically efficient in yielding biofuels.

References

1. Yadav, A., Sharma, V., & Tsai, M. L. (2023). Development of lignocellulosic biorefineries for
the sustainable production of biofuels: Towards circular bioeconomy. Bioresource Technology,
381, 129145. ISSN 0960–8524.
2. Adewuyi, A. (2022). Underutilized lignocellulosic waste as sources of feedstock for biofuel
production in developing countries. Frontiers in Energy Research, 10, 741570.
3. Kumar, A., Makishah, N. H. A., Wen, Z., Gupta, G., Pandit, S., & Prasad, R. (2022). Recent
developments in lignocellulosic biofuels, a renewable source of bioenergy. Fermentation, 8(4),
161.
4. Isikgor, F. H., & Becer, C. R. (2015). Lignocellulosic biomass: A sustainable platform for the
production of bio-based chemicals and polymers. Polymer Chemistry, 6(25), 4497–4559.
5. Gassner, M., & Maréchal, F. (2013). Increasing efficiency of fuel ethanol production from
lignocellulosic biomass by process integration. Energy & Fuels
6. Irmak, S. (2019). Challenges of biomass utilization for biofuels’, biomass for bioenergy—
recent trends and future challenges. IntechOpen.
7. Nanda, S., Mohammad, J., Reddy, S. N., Kozinski, J. A., & Dalai, A. K. (2013). Pathways of
lignocellulosic biomass conversion to renewable fuels. Biomass Conversion and Biorefinery,
4(2), 157–191 (2013).
8. Inyang, V., Laseinde, O. T., & Kanakana, G. M. (2022). Techniques and applications of ligno-
cellulose biomass sources as transport fuels and other bioproducts. International Journal of
Low-carbon Technologies, 17, 900–909.
9. Yousuf, A., Pirozzi, D., & Sannino, F. (2020). Fundamentals of lignocellulosic biomass. In
Lignocellulosic biomass to liquid biofuels (pp. 1–15). Academic Press.
10. Ghaemi, F., Abdullah, L. C., & Ariffin, H. (2019). Lignocellulose structure and the effect on
nanocellulose production. In Lignocellulose for future bioeconomy (pp. 17–30). Elsevier.
11. Thota, S. P., Bag, P. P., Vadlani, P. V., & Belliraj, S. K. (2022). Plant biomass derived multi-
dimensional nanostructured materials: a green alternative for energy storage. Engineered
Science
3 Lignocellulosic Biomass for Sustainable Production of Renewable … 85

12. Adewuyi, A. (2022). Underutilized lignocellulosic waste as sources of feedstock for biofuel
production in developing countries. Frontiers in Energy Research, 10
13. Wu, W., Li, P., Huang, L., Wei, Y., Li, J., Zhang, L., & Jin, Y. (2023). The role of lignin structure
on cellulase adsorption and enzymatic hydrolysis. Biomass, 3(1), 96–107.
14. Cai, J., He, Y., Yu, X., Banks, S. A., Yang, Y., Zhang, X., Yu, Y., Liu, R., & Bridgwater, A. V.
(2017). Review of physicochemical properties and analytical characterization of lignocellulosic
biomass. Aston University, 76, 309–322.
15. Mankar, A. R., Pandey, A., Modak, A., & Pant, K. K. (2021). Pretreatment of lignocellulosic
biomass: A review on recent advances. Bioresource Technology, 334, 125235.
16. Baruah, J., Nath, B. C., Sharma, R., Kumar, S., Deka, R. C., Baruah, D. C., & Kalita, E.
(2018). Recent trends in the pretreatment of lignocellulosic biomass for Value-Added products.
Frontiers in Energy Research, 6.
17. Qin, Z., Iqbal, I., Riaz, F., Karadağ, A., & Tabatabaei, M. (2019). Different pretreatment
methods of lignocellulosic biomass for use in biofuel production. In IntechOpen eBooks.
18. Hassan, S. S., Williams, G. A., & Jaiswal, A. K. (2018). Emerging technologies for the
pretreatment of lignocellulosic biomass. Bioresource Technology, 262, 310–318.
19. Wahid, R., Hjorth, M., Kristensen, S., & Møller, H. (2015). Extrusion as pretreatment for
boosting methane production: Effect of screw configurations. Energy & Fuels, 29(7), 4030–
4037.
20. Wei, H., YingTing, Y., JingJing, G., Wenshi, Y., & Jun-Hong, T. (2017). Lignocellulosic biomass
valorization: Production of ethanol. In Elsevier eBooks (pp. 601–604)
21. Hernández-Beltrán, J. U., Lira, I. O. H., Cruz-Santos, M. M., Saucedo-Luevanos, A.,
Hernández-Terán, F., & Balagurusamy, N. (2019). Insight into pretreatment methods of ligno-
cellulosic biomass to increase biogas yield: Current state, challenges, and opportunities. Applied
Sciences, 9(18), 3721.
22. Jin, M., Slininger, P. J., Dien, B. S., Waghmode, S. B., Moser, B. R., Orjuela, A., Da Costa
Sousa, L., & Balan, V. (2015). Microbial lipid-based lignocellulosic biorefinery: Feasibility
and challenges. Trends in Biotechnology, 33(1), 43–54.
23. Sharma, H. K., Xu, C., & Qin, W. (2019). Biological Pretreatment of lignocellulosic biomass
for biofuels and bioproducts: An overview. Waste Biomass Valor, 10, 235–251.
24. Rezania, S., Oryani, B., Cho, J., Talaiekhozani, A., Sabbagh, F., Hashemi, B., Rupani, P. F., &
Mohammadi, A. A. (2020). Different pretreatment technologies of lignocellulosic biomass for
bioethanol production: An overview. Energy, 199, 117457
25. Tan, J., Wu, H., Li, H., & Yang, S. (2022). Chapter 7—Pretreatment methods for converting
straws into fermentable sugars. In H. Li, S. Saravanamurugan, A. Pandey, S. Elumalai (Eds.),
Biomass, biofuels, biochemical (pp. 117–162). Elsevier. ISBN 9780128244197
26. Baksi, S., Saha, D., Saha, S., et al. (2023). Pre-treatment of lignocellulosic biomass: Review
of various physico-chemical and biological methods influencing the extent of biomass depoly-
merization. International Journal of Environmental Science and Technology, 20, 13895–13922.
27. Tran, T. H., Le, T. P. Q., Phong, M. T., & Nguyen, D. H. (2020). Bioethanol production from
lignocellulosic biomass. In IntechOpen eBooks.
28. Kassaye, S., Pant, K. K., & Jain, S. (2017). Hydrolysis of cellulosic bamboo biomass into
reducing sugars via a combined alkaline solution and ionic liquid pretreament steps. Renewable
Energy, 104, 177–184.
29. Abbott, A. P., Capper, G., Davies, D. B., Rasheed, R. K., & Tambyrajah, V. (2002). Novel
solvent properties of choline chloride/urea mixtures. Chemical Communications, 1, 70–71.
(Electronic supplementary information (ESI) available: Spectroscopic data). See http://www.
rsc.org/suppdata/cc/b2/b210714g/
30. Sun, Y., & Cheng, J. (2002). Hydrolysis of lignocellulosic materials for ethanol production: A
review. Bioresource Technology, 83(1), 1–11.
31. Wang, T., Liu, X., Yu, Q., Zhang, X., Qu, Y., Gao, P., & Wang, T. (2005). Directed evolu-
tion for engineering pH profile of endoglucanase III from Trichoderma reesei. Biomolecular
Engineering, 22(1–3), 89–94.
86 M. Maitra et al.

32. Stanley, D., Fraser, S., Chambers, P., Rogers, P., & Stanley, G. (2009). Generation and charac-
terisation of stable ethanol-tolerant mutants of Saccharomyces cerevisiae. Journal of Industrial
Microbiology & Biotechnology, 37(2), 139–149.
33. Wati, L. (1996). Characterisation of genetic control of thermotolerance in mutants of Saccha-
romyces cerevisiae. Ulster University. https://pure.ulster.ac.uk/en/publications/characterisa
tion-of-genetic-control-of-thermotolerance-in-mutants-5
34. Akashdeep Dey and R. Camilla Thomson. (2022). India’s biomethane generation potential
from wastes and the corresponding greenhouse gas emissions abatement possibilities under
three end use scenarios: Electricity generation, cooking, and road transport applications. Royal
Society of Chemistry.
35. Sánchez, Ó. J., & Cardona, C. A. (2008). Trends in biotechnological production of fuel ethanol
from different feedstocks. Bioresource Technology, 99(13), 5270–5295.
36. Kumar, S., Dheeran, P., Singh, S., Mishra, I. M., & Adhikari, D. K. (2015). Continuous ethanol
production from sugarcane bagasse hydrolysate at high temperature with cell recycle and in-situ
recovery of ethanol. Chemical Engineering Science, 138, 524–530.
37. Olsson, L., & Hahn-Hägerdal, B. (1996). Fermentation of lignocellulosic hydrolysates for
ethanol production. Enzyme and Microbial Technology, 18(5), 312–331.
38. Suriyachai, N. (2017). Organosolv-based fractionation and conversion of bagasse to sugars and
phenolics.
39. Wang, Y., Guo, W., Cheng, C. L., Ho, S., Chang, J. S., & Ren, N. (2016). Enhancing bio-butanol
production from biomass of Chlorella vulgaris JSC-6 with sequential alkali pretreatment and
acid hydrolysis. Bioresource Technology, 200, 557–564.
40. Wen, Z., Wu, M., Lin, Y., Yang, L., Lin, J., & Cen, P. (2014). Artificial symbiosis for acetone-
butanol-ethanol (ABE) fermentation from alkali extracted deshelled corn cobs by co-culture
of Clostridium beijerinckii and Clostridium cellulovorans. Microbial Cell Factories, 13(1).
41. Gottumukkala, L. D., Binod, P., Valappil, S. K., Mathiyazhakan, K., Pandey, A., & Sukumaran,
R. K. (2013). Biobutanol production from rice straw by a non acetone producing Clostridium
sporogenes BE01. Bioresource Technology, 145, 182–187.
42. Murphy, J. D., Drielak, E., Varma, V., Muthusamy, S., & Aslam, M. (2017). Updates on
the pretreatment of lignocellulosic feedstocks for bioenergy production—A review. Biomass
Conversion and Biorefinery, 8(2), 471–483.
43. Kim, J. N., Lee, H., Lee, S., Jae, J., & Park, Y. (2019). Overview of the recent advances in ligno-
cellulose liquefaction for producing biofuels, bio-based materials and chemicals. Bioresource
Technology, 279, 373–384.
44. Pérez, J. M., Muñoz-Dorado, J., De La Rubia, T., & Martínez, J. A. (2002). Biodegradation
and biological treatments of cellulose, hemicellulose and lignin: An overview. International
Microbiology, 5(2), 53–63.
45. Xin, F., Wu, Y., & He, J. (2014). Simultaneous fermentation of glucose and xylose to butanol
by clostridium sp. strain BOH3. Applied and Environmental Microbiology, 80(15), 4771–4778.
46. Taherzadeh, M. J., & Taherzadeh, M. J. (2016). A critical review of analytical methods
in pretreatment of lignocelluloses: Composition, imaging, and crystallinity. Bioresource
Technology, 200, 1008–1018.
47. Amiri, H., & Taherzadeh, M. J. (2018). Pretreatment and hydrolysis of lignocellulosic wastes
for butanol production: Challenges and perspectives. Bioresource Technology, 270, 702–721.
48. Chandel, A. K., Garlapati, V. K., Singh, A. K., Antunes, F. A. F., & Da Silva, S. S. (2018).
The path forward for lignocellulose biorefineries: Bottlenecks, solutions, and perspective on
commercialization. Bioresource Technology, 264, 370–381.
49. Liu, W., Chen, W., Hou, Q., Wang, S., & Liu, F. (2018). Effects of combined pretreatment of
dilute acid pre-extraction and chemical-assisted mechanical refining on enzymatic hydrolysis
of lignocellulosic biomass. RSC Advances, 8(19), 10207–10214.
50. Kumar, P., Barrett, D. M., Delwiche, M. J., & Stroeve, P. (2009). Methods for pretreat-
ment of lignocellulosic biomass for efficient hydrolysis and biofuel production. Industrial &
Engineering Chemistry Research, 48(8), 3713–3729.
3 Lignocellulosic Biomass for Sustainable Production of Renewable … 87

51. De Albuquerque Wanderley, M. C., Martin, C., De Moraes Rocha, G. J., & Gouveia,
E. R. (2013). Increase in ethanol production from sugarcane bagasse based on combined
pretreatments and fed-batch enzymatic hydrolysis. Bioresource Technology, 128, 448–453.
52. Tsai, T. Y., Lo, Y. C., Dong, C. D., Nagarajan, D., Chang, J. S., & Lee, D. J. (2020). Biobu-
tanol production from lignocellulosic biomass using immobilized Clostridium acetobutylicum.
Applied Energy, 277, 115531. ISSN 0306-2619
53. Ageitos, J. M., Vallejo, J. A., Veiga-Crespo, P., & Villa, T. G. (2011). Oily yeasts as oleaginous
cell factories. Applied Microbiology and Biotechnology, 90(4), 1219–1227.
54. Pandey, A., Soccol, C. R., & Mitchell, D. L. (2000). New developments in solid state
fermentation: I-bioprocesses and products. Process Biochemistry, 35(10), 1153–1169.
55. Leung, D. Y., Wu, X., & Leung, M. (2010). A review on biodiesel production using catalyzed
transesterification. Applied Energy, 87(4), 1083–1095.
56. Sawangkeaw, R., & Ngamprasertsith, S. (2013). A review of lipid-based biomasses as
feedstocks for biofuels production. Renewable & Sustainable Energy Reviews, 25, 97–108.
57. Kim, S., & Dale, B. E. (2004). Global potential bioethanol production from wasted crops and
crop residues. Biomass & Bioenergy, 26(4), 361–375.
58. Mosier, N. S., Wyman, C. E., Dale, B. E., Elander, R. T., Lee, Y., Holtzapple, M. T., & Ladisch,
M. R. (2005). Features of promising technologies for pretreatment of lignocellulosic biomass.
Bioresource Technology, 96(6), 673–686.
59. Hendriks, A. T. W. M., & Zeeman, G. (2009). Pretreatments to enhance the digestibility of
lignocellulosic biomass. Bioresource Technology, 100(1), 10–18.
60. Ericsson, K., Rosenqvist, H., & Nilsson, L. (2009). Energy crop production costs in the EU.
Biomass & Bioenergy, 33(11), 1577–1586.
61. Hossain, A., & Badr, O. (2007). Prospects of renewable energy utilisation for electricity
generation in Bangladesh. Renewable & Sustainable Energy Reviews, 11(8), 1617–1649.
62. Limayem, A., & Ricke, S. C. (2012). Lignocellulosic biomass for bioethanol production:
Current perspectives, potential issues and future prospects. Progress in Energy and Combustion
Science, 38(4), 449–467.
63. Kadam, K. L., & McMillan, J. R. (2003). Availability of corn stover as a sustainable feedstock
for bioethanol production. Bioresource Technology, 88(1), 17–25.
64. Kim, K. H., & Hong, J. (2001). Supercritical CO2 pretreatment of lignocellulose enhances
enzymatic cellulose hydrolysis. Bioresource Technology, 77(2), 139–144.
65. Influence of the size reduction of organic waste on their anaerobic digestion (2000). PubMed.
66. Balat, M., Balat, H., & Öz, C. (2008). Progress in bioethanol processing. Progress in Energy
and Combustion Science, 34(5), 551–573.
67. Ariunbaatar, J., Panico, A., Esposito, G., Pirozzi, F., & Lens, P. N. (2014). Pretreatment methods
to enhance anaerobic digestion of organic solid waste. Applied Energy, 123, 143–156.
68. Garrote, G., Domínguez, H., & Parajó, J. C. (1999). Hydrothermal processing of lignocellulosic
materials. European Journal of Wood and Wood Products, 57(3), 191–202.
69. Gossett, J. M., Stuckey, D. C., Owen, W. F., & McCarty, P. L. (1982). Heat treatment and
anaerobic digestion of refuse. Journal of the Environmental Engineering Division, 108(3),
437–454.
70. Behera, S., Arora, R., Nandhagopal, N., & Kumar, S. (2014). Importance of chemical
pretreatment for bioconversion of lignocellulosic biomass. Sciencedirect, 36, 91–106.
71. Samir, A., Ashour, F. H., Hakim, A. A., & Bassyouni, M. (2022). Recent advances in
biodegradable polymers for sustainable applications. Npj Mater Degrad 6, 68
72. Zhang, Y., Wang, H., Sun, X., Wang, Y., & Liu, Z. (2021). Separation and characterization
of biomass components (cellulose, hemicellulose, and lignin) from corn stalk. BioResources,
16(4), 7205–7219.
73. Rocha-Meneses, L., Ivanova, A., Atouguia, G., Ávila, I., Raud, M., Orupõld, K., & Kikas, T.
(2019). The effect of flue gas explosive decompression pretreatment on methane recovery from
bioethanol production waste. Industrial Crops and Products, 127, 66–72.
74. Dao, C. N., Mupondwa, E., Tabil, L., Li, X., & Castellanos, E. C. (2018). A review on techno-
economic analysis and life-cycle assessment of second generation bioethanol production via
biochemical processes. The Canadian Society for Bioengineering, 22–25.
88 M. Maitra et al.

75. Zhang, Q., Dong, J., Liu, Y., Wang, Y., & Cao, Y. (2016). Towards a green bulk-scale biobutanol
from bioethanol upgrading. Journal of Energy Chemistry, 25(6), 907–910.
76. Tagami-Kanada, N., Yoshikuni, K., Mizuno, S., Sawai, T., Fuchihata, M., & Ida, T. (2022).
Combustion characteristics of densified solid biofuel with different aspect ratios. Renewable
Energy, 197, 1174–1182.
77. Elkatory, M. R., Hassaan, M. A., & Nemr, A. E. (2022). Algal biomass for bioethanol and
biobutanol production. In Elsevier eBooks (pp. 251–279).
78. Kulbeik, T., Scherzinger, M., Höfer, I., & Kaltschmitt, M. (2021). Autoclave pre-treatment of
foliage—Effects of temperature, residence time and water content on solid biofuel properties.
Renewable Energy, 171, 275–286.
79. Mujtaba, M., Fraceto, L. F., Fazeli, M., Mukherjee, S., Savassa, S. M., De Medeiros, G. A.,
Pereira, A. D. E. S., Mancini, S. D., Lipponen, J., & Vilaplana, F. (2023). Lignocellulosic
biomass from agricultural waste to the circular economy: A review with focus on biofuels,
biocomposites and bioplastics. Journal of Cleaner Production, 402, 136815.
80. Cao, G., & Sheng, Y. (2016). Biobutanol production from lignocellulosic biomass: Prospective
and challenges. Journal of Bioremediation and Biodegradation, 7(4).
81. Awogbemi, O., & Von Kallon, D. V. (2023). Application of biochar derived from crops residues
for biofuel production. Fuel Communications, 15, 100088.
82. Eloka-Eboka, A. C., Maroa, S., & Taiwo, A. E. (2023). Biobutanol from agricultural and
municipal solid wastes, techno-economic, and lifecycle analysis. In Elsevier eBooks (pp. 171–
198).
83. Basak, B., Kumar, R., Bharadwaj, A. S., Kim, T. H., Kim, J. R., Jang, M., Oh, S., Roh, H., & Jeon,
B. (2023). Advances in physicochemical pretreatment strategies for lignocellulose biomass
and their effectiveness in bioconversion for biofuel production. Bioresource Technology, 369,
128413.
84. Ma, S., Li, Y., Li, J., Yu, X., Cui, Z., Yuan, X., Zhu, W., & Wang, H. (2022). Features of single
and combined technologies for lignocellulose pretreatment to enhance biomethane production.
Renewable & Sustainable Energy Reviews, 165, 112606.
85. Dempfle, D., Kröcher, O., & Luterbacher, J. S. (2021). Techno-economic assessment of
bioethanol production from lignocellulose by consortium-based consolidated bioprocessing
at industrial scale. New Biotechnology, 65, 53–60.
86. Gurunath, R. B., Gobinath, R., & Giridhar, P. V. (2022). Bioethanol production from lignocel-
lulosic biomass: Past, present and future trends. Research Journal of Biotechnology 17(10),
124–132.
87. Kumara, B., Raghavendra, C., Naik, S., Kumar, M. D., Shadambi, C. B., & Cholachagudda, N.
A. (2023). Performance study of Neem/Soyabean biofuels on 4-stroke diesel engine. Materials
Today: Proceedings, 92, 406–412.
88. Krishnan, M. G., Rajkumar, S., Thangaraja, J., & Devarajan, Y. (2023). Exploring the synergistic
potential of higher alcohols and biodiesel in blended and dual fuel combustion modes in diesel
engines: A comprehensive review. Sustainable Chemistry and Pharmacy, 35, 101180.
89. Gnansounou, E., & Raman, J. K. (2018). Environmental performances of coproducts. Appli-
cation of Claiming-Based Allocation models to straw and vetiver biorefineries in an Indian
context. Bioresource Technology, 262, 203–211.
90. Becerra-Ruiz, J. S., González-Huerta, R., Gracida, J., Amaro-Reyes, A., & Macias-Bobadilla,
G. (2019). Using green-hydrogen and bioethanol fuels in internal combustion engines to reduce
emissions. International Journal of Hydrogen Energy, 44(24), 12324–12332.
91. Mehmood, T., Nadeem, F., Meer, B., Ashraf, H., Meer, K., & Saeed, S. (2023). Biomass
valorization to biobutanol. In Elsevier eBooks (pp. 151–178).
92. Siegrist, A., Bowman, G., & Burg, V. (2022). Energy generation potentials from agricul-
tural residues: The influence of techno-spatial restrictions on biomethane, electricity, and heat
production. Applied Energy, 327, 120075.
93. Lin, C. Y., & Lu, C. (2021). Development perspectives of promising lignocellulose feedstocks
for production of advanced generation biofuels: A review. Renewable and Sustainable Energy
Reviews, 136, 110445. ISSN 1364-0321
3 Lignocellulosic Biomass for Sustainable Production of Renewable … 89

94. Zhang, R., Gao, H., Wang, Y., He, B., Lu, J., Zhu, W., Peng, L., & Wang, Y. (2023). Challenges
and perspectives of green-like lignocellulose pretreatments selectable for low-cost biofuels and
high-value bioproduction. Bioresource Technology, 369, 128315.
95. Yoo, C. G., Meng, X., Pu, Y., & Ragauskas, A. J. (2020). The critical role of lignin in ligno-
cellulosic biomass conversion and recent pretreatment strategies: A comprehensive review.
Bioresource Technology, 301, 122784.
Chapter 4
Industrial Organic Waste
and Byproducts as Sustainable Feedstock
for Bioenergy Production

Desta Getachew Gizaw, Selvakumar Periyasamy, Zinnabu Tassew Redda,


and Gurunathan Baskar

Abstract Bioenergy production has developed as a sustainable solution to address


the increasing energy demand while decreasing environmental pollution. The indus-
trial sectors are increasing along with the fast population growth and urban develop-
ments. Due to various activities, a significant amount of organic waste and byproducts
are generated in the industrial sectors. If improperly processed, this organic waste
with high nutritional content might cause environmental contamination and speed up
climate change. However, with suitable technologies and procedures, waste materials
could be used to produce valuable bioenergy sources. Converting industrial organic
waste and byproducts into bioenergy is a sustainable and environmentally friendly
approach. Besides, it offers several eco-economic benefits, such as reducing reliance
on fossil fuels, mitigating environmental pollution and reducing landfilling. This
chapter explores the potential of industrial organic waste and byproducts as feedstock
for bioenergy generation and its implications for sustainable energy production.

Keywords Organic waste · Byproducts · Industries · Feedstocks · Biofuels

D. G. Gizaw · S. Periyasamy (B)


Department of Chemical Engineering, School of Mechanical, Chemical and Materials
Engineering, Adama Science and Technology University, 1888 Adama, Ethiopia
e-mail: selvaa26kumar@gmail.com
Z. T. Redda
University of Applied Sciences (HTW) Berlin, Wilhelminenhofstraße 75A, 12459 Berlin,
Germany
School of Chemical and Bio Engineering, Addis Ababa Institute of Technology, Addis Ababa
University, King George VI St., P.O. Box 385, Addis Ababa, Ethiopia
G. Baskar
Department of Biotechnology, St. Joseph’s College of Engineering, Chennai 600119, India
S. Periyasamy
Department of Biomaterials, Saveetha Dental College and Hospitals, SIMATS, Saveetha
University, Chennai 600077, India

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2024 91
G. Baskar et al. (eds.), Circular Bioeconomy Perspectives in Sustainable Bioenergy
Production, Energy, Environment, and Sustainability,
https://doi.org/10.1007/978-981-97-2523-6_4
92 D. G. Gizaw et al.

4.1 Introduction

Due to rising population growth and changes in living standards, several industries
have increased their production capacities to keep up with demand. The number of
various organic wastes produced worldwide has dramatically increased as a result
of this [37]. Nearly 50% of the average composition of global waste, 3 M tons
per day, is organic material [23]. Significant levels of carbohydrates, lipids, proteins,
inorganic compounds, and a lot of water are present in many industrial and municipal
organic waste and byproduct streams, mostly from industries connected to the food
industry [20, 104]. Along with the dearth of virgin resources, environmental issues
related to treating and disposing organic-rich waste have long been a cause of worry
[56]. For example, the juice industry produced a huge amount of organic waste as
peels, the coffee industry produced coffee pulp as waste, and the cereal industry
produced husks. Therefore, these organic-rich waste must be treated efficiently to
convert them into sustainable bioenergy and value-added products and mitigate their
environmental impact.
When industrial organic wastes are disposed of in landfills, they release methane,
which is 34 times more potent than carbon dioxide in causing climate change and
global warming [14]. As a result, it is essential to create effective waste-to-energy
systems to improve energy recovery from the industrial organic wastes that are now
disposed of in landfills. Even though more waste is being kept out of landfills as
a whole, there is not much energy being extracted from it. Based on the need to
reduce our carbon footprint [77, 86] and the security of our fuel supply [79], high-
energy fuels such as renewable diesel [76, 107], bioethanol [25], and biogas are
being developed. Energy and value-added products recovery from organic waste
plays an important role in the urban waste hierarchy, making better use of waste, and
meeting government targets. Researchers around the world are developing different
approaches to converting organic waste into value-added products. In this regard,
industrial organic waste is an interesting and valuable renewable source of biomass
due to its biodegradability, water, and organic content, absence of competition with
food, the fact that it is unlimited if used sustainably, and, in particular, because of its
relatively high-energy density.
There are numerous technical options for generating renewable energy from
industrial organic waste. Direct combustion, anaerobic digestion, transesterification,
hydrothermal liquefaction, fermentation, pyrolysis, and thermal gasification are some
of the technical processes that enable the production of renewable energy [9, 83].
Direct burning is the most popular method for converting trash into heat or power.
In this procedure, waste-derived fuel is burned to provide energy in the presence
of additional oxygen (oxygen that gathers from the air). At high temperatures and
under controlled oxygen conditions, waste gasification transforms organic materials
into syngas composed of CO, H2 , CO2 , N2 , CH4 , etc. [35]. According to [1, 64],
pyrolysis is the thermochemical breakdown process of organic or inorganic waste at
a high temperature (430 °C) without oxygen. The process by which organic waste
4 Industrial Organic Waste and Byproducts as Sustainable Feedstock … 93

is broken down to produce biogas and biofertilizers is known as anaerobic diges-


tion. To degrade organic material, microbes must go through many activities without
oxygen. This can be applied to managing industrial, agricultural, or domestic waste
and creating clean fuels. This chapter provides a comprehensive idea of industrial
organic waste and their conversion methods to produce sustainable bioenergy.

4.2 Industrial Organic Wastes as a Feedstock

Finding a sustainable solution to the world’s rising energy consumption is now vital
because of population growth. The utilization of industrial organic waste for energy
production appears to be a solution to provide energy and lower carbon dioxide emis-
sions, as fossil fuels are scarce and non-renewable resources. Energy crops, waste
products, and other biological resources that can be utilized to produce renewable
energy and other valuable products are all included in the term biomass [38, 108].
Agricultural products like corn, sugarcane, soybean oil, and sunflower are used to
make first-generation biofuels [8, 106]. Because these biomasses are also used as
food, the phrase “food versus fuel” has been coined to describe the dilemma. Addi-
tionally, emissions of greenhouse gases (GHG) are believed to be lesser for second-
generation biofuels than first-generation fuels. For these reasons, industrial organic
wastes have gained attention due to their disponibility, including residues from the
crop, food, and oil industries.
Since they do not directly compete with food crops, industrial organic waste
streams represent a sustainable alternative to fossil-based resources. The term
“waste” refers to any organic substance that is not the primary material for which the
plants were cultivated. It also includes any byproduct derived from biomass for which
supply is significantly more than demand [23]. The food processing industries, such as
the juice, chips, meat, confectionary, and fruit industries, produce enormous amounts
of organic waste and associated effluents each year [12, 36, 100]. Different energy
sources can be made from these organic wastes. Industrial organic wastes come from
many industries and are as diverse as their origins as a result. For instance, peels,
seeds, and other organic waste are frequently produced as waste at fruit juice factories
[98]. Industry liquid byproducts like brine from pickling cucumbers and whey from
producing cheese are both frequent and act as substrates for different conversions.
Animal and post-harvest waste from agriculture is used to produce bioenergy [74,
126]. However, proteins, lipids, carbohydrates, and other components make organic
waste a viable alternative for producing lower-cost and more valuable chemicals [90,
138]. Figure 4.1 shows different types of organic waste.
Sawdust, wood chips, and abandoned logs are examples of organic wastes from
the wood processing industry that can be used as biofuel feedstocks [47, 110]. For
instance, the wood waste and sawdust produced by the paper and saw industries can be
used as boiler fuels and feedstock for ethanol manufacturing. The term “straw” refers
to the byproducts or residues left over after harvesting food crops, such as rice, wheat,
corn, beans, cotton, and sugar cane. The capability to convert corn stover, including
94 D. G. Gizaw et al.

Fig. 4.1 Different types of industrial organic wastes

the stalks, cobs, and leaves, into fermentable sugars for biobutanol generation has
also been noted [131]. For the economically viable usage of residual substrates, the
sugarcane residues, in particular sugarcane bagasse, can be a good choice for the
manufacture of bioethanol and other biofuels like biochar [15, 65].
Besides milling, malting, pearling, fermentation, heat processing, extrusion, and
puffing, there are numerous other ways to process grains used in the food industry.
Bran, germ, hull, husk, fiber, and other undesirable crop elements are produced in
considerable quantities when grains are processed using various processing tech-
niques [88]. Since all grains are preferred for consumption after milling, milling
industries account for the majority of the wastage of cereal crops [116]. The main
byproduct of grain milling is bran, which separates from the grain endosperm. It has
a rough outer shell made of coarse seed with little flour and a high protein and fiber
content. Another significant byproduct created during the wet milling of corn and
sorghum is a germ whose main component is gluten, which is high in protein [32].
On top of that, there is a significant potential for producing sustainable bioenergy
from these organic industrial wastes.
After the oil is extracted from seeds, a significant amount of processed organic
residues—known as oil cakes—are created, particularly in the oil industry [125].
Due to the high concentration of fat, oil, grease, suspended solids, and dissolved
solids in these residues, these organic wastes pollute the air, water, and solid waste.
Canola oil cake, sunflower oil cake, coconut oil cake, sesame oil cake, mustard oil
cake, palm kernel cake, soybean cake, groundnut cake, cotton seed cake, olive oil
4 Industrial Organic Waste and Byproducts as Sustainable Feedstock … 95

cake, and rapeseed cake are only a few examples of the various types of oil cakes
[102, 115]. These agro-industrial organic wastes are studied because they are very
inexpensive, contain a large number of elements, and have an endless potential to be
used as substitute substrates for fermentation.
On the other hand, enormous quantities of organic waste from the meat, poultry,
and fish processing industries are accumulating and contain perishable proteins that
may irritate people through an offensive odor [113]. According to [99], non-food-
based agro-industries produce effluents that are primarily biodegradable. Except for
the textile and tannery sectors, the effluents are non-toxic. The bulk of these wastes
are not used or handled, endangering the environment as well as the health of people
and animals. To produce biofuel, it is necessary to use these industrial organic wastes.

4.3 Bioenergy Conversation Techniques

Due to the current state of the economy and environment, there is an increasing
demand for recycling and energy conservation. Diverse technologies have been
developed and tapped into to use industrial organic waste for bioenergy produc-
tion. The method used to convert these organic wastes to energy entails turning
waste material into a variety of fuels that can be used to generate energy. Recent
years have seen an increase in the use of environmentally friendly methods to extract
and transform organic waste into chemical fuels [6, 45]. The conversion of organic
waste to energy is accomplished using one of two primary methods: thermochemical
or biochemical conversion. Various organic waste conversion methods, products,
process short descriptions, and environmental effects are presented in Table 4.1.
Biochemical conversion uses microbes or enzymes to transform organic waste into
usable energy, as opposed to thermochemical conversion, which uses heat to decom-
pose the waste’s organic components [73]. The thermochemical conversion methods
include combustion, gasification, liquefaction, and pyrolysis. The three processes of
anaerobic digestion, alcoholic fermentation, and photobiological reaction fall under
the category of biochemical conversion, on the other hand [96].

4.3.1 Thermochemical Conversion

Through the use of thermochemical technology, industrial organic waste is converted


into biochar (a solid), synthesis gas, and highly oxygenated bio-oil (a liquid). This
process needs bond-breaking and high-temperature chemical reformation. Gasifi-
cation, pyrolysis, and liquefaction are the three primary process options available
within thermochemical conversion [82]. The type and quantity of organic waste, the
selected energy source, end-use conditions, environmental principles, and financial
considerations can all impact the chosen conversion type [5]. According to several
research studies, thermal conversion technologies have attracted more attention as a
96 D. G. Gizaw et al.

Table 4.1 Organic waste conversion methods, products, process short descriptions, and environ-
mental effects
Conversion Products Descriptions Environmental References
methods effects
Pyrolysis Biochar, Thermal decomposition Higher impact [31, 87]
bio-oil and of biomass/organic waste on climate
syngas at higher temperatures in change
the absence of oxygen
Gasification Biohydrogen Produces gas molecules Lower GHG [112, 121]
and producer directly from organic emissions
gas waste at higher Lower impact
temperature on climate
change
Liquefaction Bio-oil Absence of oxygen Higher impact [4, 66]
during the on climate
thermochemical change
treatment of organic Higher energy
waste consumption
Anaerobic Biogas and Technology for AD of Lower impact [69, 117, 122]
digestion (AD) biomethane organic waste is used in a on climate
one- or two-stage method change
Lower carbon
emission and
GHG
Transesterification Biodiesel Extracted oil from Lower impact [71, 117]
organic waste converted on air emission
into biodiesel at optimal compared to
conditions incineration
Higher energy
demand
Dark fermentation Biohydrogen Microorganisms are used Lower GHG [121]
(DF) in a single-stage process emissions
to convert organic waste
into biohydrogen in the
absence of light and
oxygen

result of the availability of an industrial infrastructure to provide highly developed


thermochemical transformation equipment, a short processing time, and minimal
water use [95]. Additionally, for production, thermochemical conversion is unaf-
fected by ambient conditions. Therefore, it is essential to understand the various
thermochemical process choices to evaluate their potential in the future.
4 Industrial Organic Waste and Byproducts as Sustainable Feedstock … 97

4.3.1.1 Pyrolysis

For the production of biofuel from industrial organic waste, pyrolysis is frequently
used. It is a thermal degradation process that occurs between 400 and 900 °C without
the presence of oxygen [46]. When heated to a high temperature during the pyrolysis
process, the polymeric components of macromolecular structures are broken down
into smaller molecules and hydrocarbons, the final products. These final products
include solid coke (char), pyrolysis gas (syngas), and pyrolysis oil (bio-oil or tar).
This technique converts organic wastes into liquid fuel [41]. About 80% of the organic
waste is made up of volatile substances, with the remaining 15% being carbon and
5% being ash. Pyrolysis is a relatively new method for turning organic waste into
valuable products. Due to the inert atmosphere, pyrolysis plants emit less sulfur
and nitrogen oxides than other types of energy production [134]. Additionally, it
generates excellent solid residues. Figure 4.2 demonstrates different types of the
pyrolysis process.
Slow pyrolysis, fast pyrolysis, flash pyrolysis, and catalytic pyrolysis are the four
major types of pyrolysis. Table 4.2 shows different types of pyrolysis methods. Slow
pyrolysis is the decomposition of industrial organic wastes at a slow heating rate,
which results in greater char generation and less pyrolysis oil and gas production.
More pyrolysis oil is made possible by quick pyrolysis [135]. The heating rate is

Fig. 4.2 Illustration of different types of pyrolysis process


98 D. G. Gizaw et al.

Table 4.2 Different types of pyrolysis methods and their major features [28, 33]
Pyrolysis methods Major product Temperature (°C) Retention time Heating rate (°C/
min)
Slow pyrolysis Biochar, bio-oil, 550–600 5–30 min 5–7
syngas
Fast pyrolysis Bio-oil 550–650 1–5 s 300
Flash pyrolysis Bio-oil 450–650 <1s < 600
Catalytic pyrolysis Bio-oil 400–600 Hours–days –
Plasma pyrolysis Biochar > 1500 <1s > 1000

extremely high during flash pyrolysis. Only a few seconds or even less pass between
actions [21]. Flash pyrolysis primarily aims to increase the output of pyrolysis gas
and oil. According to [40], pyrolysis oil produced by slow, quick, and flash pyrolysis
procedures cannot be used as fuel for transportation in its pure form. These oils are
less miscible than fossil fuels because they contain a lot of water and oxygen. These
oils need to be upgraded before being used as fuel for transportation. On the other
hand, catalytic pyrolysis improves the pyrolysis oil’s quality [141]. In the pyrolysis
process, raising the temperature and adding a catalyst enhance the gas yield while
lowering the char yield. Since pyrolysis oil has a high moisture content, it cannot be
utilized directly as fuel. The quality of the pyrolysis oil can be improved via quick
pyrolysis.

4.3.1.2 Gasification

Industrial organic waste is heated during the gasification process along with the
gasifying chemicals to create syngas, as represented in Fig. 4.3. In the process of
gasification, organic wastes are transformed into syngas, a mixture of CO, CO2 ,
CH4 , NH3 , HCN, and H2 , at a high temperature (500–1800 °C), while being exposed
to a controlled amount of oxygen [48]. Air, oxygen, carbon dioxide, or steam are
used as gasification agents in a gasification process. The gasifying agent and the
ratio of gasifying agent to organic waste are carefully determined to achieve the
requisite chemical composition of syngas and maximize gasification efficiency. The
fuel from these syngas can be used to generate heat and/or power. It can also be
used as feedstock to create liquid fuel using the Fischer–Tropsch process, electricity
using a gas turbine or fuel cell, or chemical compounds like ammonia, hydrogen,
and methanol using the fuel cell [97, 111, 137].

4.3.1.3 Hydrothermal Carbonization

In pyrolysis, organic waste is heated without oxygen to break down into usable
compounds. The same procedure is known as hydrothermal carbonization or hydrous
4 Industrial Organic Waste and Byproducts as Sustainable Feedstock … 99

Fig. 4.3 Conversion of


organic waste into syngas
through gasification

pyrolysis if it is carried out in the presence of subcritical water. According to [62,


109], the main benefit of hydrothermal carbonization is that it may be utilized to effi-
ciently transform wet organic materials into carbonaceous solids, i.e., hydrochar.
It does not need to dry the materials before or during the procedure because it
can convert most input materials. The conversion of feedstocks, including indus-
trial processing organic wastes, human waste, wet animal manures, sewage sludges,
municipal solid waste, aquaculture residues, and algal residues, are examples of the
types of feedstocks that this approach may be used for. In high-pressure reactors
used for the hydrothermal carbonization process, water preserved in the liquid phase
surrounds the solid organic matter [27, 54]. The reaction temperature often governs
the kind and quantity of generated products. When the process temperature is kept
below 220 °C and matching pressures up to 20 bar are maintained, the majority of
the organic matter is transformed into solids (hydrochar) with a very small amount
of gas (1–5%). The process of producing more liquid hydrocarbons and gas fractions
at greater temperatures up to 400 °C is known as hydrothermal liquefaction [29].
When used as a solid fuel, hydrochar from industrially processed food wastes
produces more energy than the gases made when waste is converted through anaer-
obic digestion and incineration. At a relatively low temperature (180–350 °C) during
hydrothermal carbonization, carbohydrates are transformed into a carbonaceous
residue (hydrochar) [70]. Hydrochar is useful as a soil amendment, feedstock for
carbon fuel cells, and a good absorber for hazardous contaminants. This method is
particularly well suited for handling up to 90% wet organic waste. So, industrial
organic wastes can be converted effectively using this technology.
100 D. G. Gizaw et al.

4.3.2 Biochemical Conversion

4.3.2.1 Anaerobic Digestion

Industrial organic wastes are broken down naturally in landfills by microbes into
biogas, which is composed primarily of methane, carbon dioxide, and trace amounts
of other gases. This process is known as anaerobic digestion. However, this biogas
escapes and contaminates the atmosphere because it is created in an unregulated
environment (with oxygen present). The same procedure may be used to produce
biogas and other useful products if imitated in a controlled setting (without oxygen)
[24, 139]. The four steps of the anaerobic degradation process are hydrolysis, acido-
genesis, acetogenesis, and methanogenesis, as indicated in Fig. 4.4. By extracellular
enzymes and fermentative bacteria, high molecular weight organic matter (carbohy-
drates, protein, and lipids) is first broken down into tiny molecular solubility organic
matter (long chain fatty acids, glucose, and amino acids) through the hydrolysis stage
[48]. Small molecular organic matter is broken down during the acidogenesis stage
into volatile fatty acids (acetate, propionate, and butyrate) and additional byprod-
ucts such as NH3 , CO2 , and H2 S. The organic materials created in the acidogenesis
stage are subsequently digested into acetogenesis, where they produce acetate, H2 ,
CO2 , and other gases that are then utilized by methanogens to produce biogas [103].
The final products of anaerobic digestion are nitrogen-rich organic waste that can be
utilized as fertilizer and biogas that contains 60–70% methane. This technique has
been used in Europe and Asia to treat industrial organic wastes, agricultural wastes,
and wastewater sludge.

Classification of the AD Process of Organic Waste

The biological, technological, dependability, and overall performance of the AD


process are often used to categorize it [72]. Depending on the total amount of solids,
the AD process is classified as either dry or wet, according to the feeding method, it
is classified as either batch or continuous; according to the operating temperature, it
is classified as either mesophilic or thermophilic; according to the number of stages,
it is classified as either single-stage or multi-stage; and according to the digester, it is
classified as either a fixed dome, floating dome, balloon, or garage type AD process.
In contrast to a wet AD system, which has a solid content of > 15%, a dry AD process
has a solid composition of between 20 and 40% [136].
While in a continuous system, the organic waste is fed constantly in the digester,
in a batch-type AD process, the industrial organic waste is fed once in the digester,
and inoculum is introduced when it is closed for a certain amount of time. The system
is referred to as mesophilic if the operating temperature is between 20 and 40 °C,
and thermophilic if it is between 50 and 65 °C. AD can be carried out in single-stage,
two-stage, or multi-stage systems. In a single-stage system, digestion takes place in
4 Industrial Organic Waste and Byproducts as Sustainable Feedstock … 101

Fig. 4.4 Schematic representation of organic waste conversion through anaerobic digestion process

a single reactor, whereas a multi-stage system uses two or more reactors [127]. A
wide variety of trash can be treated effectively with two stages of AD.

Factors Affecting the AD Process of Organic Waste

Many chemical and physical parameters, such as seeding, stirring, temperature, pH,
C:N ratio, organic loading rate, hydraulic retention time, and volatile fatty acids, have
an impact on the AD mechanism [18, 123]. Any adjustments to these settings run the
risk of causing a breakdown in the digester process because they alter the movement
and environment of the bacteria inside. Controlling these characteristics is crucial for
maximizing biogas output. The biogas plant should operate properly and efficiently
by varying these parameters within a certain range. The organic feedstock utilized for
the digestion process that produces biogas must provide an environment that allows
the microorganisms engaged in this process to grow and function metabolically at
their best [72].
102 D. G. Gizaw et al.

4.3.2.2 Fermentation

Industrial organic waste rich in carbohydrates can be fermented to produce hydrogen.


Since organic waste is high in carbohydrates, producing hydrogen has two advan-
tages: it recovers energy and reduces waste [118, 128]. The most often employed
microbe for ethanol production is Saccharomyces cerevisiae. The primary process
by which ethanol is fermented, known as glycolysis, converts one glucose molecule
into two pyruvate molecules [60]. Under anaerobic circumstances, pyruvate is further
converted to ethanol and CO2 .
Any substance that includes sugar can undergo the fermentation process to become
ethanol [26, 59]. The three main types of feedstocks utilized to produce ethanol by
fermentation are (a) sugar, (b) starch, and (c) cellulose. While starches and celluloses
must be transformed into fermentable sugars before they can be fermented into
ethanol, sugars can be immediately fermented into ethanol [90]. Fruits, molasses,
sugar beets, and sugar cane are all sources of sugar. Molasses, a thick syrup created
during the sugar refining process, is the most popular sugar for ethanol fermentation
[52]. After being sterilized, molasses are transported to the fermentation facility.
Molasses are diluted with water because a high sugar content causes inefficient
sugar conversion and prolongs fermentation. Therefore, it needs to be converted into
fermentable sugars. Hydrolysis of starch is followed by high-temperature cooking
of the starch while -amylase is added. By dissolving the starch polymer, amylase
prevents gelatinization [68]. Further fermentation of the resultant dextrose yields
ethanol.
Organic waste is a complex mixture of fatty, starchy, and cellulosic components
that is challenging for ethanol-producing bacteria to ferment [92]. This is similar
to industrial waste that has been produced through food processing. As a result, a
pretreatment is frequently required to convert these organic wastes into fermentable
sugars, which can then be fermented to make ethanol. Heat treatment, acid hydrol-
ysis, and enzymatic hydrolysis are all possible pretreatment methods that can be used
for this purpose. An efficient pretreatment can either directly or indirectly hydrolyze
carbohydrates into fermentable sugars, prevent sugar loss, restrict inhibitory activity,
and lower energy requirements. Simultaneous saccharification and fermentation
(SSF) and separate hydrolysis and fermentation (SHF) are the procedures that are
typically used in fermentation. SSF is preferable to the SHF method, which has
historically been used to produce bioethanol, in terms of ethanol yields, cost, and
the need for an additional reactor [17, 114]. In recent years, genetic engineering has
created genetically modified organisms to improve ethanol production by reducing
the intake of microbes’ oxygen consumption. At the same time, the technology is
successful in the microbial utilization of xylose to produce ethanol by altering the
metabolic activity of specific yeast species [120].
4 Industrial Organic Waste and Byproducts as Sustainable Feedstock … 103

4.4 Biofuels from Organic Wastes

4.4.1 Liquid Biofuels

4.4.1.1 Biodiesel

In two processes, transesterification and lipid extraction from industrial organic waste
can be used to synthesize biodiesel. Organic wastes contain carbon, such as phos-
phates, lipids, vitamins, amino acids, and carbohydrates [12, 57]. Lipids can be
converted into biodiesel. For instance, the lipids that are typically included in 5–30%
of organic food processing waste should be extracted first. Therefore, removing lipids
from organic waste is the first step in turning it into biodiesel. Lipids can be extracted
using one of four methods: evaporation, enzymatic hydrolysis, Soxhlet extraction,
or supercritical extraction. Clean lipids recovered from organic waste can be used
immediately to make biodiesel through transesterification [57]. Transesterification is
a well-known process in which the oils/lipids extracted from various sources can be
converted into biodiesel with suitable short-chain alcohol and catalyst, any substance
that fastens the rate of the reaction without itself being consumed [105].
Although extraction is challenging at low values, the lipid portion of organic waste
is well adapted for conversion to fatty acid methyl ester (FAME). Oils from industrial
food wastes have recently been successfully extracted using dimethyl ether (DME)
[101]. Due to DME’s low boiling point (28 °C) and high pressure, e.g., 0.51 MPa
at 20 °C, where extraction occurs, this extraction method is appealing [98]. As a
result, the DME may be efficiently recovered. DME could extract coffee ground oils,
which comprise around 17% of the total weight [101]. Soybean and rapeseed cakes
each contain about 1% oil. Therefore, DME can be used to extract oils further when
pressing has achieved its maximum efficiency. Various industrial food waste streams
could be processed using the DME extraction method. For instance, the lipids are
left more concentrated, at 1.4% of solids post-harvest, after the fermentation of the
carbohydrates in potato peel waste [67]. This procedure might also be employed for
pomegranate seeds containing up to 20% oil by dry weight [7].

4.4.1.2 Bioethanol

A sustainable liquid fuel called bioethanol can be produced from a variety of indus-
trial organic wastes. Additionally, the cellulose component of industrial food waste
exhibits the ability to be converted into ethanol. Citrus peels are among the organic
wastes that have the highest cellulose content (37% for oranges) and lowest lignin
content (7.5% for oranges) [98]. Pectin (23% for oranges) and D-limonene are among
the fermentation inhibitors present, but removing these substances before ethanol
fermentation allows for the synergistic creation of high-value byproducts. Through
acetone, butanol, and ethanol (ABE) fermentation, readily degradable sugars from
industrial organic waste also show potential for conversion to butanol and coproducts.
104 D. G. Gizaw et al.

Butanol has an advantage over ethanol because it can be utilized in standard gasoline
vehicles without requiring engine modifications and has an energy density of 83%
that of gasoline as opposed to 65% for ethanol [2]. Recent research has focused on the
whey from cheese production and waste from baking operations, two organic wastes
that are abundant in quickly degradable sugars. Similar to making butanol from other
products, manufacturing butanol from organic waste presents the same main chal-
lenge: low fermentation concentration (30 g/L) brought on byproduct inhibition [94].
Because of its high boiling point (117 °C) and low vapor pressure (2.3 kPa), butanol
might be difficult to separate [2, 124]. These qualities increase the safety of pure
butanol while reducing its recovery efficiency by distillation. Simultaneous separa-
tion and fermentation have been suggested as a way to circumvent product inhibition,
with pervaporation and adsorption being the least energy-intensive separation tech-
niques. In general, methods for turning industrial organic waste into biofuels show
the ability to link waste sources to a dependable market. Additionally, combining
the extraction of lipids and carbohydrates often has positive synergistic effects and
provides markets for both biodiesel and bio-alcohol.

4.4.1.3 Pyrolysis Oil (Bio-oil)

A sustainable liquid fuel called bio-oil can be synthesized from a variety of indus-
trial organic wastes. Bio-oil is a complicated mixture of water, char particles, and
oxygenated hydrocarbons. Bio-oil has the potential to replace traditional petroleum
fuels as a liquid fuel because of its dark brown color [133]. Bio-oil can be utilized
as a sustainable liquid fuel because it is easily transported and stored. Additionally,
it can be utilized to create compounds. Due to its weak thermal stability, corrosive
behavior, and subpar fuel qualities, bio-oil is not yet commercially feasible. The
bio-oil, which mostly contains oxygen-containing structures, can be supplied to the
petroleum refinery as a feedstock or utilized directly as fuel in boilers for combustion
[42, 44]. Biochar, syngas, and bio-oil are the three main byproducts of pyrolysis. Fast
pyrolysis is used to create bio-oil, while slow pyrolysis is typically utilized to create
charcoal.
Bio-oil can be produced as either a stable emulsion or a separable liquid that
contains both aqueous and oil phases, depending on the feedstock’s content and
the pyrolysis process parameters. However, bio-oil has a low heating value (often
around 50% of the heating value of diesel fuel) because it contains a substantial
amount of dissolved water. Although the use of bio-oil has been tested in diesel
engines with success, there are certain drawbacks to the bio-oil, including poor cold
flowability, corrosion, and the development of gummy deposits [48, 75]. Because bio-
oil and diesel are not miscible, it is necessary to create emulsions that require pricey
surfactants to stabilize them. Bio-oils cannot be used in contemporary diesel engines
due to the corrosion and deposits they cause. As a result, bio-oil needs to be improved
before it can be utilized in transportation. Catalytic cracking or hydro-treating can
be used to upgrade bio-oil [10, 13].
4 Industrial Organic Waste and Byproducts as Sustainable Feedstock … 105

4.4.2 Gaseous Fuels

4.4.2.1 Biogas

Anaerobic digestion is used to produce one of the most significant sustainable biofuels
from industrial organic waste: biogas. It mostly consists of 65% methane and 35%
carbon dioxide, with water vapor and hydrogen sulfide traces. According to [53], it is
lighter than air by roughly 20%. Unlike liquefied petroleum gas (LPG), it cannot be
transformed into a liquid at room temperature. Bio-methane (enriched biogas) can be
compressed into cylinders for simple transportation after the CO2 has been removed.
Additionally, enriched biogas, which has the potential to be used in all applications
where CNG is now used, can be produced using CNG technology [61]. To boost the
energy content of biogas and enable high-pressure storage, CO2 is removed when
biogas is utilized as fuel [85]. A fascinating fact is that methane predominates in
biogas and natural gas production under anaerobic conditions. However, turning
dead biomass into natural gas takes a million years, whereas turning organic waste
into biogas takes just a few weeks. The components of the acidogenesis and acetoge-
nesis stages of anaerobic digestion are transformed into methane and carbon dioxide
by methanogens (methane-forming bacteria). In other words, acetogenesis further
converts alcohols and short-chain volatile acids to acetic acid and hydrogen from
VFAs, alcohol, ketones, CO2 , H2 , NH3 , and other compounds.
A few problems exist with biogas. Typically, it has a poor heating value. It gener-
ates more pollutants than natural gas when it is burnt in engines. It includes siloxanes
and, when burned, forms silica deposits. These deposits of silica harm the compo-
nents and engines. In addition, methane emissions are produced, which are thought
to be stronger greenhouse gases than carbon dioxide [34]. When in touch with human
skin, methanol is less safe than gasoline. Limit the exposure hazards, which in turn
leads to somewhat higher investment costs, this necessitates safety systems such as
closed filling systems.

4.4.2.2 Syngas

Syngas synthesis from industrial organic waste has recently attracted a lot of attention.
Synthetic gas, often known as syngas, is produced by extracting hydrogen, carbon
monoxide, and carbon dioxide from natural gas or coal. It can be used to make
compounds like methanol and ammonia. In addition to creating various byproducts
such as methanol, ethanol, hydrogen, and Fischer–Tropsch products, syngas can be
utilized directly for power generation in boilers, IC engines, gas turbines, and fuel
cells [51, 55, 63]. Syngas can be transformed into synthetic petroleum using Fischer–
Tropsch synthesis, which has uses as a fuel or lubricant [30, 39]. Syngas production
from waste and biomass has recently attracted a lot of attention. Additionally, syngas
can be created directly through gasification or in two stages by pyrolyzing organic
waste to create bio-oil, which can subsequently be gasified to create syngas.
106 D. G. Gizaw et al.

The reaction temperature significantly influenced the composition of the produced


gas, and the syngas production rate increased as the reaction temperature rose [132].
Water can increase reactivity through the steam reforming and decrease carbon
monoxide generation through the water–gas shift reaction. The amount of oxygen
enhanced methane production while simultaneously increasing hydrogen and carbon
monoxide production. In the study by [81], different agro-food wastes and fruit peels
were tested to produce hydrogen-rich syngas using supercritical water gasification at
high temperatures (400–600 °C), high pressures (23–25 MPa), and biomass-to-water
ratios (1:5 and 1:10) for a reaction time (15–45 min). With a total syngas yield of
15 mmol/g and a higher hydrogen yield of 4.8 mmol/g from coconut shells, a lower
heating value of 1595 kJ/Nm3 could be attained. Under conditions of high pressure
and temperature and the presence of a catalyst (hydrogen), syngas can be transformed
into methanol [119]. Hydrogen can be produced from syngas via steam reforming.
Organic waste with less than 35% moisture level can be gasified using subcritical
water. Supercritical water gasification can be employed since organic waste contains
more than 60% moisture. After steam reforming, the water–gas shift reaction can
improve hydrogen production even more [16, 22]. All carbon content is utilized in
the synthesis of methanol when sufficiently enough hydrogen is present under high
pressures, doubling the yield of methanol production from the same biomass.

4.4.2.3 Biohydrogen

To decrease our dependence on fossil fuels, researchers are paying more atten-
tion to biohydrogen-generating systems that can produce hydrogen from renewable
resources such as industrial organic wastes. Due to its numerous benefits, biohy-
drogen has recently emerged as one of the most potential fuels to replace fossil
fuels [50, 93]. Biohydrogen has 2.75 times more energy per unit of weight than
hydrocarbon fuels, it has a high-energy content of roughly 122 kJ/g [49]. It creates
water during combustion, making it an eco-friendly and clean fuel. For imme-
diate power generation, it can be utilized in fuel cells. Conventional methods for
producing hydrogen include partial oxidation of fossil fuels, steam reforming of
methane and other hydrocarbons, autothermal reforming, desulfurization, pyrolysis,
plasma reforming, aqueous phase reforming, and ammonia reforming. There are
other non-reforming techniques for hydrogen production, including gasification and
biological processes. Four different methods can be used to create biohydrogen: direct
photolysis of water using algae or cyanobacteria, photofermentation using photosyn-
thetic bacteria, dark fermentation using anaerobic bacteria, and hybrid systems [3,
80]. Dark fermentation is regarded as the most practical procedure because it does
not require any outside energy or light.
4 Industrial Organic Waste and Byproducts as Sustainable Feedstock … 107

4.4.3 Biochar

A carbon-rich byproduct of the pyrolysis process is biochar. Biochar, synthesis gas,


and pyrolysis oil are all products of pyrolysis [58, 89]. Due to its high acidity and
complicated composition, bio-oil cannot be utilized directly and must be upgraded.
Syngas’s use is constrained by its low yield and difficult separation and purification
process. Contrarily, biochar offers many benefits over bio-oil and syngas. Both fertil-
izer and soil conditioner can be used with it. By boosting stable soil carbon stocks
and soil carbon sequestration while lowering atmospheric CO2 concentrations, it has
the potential to be a technique for mitigating climate change. It raises agricultural
productivity, particularly on poor and damaged soils. Additionally, it increases the
soil’s ability to hold water. In terms of locking up carbon, the generation of biochar is
more effective than composting [84]. Additionally, following 10–20 years, microbial
activity in compost will liberate the carbon in it. However, the biochar-sequestered
carbon would stay stable in the soil. Additionally, it reduces emissions of green-
house gases such as nitrous oxide and methane [11, 140]. High-production bio-oil
is produced when the operating temperature is low, the heating rate is high, and
the gas residence time is short [19]. High-production syngas are produced when the
operating temperature is high, the heating rate is low, and the gas residence time
is long. On the other hand, a low heating rate and low operating temperature result
in a high output of biochar [91, 129]. As a result, the high cellulose and hemicel-
lulose feedstocks are quickly pyrolyzed to produce bio-oil and gas as the primary
products with a low biochar yield [130, 142]. High-lignin feedstocks can yield the
most biochar when pyrolyzed slowly. Therefore, whether or not slow pyrolysis is
employed depends on the choice of feedstock and the necessary balance of products,
i.e., bio-oil, syngas, and biochar [43].

4.5 Sustainability of Industrial Organic Wastes

The issues posed by industrial organic wastes and recent urbanization in industrial-
ized countries have increased the demand for natural resources and created oppor-
tunities for the use of organic waste in a variety of industries [78]. There are now
several industrial organic wastes being exploited as renewable bioenergy sources.
Various industries produce a lot of organic waste each year, and improper disposal
of this waste may pollute the environment and be harmful to human health [12].
Organic waste is typically not processed adequately, resulting in incorrect disposal
methods such as open-air burning, landfilling, and dumping in prohibited locations.
Untreated waste produces a variety of greenhouse gases, which change the climate.
Using non-renewable energy sources also contributes to the ozone layer’s depletion.
Consequently, advancing other cleaner and renewable energy is now a universal
concern.
108 D. G. Gizaw et al.

4.6 Challenges and Future Perspectives

Global population growth, agricultural operations, and food processing have led to
increased industrial organic waste generation, hygienic problems, and dangerous gas
emissions from decomposing materials. This chapter has provided sustainable and
environmentally friendly strategies for effective industrial organic waste conversion
and utilization as sustainable bioenergy. It has also been demonstrated that indus-
trial organic wastes may be transformed into bioethanol, biomethane, biohydrogen,
biobutanol, bio-oil, biochar, and biodiesel for use in fueling vehicles and generating
electricity. The creation of biofuels from organic wastes has advantages in terms
of the environment, public health, economy, industry, and technology, and when
used in transportation engines, it lowers greenhouse gas emissions, improves engine
performance, encourages smoother engine operation, and extends engine life. The
difficulties posed by the enormous volume of organic waste generated by indus-
trial activity can be avoided by transforming such organic waste into useful forms.
Creating biofuels from this waste is still an economically and environmentally sound
solution. To improve conversion rates, use less energy, and produce more goods of
higher quality, these organic wastes must first undergo pretreatment. To improve the
digestibility and biodegradability of the biomass, the pretreatment procedure breaks
down the cellulose, hemicellulose, and lignin components. It is reasonable to predict
the direction of future research given the significance of employing industrial organic
waste as a feedstock for the production of biofuels.

4.7 Conclusions

Industrial organic waste is one of the potential sources to produce various bioen-
ergy products. The waste conversion methods determine the quality and quantity
of the biofuel. However, novel and improved pretreatment procedures are required
to ensure better degradation and conversion of organic wastes to valuable products.
To better understand the intricate mechanisms and molecular kinetics of organic
waste degradation and conversion, interdisciplinary research on cost analysis, sensi-
tivity analysis, energy and exergy analysis, techno-economic assessment, and life
cycle assessments is essential. The reactor design and optimization, design param-
eters, cost, time, and manpower requirements to improve conversion efficiency and
product quality, the deployment of appropriate innovative technologies is crucial.
These include computational fluid dynamics, machine learning, artificial intelligence,
statistical modeling, and optimization tools. The successful conversion of industrial
organic waste into bioenergy could be a potential option for a sustainable circular
bioeconomy.
4 Industrial Organic Waste and Byproducts as Sustainable Feedstock … 109

References

1. Abbas-Abadi, M. S., Kusenberg, M., Zayoud, A., Roosen, M., Vermeire, F., Madanikashani,
S., Kuzmanović, M., Parvizi, B., Kresovic, U., De Meester, S., & Van Geem, K. M.
(2023). Thermal pyrolysis of waste versus virgin polyolefin feedstocks: The role of pressure,
temperature and waste composition. Waste Management, 165, 108–118.
2. Abdehagh, N., Tezel, F. H., & Thibault, J. (2014). Separation techniques in butanol production:
Challenges and developments. Biomass and Bioenergy, 60, 222–246.
3. Abdelqader, A. A., Abdelsalam, E. M., Attia, Y. A., Moselhy, M., Ali, A. S., Arisha, A.
H., & Samer, M. (2022). Application of helium-neon red laser for increasing biohydrogen
production from anaerobic digestion of biowastes. Egyptian Journal of Chemistry, 65(1),
11–17.
4. Aierzhati, A., Watson, J., Si, B., Stablein, M., Wang, T., & Zhang, Y. (2021). Development of a
mobile, pilot scale hydrothermal liquefaction reactor: Food waste conversion product analysis
and techno-economic assessment. Energy Conversion and Management: X, 10, 100076.
5. Al, A. R., Bikić, S., & Radojčin, M. (2023). Bioenergy conversion technologies: A review
and case study. Journal on Processing and Energy in Agriculture, 27(1), 30–38.
6. Alaedini, A. H., Tourani, H. K., & Saidi, M. (2023). A review of waste-to-hydrogen conversion
technologies for solid oxide fuel cell (SOFC) applications: Aspect of gasification process and
catalyst development. Journal of Environmental Management, 329, 117077.
7. Al-Maiman, S. A., & Ahmad, D. (2002). Changes in physical and chemical properties during
pomegranate (Punica granatum L.) fruit maturation. Food Chemistry, 76(4), 437–441.
8. Anand, S., Kaur, J., Sarao, L. K., & Singh, A. (2023). Agricultural residues and manures into
bioenergy. In Agroindustrial waste for green fuel application (pp. 67–87). Springer Nature
Singapore.
9. Anca-Couce, A., Hochenauer, C., & Scharler, R. (2021). Bioenergy technologies, uses, market
and future trends with Austria as a case study. Renewable and Sustainable Energy Reviews,
135, 110237.
10. Arefin, M. A., Rashid, F., & Islam, A. (2021). A review of biofuel production from floating
aquatic plants: An emerging source of bio-renewable energy. Biofuels, Bioproducts and
Biorefining, 15(2), 574–591.
11. Arif, M., Jan, T., Riaz, M., Fahad, S., Adnan, M., Amanullah, Ali, K., Mian, I.A., Khan, B., &
Rasul, F. (2020). Biochar; a remedy for climate change. In Environment, climate, plant and
vegetation growth (pp. 151–171).
12. Ashokkumar, V., Flora, G., Venkatkarthick, R., SenthilKannan, K., Kuppam, C., Stephy, G. M.,
Kamyab, H., Chen, W. H., Thomas, J., & Ngamcharussrivichai, C. (2022). Advanced technolo-
gies on the sustainable approaches for conversion of organic waste to valuable bioproducts:
Emerging circular bioeconomy perspective. Fuel, 324, 124313.
13. Atanda, L., Fraga, G. L. L., Ahmed, M. H., Alothman, Z. A., Na, J., Batalha, N., Aslam, W., &
Konarova, M. (2021). Conversion of agricultural waste into stable biocrude using spinel oxide
catalysts. Journal of Hazardous Materials, 402, 123539.
14. Atelge, M. R., Krisa, D., Kumar, G., Eskicioglu, C., Nguyen, D. D., Chang, S. W., Atabani, A.
E., Al-Muhtaseb, A. H., & Unalan, S. (2020). Biogas production from organic waste: Recent
progress and perspectives. Waste and Biomass Valorization, 11, 1019–1040.
15. Awasthi, M. K., Sindhu, R., Sirohi, R., Kumar, V., Ahluwalia, V., Binod, P., Juneja, A., Kumar,
D., Yan, B., Sarsaiya, S., & Zhang, Z. (2022). Agricultural waste biorefinery development
towards circular bioeconomy. Renewable and Sustainable Energy Reviews, 158, 112122.
16. Babatabar, M. A., & Saidi, M. (2021). Hydrogen production via integrated configuration of
steam gasification process of biomass and water-gas shift reaction: Process simulation and
optimization. International Journal of Energy Research, 45(13), 19378–19394.
17. Bakari, H., Djomdi, Falama Ruben, Z., Roger, D. D., Cedric, D., Guillaume, P., Pascal,
D., Philippe, M., & Gwendoline, C. (2023). Optimization of bioethanol production after
enzymatic treatment of sweet sorghum stalks. Waste and Biomass Valorization, 1–15.
110 D. G. Gizaw et al.

18. Bella, K., & Rao, P. V. (2021). Anaerobic digestion of dairy wastewater: Effect of different
parameters and co-digestion options—A review. Biomass Conversion and Biorefinery, 1–26.
19. Bhoi, P. R., Ouedraogo, A. S., Soloiu, V., & Quirino, R. (2020). Recent advances on catalysts
for improving hydrocarbon compounds in bio-oil of biomass catalytic pyrolysis. Renewable
and Sustainable Energy Reviews, 121, 109676.
20. Bilal, M., Mehmood, T., Nadeem, F., Barbosa, A. M., de Souza, R. L., Pompeu, G. B., Meer,
B., Ferreira, L. F. R., & Iqbal, H. M. (2022). Enzyme-assisted transformation of lignin-based
food bio-residues into high-value products with a zero-waste theme: A review. Waste and
Biomass Valorization, 1–18.
21. Chen, L., Yang, K., Huang, J., Liu, P., Yang, J., Pan, Y., Qi, F., & Jia, L. (2022). Experimental
and kinetic study on flash pyrolysis of biomass via on-line photoionization mass spectrometry.
Applications in Energy and Combustion Science, 9, 100057.
22. Chen, W. H., Lu, C. Y., Chou, W. S., Sharma, A. K., Saravanakumar, A., & Tran, K. Q.
(2023). Design and optimization of a crossflow tube reactor system for hydrogen production
by combining ethanol steam reforming and water gas shift reaction. Fuel, 334, 126628.
23. Coma, M., Martinez-Hernandez, E., Abeln, F., Raikova, S., Donnelly, J., Arnot, T. C., Allen,
M. J., Hong, D. D., & Chuck, C. J. (2017). Organic waste as a sustainable feedstock for
platform chemicals. Faraday discussions, 202, 175–195.
24. Dahl, R. (2015). A second life for scraps: Making biogas from food waste.
25. Dhandayuthapani, K., Kumar, P. S., Chia, W. Y., Chew, K. W., Karthik, V., Selvarangaraj, H.,
Selvakumar, P., Sivashanmugam, P., & Show, P. L. (2022). Bioethanol from hydrolysate of
ultrasonic processed robust microalgal biomass cultivated in dairy wastewater under optimal
strategy. Energy, 244, 122604. https://doi.org/10.1016/j.energy.2021.122604
26. Dhandayuthapani, K., Sarumathi, V., Selvakumar, P., Temesgen, T., Asaithambi, P., &
Sivashanmugam, P. (2021). Study on the ethanol production from hydrolysate derived by
ultrasonic pretreated defatted biomass of chlorella sorokiniana NITTS3. Chemical Data
Collections, 31, 100641. https://doi.org/10.1016/j.cdc.2020.100641
27. Di Giacomo, G., Gallifuoco, A., & Taglieri, L. (2016). Hydrothermal carbonization of
mixed biomass: Experimental investigation for an optimal valorisation of agrofood wastes. In
Proceedings of the 24th European biomass conference and exibition (24th EUBCE) (pp. 6–9).
28. Dickerson, T., & Soria, J. (2013). Catalytic fast pyrolysis: A review. Energies, 6(1), 514–538.
29. Dimitriadis, A., & Bezergianni, S. (2017). Hydrothermal liquefaction of various biomass and
waste feedstocks for biocrude production: A state of the art review. Renewable and Sustainable
Energy Reviews, 68, 113–125.
30. dos Santos, R. G., & Alencar, A. C. (2020). Biomass-derived syngas production via gasifi-
cation process and its catalytic conversion into fuels by Fischer Tropsch synthesis: A review.
International Journal of Hydrogen Energy, 45(36), 18114–18132.
31. Elkhalifa, S., Al-Ansari, T., Mackey, H. R., & McKay, G. (2019). Food waste to biochars
through pyrolysis: A review. Resources, Conservation and Recycling, 144, 310–320.
32. ElMekawy, A., Diels, L., De Wever, H., & Pant, D. (2013). Valorization of cereal based biore-
finery byproducts: Reality and expectations. Environmental Science & Technology, 47(16),
9014–9027.
33. Fabry, F., Rehmet, C., Rohani, V., & Fulcheri, L. (2013). Waste gasification by thermal plasma:
A review. Waste and Biomass Valorization, 4, 421–439.
34. Farghali, M., Osman, A. I., Umetsu, K., & Rooney, D. W. (2022). Integration of biogas systems
into a carbon zero and hydrogen economy: A review. Environmental Chemistry Letters, 20(5),
2853–2927.
35. Frolov, S. M., Smetanyuk, V. A., Shamshin, I. O., Sadykov, I. A., & Frolov, F. S. (2021).
Production of highly superheated steam by cyclic detonations of propane-and methane-steam
mixtures with oxygen for waste gasification. Applied Thermal Engineering, 183, 116195.
36. Gaur, V. K., Sharma, P., Sirohi, R., Awasthi, M. K., Dussap, C. G., & Pandey, A. (2020).
Assessing the impact of industrial waste on environment and mitigation strategies: A
comprehensive review. Journal of Hazardous Materials, 398, 123019.
4 Industrial Organic Waste and Byproducts as Sustainable Feedstock … 111

37. Gizaw, D. G., Periyasamy, P., Redda, Z. T., John, B. I., Mengstie, H. B., & Asaithambi, P.
(2023). A comprehensive review on sewage sludge as sustainable feedstock for bioenergy
production. Environmental Quality Management, 1–16. https://doi.org/10.1002/tqem.22116
38. Guldhe, A., Singh, B., Renuka, N., Singh, P., Misra, R., & Bux, F. (2017). Bioenergy: A
sustainable approach for cleaner environment. In Phytoremediation potential of bioenergy
plants (pp. 47–62).
39. Gupta, P. K., Kumar, V., & Maity, S. (2021). Renewable fuels from different carbonaceous
feedstocks: A sustainable route through Fischer-Tropsch synthesis. Journal of Chemical
Technology & Biotechnology, 96(4), 853–868.
40. Haghighat, M., Majidian, N., & Hallajisani, A. (2020). Production of bio-oil from sewage
sludge: A review on the thermal and catalytic conversion by pyrolysis. Sustainable Energy
Technologies and Assessments, 42, 100870.
41. Hasan, M. M., Rasul, M. G., Khan, M. M. K., Ashwath, N., & Jahirul, M. I. (2021). Energy
recovery from municipal solid waste using pyrolysis technology: A review on current status
and developments. Renewable and Sustainable Energy Reviews, 145, 111073.
42. Hu, X., & Gholizadeh, M. (2020). Progress of the applications of bio-oil. Renewable and
Sustainable Energy Reviews, 134, 110124.
43. Huang, C., Mohamed, B. A., & Li, L. Y. (2022). Comparative life-cycle assessment of pyrol-
ysis processes for producing bio-oil, biochar, and activated carbon from sewage sludge.
Resources, Conservation and Recycling, 181, 106273.
44. Ilyin, S. O., & Makarova, V. V. (2022). Bio-oil: Production, modification, and application.
Chemistry and Technology of Fuels and Oils, 58(1), 29–44.
45. Jalali, M., & Saremi, P. (2020). A study on the impact of energy conversion technologies on
food waste from macromolecules on the environment. Indian Journal of Science Research,
11(1), 45–56.
46. Jamro, I. A., Khoso, S., Shah, S. A. R., Ali, S., Ali, F., & Hassane, H. U. (2022). Investigation
of thermal decomposition and gases release from pre-drying municipal solid waste (PMSW)
via pyrolysis technology. Architecture, Civil Engineering, Environment, 15(4), 119–131.
47. Japhet, J. A., Luka, B. S., Maren, I. B., & Datau, S. G. (2020). The potential of wood and
agricultural waste for pellet fuel development in nigeria—A technical review. International
Journal of Engineering Application Science Technology, 4, 598–607.
48. Jeevahan, J., Anderson, A., Sriram, V., Durairaj, R. B., Britto Joseph, G., & Mageshwaran,
G. (2021). Waste into energy conversion technologies and conversion of food wastes into the
potential products: A review. International Journal of Ambient Energy, 42(9), 1083–1101.
49. Jeong, Y. S., Park, K. B., & Kim, J. S. (2020). Hydrogen production from steam gasification
of polyethylene using a two-stage gasifier and active carbon. Applied Energy, 262, 114495.
50. Johari, A., Singh, S., & Vidya, S. (2022). Engine performance analysis for diesel engine using
hydrogen as an alternative fuel. Materials Today: Proceedings, 56, 342–346.
51. Jones, M. P., Krexner, T., & Bismarck, A. (2022). Repurposing Fischer-Tropsch and natural
gas as bridging technologies for the energy revolution. Energy Conversion and Management,
267, 115882.
52. Kabeyi, M. J. B., & Olanrewaju, O. A. (2022). Sugarcane molasses to energy conversion for
sustainable production and energy transition. In 12th annual Istanbul international conference
on industrial engineering and operations management.
53. Kalsum, L., & Hasan, A. (2022). The effect of the packing flow area and biogas flow rate on
biogas purification in packed bed scrubber. Journal of Ecological Engineering, 23(11).
54. Kambo, H. S., Minaret, J., & Dutta, A. (2018). Process water from the hydrothermal
carbonization of biomass: A waste or a valuable product? Waste and Biomass Valorization,
9, 1181–1189.
55. Kapoor, R., Ghosh, P., Tyagi, B., Vijay, V. K., Vijay, V., Thakur, I. S., Kamyab, H., Nguyen,
D. D., & Kumar, A. (2020). Advances in biogas valorization and utilization systems: A
comprehensive review. Journal of Cleaner Production, 273, 123052.
56. Karimi, S., Mahboobi Soofiani, N., Mahboubi, A., & Taherzadeh, M. J. (2018). Use of organic
wastes and industrial by-products to produce filamentous fungi with potential as aqua-feed
ingredients. Sustainability, 10(9), 3296.
112 D. G. Gizaw et al.

57. Karmee, S. K. (2016). Liquid biofuels from food waste: Current trends, prospect and
limitation. Renewable and Sustainable Energy Reviews, 53, 945–953.
58. Karthik, V., Karuna, B., Jeyanthi, J., & Periyasamy, S. (2023). Biochar production from
Manilkara zapota seeds, activation and characterization for effective removal of Cu2+ ions
in polluted drinking water. Biomass Conversion and Biorefinery, 13(11), 9381–9395. https://
doi.org/10.1007/s13399-022-03627-2
59. Kavitha, S., Gajendran, T., Saranya, K., Selvakumar, P., & Manivasagan, V. (2021). Study
on consolidated bioprocessing of pre-treated nannochloropsis gaditana biomass into ethanol
under optimal strategy. Renewable Energy, 172, 440–452. https://doi.org/10.1016/j.renene.
2021.03.015
60. Kavitha, S., Gajendran, T., Saranya, K., Selvakumar, P., Manivasagan, V., & Jeevitha,
S. (2022). An insight-A statistical investigation of consolidated bioprocessing of Allium
ascalonicum leaves to ethanol using Hangateiclostridium thermocellum KSMK1203 and
synthetic consortium. Renewable Energy, 187, 403–416. https://doi.org/10.1016/j.renene.
2022.01.047
61. Khan, I. U., Othman, M. H. D., Hashim, H., Matsuura, T., Ismail, A. F., Rezaei-DashtArzhandi,
M., & Azelee, I. W. (2017). Biogas as a renewable energy fuel—A review of biogas upgrading,
utilisation and storage. Energy Conversion and Management, 150, 277–294.
62. Khan, T. A., Saud, A. S., Jamari, S. S., Ab Rahim, M. H., Park, J. W., & Kim, H. J. (2019).
Hydrothermal carbonization of lignocellulosic biomass for carbon rich material preparation:
A review. Biomass and bioenergy, 130, 105384.
63. Khosravani, H., Meshksar, M., Rahimpour, H. R., & Rahimpour, M. R. (2023). Introduction
to syngas products and applications. In Advances in synthesis gas: Methods, technologies and
applications (pp. 3–25). Elsevier.
64. Kim, S., Lee, Y., Lin, K. Y. A., Hong, E., Kwon, E. E., & Lee, J. (2020). The valorization of
food waste via pyrolysis. Journal of Cleaner Production, 259, 120816.
65. Koul, B., Yakoob, M., & Shah, M. P. (2022). Agricultural waste management strategies for
environmental sustainability. Environmental Research, 206, 112285.
66. Lee, U., Benavides, P. T., & Wang, M. (2020). Life cycle analysis of waste-to-energy pathways.
In Waste-to-energy (pp. 213–233). Academic Press.
67. Liang, S., & McDonald, A. G. (2014). Chemical and thermal characterization of potato
peel waste and its fermentation residue as potential resources for biofuel and bioproducts
production. Journal of Agricultural and Food Chemistry, 62(33), 8421–8429.
68. Liu, L., An, X., Zhang, H., Lu, Z., Nie, S., Cao, H., Xu, Q., & Liu, H. (2020). Ball milling
pretreatment facilitating α-amylase hydrolysis for production of starch-based bio-latex with
high performance. Carbohydrate Polymers, 242, 116384.
69. Liu, M., Ogunmoroti, A., Liu, W., Li, M., Bi, M., Liu, W., & Cui, Z. (2022). Assessment
and projection of environmental impacts of food waste treatment in China from life cycle
perspectives. Science of The Total Environment, 807, 150751.
70. Lu, X., Jordan, B., & Berge, N. D. (2012). Thermal conversion of municipal solid waste via
hydrothermal carbonization: Comparison of carbonization products to products from current
waste management techniques. Waste Management, 32(7), 1353–1365.
71. Mahmood, R., Parshetti, G. K., & Balasubramanian, R. (2016). Energy, exergy and techno-
economic analyses of hydrothermal oxidation of food waste to produce hydro-char and bio-oil.
Energy, 102, 187–198.
72. Mahmudul, H. M., Rasul, M. G., Akbar, D., Narayanan, R., & Mofijur, M. (2021). A compre-
hensive review of the recent development and challenges of a solar-assisted biodigester system.
Science of The Total Environment, 753, 141920.
73. Manikandan, S., Vickram, S., Sirohi, R., Subbaiya, R., Krishnan, R. Y., Karmegam, N.,
Sumathijones, C., Rajagopal, R., Chang, S. W., Ravindran, B., & Awasthi, M. K. (2023).
Critical review of biochemical pathways to transformation of waste and biomass into
bioenergy. Bioresource Technology, 128679.
74. Maraveas, C. (2020). Production of sustainable and biodegradable polymers from agricultural
waste. Polymers, 12(5), 1127.
4 Industrial Organic Waste and Byproducts as Sustainable Feedstock … 113

75. Masudi, A., Muraza, O., Jusoh, N. W. C., & Ubaidillah, U. (2023). Improvements in the
stability of biodiesel fuels: Recent progress and challenges. Environmental Science and
Pollution Research, 30(6), 14104–14125.
76. Mofijur, M., Siddiki, S. Y. A., Shuvho, M. B. A., Djavanroodi, F., Fattah, I. R., Ong, H. C.,
Chowdhury, M. A., & Mahlia, T. M. I. (2021). Effect of nanocatalysts on the transesterification
reaction of first, second and third generation biodiesel sources—A mini-review. Chemosphere,
270, 128642.
77. Mofijur, M. G. R. M., Rasul, A. M., Hyde, J., Azad, A. K., Mamat, R., & Bhuiya, M. M. K.
(2016). Role of biofuel and their binary (diesel–biodiesel) and ternary (ethanol–biodiesel–
diesel) blends on internal combustion engines emission reduction. Renewable and Sustainable
Energy Reviews, 53, 265–278.
78. Mohamed, B. A., Bilal, M., Salama, E. S., Periyasamy, S., Fattah, I. R., Ruan, R., Awasthi,
M. K., & Leng, L. (2022). Phenolic-rich bio-oil production by microwave catalytic pyrolysis
of switchgrass: Experimental study, life cycle assessment, and economic analysis. Journal of
Cleaner Production, 366, 132668. https://doi.org/10.1016/j.jclepro.2022.132668
79. Muhammad, G., Alam, M. A., Mofijur, M., Jahirul, M. I., Lv, Y., Xiong, W., Ong, H. C., &
Xu, J. (2021). Modern developmental aspects in the field of economical harvesting and
biodiesel production from microalgae biomass. Renewable and Sustainable Energy Reviews,
135, 110209.
80. Mukherjee, T., & Mohan, S. V. (2022). Bio-waste to hydrogen production technologies.
In Advanced biofuel technologies (pp. 389–407). Elsevier.
81. Nanda, S., Isen, J., Dalai, A. K., & Kozinski, J. A. (2016). Gasification of fruit wastes and
agro-food residues in supercritical water. Energy Conversion and Management, 110, 296–306.
82. Nkosi, N., Muzenda, E., Gorimbo, J., & Belaid, M. (2021). Developments in waste tyre ther-
mochemical conversion processes: Gasification, pyrolysis and liquefaction. RSC Advances,
11(20), 11844–11871.
83. Okolie, J. A., Epelle, E. I., Tabat, M. E., Orivri, U., Amenaghawon, A. N., Okoye, P. U., &
Gunes, B. (2022). Waste biomass valorization for the production of biofuels and value-added
products: A comprehensive review of thermochemical, biological and integrated processes.
Process Safety and Environmental Protection, 159, 323–344.
84. Oldfield, T. L., Sikirica, N., Mondini, C., López, G., Kuikman, P. J., & Holden, N. M. (2018).
Biochar, compost and biochar-compost blend as options to recover nutrients and sequester
carbon. Journal of Environmental Management, 218, 465–476.
85. Olugasa, T. T., Odesola, I. F., & Oyewola, M. O. (2014). Energy production from biogas:
A conceptual review for use in Nigeria. Renewable and Sustainable Energy Reviews, 32,
770–776.
86. Ong, H. C., Tiong, Y. W., Goh, B. H. H., Gan, Y. Y., Mofijur, M., Fattah, I. R., Chong,
C. T., Alam, M. A., Lee, H. V., Silitonga, A. S., & Mahlia, T. M. I. (2021). Recent
advances in biodiesel production from agricultural products and microalgae using ionic
liquids: Opportunities and challenges. Energy Conversion and Management, 228, 113647.
87. Opatokun, S. A., Lopez-Sabiron, A. M., Ferreira, G., & Strezov, V. (2017). Life cycle analysis
of energy production from food waste through anaerobic digestion, pyrolysis and integrated
energy system. Sustainability, 9(10), 1804.
88. Patel, A., Temgire, S., & Borah, A. (2021). Agro-industrial waste as source of bioactive
compounds and their utilization: A review. Pharma Innov, 10(5), 192–196.
89. Pawar, A., Panwar, N. L., & Salvi, B. L. (2020). Comprehensive review on pyrolytic oil
production, upgrading and its utilization. Journal of Material Cycles and Waste Management,
22, 1712–1722.
90. Periyasamy, S., Isabel, J. B., Kavitha, S., Karthik, V., Mohamed, B. A., Gizaw, D. G., Sivashan-
mugam, P., & Aminabhavi, T. M. (2023). Recent advances in consolidated bioprocessing
for conversion of lignocellulosic biomass into bioethanol–A review. Chemical Engineering
Journal, 453, 139783. https://doi.org/10.1016/j.cej.2022.139783
91. Periyasamy, S., Karthik, V., Senthil Kumar, P., Isabel, J. B., Temesgen, T., Hunegnaw, B. M.,
Melese, B. B., Mohamed, B. A., & Vo, D. V. N. (2022). Chemical, physical and biological
114 D. G. Gizaw et al.

methods to convert lignocellulosic waste into value-added products. A review. Environmental


Chemistry Letters, 20(2), 1129–1152. https://doi.org/10.1007/s10311-021-01374-w
92. Prasoulas, G., Gentikis, A., Konti, A., Kalantzi, S., Kekos, D., & Mamma, D. (2020).
Bioethanol production from food waste applying the multienzyme system produced on-site
by Fusarium oxysporum F3 and mixed microbial cultures. Fermentation, 6(2), 39.
93. Qazi, U. Y. (2022). Future of hydrogen as an alternative fuel for next-generation industrial
applications; challenges and expected opportunities. Energies, 15(13), 4741.
94. Quested, T., Ingle, R., & Parry, A. (2013). Household food and drink waste in the United
Kingdom 2012.
95. Ramesh, M., Adithya, K., Kumar, C. J., Mohan, C. G., Nalawade, J., & Prakash, R.
(2021). Thermochemical conversion methods of bio-derived lignocellulosic waste molecules
into renewable fuels. In Advanced technology for the conversion of waste into fuels and
chemicals (pp. 197–215). Woodhead Publishing.
96. Rasheed, T., Anwar, M. T., Ahmad, N., Sher, F., Khan, S. U. D., Ahmad, A., Khan, R., &
Wazeer, I. (2021). Valorisation and emerging perspective of biomass based waste-to-energy
technologies and their socio-environmental impact: A review. Journal of Environmental
Management, 287, 112257.
97. Rauch, R., Hrbek, J., & Hofbauer, H. (2016). Biomass gasification for synthesis gas production
and applications of the syngas. Advances in Bioenergy: The Sustainability Challenge, 73–91.
98. RedCorn, R., Fatemi, S., & Engelberth, A. S. (2018). Comparing end-use potential for
industrial food-waste sources. Engineering, 4(3), 371–380.
99. Rosero-Delgado, E. A., Zambrano-Arcentales, M. A., Gómez-Salcedo, Y., Baquerizo-Crespo,
R. J., & Dustet-Mendoza, J. C. (2021). Biotechnology applied to treatments of agro-
industrial wastes. In Advances in the domain of environmental biotechnology: microbiological
developments in industries, wastewater treatment and agriculture (pp. 277–311).
100. Sadh, P. K., Duhan, S., & Duhan, J. S. (2018). Agro-industrial wastes and their utilization
using solid state fermentation: A review. Bioresources and Bioprocessing, 5(1), 1–15.
101. Sakuragi, K., Li, P., Otaka, M., & Makino, H. (2016). Recovery of bio-oil from industrial
food waste by liquefied dimethyl ether for biodiesel production. Energies, 9(2), 106.
102. Sarkar, N., Chakraborty, D., Dutta, R., Agrahari, P., Sundaram, D., Singh, A., & Jacob, S.
(2021). A comprehensive review on oilseed cakes and their potential as a feedstock for inte-
grated biorefinery. Journal of Advanced Biotechnology and Experimental Therapeutics, 4(3),
376.
103. Sawyerr, N., Trois, C., Workneh, T., & Okudoh, V. (2019). An overview of biogas production:
Fundamentals, applications and future research. International Journal of Energy Economics
and Policy, 9(2), 105.
104. Selvakumar, P., & Sivashanmugam, P. (2018). Multi-hydrolytic biocatalyst from organic solid
waste and its application in municipal waste activated sludge pre-treatment towards energy
recovery. Process Safety and Environmental Protection, 117, 1–10. https://doi.org/10.1016/j.
psep.2018.03.036
105. Selvakumar, P., & Sivashanmugam, P. (2017). Optimization of lipase production from organic
solid waste by anaerobic digestion and its application in biodiesel production. Fuel Processing
Technology, 165, 1–8.
106. Selvakumar, P., Adane, A. A., Zelalem, T., Hunegnaw, B. M., Karthik, V., Kavitha, S.,
Jayakumar, M., Karmegam, N., Govarthanan, M., & Kim, W. (2022). Optimization of
binary acids pretreatment of corncob biomass for enhanced recovery of cellulose to produce
bioethanol. Fuel, 321, 124060. https://doi.org/10.1016/j.fuel.2022.124060
107. Selvakumar, P., Arunagiri, A., & Sivashanmugam, P. (2019). Thermo-sonic assisted enzy-
matic pre-treatment of sludge biomass as potential feedstock for oleaginous yeast cultivation
to produce biodiesel. Renewable Energy, 139, 1400–1411. https://doi.org/10.1016/j.renene.
2019.03.040
108. Selvakumar, P., Karthik, V., Kumar, P. S., Asaithambi, P., Kavitha, S., & Sivashanmugam, P.
(2021). Enhancement of ultrasound assisted aqueous extraction of polyphenols from waste
fruit peel using dimethyl sulfoxide as surfactant: Assessment of kinetic models. Chemosphere,
263, 128071. https://doi.org/10.1016/j.chemosphere.2020.128071
4 Industrial Organic Waste and Byproducts as Sustainable Feedstock … 115

109. Sharma, H. B., Sarmah, A. K., & Dubey, B. (2020). Hydrothermal carbonization of renewable
waste biomass for solid biofuel production: A discussion on process mechanism, the influ-
ence of process parameters, environmental performance and fuel properties of hydrochar.
Renewable and sustainable energy reviews, 123, 109761.
110. Shibu, C., Chandel, S., & Vats, P. (2023). Scaling up of wood waste utilization for sustainable
green future. In Handbook of research on sustainable consumption and production for greener
economies (pp. 358–383). IGI Global.
111. Shirazi, A., Rahbari, A., Asselineau, C. A., & Pye, J. (2019). A solar fuel plant via super-
critical water gasification integrated with Fischer-Tropsch synthesis: System-level dynamic
simulation and optimisation. Energy Conversion and Management, 192, 71–87.
112. Siddiqui, O., & Dincer, I. (2019). A well to pump life cycle environmental impact assessment
of some hydrogen production routes. International journal of hydrogen energy, 44(12), 5773–
5786.
113. Singh, N. K., Vats, A., Singh, A., Tyagi, A., Mishra, S. K., Kumar, N., & Kumar, S. (2022).
Production of bioethanol from lignocellulosic waste parali. In Role of microbes in industrial
products and processes (pp. 195–222).
114. Singh, R., Das, R., Sangwan, S., Rohatgi, B., Khanam, R., Peera, S. P. G., Das, S., Lyngdoh,
Y. A., Langyan, S., Shukla, A., & Shrivastava, M. (2021). Utilisation of agro-industrial waste
for sustainable green production: A review. Environmental Sustainability, 4(4), 619–636.
115. Siva Sankari, M., Vivekanandhan, S., Misra, M., & Mohanty, A. K. (2022). Oil cakes as
sustainable agro-industrial feedstock for biocarbon materials. ChemBioEng Reviews, 9(1),
21–41.
116. Skendi, A., Zinoviadou, K. G., Papageorgiou, M., & Rocha, J. M. (2020). Advances on the
valorisation and functionalization of by-products and wastes from cereal-based processing
industry. Foods, 9(9), 1243.
117. Sridhar, A., Kapoor, A., Kumar, P. S., Ponnuchamy, M., Balasubramanian, S., & Prabhakar,
S. (2021). Conversion of food waste to energy: A focus on sustainability and life cycle
assessment. Fuel, 302, 121069.
118. Srisowmeya, G., Chakravarthy, M., & Devi, G. N. (2020). Critical considerations in two-stage
anaerobic digestion of food waste—A review. Renewable and Sustainable Energy Reviews,
119, 109587.
119. Srivastava, R. K., Sarangi, P. K., Bhatia, L., Singh, A. K., & Shadangi, K. P. (2022). Conver-
sion of methane to methanol: Technologies and future challenges. Biomass Conversion and
Biorefinery, 12(5), 1851–1875.
120. Sundarsingh, T. J. A., Ameen, F., Ranjitha, J., Raghavan, S., & Shankar, V. (2024). Engineering
microbes for sustainable biofuel production and extraction of lipids—Current research and
future perspectives. Fuel, 355, 129532.
121. Tian, H., Li, J., Yan, M., Tong, Y. W., Wang, C. H., & Wang, X. (2019). Organic waste to
biohydrogen: A critical review from technological development and environmental impact
analysis perspective. Applied Energy, 256, 113961.
122. Tian, H., Wang, X., Lim, E. Y., Lee, J. T., Ee, A. W., Zhang, J., & Tong, Y. W. (2021). Life cycle
assessment of food waste to energy and resources: Centralized and decentralized anaerobic
digestion with different downstream biogas utilization. Renewable and Sustainable Energy
Reviews, 150, 111489.
123. Uddin, M. N., Siddiki, S. Y. A., Mofijur, M., Djavanroodi, F., Hazrat, M. A., Show, P. L.,
Ahmed, S. F., & Chu, Y. M. (2021). Prospects of bioenergy production from organic waste
using anaerobic digestion technology: A mini review. Frontiers in Energy Research, 9, 627093.
124. Ujor, V., Bharathidasan, A. K., Cornish, K., & Ezeji, T. C. (2014). Feasibility of producing
butanol from industrial starchy food wastes. Applied Energy, 136, 590–598.
125. Usman, I., Saif, H., Imran, A., Afzaal, M., Saeed, F., Azam, I., Afzal, A., Ateeq, H., Islam,
F., Shah, Y. A., & Shah, M. A. (2023). Innovative applications and therapeutic potential of
oilseeds and their by-products: An eco-friendly and sustainable approach. Food Science &
Nutrition, 11(6), 2599–2609.
116 D. G. Gizaw et al.

126. Van Nguyen, T. T., Phan, A. N., Nguyen, T. A., Nguyen, T. K., Nguyen, S. T., Pugazhendhi,
A., & Phuong, H. H. K. (2022). Valorization of agriculture waste biomass as biochar: As
first-rate biosorbent for remediation of contaminated soil. Chemosphere, 135834.
127. Van, D. P., Fujiwara, T., Tho, B. L., Toan, P. P. S., & Minh, G. H. (2020). A review of anaerobic
digestion systems for biodegradable waste: Configurations, operating parameters, and current
trends. Environmental Engineering Research, 25(1), 1–17.
128. Varanasi, J. L., & Das, D. (2020). Maximizing biohydrogen production from water hyacinth
by coupling dark fermentation and electrohydrogenesis. International Journal of Hydrogen
Energy, 45(8), 5227–5238.
129. Velusamy, K., Periyasamy, S., Kumar, P. S., Jayaraj, T., Krishnasamy, R., Sindhu, J., Sneka,
D., Subhashini, B., & Vo, D. V. N. (2021). Analysis on the removal of emerging contaminant
from aqueous solution using biochar derived from soap nut seeds. Environmental Pollution,
287, 117632. https://doi.org/10.1016/j.envpol.2021.117632
130. Velusamy, S., Subbaiyan, A., Murugesan, S. R., Shanmugamoorthy, M., Sivakumar, V.,
Velusamy, P., Veerasamy, S., Mani, K., Sundararaj, P., & Periyasamy, S. (2022). Compar-
ative analysis of agro waste material solid biomass briquette for environmental sustainability.
Advances in Materials Science and Engineering, 2022. https://doi.org/10.1155/2022/3906256
131. Veza, I., Said, M. F. M., & Latiff, Z. A. (2021). Recent advances in butanol production by
acetone-butanol-ethanol (ABE) fermentation. Biomass and Bioenergy, 144, 105919.
132. Wang, B., Gupta, R., Bei, L., Wan, Q., & Sun, L. (2023). A review on gasification of municipal
solid waste (MSW): Syngas production, tar formation, mineral transformation and industrial
challenges. International Journal of Hydrogen Energy.
133. Wang, S., Zhao, S., Uzoejinwa, B. B., Zheng, A., Wang, Q., Huang, J., & Abomohra, A.
E. F. (2020). A state-of-the-art review on dual purpose seaweeds utilization for wastewater
treatment and crude bio-oil production. Energy Conversion and Management, 222, 113253.
134. Yankovsky, S. A., Kuznetsov, G. V., Tolokolnikov, A. A., Cherednik, I. V., & Ivanov, A. A.
(2021). Experimental study of the processes of reducing the formation of sulfur oxides during
the co-combustion of particles of metalignitous coal and wood processing waste. Fuel, 291,
120233.
135. Zadeh, Z. E., Abdulkhani, A., & Saha, B. (2021). A comparative production and characteri-
sation of fast pyrolysis bio-oil from populus and spruce woods. Energy, 214, 118930.
136. Zamri, M. F. M. A., Hasmady, S., Akhiar, A., Ideris, F., Shamsuddin, A. H., Mofijur, M.,
Fattah, I. R., & Mahlia, T. M. I. (2021). A comprehensive review on anaerobic digestion of
organic fraction of municipal solid waste. Renewable and Sustainable Energy Reviews, 137,
110637.
137. Zang, G., Sun, P., Elgowainy, A. A., Bafana, A., & Wang, M. (2021). Performance and
cost analysis of liquid fuel production from H2 and CO2 based on the Fischer-Tropsch
process. Journal of CO2 Utilization, 46, 101459.
138. Zena, Y., Periyasamy, S., Tesfaye, M., Tumsa, Z., Jayakumar, M., Mohamed, B. A.,
Asaithambi, P., & Aminabhavi, T. M. (2023). Essential characteristics improvement of
metallic nanoparticles loaded carbohydrate polymeric films—A review. International Journal
of Biological Macromolecules, 124803. https://doi.org/10.1016/j.ijbiomac.2023.124803
139. Zhang, C., Su, H., Baeyens, J., & Tan, T. (2014). Reviewing the anaerobic digestion of food
waste for biogas production. Renewable and Sustainable Energy Reviews, 38, 383–392.
140. Zhang, Q., Xiao, J., Xue, J., & Zhang, L. (2020). Quantifying the effects of biochar application
on greenhouse gas emissions from agricultural soils: A global meta-analysis. Sustainability,
12(8), 3436.
141. Zhang, S., Zhang, H., Liu, X., Zhu, S., Hu, L., & Zhang, Q. (2018). Upgrading of bio-oil
from catalytic pyrolysis of pretreated rice husk over Fe-modified ZSM-5 zeolite catalyst.
Fuel Processing Technology, 175, 17–25.
142. Zhang, Y., Fan, S., Liu, T., Omar, M. M., & Li, B. (2022). Perspectives into intensification for
aviation oil production from microwave pyrolysis of organic wastes. Chemical Engineering
and Processing-Process Intensification, 176, 108939.
Chapter 5
Transesterification of Waste Cooking Oil
Through Microwave Technology: Recent
Advances and Challenges

Bunushree Behera , Kolli Venkata Supraja, S. Mari Selvam, Snehi Kinger,


and Prangya Ranjan Rout

Abstract Undeniably higher feedstock costs are hindering the commercialization of


biodiesel production. Amidst the mentioned problem, there is a current exploration of
alternatives like waste cooking oil (WCO). WCO holds the promise of substantially
reducing production costs by 60–90%, attributed to its cost-effectiveness compared
to vegetable cooking oils and energy crops. Microwave-assisted transesterification
of WCO is a time-efficient process that yields high quality and quantity of fatty
acid methyl esters (FAMEs) with a lower molar ratio of methanol to oil. Besides
the growing interest in using microwaves for WCO transesterification, a compre-
hensive review of the technology for analysing factors affecting biodiesel produc-
tion, and the challenges and limitations for process scale-up and commercializa-
tion is essential which is currently limited. The chapter illustrates the significance
of microwave heating of WCO for biodiesel production. Further, the influence of
parameters on the quality and quantity of biodiesel has also been discussed. The
process energy requirement in comparison with the conventional method has been
analysed. Challenges and limitations associated with technology commercialization
have been annotated. Overall, this chapter enables the researchers to understand the

B. Behera (B) · S. Kinger


Bioprocess Laboratory, Department of Biotechnology, Thapar Institute of Engineering and
Technology, Patiala, Punjab 147004, India
e-mail: bunushree.behera@thapar.edu
K. V. Supraja
Waste Treatment Laboratory, Department of Biochemical Engineering and Biotechnology, Indian
Institute of Technology Delhi, Hauz Khas, New Delhi 110016, India
S. Mari Selvam
Department of Biotechnology and Medical Engineering, National Institute of Technology
Rourkela, Rourkela, Odisha 769008, India
P. R. Rout
Department of Biotechnology, National Institute of Technology Jalandhar, Jalandhar,
Punjab 144027, India

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2024 117
G. Baskar et al. (eds.), Circular Bioeconomy Perspectives in Sustainable Bioenergy
Production, Energy, Environment, and Sustainability,
https://doi.org/10.1007/978-981-97-2523-6_5
118 B. Behera et al.

technology-driven biodiesel production through microwave irradiation of WCO and


identify the potential loopholes to make the process sustainable.

Keywords Biodiesel · Microwave · Sustainability · Transesterification · Waste


cooking oil

5.1 Introduction

Rapidly increasing energy demands of the ever-rising global population have led to
the over-exploitation of fossil-based resources, which are currently on the verge of
depletion. This has also led to an overwhelming increase in carbon dioxide levels in
the atmosphere, resulting in global warming and disturbance of ecological diversity.
There is a pressing demand to find an alternative form of energy from renewable
sources to sort out the above-mentioned issues [1, 2]. Among the available sources,
first-generation alternatives like solar, wind, and hydro energy require very high
capital investment. The third-generation biofuel from microalgae is currently under
investigation due to questionable economics linked with the algal productivity and
lipid yield during scale-up [3]. In case of the second-generation feedstocks often
claimed to be overshadowed by the feed versus fuel dilemma, used cooking oil, also
popularly termed waste cooking oil (WCO), has gained much attention. Cooking
oil obtained basically from plant-based sources, when utilized repeatedly due to
economic reasons, especially in hostels, food outlets, and restaurants, often results
in the mixing of oxygen with unsaturated acylglycerols, thereby forming dimeric/
polymeric acids, which causes an increase in viscosity. The change in physicochem-
ical properties, increase in acidity, and gradual browning with an unpleasant smell
make it harmful during human consumption. Thus, a considerable amount of WCO
(500,000 tonnes) is discarded into the environment annually, causing pollution of
water bodies [4]. Alternatively, the high-free fatty acids (FFAs) containing WCO
can be trans-esterified and converted into biodiesel.
Utilizing the transesterification process for converting the FFAs in WCO would
produce a composite mixture of fatty acid alkyl esters (FAAEs), popularly termed as
the “biodiesel”, with properties similar to diesel as per the international standards.
Acid/alkali-catalysed transesterification reactions under homogenous or heteroge-
neous conditions have been studied by numerous researchers [5–8] for biodiesel
production. Influence of factors like the methanol: oil ratio, reaction time, tempera-
ture, catalyst amount, and governing the yield and properties of biodiesel have been
reported by Aghel et al. [9], Gaur et al. [10], Mahesh et al. [11] and Martinez-Guerra
and Gude [12]. Talebian-Kiakalaieh et al. [13] has enlisted different acid and alka-
line homogenous catalysts for production of biodiesel from WCO. Though alkaline
catalysts are commonly utilized for transesterification of WCO due to good reaction
efficiency, and to obtain better quality biodiesel, high FFA content often might hinder
the reaction due to unwanted saponification. This might result in catalyst deactivation
5 Transesterification of Waste Cooking Oil Through Microwave … 119

and cause tri-acyl glycerol hydrolysis. Thus, it is imperative to use a two-step acid–
alkali-based approach [14], where the first process (esterification would reduce the
FFA content to 1–2%, while the latter would aid in the transesterification into FAAEs
[15]. Several improvements to the existing conventional approaches like the use of
ultrasonic [16, 17]; and microwave treatments [18, 19] applied either individually or
in combination [12] have been researched to achieve an optimum biodiesel yield with
a lower reaction temperature and in lesser time. Ultrasonic processes have shown
improved extraction and transesterification efficiency but often have a longer reac-
tion time, consume more solvent, and are energy-intensive than microwave irradia-
tion [20]. Microwave-assisted extraction and transesterification have gained massive
attention over the past few years due to uniform volumetric heating that results in
better yield and product quality within minimal processing time [21]. Other advanced
techniques of super critical solvent extraction are extremely energy intensive [22].
Microwave heating shows promising improved biodiesel production through esterifi-
cation and transesterification, warranting further research for industrial applications.
Studies by Cheng et al. [23] and Zhang et al. [24] reported that microwave-assisted
transesterification resulted in uniform heating at a lower temperature of 100 °C within
a shorter reaction time of 10–30 min. Using hetero-poly acid as catalysts, the authors
in the latter study reported fatty acid methyl esters (FAMEs) yield of 96.2%, and the
former reported an 86.7% yield. Lokman et al. [25], using microwave irradiation and
heterogeneous carbon-based sulphuric acid catalysts, reported a FAME conversion of
96% in 15 min from palm acid distillate. Microwave-assisted transesterification for
different feedstocks has also been investigated by Azcan and Yilmaz et al. [26], Gina
et al. [27], and Yadav et al. [28]. Microwave-assisted transesterification consumes
0.075 kWh energy, much lower than conventional heating that requires 0.227 kWh,
resulting in high yield of biodiesel with better FAME quality and purity [29]. Fatimah
et al. [30] reported that zirconium oxide (ZrO2 ) heterogeneous catalyst produced
approximately 96% biodiesel yield in case of both conventional and microwave-
assisted transesterification, however later consumes only 1/3rd of the time needed in
conventional method. To date, most studies carried out for microwave-assisted trans-
esterification have been chiefly done at the laboratory scale under batch operational
mode [31]. Gude et al. [32] reported that the microwave process is challenging under
batch operational conditions at the industrial scale. Tangy et al. [33] utilized SrO/
SiO2 catalyst in continuous microwave reactor and projected substantial catalytic
potential at 24.6 min leading to 99.2% FAMEs from WCO. Yang et al. [34] have
utilized magnetic SrO–ZnO/MOF catalyst for one-step biodiesel synthesis. Owing to
the presence of active components of Sr3 Fe2 O6 and ZnO, the catalyst showed basic
and acidic strength. The details on the existing state-of-art of microwave-assisted
transesterification have been reviewed by Mohamad Aziz et al. [21], Motasemi and
Ani [35], Nayak et al. [36], Nomanbhay and Ong [31] and Sajjadi et al. [37]. It
could be interpreted that microwaves offer high-efficiency biodiesel production, with
shorter processing times and improved product quality, paving the way for industrial
adoption through better understanding and optimization.
120 B. Behera et al.

Most of the above-mentioned reviews have summarized the factors influencing the
microwave-assisted process and its effect on the quantity of FAMEs and the physio-
chemical properties of biodiesel. Except for a few, a comprehensive summary of the
economics and environmental impacts is seldom discussed. To expedite the industrial
design and safer usage of microwave-assisted transesterification, an in-depth under-
standing of the influencing parameters along with the underlying mechanism and
its associated economic and environmental impacts is necessary. Thus, the present
chapter holistically and systematically reviewed the mode of action of microwaves
during the transesterification process and enlisted the influencing factors, and FAMEs
yield which could be obtained from WCO. The enviro-economic impacts linked with
the conversion aspects have been discussed in detail. Challenges encompassing the
technical and regulatory requirements are delineated. Overall, the review will act as
a baseline platform to guide potential researchers and stakeholders in implementing
WCO-based conversion into biodiesel.

5.2 Microwave-Mediated Transesterification: Modes


and Mechanism

The conversion of WCO to biodiesel through transesterification process can be


assisted through various techniques such as reactive distillation, non-catalytic super-
critical method, membrane technology, microwave, ultrasonication, biox process,
and reactive distillation [38]. Super-critical transesterification though regarded as a
neat approach, the thermal stability of fatty acid methyl esters (FAMEs) is compro-
mised due to high temperature and pressure. Also, co-solvent, co-catalyst, and sub-
critical method are undoubtedly efficient in conversion but are often limited by the
high cost of chemicals/solvents involved [22]. Among these methods, microwave-
assisted process is energy-intensive and shows higher conversion efficiency, product
quality in shorter duration. The limited solvent requirements, rapid reaction rates,
contactless uniform volumetric heating, and minimal undesired reactions are the
advantages of microwave heating over other techniques [39]. Thus, understanding the
basic mechanism of microwave technology is essential for optimal design and opera-
tion of reactors to achieve process conversion without harmful impacts of microwave
leakage.
Microwave irradiation encompasses high-frequency waves (ranging from 0.3 to
300 GHz), including infrared and radio waves, characterized by wavelengths span-
ning 0.01–1 m. Domestic microwave ovens have a frequency of 2.45 GHz, while
industrial-scale microwave ovens have a range of operation of 915–2450 GHz,
where rapid molecular vibrations of 2450 million times a second occur due
to the change in electric field [21]. Over the years, microwave irradiation has
been increasingly employed in biochemical transformations resulting in biodiesel
production, hydrothermal synthesis, carbonization, and depolymerization reactions
[40]. Microwave irradiation resulting in high-intensity molecular vibrations often
5 Transesterification of Waste Cooking Oil Through Microwave … 121

increases thermal energy resulting in more uniform heating, facilitating more straight-
forward and faster biochemical transformations. The energy content of microwaves
is usually less than most biomolecules and their intermolecular interactions [32].
This radiation induces molecular and interfacial interactions while preserving the
internal structure unchanged [41]. Microwave-based heating has several advantages
over the conventional process resulting in uniform volumetric heat transmission,
avoiding heat loss without direct contact with the source [40]. This feature results
in faster heating of catalysts, and the enhanced interaction between the reactants
for transesterification, thereby providing better efficacy and control. The mechanism
illustrating the difference between microwave and conventional thermal heat transfer
is depicted in the review by Gude et al. [32] and Mohamad Aziz et al. [21]. The
charged molecules or dipoles in the medium often collide and re-orient themselves
with the alternating electric field applied, resulting in rapid frictional interaction
generating thermal energy. The interaction of the thermal energy with the reac-
tant components causes rapid and localized superheating, which is not appropriately
measured through bulk temperature but can be quantified by conductive polarization.
Microwave technology could be utilized in the extractive process and the transester-
ification reaction. Further, microwave reactors can also combine oil extraction and
its subsequent conversion into a single step, thereby reducing the costs and energy
requirements in the process, making it more sustainable. The stoichiometric equation
for the transesterification reaction is given by Eq. 5.1 [42]

WCO + 3AC = G + 3FAE (5.1)

where AC represents alcohol and base catalyst solution, G represents glycerol,


and FAE represents a free fatty acid ester. Glycerol produced during microwave-
assisted process is much less due to higher FAME conversion efficiency. Glycerine
produced is dependent on the type of alcohol used during reaction. Glycerol has been
reported to be produced more with methanol, followed by iso-propyl alcohol, iso-
butyl alcohol, and iso-pentyl alcohol [43]. Saifuddin et al. [44] reported that even the
by-product glycerol obtained can be purified using a combined process of microwave
acidification followed by adsorption using yeast cells immobilized on chitosan.
The mechanism of base-catalysed transesterification commences with the gener-
ation of alkoxide ions (RO−) in preliminary phase. These alkoxide ions proceed to
initiate an attack on carbonyl carbon present in triglyceride resulting in tetrahedral
intermediate [42]. This intermediate undergoes a rearrangement, resulting in fatty
acid esters production through subsequent reaction with alcohol, thereby augmenting
the alkoxide ion population and contributing to the accumulation of fatty acid esters.
The pulse and continuous modes of microwave irradiation are effective for trans-
esterification reactions owing to the higher reaction rates [37]. Molecular orbital
method was employed to elucidate these phenomena. Under microwave irradiation,
the triglyceride’s structure transforms into a flatter conformation leading to a reduc-
tion in its dipole moment [31]. Furthermore, the molecular structure with the lower
dipole moment persists for a brief period following the cessation of irradiation. In the
122 B. Behera et al.

Heat
transfer

CH2OCOR1 CH3 OCOR1

CHOCOR2 CH3 OCOR2

CH2OCOR3 CH3 OCOR3

Microwave
energy
Heat
transfer

Catalyst

Waste cooking oil Methanol Biodiesel Glycerol

Fig. 5.1 Mode and mechanism of microwave-assisted transesterification of waste cooking oil

process of transesterification, a crucial step involves the transition through a tetra-


hedral state around the carboxyl carbon [45]. When triglyceride adopts a flattened
structure with reduced dipole moment, it creates a larger region above it. This struc-
tural alteration makes it more favourable for methoxy group to initiate an attack on
the carboxyl carbon. The mode and mechanism of microwave-assisted transester-
ification of WCO have been given in Fig. 5.1. In microwave-assisted conversion,
biodiesel production can be affected by limitations in mass transfer leading to inef-
ficient transport of reactants. These limitations may arise due to factors such as slow
diffusion, inadequate mixing, or restricted contact between phases, and they often
require optimization of reaction conditions, catalysts, and reactor design to over-
come. To ensure efficient mass transfer inside the reaction mixture, it is important
to optimize reaction parameters with proper mixing and stirring of the reactants,
suitable solvents can be selected, and reactant concentrations can be adjusted with
implementation of monitoring and control systems.

5.3 Factors Influencing Microwave-Assisted


Transesterification of WCO

Various process parameters like alcohol to oil ratio, reaction time, temperature, cata-
lyst concentration, and microwave power influence the transesterification reaction.
Each of these factors is interlinked, and therefore, optimization using response surface
methodology (RSM) based on full factorial design model [18], central composite
design model [46], and Box–Behnken design model [47] have been carried out by
researchers. The different factors that affect the efficiency of microwave-assisted
transesterification of WCO either positively or negatively have been given in Fig. 5.2.
5 Transesterification of Waste Cooking Oil Through Microwave … 123

Fig. 5.2 Factors affecting the microwave-assisted transesterification of waste cooking oil

5.3.1 Effect of Microwave Power and Reaction Temperature

Microwave-assisted transesterification is an endothermic process and is influenced


by reaction temperature. The viscosity of WCO reduces with increasing tempera-
ture, leading to sufficient mixing of reactants during the conversion process [48]. A
microwave power of 200–800 W corresponding to a temperature of 40–80 °C is essen-
tial for transesterification depending on the FFA content of WCO [49]. Increasing
microwave power above the optimal level can lead to the evaporation of alcohol
content, thereby causing a reduction in biodiesel yield and increasing the by-product
content. Hassan and Smith [42], developed a two-step microwave heating method,
where the reactor is initially heated up to 300 W power for 1 min to obtain a temper-
ature of 50 °C. Subsequently, the power is reduced to 100 W for 30 s to attain
60 °C, and then, the reaction time is recorded upon reaching this desired temper-
ature. Although the reaction time is significantly reduced with microwave heating,
a higher catalyst amount is needed to obtain a greater yield of biodiesel in short
time [50]. With optimal conditions, methanol:oil ratio of 7.46:1, catalyst (calcium
diglyceride) amount of 1.03 wt% at temperature of 62 °C, and highest biodiesel yield
of 94.86% are obtained in a short time of 15 min using microwave heating at 850
W [46]. In contrast, using similar reaction parameters with conventional heating, the
biodiesel yield is significantly low with 42.59%. The study also reported an activa-
tion energy of 26.56 kJ/mol for the biodiesel produced through microwave-assisted
transesterification. It is essential to maintain an optimum temperature by adjusting the
microwave power to facilitate the interaction of reactants, resulting in an increased
yield of biodiesel.
124 B. Behera et al.

5.3.2 Effect of Catalyst

Among all the factors, catalyst loading is identified as the most sensitive parameter
affecting the conversion process. The use of catalysts with a dosage of 1–10 wt% in
the transesterification process improves the process efficiency, reducing the reaction
time and temperature required for the conversion process [36]. Elevating the cata-
lyst concentration has the potential to enhance triglyceride conversion, leading to
improved ester content. Inadequate catalyst quantities might result in the incomplete
transformation of triglycerides into esters [51]. Based on its composition, the catalyst
has been divided into homogeneous, heterogeneous, and enzymatic forms.

5.3.2.1 Homogenous Catalysts

Homogenous catalyst has high efficiency with lower reaction time and tempera-
ture, but this process is costly as it involves high purification costs for separating
the catalyst from the reaction mixture [36]. Some of the acid catalysts like H2 SO4 ,
ClSO3 H, H3 CSO3 H, and HCl have been applied in transesterification process leading
to conversion efficiency of up to 85% [52]. Alkali catalysts, including KOH, NaOH,
and CH3 ONa, are soluble in alcohol as they can absorb microwave energy directly
through the –OH group, leading to disruption of the two-tier arrangement of methanol
and oil [53]. According to Ali et al. [47] and Babel et al. [52], the catalyst gener-
ated by activating limestone or coconut shell biochar with KOH exhibits insolubility
in methanol or methyl esters, facilitating their easy separation from the reaction
mixture after the transesterification process, resulting in biodiesel yield of 96% and
91%, respectively. Application of an acid or alkali catalyst in the transesterification
process is based on the FFA% present in WCO [18]. The physical and chemical
properties of WCO like viscosity, specific heat, surface tension, and %FFA content
increase during the course of frying. This is due to the long-term heating at high
temperatures (140–190 °C) and exposure to oxygen. Higher FFA content in WCO
increases the soap formation with an alkali catalyst leading to deactivation of the
catalyst. Hence, to reduce the FFA content, an acid catalyst is used in the conversion
process as a pre-treatment step [36]. Soosai et al. [54] used 5% sulphuric acid-based
esterification process to reduce FFA from 6.87 to 0.83%. For WCO with high %FFA
content, two-step microwave-assisted transesterification process can be carried out
with acid-catalysed esterification for minimizing the %FFA content, followed by
alkali-catalysed reaction for production of biodiesel with an overall conversion effi-
ciency of 95–97% [18, 54]. Some of the challenges associated with homogeneous
catalysts include higher economic costs for separation of the catalyst from the reaction
mixture and requirement of further purification steps to separate glycerol from the
ester phase [50]. Due to these drawbacks, heterogeneous catalysts have been devel-
oped to reduce the process economics as well as to improve the conversion efficiency
without requirement of additional purification steps for biodiesel production.
5 Transesterification of Waste Cooking Oil Through Microwave … 125

5.3.2.2 Heterogenous Catalysts

Heterogeneous catalysts are process-efficient, with easy separation, longer shelf


life and reduced equipment corrosion [50]. These include various hetero-polyacids
(HPA), metal oxides, acidic polymer resins, and natural substances like carbon,
zeolites, or lignin-derived materials [36]. Nazir et al. [48] utilized a heterogeneous
nano-catalyst with significantly high surface area of 30.31 m2 /g and pore size of
9.44 nm, with an acid density of 4.74 mmol/g. The catalyst had a good number
of functional groups (sulphonic acid, carboxyl, and hydroxyl) and was reported to
lose only 11% stability after 6 cycles of transesterification. Similarly, lignin-based
catalysts have also been used by Yadav et al. [55, 56]. The use of biomass derived
catalyst has been projected to make the microwave-assisted transesterification less
energy intensive [57]. Because of high basicity, researchers have also focussed on
strontium oxide (SrO) and zinc oxide (ZnO)-based bifunctional acid/base catalysts
like Sr/ZrO2 , ZnO-La2 O3 , Sr-Ti, SrO-ZnO/Al2 O3, and SrO-ZnO/MOF that showed
a transesterification efficiency of 79–98% towards conversion of highly acidic WCO
to biodiesel [34, 58]. Compared to others, SrO-based heterogeneous catalysts aid in
better penetration of microwave irradiation, thereby enhancing the transesterification
process [33]. 0.3% CaO-based catalysts, with 18:1 alcohol: oil ratio, at 270 W irradi-
ation, after 114 s have been reported to yield 97.4% biodiesel [54]. Earlier studies by
Khemthong et al. [59] have also shown the presence of 99.2% wt. of CaO in egg shells,
with high strength basic sites and stability, that would result in increased catalytic
heating during transesterification. Buasri et al. [60] projected the use of alkali-based
catalysts prepared via calcination of animal bones for microwave-assisted transester-
ification, as an economic method. Heterogeneous catalysts are found to be more effi-
cient than homogeneous catalysts considering the catalytic activity, energy efficiency,
biodiesel yield, reaction time, and process economics.

5.3.2.3 Enzymatic Catalysts

Enzymatic catalysts are mainly fungi or bacteria-derived lipase proteins that offer
easy catalyst separation from the reaction mixture, and are active at 40–60 °C
[36]. However, there are limitations to this process as it takes greater reaction time
for conversion, and there may be chances of thermal deactivation of the enzyme
[61]. When solid catalysts are used in microwave-mediated reaction processes, they
exhibit better results within shorter reaction times compared to conventional heating
methods using the similar catalysts [37]. Additionally, the microwave heating mech-
anism enhances reaction kinetics by increasing the rate constant values, improving
conversions.
126 B. Behera et al.

5.3.3 Effect of Alcohol to Oil Ratio

Methanol and ethanol are widely used alcohols for the transesterification process
because they can transform microwave energy into heat, leading to rapid boiling in a
short time [35, 62]. Maintaining an appropriate molar ratio of methanol to oil is essen-
tial for ensuring a sufficient supply of methanol to effectively break the glycerol-fatty
acid linkages [63]. Tangy et al. [33] reported that high methanol: oil ratio increases
the contact area for triglyceride moieties and alcohol facilitating transesterification.
Comparatively, higher molar ratio of methanol to oil is needed for acid-catalysed
transesterification than for alkali-catalysed reaction under the same experimental
conditions [53]. At a constant catalyst dosage of 1 wt% and fixed microwave power
of 250 W, decreasing the methanol-to-oil molar ratio and prolonging the reaction
time aims to favour a reversible reaction, consequently leading to a reduction in the
overall yield of the process [18]. To move the transesterification reaction forward,
a methanol-to-oil molar ratio of minimum 3:1 to a maximum of 18:1 is essential
depending upon the %FFA content of WCO to obtain a biodiesel yield of 85–98%
[42]. However, presence of excess methanol in the reaction mixture could lead to
slower conversion process and greater settling time, while increasing the solubility
of glycerol with the ester [29]. Enzymatic microwave transesterification processes
are greatly affected by methanol leading to reduction in the enzymatic activity. Other
solvents such as butanol or dimethyl carbonate at a molar ratio of 6:1 have been used
for transesterification to achieve a biodiesel yield of 83–94% [61].

5.3.4 Effect of Reaction Time

Conventional transesterification processes based on reactor heating require a reac-


tion time of 30 min to 180 min for biodiesel production [64]. Microwave irradiation
is attributed to the direct absorption of radiation by the polar –OH group in alcohol.
This absorption leads to immediate excitation of the –OH group by microwave radi-
ation, causing a significantly higher local temperature than its surroundings [10].
Consequently, microwave-assisted conversion offers a reaction time of 3–30 min for
complete conversion [37]. At an equivalent reaction time of 15 min, conventional
heating resulted in a biodiesel yield of 42.59%, while microwave heating showed a
significantly higher yield of 94.86% [46]. Reaction time also varies with the temper-
ature and catalyst used during the process. Acid-catalysed conversion process with
2 wt% H2 SO4 , at a reaction temperature of 60 °C and reaction time of 200 min,
resulted in a biodiesel yield of 24% only [42]. While for alkali-catalysed reaction at
a temperature of 60 °C with 1 wt% KOH, a biodiesel yield of 93–98% is achieved
within 60 min of reaction time. Often, an optimal reaction time provides an adequate
contact time between irradiation and reactants, promoting interaction between glyc-
eride moieties of oil and methanol [33]. Enzymatic microwave transesterification
processes usually require a greater reaction time of 240 min for obtaining a biodiesel
5 Transesterification of Waste Cooking Oil Through Microwave … 127

yield of up to 94%. Although the reaction time is very high, enzyme reusability and
sustainability of the process make the process more economically viable [61]. These
are the significant parameters influencing the transterification efficiency of WCO
through microwave and the few related studies reported has been given in Table 5.1.

5.4 Enviro-Economic Feasibility of Microwave-Assisted


Transesterification of WCO

Although biodiesel production from WCO is considered an acceptable option to


produce alternative energy, the scale-up requirements often need validation of their
economic feasibility and environmental impact assessment. One of the earliest studies
by Marchetti et al. [69] reported that a significant fraction of biodiesel production
costs (76–89%) are linked with feedstock/raw material prices. Thus, exploring WCO
or recycled oil can bring down the overall costs. The study also emphasized using
heterogeneous catalyst systems for transesterification since the homogenous catalyst
systems involve more equipment and process costs. The scale of operation and the
plant capacity play a significant role in deciding the selling price of WCO-based
biodiesel, either in neat or blended forms. Hussain et al. [70] reported the price
of biodiesel produced at a rate of 150 kt/year via transesterification of WCO with
15 years’ lifetime to be higher than that of conventional diesel in the United Arab
Emirates. Similar reports of biodiesel prices being 3.7 times higher than conventional
diesel were also projected by Mohammadshirazi et al. [71] for Iran. Recent studies by
Farid et al. [72] and Liu et al. [73] have conducted the techno-economic assessment
for biodiesel production from WCO. Farid et al. [72] extended the process of alkaline
transesterification resulting in 74.3% biodiesel yield with 96.9% purity, to energy and
economic feasibility analysis using engineering principles and multipliers, consid-
ering a production capacity of 3.68 kt/annum. The authors have reported a net energy
ratio (NER) of 1.15, with a net present value (NPV) of 1.43 million US $ at a 60%
internal rate of returns (IRR) for B10 blends considering a lifetime of 10 years. Liu
et al. [73] analysed the process feasibility through life-cycle cost analysis in Aspen
Plus software via simulating a 2-step transesterification consisting of pre-treatment
followed by alkali-catalysed transesterification of WCO, resulting in biodiesel with
99.6% purity. The cost of biodiesel was projected to be 65.28% higher than that of
conventional diesel. Owing to the higher content of impurities and free fatty acids
(FFA), the processing and pre-treatment costs account for 56.26% of total production
costs. Thus, the use of a single-step process co-combined with advanced techniques
like microwave-assisted transesterification under optimized conditions would help
in making the technology feasible during scale-up.
Refaat et al. [74] compared the conventional and microwave-assisted transesterifi-
cation processes of high FFA containing WCO and reported 97% and 94% reductions
in reaction and separation time, respectively, with quality similar to the biodiesel
produced during the conventional process. Reduced treatment and processing time
Table 5.1 Studies reported on microwave-assisted transesterification of waste cooking oil for biodiesel production
128

Feedstock Alcohol: oil Reaction Catalyst dosage Microwave Reactor Yield References
ratio time power/
temperature
WCO (8.17% 1st step MeOH: 35 s H2 SO4 —1 wt% 250 W, 65 °C Domestic microwave 94.6% Supraja et al.
FFA) oil—19.57:1 oven (0.082% [18]
FFA)
2nd step 30 min NaOH—2 wt%
MeOH:
oil—5:1
WCO MeOH: 55.26 min KOH-activated 900 W, 65 °C Continuous 96.65% Ali et al. [47]
oil—12.21:1 limestone—5.47 wt% microwave-assisted (0.38%
reactor (CMAR) FFA)
WCO (23% MeOH: 60 min KOH—1 wt% 400 W, 60 °C Multimode microwave 97.4% Hassan and
FFA) oil—0.3 v/v apparatus (CEM Smith [42]
MARS)
WCO MeOH: 6 min KOH—2 wt% 800 W, 80 °C Domestic microwave 92% Prafulla et al.
oil—9:1 oven [65]
WCO MeOH: 6 min BaO—2 wt% 800 W, 80 °C Domestic microwave 96% Prafulla et al.
oil—12:1 oven [65]
WCO DMC: oil—6:1 240 min Lipase 435—10 wt% 70 W Domestic microwave 94% Panadare and
oven Rathod [61]
WCO (0.46% Alcohol: 1 min KOH—1 wt% 800 W Domestic microwave 89.29% Zare et al. [66]
FFA) oil—6:1 oven
WCO (0.68% MeOH: 40 min KOH-activated coconut shell 80 °C Domestic microwave 91.3% Babel et al. [52]
FFA) oil—12:1 biochar—5 wt% oven
(continued)
B. Behera et al.
Table 5.1 (continued)
Feedstock Alcohol: oil Reaction Catalyst dosage Microwave Reactor Yield References
ratio time power/
temperature
Inedible olive MeOH: 9 min KOH—1.2 wt% 500 W Domestic microwave 93.52% Dehghan et al.
oil (2.36% oil—9:1 oven [29]
FFA)
Waste frying MeOH: 7 min NaOH—1 wt% 312 W, 65 °C Domestic microwave 88.63% Yakoob et al.
palm oil oil—12:1 oven [67]
WCO MeOH: 4.47 min NaOCH3 —0.68 wt% 720 W, 65 °C Continuous 97.13% Ali et al. [68]
oil—11.62:1 microwave-assisted
reactor (CMAR)
WCO MeOH: 15 min Lignin (–SO3 H, –COOH and 60 °C Microwave reactor 89.19% Nazir et al. [48]
oil—18:1 –OH)—15 wt% under continuous
stirring
WCO MeOH: 20 min Corn cobs-derived sulfonated 600 W, 75 °C Microwave irradiated 88.7% Rocha et al. [57]
(soyabean oil) oil—6:1 activated carbon—20 wt% unit with a reflux
condenser
WCO MeOH: 10 min Yellow horn oil utilizing 500 W, 60 °C MARS-II microwave 96.22% Zhang et al. [24]
oil—12:1 hetero-polyacid catalyst accelerated reaction
system
5 Transesterification of Waste Cooking Oil Through Microwave …

Cs2.5 H0.5 PW12 O40 —1 wt%


WCO MeOH: 20 min Guinea fowl bone-based 800 W, 65 °C Domestic microwave 95.82% Singh and
oil—18:1 heterogeneous catalyst- 4 wt% oven Sharma [49]
WCO MeOH: 15 min Calcium diglyceride—1.03 62 °C Multi-wave PRO 94.86% Gupta and
oil—7.46:1 wt% microwave reactor Rathod [46]
WCO MeOH: 5 min Natural Hydroxyapatite 800 W Domestic microwave 94% Buasri et al. [60]
oil— 18:1 (NHAp) derived from pork oven
bone—4 wt%
(continued)
129
Table 5.1 (continued)
130

Feedstock Alcohol: oil Reaction Catalyst dosage Microwave Reactor Yield References
ratio time power/
temperature
Palm oil MeOH: 4 min Industrial eggshell wastes 900 W Domestic microwave 96.7% Khemthong
oil—18:1 containing CaO—15 wt% oven et al. [59]
Non-edible MeOH: 114 s Calcium oxide (CaO)—0.3 270W Domestic microwave 97.4% Soosai et al. [54]
silk-cotton oil—18:1 wt% oven
seed oil
Oleic acid MeOH: oleic 60 min Lignin-rich biomass 50 W, 85 °C Discover SP microwave 99.01% Yadav et al. [56]
acid—16:1 (WNS-SO3 H)—9 wt% system
Oleic acid MeOH: oleic 60 min Oryza sativa catalyst—8 wt% 50 W, 80 °C Discover SP microwave 99.6% Yadav et al. [55]
acid—24:1 system
WCO MeOH: 2 min SrO—1.25 wt% 242 W, 65 °C Flow SYNTH 94.5% Tangy et al. [33]
oil—12:1 microwave system
B. Behera et al.
5 Transesterification of Waste Cooking Oil Through Microwave … 131

are expected to have an economic advantage during the overall process. Patil et al.
[19] reported lower energy requirements of 255 kJ during microwave-assisted extrac-
tion transesterification of Nannochloropsis sp. than the supercritical process requiring
600 kJ of energy. A promising aspect of using industrial-scale continuous microwave
reactors for biodiesel production with positive energy returns has been reported
by Muley and Bolder [75]. The use of single-step microwave extraction and acid-
catalysed transesterification of Rhodotorula glutinis resulted in FAMEs, with the
energy requirement of the microwave being 20% less than that of the energy content
of biodiesel [76]. The study also projected the energy return on investment to be
significantly influenced by reaction time than the process temperature. Waudby and
Zein [77] projected that the feasibility of microwave-assisted technology is possible
with the inclusion of recycled streams and an optimal combination of separation
processes. Reviews by Manikandan et al. [78], Tibesigwa et al. [79], and Nayak
et al. [36] have summarized the state-of-art linked with the techno-economic aspects
of biodiesel production. To the best of the author’s knowledge, a limited number
of studies have been done with respect to the techno-economic feasibility of the
transesterification of WCO. Nevertheless, the insights from the factors influencing
industrial-scale microwave technology for different feedstocks provide its potential
for scalability and could also be extrapolated for WCO.
Environmental assessment of WCO conversion into biodiesel is a crucial aspect to
be considered for process scale-up. Chung et al. [80] reported reduced environmental
impacts linked with using WCO as feedstock compared to that of the Jatropha due to
the absence of agricultural processing involved. Further, using an eggshell-derived
calcium oxide (CaO) based catalyst had a much lower overall environmental impact
than homogenous KOH-based transesterification due to the processing of chemi-
cals during the production and waste streams disposal involved thereof. A detailed
LCA study by Foteinis et al. [81] compared the environmental impacts of biodiesel
produced from WCO with those from the first and second generations. The biodiesel
from WCO was found to be more environmentally sustainable with 0.55 tonnes CO2
eq. and 58.37 Pt environmental footprints per tonne of biodiesel produced. Authors
have projected these values to be approximately 40% lower than the first-generation
biodiesel and almost an order of magnitude than the third-generation microalgal
biofuel. Also, a threefold reduction in environmental impacts was reported by authors
as compared to petroleum-based fuel. The non-renewable energy input during the
processing, along with the use of chemicals, is the primary cause of environmental
impacts. Iglesias et al. [82] projected that biodiesel production from WCO is environ-
mentally sustainable when operated in a centralized territory rather than at a decen-
tralized level. Though conversion of WCO is linked with a decrease in environmental
impacts, most emissions are linked with the chemicals and energy involved during
the transesterification step [83]. Alishahi et al. [84] compared the properties and the
associated environmental impacts of biodiesel produced from WCO via magnetic
stirrer esterification and microwave-assisted esterification, projecting their physico-
chemical properties and heating rate to be similar, however, the energy linked in the
latter case was 4-times lower than the former. The energy ratio of 1.04, efficiency
of 0.0178 kg/MJ, and the intensity of 56.07 MJ/kg were observed by Yari et al. [85]
132 B. Behera et al.

for the biodiesel produced from waste fish oil. Vionery-Portillo et al. [86] reported
that the B25 blend of biodiesel from WCO and conventional low-sulphur diesel used
in a 33 kW power generator had almost 39% reductions in eutrophication poten-
tial, acidification potential, abiotic depletion, and potential for human toxicity. Also,
the global warming potential declined by 36%. A decrease in NOx by 42% and a
52% increase in CO was also observed in the exhaust emissions compared to the
low-sulphur diesel. Limitations and challenges linked with the LCA of WCO into
biodiesel, along with the potential research opportunities, have been summarized in
detail by Hosseinzadeh-Bandbafha et al. [87].

5.5 Commercialization of WCO-Based Biodiesel:


Challenges and Future Outlook

WCO is undoubtedly considered an ideal feedstock for conversion into second-


generation biofuel owing to the issues linked with its subsequent use in cooking
resulting in health hazards and environmental impacts associated with improper
disposal. However, the commercialization process is hindered by several technical
and socio-economic factors.
Over the years, even though technical advances have improved yield and brought
down the price of biodiesel, still presently projected costs are higher than the existing
counterparts, thereby questioning its commercial feasibility. Microwave-based reac-
tors are often costly in terms of production, compared to the chemical synthesis reac-
tors. Scaling-up is technically hindered by the development of spark or electric arc and
sudden surge in temperature or localized hotspots due to thermal leakage [88]. Lack
of information on dielectric properties of materials makes the process of optimization
difficult, thereby declining the product yield and its subsequent properties. Shorter
wavelength, low frequency with higher penetrative capacity, might result in poten-
tial health hazards when operated for prolonged interval of time. Also, supply chain
management is another limiting factor hindering commercialization [15]. Smaller
restaurants in several countries are seldom equipped with facilities to recover and
separate the WCO for transport to adequate processing sites converting them into
biofuel. Often insufficient daily discharge, concomitant high equipment costs, and
lack of space are the primary cause of the unavailability of stringent supply chain,
especially in developing countries. Thus, a significant fraction of WCO is disposed
untreated due to the absence of stringent rules and policies. Few countries like the
USA, Japan, and EU have enforced regulatory frameworks governing the disposal
of WCO. The demand of WCO in developing countries like China, India, Indonesia,
Japan, Republic of Korea and Malaysia has increased recently. Collectively, these
countries collect 3700–5000 kT/year [89]. Liu et al. [90] during a survey in urban
areas of Beijing reported 90.14 thousand tonnes produced in 2016, out of which 24%
hotels do not submit it to the collectors, and thus 21.63 thousand tonnes are unac-
counted. With economic incentives by government of China and subsidy value of 4
5 Transesterification of Waste Cooking Oil Through Microwave … 133

Yuan RMB/kg, the above problem can be avoided. There is an ardent need to extrap-
olate these laws and regulations and impose subsidies on the collection, recovery,
and treatment facilities for WCO in developing countries. Appropriate recycling
infrastructure and sales price of biofuel would definitely attract stakeholders and
beneficiaries. The comparative analysis of standards of waste cooking oil in relation
to biodiesel has been given in Table 5.2.
Another limiting feature hindering the commercialization of WCO-based
biodiesel is its varying physicochemical properties in each batch. This variation
mainly results from the differences in the FFA content, heat exposure time, compli-
cated oil and food constituents, and their degraded counterparts. Thus, a stringent
selective procedure must be used to pre-treat the raw WCO to bring its parame-
ters under the appropriate limit before transesterification [94]. The emergence of
advanced processes like microwave-assisted techniques under optimized conditions
would aid in adequate sample preparation. Also, there is a need of better equipment
for characterization and appropriate analysis of the physiochemical properties of
samples for efficient conversion. The development of accurate temperature sensors
is required to reduce the potential errors in transesterification.
Though the energetics for microwave-assisted transesterification have been inves-
tigated, the number of studies linked with the economics and environmental impacts
of the aspects is minimal. Implementing technology to scale up WCO conversion
into biodiesel requires a stringent scrutinization of the economics involved, from
raw materials procurement and processing to its pre-treatment and microwave-based
conversion into biodiesel. Thus, more techno-economic studies must be conducted
considering microwave-assisted processes under optimized conditions at the semi-
industrial scale. Environmental impact assessment studies must be carried out without
zero-burden assumptions, as highlighted by [87], since the waste by-products gener-
ated during the transesterification process can be utilized as raw materials in other
industrial processes. The process scale-up of microwave-assisted transesterification
for substantial biodiesel production has to be extrapolated from lab-scale studies.
These studies would aid in convincing policymakers about the sustainability aspects
of microwave-based conversion of WCO into biodiesel.

5.6 Conclusions

WCO, as a second-generation feedstock for biofuel production, can curb the growing
energy demands, avoiding the unwanted environmental pollution caused by their
inadequate disposal. Microwave-based technology could facilitate the transesterifi-
cation of WCO with much lower energy requirements and in lesser process time. Opti-
mization of process conditions like the microwave power, alcohol to oil ratio, reaction
time, and catalyst dosage for microwave-assisted transesterification is expected to
improvise the process efficiency resulting in a better yield and quality of biodiesel
with FAMEs suitable for engine applications. Better conversion efficiency could be
obtained with the utilization of heterogeneous solid catalyst with capacity to adsorb
Table 5.2 Comparative analysis of biodiesel properties for waste cooking oil
134

WCO/ Density (g/ Viscosity % FFA Calorific Saponification Acid Iodine Cetane Flash Water References
Biodiesel cm3 ) (mm2 /s) value (MJ/ value (mg of value value no point content
kg) KOH/g oil) (mg (mg (°C) (%)
KOH/g KOH/g
oil) oil)
WCO – 28.8 – 44.44 186.3 17.41 – 32.48 – – Prafulla et al.
[65]
WCO 0.920 38.41 – – 204.4 2.45 – – – 0.5 Gupta and
Rathod [46]
WCO 0.910 48.10 8.17 – 193.9 16.24 193.04 – – – Supraja et al.
[18]
WCO 0.904 49.05 9.92 38.59 – 2.3 – – – – Milano et al.
[91]
WCO 0.910 37.7 3.75 – – 7.5 – – 295 0.12 Nazir et al.
[48]
WCO 0.9206 4.38 – 39.8 199.8 0.8 124.05 – 282 0.04 Azcan and
Yilmaz [26]
WCO – 4.2 1.47 – 181 2.57 103 – – 0 Selvaraj et al.
[92]
ASTM 0.87–0.89 1.9–6.0 < 0.45 < 370 < 0.5 < 115 > 47 > 130 < 0.05% Motasemi and
D6751 Ani [35]
standard for
biodiesel
EN 14,214 0.86–0.90 3.5–5.0 – – – < 0.5 < 120 > 51 > 101 < 0.05% Motasemi and
standard for Ani [35]
biodiesel
(continued)
B. Behera et al.
Table 5.2 (continued)
WCO/ Density (g/ Viscosity % FFA Calorific Saponification Acid Iodine Cetane Flash Water References
Biodiesel cm3 ) (mm2 /s) value (MJ/ value (mg of value value no point content
kg) KOH/g oil) (mg (mg (°C) (%)
KOH/g KOH/g
oil) oil)
IS 15607 0.86–0.9 2.5–6.0 – – – < 0.5 – > 51 > 120 < 0.05% Rao et al. [93]
standard for
biodiesel
Biodiesel – 2.6 – 45 – – – 46 – – Prafulla et al.
[65]
Biodiesel 0.87 4.85 – – – 0.3 – – 165 – Gupta and
Rathod [46]
Biodiesel 0.892 4.68 0.02 – 110.6 0.04 112.02 – – – Supraja et al.
[18]
Biodiesel 0.862 5.01 – 38.766 – 0.13 – – 154 – Milano et al.
[91]
Biodiesel 0.83 3.92 – – – 0.39 – 49.34 133 0.02 Nazir et al.
[48]
Biodiesel 0.8883 2.35 – 38.529 – – – – 185 0.03 Azcan and
5 Transesterification of Waste Cooking Oil Through Microwave …

Yilmaz [26]
Biodiesel – 5.2 0.48 – 186 0.96 111 – – 0 Selvaraj et al.
[92]
Biodiesel 0.87 4.2 – 36.6 – 0.34 – – 183 – Sharma et al.
[64]
135
136 B. Behera et al.

microwave energy, thereby converting them into heat. Economic and environmental
assessment studies on microwave-assisted transesterification must be focussed on
scale-up and implementation of semi-industrial or pilot-scale processes. More studies
on combining microwave-assisted transesterification with supercritical and sub-
critical methods, must be carried to achieve better diesel yield and quality. Stringent
rules and policies on properly separating WCO from the source, transportation, and
biochemical conversion would facilitate its real-time application. Attractive subsi-
dies, market initiatives, and an appropriate sale price for biodiesel would encourage
investors, thereby making upscaling and commercialization feasible.

Acknowledgements The authors acknowledge the Department of Biotechnology, Thapar Institute


of Engineering and Technology, Patiala, the Department of Biochemical Engineering and Biotech-
nology, Indian Institute of Technology Delhi, the Department of Biotechnology and Medical Engi-
neering, National Institute of Technology Rourkela and the Department of Biotechnology, and
National Institute of Technology Jalandhar for providing the requisite facilities to carry out the
study.

References

1. Behera, B., Acharya, A., Gargey, I. A., Aly, N., & Balasubramanian, P. (2019). Bioprocess
engineering principles of microalgal cultivation for sustainable biofuel production. Bioresource
Technology Reports, 5, 297–316.
2. Maity, J. P., Bundschuh, J., Chen, C. Y., & Bhattacharya, P. (2014). Microalgae for third gener-
ation biofuel production, mitigation of greenhouse gas emissions and wastewater treatment:
Present and future perspectives—A mini review. Energy, 78, 104–113.
3. Behera, B., Unpaprom, Y., Ramaraj, R., Maniam, G. P., Govindan, N., & Paramasivan, B.
(2021). Integrated biomolecular and bioprocess engineering strategies for enhancing the lipid
yield from microalgae. Renewable and Sustainable Energy Reviews, 148, 111270.
4. Yacob, M. R., Kabir, I., & Radam, A. (2015). Households willingness to accept collection
and recycling of waste cooking oil for biodiesel input in Petaling District, Selangor, Malaysia.
Procedia Environmental Sciences, 30, 332–337.
5. Ali, C. H., Asif, A. H., Iqbal, T., Qureshi, A. S., Kazmi, M. A., Yasin, S., Danish, M., & Mu, B.
Z. (2018). Improved transesterification of waste cooking oil into biodiesel using calcined goat
bone as a catalyst. Energy Sources, Part A: Recovery, Utilization, and Environmental Effects,
40(9), 1076–1083.
6. Borah, M. J., Das, A., Das, V., Bhuyan, N., & Deka, D. (2019). Transesterification of waste
cooking oil for biodiesel production catalyzed by Zn substituted waste egg shell derived CaO
nanocatalyst. Fuel, 242, 345–354.
7. Maneerung, T., Kawi, S., Dai, Y., & Wang, C. H. (2016). Sustainable biodiesel production via
transesterification of waste cooking oil by using CaO catalysts prepared from chicken manure.
Energy Conversion and Management, 123, 487–497.
8. Rezania, S., Oryani, B., Park, J., Hashemi, B., Yadav, K. K., Kwon, E. E., Hur, J., & Cho,
J. (2019). 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.
9. Aghel, B., Mohadesi, M., Ansari, A., & Maleki, M. (2019). Pilot-scale production of biodiesel
from waste cooking oil using kettle limescale as a heterogeneous catalyst. Renewable Energy,
142, 207–214.
5 Transesterification of Waste Cooking Oil Through Microwave … 137

10. Gaur, A., Mishra, S., Chowdhury, S., Baredar, P., & Verma, P. (2021). A review on factor
affecting biodiesel production from waste cooking oil: An Indian perspective. Materials Today:
Proceedings, 46, 5594–5600.
11. Mahesh, S. E., Ramanathan, A., Begum, K. M. S., & Narayanan, A. (2015). Biodiesel produc-
tion from waste cooking oil using KBr impregnated CaO as catalyst. Energy Conversion and
Management, 91, 442–450.
12. Martinez-Guerra, E., & Gude, V. G. (2014). Synergistic effect of simultaneous microwave and
ultrasound irradiations on transesterification of waste vegetable oil. Fuel, 137, 100–108.
13. Talebian-Kiakalaieh, A., Amin, N. A. S., & Mazaheri, H. (2013). A review on novel processes
of biodiesel production from waste cooking oil. Applied Energy, 104, 683–710.
14. Ho, K. C., Chen, C. L., Hsiao, P. X., Wu, M. S., Huang, C. C., & Chang, J. S. (2014). Biodiesel
production from waste cooking oil by two-step catalytic conversion. Energy Procedia, 61,
1302–1305.
15. Goh, B. H. H., Chong, C. T., Ge, Y., Ong, H. C., Ng, J. H., Tian, B., Ashokkumar, V., Lim,
S., Seljak, T., & Józsa, V. (2020). Progress in utilisation of waste cooking oil for sustainable
biodiesel and biojet fuel production. Energy Conversion and Management, 223, 113296.
16. Gupta, A. R., Yadav, S. V., & Rathod, V. K. (2015). Enhancement in biodiesel production
using waste cooking oil and calcium diglyceroxide as a heterogeneous catalyst in presence of
ultrasound. Fuel, 158, 800–806.
17. Topare, N. S., & Patil, K. D. (2021). Biodiesel from waste cooking soybean oil under
ultrasonication as an alternative fuel for diesel engine. Materials Today: Proceedings, 43,
510–513.
18. Supraja, K. V., Behera, B., & Paramasivan, B. (2020). Optimization of process variables on
two-step microwave-assisted transesterification of waste cooking oil. Environmental Science
and Pollution Research, 27, 27244–27255.
19. Patil, P. D., Gude, V. G., Mannarswamy, A., Cooke, P., Nirmalakhandan, N., Lammers, P., &
Deng, S. (2012). Comparison of direct transesterification of algal biomass under supercritical
methanol and microwave irradiation conditions. Fuel, 97, 822–831.
20. Cintas, P., Mantegna, S., Gaudino, E. C., & Cravotto, G. (2010). A new pilot flow reactor
for high-intensity ultrasound irradiation. Application to the synthesis of biodiesel. Ultrasonics
Sonochemistry, 17(6), 985–989.
21. Mohamad Aziz, N. A., Yunus, R., Kania, D., & Abd Hamid, H. (2021). Prospects and chal-
lenges of microwave-combined technology for biodiesel and biolubricant production through
a transesterification: A review. Molecules, 26(4), 788.
22. Salaheldeen, M., Mariod, A. A., Aroua, M. K., Rahman, S. A., Soudagar, M. E. M., & Fattah,
I. R. (2021). Current state and perspectives on transesterification of triglycerides for biodiesel
production. Catalysts, 11(9), 1121.
23. Cheng, J., Huang, R., Li, T., Zhou, J., & Cen, K. (2014). Biodiesel from wet microalgae:
Extraction with hexane after the microwave-assisted transesterification of lipids. Bioresource
Technology, 170, 69–75.
24. Zhang, S., Zu, Y. G., Fu, Y. J., Luo, M., Zhang, D. Y., & Efferth, T. (2010). Rapid microwave-
assisted transesterification of yellow horn oil to biodiesel using a heteropolyacid solid catalyst.
Bioresource Technology, 101(3), 931–936.
25. Lokman, I. M., Rashid, U., & Taufiq-Yap, Y. H. (2015). Microwave-assisted methyl ester
production from palm fatty acid distillate over a heterogeneous carbon-based solid acid catalyst.
Chemical Engineering and Technology, 38(10), 1837–1844.
26. Azcan, N., & Yilmaz, O. (2013). Microwave assisted transesterification of waste frying oil and
concentrate methyl ester content of biodiesel by molecular distillation. Fuel, 104, 614–619.
27. Gina, M. H., Valange, S., Barrault, J., Moreno, J. A., & López, D. P. (2014) Effect of microwave-
assisted system on transesterification of castor oil with ethanol. Universitas Scientiarum , 19(3),
193–200.
28. Yadav, G. D., Hude, M. P., & Talpade, A. D. (2015). Microwave assisted process intensification
of lipase catalyzed transesterification of 1, 2 propanediol with dimethyl carbonate for the green
synthesis of propylene carbonate: Novelties of kinetics and mechanism of consecutive reactions.
Chemical Engineering Journal, 281, 199–208.
138 B. Behera et al.

29. Dehghan, L., Golmakani, M. T., & Hosseini, S. M. H. (2019). Optimization of microwave-
assisted accelerated transesterification of inedible olive oil for biodiesel production. Renewable
Energy, 138, 915–922.
30. Fatimah, I., Rubiyanto, D., Taushiyah, A., Najah, F. B., Azmi, U., & Sim, Y. L. (2019). Use of
ZrO2 supported on bamboo leaf ash as a heterogeneous catalyst in microwave-assisted biodiesel
conversion. Sustainable Chemistry and Pharmacy, 12, 100129.
31. Nomanbhay, S., & Ong, M. Y. (2017). A review of microwave-assisted reactions for biodiesel
production. Bioengineering, 4(2), 57.
32. Gude, V. G., Patil, P., Martinez-Guerra, E., Deng, S., & Nirmalakhandan, N. (2013). Microwave
energy potential for biodiesel production. Sustainable Chemical Processes, 1, 1–31.
33. Tangy, A., Pulidindi, I. N., Perkas, N., & Gedanken, A. (2017). Continuous flow through a
microwave oven for the large-scale production of biodiesel from waste cooking oil. Bioresource
Technology, 224, 333–341.
34. Yang, J., Cong, W. J., Zhu, Z., Miao, Z. D., Wang, Y. T., Nelles, M., & Fang, Z. (2023).
Microwave-assisted one-step production of biodiesel from waste cooking oil by magnetic
bifunctional SrO–ZnO/MOF catalyst. Journal of Cleaner Production, 395, 136182.
35. Motasemi, F., & Ani, F. N. (2012). A review on microwave-assisted production of biodiesel.
Renewable and Sustainable Energy Reviews, 16(7), 4719–4733.
36. Nayak, S. N., Bhasin, C. P., & Nayak, M. G. (2019). A review on microwave-assisted trans-
esterification processes using various catalytic and non-catalytic systems. Renewable Energy,
143, 1366–1387.
37. Sajjadi, B., Aziz, A. A., & Ibrahim, S. (2014). Investigation, modelling and reviewing the effec-
tive parameters in microwave-assisted transesterification. Renewable and Sustainable Energy
Reviews, 37, 762–777.
38. Tan, S. X., Lim, S., Ong, H. C., & Pang, Y. L. (2019). State of the art review on development
of ultrasound-assisted catalytic transesterification process for biodiesel production. Fuel, 235,
886–907.
39. Usmani, Z., Sharma, M., Tripathi, M., Nizami, A. S., Gong, L., Nguyen, Q. D., Reddy, M.
S., Thakur, V. K., & Gupta, V. K. (2022). Converting biowaste streams into energy–leveraging
microwave assisted valorization technologies for enhanced conversion. Journal of the Energy
Institute, 107, 101161.
40. Gao, Y., Remón, J., & Matharu, A. S. (2021). Microwave-assisted hydrothermal treatments for
biomass valorisation: A critical review. Green Chemistry, 23(10), 3502–3525.
41. Refaat, A. A. (2010). Different techniques for the production of biodiesel from waste vegetable
oil. International Journal of Environmental Science and Technology, 7, 183–213.
42. Hassan, A. A., & Smith, J. D. (2020). Investigation of microwave-assisted transesterification
reactor of waste cooking oil. Renewable Energy, 162, 1735–1746.
43. Mazo, P. C., & Rios, L. A. (2010). Esterification and transesterification assisted by microwaves
of crude palm oil: Homogeneous catalysis. Latin American Applied Research, 40(4), 337–342.
44. Saifuddin, N., Refal, H., & Kumaran, P. (2014). Research article rapid purification of glyc-
erol by-product from biodiesel production through combined process of microwave assisted
acidification and adsorption via chitosan immobilized with yeast. Research Journal of Applied
Sciences, Engineering and Technology, 7(3), 593–602.
45. Asakuma, Y., Ogawa, Y., Maeda, K., Fukui, K., & Kuramochi, H. (2011). Effects of
microwave irradiation on triglyceride transesterification: Experimental and theoretical studies.
Biochemical Engineering Journal, 58, 20–24.
46. Gupta, A. R., & Rathod, V. K. (2018). Calcium diglyceroxide catalyzed biodiesel produc-
tion from waste cooking oil in the presence of microwave: Optimization and kinetic studies.
Renewable Energy, 121, 757–767.
47. Ali, M. A. M., Gimbun, J., Lau, K. L., Cheng, C. K., Vo, D. V. N., Lam, S. S., & Yunus, R.
M. (2020). Biodiesel synthesized from waste cooking oil in a continuous microwave assisted
reactor reduced PM and NOx emissions. Environmental Research, 185, 109452.
48. Nazir, M. H., Ayoub, M., Zahid, I., Shamsuddin, R. B., Yusup, S., Ameen, M., & Qadeer,
M. U. (2021). Development of lignin based heterogeneous solid acid catalyst derived from
5 Transesterification of Waste Cooking Oil Through Microwave … 139

sugarcane bagasse for microwave assisted-transesterification of waste cooking oil. Biomass


and Bioenergy, 146, 105978.
49. Singh, V., & Sharma, Y. C. (2017). Low cost guinea fowl bone derived recyclable heterogeneous
catalyst for microwave assisted transesterification of Annona squamosa L. seed oil. Energy
Conversion and Management, 138, 627–637.
50. Khan, H. M., Iqbal, T., Mujtaba, M. A., Soudagar, M. E. M., Veza, I., & Fattah, I. R. (2021).
Microwave assisted biodiesel production using heterogeneous catalysts. Energies, 14(23),
8135.
51. Leung, D. Y. C., & Guo, Y. (2006). Transesterification of neat and used frying oil: Optimization
for biodiesel production. Fuel Processing Technology, 87(10), 883–890.
52. Babel, S., Arayawate, S., Faedsura, E., & Sudrajat, H. (2018). Microwave-assisted transes-
terification of waste cooking oil for biodiesel production. In: Utilization and management of
bioresources: Proceedings of 6th IconSWM 2016 (pp. 165–174). Springer Singapore.
53. Hsiao, M. C., Liao, P. H., Lan, N. V., & Hou, S. S. (2021). Enhancement of biodiesel production
from high-acid-value waste cooking oil via a microwave reactor using a homogeneous alkaline
catalyst. Energies, 14(2), 437.
54. Soosai, M. R., Moorthy, I. M. G., Varalakshmi, P., & Yonas, C. J. (2022). Integrated global opti-
mization and process modelling for biodiesel production from non-edible silk-cotton seed oil
by microwave-assisted transesterification with heterogeneous calcium oxide catalyst. Journal
of Cleaner Production, 367, 132946.
55. Yadav, G., Yadav, N., & Ahmaruzzaman, M. (2023). Microwave-assisted sustainable synthesis
of biodiesel on Oryza sativa catalyst derived from agricultural waste by esterification reaction.
Chemical Engineering and Processing: Process Intensification, 187, 109327.
56. Yadav, N., Yadav, G., & Ahmaruzzaman, M. (2023). Microwave-assisted biodiesel production
using-SO3 H functionalized heterogeneous catalyst derived from a lignin-rich biomass. Science
and Reports, 13(1), 9074.
57. Rocha, P. D., Oliveira, L. S., & Franca, A. S. (2019). Sulfonated activated carbon from
corn cobs as heterogeneous catalysts for biodiesel production using microwave-assisted
transesterification. Renewable Energy, 143, 1710–1716.
58. Al-Saadi, A., Mathan, B., & He, Y. (2020). Esterification and transesterification over SrO–
ZnO/Al2 O3 as a novel bifunctional catalyst for biodiesel production. Renewable Energy, 158,
388–399.
59. Khemthong, P., Luadthong, C., Nualpaeng, W., Changsuwan, P., Tongprem, P., Viriya-Empikul,
N., & Faungnawakij, K. (2012). Industrial eggshell wastes as the heterogeneous catalysts for
microwave-assisted biodiesel production. Catalysis Today, 190(1), 112–116.
60. Buasri, A., Inkaew, T., Kodephun, L., Yenying, W., & Loryuenyong, V. (2015). Natural hydrox-
yapatite (NHAp) derived from pork bone as a renewable catalyst for biodiesel production
via microwave irradiation. In Key engineering materials (Vol. 659, pp. 216-220). Trans Tech
Publications Ltd.
61. Panadare, D. C., & Rathod, V. K. (2016). Microwave assisted enzymatic synthesis of biodiesel
with waste cooking oil and dimethyl carbonate. Journal of Molecular Catalysis B: Enzymatic,
133, S518–S524.
62. Bunushree, B., Ravichandra, B., Rangabhashiyam, S., Jayabalan, R., & Balasubramanian, P.
(2019). Qualitative analysis of biodiesel produced by alkali catalyzed transesterification of
waste cooking oil using different alcohols. Indian Journal of Chemical Technology, 25(4),
330–336.
63. Predojević, Z. J. (2008). The production of biodiesel from waste frying oils: A comparison of
different purification steps. Fuel, 87(17–18), 3522–3528.
64. Sharma, A., Kodgire, P., & Kachhwaha, S. S. (2019). Biodiesel production from waste
cotton-seed cooking oil using microwave-assisted transesterification: Optimization and kinetic
modeling. Renewable and Sustainable Energy Reviews, 116, 109394.
65. Prafulla, D. P., Veera Gnaneswar, G., Harvind, K. R., Tapaswy, M., & Shuguang, D. (2012).
Biodiesel production from waste cooking oil using sulfuric acid and microwave irradiation
processes. Journal of Environmental Protection, 3(1), 1–7.
140 B. Behera et al.

66. Zare, M., Ghobadian, B., Fayyazi, E., Najafi, G., & Hosseinzadeh, B. (2013). Microwave-
assisted biodiesel fuel production from waste cooking oil. International Journal of Agriculture
and Crop Sciences, 5(12), 1314.
67. Yaakob, Z., Ong, B. H., Satheesh Kumar, M. N., & Kamarudin, S. K. (2009). Microwave-
assisted transesterification of jatropha and waste frying palm oil. International Journal of
Sustainable Energy, 28(4), 195–201.
68. Ali, M. M., Yunus, R. M., Cheng, C. K., & Gimbun, J. (2015). Successive optimisation of waste
cooking oil transesterification in a continuous microwave assisted reactor. RSC Advances, 5(94),
76743–76751.
69. Marchetti, J. M., Miguel, V. U., & Errazu, A. F. (2008). Techno-economic study of different
alternatives for biodiesel production. Fuel Processing Technology, 89(8), 740–748.
70. Hussain, M. N., Al Samad, T., & Janajreh, I. (2016). Economic feasibility of biodiesel
production from waste cooking oil in the UAE. Sustainable Cities and Society, 26, 217–226.
71. Mohammadshirazi, A., Akram, A., Rafiee, S., & Kalhor, E. B. (2014). Energy and cost analyses
of biodiesel production from waste cooking oil. Renewable and Sustainable Energy Reviews,
33, 44–49.
72. Farid, M. A. A., Roslan, A. M., Hassan, M. A., Hasan, M. Y., Othman, M. R., & Shirai, Y. (2020).
Net energy and techno-economic assessment of biodiesel production from waste cooking
oil using a semi-industrial plant: A Malaysia perspective. Sustainable Energy Technology
Assessment, 39, 100700.
73. Liu, Y., Yang, X., Adamu, A., & Zhu, Z. (2021). Economic evaluation and production process
simulation of biodiesel production from waste cooking oil. Current Research in Green and
Sustainable Chemistry, 4, 100091.
74. Refaat, A. A., El Sheltawy, S. T., & Sadek, K. U. (2008). Optimum reaction time, performance
and exhaust emissions of biodiesel produced by microwave irradiation. International Journal
of Environmental Science and Technology, 5, 315–322.
75. Muley, P. D., & Boldor, D. (2013). Scale-up of a continuous microwave-assisted transester-
ification process of soybean oil for biodiesel production. Transactions of the ASABE, 56(5),
1847–1854.
76. Chuck, C. J., Lou-Hing, D., Dean, R., Sargeant, L. A., Scott, R. J., & Jenkins, R. W. (2014).
Simultaneous microwave extraction and synthesis of fatty acid methyl ester from the oleaginous
yeast Rhodotorula glutinis. Energy, 69, 446–454.
77. Waudby, H., & Zein, S. H. (2021). A circular economy approach for industrial scale biodiesel
production from palm oil mill effluent using microwave heating: Design, simulation, techno-
economic analysis and location comparison. Process Safety and Environmental Protection,
148, 1006–1018.
78. Manikandan, G., Kanna, P. R., Taler, D., & Sobota, T. (2023). Review of waste cooking oil
(WCO) as a feedstock for biofuel—Indian perspective. Energies, 16(4), 1739.
79. Tibesigwa, T., Olupot, P. W., & Kirabira, J. B. (2022). The critical techno-economic aspects
for production of B10 biodiesel from second generation feedstocks: A review. International
Journal of Sustainable Energy, 41(7), 751–771.
80. Chung, Z. L., Tan, Y. H., San Chan, Y., Kansedo, J., Mubarak, N. M., Ghasemi, M., & Abdullah,
M. O. (2019). Life cycle assessment of waste cooking oil for biodiesel production using waste
chicken eggshell derived CaO as catalyst via transesterification. Biocatalysis and Agricultural
Biotechnology, 21, 101317.
81. Foteinis, S., Chatzisymeon, E., Litinas, A., & Tsoutsos, T. (2020). Used-cooking-oil biodiesel:
Life cycle assessment and comparison with first-and third-generation biofuel. Renewable
Energy, 153, 588–600.
82. Iglesias, L., Laca, A., Herrero, M., & Díaz, M. (2012). A life cycle assessment comparison
between centralized and decentralized biodiesel production from raw sunflower oil and waste
cooking oils. Journal of Cleaner Production, 37, 162–171.
83. Ripa, M., Buonaurio, C., Mellino, S., Fiorentino, G., & Ulgiati, S. (2014). Recycling waste
cooking oil into biodiesel: A life cycle assessment. International Journal of Performability
Engineering, 10(4), 347.
5 Transesterification of Waste Cooking Oil Through Microwave … 141

84. Alishahi, A., Golmakani, M. T., & Niakousari, M. (2021). Feasibility study of microwave-
assisted biodiesel production from vegetable oil refinery waste. European Journal of Lipid
Science and Technology, 123(9), 2000377.
85. Yari, N., Mostafaei, M., Naderloo, L., & Safieddin Ardebili, S. M. (2022). Energy indicators for
microwave-assisted biodiesel production from waste fish oil. Energy Sources, Part A: Recovery,
Utilization, and Environmental Effects, 44(1), 2208–2219.
86. Viornery-Portillo, E. A., Bravo-Díaz, B., & Mena-Cervantes, V. Y. (2020). Life cycle assess-
ment and emission analysis of waste cooking oil biodiesel blend and fossil diesel used in a
power generator. Fuel, 281, 118739.
87. Hosseinzadeh-Bandbafha, H., Nizami, A. S., Kalogirou, S. A., Gupta, V. K., Park, Y. K., Fallahi,
A., Sulaiman, A., Ranjbari, M., Rahnama, H., Aghbashlo, M., & Peng, W. (2022). Environ-
mental life cycle assessment of biodiesel production from waste cooking oil: A systematic
review. Renewable and Sustainable Energy Reviews, 161, 112411.
88. Zhu, H., He, J., Hong, T., Yang, Q., Wu, Y., Yang, Y., & Huang, K. (2018). A rotary radiation
structure for microwave heating uniformity improvement. Applied Thermal Engineering, 141,
648–658.
89. Kristiana, T., Baldino, C., & Searle, S. (2022). An estimate of current collection and potential
collection of used cooking oil from major Asian exporting countries. Work. Pap, 21.
90. Liu, T., Liu, Y., Wu, S., Xue, J., Wu, Y., Li, Y., & Kang, X. (2018). Restaurants’ behaviour,
awareness, and willingness to submit waste cooking oil for biofuel production in Beijing.
Journal of Cleaner Production, 204, 636–642.
91. Milano, J., Ong, H. C., Masjuki, H. H., Silitonga, A. S., Chen, W. H., Kusumo, F., Dharma,
S., & Sebayang, A. H. (2018). Optimization of biodiesel production by microwave irradiation-
assisted transesterification for waste cooking oil-Calophyllum inophyllum oil via response
surface methodology. Energy Conversion and Management, 158, 400–415.
92. Selvaraj, R., Moorthy, I. G., Kumar, R. V., & Sivasubramanian, V. (2019). Microwave mediated
production of FAME from waste cooking oil: Modelling and optimization of process parameters
by RSM and ANN approach. Fuel, 237, 40–49.
93. Rao, G. L. N., Ramadhas, A. S., Nallusamy, N., & Sakthivel, P. (2010). Relationships among
the physical properties of biodiesel and engine fuel system design requirement. International
Journal of Energy and Environmental, 1(5), 919–926.
94. Mazubert, A., Poux, M., & Aubin, J. (2013). Intensified processes for FAME production from
waste cooking oil: A technological review. Chemical Engineering Journal, 233, 201–223.
Chapter 6
Green Catalysts Synthesized
from Biomass for Biodiesel Production

Amirthavalli Velmurugan, Anita R. Warrier, and Gurunathan Baskar

Abstract The impact of any processing technology on sustainable development


requires a zero-waste approach of reusing, recycling, and recovering. Biodiesel
serves as an alternative source of renewable energy to fossil fuels, and it can be
produced by transesterification of feedstock with alcohol in the presence of a cata-
lyst. Nanocatalysts are preferred in the transesterification reaction when compared
to acid, base, and enzymatic catalysts due to their high surface-to-volume ratio, high
catalytic activity, easy recoverability, simple synthesis procedure, and low toxicity.
Calcium and carbon-based green catalysts that offer a large surface area, improved
porosity, stability, and inertness can be derived from plant, agricultural, and animal
wastes. This chapter discusses the role of various green catalysts that are derived
from biomass, along with their merits and demerits for large-scale production of
biodiesel, and the possible ways of improving the yield to make biodiesel viable for
commercial applications.

Keywords Biomass · Transesterification · Biodiesel · Green catalyst

A. Velmurugan
Department of Petroleum Engineering, Academy of Maritime Education and Training, Chennai,
Tamil Nadu, India
e-mail: amirthavalli.v@ametuniv.ac.in
A. R. Warrier (B)
Nanophotonics Research Laboratory, Department of Physics, Academy of Maritime Education
and Training, Chennai, Tamil Nadu, India
e-mail: anitawarrier2@gmail.com
G. Baskar
Department of Biotechnology, St. Joseph’s College of Engineering, Chennai, Tamil Nadu, India

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2024 143
G. Baskar et al. (eds.), Circular Bioeconomy Perspectives in Sustainable Bioenergy
Production, Energy, Environment, and Sustainability,
https://doi.org/10.1007/978-981-97-2523-6_6
144 A. Velmurugan et al.

6.1 Introduction to Green Catalyst

Green catalysts are those that are synthesized from the pyrolysis of biomass which
can be from agricultural waste, plant waste, and animal waste as shown in Fig. 6.1,
and green catalysts are considered to be a good source of calcium and carbon. The
advantages of green catalysts are non-toxicity, reusability, simple catalyst synthesis
procedure, thermal stability, enhanced biodiesel yield when used in transesterifica-
tion reaction, and it does not create any harm to society as well as to the environment
[1, 2]. Moreover, there are no disposal issues since the catalyst itself is biodegradable
and excludes the production of wastewater. Catalyst derived from biomass possesses
an increased surface-to-volume ratio and finds use as a carbonaceous catalyst or as
support in biodiesel production. Agricultural waste biomass especially lignocellu-
losic, i.e., rice husk, palm kernel shell, empty fruit bunch, and walnut shell, is gaining
importance as an inexpensive and renewable feedstock for the synthesis of green cata-
lysts [3]. Moreover, green catalysts obtained from these biomasses possess increased
porosity and large surface area which is ideal for adsorption–desorption mechanism
[4]. It has been reported that bio-based catalyst derived from agricultural waste and
food waste shows high conversion efficiency in transesterification [5]. The presence
of reduced ash content in activated carbon promotes transesterification, and also, it
finds application in high pressure and temperature. Moreover, the surface properties
do not alter at increased temperature or pressure [6].
Nanoparticles synthesized from plant sources are inexpensive, eco-friendly, and
efficient and can be used for large-scale production without the requirement of any
sophisticated equipment and complex procedures when compared to physical or
chemical methods. The synthesis procedure can be carried out in a shorter duration

Fig. 6.1 Shows the various feedstocks for obtaining green catalysts from biomass
6 Green Catalysts Synthesized from Biomass for Biodiesel Production 145

of time and requires less energy [7]. The production cost of nanoparticles synthesized
using local origin plant sources is comparatively less [8].
Plants consist of alkaloids, flavonoids, terpenoids, and steroids which serve as
reducing agents in nanoparticle synthesis. It has been reported that calcium oxide
nanoparticles were synthesized from the phytochemical component of watermelon
extract by reducing and stabilizing the precursor (calcium nitrate) [9]. Synthesized
calcium oxide nanoparticles when used in the transesterification of soybean oil give
a biodiesel yield of 96.2% at 20:1 methanol-to-oil molar ratio, 8 wt% of the amount
of catalyst, and 65 °C of the reaction temperature in 2.5 h. Cholapandian et al. have
reported on the synthesis of calcium oxide nanocatalyst from Acalypha indica. A
biodiesel yield of 94.74% was obtained when the synthesized calcium oxide nanocat-
alyst was used in the transesterification of waste cooking oil at 11.8:1 methanol-to-oil
molar ratio, 2.4 wt% of the amount of catalyst, and 63.7 °C of reaction temperature
in 70 min [10]. Sumera et al. have reported on the green synthesis of cadmium oxide
nanoparticles from the leaf extract of Buxus papillosa using in situ wet impregnation
method. It has been reported that cadmium-based nanoparticle finds application in
polymerization and transesterification [11]. When the synthesized cadmium oxide
is used in the transesterification of Diospyros malabarica (Malabar Ebony) gives a
biodiesel yield of 94 wt% at 1:9 oil-to-methanol molar ratio, 0.5 wt% of the amount
of catalyst, and 90 °C of reaction temperature in 180 min [12].
Animal wastes in the form of chicken eggshell waste, waste fish scale, waste
cockle shell, waste mud scrub shell, shells of mollusk, golden apple snail shells,
ostrich eggshells, bones, oyster shells, and river snail shell ash can also serve as
suitable feedstock in the synthesis of CaO nanocatalyst for biodiesel production [13,
14]. Animal waste is biodegradable, recyclable, and non-toxic to the environment and
humans with improved catalytic efficiency [15, 16]. Animal waste excludes waste
management costs but helps in obtaining inexpensive nanocatalysts that can be used
in biodiesel production [17, 18].

6.2 Agricultural Waste

6.2.1 Rice Husk

Rice husk is an agricultural waste obtained from the rice milling process. During
the milling process, 78% of the weight is obtained as rice, broken rice, and bran,
and the remaining 22% as rice husk. Rice husk contains 15 wt% of silica, cellulose
28.7–35.6%, hemicellulose 12.0–29.3%, and lignin 15.4–20.0% [19], and it is not
recommended as fodder for animals due to its low cellulose and sugar content. Rice
husk is the hard outer covering that is developed due to the absorption of silica from
the soil to shield the rice grains within the structure. It has been reported that one
ton of rice gives 200 kg of rice husks which when combusted gives 40 kg of rice
husk ash [20]. The disposal of large quantities of rice husk in developing countries is
146 A. Velmurugan et al.

very tedious and expensive, and the majority of the rice husk is unused and heaped as
waste material or burned in the agricultural fields which further creates environmental
issues. Due to the presence of carbon in the rice husk, it serves as an important raw
material in the production of activated carbon of amorphous nature with increased
surface area after the removal of silica [21, 22]. Silica present in the rice husk is
removed in the form of sodium silicate solution after it is made to react with dilute
sodium hydroxide solution at increased temperature for a particular duration [23].
The pyrolysis of the rice husk generates bio-oil, syngas (carbon monoxide and
hydrogen), and also by-product biochar (rice husk carbon and rice husk ash) [24].
Bio-oil and syngas are used in biofuel production with a heating value of (20 MJ/kg).
Synthesis of activated carbon is carried out either directly by calcination or indirectly
by burning the rice husk into rice husk carbon and then calcinating at a known
temperature, which further undergoes chemical or physical activation followed by
drying. Rice husk ash contains amorphous silica which can be extracted by alkali
treatment [25, 26]. Rice husk ash serves as a raw material for producing silicates
and silica due to its increased silica content. Othman Ali et al. synthesized ZSM-5
zeolite using rice husk ash [27] and Jan-Jezreel F. Saceda et al. synthesized zeolite
NaY using rice husk ash. The extracted silica from rice husk ash can be used as
SBA-15 materials which are used in the production of ordered mesoporous carbon
like CMK-3 (carbon manufactured from korea-3) [28]. The ordered mesoporous
carbon finds use as catalyst supports [29], adsorbents [30], electrode materials [31],
hydrogen storage [32], sensors [33], and supercapacitors [34].
The porosity of activated carbon obtained from the carbonization step can be
further improved by using activation steps. The activation methods enhance the pore
volume and expand the diameter of the pores and surface area of activated carbon.
In the activation method, improvement in the microporous structure can be attained
by stripping out the disorganized carbon. The pores swell due to the heating of walls
among the pores. Hence, the provisional pores and microporosity are enhanced after
the activation methods. The activation step plays an important role in the preparation
of activated carbon with good porosity [35]. The pores obtained on the activated
carbon surface are distinguished into macropores greater than 25 nm, mesopores
between 1 and 25 nm, and micropores less than 1 nm [36].
The calcination of rice husk carbon needs to be carried out in a muffle furnace with
inert media (nitrogen) at a high temperature of 600–900 °C [37]. It is important to
carry out calcination in an inert atmosphere to decrease the production of ash content
and flue gas, especially carbon dioxide. The most commonly used physical activation
methods are steam and gas activation. Steam activation is normally carried out after
the thermal carbonization of biomass. Porosity developed in the biochar after thermal
carbonization further undergoes steam activation which will enhance the porosity.
The chemical reactions that occur during steam activation are [38]:

C + H2 O → CO + H2 (6.1)

CO + H2 O → CO2 + H2 (6.2)
6 Green Catalysts Synthesized from Biomass for Biodiesel Production 147

The reaction between water and carbon results in the removal of volatile matter,
the growth of new micropores as well as expansion of existing pores. Gas activa-
tion is also carried out to enhance the surface area and pore volume by the reaction
between the gas (CO2 , N2 , NH3 , air, and O2 ) and the surface of the biochar and
also yields microporous- and mesoporous-activated carbon structure. In the chem-
ical activation methods, acids such as sulfuric acid (H2 SO4 ), hydrochloric acid (HCl),
phosphoric acid (H3 PO4 ), and nitric acid (HNO3 ) or bases such as sodium hydroxide
(NaOH), potassium hydroxide (KOH), potassium carbonate (K2 CO3 ), and calcium
oxide (CaO) or salts like zinc chloride (ZnCl2 ) or oxidants like hydrogen peroxide
(H2 O2 ) and potassium permanganate (KMnO4 ) are used. The chemical activation
method gives activated carbon in reduced time with improved surface area and
mesoporous structure when compared to the physical activation.
Carbon present in the rice husk combines with oxygen molecules at high tempera-
tures to form carbon dioxide which decreases the adsorbent property and also actives
sites in the obtained biochar. Hence, the synthesis of activated carbon should be
carried out under an inert atmosphere [39] which further increases the production
cost and also needs a complicated experimental setup. Hence, a proportional integral
derivative control muffle furnace is recommended.
Rice husk char consists of 50% of ash which suppresses the pore development,
and hence, it has a low surface area. The ash present in the rice husk char is removed
by reacting it with chemicals such as zinc chloride and alkali solution (potassium
hydroxide or sodium hydroxide). The surface area of raw rice husk char varies from
25 to 28 m2 g−1 , whereas the surface area of rice husk char after removing ash by
treating it with an alkali solution (sodium hydroxide) is reported to be 162, 174, and
331 m2 g−1 . It has been reported that rice husk char on further treatment with zinc
chloride leads to a more porous-activated carbon with an enhanced surface area of
365–645 m2 g−1 [40].
Zhaoe et al. have reported that a biodiesel yield of 93.4% is obtained using highly
porous rice husk biochar loaded with 30% of CaO as a catalyst with a surface area of
28.3 m2 g−1 under optimum conditions of 9:1 methanol-to-oil ratio, 8 wt% of catalyst
in 180 min [41]. It has also been reported that rice husk biochar acts as a good carrier
and the presence of Ca–O–Si bond on the catalyst surface aids in improving the
catalytic activity and stability of the catalyst. Balkis Hazmi et al. have reported a
biodiesel yield of 98.2% using a bifunctional magnetic nanocatalyst derived from
rice husk char doped with potassium oxide and nickel oxide at 4 wt% of catalyst,
12:1 methanol-to-oil molar ratio, 65 °C of reaction temperature within a reaction
duration of 2 h [42]. Chen et al. have reported that Li-modified rice husk ash is
synthesized using rice husk ash and Li2 CO3 . The prepared catalyst with basic strength
(OH-) above 15 shows good catalytic activity and when used in transesterification of
soybean oil shows 99.5% of biodiesel under optimized conditions of 24:1 methanol-
to-oil molar ratio, 4% of the amount of catalyst, and 65 °C for a reaction duration of
3 h [20].
The usage of inexpensive rice husk-derived biomass as the catalyst raw material
can solve leftover problems of rice husk in the open atmosphere. Activated carbon
obtained from rice husks can serve as a catalyst for large-scale biodiesel production
148 A. Velmurugan et al.

[43], photocatalytic degradation of dye [44], wastewater treatment, and enhancing


power generation from fuel cells.

6.2.2 Palm Kernel Shell

It was reported by the Malaysian Palm Oil Board (MPOB) that one ton of palm fresh
fruit bunch in the milling process yields 5–7% palm kern shell (PKS), 21–22% empty
fruit bunch (EFB), and 12–16% mesocarp fiber (MF). Palm kernel shell consists of
48 wt% of lignin, 30 wt% of cellulose, and 22 wt% of hemicellulose. It mostly has 51
wt% of carbon, 39 wt% of oxygen, and minimum content of hydrogen and nitrogen
approximately 3 and 7 wt% [45].
It has been reported that biochar derived from biomass can act as the catalyst for
combining suitable functional active groups. Moreover, converting palm waste into
a heterogeneous catalyst can decrease the production cost of biodiesel as well as
decrease the health issues that can be caused due to dumped wastes. Waste generated
from palm oil industries can give a high-value activated carbon that acts as the
catalyst support for biodiesel production. Activated carbon obtained from waste
palm oil, mainly palm kern shell, is used due to its high calorific value, reduced
sulfur content and ash content, no variation in the species, enhanced shelf life, and
availability of palm waste throughout the year. Activated carbon obtained from palm
kernel shell is a highly porous carbon-rich content with an amorphous carbon form.
Activated carbon obtained from palm kernel shells is inexpensive when compared to
other agricultural byproducts. Activated carbon can be used as a catalyst provided it
should have a minimum amount of ash content and minerals.
It has been reported that palm kernel shell possesses 50 wt% of lignin which is
considered to be the maximum amount when compared with different palm waste
materials such as empty fruit bunch, mesocarp fiber, oil palm trunk, and frond (21–35
wt% of lignin) [46–49]. Enhanced yield of activated carbon can be obtained from
palm kernel shell due to the presence of complicated polymeric lignin structure
[50]. Lignin structure has benzene rings within it which provides increased chemical
stability to the decomposition reaction in comparison with the polysaccharides that
are available in cellulose and hemicellulose. Hence, lignin can be converted into
activated carbon in stable form with greater aromaticity in the carbonization method
[51], instead of being converted into bio-oil or gases like methane, carbon dioxide, and
carbon monoxide by decomposition reaction [52]. Palm kernel shell shows maximum
content of carbon which makes it a potential resource for obtaining activated carbon
with the help of pyrolysis. It also possesses increased wt% of oxygen which needs
to be removed to obtain a carbon-rich product.
Abdulla et al. have reported on the synthesis of bifunctional nanocatalyst
attained from waste palm kernel shell and synthesized catalyst (palm kernel
shell-activated carbon-K2 CO3(30%) CuO(5%) , PKSAC-KOH(30%) CuO(5%) , PKSAC-
K3 PO4(30%) CuO(5%) , and PKSAC-NaOH(30%) CuO(5%) ) along with the impregnation
of K2 CO3, KOH, K3 PO4 , Cu(NO3 )2 , and NaOH. PKSAC-K2 CO3(30%) CuO(5%) has
6 Green Catalysts Synthesized from Biomass for Biodiesel Production 149

high surface area of 438.08 m2 g−1 , pore volume of 0.367 cm3 g−1 , and pore width
of 3.8 nm. It has been reported that 95% of biodiesel yield is obtained from waste
cooking oil under optimal conditions of 5 wt% catalyst loading, 12:1 methanol-to-oil
molar ratio at 80 °C for 4 h using the mixture of potassium and copper on activated
carbon surface (PKSAC-K2 CO3(30%) CuO(5%) ) [53]. An enhanced amount of basicity
in the range of 8.866 mmolg−1 and acidity (27.016 mmol−1 ) has favored the simul-
taneous transesterification and esterification reaction. Celine et al. have reported
on the synthesis of magnetic palm kernel shell-derived catalysts (Fe-KOH-PKS)
using impregnation of potassium hydroxide followed by pyrolysis and magnetic
palm kernel shell catalyst (Fe-PKS) [54]. The synthesized catalyst shows increased
catalytic properties due to the increased magnetic saturation of 22.155 emu/g, a
basicity of 6.68 mmol/g when compared to the Fe-PKS catalyst. Fe-KOH-PKS when
used in transesterification of palm oil gives a maximum biodiesel yield of 99.43%
under optimal operating conditions of 7.5 wt% catalyst loading and 9:1 methanol-
to-oil ratio at 65 °C after 2 h. The synthesized catalyst derived from biomass helps to
obtain a huge amount of inexpensive green catalyst for large-scale biodiesel produc-
tion. Kostic et al. obtained a biodiesel yield of 99% from sunflower oil using palm
kernel shell biochar under the optimal reaction conditions of 9:1 methanol-to-oil
molar ratio, 3 wt% of catalyst, and 65 °C in 4 h [55].
The disposal of these biomass residue has become a challenging task. The palm
biomass waste is widely used in fertilizer and biogas production and also has a great
scope to get converted into a biodegradable heterogeneous catalyst for biodiesel
production and wastewater treatment due to its huge availability.

6.2.3 Sugarcane Bagasse

Sugarcane belongs to a grass family, and it is cultivated annually for its juice which
in turn is used for the production of sugar. Sugarcane bagasse the leftover part after
obtaining the juice is considered as a waste, and it consists of cellulose, hemicellulose,
and lignin. It has been reported that one metric ton of sugarcane gives 280 kg of
bagasse which finds application in the bioenergy industry after burning to supply
energy and for the production of bioethanol [56]. Sugarcane bagasse consists of
40–45% of cellulose, 30–35% of hemicellulose, and 20–30% of lignin and ash.
The cellulose consists of long chains of D-glucose units joined by glycosidic bonds,
hemicellulose consists of xylose, mannose, and glucose saccharide units, and lignin is
obtained by polymerization of aromatic alcohols [57, 58]. Mostly sugarcane bagasse
consists of 60–80% of carbohydrates which predominantly consists of carbon and
hydrogen that can be formed into activated carbon [59].
Chausali et al. have reported that the advantages of nanotechnology and biochar
are displayed together when bulk biochar is decreased to the nano biochar level [60].
Research is being widely carried out for the synthesis of carbon nanotubes, and acti-
vated carbon monoliths derived from crop residue biochar and synthesized catalysts
possess excellent properties such as good porosity, adsorption capacity, and low cost.
150 A. Velmurugan et al.

Akinfalabi et al. have synthesized bio-based sulfonated sugarcane bagasse and when
used in transesterification of palm fatty acid distillate gives a biodiesel yield of 98.6%
under optimal conditions of 10:1 methanol-to-oil molar ratio, 2 wt% of catalyst, and
60 °C reaction temperature for a duration of 1.5 h [61]. Subramaniapillai et al.
synthesized biochar from sugarcane bagasse using an activation method followed by
carbonization. Synthesized heterogeneous catalyst gives a biodiesel yield of 98.94%
when used in the transesterification of waste cooking oil at a reaction temperature
of 65 °C, 1:2 methanol-to-oil molar ratio, and 10 wt% of catalyst in 120 min [62].
Abdul Mutalib has synthesized SiO2 -rich sugarcane bagasse ash/CaO inexpensive
heterogeneous catalyst and when used in the transesterification of palm olein oil
gives a yield of 93.8% under optimum conditions of 20:1 methanol-to-oil molar
ratio, 3 h reaction time, 6 wt% catalyst amount, and 65 °C reaction temperature [63].
It has also been reported that the presence of SiO2 in the sugarcane bagasse ash
plays an important role in improving the catalytic performance and decreasing the
deactivation of the catalyst by reducing the leaching of active sites of the catalyst.
Activated carbon finds application as an absorbent material for various industrial
applications such as the removal of organic and inorganic impurities from wastewater,
and biodiesel production due to its increased thermal stability, porosity, and simple
synthesis procedure with low cost [64–67].

6.2.4 Coconut Shell

Coconut shell consists of 14 wt% of cellulose, 32 wt% of hemicellulose, and 46 wt%


of lignin [68]. It has been reported that coconut shell consists of reduced ash content,
more carbon content, and possesses increased strength and hardness which makes it
a suitable candidate for the development of catalyst.
Activated carbon can be produced from different raw carbon resources like lignite,
peat, coal, and biomass resources such as wood, sawdust, bagasse, and coconut shells
[69]. However, the abundant supply of coconut shells as a waste product from the
coconut oil and desiccated coconut industry makes the production of activated carbon
from this material more financially viable since using grain or coal as raw materials for
activated carbon will require manufacturers extra amount of money for procurement.
Furthermore, besides being an amorphous form of carbon that can absorb many gases,
vapor, and colloidal solids, coconut shell-activated carbons are advantageous over
carbons made from other materials because of their high density, high purity, and
virtually dust-free nature. These carbons are harder and more resistant to attrition.
Shobhana et al. have reported on the synthesis of Ce-supported sulfated acti-
vated carbon obtained from coconut shell. It has been reported that free fatty acid
conversion of 93% is obtained by transesterification of chicken fat and skin oil using
5Ce/ACcs-S at 90 °C of reaction temperature, 12:1 methanol-to-oil molar ratio with
3 wt% of catalyst loading in 1 h [70]. Enhanced physicochemical properties and
stability of the catalyst are attained due to the rare-earth metal oxides such as CeO2
and La2 O3 . Sulfated activated carbon possesses an increased volume of micropores
6 Green Catalysts Synthesized from Biomass for Biodiesel Production 151

and wide pore size distribution which supports adsorption when used in the produc-
tion of biodiesel. Moreover, oxygen groups that are present on the surface serve
as an attaching site for metallic precursors which further helps in the dispersion of
dopants that are added during the reaction. Increased catalytic activity is due to the
enhanced pore diameters in the catalyst which enables diffusion of reactants such
as free fatty acid and triglycerides. Activated carbon obtained from coconut shells
possesses uniform pore structure, increased density, and purity due to its hardness
and increased resistance to attrition. The maximum number of pores falls in the cate-
gory of micropores which are efficient in the removal of volatile organic compounds
and residue of pesticides [71].

6.3 Plant Waste

6.3.1 Potato Waste

It is considered to be one of the crops which are grown worldwide in the majority
and consumed by people in different forms such as potato chips, crisp, puree, and
French fries. The usage of potatoes in different forms increases the peel waste which
ranges from 15 to 40% of its original weight. China is the main producer of potatoes
with 91.8 million tonnes annually followed by India with an annual production of
50.2 million tonnes. If the potato peels are not used properly and dumped as waste,
it contributes to environmental pollution. Potato peels serve as suitable feedstock for
the production of bio-oil and biochar through the pyrolysis process. Biochar obtained
as the product of the pyrolysis process contains a biogenic potassium source which
is used for the synthesis of heterogeneous catalysts. Biodiesel conversion of 97.5%
was obtained from soybean waste cooking oil using biochar from potato peels under
optimum conditions of 9:1 methanol-to-oil molar ratio, 3 wt% of catalyst, and 60 °C
in 2 h [72].

6.3.2 Banana Waste

The banana is considered to be the earliest fruit that is familiar to humankind. Banana
is the second greatest tropical fruit cultivated in the world, and huge amount of banana
peel is obtained as waste by-product. India is considered to be the biggest producer
and consumer of bananas in the world. Indian states like Maharashtra, Assam, Tamil
Nadu, Arunachal Pradesh, Karnataka, Madhya Pradesh, and Gujarat are the major
producers of 15–20 varieties of bananas. Banana peel finds application in bioethanol
production and energy-related activities as well as acts as a biosorbent and biochar.
The usage of biochar obtained from the banana peel, trunk, and rhizome as a catalyst
for the transesterification process is gaining momentum among researchers since
152 A. Velmurugan et al.

banana peel consists of carbohydrates, minerals, and nutrients, including cations of


calcium, sodium, potassium, iron, magnesium, and anions of sulfur.
Niran et al. have synthesized banana peel pyrolyzed calcined biochar catalyst from
banana peel which is rich in potassium, calcium, magnesium, sodium, and silicon.
The obtained biochar acts as support material for the catalyst by using acid (H2 SO4 ) or
base (KOH, CaOH). Biodiesel yield of around 98% is obtained by transesterification
of soybean waste cooking oil in the presence of banana peel pyrolyzed calcined
biochar catalyst under optimal conditions of 9:1 methanol-to-oil molar ratio, 1.5
wt% of catalyst, and 60 °C reaction temperature in 2 h [73]. The synthesized catalyst
displays improved catalytic activity due to the presence of more quantity of potassium
in oxide and carbonate form. Jitjamnong et al. have reported a biodiesel yield of
99.15% from the transesterification of palm oil using a heterogeneous AC/KOH
catalyst derived from the calcination of banana peel biochar-supported K2 CO3 .The
optimal conditions for transesterification were found to be a 15:1 methanol-to-oil
molar ratio, 4 wt% of catalyst, and 65 °C reaction temperature in 120 min [74].
Sujata Brahma et al. have reported on the synthesis of biochar derived from Musa
chinensis peel, trunk, and rhizome. Biochar derived from M. chinensis peel has the
highest basicity of 0.93 mmolg−1 when compared to biochar derived from trunk
and rhizome with a basicity of 0.15 and 0.2 mmolg−1 and is micro-mesoporous.
Synthesized biochar derived from M. chinensis peel when used in transesterification
of a quinary oil mixture (soybean oil, sunflower oil, canola oil, jatropha oil, and
pongamia oil) gives a biodiesel yield of 95.82% under optimal conditions of 9:1
methanol-to-oil molar ratio, 5 wt% of catalyst, and 65 °C of reaction temperature
in 11 min [75]. Laskar has reported on the synthesis of Mangifera indica peel ash
obtained by burning the mango peel in the open air. Synthesized catalyst when used
in the transesterification of soybean oil gives a biodiesel yield of 98% under optimum
conditions of 6:1 methanol-to-oil molar ratio and 6 wt% of catalyst in 4 h under room
temperature. Enhanced catalytic activity of the synthesized catalyst with increased
surface area and porosity is due to the presence of metal oxides such as K2 O, MgO,
and CaO.
Recently ash catalyst derived from plant waste has gained momentum in biodiesel
production due to its simple preparation methods, abundance, inexpensive, and does
not harm the environment. Waste tucuma peels, Musa acuminate peduncle, Musa
acuminata peel, Musa balbisiana Colla, rice husk ash, coconut waste, rubber seed
shells, cocoa shells, and Lemna perpusilla Torrey ash are reported to be efficient
feedstocks for the synthesis of catalyst used in biodiesel production.

6.3.3 Mango Waste

Mango is grown in more than 90 countries in the world, and it is obtained in 160
varieties. India is the main producer of mango in the world. Mango is considered to
be a tropical fruit that is consumed in large quantities after banana, and residue of
mango processing (peel, kernel, and seed) consists of 35–60% of total fruit weight,
6 Green Catalysts Synthesized from Biomass for Biodiesel Production 153

hence describing high volume resource of usable bio-based chemicals and materials.
The mango peel provides 7–24% of the total weight of the fruit, which is heaped
as a waste product. Laskar et al. have reported the synthesis of a catalyst derived
from waste peels of mango (M. indica) with high basicity. The synthesized catalyst
when used in transesterification of soybean oil gives a biodiesel yield of 98% at 6:1
methanol-to-oil molar ratio and 6 wt% of catalyst at 4 h when carried out under
room temperature. Shah et al. have reported on the plasma synthesis of graphene
using mango peels [76].

6.3.4 Pomelo Waste

Pomelo is a fruit widely grown in China, and huge quantities of waste are disposed
in landfills. It has been reported that pomelo peel has been used to produce activated
carbon which finds application in the removal of organic pollutants or as anode
materials for batteries. The usage of biochar as catalyst support helps in the production
of biodiesel on a large scale and also promotes the pyrolysis process to reduce the
environmental impact. Che Zhao et al. have reported on the synthesis of biochar
loaded with 25 wt% of potassium carbonate which possesses high basic sites. The
biochar was derived from derived from pomelo peel. The synthesized catalyst when
used in the transesterification of palm oil gives a biodiesel yield of 98% under optimal
conditions of 8:1 methanol-to-oil molar ratio, 6 wt% of catalyst, and 65 °C of the
reaction temperature in 2.5 h [77].

6.3.5 Mustard Plant Waste

The mustard plant is cultivated in all places of India, central Africa, and southern
Russia, and it is considered to be one of the cultivated species back to a period of 5000
years. The oil obtained from mustard seeds is mainly used in cooking. The delicate
leaves of the immature plants are consumed as vegetables and salads. After harvesting
the mustard seeds, waste Brassica nigra plant is burnt as ash, and it is considered
to be highly basic. Biswajit Nath et al. have reported that the catalyst consists of
56.13 wt% of potassium and 26.04 wt% of calcium in the form of oxide, carbonate,
and chloride. A biodiesel yield of 98.79% is obtained while converting soybean oil
using a catalyst derived from waste B. nigra plant under optimal conditions of 12:1
methanol-to-oil molar ratio, 7 wt% of catalyst, and 320 C reaction temperature in 75
min [78].
154 A. Velmurugan et al.

6.3.6 Moringa Waste

Moringa leaves consist of micro and macronutrients such as nitrogen, phosphorus,


potassium, calcium, boron, zinc, sodium, manganese, and iron, and polyphenols,
alkaloids, tannins, ascorbic acid, and carotenoid flavonoids. The composition of
minerals and proteins varies in the plant based on their source. The leaves with
alkaline materials are considered suitable feedstock for the synthesis of catalysts.
Aleman Ramirez et al. have reported on the synthesis of moringa leaves ash and
synthesized catalyst after calcination showing the presence of inorganic carbonate
minerals of dolomite, calcite, and K2 Ca(CO3 )2 ) which helps in the transesterification
reaction. 86.7% of biodiesel yield is obtained when the synthesized catalyst is used
in the transesterification of soybean oil under optimal operating conditions of 6:1
methanol/oil molar ratio, 6 wt% catalyst, and 65 °C of reaction temperature in 120
min [79].
Activated carbon can act as suitable catalyst support for CaO, KOH, or lipase-
supported catalysts with large surface area and porosity. It can go through surface
functionalization with acid or base groups. It shows thermal, acidic, or environmental
stability as well as mechanical durability. It possesses a considerable volume of
micropores with a good pore size distribution and adsorption surface area (800–
1500 m2 /g), and it is considered to be inexpensive with inert carbon skeleton. It
shows increased catalytic activity due to its surface properties and surface oxides.
Surface oxygen groups form fastening sites for metallic precursors and metals which
characterize the properties of activated carbon as a catalyst support material.
Biochar catalysts embedded with ferrite ions are known as magnetic biochar
catalysts. It is synthesized by embedding ferrite ions into biomass accompanied by
carbonization and functionalization of the magnetic support. It has been reported that
the usage of biochar as a catalyst or catalyst support due to its reduced porosity and
adsorptive properties. Hence, biochar is sulfonated with sulfuric acid to improve the
surface area and porosity [80].
Beata et al. have reported on the transesterification of corn oil using KOH
supported on activated carbon. The author has obtained a biodiesel yield of 92%
under optimal reaction conditions of 3:1 methanol-to-oil molar ratio, 0.75 wt% of
catalyst, and 62.5 °C in 1 h of reaction time [81]. Activated carbon was synthesized
from beech tree biomass and has a low surface area with 12.4% of oxygen groups.
Ayesha Hameed et al. have reported on the synthesis of calcium-loaded activated
carbon catalyst by pyrolysis of peach shell succeeded by chemical activation with
KOH followed by calcium loading with the help of wet impregnation method. A
biodiesel yield of 96% is obtained by transesterification of waste cooking oil under
optimal reaction conditions of 8:1 methanol-to-oil molar ratio, 5 wt% of catalyst,
and 65 °C temperature in 160 min of reaction time [82].
6 Green Catalysts Synthesized from Biomass for Biodiesel Production 155

6.4 Animal Waste

6.4.1 Calcium Oxide

Waste eggshells consist of a network of protein fibers along with crystals of calcium
carbonate (CaCO3 ), magnesium carbonate, calcium phosphate, organic substances,
and water. CaCO3 on calcination at high temperatures gets converted into calcium
oxide (CaO). Eggshell consists of more than more than 90% of CaCO3 which forms a
nanoporous structure and occurs in the form of calcite. CaO catalyst synthesized from
waste eggshell, shells of mullock is non-toxic, cheap, biodegradable, non-corrosive,
highly basic, readily available, and shows good recyclability and reusability. CaO
possesses a long shelf life as a catalyst, and it can be used in moderate reaction
conditions [83]. It has been reported that basic sites of pure CaO get poisoned and
converted to calcium carbonate and calcium hydroxide when it is kept in atmospheric
air.
Catalytic activity, shelf life, and stability of CaO can be increased by doping with
rare-earth elements such as lanthanum and cerium or forming mixed metal oxides.
Improved catalytic performance of CaO was reported to be good when it is mixed
with La2 O3 but when it is exposed to air, there is a change in the structure of the
catalyst. Mixed metal oxides can also be synthesized by impregnating CaO with other
metals. Eu2 O3 /c-Al2 O3 was considered to be costlier for large-scale applications. In
the lanthanide series, praseodymium oxide is considered to be a basic catalyst with
the highest oxygen ion mobility due to which various stable phases help in the quick
changes in the oxygen state of praseodymium. Hence, praseodymium oxides find
applications as catalysts or act as promoters or stabilizers in forming a catalyst. Sana
Gohar Khan et al. have reported that 87.42% of biodiesel yield is obtained when
transesterification of castor oil is carried out using Pr doped CaO derived from waste
mussel shell at 8:1 methanol-to-oil molar ratio, 2.5 wt% of catalyst, and 60–65 °C
of the reaction temperature in a duration of 4 h.
Rahman et al. reported the synthesis of mixed metal oxides (Zn/CaO, Cu/CaO)
by incipient wet impregnation method, and the surface area of mixed metal oxides
was higher when compared to pure CaO. Synthesized mixed metal oxide Zn/CaO
when used in transesterification of eucalyptus oil gives a maximum yield of 93.8%
when compared to Cu/CaO catalyst at 6:1 methanol-to-oil ratio, 5 wt% of catalyst,
and 65 °C for a duration of 2.5 h [84]. Jiadi et al. synthesized an eggshell-derived
CaO/Au composite in which eggshell was used to adsorb Au3+ in chloroauric acid
and calcinated further. He has reported a biodiesel yield of 88.9% when eggshell-
derived CaO/Au composite was used in the transesterification of soybean oil at 12:1
methanol-to-oil molar ratio, 1 wt% of catalyst, and 70 °C for 3 h. Eggshell-derived
CaO/Au composite exhibits good catalytic activity due to the promising joining of
CaO with Au nanoparticles [85]. Kavitha et al. obtained a biodiesel yield of 96% from
dairy scum waste using a CaO catalyst under optimum conditions of 6:1 methanol-to-
oil molar ratio, 2.4 wt% of catalyst, and 65 °C of reaction temperature for a reaction
time of 3 h [86]. Chen et al. have reported on the synthesis of mixed metal oxide
156 A. Velmurugan et al.

CaO–SiO2 using CaO and Na2 SiO3 . Synthesized CaO–SiO2 catalyst when used in
the transesterification of palm oil gives FAME yield of 90.2% [87]. Shan et al. have
obtained a biodiesel yield of 92.4% from vegetable oil using a diatomite-supported
CaO catalyst with high basicity. It shows enhanced catalytic activity and reusability
when used as a catalyst in the transesterification reaction. It has also been reported
that diatomite consists of SiO2 which reacts with CaO to form a Ca–O–Si bond. It
has also been reported that Ca/Si catalyst shows improved stability due to the well
scattering of Ca compounds on the support and a stable Ca–O–Si bond is formed.
The presence of the Ca–O–Si bond improves the stability of the Ca/Si composite
catalyst and also decreases the loss of Ca2+ ions.
Krishnamurthy et al. have reported on the synthesis of CaO from snail shell waste
and obtained a biodiesel yield of 96.92% when used in the transesterification of
dairy waste scum oil under optimized conditions of 12.7:1 methanol-to-oil molar
ratio, 0.866 wt% of catalyst, and 58.56 °C reaction temperature for 119.684 min.
Kaolin a natural substance rich in calcium can also catalyze biodiesel production,
but the catalytic activity will be less when it is used alone. Liu et al. have reported
that a solid base catalyst with improved catalytic activity is synthesized by mixing
different proportions of kaolin and snail shells followed by immersion in different
KBr concentrations since K+ concentrations show improved catalytic performance.
A biodiesel yield of 98.5% was attained when the synthesized catalyst was used in
the transesterification of soybean oil under optimal conditions of 6:1 methanol-to-oil
molar ratio, 2 wt% of catalyst, and 65 °C in a duration of 2 h.

6.4.2 Chitosan

Chitosan (C56 H103 N9 O39 ) is obtained from chitin by deacetylation reaction and is
widely found in aquatic organisms such as shrimp/crab shells, algal cell membranes,
and bones of molluscus. It is widely used as a material membrane since it is biodegrad-
able, hydrophilic with enhanced chemical stability, and naturally available, inexpen-
sive, and non-toxic polymer. Chitosan consists of hydroxyl, C–O–C, and amino
groups and is insoluble in water. Amino groups that are present in chitosan act as
basic sites that support the production of biodiesel. Moreover, amino groups can
also act as cations in an acidic medium. Chitosan can be used in catalysis in various
forms such as solids, powder, films, hydrogels, fibers, flakes, membranes, or sponges.
The drawbacks of polymer-based membranes are reduced mechanical strength and
enlargement after the long-term operation, which in turn decreases the separation
performance. The ways of improving the physical properties of chitosan-based poly-
mers are inserting inorganic nanoparticles within a polymeric phase such as silica,
calcium aluminosilicate, titanium oxide, silver, graphene oxide, and titanium dioxide.
Graphene oxide is considered to be inexpensive and also possesses a large surface
area, improved mechanical strength, and consists of oxygen functional groups, e.g.,
hydroxyl, epoxide, diol, ketone, and carboxyl which react to the amine group of
chitosan to form a bond between graphene oxide and biopolymer.
6 Green Catalysts Synthesized from Biomass for Biodiesel Production 157

Maryam Helmi et al. have reported on the synthesis of NaOH/Chitosan-Fe3 O4


catalyst with magnetic properties. A biodiesel yield of 92% is obtained under optimal
reaction conditions of 6:1 methanol-to-oil molar ratio, 0.5 wt% of catalyst, and 25
°C reaction temperature in 4.5 h [88].
Leila et al. have reported on the synthesis of zeolite-chitosan biocomposite modi-
fied using Na+ . The catalytic activity of Na+ is enhanced due to the reaction of
chitosan’s amino groups and Al–O–Na groups on the zeolite surface as well as with
the physical process of Na+ in pores and cages of the composite. Biodiesel yield of
96.5% is attained by electrocatalytic transesterification of waste fried oil using the
synthesized catalyst under optimal reaction conditions of 8:1 methanol-to-oil molar
ratio, 1 wt% of catalyst, and 10 wt% of acetone as a co-solvent at room temperature
in 30 min. Wang et al. have reported on the synthesis of magnetic carbonaceous acid
(Zr–CMC–SO3 H@3Fe–C400 ) having Bronsted and Lewis sites. On esterification of
oleic acid in the presence of the catalyst gives 97% biodiesel yield, whereas when
used in the transesterification of soybean oil gives 95% biodiesel yield in the pres-
ence of the catalyst. Wang et al. synthesized a double-shell mesoporous magnetic
solid acid catalyst in which Fe3 O4 metal nanoparticles were coated with chitosan
and silica two times and then sulfonated with p-toluene sulfonic acid. A biodiesel
yield of 96.7% is obtained in the esterification of oleic acid using the synthesized
catalyst at 15:1 methanol-to-oil molar ratio and 80 °C of reaction temperature. The
improved catalytic performance of the catalyst along with chitosan is due to the
ammonia NH3 sites on the catalyst surface that are derived from the amine groups
of chitosan (Table 6.1).

6.5 Nanocatalyst Derived from Plant Pigment

Flowers, leaves, bark, and fruits of plants contain pigments and dyes like chloro-
phylls and anthocyanin which can be removed using easy procedures, and these
plant sources are considered to be inexpensive, non-toxic, eco-friendly, and widely
available. These natural dyes and pigments like anthocyanin and chlorophyll from
China rose, anthocyanins from cumin, anthocyanin from red radish, curcumin from
turmeric, and also chlorophyll from cyanobacteria like Spirulina find application as
a sensitizer in photocatalysis for the breaking down of organic pollutants. The pres-
ence of a porphyrin ring in chlorophylls serves as a light-harvesting antenna for solar
energy collection. Chlorophyll can be used for photo-driven synthesis because of
its affinity for oxygen. It has been reported that photocatalysts like titanium dioxide
along with porphyrin can be synthesized using incipient wetness impregnation proce-
dure, and synthesized photocatalyst serves as an efficient sensitizer for attaining
visible light photocatalysis when compared to synthetic dyes. It has been reported that
85% of methylene blue degradation was attained at optimum conditions of amount
of catalyst of 0.1 wt% chlorophyll, initial pH of 6, and 20 ppm of initial methylene
blue concentration for 2 h of visible light exposure duration [89]. Moreover, chloro-
phyll pigment loaded on titanium dioxide helped in improving the photocatalytic
Table 6.1 Summarizes green catalyst derived from biomass for biodiesel production along with the conversion yield and operating conditions
158

S. No. Type of Type of Catalyst Transesterification conditions Biodiesel References


biomass feedstock FS T (°C) t (min) MTOR AC yield (%)
(wt%)
1 Agricultural Rice husk Rice husk biochar Vegetable oil 65 180 9:1 8 93.4 [41]
waste loaded with 30% of
CaO
2 Agricultural Rice husk Rice husk char Used cooking 65 120 12:1 4 98.2 [42]
waste doped oil
withpotassium
oxideandnickel
oxide
3 Agricultural Sugarcane Sulfonated Palm fatty 60 90 10:1 2 98.6 [61]
waste bagasse sugarcane bagasse acid distillate
4 Agricultural Coconut Ce-supported Chicken fat 90 60 12:1 3 93 [70]
waste shell sulfated activated and skin oil
carbon
5 Plant waste Potato peels Biochar Soybean 60 120 9:1 3 97.5 [72]
waste
cooking oil
6 Plant waste Banana Heterogeneous AC/ Palm oil 65 120 15:1 4 99.15 [74]
waste KOH
catalyst-supported
K2 CO3
7 Plant waste Pomelo Biochar loaded with Palm oil 65 150 8:1 6 98 [77]
waste 25 wt% of
potassium carbonate
(continued)
A. Velmurugan et al.
Table 6.1 (continued)
S. No. Type of Type of Catalyst Transesterification conditions Biodiesel References
biomass feedstock FS T (°C) t (min) MTOR AC yield (%)
(wt%)
8 Animal waste Eggshell Eggshell-derived Soybean oil 70 180 12:1 1 88.9 [85]
CaO/Au composite
9 Animal waste Chitosan NaOH/ Waste 25 270 6:1 0.5 92 [87]
Chitosan-Fe3 O4 cooking oil
10 Plant waste Kraft lignin Potassium Rapeseed oil 65 120 15:1 3 99.6 [97]
carbonate/kraft
lignin-activated
carbon
FS feedstock, T reaction temperature, t reaction time, MTOR methanol-to-oil molar ratio, AC amount of catalyst
6 Green Catalysts Synthesized from Biomass for Biodiesel Production
159
160 A. Velmurugan et al.

efficiency for the degradation of pollutant molecules. Chlorophyll is not stable in


organic solutions. Hence, doping on a desirable substrate, and silica materials can
improve the stability. Vahhab et al. have reported that titanium oxide nanoparticles
and chlorophyll were doped to change polyvinyl alcohol/Chitosan and can serve as
a photocatalyst for the degradation of methylene blue dye. Methylene blue degrada-
tion efficiency of (96%) was attained under visible light irradiation (LED lamp 70 W
by λ is 425 nm) in 60 min using bio nanocomposite film (poly(vinyl alcohol)/TiO2 /
chitosan/chlorophyll) [90].

6.6 Kraft Lignin as a Catalyst

Lignin is considered to be an inexpensive biopolymer with various links of carbon


in the form of C–O–C and C–C. The active sites present in the lignin help in the
adsorption of contaminated molecules, and the presence of hydroxyl groups helps
in hydrophilicity and functionality [91–93]. Kraft pulping method supplies oxygen
groups for lignin which enhances the properties of adsorption [94]. The structure
and composition of kraft lignin are completely different from that of lignin [95, 96].
Hence, kraft lignin finds application in wastewater treatment as well as in biodiesel
production. Zahra Nezafat has synthesized kraft lignin-titanium oxide/ palladium
nanocomposite and used it in the degradation of methylene blue and methylene
orange. Titanium oxide nanoparticles were immobilized on kraft lignin which acts as
a substrate. Metal oxides such as titanium oxide improve the attachment of other metal
nanoparticles such as palladium and also provide oxygen functional groups which
further enhance the hydrophilicity. Palladium nanoparticles help in the removal of
organic dyes from water. Moreover, it has been reported that bare palladium nanopar-
ticles give less catalytic efficiency when compared to the attached nanoparticles. The
author has reported that the degradation of methylene blue and methyl orange was
carried out using 1 mg of KL-T/Pd catalyst in 23 and 8 s [97].
Xian-fa Li et al. have reported that potassium carbonate/kraft lignin-activated
carbon (K2 CO3 /KLC) was synthesized using an in situ method of mixing K2 CO3
with technical kraft lignin (KL) followed by activating at 800 °C for a duration of
2 h in an inert atmosphere. The synthesized catalyst K2 CO3 /KLC when used in the
transesterification of rapeseed oil gives a biodiesel yield of 99.6% at optimal 15:1
methanol-to-oil molar ratio, 3 wt% of catalyst at a reaction temperature of 65 °C in 2
h [98]. Kraft lignin is obtained from the kraft pulping process which is considered to
be inexpensive, biodegradable, and easily procurable. The synthesized catalyst has
improved structural morphology and improved scattering of active components for
biodiesel production at moderate conditions.
6 Green Catalysts Synthesized from Biomass for Biodiesel Production 161

6.7 Important Factors that Affect the Production


of Activated Carbon

1. Activation temperature
2. Activation duration
3. Chemical impregnation ratio.

6.8 Activation Temperature

Supply of heat to an impregnated material helps in improving the thermal degradation


and volatilization process which helps in the development of pores and enhancement
in surface area. The activation temperature is selected based on aspects such as the
type of precursor and the chemical agent that is used in the method. Activation
temperature for various biomass feedstock varies from 400 to 800 °C, whereas coal-
based feedstock can progress as high as 900 °C. Huiping et al. have reported on
the effect of varying activation temperature for 60 min and an impregnation ratio of
100%. An increase in activation temperature reduces the yield of activated carbon due
to the burn-off of carbon [99]. Shenghui et al. have studied the effect of carbon dioxide
activation on porous structures of coconut shell-based activated carbon. Micropore
volume, BET surface area, and total pore volume of the activated carbon were lower
at an activation temperature of 750 °C since there is a slow reaction between char
and carbon dioxide. Whereas at an activation temperature of 900 °C, there is a
substantial increase in micropore volume, surface area, and total pore volume due to
the emission of volatiles as well as an enhanced reaction between carbon and carbon
dioxide. He has obtained a micropore percentage of 87.1 at 900 °C of activation
temperature and 600 cm3 /min of CO2 flow rate [100]. Moreover, it is also reported
that when the activation temperature is increased further to 950 °C, the widening of
pores occurs rapidly when compared to the formation of new pores which results in
enhancement in pore diameter and mesopore formation which in turn further reduces
the micropore volume, micropore percentage and BET surface area, and the average
pore diameter and total volume enhances. The widening of pores is due to the rapid
reaction between carbon and carbon dioxide.

6.9 Activation Duration

The duration of the activation plays an important role in the progress of the carbon’s
porous networks. The duration should be sufficient to remove all the moisture and
volatile components present in the precursor to create pores to develop. As soon as
the volatile escape starts to end, the creation of pore structures starts to develop,
and the activation should be stopped up to this point. Increased duration results in an
increase of pores at the cost of the surface area. Control of activation time is important
162 A. Velmurugan et al.

because a shorter duration reduces energy consumption and hence is considered to


be more desirable [101].
Bin Liu et al. have found out the surface area of rice husk activated carbon formu-
lated at a fixed concentration of zinc chloride and copper chloride at an activation
temperature of 500 °C for various activation time [102]. The surface area of the micro-
pore and the surface area obtained from BET analysis show an increasing trend as the
activation time enhances from 0.5 to 1.5 h. Reduction in surface area was observed
when activation time was extended beyond 1.5 h. The activation time can favor
the pyrolysis of rice husk and the evaporation of volatiles, supporting incomplete
combustion of carbon and creating new micropores. There was an enhancement in
surface area due to the increase in the amount of micropores and mesopores. When
the activation time is increased above a certain limit, there is an increase in the
amount of evaporation of the activator. Transverse activation plays a predominant
role and damages the pore structure of activated carbon. Hence, more micropores
and macropores developed which caused a reduction in the surface area.

6.10 Chemical/Impregnation Ratio

In the process of chemical activation, the impregnation ratio plays an important role
which has an impact on the quality of the obtained carbon. The impregnation ratio
is the ratio of the weights of the chemical agent and the dry precursor. The chemical
agents that are used in the activation method pierce deeper into the structure of the
carbon which in turn results in the growth of tiny pores. The surface area of the small
pores is considered to be larger [101]. K. Y. Foo et al. have reported that on enhancing
the impregnation ratio (KOH/char) from 0.25 to 1, there is an increase in the yield of
carbon; whereas a reduction in the yield of carbon is noticed when the impregnation
ratio is increased beyond 1. It has been reported that metallic potassium structured
during the gasification process would penetrate the internal structure of the char
matrix broadening the available pores and generating new porosities. The activation
method improved by increasing the impregnation ratio. Beyond the maximum value,
there is an expansion of pores and burn off which decreases the yield of carbon. Hence,
the impregnation ratio was set at 1 for efficient activation at the lowest consumption
of the activating agent (Table 6.2).

6.11 Chemical Activation Mechanism


of the Biomass-Derived Catalyst During Pyrolysis

The chemical activation mechanism of the biomass-derived catalyst during pyrolysis


can be explained by the following steps [108]:
6 Green Catalysts Synthesized from Biomass for Biodiesel Production 163

Table 6.2 Shows the activation effects due to several activation methods
Raw material Chemical agent Activation Activation effect References
temperature
Coconut shell NaOH 700 Mesoporous structure [103]
was observed when
there was an increase
in NaOH: biochar ratio
and mesopores that are
present in activated
carbon act as an
adsorbent for the
removal of methylene
blue from aqueous
solutions
Sour cherry stone ZnCl2 700 Surface area and pore [104]
volume of activated
carbon treated with
ZnCl2 is 1704 m2 g−1
and 1.566 cm3 g−1
when compared to the
sour cherry stones
(41.54 m2 g−1 ) 0.0975
cm3 g−1 . ZnCl2 helps
in the formation of
new micropores and
also helps in the
aggregation of
micropores into
mesopores
Banana leaf K2 CO3 750 The surface area and [105]
pore volume of
activated carbon
nanosheets are 1459
m2 g−1 and 0.652 cm3
g−1 when compared to
the surface area and
pore volume of 23
m2 g−1 and a pore
volume of 0.035 cm3
g−1 for non-activated
carbon
Activation helps in the
formation of
micropores
(continued)
164 A. Velmurugan et al.

Table 6.2 (continued)


Raw material Chemical agent Activation Activation effect References
temperature
Lotus leaves KOH 800 Lotus leaf porous [106]
carbon after activation
with KOH has a
surface area of 2440
m2 g−1 when
compared to the
surface area of 2.131
for lotus leaf. The pore
size of lotus leaf
porous carbon is < 4
nm. KOH develops
micro and mesopores
on the lotus leaves.
LLPC is used in the
adsorption of RhB
Reedy grass leaves H3 PO4 500 BET surface area and [107]
pore volume of
activated carbon is
1474 m2 /g, 0.560 cm3 /
g
H3 PO4 acts as an
activating reagent,
which will increase the
yield but will also
change the thermal
degradation of the
precursor, leading to a
subsequent change in
the evolution of
porosity

1. During the pyrolysis process, KOH reacts with the oxygen-containing groups
like C=O, –OH, O–C=O, C–O, and –COOH that are present in the biomass to
create more vacancies in the biochar due to the liberation of C=O, –OH, O–C=O,
C–O, and –COOH groups from the biomass.
2. KOH also reacts with the C–C and C–H groups that are obtained from broken
carbon particles to release hydrogen protons that form vacancies.
3. OH− ions from KOH enter these vacancies and form an abundant number of new
oxygen-containing groups in the biochar given in Eq. (6.3).

OH− + C(vacancies) → (C−O) + (−OH) + C−O + (O−C=O) + (−COOH)


(6.3)
6 Green Catalysts Synthesized from Biomass for Biodiesel Production 165

4. Conversion of KOH into K2 CO3 occurs due to the reactions taking place between
KOH and oxygen-containing groups/broken carbon particles at a temperature of
400–700 °C shown in Eqs. (6.4)–(6.9) which further converts into K2 O at 800 °C
given in Eqs. (6.10)–(6.13), followed by the evolution of potassium and gaseous
products.

6KOH + 2C → 2K2 CO3 + 2K + 3H2 (6.4)

KOH + (−COOH)/(−O−C=O) → K + K2 CO3 + H2 + CO2 (6.5)

KOH + (−C=O)/(−C−O−C) → K + K2 CO3 + H2 + CO (6.6)

KOH + (−O−CH3 ) → K + K2 CO3 + H2 + CH4 (6.7)

KOH + (C−OH) → K2 CO3 + K + H2 O + H2 (6.8)

KOH + (C−H) → K2 CO3 + K + H2 (6.9)

K2 CO3 + C → K2 O + 2CO (6.10)

K2 CO3 → K2 O + CO2 (6.11)

2K + CO2 → K2 O + CO (6.12)

K2 O + C → 2K + CO (6.13)

5. These steps favored enhanced porosity and oxygen content in the biochar.
6. Pyrolytic intermediates formed during pyrolysis on reacting with free radicals
and KOH yield phenol, aromatics, and gaseous products such as hydrogen, carbon
dioxide, carbon monoxide, and methane.
7. Some of the free radicals also convert directly into gaseous products.

6.12 Future Prospects

Green catalysts derived from various biomass sources help to reduce environmental
pollution caused by the dumping of this waste. Inexpensive green catalysts synthe-
sized from various biomass will be a boon to researchers in the field of wastew-
ater treatment, biodiesel production, fuel cells, and CO2 capture. Research on the
synthesis of green catalysts derived from biomass in the form of films is important
166 A. Velmurugan et al.

since the reusability of the catalyst is comparatively easier when compared to that in
its powder form.

6.13 Conclusions

Green catalyst synthesized from various biomass serves as an inexpensive and effi-
cient catalyst in biodiesel production. This chapter discusses various biomass from
which green catalysts can be synthesized along with the composition of biomass,
ways by which activation of biomass can be carried out to improve the porosity
of the obtained green catalyst, factors that affect the production of activated carbon,
and chemical activation mechanism of the biomass-derived catalyst during pyrolysis.
The selection of the catalyst plays an important role in enhancing the conversion of
biodiesel. Biodiesel with high yield and high quality can be produced on an indus-
trial scale using inexpensive green catalysts derived from biomass such as activated
carbon, NaOH/Chitosan-Fe3 O4 catalyst, eggshell-derived CaO/Au composite, and
biochar loaded with potassium carbonate which can serve as a substitute for the fossil
fuels.

References

1. Ratan, J. K., Kaur, M., & Adiraju, B. (2018). Synthesis of activated carbon from agricultural
waste using a simple method: Characterization, parametric and isotherms study. Materials
Today: Proceedings, 5(2), 3334–3345. https://doi.org/10.1016/j.matpr.2017.11.576
2. Adegoke, K. A., & Bello, O. S. (2015). Dye sequestration using agricultural wastes as
adsorbents. Water Resources and Industry, 12, 8–24. https://doi.org/10.1016/j.wri.2015.
09.002
3. Kumar, R. J., Manjeet, K., & Bharadwaj, A. (2018). Synthesis of activated carbon from
agricultural waste using a simple method: Characterization, parametric and isotherms study.
Materials Today: Proceedings, 5(2), 3334–3345. https://doi.org/10.1016/j.matpr.2017.11.576
4. Cazetta, A. L., Vargas, A. M. M., Nogami, E. M., Kunita, M. H., Guilherme, M. R., Martins,
A. C., Silva, T. L., Moraes, J. C. G., & Almeida, V. C. (2011). NaOH-activated carbon of
high surface area produced from coconut shell: Kinetics and equilibrium studies from the
methylene blue adsorption. Chemical Engineering Journal, 174, 117–125.
5. Daimary, N., Boruah, P., Eldiehy, KhalifaSH., Pegu, T., Bardhan, P., Bora, U., Mandal, M., &
Deka, D. (2022). Musa acuminata peel: A bioresource for bio-oil and by-product utilization as
a sustainable source of renewable green catalyst for biodiesel production. Renewable Energy,
187, 450–462.
6. Baroutian, S., Aroua, M. K., Raman, A. A. A., & Sulaiman, N. M. N. (2010). Potassium
hydroxide catalyst supported on palm shell activated carbon for transesterification of palm
oil. Fuel Processing Technology, 91, 1378–1385.
7. Irshad, M. A., Nawaz, R., Rehman, M. Z., Adrees, M., Rizwan, M., Ali, S., Ahmad, S., &
Tasleem, S. (2021). Synthesis, characterization and advanced sustainable applications of
titanium dioxide nanoparticles: a review. Ecotoxicology and Environmental Safety, 212,
111978.
6 Green Catalysts Synthesized from Biomass for Biodiesel Production 167

8. Sivanesh, S., Aswin, K. N., Antony, A., SuryaVarma, M., Lakshmi, A., Kamalesh, K.,
Naageshwaran, M., Soundarya, S., & Subramanian, S. (2022). Biodiesel production from
custard apple seeds and Euglena sanguinea using CaO nano-catalyst. Bioresource Tech-
nology., 344, 126418.
9. Sahu, S., Saikia, K., Gurunathan, B., Dhakshinamoorthy, A., & Rokhum, S. L. (2023). Green
synthesis of CaO nanocatalyst using watermelon peels for biodiesel production, 547, 113342.
10. Cholapandian, K., Gurunathan, B., Rajendran, N. (2022). Investigation of CaO nanocatalyst
synthesized from Acalypha indica leaves and its application in biodiesel production using
waste cooking oil, 312, 122958.
11. Yusuff, A. S., Gbadamosi, A. O., & Popoola, L. T. (2021). Biodiesel production from trans-
esterified waste cooking oil by zinc-modified anthill catalyst: Parametric optimization and
biodiesel properties improvement. Journal of Environmental Chemical Engineering, 9(2),
104955.
12. Arshad, S., Ahmad, M., Munir, M., Sultana, S., Zafar, M., Dawood, S., Rozina, AhmadM.,
Alghamdi, S. A., Bokhari, A., Mubashir, M., Chuah, L. F., & Show, P. L. (2023). Assessing
the potential of green CdO2 nano-catalyst for the synthesis of biodiesel using non-edible seed
oil of Malabar Ebony. Fuel, 333, 126492.
13. Yaşar, F. (2019). Biodiesel production via waste eggshell as a low-cost heterogeneous catalyst:
Its effects on some critical fuel properties and comparison with CaO. Fuel, 255, 115828.
https://doi.org/10.1016/j.fuel.2019.115828
14. Mamo, T. T., & Mekonnen, Y. S. (2019). Microwave-assisted biodiesel production from
microalgae, scenedesmus species, using goat bone–made nano-catalyst. Applied Biochemistry
and Biotechnology, 190, 1147–1162. https://doi.org/10.1007/s12010-019-03149-0
15. Mosaddegh, E., & Hassankhani, A. (2014). Preparation and characterization of nano-CaO
based on eggshell waste: Novel and green catalytic approach to highly efficient synthesis of
pyrano[4,3-b] pyrans. Chinese Journal of Catalysis, 35(2), 351–356.
16. Gao, Y., & Xu, C. (2012). Synthesis of dimethyl carbonate over waste eggshell catalyst.
Catalysis Today, 190(1), 107–111. https://doi.org/10.1016/j.cattod.2011.12.004
17. Ahmad, A. L., Yasin, N. H. M., Derek, C. J. C., & Lim, J. K. (2011). Microalgae as a
sustainable energy source for biodiesel production: A review. Renewable and Sustainable
Energy Reviews, 15(1), 584–593.
18. Jazie, A. A., Sinha, H., & Pramanik, A. S. K. (2013). Transesterification of peanut and rapeseed
oils using the waste of animal bone as a cost-effective catalyst. Materials for Renewable and
Sustainable Energy, 2(11), 1–10.
19. Isikgor, F. H., & Becker, C. R. (2015). Lignocellulosic biomass: A sustainable platform for
the production of bio-based chemicals and polymers. Polymer Chemistry, 6, 4497–4559.
20. Chen, K.-T., Wang, J.-X., Dai, Y.-M., Wang, P.-H., Liou, C.-Y., Nien, C.-W., Wu, J.-S., &
Chen, C.-C. (2013). Rice husk ash as a catalyst precursor for biodiesel production. Journal of
the Taiwan Institute of Chemical Engineers, 44(4), 622–629. https://doi.org/10.1016/j.jtice.
2013.01.006
21. Liou, T.-H., & Wu, S.-J. (2009). Characteristics of microporous/mesoporous carbons prepared
from rice husk under base- and acid-treated conditions. Journal of Hazardous Materials, 171,
693–703.
22. Lin, L., Zhai, S.-R., Xiao, Z.-Y., Song, Y., An, Q.-D., & Song, X.-W. (2013). Dye adsorption
of mesoporous activated carbons produced from NaOH-pretreated rice husks. Bioresource
Technology, 136, 437–443.
23. Muniandy, L., Adam, F., Mohamed, A. R., & Ng, E.-P. (2014). The synthesis and characteri-
zation of high purity mixed microporous/mesoporous activated carbon from rice husk using
chemical activation with NaOH and KOH. Microporous and Mesoporous Materials, 197,
316–323. https://doi.org/10.1016/j.micromeso.2014.06.020
24. Quispe, I., Navia, R., & Kahhat, R. (2017). Energy potential from rice husk through direct
combustion and fast pyrolysis: A review. Waste Management, 59, 200–210. https://doi.org/
10.1016/j.wasman.2016.10.001
168 A. Velmurugan et al.

25. Krishnarao, R. V., Subrahmanyam, J., & Jagadish Kumar, T. (2001). Studies on the formation
of black particles in rice husk silica ash. Journal of European Ceramic Society, 21, 99–104.
26. Prasetyoko, D., Ramli, Z., Endud, S., Hamdan, H., & Sulikowski, B. (2006). Conversion of
rice husk ash to zeolite beta. Waste Management, 26, 1173–1179.
27. Othman Ali, I., Hassan, A. M., Shaaban, S. M., & Soliman, K. S. (2011). Synthesis and char-
acterization of ZSM-5 zeolite from rice husk ash and their adsorption of Pb2+ onto unmodified
and surfactant-modified zeolite. Separation and Purification Technology, 83, 38–44. https://
doi.org/10.1016/j.seppur.2011.08.034
28. Pimprom, S., Sriboonkham, K., Dittanet, P., Föttinger, K., Rupprechter, G., &
Kongkachuichay, P. (2015). Synthesis of copper–nickel/SBA-15 from rice husk ash catalyst
for dimethyl carbonate production from methanol and carbon dioxide. Journal of Industrial
Engineering Chemistry, 31, 156–166. https://doi.org/10.1016/j.jiec.2015.06.019
29. López, M., Palacio, R., Royer, S., Mamede, A. S., & Fernández, J. J. (2020). Mesostructured
CMK-3 carbon supported Ni–ZrO2 as catalysts for the hydrodeoxygenation of guaiacol.
Microporous Mesoporous Material, 292, 109694. https://doi.org/10.1016/j.micromeso.2019.
109694109694
30. Jeong, Y., Cui, M., Choi, J., Lee, Y., Kim, J., Son, Y., & Khim, J. (2020). Development of modi-
fied mesoporous carbon (CMK-3) for improved adsorption of bisphenol-A. Chemosphere,
238, 124559. https://doi.org/10.1016/j.chemosphere.2019.124559124559
31. Zhang, T., Qiu, H., Zhang, M., Fang, Z., Zhao, X., Wang, L., Chen, G., Wei, Y., Yue, H.,
Wang, C., & Zhang, D. (2017). A unique 2D-on-3D architecture developed from ZnMn2 O4
and CMK-3 with excellent performance for lithium-ion batteries. Carbon, 123, 717–725.
https://doi.org/10.1016/j.carbon.2017.08.013
32. Juárez, J. M., Ledesma, B. C., Costa, M. G., Beltramone, A. R., & Anunziata, O. A.
(2017). Novel preparation of CMK-3 nanostructured material modified with titania applied
in hydrogen uptake and storage. Microporous Mesoporous Materials, 254, 146–152. https://
doi.org/10.1016/j.micromeso.2017.03.056
33. Regiart, M., Fernández-Baldo, M. A., Villarroel-Rocha, J., Messina, G. A., Bertolino, F.
A., Sapag, K., Timperman, A. T., & Raba, J. (2017). Microfluidic immunosensor based on
mesoporous silica platform and CMK-3/poly-acrylamide-co-methacrylate of dihydrolipoic
acid modified gold electrode for cancer biomarker detection. Analytica Chimica Acta, 963,
83–92. https://doi.org/10.1016/j.aca.2017.01.029
34. Zhang, G., Chen, Y., Huang, K., Chen, Y., & Guo, H. (2018). CMK-3/NiCo2 S4 nanostructures
for high-performance asymmetric supercapacitors. Materials Chemistry and Physics, 220,
270–277. https://doi.org/10.1016/j.matchemphys.2018.09.008
35. Ahmed, M. J. (2016). Preparation of activated carbons from date (Phoenix dactylifera L.) palm
stones and application for wastewater treatments. Review Process Safety and Environmental
Protection, 102, 168–182. https://doi.org/10.1016/j.psep.2016.03.010
36. Huang, F. C., Lee, C. K., Han, Y. L., Chao, W. C., & Chao, H. P. (2014). Preparation of
activated carbon using micro-nano carbon spheres through chemical activation. Journal of
Taiwan Institute of Chemical Engineers, 45(5), 2805–2812.
37. Naji, S. Z., & Tye, C. T. (2022). A review of the synthesis of activated carbon for biodiesel
production: Precursor, preparation, and modification. Energy Conversion and Management:
X, 13, 100152.
38. Mazlan, M. A. F., Uemura, Y., Yusup, S., Elhassan, F., Uddin, A., & Hiwada, A. I. (2016). Acti-
vated carbon from rubber wood sawdust by carbon dioxide activation. Procedia Engineering,
148, 530–537.
39. Liu, D., Zhang, W., Lin, H., Li, Y., Lu, H., & Wang, Y. (2016). A green technology for
the preparation of high capacitance rice husk-based activated carbon. Journal of Cleaner
Production, 112, 1190–1198.
40. Ahiduzzaman, M., & Sadrul Islam, A. K. M. (2016). Preparation of porous bio-char and
activated carbon from rice husk by leaching ash and chemical activation. Springerplus, 5,
1248.
6 Green Catalysts Synthesized from Biomass for Biodiesel Production 169

41. Zhao, C., Yang, L., Xing, S., Luo, W., Wang, Z., & Lv, P. (2018). Biodiesel production by a
highly effective renewable catalyst from pyrolytic rice husk. Journal of Cleaner Production,
199, 772–780.
42. Hazmi, B., Rashid, U., Ibrahim, M. L., Nehdi, I. A., Azam, M., & Al-Resayes, S. I. (2021).
Synthesis and characterization of bifunctional magnetic nano-catalyst from rice husk for
production of biodiesel. Environmental Technology & Innovation, 21, 101296.
43. Zhao, C., Yang, L., Xing, S., Luo, W., Wang, Z., & Lv, P. (2018). Biodiesel production by a
highly effective renewable catalyst from pyrolytic rice husk. Journal of Cleaner Production,
199, 772–780. https://doi.org/10.1016/j.jclepro.2018.07.242
44. Vinayagam, M., Ramachandran, S., Ramya, V., & Sivasamy, A. (2018). Photocatalytic degra-
dation of orange G dye using ZnO/biomass activated carbon nanocomposite. Journal of
Environmental Chemical Engineering, 6(3), 3726–3734.
45. Liew, R. K., Chong, M. Y., Osazuwa, O. U., Nam, W. L., Phang, X. Y., Su Huan, M. H.,
Cheng, C. K., Chong, C. T., & Lam, S. S. (2018). Production of activated carbon as catalyst
support by microwave pyrolysis of palm kernel shell: A comparative study of chemical versus
physical activation. Research on Chemical Intermediates, 44(6), 3849–3865. https://doi.org/
10.1007/s11164-018-3388-y
46. Srimachai, T., Thonglimp, V., & Sompong, O. (2014). Ethanol and methane production from
oil palm frond by two-stage SSF. Energy Procedia, 52, 352.
47. Ang, S. K., Shaza, E. M., Adibah, Y., Suraini, A. A., & Madihah, M. S. (2013). Production of
cellulases and xylanase by Aspergillus fumigatus SK1 using untreated oil palm trunk through
solid state fermentation. Process Biochemistry, 48(9), 1293–1302.
48. Ishola, M. M., Isroi, T., & Mohammad, J. (2014). Effect of fungal and phosphoric acid
pretreatment on ethanol production from oil palm empty fruit bunches (OPEFB). Bioresource
Technology, 165, 9–12. https://doi.org/10.1016/j.biortech.2014.02.053
49. Saidu, M., Yuzir, A., Salim, M. R., Salmiati, S. A., & Abdullah, N. (2014). Biological
pre-treated oil palm mesocarp fiber with cattle manure for biogas production by anaerobic
digestion during acclimatization phase. International Biodeterioration & Biodegradation, 95,
189–194. https://doi.org/10.1016/j.ibiod.2014.06.014
50. Neutelings, G. (2011). Lignin variability in plant cell walls: Contribution of new models.
Plant Science, 181(4), 379–386.
51. Cao, J., Xiao, G., Xu, X., Shen, D. K., & Jin, B. (2013). Study on carbonization of lignin
by TG-FTIR and high-temperature carbonization reactor. Fuel Processing Technology, 106,
41–47. https://doi.org/10.1016/j.fuproc.2012.06.016
52. Collard, F.-X., & Blin, J. (2014). A review on pyrolysis of biomass constituents: Mechanisms
and composition of the products obtained from the conversion of cellulose, hemicelluloses,
and lignin. Renewable and Sustainable Energy Reviews, 38, 594–608. https://doi.org/10.1016/
j.rser.2014.06.013
53. Abdullah, R. F., Rashid, U., Taufiq-Yap, Y. H., Ibrahim, M. L., Ngamcharussrivichai, C., &
Azam, M. (2020). Synthesis of bifunctional nanocatalyst from waste palm kernel shell and
its application for biodiesel production. Royal Society of Chemistry, 10, 27183–27193.
54. Goh, C. M. H., Tan, Y. H., Mubarak, N. M., Kansedo, J., Rashid, U., Khalid, M., & Walvekar,
R. (2021). Synthesis of magnetic basic palm kernel shell catalyst for biodiesel production and
characterization and optimization by Taguchi method. Applied Nanoscience, 12, 3721–33.
https://doi.org/10.1007/s13204-021-01815-6
55. Kostić, M. D., Bazargan, A., Stamenković, O. S., Veljković, V. B., & McKay, G. (2016).
Optimization and kinetics of sunflower oil methanolysis catalyzed by calcium oxide-based
catalyst derived from palm kernel shell biochar. Fuel, 163, 304–313. https://doi.org/10.1016/
j.fuel.2015.09.042
56. Cheah, K. Y., Toh, T. S., & Koh, P. M. (2010). Palm fatty acid distillate biodiesel: Next
generation palm biodiesel. International News on Fats, Oils, and Related Materials, 21, 264–
266.
57. Betancur, G. J. V., & Pereira, N. (2010). Sugar cane bagasse as feedstock for second generation
ethanol production. Part I: Diluted acid pretreatment optimization. Electronic Journal of
Biotechnology, 13(3), 1–9.
170 A. Velmurugan et al.

58. Kumar, A., Negi, Y. S., Choudhary, V., & Bhardwaj, N. K. (2020). Characterization of cellulose
nanocrystals produced by acid-hydrolysis from sugarcane bagasse as agro-waste. Journal of
Materials Physics and Chemistry, 2(1), 1–8.
59. Rezende, C. A., De Lima, M., Maziero, P., Deazevedo, E., Garcia, W., & Polikarpov, I.
(2011). Chemical and morphological characterization of sugarcane bagasse submitted to a
delignification process for enhanced enzymatic digestibility. Biotechnology for Biofuels, 4(1),
1–19.
60. Chausali, N., Saxena, J., & Prasad, R. (2021). Nanobiochar and biochar based nanocomposites:
Advances and applications. Journal of Agriculture and Food Research, 5, 100191.
61. Akinfalabi, S.-I., Rashid, U., Ngamcharussrivichai, C., & Nehdi, I. A. (2020). Synthesis
of reusable biobased nano-catalyst from waste sugarcane bagasse for biodiesel produc-
tion. Environmental Technology & Innovation, 18, 100788. https://doi.org/10.1016/j.eti.2020.
100788
62. Niju, S., Kanna, S. K. A., Ramalingam, V., Kumar, M. S., & Balaji, M. (2019). Sugarcane
bagasse derived biochar—A potential heterogeneous catalyst for transesterification process.
Energy Sources, Part A: Recovery, Utilization, and Environmental Effects, 45(4), 9815–9826.
https://doi.org/10.1080/15567036.2019.1680771
63. Mutalib, A. A. A., Ibrahim, M. L., Matmin, J., Kassim, M. F., Mastuli, M. S., Taufiq-Yap, Y.
H., Shohaimi, N. A. M., Islam, A., Tan, Y. H., & Kaus, N. H. M. (2020). SiO2 -rich sugar cane
bagasse ash catalyst for transesterification of palm oil. BioEnergy Research, 13, 986–997.
https://doi.org/10.1007/s12155-020-10119-6
64. El Nemr, A., El-Sikaily, A., & Khaled, A. (2010). Modeling of adsorption isotherms of
methylene blue onto rice husk activated carbon. Egyptian Journal of Aquatic Research, 36(3),
403–425.
65. El Nemr, A., El Sadaawy, M. M., Khaled, A., & El Sikaily, A. (2014). Adsorption of the
anionic dye direct red 23 onto new activated carbons developed from Cynara cardunculus:
Kinetics, equilibrium and thermodynamics. Blue Biotechnology Journal, 3(1), 121–142.
66. Shoaib, A. G. M., El-Sikaily, A., El Nemr, A., Mohamed, A.E.-D.A., & Hassan, A. A. (2020).
Preparation and characterization of highly surface area activated carbons followed type IV
from marine red alga (Pterocladia capillacea) by zinc chloride activation. Biomass Convers
Biorefinery, 12, 2253–2265. https://doi.org/10.1007/s13399-020-00760-8
67. Shoaib, A. G. M., El-Sikaily, A., El Nemr, A., Mohamed, A.E.-D.A., & Hassan, A. A. (2020).
Testing the carbonization condition for high surface area preparation of activated carbon
followed type-IV from green alga Ulva lactuca. Biomass Conversion and Biorefinery, 12,
3303–3318. https://doi.org/10.1007/s13399-020-00823-w
68. Sekhon, S. S., Kaur, P., & Park, J.-S. (2021). From coconut shell biomass to oxygen reduction
reaction catalyst: Tuning porosity and nitrogen doping. Renewable and Sustainable Energy
Reviews, 147, 111173. https://doi.org/10.1016/j.rser.2021.111173
69. Ioannidou, O., & Zabaniotou, A. (2006). Agricultural residues as precursors for activated
carbon production—A review. Renewable and Sustainable Energy Reviews, 11(9), 1966–
2005.
70. Shobhana-Gnanaserkhar, A.-M., AbdulKareem-Alsultan, G., Sivasangar-Seenivasagam, S.
M., & Izham, T.-Y. (2020). Biodiesel production via simultaneous esterification and transes-
terification of chicken fat oil by mesoporous sulfated Ce supported activated carbon. Biomass
and Bioenergy, 141, 105714.
71. Mozammel, H. M., Masahiro, O., & Bhattacharya, S. C. (2002). Activated charcoal from
coconut shell using ZnCl2 activation. Biomass and Bioenergy, 22(5), 397–400. https://doi.
org/10.1016/s0961-9534(02)00015-6
72. Daimary, N., Eldiehy, K. S. H., Boruah, P., Deka, D., Bora, U., & Kakati, B. K. (2022).
Potato peels as a sustainable source for biochar, bio-oil and a green heterogeneous catalyst
for biodiesel production. Journal of Environmental Chemical Engineering, 10(1), 107108.
73. Daimary, N., Boruah, P., KhalifSHEldiehy, T. P., Bardhan, P., Bora, U., Mandal, M., & Deka,
D. (2022). Musa acuminata peel: A bioresource for bio-oil and by-product utilization as a
sustainable source of renewable green catalyst for biodiesel production. Renewable Energy,
187, 450–462.
6 Green Catalysts Synthesized from Biomass for Biodiesel Production 171

74. Jitjamnong, J., Thunyaratchatanon, C., Luengnaruemitchai, A., Kongrit, N., Kasetsomboon,
N., Sopajarn, A., & Khantikulanon, N. (2020). Response surface optimization of biodiesel
synthesis over a novel biochar-based heterogeneous catalyst from cultivated (Musa sapientum)
banana peels. Biomass Conversion and Biorefinery, 11, 2795–2811. https://doi.org/10.1007/
s13399-020-00655-8
75. Brahma, S., Basumatary, B., Basumatary, S. F., Das, B., Brahma, S., Rokhum, S. L., &
Basumatary, S. (2023). Biodiesel production from quinary oil mixture using highly efficient
Musa chinensis based heterogeneous catalyst. Fuel, 336, 127150.
76. Shah, J., Janneth Loez-Mercado, M., Carreon, G., Lopez-Miranda, A., & Carreon, M. L.
(2018). Plasma synthesis of graphene from mango peel. ACS Omega, 3(1), 455–463. https://
doi.org/10.1021/acsomega.7b01825
77. Zhao, C., Lv, P., Yang, L., Xing, S., Luo, W., & Wang, Z. (2018). Biodiesel synthesis over
biochar-based catalyst from biomass waste pomelo peel. Energy Conversion and Management,
160, 477–485. https://doi.org/10.1016/j.enconman.2018.01.059
78. Nath, B., Das, B., Kalita, P., & Basumatary, S. (2019). Waste to value addition: Utilization
of waste Brassica nigra plant derived novel green heterogeneous base catalyst for effective
synthesis of biodiesel. Journal of Cleaner Production, 239, 118112.
79. Aleman-Ramireza, J. L., Moreiraa, J., Torres-Arellanoa, S., Adriana Longoria, C., Okoyeb,
P. U., & Sebastian, P. J. (2021). Preparation of a heterogeneous catalyst from Moringa leaves
as a sustainable precursor for biodiesel production. Fuel, 284, 118983.
80. Rakopoulos, C. D., Antonopoulos, K. A., Rakopoulos, D. C., Hountalas, D. T., & Giakoumis,
E. G. Comparative performance and emissions study of a direct injection diesel engine using
blends of diesel fuel with vegetable oils or bio-diesels of various origins. Energy Conversion
and Management, 47(18–19), 3272–3287.
81. Narowska, B., Kułażyński, M., Łukaszewicz, M., & Burchacka, E. (2019). Use of activated
carbons as catalyst supports for biodiesel production. Renewable Energy, 135, 176–185.
82. Hameed, A., Naqvi, S. R., Sikandar, U., & Chen, W.-H. (2022). One-step biodiesel production
from waste cooking oil using cao promoted activated carbon catalyst from Prunus persica
seeds. Catalysts, 12, 592. https://doi.org/10.3390/catal12060592
83. Kouzu, M., Kasuno, T., Tajika, M., Sugimoto, Y., Yamanaka, S., & Hidaka, J. (2008). Calcium
oxide as a solid base catalyst for transesterification of soybean oil and its application to
biodiesel production. Fuel, 87, 2798–2806.
84. Rahman, W. U., Fatima, A., Anwer, A. H., Athar, M., Khan, M. Z., Khan, N. A., & Halder, G.
(2018). Biodiesel synthesis from eucalyptus oil by utilizing waste egg shell derived calcium-
based metal oxide catalyst. Process Safety and Environmental Protection, 122, 313–319.
https://doi.org/10.1016/j.psep.2018.12.015
85. Liu, J., Liu, M., Chen, S., Wang, B., Chen, J., Yang, D.-P., Zhang, S., & Du, W. (2020). Conver-
sion of Au(III)-polluted waste eggshell into functional CaO/Au nanocatalyst for biodiesel
production. Green Energy & Environment, 7(2), 352–359. https://doi.org/10.1016/j.gee.2020.
07.019
86. Kavitha, V., Geetha, V., Jacqueline, P., & Jennita. (2019). Production of biodiesel from dairy
waste scum using eggshell waste. Process Safety and Environmental Protection, 125, 279–
287.https://doi.org/10.1016/j.psep.2019.03.021
87. Chen, G., Shan, R., Li, S., & Shi, J. (2015). A biomimetic silicification approach to synthesize
CaO–SiO2 catalyst for the transesterification of palm oil into biodiesel. Fuel, 153, 48–55.
https://doi.org/10.1016/j.fuel.2015.02.109
88. Helmi, M., & Hemmati, A. (2021). Synthesis of magnetically solid base catalyst of NaOH/
Chitosan-Fe3 O4 for biodiesel production from waste cooking oil: Optimization, kinetics and
thermodynamic studies. Energy Conversion and Management, 248, 114807.
89. Krishnan, S., & Shriwastav, A. (2020). Application of TiO2 nanoparticles sensitized with
natural chlorophyll pigments as catalyst for visible light photocatalytic degradation of
methylene blue. Journal of Environmental Chemical Engineering, S2213-3437(20), 31048-4.
172 A. Velmurugan et al.

90. Soltaninejad, V., & Maleki, A. (2021). A green, and eco-friendly bionanocomposite film
(poly(vinyl alcohol)/TiO2 /chitosan/chlorophyll) by photocatalytic ability, and antibacte-
rial activity under visible-light irradiation. Journal of Photochemistry & Photobiology, A:
Chemistry, 404, 112906.
91. Ge, Y., & Li, Z. (2018). Application of lignin and its derivatives in adsorption of heavy metal
ions in water: A review. ACS Sustainable. Chemistry and Engineering, 6, 7181–7192.
92. Kajita, S., Hishiyama, S., Tomimura, Y., Katayama, Y., & Omori, S. (1997). Structural charac-
terization of modified lignin in transgenic tobacco plants in which the activity of 4-coumarate:
Coenzyme A ligase is depressed. Plant Physiology, 114, 871–879.
93. Li, Z., Zhang, J., Qin, L., & Ge, Y. (2018). Enhancing the antioxidant performance of lignin
by enzymatic treatment with laccase. ACS Sustainable Chemistry and Engineering, 6, 2591–
2595.
94. Mohan, D., Pittman, C. U., Jr., & Steele, P. H. (2006). Single, binary and multi-component
adsorption of copper and cadmium from aqueous solutions on Kraft lignin-a biosorbent.
Journal of Colloid and Interface Science, 297, 489–504.
95. Schlee, P., Hosseinaei, O., Baker, D., Landmér, A., Tomani, P., Mostazo-López, M. J., Cazorla-
Amorós, D., Herou, S., & Titirici, M. M. (2019). From waste to wealth: from kraft lignin to
free-standing supercapacitors. Carbon, 145, 470–480.
96. Chatterjee, S., & Saito, T. (2015). Lignin-derived advanced carbon materials. ChemSusChem-
Chemistry Europe, 8, 3941–3958.
97. Nezafat, Z., Mohazzab, B. F., Jaleh, B., Nasrollahzadeh, M., Baran, T., & Shokouhimehr,
M. (2021). A promising nanocatalyst: Upgraded kraft lignin by titania and palladium
nanoparticles for organic dyes reduction. Inorganic Chemistry Communications, 130, 108746
98. Li, X.-f, Zuo, Y., Zhang, Y., Fu, Y., & Guo, Q.-X. (2013). In situ preparation of K2 CO3
supported Kraft lignin activated carbon as a solid base catalyst for biodiesel production. Fuel,
113, 435–442.
99. Zhang, H., Yan, Y., & Yang, L. (2010). Preparation of activated carbon from sawdust by zinc
chloride activation. Adsorption, 16(3), 161–166. https://doi.org/10.1007/s10450-010-9214-5
100. Guo, S., Peng, J., Li, W., Yang, K., Zhang, L., Zhang, S., & Xia, H. (2009). Effects of CO2
activation on porous structures of coconut shell-based activated carbons. Applied Surface
Science, 255(20), 8443–8449. https://doi.org/10.1016/j.apsusc.2009.05.150
101. Gratuito, M. K. B., Panyathanmaporn, T., Chumnanklang, R. A., Sirinuntawittaya, N., &
Dutta, A. (2008). Production of activated carbon from coconut shell: Optimization using
response surface methodology, 99(11), 4887–4895. https://doi.org/10.1016/j.biortech.2007.
09.042
102. Liu, B., Gu, J., & Zhou, J. (2016). High surface area rice husk-based activated carbon prepared
by chemical activation with ZnCl2 –CuCl2 composite activator. Environmental Progress &
Sustainable Energy, 35(1), 133–140. https://doi.org/10.1002/ep.12215
103. Cazetta, A. L., Vargas, A. M. M., Nogami, E. M., Kunita, M. H., Guilhermea, M. R., Martins,
A. C., Silva, T. L., Moraes, J. C. G., & Almeida, V. C. (2011). NaOH-activated carbon of
high surface area produced from coconut shell: Kinetics and equilibrium studies from the
methylene blue adsorption. Chemical Engineering Journal, 174, 117–125.
104. Angin, D. (2014). Production and characterization of activated carbon from sour cherry stones
by zinc chloride. Fuel, 115, 804–811.
105. Roy, C. K., Shah, S. S., Reaz, A. H., Sultana, S., Chowdhury, A.-N., Firoz, S. H., Zahir,
M. H., Qasem, M. A. A., & Aziz, M. A. (2020). Preparation of hierarchical porous activated
carbon from banana leaves for high-performance supercapacitor: Effect of type of electrolytes
on performance. Chemistry—An Asian Journal, 16(4), 296–308. https://doi.org/10.1002/asia.
202001342
106. Li, A., Huang, W., Qiu, N., Mou, F., & Wang, F. (2020). Porous carbon prepared from
lotus leaves as potential adsorbent for efficient removal of rhodamine B. Materials Research
Express, 7, 055505.
107. Xu, J., Chen, L., Qu, H., Jiao, Y., Xie, J., & Xing, G. (2014). Preparation and characterization
of activated carbon from reedy grass leaves by chemical activation with H3 PO4 . Applied
Surface Science, 320, 674–680.
6 Green Catalysts Synthesized from Biomass for Biodiesel Production 173

108. Chen, W., Gong, M., Li, K., Xia, M., Chen, Z., Xiao, H., Fang, Y., Chen, Y., Yang, H., &
Chen, H. (2020). Insight into KOH activation mechanism during biomass pyrolysis: Chemical
reactions between O-containing groups and KOH. Applied Energy, 278, 115730. https://doi.
org/10.1016/j.apenergy.2020.115730
Chapter 7
Sustainable Solutions for Bioenergy
Production from Hospital-Based Plastic
Waste—Thinking Beyond Landfills

Patitapaban Dash, Chirasmita Mohanty, Pratyush Kumar Das,


Joseph M Anto Simon, Debasish Sahoo, and Gurunathan Baskar

Abstract The use of plastics and the growth of modern plastic industries began in
the early twentieth century, thus providing the human populace with a very useful tool
that found widespread applications in various sectors. Mostly synthesized from petro-
chemicals, plastic’s high durability makes it one of the most widely used synthetic
materials. Its characteristic properties like lightweight, ability to be molded, insu-
lating ability, and resistance against degradation have made it an ideal choice for
being used to manufacture different consumer goods, medical equipment, and safety
gears. The hospital sector is one of the largest users of plastic products owing to their
non-reactive and durable nature simultaneously producing a proportionate amount
of disposable plastic waste. The COVID pandemic saw a huge rise in the use of
plastic-based materials in the healthcare sector, thus increasing the waste load in
the environment. The long-term environmental persistence of plastics due to their

P. Dash · C. Mohanty
Centre for Biotechnology, Siksha ‘O’ Anusandhan (Deemed to be University), Bhubaneswar,
Odisha 751003, India
P. K. Das (B)
Department of Phytopharmaceuticals, School of Agricultural Engineering and Bioprocessing
(SoABE), Centurion University of Technology and Management, Paralakhemundi,
Odisha 761200, India
e-mail: pratyush.das@cutm.ac.in
M. A. S. Joseph
Department of Biotechnology, Sri Krishna Arts and Science College, Bharathiar University,
Coimbatore, Tamil Nadu 641008, India
D. Sahoo
BioInnovale Lifescience (P) Ltd, Bengaluru, Karnataka, India
G. Baskar
Department of Biotechnology, St. Joseph’s College of Engineering, Chennai, Tamil Nadu, India
C. Mohanty
Department of Biotechnology, School of Bio Sciences and Technology, Vellore Institute of
Technology, Vellore 632014, Tamil Nadu, India

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2024 175
G. Baskar et al. (eds.), Circular Bioeconomy Perspectives in Sustainable Bioenergy
Production, Energy, Environment, and Sustainability,
https://doi.org/10.1007/978-981-97-2523-6_7
176 P. Dash et al.

non-biodegradable nature coupled with improper disposal and conventional waste


management systems have led to the accumulation of these synthetic polymers in
different environmental matrices, thereby causing pollution. It is essential to devise
sustainable solutions to reduce the load of plastic waste in the environment, thereby
nullifying its toxicity. In this context, the conversion of plastic waste into suitable
energy forms is suggested as a means to sustainably reduce the plastic contami-
nants in the environment and parallelly supply energy alternatives to the dwindling
conventional energy resources.

Keywords Bioenergy · Environmental toxicity · Marine pollutants · Pyrolysis of


plastic · Sustainable solutions · Valorization of plastic

7.1 Introduction

Plastics are one of the most important inventions ever come across by human civi-
lization. Plastics are generally made up of synthetic polymers, with several fasci-
nating properties that make them essential for domestic and industrial applications.
Diverse groups of polymers contribute to plastics manufacturing, making them cheap,
lighter, durable, resistant to corrosion, and with high insulating (thermal and elec-
trical) properties [1]. These properties have made plastics a cost-effective option
for manufacturing a variety of products [2]. There are several hundreds of plastic
materials that are commercially available but only a few of them qualify as ther-
moplastics for commodity level, mainly attributed to their high volume and cheaper
price. The commodity plastics are mainly comprised of low-density polyethylene
(LDPE), high-density polyethylene (HDPE), polyvinylchloride (PVC), polystyrene
(PS), polypropylene (PP), and polyethylene terephthalate (PET). These plastics
contribute to 90% of global plastic production.
Plastics due to their sterility and lightweight are the foremost choice for the
production of a maximum of hospital-related equipment including personal protec-
tive equipment (PPEs). The COVID-19 pandemic situation witnessed a steep rise in
the demand for PPEs and thus that of plastics. The large-scale use and disposal of
PPEs and other COVID-19 wastes has created real havoc in the environment [3], thus
piling up a huge number of plastic wastes. The highly recalcitrant nature of plastics
has made it difficult to manage the waste load as they persist in the environment for
several years [4], thus leading to environmental pollution [5]. The slow degradation
of plastics in the environment under the influence of abiotic factors like temperature,
pressure, light, moisture, and UV further compounds the problem, thereby endan-
gering flora and marine fauna along with raising risks to human health. Conventional
methods of plastic waste management fail to provide a sustainable solution and thus
need further attention and research in this regard. Valorization of plastic wastes
into bioenergy is a sustainable way to handle the waste load in the environment.
Conversions of plastic waste into different forms of chemicals or fuels can be a good
option.
7 Sustainable Solutions for Bioenergy Production from Hospital-Based … 177

Conversion of plastic wastes into useful energy forms can be possible by utilizing
several techniques [6]. Heat treatment and oxidation technologies have been found
useful in the conversion of plastic wastes into fuels. Plastic waste can be easily
converted into various energy-usable forms like liquid fuels [7] and combustible
gases. Organic by-products like carboxylic and sulfonic acids can also be used as a
source of carbon in the chemical industries [8].
Pyrolysis is an advanced technique that seems promising as far as sustainable
energy production is concerned. The chapter discusses the importance of plastics
in the healthcare sector and the burden created due to the disposal of such plastics
on the environment. The conventional plastic waste management methods are also
discussed along with their various shortcomings. Pyrolysis of plastic wastes into
energy/fuel sources has been emphasized.

7.2 The Intangible Relationship Between the Healthcare


Sector and Plastics

The human race has been using medical devices for thousands of years. Metal was
the primary material used to create medical equipment during the early periods of
history. Ceramics began to be employed in the late 1920s, and subsequently, glass
entered the picture. However, some of the problems with these materials paved the
door for the use of plastics in medical care. Despite the fact that scientists have been
experimenting with plastics since the early nineteenth century, the medical business
was not completely transformed by plastics until the middle of the twentieth century.
According to a report by Market and Markets, the global market for medical plastics
was valued at roughly USD 22.8 billion in 2019 and is projected to reach USD 31.7
billion, at a CAGR of 6.8% by 2024 (Market and Markets 2021). Plastics are utilized
in many different applications, such as MRI casings, surgical equipment that replace
ceramic and other metals, bedpans, and implants (Table 7.1).
Numerous medical manufacturers treat plastic with antimicrobial coatings that
can resist or kill bacteria, lowering the risk of infection and preventing cross-
contamination. The following is a list of some of the polymers used in medical
applications: polyethylene (PE), polyvinylchloride (PVC), polycarbonate (PC),
polypropylene (PP), polystyrene (PS), polyethylene terephthalate (PET), polyamide
(PA), acrylonitrile butadiene styrene (ABS), and polyurethane (PU).
In the modern world, plastics and synthetic polymers have assimilated into every
aspect of our existence. Every type of plastic that has been manufactured up to this
point is still in use, posing a continuing threat to the environment [4]. The COVID-19
pandemic has exacerbated the problems associated with managing plastic garbage,
which were already a problem [10]. There is a consistent rise in COVID-19 cases,
particularly with the emergence of the second wave in India. This has highlighted
the pressing requirement for significant amounts of personal protective equipment
(PPE).
178 P. Dash et al.

Table 7.1 Plastics used in medical applications


Property Applications in Types of plastics Medical device
medical devices applications
(percentage usage)
Commodity plastics 70% of all plastics Polyethylene Tubing
Polypropylene Films, Packaging
Polystyrene Connectors
Polyvinylchloride Labware
IV bags
Catheters
Face masks
Drug-delivery
components
Housings
Luers
Membranes
Sutures
Engineering 20% of all plastics Polyamides, Nylons Surgical instruments,
thermoplastics Polyesters Balloons
Polycarbonates Blood set components,
Polyurethanes Blood bowls
Acrylics Blood oxygenators,
Acetals Syringes
Moving parts and
components
Luers
Catheters
High-temperature 10% of all plastics Polyimides Surgical instruments,
engineering Polyetherimides Surgical trays
thermoplastics and Polysulfones Syringes
other polymers Polyether ether ketone Implants
Polyphenylene sulfide Dentals implants
Fluoropolymers Bone implants
Liquid crystalline Moving parts and
polymers components
Biopolymers High precision parts
Thermosets and Electronic components
adhesives Luers
Bioresorbable sutures
Source Sastri [9]

Despite having a significant amount of plastic in PPE, they appear to be the most
reliable and cost-effective way to prevent virus transmission. PPE has become essen-
tial for healthcare and frontline workers due to the nature of virus transmission from
person to person, which is accomplished by aerosols. More asymptomatic instances
have led to a surge in demand for PPE kits, particularly single-use disposable face
masks, face shields, and gloves, without adequate attention to proper disposal mech-
anisms. Consequently, there has been a sharp increase in global manufacturing and
production of these items [11, 12].
7 Sustainable Solutions for Bioenergy Production from Hospital-Based … 179

7.3 The Burden of Hospital-Generated Plastic Waste

Plastics due to their high durability and sterility have been the most preferred substrate
for use in the field of healthcare. With the onset of COVID-19, plastics have become
more important as a day-to-day material in hospitals. The frequent and large-scale
use of face masks during the COVID regime has been found to largely contribute to
global plastic pollution. The problem is even aggravated due to the improper disposal
of face masks by human communities. According to Shiferie [13], improper disposal
of 1% of face masks will approximately lead to 10 million masks accumulating in
the environment on a monthly basis, thus leading to pollution. These 10 million
masks roughly account for 40,000 kg of plastic waste in the environment assuming
the weight of a single face mask to be 4 g [14].
The healthcare sector is one of the largest consumers of durable plastics and thus
produces significantly larger amounts of plastic waste. To make it more worrisome,
the onset of COVID-19 even has compounded the problem with the generation of
huge volumes of plastic waste. Global plastic waste generation was estimated to
be 380 million tons in 2018 [15]. Considering the rate of plastic consumption by
the biomedical sector during the pandemic, the generation of waste thereof was
estimated to double in 2020 [16]. Figure 7.1 provides a detailed overview of plastic
waste generation from 2010 to 2018 and estimated values for 2019–20 and 2050.
India itself accounted for 20,400 kg of plastic waste generation from its hospitals
in 2005 which further increased to 198,766 kg in 2019, an increase of approximately
89% (Fig. 7.2). The onset of COVID-19 even worsened the problem.

Fig. 7.1 Global plastic waste generation from 2010 to 2018, and estimation for 2019–20, and 2050.
Source Khoo et al. [17]
180 P. Dash et al.

Fig. 7.2 Plastic waste generation from hospitals in India (2005–2019). Source Rai et al. [18]

Wuhan in China, the epicenter of COVID-19 witnessed the generation of high


amounts of PPE plastics on a daily basis which was almost six times higher than that
of the pre-pandemic period [11]. However, recent estimates suggest the generation
of 685 kg of medical waste generated per day for every 100 confirmed COVID-19
cases [16, 19, 20]. Peng et al., estimated that two-thirds of the global plastic waste
was from hospitals with Asian countries accounting for 72% of the global plastic
waste burden [21]. Moreover, on a global scale, plastic waste disposed of by 193
countries in relation to COVID-19 was also found to breach the 8.4 million tons
mark by August 23, 2021. The hospital-generated plastic waste during COVID-19
was mostly from the protective gear being used by the frontline healthcare workers.

7.4 Conventional Methods of Plastic Waste Management

Hospital-generated plastics need proper segregation and treatment prior to disposal.


However, humans out of their unawareness or negligence generally tend to mix these
wastes with other garbage items, thus making the task of disposal quite difficult.
Plastics in general play a very important role in the healthcare sector, and replacing
them seems impossible as of now. However, the management of plastic waste must
be taken up with the utmost necessity to maintain a sustainable environment.
The three conventional methods for the management of plastic waste include
incineration, landfilling, and recycling. The process of recycling can be divided into
four types (primary, secondary, tertiary, and quaternary) and thus lead to different
products (Fig. 7.3).
7 Sustainable Solutions for Bioenergy Production from Hospital-Based … 181

Fig. 7.3 Different recycling processes and the nature of their corresponding product obtained.
Source Sastri [9]

However, plastic waste from the healthcare sector is mostly contaminated and
requires prior proper sterilization. The purity of plastics is also a major concern
and one of the bottlenecks in the recycling process. Plastics composed of mixed
polymeric structures or certain other additives are difficult to recycle [10]. Moreover,
mixed polymeric plastics have a very low recyclable value due to their low-quality
post recycling [22]. Several limitations of the recycling processes leave out only the
options of incineration and landfilling.
A thermal treatment process such as incineration is another method that utilizes
high temperatures to decontaminate as well as reduce the volume of plastic waste
generated by the healthcare sector. The process results in the complete combustion
of the plastics resulting in the formation of water and carbon dioxide (CO2 ) [23].
Moreover, this process also helps in the conversion of these wastes into energy
particularly in the form of heat [24]. The incineration process leads to the release of
a high amount of CO2 , thus leading to atmospheric pollution and global warming
[25]. Not only CO2 but incineration of plastic wastes is also involved in the release
of toxic pollutants into the atmosphere such as particulate matter (PM), volatile
organic compounds (VOC), carbon monoxide (CO), and nitrogen oxides (NOx). The
incineration processes no doubt reduce the load of plastic waste to a much greater
182 P. Dash et al.

extent but simultaneously also contribute toward environmental pollution [26], thus
raising the risk level on public health [2].
Landfilling is the most common method followed for the disposal of plastic waste
generated from hospitals and healthcare centers. It is widely followed in most under-
developed and developing nations due to its low cost of operation. The onset of
COVID-19 saw a large-scale generation of plastic waste from the healthcare sector
which was mostly disposed of in landfills due to a lack of proper incineration facil-
ities with high capacities. Plastic wastes at landfill sites tend to degrade naturally
under the influence of environmental temperature, atmospheric pressure, moisture,
and sunlight but in a very slow manner. However, the rate at which plastic wastes are
produced in the environment superexceeds the rate of their degradation, thus leading
to large-scale accumulation. Slow degradation of plastic wastes from landfill sites
also contributes to the global greenhouse gas pool in the form of CO2 , CO, and others
[27]. The long persistence of these wastes also results in these toxic gases remaining
in the surrounding air for a quite long period [28].

7.5 Impact of Plastic Waste on the Environment

Plastic waste being persistent in the environment for a longer period of time poses
several environmental concerns. Plastics degrade very slowly under the influence of
factors like temperature, pressure, light, reaction with chemicals, and some microbial
actions [20]. Plastics being a complex mixture of several chemical compounds mainly
contribute to chemical pollution of the environment. Chemicals from plastics due
to weathering can leach into the soil, thereby affecting the soil profile as well as
groundwater. Certain volatile chemicals present in plastic wastes pose the chances
of being released into the atmosphere, thus reducing the air quality. Plastics because
of their intrinsic properties like high surface area and presence of certain functional
groups can also adsorb certain toxic pollutants from the environment, thus increasing
the environmental toxicity load [29]. Chemical pollution due to plastic waste is more
pronounced in aquatic bodies as compared to terrestrial ecosystems. This may be
attributed to the lower temperature of the water as compared to the land system. A
lower temperature ensures slow degradation of the waste, thus making it persistent
for a longer period [30]. Plastics can further undergo degradation to form micro and
nano plastics which adsorb other chemical pollutants from the environment [31].
Management of plastic waste can also lead to deterioration of the air quality, thus
causing severe pollution of the atmosphere. The incineration of plastic wastes leads
to the generation of several toxic pollutants including greenhouse gases (GHGs)
and particulate matter (PM). Besides this, particular waste management process also
releases several volatile toxins into the atmosphere [12]. Incineration of PPE wastes
during COVID-19 has been found to generate particulate matter in the dimensions
of 10 µm (PM10 ) and 2.5 µm (PM2.5 ).
Plastic wastes are also a major source of marine pollutants with recent data
suggesting approximately 25.9 kilotons (kt) of these being released into the marine
7 Sustainable Solutions for Bioenergy Production from Hospital-Based … 183

environment during the pandemic [21]. The initial phases of COVID-19 during 2020
witnessed the disposal of 1.56 billion face masks in the environment which roughly
weighed around 5.66 million tons (Mt) and were disposed of offshore [32]. Adsorp-
tion of various heavy metals onto the disposed PPEs in water bodies can create a
disturbance in aquatic environmental health [33]. Plastic particles generated from
disposable hospital-generated plastic wastes can also be consumed by marine organ-
isms, thus leading to higher mortality rates [34]. Plastic particles are known to exert
ecotoxic, cytotoxic, and genotoxic effects on marine organisms. Plastic wastes result
in the generation of reactive oxygen species (ROS), thus creating oxidative stress
and disturbance in the DNA repair mechanism [35]. Plastic particles in the water
also prevent the interaction of oxygen with the aquatic bodies, thus creating an
oxygen-depleted environment and subsequent death of the aquatic flora and fauna.
Plastic pollution also leads to microbial pollution which may be attributed to the
environmental interaction between the microbes and the plastics. Plastics provide
a solid surface for the formation of microbial biofilms and can facilitate horizontal
gene transfer among microbial species, thus leading to increased pathogenicity [36].

7.6 Valorization of Plastic Wastes into Bioenergy

The approach of using landfills appeared inadequate for managing the accumulation
of contagious COVID-19-related medical waste. Budiman and Ardiansyah, stated
that in May 2020, the stacking of COVID-19 medical garbage collapsed the land-
fill’s wall, letting tons of waste enter the Cisadane River [37]. This improper waste
handling led to contamination of the main river, affecting the public water source.
According to Corburn et al. [38], the improper disposal of plastic and medical waste
through landfills and local burning exacerbated waste problems in Indian munici-
palities during the pandemic and contributed to the spread of the virus. As a result,
there exists an issue in managing unconventional waste sustainably while reducing
carbon emissions, minimizing secondary virus transmission, and mitigating poten-
tial health risks. Moreover, in the absence of conventional waste management and
emergency strategies to address the epidemic, the world could face devastating conse-
quences. Given the adverse environmental effects of plastic waste and its disposal
methods, innovative environmentally friendly approaches for plastic waste treatment
need to be developed [39]. The dwindling energy scenario which is responsible for
the creation of an energy crisis on a global basis needs to be countered by devising
alternative sources of energy generation [40–43]. In this regard, the valorization
of plastic medical waste will not only solve the environmental problem but also
suffice for the ever-increasing energy needs. Pyrolysis is one of the most advanced
technologies currently being researched and poses the ability to sustainably convert
medical-based plastic wastes into suitable bioenergy forms. The current section high-
lights the role of pyrolysis in the conversion of these plastic wastes into bioenergy
and also discusses several advancements in the field.
184 P. Dash et al.

7.6.1 Pyrolysis

The pyrolysis approach piqued the researchers’ interest because it offers high waste
decomposition, a simple process, and promotes green waste management over current
incineration practices. Pyrolysis involves a thermochemical transformation of plastic
waste at high temperatures in an inert deoxygenated environment, resulting in the
conversion of the waste into lower molecular weight compounds [44].
In a prior study by Xue et al. [45], pyrolysis was recognized as a viable and efficient
technique for managing industrial and medical solid waste. It serves as an alternative
method for thermal disinfection of COVID-19-related waste. The heat and pressure
used in waste pyrolysis result in combustible lower molecular weight molecules,
primarily gases like hydrogen, methane, and carbon monoxide [46]. Liquids are
predominantly formed at high operating temperatures, including methanol, acetone,
acetic acid, acetaldehyde, tar, solvent oil, and other organic materials [44]. Addition-
ally, low-temperature pyrolysis yields solid end-products like coke, char, and carbon
black (Fig. 7.4). The collected end-products are repurposed, thus reducing pollution.
This approach is highlighted as a great solution for transforming plastic waste into
usable by-products [47]. Vivero et al. [48] emphasized that pyrolysis is optimal for
converting plastic waste into hydrocarbon blends, aiding in recycling and reducing
PPE plastic trash.
It aids in the reduction of plastic waste, the recovery of chemicals, and the substi-
tution of fuels and other virgin materials. Table 7.2 compares the solid yields of
pyrolysis of polypropylene, including char, coke, and waxy residue. According to
the published study, polypropylene plastic waste showed slow pyrolysis with a 5 °C/
min heating rate, resulting in over 2.67% char product yield [49]. Ahmad et al. [50]
studied pyrolysis temperature effects on PP waste, ranging from 250 to 400 °C,
and achieved the highest conversion at 250 °C. In terms of char yields, the yield of
solid residue dropped with increasing temperature, from 13.68 to 5.7%. Witkowski
et al. [51] suggested that this decrease in char output at higher temperatures is linked
to the primary depolymerization of plastics and the secondary depolymerization of
produced char.

7.6.1.1 Cold Plasma Pyrolysis

Recent research has explored plasma technology’s application in pyrolysis systems.


Cold plasma pyrolysis presents an effective approach to managing infectious medical
plastic waste, aided by high pyrolytic temperatures and a sophisticated cooling
system. This method utilizes excited hot electrons generated from electrodes to effi-
ciently break down plastic chemical bonds under controlled conditions and in less
time than traditional pyrolysis [58]. However, this approach is energy-intensive due
to the advanced cooling system and higher process temperatures required to ther-
mally break down plastic waste into lighter hydrocarbons. Globally, it is crucial to
7 Sustainable Solutions for Bioenergy Production from Hospital-Based … 185

Fig. 7.4 Steps in pyrolytic conversion of plastic wastes into valuable energy forms. (Step-1:
Collected plastic wastes are fed into a hopper which is then passed onto an extruder. Step-2: The
plastic wastes are converted into smaller fragments or particles in the extruder and pushed forward
into the pyrolysis reactor. Step-3: The pyrolysis reactor degrades the plastic wastes under high
temperatures and in the absence of oxygen. The vapors from the pyrolysis reactor move forward
to the condenser. Step-4: The condenser primarily produces gases like H2 , CH4 , and CO. Step-6:
Moreover, condensation also produces a mixture of liquid fuels if the pyrolysis process involves
very high temperatures. Step-7: While, a low-temperature pyrolysis yields coke, char, and carbon
black that can be used as a solid fuel. Step-8: The final step involves the use of a separation chamber
or separator that separates different liquid fuels along with the release of any gases)

Table 7.2 Pyrolysis of polypropylene under different conditions and respective char yield
Reactor Temperature (°C) Other conditions Char yield (wt%) References
Batch reactor 380 Pressure—1 atm 13.30 [52]
Heating rate—3 °C/
min
Batch reactor 740 Heating rate—10 °C/ 1.60 [53]
min
Fluidized bed 703 Feed rate—7 g/min 6.90 [54]
reactor
Semi-batch reactor 450 Pressure—1 atm 3.60 [55]
Heating rate—25 °C/
min
Fixed bed and batch 400 Heating rate—25 °C/ 2.50 [49]
reactor min
Residence
time—45 min
Fixed bed reactor 900 Residence 2.80 [56]
time—21 min
Microwave-assisted 250 HZSM-5 catalyst 1.54 [57]
semi-batch reactor
186 P. Dash et al.

adopt economically simple processes and energy-saving technologies for handling


extensive plastic waste production.

7.6.1.2 Microwave-Assisted Pyrolysis

This process involves the thermal decomposition of plastic wastes at higher temper-
atures of 573–1073 K with the assistance of microwave irradiation, thus producing
biogas and bio-oils [56, 57]. Microwave irradiation-based pyrolysis is characterized
by faster heating rates, low operating temperatures, and a very short reaction time
[59]. Materials like activated carbon, graphene oxide, and silicon dioxide are used as
absorbents during microwave irradiation to improve the heating rate, thus achieving
high temperatures within a very short time as compared to conventional techniques
[60]. Co-pyrolysis of waste plastics and waste cooking oil at 500 °C carried out
through microwave irradiation has been found to produce a high yield of liquid fuel
within 10 min. The liquid fuel thus obtained demonstrated characteristic similarity
with that of diesel used in transportation. The fuel besides exhibiting higher stability
also remained free from nitrogen and sulfur and had a very low oxygen content.
Moreover, the liquid fuel had much high energy content of 42–46 MJ/kg [61].

7.6.1.3 Plasma Pyrolysis

The plasma pyrolysis or plasma gasification technique for managing plastic wastes
is mostly being followed in the countries of the European Union [62]. In this process
(Fig. 7.5), gasification of the plastic wastes is mostly carried out by thermal plasma-
based heating generated by DC current-operated arc plasma torches. This operates
at a very high reaction temperature of 2000–14,000 °C and a short residence time
of less than 30 min. Any carbonaceous materials including plastics when fed into
the plasma reactor are heated by the arcs, thus liberating volatile matter which when
cracked leads to the formation of syngas and methane [63].
A recent study explored a novel hydrothermal processing technology, which effec-
tively addresses medical plastic waste using a high-pressure system. This method
employs a hot compressed aqueous medium at elevated pressures to yield monomers,
chemicals, hydrochar, and crude oil, which can be used in plastic production and other
applications [64]. Moreover, hydrothermal processing contributes to lower oxygen
content, improved energy content in liquid crude oil, cost-effective processing,
and efficient storage. This alternative approach holds promise for reducing plastic
pollution and addressing environmental concerns.
7 Sustainable Solutions for Bioenergy Production from Hospital-Based … 187

Fig. 7.5 Plasma pyrolysis of


plastic waste leads to the
generation of syngas and
methane

7.7 Challenges and Future Prospects

Taking into consideration the vast amount of plastic waste generated by the hospital
and healthcare sector and the pollution caused by them, an advanced technique like
pyrolysis offers a sustainable alternative to the problem. Pyrolysis not only converts
these plastic wastes into energy-usable and other valuable forms, but also reduces
the risk of pollution as observed in the case of conventional treatment methods.
Pyrolysis of plastic wastes leads to the formation of liquid, gaseous, and solid fuels
of high importance and energy yield. However, valorization of the hospital-based
plastic waste mostly comes with several challenges.
The selection of feedstock is a major challenge in the pyrolysis process. Heteroge-
neous plastic materials create a problem in the pyrolysis process. Thermal degrada-
tion of materials made of PET and PVC hinders the conversion process. Chlorinated
hydrocarbons result due to thermal degradation of PVC along with the formation of
HCl. These by-products in turn corrode the reactor and make the oil halogenated.
Phthalic acid is formed as a result of PET degradation leading to the clogging of the
reactor pipes and deterioration in the oil quality [65].
Plastic waste undergoing pyrolysis needs prior pretreatment to remove other impu-
rities like glass, metals, and wood. Moreover, plastic waste generated from hospital
settings needs to be pretreated to remove any microbial contamination. Hospital-
generated plastic wastes pose the threat of infecting persons employed in the collec-
tion, sorting, and processing of the wastes and must be taken utmost care. Plastic
wastes obtained from hospital settings also vary in size as per their utilities. Pretreat-
ment is also required to bring them to a uniform size, thus adding extra cost to the
entire process.
188 P. Dash et al.

Another major problem is the formation of wax as a final product in the pyrolysis
process. Reactors with inefficient recovery systems can lead to the scaling of wax on
the inner surface of the condensing systems, thus making the recovery process too
tedious [66].
Pyrolysis liquids are quite unstable and often tend to repolymerize, thus making
their storage difficult for a longer time period. Post-treatment techniques like blending
and dewaxing are necessary to maintain the quality of the liquid fuels for a longer
period [67, 68].
The conversion of hospital-based plastic wastes into useful bioenergy forms
provides a lot of research options for the overall advancement of the technology
as well as to reduce its operational costs. Further, research in this regard is highly
essential for the sustainable reduction of these wastes in the environment in a more
efficient manner.

7.8 Conclusions

Plastics being recalcitrant in nature are a hazard to the environment as well as their
biotic components. However, looking at their usefulness, social, and economic rele-
vance, it is quite impossible to limit or discontinue their uses as of now. Improper
disposal and management of plastic load in the environment can be really troublesome
and pose several health hazards to the living components including humans. Proper
management of plastic waste by means of converting it into sustainable energy forms
can be a win–win situation. Pyrolysis of plastic wastes into energy-usable forms can
not only contemplate the energy crisis situation but also lead to the formation of
a sustainable environment. At present, the management of plastic waste involves
a lot of expenditures. These expenditures could be reversed by the conversion of
these wastes into usable forms such as bioenergy. The pyrolysis technique despite
its several advantages also faces numerous challenges which include the selection
of feedstocks, pretreatment processes involved, and stability concerns. As such, it
opens up wide avenues for research in the field with the sole aim of making the
technique more sustainable and cost-effective.

References

1. Kumar, B., Pundir, A., Mehta, V., Singh, B., & Solanki, R. (2020). A review paper on plastic,
its variety, current scenario and its waste management. Plant Arch, 20, 53–56.
2. Er, C. T., Sen, L. Z., Srinophakun, P., & Wei, O. C. (2023). Recent advances and chal-
lenges in sustainable management of plastic waste using biodegradation approach. Bioresource
Technology, 128772. https://doi.org/10.1016/j.biortech.2023.128772
3. Zarocostas, J. (2022). WHO concerned over COVID-19 health-care waste. The Lancet,
399(10324), 507. https://doi.org/10.1016/S0140-6736(22)00225-2
7 Sustainable Solutions for Bioenergy Production from Hospital-Based … 189

4. Gewert, B., Plassmann, M. M., & MacLeod, M. (2015). Pathways for degradation of plastic
polymers floating in the marine environment. Environmental Science: Processes & Impacts,
17(9), 1513–1521. https://doi.org/10.1039/C5EM00207A
5. Lee, J., Park, H. J., Moon, M., Lee, J. S., & Min, K. (2021). Recent progress and challenges
in microbial polyhydroxybutyrate (PHB) production from CO2 as a sustainable feedstock:
A state-of-the-art review. Bioresource Technology, 339, 125616. https://doi.org/10.1016/j.bio
rtech.2021.125616
6. Chen, X., Wang, Y., & Zhang, L. (2021). Recent progress in the chemical upcycling of plastic
wastes. Chemsuschem, 14(19), 4137–4151.
7. Papari, S., Bamdad, H., & Berruti, F. (2021). Pyrolytic conversion of plastic waste to value-
added products and fuels: A review. Materials, 14(10), 2586.
8. Li, N., Liu, H., Cheng, Z., Yan, B., Chen, G., & Wang, S. (2022). Conversion of plastic waste
into fuels: A critical review. Journal of Hazardous Materials, 15(424), 127460. https://doi.org/
10.1016/j.jhazmat.2021.127460
9. Sastri, V. R. (2021). Plastics in medical devices: Properties, requirements, and applications.
William Andrew.
10. Inamdar, I. (2022). Recycling of plastic wastes generated from COVID-19: A comprehen-
sive illustration of type and properties of plastics with remedial options. Science of The Total
Environment, 155895. https://doi.org/10.1016/j.scitotenv.2022.155895
11. Adyel, T. M. (2020). Accumulation of plastic waste during COVID-19. Science, 369(6509),
1314–1315. https://doi.org/10.1126/science.abd9925
12. Parashar, N., & Hait, S. (2021). Plastics in the time of COVID-19 pandemic: Protector or
polluter? Science of the Total Environment, 759, 144274. https://doi.org/10.1016/j.scitotenv.
2020.144274
13. Shiferie, F. (2021). Improper disposal of face masks during COVID-19: Unheeded public
health threat. The Pan African Medical Journal, 38. https://doi.org/10.11604/pamj.2021.38.
366.29063
14. Patrício Silva, A. L., Prata, J. C., Walker, T. R., Duarte, A. C., Ouyang, W., Barcelo, D., &
Rocha-Santos, T. (2021). Increased plastic pollution due to COVID-19 pandemic: Challenges
and recommendations. Chemical Engineering Journal, 405, 126683. https://doi.org/10.1016/
j.cej.2020.126683
15. Shams, M., Alam, I., & Mahbub, M. S. (2021). Plastic pollution during COVID-19: Plastic
waste directives and its long-term impact on the environment. Environmental advances, 5,
100119. https://doi.org/10.1016/j.envadv.2021.100119
16. Khoo, K. S., Ho, L. Y., Lim, H. R., Leong, H. Y., & Chew, K. W. (2021). Plastic waste associated
with the COVID-19 pandemic: Crisis or opportunity? Journal of hazardous materials, 417,
126108. https://doi.org/10.1016/j.jhazmat.2021.126108
17. Rai, P. K., Sonne, C., Song, H., & Kim, K. H. (2023). Plastic wastes in the time of COVID-19:
Their environmental hazards and implications for sustainable energy resilience and circular
bio-economies. Science of The Total Environment, 858, 159880. https://doi.org/10.1016/j.sci
totenv.2022.159880
18. Achra, A., Mahajan, R. K., & Sahoo, S. (2021). The changing pattern of the quantum of biomed-
ical waste generated from a tertiary care hospital in Delhi. Journal of Laboratory Physicians,
13(01), 080–083. https://doi.org/10.1055/s-0041-1723056
19. Patricio Silva, A. L., Prata, J. C., Duarte, A. C., Soares, A. M., Barceló, D., & Rocha-Santos, T.
(2021). Microplastics in landfill leachates: The need for reconnaissance studies and remediation
technologies. Case Studies in Chemical and Environmental Engineering, 3, 100072.
20. Patricio Silva, A. L., Prata, J. C., Duarte, A. C., Barcelò, D., & Rocha-Santos, T. (2021). An
urgent call to think globally and act locally on landfill disposable plastics under and after covid-
19 pandemic: Pollution prevention and technological (Bio) remediation solutions. Chemical
Engineering Journal, 426, 131201. https://doi.org/10.1016/j.cej.2021.131201
21. Peng, Y., Wu, P., Schartup, A. T., & Zhang, Y. (2021). Plastic waste release caused by COVID-
19 and its fate in the global ocean. Proceedings of the National Academy of Sciences, 118(47),
e2111530118. https://doi.org/10.1073/pnas.2111530118
190 P. Dash et al.

22. Lange, J. P. (2021). Managing plastic waste\\sorting, recycling, disposal, and product redesign.
ACS Sustainable Chemistry & Engineering, 9(47), 15722–15738. https://doi.org/10.1021/acs
suschemeng.1c05013
23. Evode, N., Qamar, S. A., Bilal, M., Barceló, D., & Iqbal, H. M. (2021). Plastic waste and
its management strategies for environmental sustainability. Case Studies in Chemical and
Environmental Engineering, 4, 100142. https://doi.org/10.1016/j.cscee.2021.100142
24. Klemeš, J. J., Van Fan, Y., Tan, R. R., & Jiang, P. (2020). Minimising the present and future
plastic waste, energy and environmental footprints related to COVID-19. Renewable and
Sustainable Energy Reviews, 127, 109883. https://doi.org/10.1016/j.rser.2020.109883
25. Devasahayam, S., Raju, G. B., & Hussain, C. M. (2019). Utilization and recycling of end
of life plastics for sustainable and clean industrial processes including the iron and steel
industry. Materials Science for Energy Technologies, 2(3), 634–646. https://doi.org/10.1016/j.
mset.2019.08.002
26. Jeswani, H., Krüger, C., Russ, M., Horlacher, M., Antony, F., Hann, S., & Azapagic, A. (2021).
Life cycle environmental impacts of chemical recycling via pyrolysis of mixed plastic waste in
comparison with mechanical recycling and energy recovery. Science of the Total Environment,
769, 144483. https://doi.org/10.1016/j.scitotenv.2020.144483
27. Royer, S. J., Ferrón, S., Wilson, S. T., & Karl, D. M. (2018). Production of methane and ethylene
from plastic in the environment. PLoS ONE, 13(8), e0200574. https://doi.org/10.1371/journal.
pone.0200574
28. Prata, J. C., Silva, A. L., Walker, T. R., Duarte, A. C., & Rocha-Santos, T. (2020). COVID-
19 pandemic repercussions on the use and management of plastics. Environmental Science &
Technology, 54(13), 7760–7765. https://doi.org/10.1021/acs.est.0c02178
29. Kumar, H., Azad, A., Gupta, A., Sharma, J., Bherwani, H., Labhsetwar, N. K., & Kumar, R.
(2021). COVID-19 creating another problem? Sustainable solution for PPE disposal through
LCA approach. Environment, Development and Sustainability, 23, 9418–9432. https://doi.org/
10.1007/s10668-020-01033-0
30. Duan, J., Bolan, N., Li, Y., Ding, S., Atugoda, T., Vithanage, M., Sarkar, B., Tsang, D. C., &
Kirkham, M. B. (2021). Weathering of microplastics and interaction with other coexisting
constituents in terrestrial and aquatic environments. Water Research, 196, 117011. https://doi.
org/10.1016/j.watres.2021.117011
31. Kutralam-Muniasamy, G., Pérez-Guevara, F., & Shruti, V. C. (2021). A critical synthesis of
current peer-reviewed literature on the environmental and human health impacts of COVID-19
PPE litter: New findings and next steps. Journal of Hazardous Materials, 422, 126945. https://
doi.org/10.1016/j.jhazmat.2021.126945
32. Yuan, X., Wang, X., Sarkar, B., & Ok, Y. S. (2021). The COVID-19 pandemic necessitates
a shift to a plastic circular economy. Nature Reviews Earth & Environment, 2(10), 659–660.
https://doi.org/10.1038/s43017-021-00223-2
33. De-la-Torre, G. E., Dioses-Salinas, D. C., Pizarro-Ortega, C. I., Severini, M. D., López, A. D.,
Mansilla, R., Ayala, F., Castillo, L. M., Castillo-Paico, E., Torres, D. A., & Mendoza-Castilla,
L. M. (2022). Binational survey of personal protective equipment (PPE) pollution driven by
the COVID-19 pandemic in coastal environments: Abundance, distribution, and analytical
characterization. Journal of Hazardous Materials, 426, 128070. https://doi.org/10.1016/j.jha
zmat.2021.128070
34. Ray, S. S., Lee, H. K., Huyen, D. T., Chen, S. S., & Kwon, Y. N. (2022). Microplastics waste
in environment: A perspective on recycling issues from PPE kits and face masks during the
COVID-19 pandemic. Environmental Technology & Innovation, 26, 102290. https://doi.org/
10.1016/j.eti.2022.102290
35. Tagorti, G., & Kaya, B. (2022). Genotoxic effect of microplastics and COVID-19: The hidden
threat. Chemosphere, 286, 131898. https://doi.org/10.1016/j.chemosphere.2021.131898
36. Rogers, K. L., Carreres-Calabuig, J. A., Gorokhova, E., & Posth, N. R. (2020). Micro-by-micro
interactions: How microorganisms influence the fate of marine microplastics. Limnology and
Oceanography Letters, 5(1), 18–36. https://doi.org/10.1002/lol2.10136
7 Sustainable Solutions for Bioenergy Production from Hospital-Based … 191

37. Budiman, Y. C., & Ardiansyah, T. (2020). In Indonesia, coronavirus floods Cisadane River with
extra hazard: Medical waste. TheJakartaPost. https://www.thejakartapost.com/news/2020/09/
01/in-indonesia-coronavirus-floods-cisadane-river-with-extra-hazardmedical-waste.html
38. Corburn, J., Vlahov, D., Mberu, B., Riley, L., Caiaffa, W. T., Rashid, S. F., Ko, A., Patel, S.,
Jukur, S., Martínez-Herrera, E., & Jayasinghe, S. (2020). Slum health: Arresting COVID-19 and
improving well-being in urban informal settlements. Journal of Urban Health, 97, 348–357.
https://doi.org/10.1007/s11524-020-00438-6
39. Harussani, M. M., Sapuan, S. M., Rashid, U., Khalina, A., & Ilyas, R. A. (2022). Pyrolysis
of polypropylene plastic waste into carbonaceous char: Priority of plastic waste management
amidst COVID-19 pandemic. Science of The Total Environment, 803, 149911. https://doi.org/
10.1016/j.scitotenv.2021.149911
40. Das, P. K., Das, B. P., Dash, P., & Gurunathan, B. (2022). Production of biofuel from genetically
modified microalgal biomass and its effects on environment and public health. In Biofuels and
bioenergy (pp. 505–519). Elsevier. https://doi.org/10.1016/B978-0-323-85269-2.00004-6
41. Das, P. K., Das, B. P., & Dash, P. (2021). Application of nanotechnology in the production of
bioenergy from algal biomass: Opportunities and challenges. Nanomaterials, 355–377. https://
doi.org/10.1016/B978-0-12-822401-4.00024-6
42. Das, P. K., Das, B. P., & Dash, P. (2020). Potentials of postharvest rice crop residues as a source
of biofuel. In Refining biomass residues for sustainable energy and bioproducts (pp. 275–301).
Academic Press. https://doi.org/10.1016/B978-0-12-818996-2.00013-2
43. Das, P. K., Das, B. P., & Dash, P. (2019). Role of energy crops to meet the rural energy needs:
An overview. Biomass Valorization to Bioenergy, 12, 11–30. https://doi.org/10.1007/978-981-
15-0410-5_2
44. Verma, A., Budiyal, L., Sanjay, M. R., & Siengchin, S. (2019). Processing and characteri-
zation analysis of pyrolyzed oil rubber (from waste tires)-epoxy polymer blend composite
for lightweight structures and coatings applications. Polymer Engineering & Science, 59(10),
2041–2051. https://doi.org/10.1002/pen.25204
45. Xue, Y., Kelkar, A., & Bai, X. (2016). Catalytic co-pyrolysis of biomass and polyethylene in
a tandem micropyrolyzer. Fuel, 166, 227–236. https://doi.org/10.1016/j.fuel.2015.10.125
46. Kairytė, A., Kremensas, A., Vaitkus, S., Członka, S., & Strąkowska, A. (2020). Fire suppression
and thermal behavior of biobased rigid polyurethane foam filled with biomass incineration
waste ash. Polymers, 12(3), 683. https://doi.org/10.3390/polym12030683
47. Worrell, W. A., & Vesilind, P. A. (2012). Integrated solid waste management. In Solid waste
engineering (2nd ed.). Cengage Learning.
48. Vivero, L., Barriocanal, C., Alvarez, R., & Diez, M. A. (2005). Effects of plastic wastes on
coal pyrolysis behaviour and the structure of semicokes. Journal of Analytical and Applied
Pyrolysis, 74(1–2), 327–336. https://doi.org/10.1016/j.jaap.2004.08.006
49. Sogancioglu, M., Yel, E., & Ahmetli, G. (2020). Behaviour of waste polypropylene pyrolysis
char-based epoxy composite materials. Environmental Science and Pollution Research, 27,
3871–3884. https://doi.org/10.1007/s11356-019-07028-3
50. Ahmad, I., Khan, M. I., Khan, H., Ishaq, M., Tariq, R., Gul, K., & Ahmad, W. (2015). Pyrolysis
study of polypropylene and polyethylene into premium oil products. International Journal of
Green Energy, 12(7), 663–671. https://doi.org/10.1080/15435075.2014.880146
51. Witkowski, A., Stec, A. A., & Hull, T. R. (2016). Thermal decomposition of polymeric
materials. In SFPE handbook of fire protection engineering (pp. 167–254).
52. Sakata, Y., Uddin, M. A., & Muto, A. (1999). Degradation of polyethylene and polypropylene
into fuel oil by using solid acid and non-acid catalysts. Journal of analytical and Applied
Pyrolysis, 51(1–2), 135–155. https://doi.org/10.1016/S0165-2370(99)00013-3
53. Demirbas, A. (2004). Pyrolysis of municipal plastic wastes for recovery of gasoline-range
hydrocarbons. Journal of Analytical and Applied Pyrolysis, 72(1), 97–102. https://doi.org/10.
1016/j.jaap.2004.03.001
54. Jung, S. H., Cho, M. H., Kang, B. S., & Kim, J. S. (2010). Pyrolysis of a fraction of waste
polypropylene and polyethylene for the recovery of BTX aromatics using a fluidized bed reactor.
Fuel Processing Technology, 91(3), 277–284. https://doi.org/10.1016/j.fuproc.2009.10.009
192 P. Dash et al.

55. Abbas-Abadi, M. S., Haghighi, M. N., Yeganeh, H., & McDonald, A. G. (2014). Evaluation
of pyrolysis process parameters on polypropylene degradation products. Journal of Analytical
and Applied Pyrolysis, 109, 272–277. https://doi.org/10.1016/j.jaap.2014.05.023
56. Liu, X., Burra, K. G., Wang, Z., Li, J., Che, D., & Gupta, A. K. (2020). On deconvolution
for understanding synergistic effects in co-pyrolysis of pinewood and polypropylene. Applied
Energy, 279, 115811. https://doi.org/10.1016/j.apenergy.2020.115811
57. Duan, D., Wang, Y., Dai, L., Ruan, R., Zhao, Y., Fan, L., Tayier, M., & Liu, Y. (2017).
Ex-situ catalytic co-pyrolysis of lignin and polypropylene to upgrade bio-oil quality by
microwave heating. Bioresource Technology, 241, 207–213. https://doi.org/10.1016/j.biortech.
2017.04.104
58. Yao, L., King, J., Wu, D., Chuang, S. S., & Peng, Z. (2021). Non-thermal plasma-assisted
hydrogenolysis of polyethylene to light hydrocarbons. Catalysis Communications, 150,
106274. https://doi.org/10.1016/j.catcom.2020.106274
59. Loy, A. C., Quitain, A. T., Lam, M. K., Yusup, S., Sasaki, M., & Kida, T. (2019). Develop-
ment of high microwave-absorptive bifunctional graphene oxide-based catalyst for biodiesel
production. Energy Conversion and Management, 180, 1013–1025. https://doi.org/10.1016/j.
enconman.2018.11.043
60. Bu, Q., Chen, K., Xie, W., Liu, Y., Cao, M., Kong, X., Chu, Q., & Mao, H. (2019). Hydrocarbon
rich bio-oil production, thermal behavior analysis and kinetic study of microwave-assisted co-
pyrolysis of microwave-torrefied lignin with low density polyethylene. Bioresource Technology,
291, 121860. https://doi.org/10.1016/j.biortech.2019.121860
61. Mahari, W. A., Chong, C. T., Cheng, C. K., Lee, C. L., Hendrata, K., Yek, P. N., Ma, N.
L., & Lam, S. S. (2018). Production of value-added liquid fuel via microwave co-pyrolysis of
used frying oil and plastic waste. Energy, 162, 309–317. https://doi.org/10.1016/j.energy.2018.
08.002
62. Mączka, T., Śliwka, E., & Wnukowski, M. (2013). Plasma gasification of waste plastics. Journal
of Ecological Engineering, 14(1), 33–9. https://doi.org/10.5604/2081139X.1031534
63. Al Rayaan, M. B. (2021). Recent advancements of thermochemical conversion of plastic waste
to biofuel-A review. Cleaner Engineering and Technology, 2, 100062. https://doi.org/10.1016/
j.clet.2021.100062
64. Hongthong, S., Leese, H. S., & Chuck, C. J. (2020). Valorizing plastic-contaminated waste
streams through the catalytic hydrothermal processing of polypropylene with lignocellulose.
ACS Omega, 5(32), 20586–20598.
65. Qureshi, M. S., Oasmaa, A., Pihkola, H., Deviatkin, I., Tenhunen, A., Mannila, J., Minkkinen,
H., Pohjakallio, M., & Laine-Ylijoki, J. (2020). Pyrolysis of plastic waste: Opportunities and
challenges. Journal of Analytical and Applied Pyrolysis, 152, 104804. https://doi.org/10.1016/
j.jaap.2020.104804
66. Scheirs, J., & Kaminsky, W. (2006). Feedstock recycling and pyrolysis of waste plastics. Wiley.
67. Butler, E., Devlin, G., & McDonnell, K. (2011). Waste polyolefins to liquid fuels via pyrolysis:
Review of commercial state-of-the-art and recent laboratory research. Waste and Biomass
Valorization., 2, 227–255.
68. Lam, S. S., Mahari, W. A., Ok, Y. S., Peng, W., Chong, C. T., Ma, N. L., Chase, H. A., Liew, Z.,
Yusup, S., Kwon, E. E., & Tsang, D. C. (2019). Microwave vacuum pyrolysis of waste plastic
and used cooking oil for simultaneous waste reduction and sustainable energy conversion:
Recovery of cleaner liquid fuel and techno-economic analysis. Renewable and Sustainable
Energy Reviews, 115, 109359. https://doi.org/10.1016/j.rser.2019.109359
Chapter 8
Functionalized Biochar for Green
and Sustainable Production of Biodiesel
Hlawncheu Zohmingliana, Joseph V. L. Ruatpuia,
and Samuel Lalthazuala Rokhum

Abstract Biomass is a renewable energy source generated from agricultural waste,


solid waste, etc. In addition to being utilized extensively for storing energy and gas,
biomass is a major source of catalysts for biodiesel production. Biochar is a by-
product of thermochemically degrading biomass through processes like pyrolysis,
hydrothermal conversion, carbonization, etc., which has great potential as an energy
storage material, a precursor to nanotubes and graphene, and as a catalyst for various
chemical reactions. The high porosity, large surface area, low cost, and ease of
generation make biochar a great contender for a catalyst. The post modification
includes the physical or chemical activation which makes the surface area larger by
increasing the pores. It also includes functionalization by adding a specific functional
group for a particular chemical reaction. Acid-functionalized and base-functionalized
biochar can be used for the transesterification of triglycerides and the esterification
of free fatty acids. This production of biodiesel using functionalized biochar as
a catalyst can be enhanced by optimizing different reaction parameters like time,
temperature, etc. Overall, functionalized biochar is a strong candidate as a catalyst
for the production of biodiesel.

Keywords Biomass · Biodiesel · Pyrolysis · Hydrothermal · Trans/esterification ·


Biochar

8.1 Introduction

With the depletion of fossil fuel stock and regard for the environmental impact like
rising temperature, air pollution, melting of ice, etc., there has been research for
finding an alternative energy source for more than a decade [1]. Although renewable
energy like solar, wind, geothermal, and hydropower were studied and have been used
as a source of energy in many places, the inconsistent energy output, the impact on

H. Zohmingliana · J. V. L. Ruatpuia · S. L. Rokhum (B)


Department of Chemistry, National Institute of Technology Silchar, Silchar 788010, India
e-mail: rokhum@che.nits.ac.in

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2024 193
G. Baskar et al. (eds.), Circular Bioeconomy Perspectives in Sustainable Bioenergy
Production, Energy, Environment, and Sustainability,
https://doi.org/10.1007/978-981-97-2523-6_8
194 H. Zohmingliana et al.

the environment, and the irregular pattern posed an issue on the availability through
the year [2]. Because biomass is readily available and provides energy in all states of
physical existence, it is thought to be a good replacement for fossil fuels in light of all
these challenges and problems [3]. Biomass is a naturally occurring organic resource
that is renewable and derived from plant and animal waste [4]. This biomass can be
transformed into fuel, energy storage material, a catalyst for different chemical reac-
tions, and some compounds with added value [5]. The different possible conversion
method includes pyrolysis, gasification, hydrothermal conversion, etc. The products
from the degradation of biomass can be biochar (solid), crude (liquid), and syngas
(gas) [6].
When biomass is broken down, a carbon-rich material called biochar results.
There are two main ways to obtain it: pyrolysis, which includes heating the biomass
in the absence of oxygen, or hydrothermal conversion, which requires employing
high temperatures and pressure in an environment based on water [7]. Although
pyrolysis is one of the common methods for biochar production, the necessity for
the raw material to be moisture-free and should be done in the absence of oxygen
makes the hydrothermal process more convenient where the process does not need
moisture removal [8]. In comparison with other carbonaceous material from other
chemical process, biochar is much more inexpensive and easy to produce [9]. The
biochar still needs to be activated to increase the surface area by physical or chemical
methods which will enhance the activity of the biochar [10]. The functionalization
approach has also been investigated for further use of the biochar as a catalyst. The
biochar produced differs significantly in its physicochemical properties depending
on the raw biomass, activation process, conversion method, and functionalization
[11]. The activated and functionalized biochar has been widely used for a variety
of applications, including gas storage, energy storage, supercapacitors, wastewater
treatment, and, more importantly, as a versatile catalyst and catalyst support due to
its large surface area and abundance of functional groups [12].
Biodiesel are produced from plants and vegetable oil and as well as animal oil
as a feedstock [13]. This oil mainly consists of mixtures of triglycerides and free
fatty acids depending on the quality of the oil. This oil reacts with low molecular
weight alcohol such as methanol, ethanol, and propanol [14]. For practical purposes,
methanol is commonly used as it is the least expensive. The reaction is known as
transesterification which produces a mixture of fatty acids methyl ester and glycerol
[15] (Fig. 8.1).

CH2 OOC R1 CH2 OH


R1 COO R'
Catalyst
CH OOC R2 + 3R'OH R2 COO R' + CH OH

R3 COO R' CH2 OH


CH2 OOC R3

Glyceride Alcohol Esters Glycerol

Fig. 8.1 Transesterification of glyceride


8 Functionalized Biochar for Green and Sustainable Production of Biodiesel 195

Catalyst
R1 COOH + R'OH R1 COO R' + H2O
Fatty acid Alcohol Ester Water

Fig. 8.2 Esterification of fatty acids

Free fatty acids (FFA) and alcohol can be used to create biodiesel by catalyzing
an esterification reaction that yields both biodiesel and water [16] (Fig. 8.2).
The transesterification process can be catalyzed by various different catalysts
through different processes: (1) homogeneous alkali, (2) homogeneous acid, (3)
heterogeneous basic catalyst, and (4) heterogeneous acid catalyst [17–20]. For the
industrial purposes, the homogeneous alkali-catalyzed method is widely used. It
produced a high yield within a very short period, but it is the most expensive
process since the separation of the homogeneous catalyst is very hard and needs
a lot of washing to produce high purity [21]. Due to their limited tolerance for water
and free fatty acids (FFA), basic catalysts provide difficulties in the manufacture of
biodiesel. Due to the need for pure vegetable oil as a feedstock, the procedure is,
therefore, more expensive overall [17]. The homogeneous acid catalyst improves the
transesterification with low quality oil and high FFA content of up to 5 wt.% [22].
However, to get a conversion of more than 95%, the reaction took more than 48 h
in mild condition [23]. Both the homogeneous alkali and acidic processes require a
significant amount of water to separate the catalyst, which increases the amount of
waste produced and calls for a more involved procedure. Heterogeneous catalysts are
the most economical since they can be easily separated from the product and can be
reused without much loss of activity [24, 25]. It does not require water for separation
and neutralization and hence has a much easier process. Between the basic and acidic
heterogeneous catalysts, the acid catalyst is significantly more effective because it
can concurrently catalyze the esterification and transesterification reactions, enabling
the processing of feedstocks with lower quality and cheaper cost [26].

8.2 Biochar Production

The conversion of biomass to biochar generally includes methods such as pyrol-


ysis, carbonization, hydrothermal process, liquefaction, torrefaction, and liquefac-
tion [28]. They are divided into two major types depending on whether water is
added as a solvent or catalyst and whether it is absent in the process. The amount
and properties of the biochar produced depend on the feedstock biomass as well as
the type of conversion process used. In this chapter, we will discuss only pyrolysis
and hydrothermal conversion (Fig. 8.3).
196 H. Zohmingliana et al.

Fig. 8.3 Methods involved in producing biochar-based catalyst. Reproduced from Ref. [27]

8.2.1 Pyrolysis

Pyrolysis, which is the heating of organic material waste to some high temperature
in the absence of oxygen, is one of the method to produce biochar [29, 30]. Co-
pyrolysis is a method that involves simultaneously pyrolyzing multiple feedstocks,
such as plastic waste and biomass, to create a variety of useful products. This method
enables the synthesis of solid carbon, hydrogen, value-added compounds, and fuels
[31]. The pyrolysis is divided into two categories based on the heating rate: slow
pyrolysis (heating rates of 0.01–1 ºC/s) [32] and rapid pyrolysis (heating rates of 10–
1000 ºC/s) [33]. The average heating range is 400–600 ºC, and the residence duration
in an atmosphere without oxygen ranges from a few minutes to several hours [32, 33].
The slow pyrolysis maximizes the biochar [34], while the fast pyrolysis increases
the bio-oil yield [35].
The biochar produced by pyrolysis are termed as pyrochar [36]. Biomass pyrolysis
involves free radical reaction which is very reactive, and hence, the overall reaction
is a very complex process. Free radicals produced during the pyrolysis process can
participate in several chemical processes, such as free radical substitution, free radical
addition, and coupling reactions. These interactions can take place between the free
radicals themselves or with other substances already in the system. As a result,
a variety of products can be created, including syngas, biochar, and bio-oil [37].
The pyrolysis parameters, including temperature, heating rate, residence time, and
the kind of feedstock, can have a big impact on the properties of the biochar that
is produced, including its stability, carbon content, surface area, capacity to retain
nutrients, pH, and cation exchange capacity [32].
Mechanism of Biochar Formation by Pyrolysis
The mechanism of biochar synthesis during pyrolysis involves several complex
thermal and chemical processes. The three main stages of the process are drying,
pyrolysis, and carbonization [38]. Following is a general description of each phase:
8 Functionalized Biochar for Green and Sustainable Production of Biodiesel 197

I. Drying: In the beginning, the feedstock is heated to remove moisture. At the


relatively low temperature, which is typically below 100 ºC, the main goal is to
eliminate the water content from the material [39].
II. Pyrolysis: The temperature is raised to start the pyrolysis process once the drying
stage is finished, typically between 200 and 500 ºC. When there is little or no
oxygen available, pyrolysis takes place. This phase sees the feedstock’s organic
molecules decompose as a result of intricate chemical interactions [35].
(a) Volatile Release: Volatile substances from the feedstock, including water,
gases (methane, ethylene, carbon monoxide, etc.), and tars, are released as
the temperature rises. This process is frequently known as devolatilization
[40].
(b) Fragmentation and Rearrangement: Thermal decomposition is used to
further break down the solid residue. Due to reactions like cracking,
dehydrogenation, decarboxylation, and depolymerization, bigger organic
molecules become fragmented at high temperatures. As a result, numerous
intermediate chemicals are created [37].
(c) Char Formation: Carbonaceous elements start to build and turn into a solid
residue known as biochar as the thermal decomposition process advances.
Rearranging and condensing the residual carbon-rich pieces during the
carbonization process results in the production of a stable carbon matrix
[41].
III. Carbonization: The maturation and stabilization of the carbonaceous structure
continue during the production of biochar. As the temperature rises, usually
above 500 ºC, and the thermal breakdown reactions become more constrained,
this happens [42]. The carbonization procedure increases the biochar’s carbon
content, lowers volatile matter, and improves its stability [43].
Based on variables including feedstock content, heating rate, temperature, and
residence time, the precise reaction routes and mechanisms during pyrolysis can
change. These factors may affect the amount of volatile emission, the level of
carbonization, and the ultimate characteristics of the biochar produced [43]. It is
crucial to remember that pyrolysis is a complicated process, and current research is
currently being done to determine the precise mechanisms and reaction routes. To
make biochar with desirable qualities for uses like soil amendment, carbon sequestra-
tion, and energy production, researchers are still looking into and adjusting pyrolysis
settings (Fig. 8.4).

8.2.2 Hydrothermal Conversion

A technique known as hydrothermal conversion uses high-temperature and high-


pressure water to transform biomass or other organic materials into valuable goods
[45]. It is a type of thermochemical conversion in which several chemical processes
198 H. Zohmingliana et al.

Fig. 8.4 Possible formation pathway of pyrochar. Reproduced from Ref. [44] with permission from
the Royal Society of Chemistry
8 Functionalized Biochar for Green and Sustainable Production of Biodiesel 199

are carried out using water as a medium. Hydrothermal conversion is commonly


selected as the preferred method when the feedstock cannot be pyrolyzed or has a
high moisture content that would require a considerable energy input for drying [46].
Biomass or organic materials are frequently heated to temperatures exceeding
180 ºC and compressed above atmospheric pressure during hydrothermal conver-
sion. High-pressure water works as a catalyst and solvent to help break down
complex chemical compounds into smaller ones. Three different processes can occur:
gasification (gas), liquefaction (liquid), and carbonization (solid) [47].

8.2.2.1 Hydrothermal Carbonization

Biomass or organic waste is transformed into hydrochar or biocarbon through the


thermochemical process of hydrothermal carbonization (HTC) [48]. Because it
occurs in the presence of water under intense pressure and heat, it is frequently
referred to as “wet carbonization.” Biomass or organic waste is often combined
with water and heated to temperatures between 180 and 250 °C under pressures
between 10 and 25 bar during the hydrothermal carbonization process. The presence
of water and the high temperature and pressure conditions encourage the breakdown
of complex organic molecules in the feedstock and start several chemical processes
[49].
Several concurrent reactions take place during hydrothermal carbonization. The
biomass is initially partially hydrolyzed, which results in the creation of water-soluble
molecules including sugars and organic acids. These soluble molecules go through
additional reactions as the process progresses, such as dehydration, decarboxyla-
tion, and polymerization. This causes the production of solid hydrochar, a carbon-
rich substance that resembles char [50]. The hydrochar created by HTC has several
advantageous qualities. Its high carbon content, which normally ranges from 50 to
90%, is what makes it unique. Similar to coal, it has a high energy density and can
be utilized as a solid fuel for the production of heat and power [51]. In addition,
hydrochar is stable and has a low oxygen content, which makes it more resistant to
biodegradation and less prone to combustion than raw biomass [52].

8.2.2.2 Hydrothermal Liquefaction (HTL)

HTL entails converting biomass into a liquid substance known as bio-oil or crude
[53]. In this procedure, the complex organic molecules in the biomass are broken
down by the high-temperature and high-pressure water, creating a mixture of liquid
organic chemicals. By further refining the created bio-oil, important items including
transportation fuels, chemicals, and bio-based materials can be obtained [54].
200 H. Zohmingliana et al.

8.2.2.3 Hydrothermal Gasification (HTG)

Biomass is transformed into a mixture of gases, chiefly hydrogen (H2 ), carbon


monoxide (CO), and methane (CH4 ), in the process of hydrogen-to-gas (HTG) [55].
Syngas, also known as synthesis gas, is created as a result of the organic material’s
reaction with the high-pressure water. Syngas is a feedstock that can be used to make
chemicals, fertilizers, and biofuels.
Compared to other methods of biomass conversion, hydrothermal conversion has
several benefits. Agricultural waste, energy crops, algae, and organic waste are just
a few of the many feedstocks it can process. When compared to traditional pyrol-
ysis or gasification, it is a comparatively quick process and may be carried out at
lower temperatures. Additionally, the breakdown of biomass components is aided
by the high-pressure water conditions, allowing for a more thorough conversion and
minimizing the production of char or solid leftovers [48].
Mechanism of Hydrochar Formation by Hydrothermal Carbonization
The process of hydrothermal carbonization (HTC), which produces hydrochar,
comprises several intricate processes that take place in high-temperature, high-
pressure water environments [56]. Numerous crucial processes are thought to play a
role in the creation of hydrochar, even though the precise mechanisms are not entirely
known. Here are a few of the suggested mechanisms:
I. Hydrolysis: Complex polysaccharides, lignin, and other organic components
are hydrolyzed in the early stages of HTC when biomass or organic waste
is subjected to hot water. These large molecules undergo hydrolysis, which
reduces them to smaller, water-soluble substances like sugars and organic acids.
These soluble molecules are capable of additional reactions to produce other
products [57].
II. Dehydration: Water removal from the hydrolysis products causes dehydration
processes to take place. Furans, aldehydes, ketones, and other reactive inter-
mediates are created as a result of the elimination of water molecules. The
carbonization process is aided by these chemicals’ ability to generate more
durable aromatic structures through subsequent reactions [57].
III. Decarboxylation: Decarboxylation reactions include the removal of carboxyl
groups (–COOH) from organic acids and other carboxyl-containing substances.
This process results in the production of aromatic chemicals and the release of
carbon dioxide (CO2 ). Decarboxylation increases the carbon content of the final
hydrochar product, which helps to carbonize the biomass [57].
IV. Polymerization: Hydrochar is created primarily by polymerization processes.
When the reactive intermediates generated by hydrolysis, dehydration, and
decarboxylation are subjected to polymerization, they interact with one another
to create bigger, more complex molecules. Cross-linked structures are created
as a result of polymerization events, which also help the hydrochar to become
more aromatic [57].
8 Functionalized Biochar for Green and Sustainable Production of Biodiesel 201

V. Condensation: In condensation reactions, smaller molecules create chem-


ical interactions with one another to generate larger, more condensed struc-
tures. These reactions aid in the carbonization process and the synthesis of
molecules rich in aromatic carbon. Several reactive intermediates, including
furans, phenols, and aldehydes, can undergo condensation reactions with one
another [57].
VI. Carbonization and Solidification: The organic compounds continue to change
as the reactions advance, and eventually, the system reaches a point where the
hydrochar starts to harden. A carbon-rich matrix with a porous structure is
created during this solidification process. Temperature, pressure, reaction time,
and the makeup of the biomass fuel are all elements that affect the solidification
process [56].
The precise reaction pathways and processes can differ depending on the type of
biomass feedstock, the reaction environment, and the experimental setup, which is
vital to highlight. To optimize the HTC process and enhance the quality and attributes
of the hydrochar produced, researchers are still investigating the intricate principles
of hydrochar creation.

8.3 Feedstock for Biochar Production

Wood chips and sawdust, which include lignocellulosic components, are the most
frequently utilized feedstock. Different lignin, hemicellulose, and cellulose types can
be found in different types of agricultural and forestry waste [58]. These leftovers,
which are frequently not used after harvesting, might be used to make biochar. Nut
shells like those from almonds, walnut, and coconut can be utilized as raw materials
to make biochar. These shells can be used to create biochar, even though they are
generally dumped as waste [59]. Livestock waste, such as poultry litter, cow dung,
and pig manure, can be used as feedstocks for biochar production [60]. This not only
helps in waste management but also produces biochar with added nutrient content.
Fast-growing energy crops like bamboo, miscanthus, and switchgrass can be used as
feedstocks for biochar production. These crops are specifically grown for biomass
production and can be sustainably harvested for biochar [61].
Organic waste from households, restaurants, and other urban sources can be used
as feedstocks for biochar production. This helps in waste diversion and reduces the
amount of waste sent to landfills [62]. Certain types of algae and aquatic plants, such
as water hyacinth and duckweed, can be used as feedstocks for biochar production.
These plants are often considered invasive species and can be harvested to produce
biochar [63]. However, animal waste and aquatic materials that have low lignin and
cellulose content have not been studied extensively due to their difficult handling
condition. Different biomass with different chemical compositions (lignin, hemicel-
lulose, and cellulose) are thought to have a significant impact on the properties of the
biochar produced, surface area, reactivity with chemical activating agent, and yields.
202 H. Zohmingliana et al.

The choice of feedstock depends on factors such as availability, cost, and desired
biochar characteristics.

8.4 Activation and Functionalization of Biochar

The biochar produced after just pyrolysis and hydrothermal carbonization may not
have enough surface area and large porosity and hence may not serve well for appli-
cation. Biochar activation refers to a process that enhances the physical and chem-
ical properties of biochar, making it more effective for specific applications [42].
Activated biochar has increased surface area, pore volume, and reactivity, which
improves its ability to adsorb and retain nutrients, pollutants, and other substances
[64]. In addition to its core function, activated biochar has a wide range of uses due
to its improved surface and adsorption qualities. It can be used in a variety of fields,
including environmental adsorbents, soil amendment, carbon sequestration agents,
and activated carbon.

8.4.1 Types of Biochar Activation Method

8.4.1.1 Physical Activation

Physical activation involves subjecting biochar to high temperatures above 700 ºC


in the presence of gases, typically steam or carbon dioxide. This process, known as
steam or CO2 activation, causes the carbon matrix to expand, creating additional pores
and increasing surface area [8]. Removing the unstructured parts, this disintegration
only leaves the carbon structure with its tiny pores. As a result, the interior surface area
of the activated biochar increases. The next stage of the activation process involves
the depletion of crystalloid carbon, which is defined as carbonized material or carbon
with small pores. Because of this depletion, the activated biochar develops bigger
pores. The pores in the biochar that are loaded with carbon volatiles are opened under
precise circumstances during the activation process. These volatiles can diffuse with
nearby pores thanks to this opening. These occurrences result in a large increase in
the biochar’s specific surface area, the production of micropores, and a noticeable
decrease in the concentration of mesopores [42] (Fig. 8.5).
The yield of activated carbon may decrease as a result of the increased danger of
combustion from air activation processes that take longer to complete. The likelihood
that the carbonaceous material will reach its ignition point and undergo combustion
is increased by the lengthier exposure to air. As a result, it is essential to maintain the
temperature and exposure time to keep the pore volume and specific surface area. In
general, activation by air is less expensive than activation by CO2 , which produces
more pores and eventually increases the biochar’s surface area [8].
8 Functionalized Biochar for Green and Sustainable Production of Biodiesel 203

Fig. 8.5 SEM image of a steam-activated wood apple b H2 SO4 impregnated wood apple.
Reproduced from Ref. [65]

The type of biomass used, the kind of gas used during activation, and the partic-
ular activation circumstances used can all have a major impact on the features of
activated biochar. According to Nabais et al. [11], the use of steam increased the
activation rate of biochar produced from the pyrolysis of coffee endocarp, whereas
CO2 increased the specific area and pore size. Reinoso et al. [66] used steam and
CO2 to activate biochar from olive seed and compared the results. According to their
studies, activated carbon made with pure steam as the activation agent has smaller
micropore volume than activated carbon made with CO2 . Additionally, it has been
found that steam activation produces meso- and macro-porosity more efficiently than
CO2 activation, resulting in larger pores. A study was done by Koltowski et al. [10]
that looked at activating biochar made from willow that had undergone slow pyrol-
ysis. The study’s goal was to look at how CO2 and steam as activation agents affected
the properties of the activated biochar that resulted. They have reported that the steam
activation and CO2 activation have improved the surface area, and the steam activa-
tion have higher surface areas (840.6 m2 g−1 ) than CO2 activation (512.0 m2 g−1 ).
Additionally, according to the research results, steam-activated biochar showed larger
pore diameters and a greater specific surface area than CO2 -activated biochar under
similar activation settings. This suggests that the steam activation procedure resulted
in the formation of a more porous and structurally varied biochar material.

8.4.1.2 Chemical Activation

Chemical activation involves first impregnating the biochar made from biomass with
an activating chemical, such as KOH, NaOH, ZnCl2 , K2 CO3 , H2 SO4 , and H3 PO4 ,
before heating it at a high temperature while being surrounded by inert gas [8, 67,
68]. Although chemical activation has drawbacks like corrosion of the apparatus,
difficult recovery of the chemical, and cost of chemicals, it has a lot of advantages like
high efficiency, a larger surface area of more than 4000 m2 g-1, controlled porosity,
low-temperature range, and higher carbon yield. The temperature utilized during
204 H. Zohmingliana et al.

chemical activation, the kind of chemical agent used, the kind of biomass from
which the biochar is formed, and the concentration of the chemical agent are all
variables that affect the physical qualities of biochar that have undergone chemical
activation [42, 69].
The production of biochar with a significant surface area and porosity has been
seen when potassium hydroxide (KOH) is used as an activating agent in the chem-
ical activation process. This is due to the cooperative actions of numerous mecha-
nisms, such as metallic K+ intercalation, physical activation, chemical erosion, and
lattice expansion of carbon [42]. According to research done by Ros et al. [70],
using phosphoric acid (H3 PO4 ) to sludge char as an activating agent did not signifi-
cantly improve the specific surface area or pore volume. Similar to how alkali metal
activating agents increased specific surface area and pore volume, activation with
carbon dioxide (CO2 ) had a much lower effect. Two activators, zinc chloride (ZnCl2 )
and ammonium chloride (NH4 Cl), were used in a different investigation to activate
biochar made from mandarin peel. A tubular furnace heated to 700 ºC was used for
the activation procedure [71]. It was found out that although NH4 Cl increased the
surface area to 181 m2 g−1 , it was not comparable to the surface area increased by
using ZnCl2 to 1085 m2 g−1 .
In a study by Park et al. [72], they used different chemical activators, such as HCl,
H2 SO4 , H3 PO4 , KOH, MgO, ZnCl2 , and K2 SO4 , to activate sesame straw biochar.
Their study’s goal was to find out how the various activation agents affected the
amount of phosphorus that could be absorbed by the resulting activated biochar.
The research results showed that ZnCl2 and MgO were more effective than the
other chemical activating agents at increasing the phosphorus adsorption capacity
of sesame straw biochar than the other activating agents evaluated. Particularly, the
ZnCl2 -activated biochar had the highest phosphorus adsorption capability, with a
staggering maximum of 15 g per kilogram (g kg−1 ). It is therefore a highly effective
phosphorus adsorbent, capable of adsorbing up to 15 g of phosphorus for every
kilogram of ZnCl2 -activated biochar [72].

8.4.1.3 Biological Activation

When microorganisms are inoculated or added to biochar to encourage biological


activity and nutrient cycling in the soil, this process is referred to as biological
activation. In this procedure, the biochar is “primed” or “pre-charged” with microbes
that can colonize its surface and foster microbial activity. Biological activation may
involve the employment of advantageous bacteria, fungi, and other soil microbes.
They can be obtained from vermicompost, compost, or other organic materials rich
in microorganisms. The activation process typically involves mixing biochar with
the microbial inoculant and allowing it to interact and colonize the biochar surface.
The removal of lignin from biowaste and the hydrolysis of cellulose using cellulose
enzymes created highly microporous biochar with a larger surface area. The addition
of paraffin through vacuum impregnation then created a 3D porous structure with a
high energy capacity of 141.47 J/g, demonstrating that bio activation of biowaste can
8 Functionalized Biochar for Green and Sustainable Production of Biodiesel 205

create a potential carbon source for energy applications [73]. A highly porous carbon
material with a remarkable surface area of 1831 m2 /g, mostly made up of micro-
and mesopores, was produced when lignin was activated utilizing green bacterial
processes. For the material to function well in many applications, its broad porosity
is essential. The fungus Trichoderma was grown in a mixture of wheat straw and food
scraps. The obtained biomass underwent a carbonization procedure before being
heated to high temperatures between 800 and 900 ºC for activation. This procedure
produced a carbon substance with a remarkable surface area of about 4000 m2 /g. Its
volume of 2.37 cm3 /g further demonstrated the material’s porosity and potential for
effective charge storage [74].

8.4.2 Biochar Functionalization

The biochar obtained from biomass usually has low surface functionality with only
limited C–O, C=O, and OH groups present and possessing very limited porosity and
surface area (usually < 150 m2 /g) [75]. The surface functionality, however, can be
tuned very easily which offers a promising platform for synthesizing various func-
tional materials. In general, biochar can gain surface activity via surface modification
or active substance deposition [64].
Biochar modification is a flexible method for adjusting its characteristics for
certain uses. One efficient technique involves adding particular functional groups,
especially acid groups, to the surface of the organic biochar. Sulfonation with concen-
trated sulfuric acid (H2 SO4 ) or its derivatives is one of the most often used processes
for modifying biochar. Using strong sulfuric acid or an acid derivative, such as fuming
sulfuric acid (oleum), the biochar is subjected to this process to add sulfonic acid
functional groups (–SO3 H) onto the surface of the material [42]. Other weak acid
functional groups, such as carboxylic acid groups (–COOH), can also be introduced to
the biochar surface during alteration in addition to sulfonic acid groups (–SO3 H). The
addition of –COOH groups can increase the modified biochar’s catalytic activity and
reactivity, making it a flexible solid acid catalyst for a variety of chemical reactions
[76].
Solid acids derived from biochar that have functional groups added, including –
SO3 H and –COOH, demonstrate a number of key benefits that make them attractive
substitutes for conventional mineral acid catalysts in a variety of catalytic reactions
[77]. By subjecting biochar to gaseous sulfur trioxide (SO3 ) concentrations greater
than 20%, a process known as “gaseous sulfonation” is used to alter the material.
Between 25 and 150 °C, relatively low temperatures are commonly used for this
operation. The biochar interacts with the SO3 gas during the sulfonation process,
which results in the insertion of sulfonic acid functional groups (–SO3 H) onto its
surface [78]. Another way to alter biochar is by treating it with a liquid sulfonating
agent, usually concentrated sulfuric acid (H2 SO4 ) or its derivatives. Liquid sulfona-
tion typically makes it easier for the surface area of the biochar to expand more than
gaseous sulfonation does.
206 H. Zohmingliana et al.

There are several methods for chemical functionalization of biochar, and here are
a few commonly employed techniques:
I. Acid Treatment: Acids like nitric acid (HNO3 ) and hydrochloric acid (HCl) can
be used to treat biochar. By adding carboxylic acid (–COOH) functional groups
to the surface of the biochar during acid treatment, it becomes more hydrophilic
and has a greater chance of interacting with other molecules.
II. Base Treatment: Similar to acid treatment, biochar can be treated with bases like
sodium hydroxide (NaOH) or potassium hydroxide (KOH). Basic functional
groups (–OH) are added to the surface of biochar during base treatment, altering
its properties and reactivity.
III. Silanization: It is possible to silanize biochar by using organosilanes, which
contain one silicon atom and organic functional groups. This process enables
the addition of additional functional groups to the surface of the biochar, such as
amino (–NH2 ) or alkyl (–CH3 ) groups. For adapting biochar’s surface chemistry
for specific applications, silanization offers a flexible technique.
IV. Ozonation: The treatment of charcoal with ozone (O3 ) gas is known as ozona-
tion. This process exposes the biochar surface to oxygen-containing func-
tional groups, such as carbonyl (–C=O) or hydroxyl (–OH) groups. Ozonation,
which raises the surface reactivity of biochar, can increase its ability to absorb
substances.
V. Plasma Treatment: Plasma is used to treat biochar by exposing it to a low-
temperature plasma discharge. Radicals, a class of extremely reactive species
created by plasma, can interact with biochar’s surface to introduce different
functional groups. Plasma treatment is an efficient and well-controlled method
of functionalizing biochar.
The chemical functionalization of biochar can alter its surface chemistry,
hydrophobicity/hydrophilicity, porosity, and reactivity, making it suitable for diverse
applications. Functionalized biochar has been used in environmental remediation,
water treatment, soil improvement, energy storage, and as catalyst supports, among
other areas. The specific functionalization method chosen depends on the desired
properties and targeted applications of the biochar (Fig. 8.6).

Fig. 8.6 Model of porous


biochar containing different
functional groups.
Reproduced from Ref. [77]
8 Functionalized Biochar for Green and Sustainable Production of Biodiesel 207

8.5 Biodiesel Production by Biochar-Based Catalyst

A sustainable and eco-friendly substitute for petroleum diesel is biodiesel. It is


a sustainable fuel made by esterification and transesterification of animal fats,
vegetable oils, or used cooking oil with alcohols like methanol or ethanol. These
processes produce a mixture of mono-alkyl esters, which combine to generate
biodiesel [79]. Esterification and transesterification are the two primary types of
processes used in the manufacturing of biodiesel to transform vegetable or animal
fats into mono-alkyl esters. These procedures necessitate the use of a catalyst and
require the conversion of triglycerides and free fatty acids into biodiesel. Both
acidic and basic catalysts can aid in the transesterification and esterification proce-
dures required for the manufacture of biodiesel. Numerous heterogeneous and
homogeneous catalysts have been employed for this purpose [80].
In reality, homogeneous catalysts allow for rapid and complete conversion of
feedstocks while also offering a high level of reaction efficiency for the production
of biodiesel. But their major drawback is that it takes a lot of work to purify the
product and is difficult to reuse, which raises production expenses [81]. Accord-
ingly, several carbon-based acid and basic catalyst as well as heterogeneous acid
and basic catalyst have been employed to produce biodiesel. These catalysts include
CaO, MgO, Amberlyst-15, TiO2 /ZrO2 , Al2 O3 /ZrO2 , and WO3 /ZrO2 [82]. The use
of carbon-based biochar as a catalyst has many benefits. These advantages include
its affordability, substantial surface area, wide accessibility, and capacity to tailor
its surface functional groups to particular applications. The biochar is activated by
some activating agent which increase the porosity and hence the surface area after
which it is subjected for functionalization and make an acid or basic catalyst, which
is used for biodiesel production [83].
Acid-functionalized biochar catalysts and base-functionalized biochar catalysts
are two different types of biochar catalysts that can be useful in the generation of
biodiesel. Sulfonated biochars have been the most popular heterogeneous catalysts
for this use among them. Sulfonic acid groups (–SO3 H) are added to the surface
of biochar during the sulfonation process. This alteration increases the biochar’s
catalytic activity and makes it a powerful solid acid catalyst for the synthesis of
biodiesel. Sulfonation can be accomplished in a variety of ways. One method includes
heating a concentrated solution of sulfuric acid (H2 SO4 ) and soaking the charcoal in
it. The sulfonic acid groups are immobilized on the surface of the biochar during this
procedure. Another way to sulfonate biochar is to expose it to gaseous sulfur trioxide
(SO3 ). This exposure causes -SO3 H groups to be incorporated onto the surface of
the biochar, turning it into a powerful solid acid catalyst. These –SO3 H groups can
catalyze the esterification of free fatty acids (FFAs) or promote the transesterification
reactions of triglycerides with alcohols to form biodiesel [79] (Fig. 8.7).
Rokhum et al. [84] synthesized a carbon-based solid acid catalyst from glucose
via a an efficient one-pot hydrothermal carbonization-sulfonation without the need
of high temperature. The catalyst is combined with aromatic structure of hydrophilic
–OH and –COOH groups and a high density of –SO3 H centers. In their study, the
208 H. Zohmingliana et al.

Fig. 8.7 Representation of the one-pot, multistep conversion of glucose to the mesoporous
SAFACAM pictured in the reagent bottle (red circle = SO3 H). Reproduced from Ref. [84]

relevance of the catalyst was tested with oleic acid, where the level of SO3 H (0.81–
1.29 mmolg−1 ) proved synthetically tunable. In a study, Kasner et al. synthesized
sulfonated biochar using two different techniques: (a) concentrated sulfuric acid
(H2 SO4 ) and (b) gaseous sulfur trioxide (SO3 ). Both of these techniques were used
to add sulfonic acid groups (–SO3 H) to the biochar’s surface, creating sulfonated
biochar catalysts. Their investigation proved that the biochar matrix’s surface area
and pore volume increased as a result of sulfonation using concentrated sulfuric
acid (H2 SO4 ). This indicates that there were more active sites and available area for
catalytic reactions in the improved biochar. The biochar’s surface was successfully
increased throughout the H2 SO4 sulfonation procedure, making catalytic activities
easier [78]. In their studies, they found that the sulfonated biochar catalyst had impres-
sive catalytic activity for the esterification of fatty acids during their investigations.
The outcomes demonstrated a remarkable conversion rate of the fatty acids, which
produced the matching fatty acid methyl esters (FAMEs) at a rate of about 90–100%.
It only took between 30 and 60 min for this high conversion rate to be achieved. The
esterification processes were carried out in the 55–60 ºC temperature range.
González et al. [85] synthesized biochar using oat hull as a feedstock, and then,
they used concentrated sulfuric acid (H2 SO4 ) to sulfonate the finished product. In
order to perfect the alteration, they tested with different temperatures throughout
the sulfonation process in a microwave reactor. After getting sulfonated biochar
catalysts with various levels of sulfonation, researchers looked into how well these
8 Functionalized Biochar for Green and Sustainable Production of Biodiesel 209

substances worked as catalysts for the synthesis of biodiesel. They specifically


employed discarded cooking oil as the feedstock for the synthesis of biodiesel.
The same microwave reactor was used to produce biodiesel from start to finish,
including the esterification of free fatty acids and the transesterification of triglyc-
erides. Dehkhoda et al. [16] sulfonated hardwood char using concentrated sulfuric
acid and fuming sulfuric acid, activated it chemically with 10 M KOH to increase
the specific surface area, and applied it to the transesterification of vegetable oil and
the esterification of FFA. Li et al. [83] sulfonated pyrolyzed rice hulk by concen-
trated sulfuric acid and used the prepared catalyst for simultaneous esterification and
transesterification to produced biodiesel from waste cooking oil. In their study, they
have showed that the yield of fatty acid methyl ester is 87.57% after 15 h.
In fact, base-functionalized biochar catalysts have been used in addition to acid-
functionalized biochar catalysts to produce biodiesel. CaO/biochar and KOH/biochar
are two examples of such catalysts [86, 87]. Using a CaO-based palm kernel shell
biochar catalyst, McKay and colleagues created a new catalyst for their study. This
base-functionalized biochar catalyst was created by impregnating biochar made from
palm kernel shells in calcium oxide (CaO). They examined the effectiveness of the
CaO-based biochar catalyst in the transesterification reaction for the creation of
fatty acid methyl esters (FAMEs) from sunflower oil. In their investigation, it was
found that under particular ideal conditions, they were able to produce an amazing
99% yield of fatty acid methyl esters (FAMEs). A 3 wt.% catalyst loading, 65 ºC
reaction temperature, and a 9:1 (MTOR) were all used to achieve this high yield.
Wang et al. [81] also synthesized a series of peat biochar supported K2 CO3 catalyst
using wet impregnation method with different K2 CO3 loading (20–40 wt.%) [88].
Table 8.1 presents the production of biodiesel from different feedstock using different
biochar-based acid and basic catalyst.

8.5.1 Comparison of Biochar and Functionalized Biochar


in Biodiesel Production

In summary, the comparison between biochar and functionalized biochar as catalysts


for biodiesel production highlights the superior catalytic efficiency of the latter. As
research progresses, optimizing reaction parameters, such as temperature and dura-
tion, promises to further enhance the efficiency of biodiesel production when func-
tionalized biochar is utilized. This innovation holds great promise for the sustainable
and cost-effective production of biodiesel, contributing to a more environmentally
friendly and economically viable energy source. From entries 1 and 12, we can see
that without sulfonation biochar the biodiesel yield were decreased compared to
entries (2–11, 13, and 14) having sulfonated biochar. For entry 15, although it is
non-sulfonated biochar, the yield of biodiesel was quite high; this can be due to very
high reaction temperature for transesterification reaction.
210 H. Zohmingliana et al.

Table 8.1 Biodiesel production under the catalysis of various biochar-based acid or basic catalyst
Sl. Biochar Functionalization Feedstock Reaction Biodiesel References
No. condition conditiona yield (%)
Oat hull None Waste cooking 140, 0.5 0.5 [85]
biochar oils (microwave),
10, 10:1
Oat hull Sulfonated with Waste cooking 140, 0.5 28 [85]
biochar H2 SO4 in oils (microwave),
microwave 10, 10:1
reactor at 100 ºC
for 30 min
Oat hull Sulfonated with Waste cooking 140, 0.5 72 [85]
biochar H2 SO4 in oils (microwave),
microwave 10, 10:1
reactor at 140 ºC
for 30 min
Biochar 20 g of biochar Vegetables oils 60, 3, 5, 77–89 [16]
from fast sulfonated with EtOH/oil
pyrolysis of 200 mL of (18:1)
different H2 SO4 at 150 ºC
feedstocks for 24 h
Biochar Sulfonated with 5% palmitic 65, 3, 20:1 > 80 (FFA [78]
from peanut H2 SO4 at 100 ºC and stearic conversion)
hulls, pine for 12–18 h or acid
pellets, and gaseous SO3 at
wood chips room
temperature for
six days
Wood Sulfonated with Canola oil 65, 12, 5, 44.2 [89]
biochar fuming H2 SO4 at 15:1
150 ºC for 15 h
Rice husk Sulfonated with Waste cooking 110, 15, 5, 87.6 [83]
biochar H2 SO4 at 90 ºC oils 20:1 (FAME
for 0.5 h yield)
> 98 (FFA
conversion)
Rice husk Sulfonated with Oleic acid 110, 2, 5, 4:1 98.7 [90]
biochar H2 SO4 at 70–150
ºC for 0.25–4 h
Corn straw Sulfonated with Oleic acid 60, 4, 7, 7:1 98 [91]
biochar fuming H2 SO4 at
180 ºC for 4 h
Palm kernel Loaded with Sunflower oil 60, 5, 3, 9:1 99 [86]
shell CaO at a total
biochar basicity of
0.516 mmol g−1
(continued)
8 Functionalized Biochar for Green and Sustainable Production of Biodiesel 211

Table 8.1 (continued)


Sl. Biochar Functionalization Feedstock Reaction Biodiesel References
No. condition conditiona yield (%)
Palm kernel Loaded with Sunflower oil 60, 5, 3, 9:1 99 [87]
shell 50.6% of CaO
biochar
Peat None Palm oil 65, 1.5, 5, 10.5 [81]
biochar 8:1
Peat Loaded with Palm oil 65, 1.5, 5, 95.2–98.6 [81]
biochar 20–40 wt.% of 8:1
K2 CO3
Flamboyant Impregnated in Hevea 60, 1, 3.5, 89.3 [88]
pods KOH solution brasiliensis oil 15:1
biochar
Biochar None Waste cooking 300, 1, 50:1 ~ 90 [92]
from maize oil
residue
a Temperature (ºC), time (in hour), Catalyst loading (wt.%), methyl-to-oil ratio (unless otherwise
stated)

8.5.2 Catalyst Reusability and Kinetics Study

Recycling poses a significant concern due to the limited availability of resources. In


industrial settings, the imperative is to utilize catalysts that have a prolonged lifespan
and can be easily retrieved. Heterogeneous catalysts enable straightforward recy-
cling, potentially lowering the overall expenses of chemical processes and enhancing
environmental conditions. According to Gonzales et al. [85] findings, when the cata-
lyst was employed for successive cycles, both with and without hexane washing, there
was a notable decline in catalytic activity. Specifically, the FAME yield dropped from
approximately 86% during the initial cycle to approximately 14% by the third cycle,
signifying an 84% reduction in catalytic efficiency. Conversely, when the catalyst was
subjected to hexane washing, its performance exhibited a more modest decrease. The
FAME yield began at around 90% during the first cycle and declined to approximately
60% after the third cycle, resulting in only a 33% decrease in activity. In another
study by Wang et al. [81], the catalyst demonstrated exceptional catalytic perfor-
mance, resulting in a remarkable 98.6% biodiesel yield, primarily attributed to its
superior total basicity. Additionally, their investigation into the catalyst’s reusability
revealed that it maintained favorable catalytic activity, yielding above 81% biodiesel
even after undergoing nine successive uses. This finding confirm that biochar catalyst
has great potential for biodiesel production.
In the presence of excess methanol, the esterification and transesterification reac-
tions used in biodiesel production exhibit pseudo-first-order kinetics, which permits
the neglect of the reverse process [84, 86]. The activation energies reported in the liter-
ature vary for different biochar catalysts when using various feedstocks, including
212 H. Zohmingliana et al.

sunflower oil (108.8 kJ/mol), waste frying oil (79–84 kJ/mol), jatropha oil (29.5 kJ/
mol), and canola oil (102–136 kJ/mol) [86]. Hence, the response rate (r) can be
expressed using Eq. 8.1:

d[O]
r=− = k[O] (8.1)
dt
where [O] stands for the oil concentration, t for the reaction time, and k for the
rate constant. The methyl ester conversion at time t was measured from Eq. 8.2 to
determine first-order rate constants. The Arrhenius equation (Eq. 8.3) was used to
determine the activation energy (E a ) for the transesterification processes using the
rate constants.

− ln(1 − X ) = kt (8.2)

Ea
ln k = − + ln P (8.3)
RT

8.6 Future Prospect

The expansion of the utilization of biomass as a catalyst source in the production


of biodiesel appears to be on the horizon. A multipurpose substance that can be
further improved for catalytic uses is biochar, a by-product of biomass breakdown.
As research advances, it is foreseeable that optimization of reaction parameters such
as temperature and duration will enhance the efficiency of biodiesel production,
utilizing functionalized biochar as a catalyst. This will result in processes that are
both economically feasible and environmentally sustainable. Functionalized biochar,
particularly acid-functionalized and basic-functionalized variants, shows potential as
a catalyst for chemical reactions. These materials can find applications in various
processes, including transesterification and esterification, contributing to the sustain-
able production of biodiesel and other chemicals. Further studies could uncover new
ways to improve its catalytic capabilities.

8.7 Conclusions

Biochar emerges as a suitable low-cost and sustainable catalyst and catalyst support
for chemical synthesis, biofuel production, energy storage, gas adsorption, etc.,
giving comparable or superior performance to the commercial one. In this review, we
discuss the biochar produced by pyrolysis and hydrothermal processes. The physic-
ochemical properties of the biochar depend on the feedstock, temperature, time, and
8 Functionalized Biochar for Green and Sustainable Production of Biodiesel 213

the process of formation. The surface area of the biochar is not high enough, and due
to the absence of a functional group, the biochar further needs to be activated and
functionalized to be used as a catalyst. The surface area and porosity of the biochar
can be enhanced to different extents depending on the activation methods employed.
Furthermore, depending on the intended application of the biochar, it can be further
modified with specific functional groups or substances to carry out particular tasks,
such as selective adsorption and catalysis. The biochars that have been activated and
functionalized demonstrate significant promise for serving as versatile catalysts or
catalyst supports in biomass upgrading. These biochars have been effectively utilized
in various applications such as carbohydrate hydrolysis and dehydration, biodiesel
production, biomass pyrolysis, gasification, and bio-oil upgrading. Further investi-
gation has to be done in developing the catalytic properties of biochar which will
be important to designing active, selective, and stable biochar catalysts. Further-
more, for biochar to be a substitute for industrial heterogeneous catalysts, it has to
be producible on an industrial scale.

References

1. Demirbaş, A. (2001). Biomass resource facilities and biomass conversion processing for fuels
and chemicals. Energy Conversion and Management, 42, 1357–1378. https://doi.org/10.1016/
S0196-8904(00)00137-0
2. Kunkes, E. L., Simonetti, D. A., West, R. M., Serrano-Ruiz, J. C., Gärtner, C. A., & Dumesic,
J. A. (2008). Catalytic conversion of biomass to monofunctional hydrocarbons and targeted
liquid-fuel classes. Science (80-.), 322, 417–421. https://doi.org/10.1126/SCIENCE.1159210/
SUPPL_FILE/KUNKES.SOM.PDF
3. Ruatpuia, J. V. L., Changmai, B., Pathak, A., Alghamdi, L. A., Kress, T., Halder, G., Wheatley,
A. E. H., & Rokhum, S. L. (2023). Green biodiesel production from Jatropha curcas oil using
a carbon-based solid acid catalyst: A process optimization study. Renewable Energy, 206,
597–608. https://doi.org/10.1016/J.RENENE.2023.02.041
4. Soffian, M. S., Abdul Halim, F. Z., Aziz, F., Rahman, M. A., Mohamed Amin, M. A., &
Awang Chee, D. N. (2022). Carbon-based material derived from biomass waste for wastew-
ater treatment. Environmental Advances, 9, 100259. https://doi.org/10.1016/J.ENVADV.2022.
100259
5. Zhou, C. H., Xia, X., Lin, C. X., Tong, D. S., & Beltramini, J. (2011). Catalytic conversion of
lignocellulosic biomass to fine chemicals and fuels. Chemical Society Reviews, 40, 5588–5617.
https://doi.org/10.1039/C1CS15124J
6. Ma, L., Wang, T., Liu, Q., Zhang, X., Ma, W., & Zhang, Q. (2012). A review of thermal–
chemical conversion of lignocellulosic biomass in China. Biotechnology Advances, 30, 859–
873. https://doi.org/10.1016/J.BIOTECHADV.2012.01.016
7. Sri Shalini, S., Palanivelu, K., Ramachandran, A., & Raghavan, V. (2020). Biochar from
biomass waste as a renewable carbon material for climate change mitigation in reducing green-
house gas emissions—A review. Biomass Conversion and Biorefinery, 115(11), 2247–2267.
https://doi.org/10.1007/S13399-020-00604-5
8. Cha, J. S., Park, S. H., Jung, S. C., Ryu, C., Jeon, J. K., Shin, M. C., & Park, Y. K. (2016). Produc-
tion and utilization of biochar: A review. Journal of Industrial and Engineering Chemistry, 40,
1–15. https://doi.org/10.1016/J.JIEC.2016.06.002
214 H. Zohmingliana et al.

9. Maroušek, J., Vochozka, M., Plachý, J., & Žák, J. (2017). Glory and misery of biochar. Clean
Technologies and Environmental Policy, 19, 311–317. https://doi.org/10.1007/S10098-016-
1284-Y/METRICS
10. Kołtowski, M., Charmas, B., Skubiszewska-Zięba, J., & Oleszczuk, P. (2017). Effect of biochar
activation by different methods on toxicity of soil contaminated by industrial activity. Ecotox-
icology and Environmental Safety, 136, 119–125. https://doi.org/10.1016/J.ECOENV.2016.
10.033
11. Nabais, J. M. V., Nunes, P., Carrott, P. J. M., Ribeiro Carrott, M. M. L., García, A. M., Díaz-Díez,
M. A. (2008). Production of activated carbons from coffee endocarp by CO2 and steam acti-
vation. Fuel Processing Technology, 89, 262–268. https://doi.org/10.1016/J.FUPROC.2007.
11.030
12. Uchimiya, M., Pignatello, J. J., White, J. C., Hu, S. L., & Ferreira, P. J. (2017). Surface
interactions between gold nanoparticles and biochar. Science Reports, 71(7), 1–9. https://doi.
org/10.1038/s41598-017-03916-1
13. Bhuiya, M. M. K., Rasul, M. G., Khan, M. M. K., Ashwath, N., Azad, A. K., & Hazrat, M.
A. (2014). Second generation biodiesel: potential alternative to-edible oil-derived biodiesel.
Energy Procedia, 61, 1969–1972. https://doi.org/10.1016/J.EGYPRO.2014.12.054
14. Meher, L. C., Vidya Sagar, D., & Naik, S. N. (2006). Technical aspects of biodiesel production
by transesterification—A review. Renewable and Sustainable Energy Reviews, 10, 248–268.
https://doi.org/10.1016/J.RSER.2004.09.002
15. Demirbas, A. (2008). Comparison of transesterification methods for production of biodiesel
from vegetable oils and fats. Energy Conversion and Management, 49, 125–130. https://doi.
org/10.1016/J.ENCONMAN.2007.05.002
16. Dehkhoda, A. M., West, A. H., & Ellis, N. (2010). Biochar based solid acid catalyst for biodiesel
production. Applied Catalysis, A: General, 382, 197–204. https://doi.org/10.1016/J.APCATA.
2010.04.051
17. Freedman, B., Pryde, E. H., & Mounts, T. L. (1984). Variables affecting the yields of fatty
esters from transesterified vegetable oils. Journal of the American Oil Chemists Society, 61,
1638–1643. https://doi.org/10.1007/BF02541649/METRICS
18. West, A. H., Posarac, D., & Ellis, N. (2008). Assessment of four biodiesel production processes
using HYSYS plant. Bioresource Technology, 99, 6587–6601. https://doi.org/10.1016/J.BIO
RTECH.2007.11.046
19. Saka, S., & Kusdiana, D. (2001). Biodiesel fuel from rapeseed oil as prepared in supercritical
methanol. Fuel, 80, 225–231. https://doi.org/10.1016/S0016-2361(00)00083-1
20. López, D. E., Goodwin, J. G., Bruce, D. A., & Lotero, E. (2005). Transesterification of triacetin
with methanol on solid acid and base catalysts. Applied Catalysis, A: General, 295, 97–105.
https://doi.org/10.1016/J.APCATA.2005.07.055
21. Zhang, Y., Dubé, M. A., McLean, D. D., & Kates, M. (2003). Biodiesel production from waste
cooking oil: 1. Process design and technological assessment. Bioresource Technology, 89, 1–16.
https://doi.org/10.1016/S0960-8524(03)00040-3
22. Lam, M. K., Lee, K. T., & Mohamed, A. R. (2010). Homogeneous, heterogeneous and enzy-
matic catalysis for transesterification of high free fatty acid oil (waste cooking oil) to biodiesel:
A review. Biotechnology Advances, 28, 500–518. https://doi.org/10.1016/J.BIOTECHADV.
2010.03.002
23. Canakci, M., & Van Gerpen, J. (1999). Biodiesel production via acid catalysis. Transactions
of the ASAE, 42, 1203–1210. https://doi.org/10.13031/2013.13285
24. Warabi, Y., Kusdiana, D., & Saka, S. (2004). Reactivity of triglycerides and fatty acids of
rapeseed oil in supercritical alcohols. Bioresource Technology, 91, 283–287. https://doi.org/10.
1016/S0960-8524(03)00202-5
25. Ruatpuia, J. V. L., Halder, G., Mohan, S., Gurunathan, B., Li, H., Chai, F., Basumatary, S., &
Lalthazuala, S. (2023). Microwave-assisted biodiesel production using ZIF-8 MOF-derived
nanocatalyst : A process optimization, kinetics, thermodynamics and life cycle cost analysis.
Energy Conversion and Management, 292, 117418. https://doi.org/10.1016/j.enconman.2023.
117418
8 Functionalized Biochar for Green and Sustainable Production of Biodiesel 215

26. Furuta, S., Matsuhashi, H., & Arata, K. (2004). Biodiesel fuel production with solid superacid
catalysis in fixed bed reactor under atmospheric pressure. Catalysis Communications, 5, 721–
723. https://doi.org/10.1016/J.CATCOM.2004.09.001
27. Velusamy, K. Devanand, J., Senthil Kumar, P., Soundarajan, K., Sivasubramanian, V., Sindhu,
J., & Vo, D. V. N. (2021). A review on nano-catalysts and biochar-based catalysts for biofuel
production. Fuel, 306, 121632. https://doi.org/10.1016/J.FUEL.2021.121632
28. Jha, S., Nanda, S., Acharya, B., & Dalai, A. K. (2022). A review of thermochemical conversion
of waste biomass to biofuels. Energies, 15, 6352. https://doi.org/10.3390/EN15176352
29. Han, T. U., Kim, Y. M., Watanabe, A., Teramae, N., Park, Y. K., & Kim, S. (2017). Pyrolysis
kinetic analysis of poly(methyl methacrylate) using evolved gas analysis-mass spectrometry.
Korean Journal of Chemical Engineering, 34, 1214–1221. https://doi.org/10.1007/S11814-
016-0354-5/METRICS
30. Heidari, A., Stahl, R., Younesi, H., Rashidi, A., Troeger, N., & Ghoreyshi, A. A. (2014). Effect
of process conditions on product yield and composition of fast pyrolysis of Eucalyptus grandis
in fluidized bed reactor. Journal of Industrial and Engineering Chemistry, 20, 2594–2602.
https://doi.org/10.1016/J.JIEC.2013.10.046
31. Wang, Z., Burra, K. G., Lei, T., & Gupta, A. K. (2021). Co-pyrolysis of waste plastic and solid
biomass for synergistic production of biofuels and chemicals—A review. Progress in Energy
and Combustion Science, 84, 100899. https://doi.org/10.1016/J.PECS.2020.100899
32. Ronsse, F., van Hecke, S., Dickinson, D., & Prins, W. (2013). Production and characterization of
slow pyrolysis biochar: Influence of feedstock type and pyrolysis conditions. GCB Bioenergy,
5, 104–115. https://doi.org/10.1111/GCBB.12018
33. Bridgwater, A. V. (2012). Review of fast pyrolysis of biomass and product upgrading. Biomass
and Bioenergy, 38, 68–94. https://doi.org/10.1016/J.BIOMBIOE.2011.01.048
34. Williams, P. T., & Besler, S. (1996). The influence of temperature and heating rate on the
slow pyrolysis of biomass. Renewable Energy, 7, 233–250. https://doi.org/10.1016/0960-148
1(96)00006-7
35. Bridgwater, A. V., Meier, D., & Radlein, D. (1999). An overview of fast pyrolysis of biomass.
Organic Geochemistry, 30, 1479–1493. https://doi.org/10.1016/S0146-6380(99)00120-5
36. Wagner, A., & Kaupenjohann, M. (2014). Suitability of biochars (pyro- and hydrochars) for
metal immobilization on former sewage-field soils. European Journal of Soil Science, 65,
139–148. https://doi.org/10.1111/EJSS.12090
37. Demirbaş, A. (2000). Mechanisms of liquefaction and pyrolysis reactions of biomass. Energy
Conversion and Management, 41, 633–646. https://doi.org/10.1016/S0196-8904(99)00130-2
38. kyu Choi, M., Park, H. C., & Choi, H. S. (2018). Comprehensive evaluation of various pyrolysis
reaction mechanisms for pyrolysis process simulation. Chemical Engineering and Processing:
Process Intensification, 130, 19–35. https://doi.org/10.1016/J.CEP.2018.05.011
39. Lindfors, C., Kuoppala, E., Oasmaa, A., Solantausta, Y., & Arpiainen, V. (2014). Fractionation
of bio-oil. Energy and Fuels, 28, 5785–5791. https://doi.org/10.1021/EF500754D
40. Yu, J., Lucas, J. A., & Wall, T. F. (2007). Formation of the structure of chars during devolatiliza-
tion of pulverized coal and its thermoproperties: A review. Progress in Energy and Combustion
Science, 33, 135–170. https://doi.org/10.1016/J.PECS.2006.07.003
41. Wang, T., Zhai, Y., Zhu, Y., Li, C., & Zeng, G. (2018). A review of the hydrothermal carboniza-
tion of biomass waste for hydrochar formation: Process conditions, fundamentals, and physic-
ochemical properties. Renewable and Sustainable Energy Reviews, 90, 223–247. https://doi.
org/10.1016/J.RSER.2018.03.071
42. Liu, W. J., Jiang, H., & Yu, H. Q. (2015). Development of biochar-based functional materials:
Toward a sustainable platform carbon material. Chemical Reviews, 115, 12251–12285. https://
doi.org/10.1021/acs.chemrev.5b00195
43. Tomczyk, A., Sokołowska, Z., & Boguta, P. (2020). Biochar physicochemical properties:
Pyrolysis temperature and feedstock kind effects. Reviews in Environmental Science &
Biotechnology, 19, 191–215. https://doi.org/10.1007/S11157-020-09523-3/TABLES/3
44. Cao, X., Sun, S., & Sun, R. (2017). Application of biochar-based catalysts in biomass upgrading:
A review. RSC Advances, 7, 48793–48805. https://doi.org/10.1039/C7RA09307A
216 H. Zohmingliana et al.

45. Kang, S., Li, X., Fan, J., & Chang, J. (2013). Hydrothermal conversion of lignin: A review.
Renewable and Sustainable Energy Reviews, 27, 546–558. https://doi.org/10.1016/J.RSER.
2013.07.013
46. Carpenter, D., Westover, T. L., Czernik, S., & Jablonski, W. (2014). Biomass feedstocks for
renewable fuel production: A review of the impacts of feedstock and pretreatment on the yield
and product distribution of fast pyrolysis bio-oils and vapors. Green Chemistry, 16, 384–406.
https://doi.org/10.1039/C3GC41631C
47. Tekin, K., Karagöz, S., & Bektaş, S. (2014). A review of hydrothermal biomass processing.
Renewable and Sustainable Energy Reviews, 40, 673–687. https://doi.org/10.1016/J.RSER.
2014.07.216
48. Khan, T. A., Saud, A. S., Jamari, S. S., Rahim, M. H. A., Park, J. W., & Kim, H. J. (2019).
Hydrothermal carbonization of lignocellulosic biomass for carbon rich material preparation:
A review. Biomass and Bioenergy, 130, 105384. https://doi.org/10.1016/J.BIOMBIOE.2019.
105384
49. Lucian, M., & Fiori, L. (2017). Hydrothermal carbonization of waste biomass: process design,
modeling, energy efficiency and cost analysis. Energies, 10, 211. https://doi.org/10.3390/EN1
0020211
50. Cao, Y., He, M., Dutta, S., Luo, G., Zhang, S., & Tsang, D. C. W. (2021). Hydrothermal
carbonization and liquefaction for sustainable production of hydrochar and aromatics. Renew-
able and Sustainable Energy Reviews, 152, 111722. https://doi.org/10.1016/J.RSER.2021.
111722
51. Xiao, L. P., Shi, Z. J., Xu, F., & Sun, R. C. (2012). Hydrothermal carbonization of lignocellulosic
biomass. Bioresource Technology, 118, 619–623. https://doi.org/10.1016/J.BIORTECH.2012.
05.060
52. Román, S., Nabais, J. M. V., Laginhas, C., Ledesma, B., & González, J. F. (2012). Hydrothermal
carbonization as an effective way of densifying the energy content of biomass. Fuel Processing
Technology, 103, 78–83. https://doi.org/10.1016/J.FUPROC.2011.11.009
53. Chan, Y. H., Yusup, S., Quitain, A. T., Uemura, Y., & Sasaki, M. (2014). Bio-oil production
from oil palm biomass via subcritical and supercritical hydrothermal liquefaction. Journal of
Supercritical Fluids, 95, 407–412. https://doi.org/10.1016/J.SUPFLU.2014.10.014
54. Pinheiro Pires, A. P., Arauzo, J., Fonts, I., Domine, M. E., Fernández Arroyo, A., Garcia-
Perez, M. E., Montoya, J., Chejne, F., Pfromm, P., Garcia-Perez, M. (2019). Challenges and
opportunities for bio-oil refining: A review. Energy & Fuels, 33, 4683–4720. https://doi.org/
10.1021/ACS.ENERGYFUELS.9B00039
55. Kruse, A. (2009). Hydrothermal biomass gasification. Journal of Supercritical Fluids, 47,
391–399. https://doi.org/10.1016/J.SUPFLU.2008.10.009
56. Funke, A., & Ziegler, F. (2010). Hydrothermal carbonization of biomass: A summary and
discussion of chemical mechanisms for process engineering. Biofuels, Bioproducts and
Biorefining, 4, 160–177. https://doi.org/10.1002/BBB.198
57. Coronella, C. J., Lynam, J. G., Reza, M. T., & Uddin, M. H. (2014). Hydrothermal carbonization
of lignocellulosic biomass (pp. 275–311). https://doi.org/10.1007/978-3-642-54458-3_12
58. Yousuf, A., Pirozzi, D., & Sannino, F. (2020). Fundamentals of lignocellulosic biomass. Ligno-
cellulosic biomass to liquid biofuels (pp. 1–15). https://doi.org/10.1016/B978-0-12-815936-1.
00001-0
59. Alfattani, R., Shah, M. A., Siddiqui, M. I. H., Ali, M. A., & Alnaser, I. A. (2021). Bio-char
characterization produced from walnut shell biomass through slow pyrolysis: sustainable for
soil amendment and an alternate bio-fuel. Energies, 15, 1. https://doi.org/10.3390/EN15010001
60. Maj, I. (2022). Significance and challenges of poultry litter and cattle manure as sustainable
fuels: A review. Energies, 15, 8981. https://doi.org/10.3390/EN15238981
61. Singh, A., Nanda, S., & Berruti, F. (2020). A review of thermochemical and biochemical
conversion of Miscanthus to biofuels. In Biorefinery of alternative resources: targeting green
fuels and platform chemicals (pp. 195–220). https://doi.org/10.1007/978-981-15-1804-1_9/
COVER
8 Functionalized Biochar for Green and Sustainable Production of Biodiesel 217

62. O’Connor, J., Hoang, S. A., Bradney, L., Dutta, S., Xiong, X., Tsang, D. C. W., Ramadass, K.,
Vinu, A., Kirkham, M. B., & Bolan, N. S. (2021). A review on the valorisation of food waste
as a nutrient source and soil amendment. Environmental Pollution, 272, 115985. https://doi.
org/10.1016/J.ENVPOL.2020.115985
63. Singh, P., Sharma, S., & Dhanorkar, M. (2022). Aquatic plant biomass-derived porous carbon:
Biomaterials for sustainable waste management and climate change mitigation. International
Journal of Environmental Science and Technology, 2022, 1–16. https://doi.org/10.1007/S13
762-022-04601-1
64. Ahmed, M. B., Zhou, J. L., Ngo, H. H., Guo, W., & Chen, M. (2016). Progress in the prepa-
ration and application of modified biochar for improved contaminant removal from water
and wastewater. Bioresource Technology, 214, 836–851. https://doi.org/10.1016/J.BIORTECH.
2016.05.057
65. Ghosh, N., Rhithuparna, D., Khatoon, R., Rokhum, S. L., & Halder, G. (2023). Sulphonic-acid
functionalized novel Limonia acidissima carbonaceous catalyst for biodiesel synthesis from
Millettia pinnata oil: optimization, kinetics, thermodynamics and cost analysis. Journal of
Cleaner Production, 394, 136362. https://doi.org/10.1016/J.JCLEPRO.2023.136362
66. Rodríguez-Reinoso, F., Molina-Sabio, M., & González, M. T. (1995). The use of steam and
CO2 as activating agents in the preparation of activated carbons. Carbon N. Y., 33, 15–23.
https://doi.org/10.1016/0008-6223(94)00100-E
67. Yorgun, S., Vural, N., & Demiral, H. (2009). Preparation of high-surface area activated carbons
from Paulownia wood by ZnCl2 activation. Microporous and Mesoporous Materials, 122,
189–194. https://doi.org/10.1016/J.MICROMESO.2009.02.032
68. Angin, D., Altintig, E., & Köse, T. E. (2013). Influence of process parameters on the surface
and chemical properties of activated carbon obtained from biochar by chemical activation.
Bioresource Technology, 148, 542–549. https://doi.org/10.1016/J.BIORTECH.2013.08.164
69. Azargohar, R., & Dalai, A. K. (2008). Steam and KOH activation of biochar: experimental and
modeling studies. Microporous and Mesoporous Materials, 110, 413–421. https://doi.org/10.
1016/J.MICROMESO.2007.06.047
70. Ros, A., Lillo-Ródenas, M. A., Fuente, E., Montes-Morán, M. A., Martín, M. J., & Linares-
Solano, A. (2006). High surface area materials prepared from sewage sludge-based precursors.
Chemosphere, 65, 132–140. https://doi.org/10.1016/J.CHEMOSPHERE.2006.02.017
71. Park, H., Kim, J., Lee, Y. G., & Chon, K. (2021). Enhanced adsorptive removal of dyes using
mandarin peel biochars via chemical activation with NH4 Cl and ZnCl2 . Water (Switzerland),
13, 1495. https://doi.org/10.3390/W13111495/S1
72. Park, J. H., Ok, Y. S., Kim, S. H., Cho, J. S., Heo, J. S., Delaune, R. D., & Seo, D. C. (2015).
Evaluation of phosphorus adsorption capacity of sesame straw biochar on aqueous solution:
Influence of activation methods and pyrolysis temperatures. Environmental Geochemistry and
Health, 37, 969–983. https://doi.org/10.1007/S10653-015-9709-9/TABLES/2
73. Tian, S., Yang, R., Pan, Z., Su, X., Li, S., Wang, P., & Huang, X. (2022). Anisotropic reed-stem-
derived hierarchical porous biochars supported paraffin wax for efficient solar-thermal energy
conversion and storage. Journal of Energy Storage, 56, 106153. https://doi.org/10.1016/J.EST.
2022.106153
74. Venkatachalam, C. D., Sekar, S., Sengottian, M., Ravichandran, S. R., & Bhuvaneshwaran, P.
(2023). A critical review of the production, activation, and morphological characteristic study
on functionalized biochar. Journal of Energy Storage, 67, 107525. https://doi.org/10.1016/J.
EST.2023.107525
75. Mullen, C. A., Boateng, A. A., Goldberg, N. M., Lima, I. M., Laird, D. A., & Hicks, K. B.
(2010). Bio-oil and bio-char production from corn cobs and stover by fast pyrolysis. Biomass
and Bioenergy, 34, 67–74. https://doi.org/10.1016/J.BIOMBIOE.2009.09.012
76. Deng, A., Lin, Q., Yan, Y., Li, H., Ren, J., Liu, C., & Sun, R. (2016). A feasible process
for furfural production from the pre-hydrolysis liquor of corncob via biochar catalysts in a
new biphasic system. Bioresource Technology, 216, 754–760. https://doi.org/10.1016/J.BIO
RTECH.2016.06.002
218 H. Zohmingliana et al.

77. Lee, J., Kim, K. H., & Kwon, E. E. (2017). Biochar as a catalyst. Renewable and Sustainable
Energy Reviews, 77, 70–79. https://doi.org/10.1016/J.RSER.2017.04.002
78. Kastner, J. R., Miller, J., Geller, D. P., Locklin, J., Keith, L. H., & Johnson, T. (2012). Catalytic
esterification of fatty acids using solid acid catalysts generated from biochar and activated
carbon. Catalysis Today, 190, 122–132. https://doi.org/10.1016/J.CATTOD.2012.02.006
79. Konwar, L. J., Boro, J., & Deka, D. (2014). Review on latest developments in biodiesel produc-
tion using carbon-based catalysts. Renewable and Sustainable Energy Reviews, 29, 546–564.
https://doi.org/10.1016/J.RSER.2013.09.003
80. Rafi, J. M., Rajashekar, A., Srinivas, M., Rao, B. V. S. K., Prasad, R. B. N., & Lingaiah, N.
(2015). Esterification of glycerol over a solid acid biochar catalyst derived from waste biomass.
RSC Advances, 5, 44550–44556. https://doi.org/10.1039/C5RA06613A
81. Wang, S., Zhao, C., Shan, R., Wang, Y., & Yuan, H. (2017). A novel peat biochar supported
catalyst for the transesterification reaction. Energy Conversion and Management, 139, 89–96.
https://doi.org/10.1016/J.ENCONMAN.2017.02.039
82. Balajii, M., & Niju, S. (2019). Biochar-derived heterogeneous catalysts for biodiesel produc-
tion. Environmental Chemistry Letters, 17, 1447–1469. https://doi.org/10.1007/S10311-019-
00885-X/TABLES/10
83. Li, M., Zheng, Y., Chen, Y., & Zhu, X. (2014). Biodiesel production from waste cooking
oil using a heterogeneous catalyst from pyrolyzed rice husk. Bioresource Technology, 154,
345–348. https://doi.org/10.1016/J.BIORTECH.2013.12.070
84. Rokhum, S. L., Changmai, B., Kress, T., & Wheatley, A. E. H. (2022). A one-pot route to
tunable sugar-derived sulfonated carbon catalysts for sustainable production of biodiesel by
fatty acid esterification. Renewable Energy, 184. https://doi.org/10.1016/j.renene.2021.12.001
85. González, M. E., Cea, M., Reyes, D., Romero-Hermoso, L., Hidalgo, P., Meier, S., Benito, N., &
Navia, R. (2017). Functionalization of biochar derived from lignocellulosic biomass using
microwave technology for catalytic application in biodiesel production. Energy Conversion
and Management, 137, 165–173. https://doi.org/10.1016/J.ENCONMAN.2017.01.063
86. Kostić, M. D., Bazargan, A., Stamenković, O. S., Veljković, V. B., & McKay, G. (2016).
Optimization and kinetics of sunflower oil methanolysis catalyzed by calcium oxide-based
catalyst derived from palm kernel shell biochar. Fuel, 163, 304–313. https://doi.org/10.1016/
J.FUEL.2015.09.042
87. Bazargan, A., Kostić, M. D., Stamenković, O. S., Veljković, V. B., & McKay, G. (2015). A
calcium oxide-based catalyst derived from palm kernel shell gasification residues for biodiesel
production. Fuel, 150, 519–525. https://doi.org/10.1016/J.FUEL.2015.02.046
88. Dhawane, S. H., Kumar, T., & Halder, G. (2015). Central composite design approach towards
optimization of flamboyant pods derived steam activated carbon for its use as heterogeneous
catalyst in transesterification of Hevea brasiliensis oil. Energy Conversion and Management,
100, 277–287. https://doi.org/10.1016/j.enconman.2015.04.083
89. Yu, J. T., Dehkhoda, A. M., & Ellis, N. (2010). Development of biochar-based catalyst for
transesterification of canola oil. Energy and Fuels, 25, 337–344. https://doi.org/10.1021/EF1
00977D
90. Li, M., Chen, D., & Zhu, X. (2013). Preparation of solid acid catalyst from rice husk char
and its catalytic performance in esterification. Chinese Journal of Catalysis, 34, 1674–1682.
https://doi.org/10.1016/S1872-2067(12)60634-2
91. Liu, T., Li, Z., Li, W., Shi, C., & Wang, Y. (2013). Preparation and characterization of biomass
carbon-based solid acid catalyst for the esterification of oleic acid with methanol. Bioresource
Technology, 133, 618–621. https://doi.org/10.1016/J.BIORTECH.2013.01.163
92. Lee, J., Jung, J. M., Oh, J. I., Ok, Y. S., Lee, S. R., & Kwon, E. E. (2017). Evaluating the effective-
ness of various biochars as porous media for biodiesel synthesis via pseudo-catalytic transes-
terification. Bioresource Technology, 231, 59–64. https://doi.org/10.1016/J.BIORTECH.2017.
01.067
Chapter 9
Versatile Pretreatment Approaches
to Improve the Bioethanol Production
from Various Biomass Feedstocks

R. Rajesh Kannan, V. Saravanan, M. Rajasimman,


Panchamoorthy Saravanan, and Gurunathan Baskar

Abstract The result in climate change, fossil resource depletion, and transportation
alternatives to petroleum are in great demand. The structural components belonging
to lignocellulosic biomass like lignin, cellulose, hemicellulose and as well as tech-
nological unit phases like pretreatment and as well as hydrolysis of enzymes are
discussed. The sole purpose of pretreatment stage is to increase the surface area of
carbohydrate that is available for the enzymatic saccharification and also lower the
inhibitor concentrations. The review focuses on bioethanol production from lignocel-
lulosic biomass derived from plants. For synthesizing effective bioethanol, diverse
lignocellulosic biomass like feedstocks of wood, agricultural wastes, and marine
algae are used as substrates, and different enzyme techniques are used in hydrol-
ysis protocols. This paper provides a detailed review in the synthesizing of ligno-
cellulosic bioethanol via biochemical pathway while focusing on the widely used
pretreatment technologies as well as important enzymatic hydrolysis and fermenta-
tion operational parameters in accordance to the yield of sugar and ethanol. Also, the
importance of immobilization is discussed in this study as well as numerous detox-
ification processes for eliminating hazardous substances. According to this review,
modified microbes could potentially be employed to boost fermentation of glucose
and xylose, cellulolytic enzyme synthesis and also stress response.

Keywords Bioethanol · Pretreatment techniques · Lignocellulosic biomass ·


Sugar · Starch

R. Rajesh Kannan · V. Saravanan (B) · M. Rajasimman


Department of Chemical Engineering, Annamalai University, Chidambaram 608002, Tamil Nadu,
India
e-mail: sarav304@gmail.com
P. Saravanan
Department of Petrochemical Technology, UCE—BIT Campus, Anna University,
Tiruchirappalli 620024, Tamil Nadu, India
G. Baskar
Department of Biotechnology, St. Joseph’s College of Engineering, Chennai, Tamil Nadu, India

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2024 219
G. Baskar et al. (eds.), Circular Bioeconomy Perspectives in Sustainable Bioenergy
Production, Energy, Environment, and Sustainability,
https://doi.org/10.1007/978-981-97-2523-6_9
220 R. Rajesh Kannan et al.

9.1 Introduction

Renewable and sustainable energy sources which include bioethanol, biodiesel,


biohydrogen, and methane are obtained from various raw resources to address the
energy and environmental issues. Based on the raw materials used from a scien-
tific standpoint, all biofuels are divided into four types [1]. Food-based agricultural
waste namely starch (corn, wheat, cassava, sugarcane) yields biofuels which are
categorized as first-generation biofuels since they are in competence with arable
land as well as feed supply [2]. These generation biofuels will most certainly consti-
tute a significant share of worldwide biofuel production in 2021. In the coming
years, second-generation biofuels which are based on lignocellulosic-based mate-
rials could be a potential fuel supplement. Biofuels of the next generation include
algal biofuels. Fourth generation of biofuels is made from genetically modified algae
having high lipid content [3]. In 2023, global bioethanol output is predicted to reach
150 billion litres, from roughly 120 billion litres in 2019. The great majority of
global bioethanol production is from the United States. Due to the biofuel policies of
both the United States’ and the European Union’s which stimulate the development
of cellulose-based biofuels worldwide, the cellulosic bioethanol commercialization
from agricultural wastes is predicted to enhance second-involvement generations in
the global bioethanol market [4–8].

9.1.1 First-Generation Bioethanol

First-generation bioethanol is a liquid biofuel generated from high-starch, high-


sugar agricultural crops that are intended for use in automobiles [9]. Because
Saccharomyces cerevisiae cannot break down complex carbohydrates, both starchy
and lignocellulosic components must be hydrolyzed [10]. The synthesis of first-
generation bioethanol and its raw materials are shown in Fig. 9.1. As a result, there
are severe worries about large-scale production’s social and environmental effects
[11]. The production of first-generation biofuel demands for water and arable land
with food production due to over fertilization and may lead to depletion of neces-
sary resources which includes shortage of water availability and degradation of soil
and water resources [12]. Also, food prices for grains, crops, vegetable oils, and
livestock feed may rise as a result. Furthermore, emission of greenhouse gases is
lower for biofuels than fossil fuel combustion [13]. Biodiesel which is created by
transesterifying vegetable oils, aside from food-based bioethanol, is also another
first-generation biofuel [9].
9 Versatile Pretreatment Approaches to Improve the Bioethanol … 221

Fig. 9.1 Raw materials used


for first generation of
bioethanol production

9.1.2 Second-Generation Bioethanol

The biofuels of second and later generations including bioethanol do not compete
with feed supply because they are created from non-food basic elements [14]. The
primary feedstock used for the production of second-generation bioethanol is ligno-
cellulosic biomass, however, industrial wastes like whey [10] or crude glycerol can
also be employed. In most cases, such biomass is both cheap and plentiful in a given
location [15–20]. The Lignocellulose is both a renewable and sustainable carbon
source that can be found in a wide range of raw materials from plants [16, 21–28] as
given in Fig. 9.2. The availability of amount of lignocellulosic biomass is predeter-
mined by climate conditions. Converting the lignocellulose to reducing sugars is the
most challenging compared to starch conversion [29–36]. Researchers have inves-
tigated the usage of many varieties of plant biomass in the production of biofuels
(Fig. 9.3). The biofuels of second and later generations which includes bioethanol
do not compete with the feed supply because they are created from non-feed basic
elements [14].

Fig. 9.2 The first-generation bioethanol production process


222 R. Rajesh Kannan et al.

Fig. 9.3 Second-generation bioethanol manufacturing procedures

9.1.3 Third-Generation Bioethanol

The production of microalgae or unicellular microorganisms (such as Cyanidium,


Caldarium, or Synechococcus) belonging to eukaryotes and prokaryotes (cyanobac-
teria) is the third generation of biofuels [36]. Active microalgal biomass such as
live biocatalysts can use nutrients namely carbon, nitrogen, phosphate, or sulphur
from industrial effluents as substrates to produce high biomass concentrations. Indus-
trial power plant effluent gases, wastewater, organic waste hydrolysis products, and
digestate are among the waste streams (waste from biogas production). As a result,
manufacturing third-generation biofuels can aid in the reduction of waste streams
from a wide range of industries. Algae, for example, are an excellent alternative fuel
because of their high lipid and carbohydrate content. By alkaline transesterification
with methanol, microalgae oils provide a possible raw material to vegetable oils
commonly used in the production of biodiesel [34]. Technology developments such
as enhanced oil extraction from cellular biomass are required for the manufacturing of
biodiesel from microalgae. Microalgae residual biomass can be converted into envi-
ronmentally friendly and energy carriers for a long-term (biofuels) like bioethanol
and biogas as well as used for fermentation of methane and biohydrogen gener-
ation [36]. Anaerobic fermentation of starch as well as starch-like polysaccharides
produces bioethanol from algal biomass. Starch is a conserved element in microalgae
like Cryptophyta and Chlorophyta [36]. Despite the process’s intricacy, yields are
produced at a smaller rate for producing bioethanol from biomass. The by-product
biohydrogen is a subsidiary of production of methane from manufactured organic
acids during the acidogenic phase of anaerobic digestion from organic waste transfor-
mation [36]. Nevertheless, due to the biohydrogen amount manufactured by algae is
currently restricted, the biohydrogen manufacturing process must be enhanced [36].
Algal cultivation is straightforward and can be done in various water environments
[37].
The raw materials of bioethanol can be made from various origins of biomass.
They usually fall into three categories as depicted in Fig. 9.4. The availability of
feedstock containing sugar and starch (first generation) is, however, affected by their
9 Versatile Pretreatment Approaches to Improve the Bioethanol … 223

Fig. 9.4 Raw materials utilized in the manufacturing of bioethanol

usage as feed. A possible feedstock for the manufacture of bioethanol is Lignocel-


lulosic biomass (second generation), due to its low capital, availability, and large
dispersion. They are also not competitive with crops used for feed [27].

9.2 Raw Materials for Bioethanol Production

9.2.1 Sugar-Containing Raw Materials

Sugar cane and beet are the two most appropriate sugar-producing plants on earth.
Sugar cane is responsible for global sugar output, while sugar beet is responsible for
one-third of left [28] as given in Table 9.1. Most Saccharomyces species generate
the enzyme invertase, which can efficiently hydrolyze them. As a result, bioethanol
synthesis from sugar (sucrose) feedstocks does not require pretreatment, making
it more practical than starch feedstocks [28]. For sugar extraction, sugar crops are
needed to be milled and then transferred to a fermentation medium, where ethanol
is extracted directly from the juice or molasses [29]. Table 9.1 shows the major
achievement and investigation of ethanol production from various sources of sugar-
containing raw materials and its composition.

9.2.2 Starch-Containing Raw Materials

The raw materials from starch for production of bioethanol are Cassava tubers
comprised of a mass of 80% carbohydrate content and mass protein content
of less than 1%. To process cassava tubers for bioethanol production, phases
like dusting, flaking, fragmentation, and aeration are included. After that, the
bioethanol is processed from dried cassava chips [5]. Linear (amylose) and branching
(amylopectin) polyglucans both acting as a mixture makes up starch. -amylase is a
starch hydrolase enzyme that works on the -1, 4 but not the -1, 6 links in amylopectin
[37]. The hydrolysis of starch (mostly by -amylase and glucoamylase) into glucose
224 R. Rajesh Kannan et al.

Table 9.1 Sugar-containing raw material for ethanol production and its chemical composition
Name of the Major Major Freesugar Moisture References
crops investigation achievements (%) (%)
Sugarcane Juices were The highest 12–17.6 68.0–82.4 [31, 32]
(Saccharum studied (i) ethanol
officinarum) without adding concentration
supplement, (ii) (39.4–42.1 g/L)
with addition of was found in (iii),
0.5% yeast while the lowest
extract, and (iii) was 11.0 g/L and
with addition of found in (i)
yeast extract,
thiamine, and
micronutrients
Enrichment Isolation and
technique was selection of
applied to isolate thermo tolerant
thermo tolerant yeast strain that
yeast from cane produced high
juice for ethanol concentrations of
production at ethanol at both 40
elevated and 45 °C from
temperature the juice
Juices were Higher (30.0%)
supplemented and ethanol was found
cells were with adapted cells
adapted to than non-adapted
galactose medium cells
Sugarbeet Juices were Ethanol 16.5 82.6 [8, 33]
(Beta ultrafiltered and concentration was
vulgaris) were found to be
supplemented 85.0–87.0 g/L
with mineral salt The highest
Flocculating and ethanol yield
non-flocculating (0.44 g/g) with the
yeasts along with lowest
Z. mobilis were productivity
used through (0.08 g/h/L) was
immobilization found by Z.
on loofa sponge mobilis
(continued)
9 Versatile Pretreatment Approaches to Improve the Bioethanol … 225

Table 9.1 (continued)


Name of the Major Major Freesugar Moisture References
crops investigation achievements (%) (%)
Sweet Five different Keller variety 16–21.8 78.2
sorghum genetic varieties produced the
(Sorghum (Keller, BJ 248, highest ethanol
bicolor) SSV84, Wray, (9.0%,w/v)
and NSSH104)
were investigated
for ethanol
production
Watermelon Juice was used as As much as 25.0% 70–10 90–93
(Citrullus diluent, additive, w/v sugar was
lanatus) and nitrogen fermented at
source for pH3.0 (ethanol
processed sugar yield was
0.41–0.46 g/g) or
up to 35.0%w/v
sugar at pH 5.0
with an ethanol
yield of
0.36–0.41 g/g

syrup is then converted to ethanol by yeast Saccharomyces cerevisiae, required for the
production of bioethanol from starch-containing feedstocks. When compared with
bioethanol synthesis from sugar-containing feedstocks, this step is more expensive
[38]. Bacteria such as Bacillus licheniformis as well as genetically altered organisms
produce α-amylase and glucoamylase [39, 40].
The Saccharomyces cerevisiae yeast transforms glucose to ethanol under anaer-
obic circumstances. In terms of mass conversion, the efficiency is about 51%, in the
highest yield of glucose into ethanol. Glucose is needed for yeast for the development
of cell and production of other metabolic products, and its efficiency in optimum
conversion is also reduced. In practise, glucose is converted to ethanol at a mass
conversion rate of 40–48% [41]. Starch-derived ethanol enhances enzyme applica-
tion and yeast strains with high ethanol tolerance when compared to sugar-derived
ethanol [42].

9.2.3 Lignocellulose-Containing Raw Materials

Due to being renewable and non-competitive with food crops, bioethanol production
from lignocellulosic biomass is more appealing and sustainable. Additionally, the
acquisition of usage of bioethanol from lignocellulosic biomass is associated with
a large decrease in greenhouse gas emissions [44]. Unlike fossil fuels, lignocellu-
losic biomass is uniformly spread throughout the world, assuring supply security
226 R. Rajesh Kannan et al.

Fig. 9.5 Raw materials for bioethanol production comprise lignocellulosic materials

by utilizing local energy sources [38]. It can also be created using a wide variety of
wastes or harvested from the forest and is often less expensive compared to feedstocks
including sugar or starch, which necessitates a comprehensive agricultural breeding
approach [3, 45]. Many different types of raw materials containing lignocellulosic
components are depicted in Fig. 9.5.
Lignocellulosic biomass has 43% cellulose, 27% lignin, 20% hemicellulose, and
10% of other constituents when calculated for an average [3]. Figure 9.5 depicts
the procedures where lignocellulosic biomass is required in the manufacture of
bioethanol. Because lignocellulosic biomass has a more heterogeneous structure
than the homogeneous and similar raw materials which is used in chemical industry,
it needs more sophisticated chemical operations [46, 47]. Furthermore, lignocellu-
losic crop harvesting is often not possible all year, making biomass suppliers’ duties
more difficult [48–50]. As a result, this problem must be solved by the stabilization
of biomass in such a way, especially for the availability of long-term storage and
sustaining continuous biorefinery operation for the coming years [51–56].

9.3 Pretreatment Overview

Lignocellulosic biomass, such as wood, grass, agricultural, and forest residues,


are potential resources for the production of bioethanol. The current biochemical
process of converting biomass to bioethanol typically consists of three main steps:
pretreatment, enzymatic hydrolysis, and fermentation. For this process, pretreat-
ment is probably the most crucial step since it has a large impact on the efficiency
of the overall bioconversion. The aim of pretreatment is to disrupt recalcitrant struc-
tures of cellulosic biomass to make cellulose more accessible to the enzymes that
convert carbohydrate polymers into fermentable sugars [57–59]. Different pretreat-
ment methods, including dilute acid pretreatment, steam explosion pretreatment,
organosolv, liquid hot water, ammonia fibre expansion, soaking in aqueous ammonia,
sodium hydroxide/lime pretreatments, and ozonolysis are intensively introduced
and discussed. In this minireview, the key points are focused on the structural
changes primarily in cellulose, hemicellulose, and lignin during the above-leading
pretreatment technologies [60–68].
9 Versatile Pretreatment Approaches to Improve the Bioethanol … 227

The pretreatment methods can be classified into different categories according to


various criteria [69–75]. It is common to differentiate pretreatment methods that are
efficient at high, low, or neutral pH. Basically, low pH methods require addition of
acids to increase the hydrolytic capacity while higher pH methods need pH-adjusting
agents such as sodium hydroxide or ammonia. The neutral pH methods, mainly liquid
hot water (LHW) pretreatment, simply apply water in the process. However, the
substrate medium in the neutral methods is weakly acidic due to the release of organic
acids from the biomass during the pretreatments. Another way to do the grouping is
based on the principal mechanism acting during pretreatment. As such, the methods
can be classified as physical, physicochemical, and chemical pretreatments. In this
review, mainly the first classification is used. The acidic pretreatments include dilute
acid pretreatment, steam explosion pretreatment, and organosolv pretreatment. The
neutral pretreatment is liquid hot water, and the sodium hydroxide/lime pretreat-
ment and ammonia fibre expansion are classified into the alkaline pretreatments
(Fig. 9.6). Besides the above methods, pretreatment of biomass with high-pressure
ozone (ozonolysis) was recently developed in bioethanol production [70–79]. In this
review, a brief process description is first given with recent developments for each
pretreatment, followed by discussion of the technology’s advantages and disadvan-
tages. Different feedstocks, including herbaceous/agricultural residues, hardwood
and softwood biomasses, applied in these technologies are highlighted (Figs. 9.7
and 9.8). The key points are focused on the structural changes primarily in cellulose,
hemicellulose, and lignin during the above leading pretreatment technologies [6–10].

Fig. 9.6 The benefits and drawbacks of pretreatment of lignocellulose-containing raw materials
for bioethanol production
228 R. Rajesh Kannan et al.

Fig. 9.7 Chemical treatment methods for bioethanol production

Fig. 9.8 Physicochemical processing types of lignocellulose-containing raw materials

9.3.1 Biological Processing of Lignocellulose-Containing


Raw Materials

Comparing other existing pretreatments, biological pretreatments are regarded as


illustrated in Fig. 9.9.
The employment of enzymes to hydrolyze raw lignocellulosic materials is also part
of biological pretreatment. The rate at which lignocellulosic feedstock hydrolyzes
into fermentable sugars determines the overall bioprocess efficiency [44]. Cellu-
lose enzymatic hydrolysis by cellulase allows reducing sugars to be fermented into
ethanol by the usage of yeasts or bacteria [57]. Endoglucanases, exoglucanases, and
-glucosidases are the three major groups of cellulases involved in the hydrolysis of
cellulose [9–12]. The two existing phases of enzymatic hydrolysis are primary and
secondary. During the first hydrolysis stage, liberation of oligosacharides into liquid

Fig. 9.9 Advantages of biological pretreatment for bioethanol production


9 Versatile Pretreatment Approaches to Improve the Bioethanol … 229

phase are the work of both endoglucanases and exoglucanases on the solid substrate
surface. During secondary hydrolysis step, Oligosacharides are further degraded to
cellobiose and glucose [44].

9.4 Production of Bioethanol

9.4.1 Using Sugar-Containing Raw Materials

The microbe Saccharomyces cerevisiae is commonly used for the production of


bioethanol from sugar-containing feedstocks because of its ability that break down
sucrose to hexoses (glucose and fructose) (see Fig. 9.10). Because S. cerevisiae cells
need oxygen of very small amounts for synthesizing fatty acid and sterol during
bioethanol generation, important bioprocess parameter includes aeration (3–7). S.
cerevisiae cannot bear increased concentrations of sugar and salts, as well as elevated
temperatures in the medium. The media containing cane molasses exhibits high
osmolarity, which limits ethanol formation, due to sugar and salt concentrations in
the medium. A large quantity of tests had been done to develop strains of S. cerevisiae
that are tolerant to high salt and temperature (7).
In the continuous bioprocesses for the production of bioethanol, concentration
of immobilized cells is relatively high, and bioprocesses can be easily controlled
at higher dilution rates, resulting in developed bioprocess productivities [41]. The
immobilization approaches are classified by the following members as shown in
Fig. 9.11.
Yeast cell trapping or mechanical separation is generally less effective when
compared to surface adsorption immobilization. Although some of the system is

Fig. 9.10 The bioethanol production process using sugar as a starting material
230 R. Rajesh Kannan et al.

Fig. 9.11 Immobilization procedures employed in the synthesis of bioethanol

washed out of some yeast cells, surface adsorption studies have shown unaffected
development in yeast cell [18]. Efficacy of self-flocculating yeast cells in producing
bioethanol was comparable to yeast cells immobilized on supporting materials. More-
over, bioprocess is simpler as well as more cost-effective than when the yeast cell is
immobilized on supporting materials due to no usage of supporting material.

9.4.2 Using Starch-Containing Raw Components

Yeasts (Saccharomyces cerevisiae, Saccharomyces pastorianus, Schizosaccha-


romyces pombe, and Kluyveromyces sp., etc.) that are competent in metabolizing
starch hydrolysates are used in the dry-grind (67%) and wet-mill (33%) processes for
bioethanol synthesis from maize starch, respectively [11]. Production of bioethanol
in the United States often uses dry milling due to its operating costs and lower capital
[11]. The whole corn is ground and formed into a mash by combining it with water.
Furthermore, in a jet cooker which is set to 80–90 °C, the mash is cooked for 15–
20 min. -amylase is used to enhance liquefaction during jet cooking (in very small
doses). In secondary liquefaction, more -amylase is added, which takes 90 min at
95°.
A few adjustments to the dry-grind process were made before fermentation, to
retrieve the maize germ and fibre [19]. A variety of value-added coproducts (such as
fibre, germ, starch, and gluten) are produced using wet milling before fermentation,
making the process energy-efficient and cost-effective [11]. The requirements of wet
milling procedures are clean, steeped, and degermed maize to efficiently procure the
extraction of oil from corn germ. The defibration of maize is done in order to extract
the fibres, gluten, and starch. The bioethanol manufacturing process includes the
same phases as the dry-grind method [12]. The yield of wet-mill is given as 0.2919,
while the dry-grind yield is 0.3235 per kg of corn [11]. For bioethanol production,
the technology which includes the growth of yeast (from active dry yeasts) in the
bioreactor during early saccharification is also used. Simultaneous saccharification,
yeast proliferation, and fermentation are some of the names used for this method
(SSYPF) [12].
9 Versatile Pretreatment Approaches to Improve the Bioethanol … 231

Fig. 9.12 The steps required in producing bioethanol from lignocellulosic raw materials

9.4.3 Using Lignocellulose-Containing Raw Materials

The conversion of lignocellulosic raw materials to fermentable sugars has been an


extensive research subject in the biorefining field. Despite the high level of interest
and advancement in lignocellulosic bioethanol research and development, several
hurdles remain [26, 56]. Figure 9.12 depicts the utilized processes in the manufacture
of bioethanol using raw materials which contains lignocellulose.
The lignocellulosic raw material pretreatment for the production of bioethanol
was covered in the preceding section. As a result, the fermentation of lignocel-
lulosic hydrolysates will be the emphasis of this section. During fermentation of
glucose to ethanol, a number of microorganisms are used, the most common of
which being the Saccharomyces cerevisiae. Unlike monosaccharides and disaccha-
rides, pentoses are incapable of being broken down. Furthermore, S. cerevisiae is
unable to digest cellulose and hemicellulose directly [23]. Recombinant DNA tech-
nology or pentose-fermenting bacteria (e.g. Pichia stipitis, Pachysolen tannophilus,
and Candida shehatae) is a certain procedure utilized to circumvent this constraint
[22]. These pentose-fermenting bacteria produce ethanol at a rate, at least five times
slower than S. cerevisiae from glucose. Furthermore, the tolerance is 2–4 times lower
for these microbes in the presence oxygen and ethanol [21–23].
Arabinose-metabolizing genes obtained from yeasts like Candida aurigiensis
have been inserted into S. cerevisiae and, and as a result, there have recently been a
lot of efforts to find an optimal microbe which produces ethanol directly from any
form of carbohydrate [11].
The efficiency of ethanol extraction using vegetable oils, such as coconut, olive,
safflower, and castor oil, and their derivatives, alcohols, and esters was also examined.
232 R. Rajesh Kannan et al.

These oils were compared with the following esters: methyl laurate, methyl oleate,
methyl linoleate, and methyl ricinoleate, and alcohols: lauryl (1-dodecanol), oleyl,
and ricinoleyl.

9.5 Future Direction and Perspective

Ethanol is a fundamental chemical product, an important solvent, an industrial


building block, and a promising renewable fuel. The chemical equilibrium conver-
sion of the hydration of ethylene decreases at high temperatures to increase the rate
of reaction. The excessive amount of energy to heat gases generates high pressure
and utilizes crude oil, which is a nonrenewable resource. Ethanol synthesis from
biomass residue produces a huge amount of CO2 with a high cost of operation
because of expensive machinery and fuel. The ethanol production from total sugars
in lignocellulosic materials is inhibited by the action of pentose sugars, which are not
fermentable by the brewer’s yeast, i.e. S. cerevisiae, during the hydrolysis of hemi-
cellulose. Although the fermentation route has been commercially realized, the cost
of operation for this process is expensive due to the energy-intensive distillation steps
at high dilution rates, which allow for high productivities, and the incomplete utiliza-
tion of substrate, resulting in low yields to meet the market demand. Creating highly
operative and selective nanocatalysts for the electrochemical CO2 reduction reaction
remains a major issue in the electrochemical CO2 reduction process. The reaction
mechanism is also closely associated with multiple catalytic active sites controlling
every elementary reaction over catalysts. Thus, catalysts for optimum active sites
should be designed and prepared to improve CO2 conversion and ethanol selectivity.
The formation rate of ethanol increases with increasing reaction temperature because
thermodynamic equilibrium is still not reached. However, many challenges, e.g. the
control of the ratio of multiple active sites, regulation of interface sites, and device of
catalyst structures need to be explored in future research to improve the conversion
and selectivity in ethanol production.

9.6 Conclusions

Bioethanol appears to be a viable solution to the current fuel crisis based on the
information presented. Bioethanol separation and purification, as well as renewable
biomass pretreatment, cellulase production, and sugar co-fermentation (pentose and
hexose), have all made significant progress in recent years. But bioethanol is still not
cost-competitive with the fossil fuels (the only exception being bioethanol produced
from sugar cane in Brazil). Furthermore, the difficult issue at hand is the lowering of
cost of bioethanol production. As a result, the concept of biorefinery is required to
better utilize renewable feedstocks and to produce additional value-added coproducts
(such as bio-based materials from lignin) that minimize the bioethanol production
9 Versatile Pretreatment Approaches to Improve the Bioethanol … 233

cost. As a result, the comparison of bioethanol to fossil fuels proves that bioethanol
will be a more cost-effective one. Based on the presented data, it is obvious that
bioethanol can be an alternative solution for the current fuel issue. The biggest
challenge remains how to reduce the production cost of bioethanol. Therefore, the
biorefinery concept is needed to utilize renewable feedstocks more comprehensively
and to manufacture more value-added coproducts (e.g. bio-based materials from the
lignin) that would reduce the cost of bioethanol production. This will make bioethanol
more economically competitive than fossil fuels.

References

1. Bhaskar, T., Bhavya, B., Singh, R., Naik, D. V., Kumar, A., & Goyal, H. B. (2011). Thermo-
chemical conversion of biomass to biofuels. In A. Pandey, C. Larroche, S. C. Ricke, C. G.
Dussap, & E. Gnansounou (Eds.), Biofuels—Alternative feedstocks and conversion processes
(pp. 51–77). Academic Press. https://doi.org/10.1016/B978-0-12-385099-7.00003-6
2. Forster, P., Ramaswamy, V., Artaxo, P., Berntsen, T., Betts, R., & Fahey, D. W. (2007). Changes
in atmospheric constituents and in radiative forcing. In S. Solomon, D. Qin, M. Manning, Z.
Chen, M. Marquis, & K. B. Averyt (Eds.), Climate change 2007: The physical science basis
(pp. 129–234). Cambridge University Press.
3. Cherubini, F., & Strømman, A. H. (2011). Principles of biorefining. In A. Pandey, C. Larroche,
S. C. Ricke, C. G. Dussap, & E. Gnansounou (Eds.), Biofuels—Alternative feedstocks and
conversion processes (pp. 3–24). Academic Press. https://doi.org/10.1016/B978-0-12-385099-
7.00001-2
4. Lange, J. P. (2007). Lignocellulose conversion: An introduction to chemistry, process and
economics. Biofuels, Bioproducts and Biorefining, 1(1), 39–48. https://doi.org/10.1002/bbb
5. Khanal, S. K. (2008). Anaerobic biotechnology for bioenergy production: Principles and
applications. Wiley-Blackwell. https://doi.org/10.1002/9780813804545
6. Yüksel, F., & Yüksel, B. (2004). The use of ethanol–gasoline blend as a fuel in an SI engine.
Renewable Energy, 29(7), 1181–1191. https://doi.org/10.1016/j.renene.2003.11.012
7. World Bioenergy Association. (2017). WBA Global Bioenergy Statistics. www.worldbioenergy.
org/global-bioenergy-statistics
8. Roadmap, B. (2012). German Federal Government action plans for the material and ener-
getic utilisation of renewable raw materials. https://www.bmbf.de/pub/Roadmap_Biorefiner
ies_eng.pdf
9. Kamm, B., Kamm, M., Gruber, P. R., & Kromus, S. (2008). Biorefinery—Industrial processes
and products: Status quo and future directions (pp. 1–40). Weinheim: Wiley-VCH Verlag
GmbH. https://doi.org/10.1002/9783527619849
10. Cherubini, F. (2010). The biorefinery concept: Using biomass instead of oil for producing
energy and chemicals. Energy Conversion and Management, 51(7), 1412–1421. https://doi.
org/10.1016/j.enconman.2010.01.015
11. Grilc, M., & Likozar, B. (2017). Levulinic acid hydrodeoxygenation, decarboxylation and
oligmerization over NiMo/Al2 O3 catalyst to bio-based value-added chemicals: Modelling of
mass transfer, thermodynamics and micro-kinetics. ChemEng J, 330, 383–397. https://doi.org/
10.1016/j.cej.2017.07.145
12. Huš, M., Bjelić, A., Grilc, M., & Likozar, B. (2018). First-principles mechanistic study of ring
hydrogenation and deoxygenation reactions of eugenol over Ru(0001) catalysts. Journal of
Catalysis, 358, 8–18. https://doi.org/10.1016/j.jcat.2017.11.020
13. Bjelić, A., Grilc, M., & Likozar, B. (2018). Catalytic hydrogenation and hydrodeoxygenation
of lignin-derived model compound eugenol over Ru/C: Intrinsic microkinetics and transport
234 R. Rajesh Kannan et al.

phenomena. Chemical Engineering Journal, 333, 240–259. https://doi.org/10.1016/j.cej.2017.


09.135
14. Kromus, S., Wachter, B., Koschuh, W., Mandl, M., Krotscheck, C., & Narodoslawsky, M.
(2004). The green biorefinery Austria—Development of an integrated system for green biomass
utilization. Chemical Biochemical Engineering Quartely, 18(1), 7–12.
15. Koller, M., Bona, R., Hermann, C., Horvat, P., Martinz, J., & Neto, J. (2005). Biotechnological
production of poly(3-hydroxybutyrate) with Wautersia eutropha by application of green grass
juice and silage juice as additional complex substrates. Biocatalysis Biotransformation, 23(5),
329–337. https://doi.org/10.1080/10242420500292252
16. Khuong, L. S., Masjuki, H. H., Zulkifli, N. W. M., NizaMohamad, E., Kalam, M. A., &
Alabdulkarem, A. (2017). Effect of gasoline-bioethanol blends on the properties and lubrication
characteristics of commercial engine oil. RSC Advances, 7(25), 15005–15019. https://doi.org/
10.1039/C7RA00357A
17. OECD/Food and Agriculture Organization of the United Nations. (2015). OECD-FAO
Agricultural Outlook 2015. OECD Publishing.
18. European Biofuels Technology Platform. (2010). Strategic research agenda 2010 update: Inno-
vation driving sustainable biofuels. http://www.biofuelstp.eu/srasdd/SRA_2010_update_web.
pdf
19. Van Thuijl, E., & Deurwaarder, E. P. (2006). European biofuels policies in retrospect. Petten,
the Netherlands: Energy Research Centre of the Netherlands (ECN). https://www.ecn.nl/pub
lications/PdfTetch.aspx?nr=ECN-C-06-016
20. Prieur-Vernat, A., & His, S. (2007). Biofuels worldwide. Panorama 2007. IFP-Innovation
Energy Environment. http://www.firp.ula.ve/archivos/material_web_4xx/07_IFP_Biofuels.
pdf
21. Li, K., Liu, S., & Liu, X. (2014). An overview of algae bioethanol production. International
Journal of Energy Research, 38(8), 965–977. https://doi.org/10.1002/er.3164
22. Balat, M., & Balat, H. (2009). Recent trends in global production and utilization of bio-ethanol
fuel. Applied Energy, 86(11), 2273–2282. https://doi.org/10.1016/j.apenergy.2009.03.015
23. Festel, G., Würmseher, M., Rammer, C., Boles, E., & Bellof, M. (2013) Modelling production
cost scenarios for biofuels and fossil fuels in Europe. Discussion paper No. 13–075. ZEW—
Centre for European Economic Research. http://ftp.zew.de/pub/zew-docs/dp/dp13075.pdf
24. Zhao, L., Zhang, X., Xu, J., Ou, X., Chang, S., & Wu, M. (2015). Techno-economic analysis
of bioethanol production from lignocellulosic biomass in China: Dilute-acid pretreatment and
enzymatic hydrolysis of corn stover. Energies, 8(5), 4096–4117. https://doi.org/10.3390/en8
054096
25. What determines retail prices for gasoline and diesel? Volta Oil (2018). https://www.voltaoil.
com/what-makes-up-retail-price-for-gasoline/
26. Mussatto, S. I., Dragone, G., Guimarães, P. M. R., Silva, J. P. A., Carneiro, L. M., & Roberto,
I. C. (2010). Technological trends, global market, and challenges of bio-ethanol production.
Biotechnology Advances, 28(6), 817–830. https://doi.org/10.1016/j.biotechadv.2010.07.001
27. Tomás-Pejó, E., Alvira, P., Ballesteros, M., & Negro, M. J. (2011). Pretreatment technologies
for lignocellulose-to-bioethanol conversion. In A. Pandey, C. Larroche, S. C. Ricke, C. G.
Dussap, & E. Gnansounou (Eds.), Biofuels—Alternative feedstocks and conversion processes
(pp. 149–176). Academic Press. https://doi.org/10.1016/B978-0-12-385099-7.00007-3
28. Linoj, K. N. V., Dhavala, P., Goswami, A., & Maithel, S. (2006). Liquid biofuels in South Asia:
Resources and technologies. Asian Biotechnology and Development Review, 8(2), 31–49.
29. İçöz, E., Tuğrul, K. M., Saral, A., & İçöz, E. (2008). Research on ethanol production and use
from sugar beet in Turkey. Biomass and Bioenergy, 33(1), 1–7. https://doi.org/10.1016/j.bio
mbioe.2008.05.005
30. Tabak, J. (2009). Biofuels. Infobase Publishing.
31. Pavlečić, M., Vrana, I., Vibovec, K., IvančićŠantek, M., Horvat, P., & Šantek, B. (2010).
Ethanol production from different intermediates of sugar beet processing. Food Technology
and Biotechnology, 48(3), 362–367.
9 Versatile Pretreatment Approaches to Improve the Bioethanol … 235

32. Pavlečić, M., Rezić, T., IvančićŠantek, M., Horvat, P., & Šantek, B. (2017). Bioethanol produc-
tion from raw sugar beet cossettes in horizontal rotating tubular bioreactor. Bioprocess and
Biosystems Engineering, 40(11), 1679–1688. https://doi.org/10.1007/s00449-017-1823-x
33. Senthilkumar, V., & Gunasekaran, P. (2009). Bioethanol from biomass production of ethanol
from molasses. In A. Pandey (Ed.), Handbook of plant-based biofuels (pp. 73–86). CRC Press.
34. Manikandan, K., Saravanan, V., & Viruthagiri, T. (2008). Kinetics studies on ethanol production
from banana peel waste using mutant strain of Saccharomyces cerevisiae. Indian Journal of
Biotechnology, 1(7), 83–88.
35. Jobling, S. (2004). Improving starch for food and industrial applications. Current Opinion in
Plant Biology, 7(2), 210–218. https://doi.org/10.1016/j.pbi.2003.12.001
36. Solomon, B. D., Barnes, J. R., & Halvorsen, K. E. (2007). Grain and cellulosic ethanol: History,
economics, and energy policy. Biomass and Bioenergy, 31(6), 416–425. https://doi.org/10.
1016/j.biombioe.2007.01.023
37. Mousdale, D. M. (2008). Biofuels. Biotechnology, chemistry and sustainable development.
CRC Press.
38. Soccol, C. R., Faraco, V., Karp, S., Vandenberghe, L. P. S., Thomaz-Soccol, V., &
Woiciechowski, A. (2011). Lignocellulosic bioethanol: Current status and future perspectives.
In A. Pandey, C. Larroche, S. C. Ricke, C. G. Dussap, & E. Gnansounou (Eds.), Biofuels—
Alternative feedstocks and conversion processes (pp. 101–122). Academic Press. https://doi.
org/10.1016/B978-0-12-385099-7.00005-X
39. Pandey, A., Nigam, P., Soccol, C. R., Soccol, V. T., Singh, D., & Mohan, R. (2000). Advances
in microbial amylases. Biotechnology and Applied Biochemistry, 31(Pt 2), 135–152. https://
doi.org/10.1042/BA19990073
40. Shigechi, H., Koh, J., Fujita, Y., Matsumoto, T., Bito, Y., & Ueda, M. (2004). Direct produc-
tion of ethanol from raw corn starch via fermentation by use of a novel surface-engineered
yeast strain codisplaying glucoamylase and α-amylase. Applied and Environment Microbiology,
70(8), 5037–5040. https://doi.org/10.1128/AEM.70.8.5037-5040.2004
41. Lee, S., Speight, J. G., & Loyalka, S. K. (Eds.) (2007). Handbook of alternative fuel
technologies. CRC Press. https://doi.org/10.1201/9781420014518
42. Schubert, C. (2006). Can biofuels finally take center stage? Nature Biotechnology, 24, 777–784.
https://doi.org/10.1038/nbt0706-777
43. Tanadul, O. U., VanderGheynst, J. S., Beckles, D. M., Powell, A. L. T., & Labavitch, J. M.
(2014). The impact of elevated CO2 concentration on the quality of algal starch as a potential
biofuel feedstock. Biotechnology and Bioengineering, 111(7), 1323–1331. https://doi.org/10.
1002/bit.25203
44. Binod, P., Sindhu, R., Singhania, R. R., Vikram, S., Devi, L., & Nagalakshmi, S. (2010).
Bioethanol production from rice straw: An overview. Bioresource Technology, 101(13), 4767–
4774. https://doi.org/10.1016/j.biortech.2009.10.079
45. Quintero, J. A., Rincón, L. E., & Cardona, C. A. (2011). Production of bioethanol from agroin-
dustrial residues as feedstocks. In A. Pandey, C. Larroche, S. C. Ricke, C. G. Dussap, & E.
Gnansounou (Eds.), Biofuels—Alternative feedstocks and conversion processes (pp. 251–285).
Academic Press. https://doi.org/10.1016/B978-0-12-385099-7.000011-5
46. Dale, B. E., & Kim, S. (2010). Biomass refining global impact—The biobased economy of
the 21st century. In B. Kamm, P. R. Gruber, & M. Kamm (Eds.), Biorefineries—Industrial
processes and products (Status quo and future directions) (pp. 41–66). Wiley-VCH.
47. Hatti-Kaul, R. (2010). Biorefineries—A path to sustainability? Crop Science 50(Suppl_
1):S152–S156. https://doi.org/10.2135/cropsci2009.10.0563
48. Mansfield, S. D., Mooney, C., & Saddler, J. N. (1999). Substrate and enzyme characteristics
that limit cellulose hydrolysis. Biotechnology Progress, 15(5), 804–816. https://doi.org/10.
1021/bp9900864
49. Kumar, R., & Wyman, C. E. (2010). Key features of pretreated lignocellulosic biomass solids
and their impact on hydrolysis. In K. Waldron (Ed.), Bioalcohol production—Biochemical
conversion of lignocellulosic biomass. Series in Energy: Number 3 (pp. 73–121). Woodhead
Publishing.
236 R. Rajesh Kannan et al.

50. Laureano-Perez, L., Teymouri, F., Alizadeh, H., & Dale, B. E. (2005). Understanding factors
that limit enzymatic hydrolysis of biomass. Applied Biochemistry Biotechnology, 121–124(1–
3), 1081–1099. https://doi.org/10.1385/ABAB:124:1-3:1081
51. Mosier, N., Wyman, C., Dale, B. D., Elander, R., Lee, Y. Y., & Holtzapple, M. (2005). Features
of promising technologies for pretreatment of lignocellulosic biomass. Bioresource Technology,
96(6), 673–686. https://doi.org/10.1016/j.biortech.2004.06.025
52. Pan, X., Xie, D., Gilkes, N., Gregg, D. J., & Saddler, J. N. (2005). Strategies to enhance the enzy-
matic hydrolysis of pretreated softwood with high residual lignin content. Applied Biochemistry
Biotechnology, 121–124, 1069–1079. https://doi.org/10.1385/ABAB:124:1-3:1069
53. Chandra, R. P., Bura, R., Mabee, W. E., Berlin, A., Pan, X., & Saddler, J. N. (2007). Substrate
pretreatment: The key to effective enzymatic hydrolysis of lignocellulosics? In: L. Olsson (Ed.)
Biofuels: Advances in biochemical engineering/biotechnology (vol. 108, pp. 67–93). Springer.
https://doi.org/10.1007/10_2007_064
54. Huang, H. J., Ramaswamy, S., Tschirner, U. W., & Ramarao, B. V. (2008). A review of separa-
tion technologies in current and future biorefineries. Separation and Purification Technology,
62(1), 1–21. https://doi.org/10.1016/j.seppur.2007.12.011
55. Kucera, D., Benesova, P., Ladicky, P., Pekar, M., Sedlacek, P., & Obruca, S. (2017). Production
of polyhydroxyalkanoates using hydrolyzates of spruce sawdust: Comparison of hydrolyzates
detoxification by application of overliming, active carbon, and lignite. Bioengineering (Basel),
4(2), 53. https://doi.org/10.3390/bioengineering4020053
56. Taherzadeh, M. J., & Karimi, K. (2008). Pretreatment of lignocellulosic wastes to improve
ethanol and biogas production: A review. Int J MolSci, 9(9), 1621–1651. https://doi.org/10.
3390/ijms9091621
57. Sun, Y., & Cheng, J. (2002). Hydrolysis of lignocellulosic materials for ethanol production: A
review. Bioresource Technology, 83(1), 1–11. https://doi.org/10.1016/S0960-8524(01)00212-7
58. Talebnia, F., Karakashev, D., & Angelidaki, I. (2010). Production of bioethanol from wheat
straw: An overview on pretreatment, hydrolysis and fermentation. Bioresource Technology,
101(13), 4744–4753. https://doi.org/10.1016/j.biortech.2009.11.080
59. Karunanithy, C. (2010). Effect of extruder parameters and moisture content of switch-
grass, prairie cord grass on sugar recovery from enzymatic hydrolysis. Applied Biochemistry
Biotechnology, 162(6), 1785–1803. https://doi.org/10.1007/s12010-010-8959-3
60. Yoo, J., Alavi, S., Vadlani, P., & Amanor-Boadu, V. (2011). Thermo-mechanical extrusion
pretreatment for conversion of soybean hulls to fermentable sugars. Bioresource Technology,
102(16), 7583–7590. https://doi.org/10.1016/j.biortech.2011.04.092
61. Kootstra, A. M. J., Beeftink, H. H., Scott, E. L., & Sanders, J. P. M. (2010). Comparison of dilute
mineral and organic acid pretreatment for enzymatic hydrolysis of wheat straw. Biochemical
Engineering Journal, 46(2), 126–131. https://doi.org/10.1016/j.bej.2009.04.020
62. Carvalheiro, F., Duarte, L. C., & Gírio, F. M. (2008). Hemicellulose biorefineries: A review on
biomass pretreatments. Journal of Scientific and Industrial Research (India), 67, 849–864.
63. Cheng, Y. S., Zheng, Y., Yu, C. W., Dooley, T. M., Jenkins, B. M., & VanderGheynst, J. S.
(2010). Evaluation of high solids alkaline pretreatment of rice straw. Applied Biochemistry and
Biotechnology, 162(6), 1768–1784. https://doi.org/10.1007/s12010-010-8958-4
64. Kumar, P., Barrett, D. M., Delwiche, M. J., & Stroeve, P. (2009). Methods for pretreat-
ment of lignocellulosic biomass for efficient hydrolysis and biofuel production. Industrial
and Engineering Chemistry Research, 48(8), 3713–3729. https://doi.org/10.1021/ie801542g
65. García-Cubero, M. A., González-Benito, G., Indacoechea, I., Coca, M., & Bolado, S.
(2009). Effect of ozonolysis pretreatment on enzymatic digestibility of wheat and rye straw.
BioresourceTechnology, 100(4), 1608–1613. https://doi.org/10.1016/j.biortech.2008.09.012
66. Silverstein, R. A., Chen, Y., Sharma-Shivappa, R. R., Boyette, M. D., & Osborne, J. (2007). A
comparison of chemical pretreatment methods for improving saccharification of cotton stalks.
Bioresource Technology, 98(16), 3000–3011. https://doi.org/10.1016/j.biortech.2006.10.022
67. Barros, R. R., Paredes, R. D. S., Endo, T., da Silva Bon, E. P., & Lee, S. H. (2013). Association
of wet disk milling and ozonolysis as pretreatment for enzymatic saccharification of sugarcane
bagasse and straw. Bioresource Technology, 136, 288–294. https://doi.org/10.1016/j.biortech.
2013.03.009
9 Versatile Pretreatment Approaches to Improve the Bioethanol … 237

68. Olivier-Bourbigou, H., Magna, L., & Morvan, D. (2010). Ionic liquids and catalysis: Recent
progress from knowledge to applications. Applied Catalysis A, 373(1–2), 1–56. https://doi.org/
10.1016/j.apcata.2009.10.008
69. Hayes, D. J. (2009). An examination of biorefining processes, catalysts and challenges.
Catalysis Today, 145(1–2), 138–151. https://doi.org/10.1016/j.cattod.2008.04.017
70. Banerjee, C., Mandal, S., Ghosh, S., Kuchlyan, J., Kundu, N., & Sarkar, N. (2013). Unique char-
acteristics of ionic liquids comprised of long-chain cations and anions: A new physical insight.
Journal of Physics Chemistry B, 117(14), 3927–3934. https://doi.org/10.1021/jp4015405
71. Latif, M. N., Wan Isahak, W. N. R., Samsuri, A., Hasan, S. Z., Manan, W. N., & Yaakob, Z.
(2023). Recent advances in the technologies and catalytic processes of ethanol production.
Catalysts, 13(7), 1093. https://doi.org/10.3390/catal13071093
72. Shukla, A., Kumar, D., Girdhar, M., Kumar, A., Goyal, A., Malik, T., & Mohan, A. (2023).
Strategies of pretreatment of feedstocks for optimized bioethanol production: distinct and
integrated approaches. Biotechnology for Biofuels and Bioproducts, 16, 44.
73. Klongklaew, A., Unban, K., Kalaimurugan, D., Kanpiengjai, A., Azaizeh, H., Schroedter, L.,
Venus, J., & Khanongnuch, C. (2023). Bioconversion of dilute acid pretreated corn stover to
L-lactic acid using co-culture of furfural tolerant Enterococcus mundtii WX1 and Lactobacillus
rhamnosus SCJ9. Fermentation, 9, 112.
74. Gong, J.-S.-Q., Su, J.-E., Cai, J.-Y., Zou, L., Chen, Y., Jiang, Y.-L., & Hu, B.-B. (2023).
Enhanced enzymolysis and bioethanol yield from tobacco stem waste based on mild synergistic
pretreatment. Frontiers in Energy Research, 10, 989393.
75. Usino, D. O., Sar, T., Ylitervo, P., & Richards, T. (2023). Effect of acid pretreatment on the
primary products of biomass fast pyrolysis. Energies, 16, 2377.
76. Abdurrahman, A., Richard, P. I., Galadima, A. U., Muhammad, A., & Adamu, M. (2022). Dilute
sulphuric acid pre-treatment for efficient production of bioethanol from sugarcane bagasse
using Saccharomyces cerevisiae. Journal of Biotechnology, 1, 56–65.
77. Zhao, H., Baker, G. A., & Cowins, J. V. (2010). Fast enzymatic saccharification of switchgrass
after pretreatment with ionic liquids. Biotechnology Progress, 26(1), 127–133. https://doi.org/
10.1002/btpr.331
78. Nguyen, T. A. D., Kim, K. R., Han, S. J., Cho, H. Y., Kim, J. W., & Park, S. M. (2010).
Pretreatment of rice straw with ammonia and ionic liquid for lignocellulose conversion to
fermentable sugars. Bioresource Technology, 101(19), 7432–7438. https://doi.org/10.1016/j.
biortech.2010.04.053
79. da Costa Lopes, A. M., João, K. G., Rubik, D. F., Bogel-Łukasik, E., Duarte, L. C., &
Andreaus, J. (2013). Pre-treatment of lignocellulosic biomass using ionic liquids: Wheat straw
fractionation. Bioresource Technology, 142, 198–208. https://doi.org/10.1016/j.biortech.2013.
05.032
80. Park, N., Kim, H. Y., Koo, B. W., Yeo, H., & Choi, I. G. (2010). Organosolv pretreatment with
various catalysts for enhancing enzymatic hydrolysis of pitch pine (Pinus rigida). Bioresource
Technology, 101(18), 7057–7064. https://doi.org/10.1016/j.biortech.2010.04.020
81. Mesa, L., González, E., Cara, C., González, M., Castro, E., & Mussatto, S. I. (2011). The effect
of organosolv pretreatment variables on enzymatic hydrolysis of sugarcane bagasse. Chemical
Engineering Journal, 168(3), 1157–1162. https://doi.org/10.1016/j.cej.2011.02.003
Chapter 10
Biorefinery Avenues for Processing
Urban Solid Waste: Potential
for Value-Added Chemicals and Energy

Swapna Gade, Yuvraj Patil, and Bhalchandra Bhanage

Abstract Global urbanization against rural settlements has been on a steady rise
over the past 6 decades. Nearly one-third of the entire population of India residing in
urban locations (roughly 5.5% of available residential land) has placed an unprece-
dented burden on the municipal economy, energy utilization, and waste manage-
ment systems. Sustainability as regards resources and municipal solid waste (MSW)
management remains a challenge. The focus of this chapter concerns these chal-
lenges, the various avenues for processing MSW, biorefinery approaches for de-novo
generation of commercial products, energy, etc. Additionally, the sustainability of
these approaches, cost to the taxpayer, and the conversion of MSW to value-added
resources or energy are also discussed. Harnessing bioenergy from MSW biomass
may be an attractive route for addressing the ongoing global energy crisis. Carbon
dioxide can be trapped via reactive amines to reduce emissions and generate value-
added compounds such as urea. Certain wastes within MSW can be processed to
obtain glycerol-based compounds which have wide applications in industry. Segre-
gated waste management strategies will help in selectively processing organic wastes
using the biorefinery route, which entails broad-spectrum cellular/enzymatic tools to
break down the organic material. Systematic management of MSW can be expected
to be a major resource salvage route. Harvesting MSW is thus proposed here to be
a vital mechanism to generate a circular bioeconomy aimed at sustainability and
economic viability.

S. Gade
Chemical Engineering and Process Development Division, National Chemical Laboratory, Homi
Bhaba Road, Pashan, Pune 411008, India
Y. Patil
School of Health Sciences and Technology, Dr. Vishwanath Karad MIT World Peace University,
Pune 411038, India
B. Bhanage (B)
Department of Chemistry, Institute of Chemical Technology, Nathalal Parekh Marg, Matunga,
Mumbai 400019, India
e-mail: bm.bhanage@gmail.com

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2024 239
G. Baskar et al. (eds.), Circular Bioeconomy Perspectives in Sustainable Bioenergy
Production, Energy, Environment, and Sustainability,
https://doi.org/10.1007/978-981-97-2523-6_10
240 S. Gade et al.

Keywords Value-added chemicals · Biorefinery-enriched Economy · Municipal


solid waste processing

10.1 Introduction

Global fuel crisis, resource restrictions, growing human population, and the burden
put on urban economies and administration are unprecedented in human history.
Over the last 3 decades, attempts have been made to look at alternatives for resources,
especially those that are finite, such as fossil fuel. Resources for powering society
and commerce in general may stem from potentially renewable origins, provided
these are handled appropriately. An example of such a resource is plant-based fiber
or oil which has several applications in fabrics or biodiesel domain, for instance.
Population burden-wise, the United Nations World Urbanization Prospects report
indicates that nearly one-third of the population of India has come to reside in urban
locations which constitute about 5.5% of available residential (non-agricultural land)
land. The phenomenon has placed a burden on the municipal economy and waste
management systems leading to several challenges for municipal solid waste (MSW)
solutions.

10.1.1 Challenges of a Post-covid World: Energy Crisis,


Lower CO2 Emission Milestones, and Poor Resource
Availability

The past decade has seen multiple conflicts in the Eurasian region particularly mapped
as a major crude oil and natural gas resource. As a direct consequence, Europe and
other regions have seen fuel restrictions and inflation in fuel prices [1, 2]. As the global
commerce network continues to be intimately linked, such regional disturbances
translate to global upsets that sustain long-drawn economic and fuel crises. With the
ever-looming global warming and climate change scenario, developed nations are
also pushing the envelope on curbing carbon footprints across the globe [3, 4]. The
Paris agreement outlines a 43% decrease in greenhouse gas (GHG) emissions by
2030 if we are to prevent global warming of the planet exceeding 1.5 ºC, to avoid far-
reaching climatic disasters. For instance, with a 2 ºC increase in global temperature,
coral reefs are in imminent threat of annihilation [4]. Thus, with two major crises and
an active international policy for reduced use of carbon fuel, significant work and
discussion have been carried out that covers alternative and sustainable approaches to
energy, and resource development as well as the framework that these approaches are
based upon. A practical approach to a more sustainable future lies in managing the
spent resources or waste, or in other words, to implement a circular economy [5]. The
current chapter discusses avenues of managing a crucial aspect of waste generated and
10 Biorefinery Avenues for Processing Urban Solid Waste: Potential … 241

its potential processing in biorefineries to produce value-added products. This circular


approach or, a circular economy, aims at reutilizing the waste to resupply value-added
products back into the economy and manage resource restriction as well as salvage
compounds that would normally be lost [5]. In a review by Ferreira et al., a circular
bioeconomy idea is proposed to rely on the utilization of waste(s) as a feedstock for
different biotechnological processes to efficiently obtain a broad-spectrum yield of
biochemical components through a biorefinery approach [6].

10.1.2 Urban Ecosystems, Municipal Waste,


and Sustainability

MSW, in the context of this chapter, is fundamentally different from other anthropo-
logical wastes. It does not bear the same constituents as agricultural or forest waste.
The diverse nature of waste presents multiple issues that require attention in the
following areas: Waste collection/segregation, Waste Processing or Recycling, and
Material viability for Biorefinery [5, 6]. The composition of MSW is dynamic in
nature and varies by region and culture. The changing nature also extends to the
temporal dimension and therefore requires periodic review for optimal processing
of MSW. The typical composition of MSW has six recurring waste matter groups;
food, paper, wood, metals, rubber/plastics/fabric, glass, and a miscellaneous umbrella
group [7, 8]. Figure 10.1 depicts the mean proportions of the waste categories across
different global locations. As a general rule, the amount of waste generated increases
in proportion to the national GDP, but local factors may drive the proportion of the
waste. As observed, a few peculiarities emerge from the composition study, namely,
(a) irregular wastage reported for wood and paper, (b) reported data for food/kitchen
waste, and (c) the broad miscellaneous category which may contain key elements
that need attention.
China and by extension, Asian nations such as India appear to have a large propor-
tion of food wastage which could be considered as biomass. While paper waste
appears in all regions, wood waste is limited or virtually absent in European/Asian
data. Observations such as these impact the broader consideration of biorefinery
approaches because of the implied availability or limitation of the feedstock.
Interestingly, according to a report by Ministry of Housing and Urban Affairs
(2021), India demonstrates a food waste of about 45%, like the EU model. A similar
profile for glass, metals, wood (construction), etc., is also noted. The rubber/plastic/
fabric waste is however the highest and on par or greater than that demonstrated
by China. It may be speculated that being Asian super-centers of consumer goods
production, India and China may show similar tendencies of resource wasting when
it comes to plastic, rubber, or fabrics.
242 S. Gade et al.

Fig. 10.1 Typical composition of MSW and regional variation. Data Adapted from [7, 8] and
MoHUA Report on Circular Economy and Waste Management (2021). a Typical MSW composition
as observed across developed nations and in China, b Denotes break-away Pie chart data depicting
India’s average MSW composition

10.1.3 Municipal Waste Management and Challenges


with Solid Wastes

Management of MSW is a challenging task, especially when it is associated with


different waste categories. The task is even more tedious in the developing countries
on account of randomization of waste collection and non-adherence to segregation
rules [7, 8]. Problems that are frequently linked with poor management of MSW are
given below:
• Poor solid waste management and treatment,
• Unregulated waste dumping,
• Insufficient operator skills,
• Lack of responsiveness of health hazards,
• Unmanaged waste assortment system,
• Lack of awareness about recycling.
These mentioned difficulties are typically found in middle/poor-income countries.
Urbanization in such countries is frequently without appropriate planning. Rapid
proliferation of industries without proper planning is also one of the reasons leading
to poor waste management [3]. Occurrences such as these are critical in importance in
low-income to mid-income nations, where unregulated migration from rural zones to
urban locales is normal, urban expansion sans adequate city-planning, and hastened
industrialization with lacking infrastructure and supporting networks. Uncontrolled
10 Biorefinery Avenues for Processing Urban Solid Waste: Potential … 243

population upticks, as mentioned earlier and also noted by the United Nations World
Urbanization Prospects report (2018), are also the cause of some of the ill effects.

10.2 Conventional Approaches to MSW Management

MSW is an urban planning problem of great proportions and one which grows expo-
nentially with a greater standard of living and growing economy. While the nature
of MSW may not be altered significantly across economic statures, the volume of
MSW generated can and will certainly increase [5, 7]. Typical approaches toward
managing MSW or garbage in general include the following.
A. Landfill
The earliest findings of landfills from the Roman era highlight the traditional
municipal approach to disposing of garbage by accumulating wastes and burying
or covering it [9]. As noted in Fig. 10.2, many countries prefer to use landfill as
a viable option for MSW disposal. With relatively available non-agricultural land
space landfill is opted over any of the other MSW management alternatives. The
preference for landfill also loosely correlates with the economic status and political
bent of the nation. It is interesting to note that economic analysis of landfill utility
also includes cost payable as landfill gate fees, the latter being of an incremental
nature. This implies that not only is landfill space limited, but that it comes at a
premium, a cost which is undoubtedly passed on the taxpayer [5].
Nations such as Japan have virtually no space for landfill, leading the charge
for alternative MSW processing. One such avenue is MSW incineration, which is
considered as energy retrieval [7, 10, 11].
2 Incineration
As noted in Fig. 10.2, Japan as a first world nation with a large economy shows
a surprising dependence on incineration of MSW. While incineration undoubtedly
dramatically reduces the volume of MSW, it is associated with thermal inputs for
the municipal system. MSW incineration has been attributed to significant electricity
generation and supply for district heating [10]. Incineration is an advanced thermal

Fig. 10.2 Fate of municipal solid waste across different nations


244 S. Gade et al.

Fig. 10.3 Schematic map of MSW incineration

process carried out at 850 °C in presence of oxygen. Further waste is then transformed
to flue gas i.e., carbon dioxide (CO2 ), water, and non-combustible materials named
bottom ash. Flue gas is useful in generating electricity and heat via boiler and after
dusting the flue gas, which is emitted to the atmosphere. Figure 10.3 shows the
schematic map of MSW Incineration. As an island nation, Japan can optimize MSW
disposal while addressing its energy demands. Incineration has potent drawbacks,
such as release of toxic metal and similar material particles which may lead to cancer
[11].
An incineration process producing heat which has a thermal efficacy of around
80–90% can be used for electricity generation at around 17–30% crude effective-
ness. The competence of an incinerator for energy corporations is inferior to gas-fired
power station or a large coal. Normally, the efficiency of coal fire power station is in
between 33 and 38% while gas turbine power station electric power efficiency could
be ~ 50%. However, technical issues may arise in the case of larger-scale coal and
gas-fired power plants. Whereas, the energy generated from an incinerator can be
used to produce steam for heating, and the productivity is like a boiler fired by natural
oil/gas. It is possible that, in the future, the efficiency of electricity generation using
incineration will increase, given the trend in other solid fuel applications involving
more severe operating conditions at higher temperature [7, 10]. It is conceivable, in
our opinion that, the future may hold greater upward mobility of the electricity gener-
ating potential, based on the rising trend of solid fuel applications which frequently
involve more severe operating conditions at higher temperatures [7].
10 Biorefinery Avenues for Processing Urban Solid Waste: Potential … 245

3 Composting
Composting is biological decomposition of the organic fraction which can be
done under precise conditions. Typically, end-to-end composting is executed as—
compost preparation—decomposition-post processing and marketing [12]. It is the
most common waste treatment process for organic waste segment or for garden waste,
and as such is suitable for organic and biodegradable waste. Routinely, large swaths
of land area are needed for composting and are subsequently locked up in the activity.
Furthermore, it suffers a critical drawback that, with this process, we cannot control
emission of gases. The implication here is that green routes are not followed making
this a less satisfactory and a potentially unsustainable method for waste disposal.
MSW composting is a relatively new practice for conversion of MSW into a value-
added soil additive. As would be expected from compost, MSW composting has been
shown to enhance growth characteristics of plants such as wheat, Swiss chard and
basil, among others [12, 13]. Interestingly, it is not noted to affect plant growth
negatively by potential addition of heavy metals. Furthermore, heavy metals may
possibly remain bound to the organic fraction of MSWC. A study conducted in Turkey
has also shown MSW composting to have remarkable properties of sequestration of
metals like nickel and cadmium in contaminated soil samples [14]. Unfortunately,
despite definitive advantages to composting of MSW, relatively few nations have
leaned toward adoption of the approach (see Fig. 10.2).
4 Recycling
While recycling has been promoted for a long time, prior to the MSW management
crisis, recent government initiatives and incentives are a strong driver for MSW recy-
cling of plastics, etc. Despite initiatives, though, without public involvement, recy-
cling may not broadly benefit MSW management [15, 16]. As discussed earlier, devel-
oped nations such as the United States demonstrate a greater paper waste compared to
other nations. Consequently, recycling has indeed emerged as a mitigation remedy
for MSW management [7]. On the other hand, countries such as Bulgaria with a
missing axis of recycling, composting, or incineration (energy recovery) are poor
candidates for implementing circular economy (see Fig. 10.2).
It has been observed that the lower-income families produce comparatively more
organic food waste. Large amounts of glass, plastics, wood, and metals-based waste
are frequently noted as the waste output of higher-income families. The ability to
reprocess waste depends on the MSW composition. For example, in low-income
nations, organic waste can be utilized for making cheap fertilizer by composting
it [17]. Effective collection, separation of waste, and further processing are directly
proportional to the wealth of the community. Countries that are aware about the effec-
tive circular economy are therefore paying a major premium on waste management,
waste collection, and transportation. A few examples of waste collection percentage
in different countries as portrayed by the World Bank report on Wastes (2012), high-
light the dependency of economy and MSW management: Philippines and Sri Lanka
(40%), Paraguay and Vietnam (50%), India (70%), Malaysia, Switzerland, Germany,
Japan, and United States (100%).
246 S. Gade et al.

As Indian commerce catches up to the Chinese phenomenon of cheap, mass


production of plastic (and other materials) commodities, there is an inclement rise
in plastic wastes to the tune of nearly 15 K tons per day. Plastics add up to nearly
8% of all solid wastes, especially in cities such as Ahmedabad, Kolkata, and Delhi.
Plastics are considered to be process-competent for recycling and to some extent for
production of alternative fuel via pyrolysis.
Like Rishikesh, other Indian municipalities have invested in and explored the
potential of MSW management with the view of creating circular economy. Indore,
as reported by TERI (2018), which has recently gained the unique standing of being
the cleanest city in India, also manages its MSW by segregation of waste at the
source, followed by composting and fermentation to make manure and biofuel in the
form of CNG.
Rawat and Daverey have reported on Classification of household solid waste and
existing position of MSW management in Rishikesh and Uttarakhand, in India [18].
Based on broader review of literature concerning MSW, Table 10.1 displays various
components that are grouped under the varied waste categories.
The presence of paper-based items in the grey zone is a result of product contam-
ination, either with chemicals, potentially toxic dyes, inks, etc. They have associated
the kitchen and other waste generated from houses, and the rates of production of
waste are diverse worldwide. From Table 10.2, it can be seen that the per head left-
over production rate in Rishikesh is almost identical to other metro cities of evolving
countries such as Suzhou, Beijing. The waste production is very large in Mexicali
city (0.981 kg/c/d), as compared to metropolitan city [18].
Ding et al. have published a review on management and technologies used for the
treatment and utilization of resources of MSW. They have also made a comparative

Table 10.1 Classification of MSW


Types MSW composition Collection
color code
Kitchen Vegetable stems, animal skeleton viscera, leftovers, plant waste, Green
waste kitchen oil, fruit waste, etc.
Hazardous Waste tube bulb, scrap printer cartridges, batteries, waste electronics, Red
waste hair dye, overdue drugs and cosmetics, waste paint can, pesticide
products container, etc.
Recyclable Metal—pop-tin cans waste, toothpaste cover, etc. Blue
Paper—newspapers, cardboard waste, book pages, cartons etc.
Glass—various colored/non-colored glass bottles, mirror glass, etc.
Textile—school bags, waste clothes, etc.
Plastics—disposable plastic tableware, plastic bottles, plastic cup,
plastic foam, plastic packaging, etc.
Other Contaminated plastic bags, cigarette, dirty papers, shell covers, dust, Gray/
waste non-recyclable glass, waste/broken ceramics, contaminated diapers, Black
etc.
Bold text: Indicate material categories which form the basis of composition of MSW. The categorical
proportions vary from region to region
10 Biorefinery Avenues for Processing Urban Solid Waste: Potential … 247

Table 10.2 Household waste production rate and influence of food waste (%) from diverse cities
of world
City (country) Household waste generation (kg/c/d) Food waste (%) References
Wales (England) – 26.79 [19]
Dublin (Ireland) – 40.48 [20]
Takoradi (Ghana) 0.7 61 [21]
Mexicali (Mexico) 0.98 – [22]
Bolgatanga (Ghana) 0.21 61 [21]
Suzhou (China) 0.28 65.7 [23]
Beijing (China) 0.23 69.3 [24]
Rishikesh (India) 0.26 57.3 [18]

study of international regions waste management [8]. The authors have presented
different aspects of eight different cities in China, mainly situated in coastal regions.
The study is based on gross domestic production, MSW generation, MSW segre-
gation (collection, transport, separation, composition), and acts and regulations
related to the same. The study also includes advancement in technologies in MSW
source consumption for waste-to-material and waste-to-energy. MSW is mainly sepa-
rated into household food waste, recyclable waste, harmful waste, perishable waste-
land, and other waste. The review includes different types of MSW separation trends
that have been followed in China, mainly they are of four types which are collected
in different containers (Table 10.2) [8].
In other examples of countries/cities with MSW management systems, we have
Indonesia, where 60–70% of the produced waste is moved to landfills, whereas the
remaining 30–40% of waste generated is independently managed by the community,
burned it, or dropped in rivers [25]. The municipal waste in Berlin is mostly separated
into five types: trade waste (10.06%), domestic waste (73.80%), bulky waste (6.95%),
commercial waste (2.30%), and road sweepings (3.78%) [8].
Tokyo has an extremely organized waste collection strategy which is the most
important part of waste collection that has been done using the slogan “separation
is resource and mixing is waste” [26]. Waste management in Tokyo has four large
categories: flammable waste (3.3 million tons), non-flammable left-over (0.1 million
tons), recyclable matter (0.5 million tons), and bulky waste (0.08 million tons) [8, 26].
In the case of Singapore, MSW is separated in two main systems, general waste
collection system and public waste collection scheme. General waste collection
system attends industrial and commercial locations. These locations solid waste
includes organic waste, inorganic waste, sludge, and also grease. All these wastes
are collected by the licensed, general waste collection system is maintained by
private waste collectors. The public waste collection scheme mainly targets the MSW
produced by general household waste, garden waste, furnishings and domestic infras-
tructure waste, and remnants of animals. The accumulated waste is moved to different
processing sites for reprocessing and handling. Organic/inorganic wastes can be
248 S. Gade et al.

sorted and further transferred to incinerators or transfer stations for handling post-
contracting. Advanced technologies are also included in the waste collection systems
such as pneumatic waste collection instrumentation, applied in novel buildings and
private housing lands in Singapore. To transport the waste, underground pipes are
used with vacuum suction to reach the assembly point of waste. This method is more
hygienic and productive as compared to the above two methods [27, 28].
Biorefineries are increasingly being viewed as a viable commercial portal for mass
valorization of waste products [29]. The chief reason for this change of perspective is
not just resource limitation, but environmental concerns as well as central policies on
sustainability and economy. The latter frequently dictates scientific approaches and
broad adoption of the same, for instance, Fig. 10.1 depicts various approaches toward
urban waste management in different countries. It is interesting to note a geograph-
ical disparity in policy approaches as noted in the depiction [29]. Eastern European
nations rely heavily on the basic waste management solution of landfill while western
nations may employ more sustainable approaches of composting and recycling waste.
Although some of these western nations may also resort to waste negation by incin-
eration, this approach may be driven by lack of space for either landfill or recycling/
composting activities. Greece and Portugal display differences in MSW manage-
ment from other western European nations, possibly attributed to national economic
priorities and GDP. The United States and Japan as non-continental comparators are
interesting in that they show opposing preferences for land-use while at the same
time favor sustainable MSW management. Japan’s surprising reliance on incinera-
tion may be directly related to restriction of space availability in the island nation
[29].
Ghumra and colleagues discuss the ramifications of MSW management, in partic-
ular the valorization of MSW attributed to organic origin. The organic fraction or
food-sourced waste is a large potential source of ethanol, butanol, organic acids,
and gases. The biorefinery here is facilitated by pre-treatment of waste followed by
thermochemical or biochemical mass conversion. Subsequently, fermentation of the
retrieved sugars, amino acids, and other compounds results in value-added products.
Organic components of MSW are typically a complex mixture of cellulose, carbohy-
drates, proteins, and lipids. With regard to thermochemical mass conversion, Plasma
gasification is one of the advanced technologies being tested for altering MSW into
circular economy. This advanced technology needs much attention, training, and
need to spread awareness among research and Municipal bodies who are working
on waste-to-circular economy program. Need to do much attention and research for
successful commercialization [28].
Plasma gasification can be an effective technique for the disposal of MSW and
could be much more beneficial for the waste-to-value circular economy. Successful
industrial application of this process is needed. Since last decade, few research groups
have been studying on application of plasma gasification, the problem associated with
this process, etc. [28, 30–33].
10 Biorefinery Avenues for Processing Urban Solid Waste: Potential … 249

10.3 Biorefinery to Bioeconomy

To improve circular economy, we need to reconsider organic/bio-waste and its poten-


tial role in circular bioeconomy. We need to improve eco-friendly utilization of
resources. It may well influence the reduction in municipal waste, carbon demand,
and greenhouse gases.
Bioeconomy structures surpass circular economic points that contain product or
facility zones. It is mainly determined by vast number of biological processes, animal
waste, fisheries, forestry, food waste, and animal feedstuff. These biomass wastes
are used in various applications. In chemical processes and aquaculture (aquarium),
etc. biomass waste is used. Furthermore, decaying crops and foodstuff can go back
into the biological and nutrient cycle. Recycling of animal fats, bioplastic, and
vegetable waste can be converted into oleo-chemicals which are also expected as
a part of circular economy [34, 35] (Fig. 10.4). Biorefinery approaches take advan-
tage of enzymes such as proteases, cellulases, amylases, and lipases to breakdown
the biomass to yield fermentable sugars, peptides, amino acids, and free fatty acids
as building blocks that are vital for valorization [36].
Though the idea of a similarity of biorefinery and petroleum refineries, the trans-
formations are vital (Fig. 10.5). Throughout the petrochemical refinery procedure,
fossil routes are converted into energy (e.g., fuel) also the list of the other products
like Plastic, chemicals, etc., In case of a biorefinery, a widespread range of bio-based

Fig. 10.4 Biorefinery approach to process municipal solid waste


250 S. Gade et al.

Fig. 10.5 Comparison of a biorefinery and a petroleum refinery; schematic based on [29]

goods like biochemicals, food and feed, bioenergy, biomaterials, etc. are formed from
the biomass feedstock.

10.4 Value-Added Products from MSW Using Biorefinery


Approaches

As discussed earlier, various thermochemical methods have been used for waste-
to-value processing and waste disposal, as established by several studies on steam
gasification [37], incineration [38], and pyrolysis [39]. As a distinct methodology, the
biorefinery approach is rapidly gaining traction for use in managing MSW and wastes
classified as wet waste, such as algal waste, etc. Under this process, biomass is first
separated from MSW and used in bioenergy production by biorefinery [40]. Gasifica-
tion [28] or and other technologies are established [41] and have been demonstrated
for use with biomass-derived feedstocks.
Biorefinery can be generally regarded as a complex of approaches that, (a) sepa-
rate biomass/materials that can be subject to processing, (b) primary processing
that helps derivatize feedstock which can be used in conventional fermentation,
(c) advanced processing which is aimed at generating value-added products, and
(d) processing to maximize extraction of usable compounds with commercial
value. Consequently, biorefinery is fundamentally distinct from the above-mentioned
10 Biorefinery Avenues for Processing Urban Solid Waste: Potential … 251

avenues of processing municipal solid waste. An interesting example of use of biore-


finery in conjunction with an indispensable industry is the integration of biorefinery
concept with paper processing plants. Wood pulping and paper production involve
manipulation of ligno-cellulosic contents of the starting material. Post lignin removal,
cellulose, and hemi-cellulosic content is capable of being harnessed for fermenta-
tion yielding fermentable sugars along with other value-added compounds including
pharmaceutical ingredients such as mannitol.
Another instance of biorefinery approach in managing MSW is the process engi-
neering for using a myriad methodological technique such as a microwave or ultra-
sonication assisted extraction, supercritical or hydrothermal methods for extraction
of compounds from food waste. Food waste from markets, domestic, and public
eatery waste, etc., is a veritable resource for not just starches, sugars, lipids, cellu-
lose, and lignin but also organic constituents, biomass for fermentation and poten-
tially salvaging of vitamins, co-factors, and biopolymers. Pretreatment with super-
critical aqueous media and subsequent enzymatic or catalytic treatment help to extract
fermentable sugars to further yield bioethanol, methane, lactic acid etc. Lipids are
already being utilized to generate biodiesel, which is primarily an ester-based fuel
and subsequently glycerol as well.
Various thermochemical methods have been used for waste-to-value processing
and waste disposal, as established by several studies on steam gasification [37], incin-
eration [38], and pyrolysis [39]. Biomass is first separated from MSW, it can be used
in bio-energy production by biorefinery [40]. Gasification [28] and other technolo-
gies are established [41]. As mentioned earlier, within the process of MSW disposal
under the biorefinery umbrella, biomass is first parted from MSW. Further, it can
be utilized to produce bioenergy through biorefinery [40], gasification [28], or other
technologies [41]. The bioenergy produced from such a source can be in the form of
biofuel, biogas, crude oil, and solid fuel. These alternatives are substitutes for fossil
fuels, which could help in reducing emission of greenhouse gas [42]. Furthermore,
bio ethanol, bio-methanol, methane, and even hydrogen can be produced from biogas
[43].
Utilization of waste food (organic waste) into a source of renewable energy has
converted the food sector to a possibly practicable carbon economy [44]. To achieve
circular bioeconomy, removal of bioactive and economically vital compounds from
the MSW and application of the wastes to generate biofuels are substitutions to the
remaining fossil fuels.
As described earlier, biomass is essentially any organic matter that is available
as a renewable source, including certain organic industrial wastes, forest wastes,
agricultural crop waste, and municipal waste. Specifically, referring to agricultural
crops, plants such as switchgrass, jatropha like plants are known for providing energy
sources, via fermentation and biorefinery. Primary biomass components vary by
type and source of biomass. Mainly lignocellulose, sugar, triglyceride, and starch
are known for biorefinery applications. Chemical structures of those compounds are
shown in Fig. 10.6 [45].
Food waste includes lipids, proteins, and carbohydrates. These compounds can be
used as a reactant for the manufacture of various bioactive compounds. For example,
252 S. Gade et al.

Fig. 10.6 Biorefinery value-added products

animal wastes can be the source of gelatin, polypeptides, and collagen. Dairy waste
contains lactose and casein which can be used after extraction treatment on source
[46]. Seafood waste is a source of protein isolates and hydrolysates. Kitchen waste
includes cellulose, flavonoids, anthocyanins, etc. [46, 47].

10.5 Recent Trends in Biomass Conversion

MSW compositions from Asian nations like India and China reveal a large segment
for food wastes, in addition to other organic wastes. As described earlier, a more
appropriate resource-recouping ideology is to adopt the bio-process/biorefinery
route. Literature review reveals certain insights into the contemporary trends in this
domain. While the area of municipal waste management is a stand-alone challenge,
the proposed integration with biorefinery and circular economy will set the stage for
a sustainable industry and civil management program.
Inclusivity and Fundamental Rethinking: As with a radical reimagining of a
concept, broad-sweeping changes need to be enacted for promoting and enabling
circular bioeconomy. To this effect, floating an educational outline for the propaga-
tion of the concept is needed. To achieve the goal, it is vital to extend and venture
beyond any single domain and implement design thinking for innovations that may
10 Biorefinery Avenues for Processing Urban Solid Waste: Potential … 253

bring commercial viability. True multi-disciplinary blending may be required to


improve success in the MSW-Biorefinery field. Experts from areas such as law,
policy-making, industry, STEM fields, and even social sciences would be required
to work together to perceive and purview issues in the field and to devise custom
solutions for the same.
Global Partnerships: Unsurprisingly, the key to success in complex, globally rele-
vant problems is collaboration. Due to the unique geographical, cultural, or scientific
dispositions, some nations can be better situated for economic and timely solutions to
specific issues. For instance, if Bulgaria were to exploit composting instead of relying
exclusively on landfill strategies for MSW management, an alliance with Germany as
a regional partner would be highly effective. Beyond practical alliances, the collabo-
rations are envisioned to educate all stakeholders to the nature of natural challenges
occurring, which are the primary drivers of this sea change [48]. Climate change and
efforts to mitigate the same should be the prime focus of such collaborations.
Goal Setting: Not unlike personal development strategies, nations need to catego-
rize goals for achieving sustainability and circular bioeconomy. The establishment of
bioeconomy linked to sustainable development goals (SDG) can vary substantially,
depending on the strategic objectives adopted by a given nation for its development
and goal. It follows that the nation’s SDG program is intricately linked to the choice
for enabling bioeconomy [49].
Administrative Accountability: As with inclusivity, lawmakers and administra-
tions are deeply involved with enabling and maintaining a circular bioeconomy.
The government must take initiatives in MSW regulation and drive nation-wide
responsibility in MSW collection, classification, and disposal route.
Green Route: Beyond cliché, the process of bioeconomy should follow a green
route, at least in part. The green pathway is primarily to promote biomass conver-
sions such as waste-to-material and waste-to-economy. While the latter is broad in
scope, the idea of enacting value-added-services/products is central to this theme.
The accompanying theme is hence to minimize waste in the first place, judicious use
of resources, re-use and recycling of resources as a natural pathway in civil decisions
[8].
The summarized snapshot of the insights gained from this review is depicted in
Fig. 10.7, showing country/region-specific MSW generation pattern. Depending on
the proportions of specific wastes, the regional MSW management mechanism can
be optimized to serve, for instance, higher organic waste by the biorefinery pathway.
Organic wastes can be sorted for maximum efficiency of enzymatic pre-processing
and degradation, prior to fermentation or bio-esterification (biodiesel synthesis). In an
ideal example of multi-modal integration, heat released from incineration of wastes
can be used to drive boilers/distillation unit for separation of high-grade biodiesel.
Resulting gaseous wastes can be systematically processed to reduce CO2 emissions
as well as produce value-added products (VAPs) such as urea, glycidol, and glycerol
carbonate. Most VAPs are circularly re-employed in civil and agricultural domains
which feedback into the loop. A well-regulated MSW management pathway and
auxiliary domains, as indicated by several initiatives and studies, will ultimately
reduce waste production and enable sustainability [50–58].
254 S. Gade et al.

Fig. 10.7 Municipal solid waste management with value-added products in a circular bioeconomy

10.6 Conclusions

The global economy and resource pool situation may have long passed their peak
and with the specter of fossil fuel exhaustion, urban localization of population, and
unmanageable municipal waste scenarios, serious attention needs to be devoted to
managing the municipal solid wastes. This chapter highlights and emphasizes the
need for a circular economy revolving around urban waste and specifically a circular
bioeconomy, owing to various value-added products that may be reaped from judi-
cious management of certain MSW materials. Various MSW management models
with the outlook of circular bioeconomy have been broadly studied. Biorefinery plays
a pivotal role in circular economy. With proper methodology, technology biorefinery
with a zero-waste goal can be achieved. It should be implemented in various indus-
trial areas, societies, municipal garbage collection areas, etc. Sustainability of the
mechanism of MSW management and the establishment of the circular economy will
reduce the impact of MSW pressures on the taxpayer as well as the municipalities
in question. The successful management of MSW in Indian cities like Rishikesh and
Indore lays a roadmap for the rest of the country to implement best MSW manage-
ment practices and maximize utilization of the biorefinery pathway to a sustainable
and better economy. Beyond India, a broader study of nations like China, Japan, the
United States, and countries within the European Union provide an insight into the
attitudes of their citizens as well as fundamental differences which result in differing
patterns of consumption and therefore wastage.
As proposed in this work, multiple overlapping approaches may be entertained to
institutionalize the circular bioeconomy, for instance, adoption of the carbon seques-
tration technologies as illustrated in Fig. 10.2 in combination with not just MSW
incineration but also biodiesel fueled industries to minimize the impact of greenhouse
10 Biorefinery Avenues for Processing Urban Solid Waste: Potential … 255

gases while safely managing anthropogenic organic wastes. Biodiesel production in


turn generates glycerol and other compounds which can be processed to obtain higher
value-added products.
A detailed study of the existing data and literature show that the widespread
adoption of biorefineries to enable circular bioeconomy is still largely affected by
socio-economic perceptions and social “acceptability,” virtually creating a barrier
for the path ahead. Such a stalemate requires intervention at the highest levels and
hence support from policymakers and scientists alike. The potential value provided
by organic MSW and associated biomass has taken the biorefinery role past its
boundaries, well on the way to installing the much-desired circular bioeconomy and
adding value to even secondary products.

References

1. Guan, Y., Yan, J., Shan, Y., Zhou, Y., Hang, Y., et al. (2023). Nature Energy, 8, 304–316.
2. Tollefson, J. (2022). What the war in Ukraine means for energy, climate and food. In Nature
news feature (pp. 232–233). Nature Publishing Group.
3. Edenhofer, O. (2015). Climate change 2014: Mitigation of climate change. Cambridge
University Press.
4. Change, I. C. (2014). Contribution of working group III to the fifth assessment report of the
intergovernmental panel on climate change, 1454, 147.
5. Peter, D., Rathinam, J., & Vasudevan, R. T. (2022). Biotechnology for zero waste (pp. 409–420).
6. Ferreira, A., Ribeiro, B., Ferreira, A. F., Tavares, M. L. A., & Vladic, J., et al. (2019). Biofuels,
Bioproducts and Biorefining, 13, 1169–1186
7. Azapagic, A. (2010). Sustainable development in practice (pp. 261–325).
8. Ding, Y., Zhao, J., Liu, J.-W., Zhou, J., Cheng, L., et al. (2021). Journal of Cleaner Production,
293, 126144.
9. Gadot, Y. (2022). The landfill of early Roman Jerusalem: The 2013–2014 excavations in area
D3. Penn State University Press.
10. Ozturk, M., & Dincer, I. (2020). Greenhouse Gases: Science and Technology, 10, 855–864.
11. Denison, R. A., & Silbergeld, E. K. (1988). Risk Analysis, 8, 343–355.
12. Lakhdar, A., Iannelli, M. A., Debez, A., Massacci, A., Jedidi, N., & Abdelly, C. (2010). Journal
of the Science of Food and Agriculture, 90, 965–971.
13. Zheljazkov, V. D., & Warman, P. R. (2004). Journal of Environmental Quality, 33, 542–552.
14. Elmaslar, Ö. E. (2015). Environmental Progress & Sustainable Energy, 34, 1372–1378.
15. Hamburg, K. T., Haque, C. E., & Everitt, J. C. (1997). Canadian Geographies/Géographies
canadiennes, 41, 149–165.
16. Giordano, N. (1993). European Environment, 3, 14–17.
17. Periathamby, A. (2011). Waste. In T. M. Letcher & D. A. Vallero (Eds.) (pp. 109–125).
Academic Press.
18. Rawat, S., & Daverey, A. (2018). Environmental Engineering Research, 23, 323–329.
19. Burnley, S. J., Ellis, J. C., Flowerdew, R., Poll, A. J., & Prosser, H. (2007). Resources
Conservation and Recycling, 49, 264–283.
20. Dennison, G. J., Dodd, V. A., & Whelan, B. (1996). Resources Conservation and Recycling,
17, 227–244.
21. Miezah, K., Obiri-Danso, K., Kádár, Z., Fei-Baffoe, B., & Mensah, M. Y. (2015). Waste
Management, 46, 15–27.
22. Ojeda-Benítez, S., Vega, C. A.-D., & Marquez-Montenegro, M. Y. (2008). Resources,
Conservation and Recycling, 52, 992–999.
256 S. Gade et al.

23. Gu, B., Wang, H., Chen, Z., Jiang, S., Zhu, W., et al. (2015). Resources Conservation and
Recycling, 98, 67–75.
24. Qu, X.-Y., Li, Z.-S., Xie, X.-Y., Sui, Y.-M., Yang, L., & Chen, Y. (2009). Waste Management,
29, 2618–2624.
25. Damanhuri, E., Wahyu, I. M., Ramang, R., & Padmi, T. (2009). Journal of Material Cycles
and Waste Management, 11, 270–276.
26. Matsumoto, S. (2011). Resources Conservation and Recycling, 55, 325–334.
27. Zhang, D., Keat, T. S., & Gersberg, R. M. (2010). Waste Management, 30, 921–933.
28. Munir, M. T., Mardon, I., Al-Zuhair, S., Shawabkeh, A., & Saqib, N. U. (2019). Renewable
and Sustainable Energy Reviews, 116, 109461.
29. Clark, J., & Deswarte, F. (2015). Introduction to chemicals from biomass (pp. 1–29).
30. Gomez, E., Rani, D. A., Cheeseman, C. R., Deegan, D., Wise, M., & Boccaccini, A. R. (2009).
Journal of Hazardous Materials, 161, 614–626.
31. Sanlisoy, A., & Carpinlioglu, M. O. (2017). International Journal of Hydrogen Energy, 42,
1361–1365.
32. Changming, D., Chao, S., Gong, X., Ting, W., & Xiange, W. (2018). Waste Management, 77,
373–387.
33. Fabry, F., Rehmet, C., Rohani, V., & Fulcheri, L. (2013). Waste and Biomass Valorization, 4,
421–439.
34. Giampietro, M. (2019). Ecological Economics, 162, 143–156.
35. Stegmann, P., Londo, M., & Junginger, M. (2020). Resources Conservation & Recycling: X,
6, 100029.
36. Ghumra, D. P., Rathi, O., Mule, T. A., Khadye, V. S., Chavan, A., et al. (2022). Biofuels.
Bioproducts and Biorefining, 16, 877–890.
37. Arena, U. (2012). Waste Management, 32, 625–639.
38. Abd Kadir, S. A. S., Yin, C.-Y., Rosli Sulaiman, M., Chen, X., & El-Harbawi, M. (2013).
Renewable and Sustainable Energy Reviews, 24, 181–186.
39. Velghe, I., Carleer, R., Yperman, J., & Schreurs, S. (2011). Journal of Analytical and Applied
Pyrolysis, 92, 366–375.
40. Bolan, N. S., Thangarajan, R., Seshadri, B., Jena, U., Das, K. C., et al. (2013). Bioresource
Technology, 135, 578–587.
41. Dehkordi, S., Taghipour, A., Ferdowsi, A., Shumal, M., & Dehnavi, A. (2019). Renewable and
Sustainable Energy Reviews, 119, 109586.
42. You, S., Tong, H., Armin-Hoiland, J., Tong, Y. W., & Wang, C.-H. (2017). Applied Energy,
208, 495–510.
43. Furtado Amaral, A., Previtali, D., Bassani, A., Italiano, C., Palella, A., et al. (2020). Energy,
203, 117820.
44. Girotto, F., Alibardi, L., & Cossu, R. (2015). Waste Management, 45, 32–41.
45. Mailaram, S., Kumar, P., Kunamalla, A., Saklecha, P., & Maity, S. K. (2021). Sustainable fuel
technologies handbook (51–87). https://doi.org/10.1016/B978-0-12-822989-7.00003-2
46. Anal, A. K. (2017). Food processing by-products and their utilization (pp. 1–10).
47. Tsouko, E., Alexandri, M., Fernandes, K. V., Guimarães Freire, D. M., Mallouchos, A., &
Koutinas, A. A. (2019). Food Technol Biotechnol, 57, 29–38.
48. Aguilar, A., & Patermann, C. (2020). New Biotechnology, 59, 20–25.
49. Calicioglu, Ö., & Bogdanski, A. (2021). New Biotechnology, 61, 40–49.
50. Sacchi, S., Lotti, M., & Branduardi, P. (2021). New Biotechnology, 60, 72–75.
51. Bhattacharya, R. R. N., Chandrasekhar, K., Deepthi, M. V., Roy, P., & Khan, A. (2018). TERI-
challenges and opportunities in plastic waste management in India (p. 20).
52. Mongkhonsiri, G., Charoensuppanimit, P., Anantpinijwatna, A., Gani, R., & Assabumrungrat,
S. (2020). Journal of Cleaner Production, 255, 120278.
53. Sarangi, P. K., Singh, A. K., Sonkar, S., Shadangi, K. P., Srivastava, R. K., Gupta, V. K., Parikh,
J., Sahoo, U. K., & Govarthanan, M. (2023). Industrial Crops and Products, 205, 117488.
54. Department for Environment Food and Rural Affairs. (2013). Waste technology briefs (p. 56).
www.defra.gov.uk
10 Biorefinery Avenues for Processing Urban Solid Waste: Potential … 257

55. IPPC reference document on best available technologies for waste incineration, August 2006.
56. United Nations, Department of Economic and Social Affairs, Population Division. (2019).
World urbanization prospects: The 2018 revision (ST/ESA/SER.A/420). United Nations.
57. Hoornweg, D., & Bhada-Tata, P. (2012). What a waste: A global review of solid waste
management. World Bank—Urban Development Series.
58. Ministry of Housing and Urban Affairs, Government of India. (2021). Circular economy in
municipal solid and liquid waste (p. 112). MoHUA/2021/June/30.
Chapter 11
Pyrolytic Conversion of Heterogenic
Natural Waste Biomass from Rural
Communities with Concomitant
Valorization

M. Anil Kumar, Pareshkumar G. Moradeeya, K. Manikanda Bharath,


P. Jakulin Divya Mary, and K. S. Giridharan

Abstract The challenge of waste disposal and management has become more promi-
nent with increasing awareness. In rural areas, natural waste management is carried
out within available space constraints. The conventional methods used for solid
waste disposal include burning, landfilling, and composting, often regarded as the
most effective and widely used practices. There is a significant concern over the
adverse environmental consequences of implementing non-sustainable practices in
solid waste management. This chapter aims to comprehensively analyse the many
sources of solid waste, focusing on lignocellulosic biomasses originating from rural
areas, mainly within the Indian sub-continent. Pyrolytic conversion is a thermal
process that transforms waste materials into three distinct forms: solid residue known
as char, liquid fuel, and gaseous emissions. Recent research has concentrated on
using food waste as an energy source (e.g. for the manufacture of bioethanol and
biodiesel) rather than disposing of and decomposing it. Utilizing a biorefinery or
biotechnology, organic waste can also be used to produce valuable organic chemi-
cals, such as succinic acid and/or bioplastics. The chapter further provides a detailed
explanation of mapping secondary data from different departments and contrasts it
with previously conducted research studies. This chapter is intended to summarise
the facts on the efficient use of biochar for pollution control and environmental
management.

Keywords Biochar · Pyrolysis · Rural · Solid wastes · Thermo-chemical ·


Valorization

M. Anil Kumar (B) · K. Manikanda Bharath · P. Jakulin Divya Mary · K. S. Giridharan


Centre for Rural and Entrepreneurship Development, National Institute of Technical Teachers
Training and Research, Chennai, Tamil Nadu 600113, India
e-mail: anilkumar@nitttrc.edu.in
P. G. Moradeeya
Department of Environmental Science and Engineering, Marwadi University, Rajkot,
Gujarat 360003, India

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2024 259
G. Baskar et al. (eds.), Circular Bioeconomy Perspectives in Sustainable Bioenergy
Production, Energy, Environment, and Sustainability,
https://doi.org/10.1007/978-981-97-2523-6_11
260 M. Anil Kumar et al.

11.1 Introduction

The effective and sustainable management of biomass waste has become more signif-
icant within the framework of environmental preservation and the transition towards
an environmentally friendly economy [51]. Rural areas often generate significant
volumes of different natural waste biomass originating from a range of sources,
including agriculture, forestry, and other organic materials. Inadequate handling of
biomass waste can lead to the contamination of the environment, the release of
greenhouse gases, and the potential for health risks [43]. However, it is important
to note that these waste materials possess the potential to be converted into useful
goods through the utilization of pyrolytic processes. Various agriculture and forestry
methods in rural communities worldwide result in significant amounts of organic
waste biomass being generated [27]. Rural regions often exhibit a diverse range of
biomass resources, including agricultural residues, forest debris, and organic wastes.
According to Schipfer et al. [60], inadequate management of resources can lead
to both environmental contamination and economic losses. Pyrolysis is a chemical
process that facilitates the conversion of various biomass feedstocks into three distinct
products, namely biochar, bio-oil, and syngas. The implementation of this conver-
sion yields numerous benefits for both the environment and the local economy, as
evidenced by studies conducted by Chormare et al. [20] and Seah et al. [61]. The
accumulation of biomass poses environmental concerns and represents a missed
opportunity for the valorization of resources [14]. Pyrolytic conversion is a thermo-
chemical process characterized by the application of heat to biomass in the absence
of oxygen, as described by Sobamowo and Ojolo [62]. The aforementioned study
conducted by Kassim et al. [33] demonstrates the promising capability of this proce-
dure to transform diverse forms of organic waste biomass into valuable commodities,
while also efficiently mitigating environmental issues.
The study conducted by Bandgar et al. [10] examined the thermochemical process
of various natural waste biomass derived from rural areas, as well as the following
valorization of the resulting products. The term “heterogenic” is used to empha-
size the diverse and mixed nature of biomass, encompassing agricultural wastes,
forest trash, and other organic components [13]. The implementation of pyrolysis
provides these communities with an effective means of managing their trash in
an environmentally sustainable manner, while also taking advantage of economic
opportunities through the extraction of valuable products from the process [5]. This
article examines the most recent advancements in pyrolysis technology, the opti-
mization of the process, and the utilization of the resulting products. The paper
also investigates the challenges and opportunities associated with the implementa-
tion of pyrolytic conversion systems in rural regions, taking into account the unique
situations, socioeconomic considerations, and environmental variables [53].
The global yearly production of agricultural biomass is estimated to be over 140
billion metric tonnes. A significant quantity of biomass has the potential to produce
enormous quantities of valuable energy and minerals. A considerable proportion
of the global populace residing in economically disadvantaged nations persists in
11 Pyrolytic Conversion of Heterogenic Natural Waste Biomass from Rural … 261

confronting a dearth of energy accessibility, impacting approximately 1.6 billion


individuals. One possible approach involves the conversion of agricultural biomass
waste into energy, as proposed by Mishra et al. [46]. The aforementioned method-
ology possesses the capacity to significantly diminish dependence on non-renewable
energy sources, alleviate the emission of greenhouse gases, and offer a viable and
enduring energy solution. The utilization of biomass waste has considerable promise
as a feasible source of raw materials for both extensive industrial manufacturing
and localized small-scale businesses [12]. Biomass include residual materials from
agricultural and animal husbandry practices, such as discarded stalks, straws, leaves,
roots, husks, nut and seed shells, scrap wood, and animal excrement. The utilization
of waste biomass is highly regarded due to its abundant availability, renewable nature,
and cost-efficiency [32]. Nations are currently engaged in the active exploration of
alternative energy sources as a means to reduce the release of carbon dioxide and
other greenhouse gases, in response to the worldwide endeavour to address and miti-
gate the impacts of climate change. The utilization of biomass as a source of energy
generation has minimal implications for the global carbon cycle, thereby aiding in
the reduction of our reliance on fossil fuels. The simultaneous advantages of this
duality are beneficial for both the economy and the environment.
Pyrolytic conversion refers to a thermochemical process wherein biomass under-
goes heating under conditions of limited oxygen, leading to the decomposition of
the biomass into various by-products such as biochar, bio-oil, and syngas [7]. The
heterogeneous natural waste biomass found in rural areas encompasses a range of
organic waste components, including agricultural residues, forest remnants, animal
excrement, and similar items [11, 33]. The pyrolysis process entails the concurrent
generation of many useful chemicals. For instance, the utilization of biochar as a
soil amendment, the augmentation of bio-oil for the production of biofuels, and the
harnessing of syngas for the generation of heat and power have been identified as
potential applications [14].
The pyrolytic conversion process entails the conversion of diverse natural waste
biomass derived from rural regions, with the concomitant extraction of value from this
biomass [71]. This study examines the latest advancements in pyrolysis technology,
the process of selecting appropriate feedstock, the optimization of the pyrolysis
process, and the effective utilization of pyrolysis products. Furthermore, this study
investigates the social and environmental implications associated with the utiliza-
tion of pyrolytic conversion systems in rural regions [13]. There is a rising concern
regarding the degradation of the environment and a pressing need for sustainable
energy solutions. As a result, there has been a notable increase in the exploration of
renewable and eco-friendly alternatives for the utilization of biomass. Biomass, as an
abundant and diversified natural resource, holds considerable promise in its capacity
to contribute to a sustainable energy framework. The utilization of waste biomass
derived from agricultural, forestry, and organic origins is widespread among rural
communities.
However, the efficient management and utilization of this diverse biomass are
often overlooked, leading to environmental pollution, and missed opportunities for
resource recovery [11, 33, 63]. The utilization of pyrolytic conversion technology
262 M. Anil Kumar et al.

offers a feasible approach to efficiently address the simultaneous issues of waste


management and energy demands in rural areas [22]. Pyrolysis is a thermally driven
chemical process whereby biomass undergoes transformation into several beneficial
outputs, such as charcoal, bio-oil, and syngas, through exposure to elevated temper-
atures inside an oxygen-limited environment. This methodology has several advan-
tages, including the production of biochar, which can be utilized as a soil supplement
to enhance soil fertility and sequester carbon, the conversion of bio-oil into biofuels,
and the utilization of syngas for heat and power generation. The utilization of various
types of natural waste biomass derived from rural communities through the process
of pyrolytic conversion presents a unique opportunity for the establishment of a
circular economy [19]. The aforementioned methodology entails the transformation
of waste materials into valuable resources, hence promoting the principles of environ-
mental sustainability and facilitating the growth of local economies. The utilization
of underutilized waste biomass is an opportunity for rural communities to reduce
their reliance on fossil fuels, mitigate the emission of greenhouse gases, and enhance
their socioeconomic resilience.
The main aim of this chapter is to provide important insights into the pyrolytic
conversion of diverse natural waste biomass derived from rural communities [27].
Furthermore, it aims to emphasize the importance of simultaneous valorization within
the framework of sustainable waste management and resource utilization. Further-
more, the primary objective of this study is to examine the recent progress and
scholarly investigations pertaining to the pyrolytic conversion of various types of
organic waste biomass derived from rural areas. Additionally, the study seeks to
assess the potential of this conversion process in terms of generating economic value,
as highlighted by my researcher [32, 48, 56, 63].
The objective of this study is to provide a comprehensive analysis of the poten-
tial advantages and drawbacks associated with this distinctive technique through
an examination of prominent research findings, technical advancements, and chal-
lenges. Furthermore, the main aim of this study is to identify regions with a dearth
of information and possible directions for future research. These endeavours possess
the capacity to augment the sustainable utilization of biomass resources, stimulate
rural development, and promote the advancement of environmental sustainability.
The main aim of this study is to provide a significant contribution to the advance-
ment of renewable energy technologies, the improvement of rural livelihoods, and
the conservation of natural resources for the benefit of future generations.

11.2 Circularity of Biomass Waste and Plastic Waste


via Co-pyrolysis

The increasing global concern about environmental degradation and the deple-
tion of finite resources has spurred the exploration of sustainable strategies for
waste management and energy production [48, 60]. The handling of biomass waste
11 Pyrolytic Conversion of Heterogenic Natural Waste Biomass from Rural … 263

and plastic rubbish presents significant challenges in the field of waste manage-
ment. Nevertheless, the utilization of innovative technologies like co-pyrolysis [43]
presents the possibility of circularity. Circularity refers to a deliberate approach aimed
at creating self-sustaining systems that effectively reuse or convert waste materials
into useful resources. This technique serves to reduce environmental damage and
promote the adoption of sustainable practices. According to Bhatt et al. [12], pyrol-
ysis, a thermochemical process, offers a promising solution for effectively tackling
the challenges associated with both biomass and plastic waste. By means of co-
processing the waste streams, the method of pyrolysis has the capacity to diminish
the quantities of garbage and concurrently produce useful commodities, including
biofuels, charcoal, and valuable chemicals. Co-pyrolysis refers to a thermochemical
process in which biomass and plastic waste are subjected to pyrolysis either concur-
rently or consecutively [28, 49]. The aforementioned procedure yields significant
outputs, such as biofuels, charcoal, and various chemical compounds. This process
enables the efficient utilization of biomass and plastic waste, thereby limiting envi-
ronmental impacts and producing a self-sustaining system in which waste materials
are transformed into valuable resources.

11.2.1 Importance of Circular Bioeconomy

The contemporary global community is presently faced with complex challenges,


including climate change, resource depletion, and environmental degradation. The
circular bioeconomy has emerged as a transformative solution within this context,
effectively addressing the pressing concerns at hand [19]. The concept of the circular
bioeconomy combines the principles of circular economy with the responsible
utilization of biological resources, presenting a holistic and innovative strategy for
promoting a resilient and regenerative society [9]. The circular bioeconomy places
significant focus on the efficient utilization of renewable biological resources, such as
biomass, agricultural produce, and organic waste, with the aim of generating value-
added products, energy, and services [12]. The main goal is to reduce reliance on finite
resources, mitigate the impacts of climate change, promote biodiversity conserva-
tion, and foster the creation of environmentally sustainable innovations. The circular
bioeconomy seeks to transition away from a linear framework typified by the “take-
make-dispose” approach, towards a regenerative system that prioritizes the reuse,
repurposing, and regeneration of resources [19]. The circular bioeconomy’s impor-
tance in addressing pressing global issues has been discussed by Lopes et al. [39].
This research investigates the various aspects of circularity and bio-based practices,
highlighting their potential benefits for the environment, economy, and society. The
assessment will additionally assess the alignment between the circular bioeconomy
and sustainable development goals, as well as its role in bolstering resilience in the
context of worldwide crises. Through a comprehensive examination of existing liter-
ature and the integration of findings from multiple scholarly investigations, the prin-
cipal aim of this research endeavour is to provide a comprehensive understanding of
264 M. Anil Kumar et al.

the significance of the circular bioeconomy and its potential to foster transformative
and innovative progress towards a sustainable future.

11.2.1.1 Significance of Globular Bioeconomy

Three key concepts can be employed to mitigate the ecological consequences of


human activities: the circular economy, the bioeconomy, and the bioeconomy of
circularity [65]. These concepts can be employed in the pursuit of establishing
a sustainable human civilization and safeguarding the environment. The circular
economy is characterized by the utilization of non-hazardous materials and energy
sources that do not exhaust natural resources. It aims to facilitate the reduction, mini-
mization, and eventual closure of all material resource loops. Within the societal,
economic, and environmental framework, the concept of “bioeconomy” pertains to
the transformation of renewable biological resources into biofuels, bioproducts, and
biopower in a manner that ensures long-term sustainability [58, 65]. The circular bioe-
conomy, alternatively referred to as the bio-based circular carbon economy, repre-
sents a comprehensive framework that effectively integrates the benefits associated
with both of these ideas. This paradigm is alternatively referred to as the circular
carbon economy. The circular bioeconomy is strongly dependent on carbon as a
fundamental constituent. According to Seah et al. [61] and Tan and Lamers [67], in
order to achieve a sustainable future, it is imperative to undertake measures such as
the implementation of strategies to “close the carbon cycle” and the development of
carbon sinks inside the technosphere.

11.2.2 Biomass Waste

Biomass waste refers to organic resources that are discarded or classified as waste
materials due to various human activities, such as agriculture, forestry, industrial
processes, and municipal waste [46]. The organic constituents being examined are
derived from organisms that are presently extant or have recently become extinct,
encompassing flora, fauna, and microorganisms. In contrast to the protracted produc-
tion process spanning millions of years from decomposed organic material that fossil
fuels experience, biomass waste is a renewable and sustainable resource that may
be continuously supplied by inherent natural mechanisms. Biomass waste, a signif-
icant by-product arising from our activities and natural phenomena, exhibits enor-
mous promise as a valuable asset for the promotion of sustainable development and
the preservation of the environment. Biomass waste encompasses a diverse range
of organic materials that are derived from living creatures or have been recently
produced by them. This category encompasses a range of materials, such as agri-
cultural residues, forestry by-products, organic substances, and other biodegradable
entities. In contrast to finite fossil fuels, biomass waste presents itself as a viable and
11 Pyrolytic Conversion of Heterogenic Natural Waste Biomass from Rural … 265

ecologically sustainable option due to its renewability and lack of contribution to


climate change when subjected to combustion [43].
Only forty per cent of the trash that is produced by the agriculture industry
each year is recycled [35, 69]. This results in an annual production of around 140
billion tonnes of biomass leftovers. Only 4.5% of the lignocellulosic biomass that is
produced each year from forestry and agricultural leftovers is exploited [61]. This
waste contributes approximately 181.5 billion tonnes of lignocellulosic biomass. In
addition to the biomass that is obtained from the agricultural and forestry industries,
an additional significant source of biomass is the organic waste that is produced as
a result of human activities. As the human population continues to rise, the amount
of garbage produced across the globe has also increased, becoming a problem on a
global scale. According to Maalouf and Mavropoulos [40], billion tonnes of munic-
ipal solid trash are generated every year all over the world. However, a sizeable
amount of this rubbish is still not collected by local governments. According to Seah
et al. [61], the collected municipal solid trash, only around 19% is recycled, 11%
is transformed to electricity, and the other 70% is placed in landfills and dumpsites.
Organic trash, which is also known as biodegradable waste, is a word that is used
to describe the undesired natural products that are produced by animals and plants.
Human waste, sewage, food waste, green trash, abattoir waste, wet market waste, and
other types of waste are included in this category of undesirable items. These wastes
pose a risk to human health if they are not effectively managed because they can
cause unpleasant odours to be released, they can encourage the growth of pathogenic
coliform bacteria, and they can make it easier for illnesses to be passed from one
person to another [8, 52, 68]. An example of this would be the dangerous bacteria
Escherichia coli (EC), which produces the toxins and has been linked to a significant
number of cases of food-borne illness in human beings. The majority of EC origi-
nates from ruminants, with cattle being the most important contributor. Raw meat
items, along with freshly harvested fruits and vegetables, have been discovered to be
contaminated with EC. Diseases brought on by EC range from relatively harmless
bloody diarrhoea (haemorrhagic colitis) to the potentially lethal haemolytic uremic
syndrome [30, 55]. The landfills receive a significant amount of organic waste, most
of which is comprised of discarded food. According to research conducted by the
Food and Agriculture Organisation of the United Nations [31, 59], each year, the
food supply chain results in the loss of 1.3 billion tonnes of food, which is equiv-
alent to one-third of the world’s total food output. According to Seah et al. [61],
the spoiled food is either thrown out at the consumer level or lost at the harvesting
and processing stages. According to research by Melikoglu et al. [45] and Schanes
et al. [59] up to 95% of the food that is thrown away ends up in landfills, where it
serves as a source of methane and other greenhouse gases. According to Schanes
et al. [59], this not only helps to ensure the availability of food but also contributes
to the acceleration of climate change. Therefore, it is of the utmost importance that
global biomass wastes (including forestry residues, agricultural residues, and organic
wastes) be treated in an appropriate and secure manner in order to minimise the risks
to people’s health, decrease their impact on the ecosystem, prevent the depletion of
resources, and recycle plant material waste products [23, 37, 70].
266 M. Anil Kumar et al.

The environmental challenges linked to biomass waste encompass the build-up


and improper disposal of this waste, leading to the release of greenhouse gases,
contamination of soil and water, and degradation of habitats. However, the diligent
management and utilization of biomass waste can offer various benefits, such as the
production of sustainable energy, improvement of soil fertility, reduction of waste,
and mitigation of climate change. When effectively managed, biomass waste can
be utilized as a valuable resource, offering various environmental, economic, and
social benefits [39]. Communities have the potential to adopt more sustainable and
circular resource use and waste management strategies by using biomass waste. The
present study investigates diverse approaches for the utilization of biomass waste,
encompassing biomass-to-energy technology, composting, and biochar production.
Moreover, this review aims to examine the diverse challenges faced in the manage-
ment of biomass waste, spanning technological, environmental, and policy-related
considerations.

11.2.3 Plastic Waste

The problem of plastic trash has emerged as a significant environmental issue


in modern times, with far-reaching impacts on ecosystems, human welfare, and
the global environment as a whole. Plastics, widely recognized for their dura-
bility and versatility, have become essential components of our daily lives, serving
various purposes in packaging, consumer goods, electronics, and various industrial
domains [12, 28]. However, the widespread utilization and improper disposal of
these substances have resulted in the accumulation of plastic waste in many settings,
including landfills, oceans, and natural ecosystems, hence leading to substantial
ecological consequences.
Plastic waste has been piling up at an alarming rate in the contemporary world
as a consequence of the meteoric rise in the manufacture of plastic and the pitiful
rate at which plastic gets recycled. According to Maquart et al. [44], around 55%
of the world’s plastic trash is disposed of in landfills or the natural environment.
Every year, one million tonnes of waste are discharged into the ocean, with the
majority of that waste coming from Asian countries [50]. According to estimates
provided by Lebreton and Andrady [36], the global production of MPW in 2015 was
anywhere from 60 and 99 million metric tonnes (Mt). This quantity might triple to
155–265 Mt y−1 by the year 2060 if everything continues as it has been up until that
point [36]. Even in the next years, the load of future MPW will still be dispropor-
tionately large in the continents of Africa and Asia. According to Kibria et al. [34],
the large quantity of waste plastics has a negative impact on the ecosystem due to
the contamination of the soil caused by landfilling, the contamination of the marine
environment caused by ocean dumping, and the pollution of the air caused by open
dumping. Plastics have been responsible for a rise in the amount of pollution caused
by plastics and have contributed to a wide variety of significant environmental issues.
11 Pyrolytic Conversion of Heterogenic Natural Waste Biomass from Rural … 267

Plastic wastes can seep from landfills and bioaccumulate; these wastes contain
flame retardants, bisphenol A (BPA), phthalates, and heavy metals such as lead
and cadmium. Because of this, the consumption of marine species by humans has
been linked to an increased risk of developing cardiovascular disease, reproductive
problems, and obesity [16]. Therefore, it is absolutely necessary to implement plastic
waste management solutions such as recycling or conversion to energy so that each
plastic product can be used to its maximum potential and circularity can be achieved.
The nation produces 160,038.9 TPD of solid waste, of which 152,749.5 TPD is
collected with a collection efficiency of 95.4%, 79,956.3 TPD (50%) is processed,
and 29,427.2 TPD (18.4%) is landfilled. The collection efficiency for solid waste is
95.4% and unaccounted for is 50655.4 TPD, equal to 31.7% of the total quantity of
solid waste produced. Its state-wise details are provided in Table 11.1.
An analysis has been conducted on the data pertaining to solid waste management
(SWM) for the period spanning from 2015 to 2021 is given in Table 11.2.
The extended durability of plastics in the natural environment, with several
centuries of degradation processes, has prompted apprehension regarding their
detrimental impacts on fauna, aquatic organisms, and the ecological food web.
Microplastics, which are little plastic particles that emerge due to the degrada-
tion of larger plastic objects, have permeated even the most isolated regions of the
planet, presenting potential hazards to both human well-being and biodiversity. This
study investigates the diverse environmental consequences of plastic trash, encom-
passing marine pollution, habitat degradation, and the potential implications for
climate change. Furthermore, this analysis examines the socioeconomic dimensions
of managing plastic trash, considering the financial implications of waste manage-
ment systems and the possibilities for implementing circular economy strategies [19].
The text also delves into examining the role played by technology and innovation in
managing plastic waste. This includes an exploration of the progress made in recy-
cling technologies and processes that convert plastic into energy. Gaining compre-
hension of the obstacles and investigating prospective remedies is vital to formulate
efficacious policies, advancing public consciousness, and cultivating cooperative
endeavours to achieve a sustainable future devoid of plastic.

11.2.4 Co-pyrolysis of Biomass and Plastic

The co-pyrolysis of biomass and plastic is a recently developed and promising


thermochemical process that involves the simultaneous or sequential pyrolysis of
these different types of feedstocks [1, 49]. Pyrolysis is a thermochemical conversion
process that takes place in the absence of oxygen, resulting in the production of useful
by-products such biochar, bio-oil, and syngas. Number of benefits can be obtained
from the simultaneous thermal breakdown of plastic waste and biomass, including
the co-production of energy and high-value products, as well as possible uses that
effectively solve waste management issues and advance the development of a more
sustainable and circular economic model [46]. Table 11.3 gives the different reports
268 M. Anil Kumar et al.

Table 11.1 Overall solid waste management status in India


No. State TPD of wastes
Solid waste Collected Treated Landfilled
generated
1. Andhra Pradesh 6898 6829 1133 205
2. Arunachal 236.51 202.11 Nil 27.5
Pradesh
3. Assam 1199 1091 41.4 0
4. Bihar 4281.27 4013.55 Not No
provided
5. Chhattisgarh 1650 1650 1650 0
6. Goa 226.87 218.87 197.47 22.05
7. Gujarat 10,373.79 10,332 6946 3385.82
8. Haryana 5352.12 5291.41 3123.9 2167.51
9. Himachal 346 332 221 111
Pradesh
10. Jammu & 1463.23 1437.28 547.5 376
Kashmir
11. Jharkhand 2226.39 1851.65 758.26 1086.33
12. Karnataka 11,085 10,198 6817 1250
13. Kerala 3543 964.76 2550 Not provided
14. Madhya Pradesh 8022.5 7235.5 6472 763.5
15. Maharashtra 22,632.71 22,584.4 15,056.1 1355.36 (unscientifically
disposed = 6221.5)
16. Manipur 282.3 190.3 108.6 81.7
17. Meghalaya 107.01 93.02 9.64 83.4
18. Mizoram 345.47 275.92 269.71 0
19. Nagaland 330.49 285.49 122 7.5
20. Odisha 2132.95 2097.14 1038.31 1034.33
21. Punjab 4338.37 4278.86 1894.04 2384.82
22. Rajasthan 6897.16 6720.476 1210.46 5082.16
23. Sikkim 71.9 71.9 20.35 51.55
24 Tamil Nadu 13,422 12,844 9430.35 2301.04
25 Telangana 9965 9965 7530 991
26 Tripura 333.9 317.69 214.06 12.9
27 Uttarakhand 1458.46 1378.99 779.85 –
28 Uttar Pradesh 14,710 14,292 5520 0
29 West Bengal 13,709 13,356 667.6 202.23
30 Andaman and 89 82 75 7
Nicobar Islands
31 Chandigarh 513 513 69 444
(continued)
11 Pyrolytic Conversion of Heterogenic Natural Waste Biomass from Rural … 269

Table 11.1 (continued)


No. State TPD of wastes
Solid waste Collected Treated Landfilled
generated
32 DDDNH 267 267 237 14.5
33 Delhi 10,990 10,990 5193.57 5533
34 Lakshadweep 35 17.13 17.13 Nil
35 Puducherry 504.5 482 36 446
Total 160,038.9 152,749.5 79,956.3 29,427.2
Source Annual Report on Solid Waste Management (2020–21), CPCB, Delhi

Table 11.2 Trends in SWM


Year SW generation per capita (g/d) Gap in SWM (%)
of last six years (2015–21) in
India 2015–16 118.68 40.99
2016–17 132.78 79.78
2017–18 98.79 9.38
2018–19 121.54 30.35
2019–20 119.26 25.82
2020–21 119.07 31.70

on the co-pyrolysis of waste biomass and biomass. The escalating environmental


concerns at a worldwide level, such as the widespread occurrence of plastic pollu-
tion and the increase in garbage generation, have inspired a quest for innovative and
environmentally friendly strategies for waste management.
The co-pyrolysis of biomass and plastic waste has evolved as an innovative tech-
nology that presents a potential solution to tackle these pressing concerns concur-
rently [60]. The thermochemical process described herein entails the concurrent or

Table 11.3 Co-pyrolysis of natural biomass and plastic waste


Feedstock Heating rate (°C/min) Liquid yield (wt%) References
PP and saw dust 30 24–48 Ye et al. [75]
Palm shell and PS 61.6 Abnisa et al. [2]
Rice husk and PP – 74.0 Costa et al. [21]
Red oak and HDPE 57.6 Xue et al. [73]
Lignin and PP – 47.4 Duan et al. [24]
Waste cooking oil and 50 62 Mahari et al. [41]
waste plastic
Grape seeds and PS 100 56 Sanahuja-Parejo et al.
[57]
Rice straw and PP 35.7 Suriapparao et al. [66]
270 M. Anil Kumar et al.

sequential pyrolysis of biomass and plastic waste, yielding valuable commodities


such as biochar, bio-oil, and syngas. The co-pyrolysis of biomass and plastic waste
presents several advantages compared to separate pyrolysis methods [18, 49]. This
approach not only redirects plastic waste away from landfills and incineration facili-
ties but also utilises biomass waste’s energy and high carbon content. Implementing
this dual waste valorisation strategy effectively contributes to establishing a circular
economy by converting waste materials into valuable resources [19]. This technique
not only mitigates environmental pollution but also enhances resource efficiency.
When two components combine to modify hydrogen and oxygen levels, synergistic
effects result. Plastic wastes’ high hydrogen/carbon and low oxygen/carbon rates
function synergistically with lignocellulosic biomass’s high and low.
One potential strategy for addressing the growing issue of plastic waste manage-
ment involves the utilization of co-pyrolysis with biomass to convert plastic trash
into bio-oil and biochar. Pyrolysis is a thermochemical phenomenon characterized
by the thermal decomposition of substances at elevated temperatures, occurring in
an oxygen-deprived environment, resulting in the generation of valuable end prod-
ucts. The aforementioned products encompass gas, liquid bio-oil, and solid charcoal
[25]. The pyrolysis process of waste plastic is influenced by several process factors,
including high temperature and high pressure, leading to different oil yields [4, 6].
The process of plastic pyrolysis yields an oil with favourable attributes, such as a
significant heating value and a substantial proportion of hydrocarbons, which are akin
to those found in normal gasoline. In contrast, the process of plastic pyrolysis yields
significant amounts of ash, and the resultant oil exhibits elevated levels of oxygen and
moisture, as well as increased corrosiveness and viscosity [4, 6, 64]. Nevertheless,
the adverse characteristics mentioned above can be mitigated with the incorporation
of organic matter obtained from the biomass source, as demonstrated by Esso et al.
[26] in their study on co-pyrolysis. The findings of recent studies conducted by Esso
et al. [26] and Wang et al. [72] indicate that the co-pyrolysis of biomass with plastic
has been shown to significantly improve the quality and yield of biofuels, without
any observed adverse consequences, in comparison with the pyrolysis of biomass
alone.
Co-pyrolysis is widely considered to be a straightforward and safe process for the
generation of biofuels of superior quality, eliminating the necessity for bio-oil treat-
ment through high-pressure hydrogenation [15, 3, 73, 80]. Yaman et al. [74] suggest
that an additional advantage of co-pyrolysis lies in its capacity to facilitate the retrieval
of significant molecules, hence concurrently diminishing dependence on fossil fuels.
The customization of co-pyrolysis products according to specific application require-
ments has been explored in several studies [19, 29, 42, 47, 54]. Several examples of
uses include gas turbines, diesel engines, and generators. In contrast to the fractional
distillation process employed for fossil fuels, co-pyrolysis technologies yield reduced
amounts of carbon emissions. The reason for this is that co-pyrolysis does not neces-
sitate the utilization of solvents, has the ability to operate with or without catalysts,
and reduces the quantity of hydrogen needed for the hydrotreatment of crude bio-oil
[17, 77, 78]. The co-pyrolysis technique yields biodiesel with equivalent characteris-
tics to petroleum diesel. Moreover, the biodiesel’s low viscosity enhances its potential
11 Pyrolytic Conversion of Heterogenic Natural Waste Biomass from Rural … 271

for many applications and augments its combustion efficiency. This finding provides
evidence that the utilization of biofuels generated via co-pyrolysis has promise in
serving as feasible alternatives to conventional fossil fuels.
Nonetheless, certain obstacles arise in the context of co-pyrolysis and these
issues encompass optimizing process conditions, ensuring compatibility among
various feedstocks as well as establishing suitable technology and policy frameworks.
Continued research and development endeavours are currently being undertaken to
address these obstacles and harness the complete capabilities of co-pyrolysis as an
environmentally sound waste management and energy generation technology. The
simultaneous thermal decomposition of biomass and plastic presents a potentially
advantageous method for transforming a wide range of waste materials into valuable
commodities, thereby tackling the complexities associated with waste management
and promoting the principles of sustainability and circularity within the economy.
Through ongoing research and substantial investment in this technology, co-pyrolysis
can assume a substantial role in valorizing waste materials and generating renewable
energy.

11.2.5 Influential Parameters in Co-pyrolysis

The term “co-pyrolysis” refers to the concurrent thermal decomposition of various


materials, resulting in the production of a wide range of products. The process of
co-pyrolysis has a significant influence on the quantity, content, and properties of
the resulting products. There exist several notable aspects in co-pyrolysis that have
the potential to influence the overall process and its outcomes [14]. Pyrolysis is a
thermal decomposition process that occurs in the absence of oxygen, resulting in the
production of various by-products such as biochar, bio-oil, and syngas. The utilization
of co-pyrolysis has generated significant attention owing to its numerous advantages
in comparison with pyrolysis employing a singular feedstock. The aforementioned
benefits entail enhanced production output and quality, heightened energy efficiency,
and the utilization of waste resources for added value. The identification and analysis
of relevant parameters in co-pyrolysis provide valuable insights for the optimization
of thermochemical conversion processes [38].
The composition of the feedstock mixture has an impact on the distribution of
the end product. The attainment of an optimal feedstock ratio holds promise for
generating synergistic effects, hence enhancing the overall efficiency of the process.
The aforementioned contributions facilitate a deeper understanding of the complex
relationship between feedstocks, reaction conditions, and resulting products, hence
promoting the development of co-pyrolysis technologies that are more efficient and
sustainable. However, it is crucial to engage in ongoing research and experimen-
tation in order to improve our comprehension of the key elements associated with
co-pyrolysis and make progress in this field [79]. The optimization of crucial param-
eters in co-pyrolysis has the potential to promote the development of effective and
economically viable approaches for producing value biofuels, boosting soil quality
272 M. Anil Kumar et al.

through biochar, and deriving other valuable products from renewable resources. The
temperature at which a reaction is conducted has a critical role in determining the
rate at which the reaction proceeds as well as the amount of desired products that
are formed. Higher temperatures have the ability to promote secondary reactions
and cause alterations in the distribution of products, while lower temperatures may
potentially result in enhanced char formation. The choice of reactor type, such as
fixed-bed, fluidized-bed, or rotary kiln, can have a significant impact on the heat and
mass transport characteristics, ultimately affecting the yields and compositions of
the final products. The reaction pathways and product yields can be influenced by
the pressure conditions experienced during co-pyrolysis. Various external forces can
lead to the production of a variety of gaseous and liquid substances.
The duration of the feedstock’s exposure to the pyrolysis temperature directly
influences the magnitude of the pyrolysis reactions. Increased residence periods
can result in more significant reaction extents and alterations in the distribution
of products. The energy consumption during pyrolysis and the resulting product
distribution can be significantly influenced by the moisture content present in the
feedstock, particularly in the case of feedstocks derived from biomass [14]. In order
to achieve sustainability and promote eco-friendliness, co-pyrolysis processes must
consider environmental factors such as greenhouse gas emissions, waste minimisa-
tion, and energy utilization. Pre-treatment techniques, such as drying or grinding,
can potentially alter the feedstock’s characteristics, hence impacting the behaviour
of co-pyrolysis. The utilisation of catalysts or additives during co-pyrolysis can alter
the pathways of the reaction and the selectivity of the resulting products, thereby
enhancing both the quality and quantity of the products obtained. The rate of temper-
ature increase during co-pyrolysis affects the kinetics and outcomes of various reac-
tions and the composition of the finished product. The concurrent thermal decom-
position of biomass and plastic has the potential to decrease the ash content of the
resulting biochar. The inhibitory effect of plastic on ash formation during pyrolysis
produces a carbonaceous substance with higher purity.

11.3 Choice of Feedstocks

The careful choice of feedstocks is a crucial aspect of the co-pyrolysis process as it


has a direct influence on the product distribution and its characteristics. The selection
of feedstocks should take into consideration many factors such as the availability and
cost of the feedstocks, the desired end products, and the overall sustainability of the
process [39]. Proximate analysis is a method employed to ascertain the quantities of
moisture, volatile matter, fixed carbon, and ash present in the feedstock. The ultimate
analysis provides data on the elemental composition of the feedstock, encompassing
carbon, hydrogen, oxygen, nitrogen, and sulphur. The aforementioned evaluations
are crucial for acquiring a comprehensive comprehension of the energy content of
the feedstock and its subsequent behaviour throughout the pyrolysis process. When
11 Pyrolytic Conversion of Heterogenic Natural Waste Biomass from Rural … 273

making choices for feedstocks for co-pyrolysis, it is crucial to conduct a compre-


hensive examination of the feedstock’s chemical composition, accessibility, cost-
effectiveness, suitability for sustainable utilization, and compatibility with the desired
final products. The customization of co-pyrolysis processes has been shown to effec-
tively produce valuable products while simultaneously solving environmental and
economic challenges [49]. This objective is achieved by a meticulous process of
feedstock selection and the optimization of product ratios.

11.3.1 Biomass Feedstock

Biomass feedstocks encompass a wide array of organic resources, which encompass,


but are not restricted to, wood, agricultural residues (e.g. straw and husks), energy
crops (e.g. switchgrass), and organic waste (e.g. food waste and yard trimmings)
[12]. Biomass possesses the inherent advantage of being a renewable resource and
exhibits the potential to attain carbon neutrality, rendering it a favourable choice
for co-pyrolysis from an environmental perspective. The potential of biomass feed-
stock in influencing the global transition towards sustainable and renewable energy
systems holds significant promise. The importance of biomass feedstock is antic-
ipated to persist as a vital and sustainable resource for bioenergy, encompassing
biofuels such as bioethanol and biodiesel, as well as biogas. The advancement of
technology and research is anticipated to bolster the efficiency and viability of
biomass-derived energy production, hence augmenting its competitiveness relative
to conventional fossil fuels. The incorporation of biomass feedstock holds consid-
erable significance within the context of the circular economy paradigm, wherein
organic waste materials undergo a conversion process to yield valuable commodi-
ties including biochar, bioplastics, and biochemicals. The implementation of this
integration will enable the enhancement of resource utilization and the mitigation
of waste generation. According to Lopes et al. [39], the implementation of sustain-
able strategies in the management of biomass feedstock has promise for making a
substantial contribution to the process of carbon sequestration. The reason for this
phenomenon is that the increase in biomass promotes the uptake of carbon dioxide
from the atmosphere.
Integrating bioenergy production with carbon capture and storage (BECCS) has
the potential to result in a net removal of carbon dioxide from the atmosphere,
contributing to the mitigation of climate change. This initiative provides decentralized
energy solutions explicitly focusing on rural and isolated regions. The use of locally
accessible biomass resources has the potential to enhance energy accessibility and
decrease dependence on centralized power networks. The primary objective of the
research and development activities will be to investigate and analyse different feed-
stock sources for biomass conversion. This encompasses many sources of biomass,
such as non-food biomass, algae, agricultural wastes, and waste materials, mitigating
the risk of rivalry with food production. Biomass feedstock is paramount in facil-
itating the shift towards a sustainable energy paradigm. Using this resource as a
274 M. Anil Kumar et al.

renewable and carbon–neutral source shows potential in mitigating greenhouse gas


emissions, enhancing energy security, and facilitating rural development. In order
to fully harness the potential of biomass feedstock, it is imperative to effectively
tackle the difficulties associated with land usage, feedstock supply, and environ-
mental impact. The implementation of sustainable practices, such as the adoption
of responsible land management techniques, the practice of crop rotation, and the
conversion of waste into energy, will play a crucial role in guaranteeing the enduring
sustainability of biomass feedstock.

11.3.2 Plastic Feedstock

Plastic waste, including polyethylene, polypropylene, and polystyrene, is commonly


utilized as a feedstock for co-pyrolysis. Plastic garbage not only plays a significant
role in promoting efficient waste management strategies but also presents the poten-
tial for the production of valuable hydrocarbon-rich goods. There exists a necessity
to develop scientifically engineered landfills in order to facilitate appropriate waste
management, as the current practice involves the disposal of garbage in dumpsites.
Based on the data provided by State Pollution Control Boards (SPCBs) and Pollu-
tion Control Committees (PCCs), it has been determined that the country possesses
a total of 3184 landfill sites. Out of the total, a total of 234 dumpsites have undergone
reclamation attempts, and the remaining 8 have been converted into landfills. Uttar
Pradesh exhibits the highest prevalence of dumpsites, with a cumulative count of 609,
while Madhya Pradesh and Maharashtra trail behind with 326 and 237 dumpsites,
respectively. In relation to the reclamation of dumpsites, the state of Maharashtra
has witnessed the most number of sites being reclaimed, with a total of 141. This is
followed by Madhya Pradesh with 50 sites, Tamil Nadu with 23 sites, and Telangana
with 6 sites, in descending order. In the state of Andhra Pradesh, three dumpsites
have undergone a transformation into landfills. Furthermore, it should be noted that
a landfill has been established in each of the states of Meghalaya, Rajasthan, Sikkim,
Telangana, and Chandigarh. Moreover, Fig. 11.1 illustrates the overall solid waste
deposition and market share of the waste problem in India.
The inclusion of plastic waste in the pyrolysis process has been observed to
positively impact the production of biochar by facilitating the formation of addi-
tional char. The outcome of this process yields a greater overall biochar production
compared to the exclusive utilization of biomass pyrolysis. The presence of a signifi-
cant amount of hydrocarbons in plastic waste has the potential to augment the carbon
content of biochar, hence enhancing its stability and ability to sequester carbon over
an extended period of time. The future prospects for plastic feedstock involve the
analysis of environmental challenges linked to plastic waste, as well as the investiga-
tion of innovative approaches for converting plastics into valuable commodities. The
examination of recycling technologies, including chemical recycling and pyrolysis,
as means to convert plastic waste into primary resources for the production of new
plastics, fuels, and other important commodities, has garnered significant attention
11 Pyrolytic Conversion of Heterogenic Natural Waste Biomass from Rural … 275

Fig. 11.1 Economic prospectives of the total waste and market share

and generated a sense of optimism. According to Schipfer et al. [60], these technolo-
gies have the ability to alleviate the environmental impacts associated with plastic
waste and promote the adoption of a circular economy framework. The progress in
the development of bio-based and biodegradable plastics holds promise for reducing
reliance on polymers derived from fossil fuels and addressing their environmental
impacts. The anticipated rise in acceptance and popularity is attributed to the use
of sustainable plastic manufacturing methods that employ renewable feedstocks,
namely those from plant-based sources. The utilization of plastic feedstock in waste-
to-energy conversion methods, such as co-pyrolysis with biomass, has the potential
to augment the valorization of trash and facilitate the retrieval of energy resources.
The aforementioned technology possesses the capacity to aid the amelioration of the
accumulation of plastic trash in landfills and promote the progression of renewable
energy generation. The transmission of knowledge regarding the ecological conse-
quences linked to plastic waste and the unexplored possibilities of plastic feedstock
in waste management and energy production should be given utmost priority [76].
The dissemination of knowledge to the general population about the responsible
utilisation of plastic, recycling practices, and the reduction of waste would facilitate
modifications in behaviour and contribute to the advancement of sustainable solu-
tions. The pressing environmental issues associated with plastic waste need creative
276 M. Anil Kumar et al.

strategies to transform plastic trash into valuable resources. By adopting sophisti-


cated recycling technologies, using sustainable plastic manufacturing techniques,
and implementing waste-to-energy conversion processes, the plastic feedstock may
be effectively used as a valuable resource within the framework of circular economy
practises [19]. These endeavours mitigate plastic pollution, enhance resource effi-
ciency, and foster the development of renewable energy. In order to fully harness the
potential of plastic feedstock, it is essential to foster collaborative endeavours among
governmental bodies, industrial sectors, research institutes, and consumers. Mini-
mizing plastic waste and promoting environmental sustainability may be achieved
by implementing policies, increasing investment in research and development, and
widespread adoption of sustainable practices [76]. These factors will be crucial in
designing a future where plastic feedstock supports environmental sustainability.
By reconceptualizing the function of plastic garbage as an asset, we may foster a
future characterized by increased durability and conscientiousness, benefiting future
generations.

11.3.3 Biomass-Plastic Feedstock Blending

The utilization of biomass and plastic waste as co-feedstocks in co-pyrolysis repre-


sents a novel approach, wherein these distinct materials are combined and subjected
to thermal degradation in the absence of oxygen. The concurrent thermal decomposi-
tion of these particular sources of biomass offers a promising prospect for the utiliza-
tion of discarded resources and the generation of environmentally friendly power.
The composition of feedstocks plays a significant role in determining the outcomes
of the pyrolysis process and the resulting yields of products. Biomass feedstocks
encompass a plentiful supply of cellulose, hemicellulose, and lignin, while plastic
waste predominantly consists of extended hydrocarbon chains. Different types of
feedstocks can lead to a diverse array of products. In recent years, there has been a
significant increase in the adoption of biomass conversion technologies, including
the combustion of rice husk and sugarcane bagasse, as well as the gasification of
diverse agricultural leftovers. Nevertheless, a considerable proportion of this invalu-
able resource is still being wasted as a result of its utter indifference or unregulated
combustion in agricultural fields. The aforementioned issue is notably widespread in
poor nations, when the lack of a suitable institutional structure impedes endeavours
aimed at mitigating pollution.
The typical practice of immediately incinerating agricultural leftovers is a
commonly utilized technique, it results in the emission of air pollutants, which can
have adverse impacts on both human health and the environment. The inadequate
utilization of biomass, a resource that is capable of renewal, presents a number of
issues. The issue under consideration relates to the efficient utilization of biomass
as a feasible energy source and other necessary commodities. Biomass feedstocks,
such as agricultural residues, energy crops, and organic waste, are widely recognized
as renewable and abundant sources of carbon-based materials. On the other hand,
11 Pyrolytic Conversion of Heterogenic Natural Waste Biomass from Rural … 277

plastic waste, which poses a notable ecological concern, is comprised of elongated


hydrocarbon chains derived from polymers sourced from petroleum. The combi-
nation of biomass and plastic feedstocks produces synergistic effects, resulting in
enhanced process efficiency and unique product outcomes.
The interplay between biomass and plastic during the process of pyrolysis can
impact the rate at which reactions occur and the distribution of products, ultimately
leading to enhanced yields and modified characteristics of the resultant products
compared to pyrolysis conducted using each feedstock separately. Combining these
feedstocks and putting them into pyrolysis, various valuable products may be derived,
including bio-oil, biochar, and syngas. The combination of biomass and plastic
feedstocks for co-pyrolysis has substantial promise in waste valorisation, enhanced
product yields, and the generation of bio-oil and biochar of superior quality. Gin
et al. [28] examined the impact of the initial plastic-biomass mixture to catalyst ratio,
while others examined catalytic pyrolysis of plastic alone. To lower operating costs,
pyrolysis of high-density polyethylene utilizing zeolite as the catalyst can be done
with low catalyst content. The study found that reducing catalyst content reduces
system activity, but raising process temperature can compensate. Research indicates
that the catalyst amount used in co-pyrolysis of plastic and biomass impacts product
yield and distribution by causing secondary cracking of pyrolyzates, resulting in the
production of polyaromatic hydrocarbons and hydrogenated bio-oil fractions used
in conventional fuels.
The utilization of biomass presents numerous benefits. Biomass, particularly the
biomass resources obtained from agricultural activities, is a viable and plentiful kind
of renewable energy. The utilization of this chemical does not result in more carbon
dioxide being released into the Earth’s atmosphere, rendering it a promising substi-
tute for fossil fuels and a valuable instrument in the mitigation of greenhouse gas
emissions. The utilization of biomass has promise in enhancing farmers’ economic
resources while ensuring the preservation of their ability to cultivate vital food and
non-food crops. This issue is currently under serious deliberation over prioritiza-
tion. However, additional research is necessary to improve the selection of feedstock
combinations, develop appropriate catalysts and additives, and address the economic
and environmental considerations related to the extensive adoption of this method-
ology. The combination of biomass and plastic feedstock exhibits significant poten-
tial in the realm of sustainable waste management and the production of renewable
energy. However, the combination of biomass and plastic feedstocks in co-pyrolysis
presents certain challenges that need to be resolved. The aforementioned concerns
encompass determining appropriate ratios for the feedstock, addressing potential
by-products, such as halogens derived from plastics, and assessing the economic
feasibility and environmental implications of this methodology.
278 M. Anil Kumar et al.

11.4 Synergistic Effects of Co-pyrolysis

The phenomenon of synergistic effects in co-pyrolysis refers to the enhanced perfor-


mance and benefits that arise from the simultaneous pyrolysis of multiple diverse
feedstocks. The co-pyrolysis process entails the thermal decomposition of feed-
stocks, which can lead to improved reactions, increased product yields, and enhanced
properties of the resulting products when compared to the pyrolysis of individual
feedstocks. The phenomenon of synergy in co-pyrolysis arises from the intrinsic
complementarity observed among various feedstocks, which possess unique chem-
ical compositions and features. The simultaneous pyrolysis of different feedstocks
can have a reciprocal influence on their individual characteristics, leading to a thermal
conversion procedure characterized by improved efficiency and efficacy. Overall,
the synergistic impacts of co-pyrolysis offer favourable potential for augmenting the
efficacy of transforming diverse feedstocks into valuable commodities. However,
the realization of these benefits relies on a thorough comprehension of feedstock
interactions and the careful selection of feedstock combinations.
The progress made in co-pyrolysis research holds great significance in the
advancement of sustainable and efficient pyrolysis methods for a range of appli-
cations, such as bioenergy production, waste disposal, and soil improvement. The
incorporation of synergistic effects will play a pivotal role in attaining these aims.
The utilization of biomass and plastic feedstocks in a mutually beneficial manner
holds promise for the production of bio-oils with enhanced energy density. The
observed rise in calorific value can be attributed to the integration of hydrocarbons
derived from the plastic feedstocks into the bio-oil. The integration of biomass and
plastic feedstock blending holds considerable promise in addressing the growing
complexities associated with plastic waste management and the demand for envi-
ronmentally viable energy options. The utilization of biomass and plastic feedstocks
in co-pyrolysis represents an innovative and environmentally aware approach to
converting waste materials into valuable resources. The co-pyrolysis process, namely
the combination of biomass and plastic feedstocks, has demonstrated the manifes-
tation of synergistic effects. The aforementioned impacts have the potential to yield
increased bio-oil output, improved characteristics, and tailored biochar properties.
As previously said, the advantages demonstrate the ability of this approach to make
a substantial contribution to the production of sustainable and efficient biofuels, the
valorization of waste, and the promotion of environmental stewardship.

11.4.1 Synergistic Mechanism

The phenomenon in which the combined effect of multiple components or factors


exceeds the sum of their individual effects is commonly known as the synergistic
mechanism. The concept of synergistic mechanism holds significant importance and
11 Pyrolytic Conversion of Heterogenic Natural Waste Biomass from Rural … 279

can be observed across several scientific and technical contexts. The statement under-
scores the significance of collaboration and the mutually beneficial improvement of
separate elements. The concept of synergy is widely acknowledged across diverse
disciplines, encompassing chemistry, biology, ecology, and the social sciences. In
the fields of chemistry and biology, the interplay of various chemicals or molecules
can give rise to effects that exhibit greater potency or novelty compared to the effects
produced by individual substances in isolation.
In the field of ecology, the coexistence of different species can give rise to a
harmonious ecosystem in which each species plays a role in supporting the health and
functioning of others, so enhancing the overall stability and resilience of the system.
Gaining comprehension of synergistic mechanisms inside industrial applications
and devising strategies to harness their potential are pivotal measures in enhancing
process efficiency, facilitating product development, and refining system architecture.
In the context of co-pyrolysis, as discussed earlier, the amalgamation of biomass
and plastic feedstocks has the potential to generate synergistic outcomes, hence
enhancing the overall efficiency and product yields of the pyrolysis procedure as
shown in Fig. 11.2.
The synergistic process encompasses not only positive interactions but also nega-
tive interactions, when the combined effect is more detrimental than the total of
the separate effects. The utilization of the term “antagonistic mechanism” may be
deemed suitable for application in such circumstances. The concept of synergistic
processes underscores the importance of holistic consideration of a system, rather
than focusing solely on its individual components. In order to enhance their ability to
develop more efficient and effective solutions, scientists and engineers might strate-
gically leverage synergistic interactions. This phenomenon has the potential to drive
advancements across multiple industries and facilitate the progress of sustainable
and innovative technologies.

Fig. 11.2 Waste utilization using synergistic thermochemical co-processing


280 M. Anil Kumar et al.

11.4.2 Impact on Quality and Yield of Biofuels

The potential influence of synergistic effects during co-pyrolysis, particularly in the


blending of biomass and plastic feedstocks, has the capacity to significantly impact
the quality and yield of the resulting biofuels. The inclusion of plastic feedstocks
in the co-pyrolysis process has the potential to enhance the yield of bio-oil. During
the pyrolysis process, plastics, which contain a significant amount of hydrocarbons,
introduce additional volatile compounds into the mixture. Consequently, the overall
bio-oil production is greater in comparison with biomass pyrolysis. The presence
of biomass and plastic feedstocks may result in a reduction of oxygen-containing
molecules in the bio-oil compared to previous conditions.
As a result, the process yields bio-oil with higher energy content, increased
stability, reduced acidity, and improved properties, rendering it highly suitable for
utilization as a renewable fuel source. The inclusion of plastic in the feedstock mix
during the co-pyrolysis process has the potential to mitigate the formation of exces-
sive char. As a consequence of this phenomenon, there is a potential decrease in
the quantity of solid char generated, leading to an increased production of valuable
liquid goods, notably bio-oil. By adjusting the feedstock ratio and modifying the
processing conditions, it is possible to tailor the composition of the resulting bio-oil
to meet specific application requirements. This allows for the production of biofuels
that possess desired characteristics suitable for transportation or industrial purposes.
This characteristic enables the utilization of bio-oil in a diverse range of applica-
tions. The co-pyrolysis process entails the interaction between biomass and plastic,
which has the potential to alter the properties of the resulting biochar. This might
potentially lead to the production of biochar characterized by an increased carbon
content, reduced ash content, and improved surface area. Consequently, it would be
more suitable for enhancing soil quality and facilitating carbon sequestration. It is
imperative to emphasize that the influence of synergistic effects on the quality and
quantity of biofuels generated through co-pyrolysis is governed by a multitude of
criteria. The elements encompassed in this study comprise the composition of the
feedstock, the ratio of feedstock, the temperature at which pyrolysis occurs, and
the potential influence of catalysts or additives. Therefore, it is imperative to opti-
mize these parameters in order to fully exploit the potential of the biomass-plastic
feedstock blending process and achieve the desired outcomes in biofuel production.

11.4.3 Impact on Quality and Yield of Biochar

The potential impact of synergistic effects during co-pyrolysis, namely the combi-
nation of biomass and plastic feedstocks, can also have implications for the charac-
teristics and quantity of the resulting biochar. Biochar is generated by the process
of biomass pyrolysis, resulting in the formation of a solid carbonaceous material.
Biochar has the potential to serve as a soil amendment, a carbon sequestration agent,
11 Pyrolytic Conversion of Heterogenic Natural Waste Biomass from Rural … 281

or a precursor for several other applications. Enhancing the surface area of biochar
has the potential to enhance its ability to absorb and retain water and nutrients, hence
augmenting its efficacy as a constituent of soil amendments. The combination of
biomass and plastic feedstock can result in modified attributes of biochar, perhaps
enhancing its capacity for nutrient retention. Biochar is found to be a more effec-
tive soil conditioner, hence promoting plant growth and enhancing soil fertility. The
composition of biochar can be influenced by the ratio of feedstocks employed in the
co-pyrolysis process as well as the processing parameters. The interplay between the
two components, which adjusts the amount of hydrogen and oxygen in the system,
is what creates the synergistic effects and makes them possible. For instance, the
high hydrogen/carbon (H/C) ratios of plastic wastes and the low oxygen/carbon (O/
C) ratios of lignocellulosic biomass operate synergistically together. Similarly, the
high oxygen/carbon (O/C) ratios of lignocellulosic biomass and the low hydrogen/
carbon (H/C) ratios of lignocellulosic biomass also work well together.
The characteristics of biochar can be adjusted to meet the specific needs of a
given soil type or agricultural pursuit through the manipulation of these variables. It
is imperative to bear in mind that the impact of synergistic effects on the quality and
productivity of biochar is contingent upon various factors. The factors encompassed
in this category consist of the various types of biomass and plastic feedstocks utilized,
the ratio of feedstocks employed, the temperature at which the pyrolysis process is
conducted, and the potential inclusion of catalysts or additives. It is imperative to
optimize these parameters in a suitable manner in order to fully exploit the potential
benefits that may be derived from the combination of biomass and plastic feedstock
in the process of biochar synthesis. The aforementioned advantages exemplify the
potential of this approach in producing biochar that may be tailored to improve soil
health and trap carbon, thereby promoting environmentally sustainable agriculture
and conserving natural resources.

11.5 Conclusions and Future Perspectives

Pyrolytic conversion and valorization of rural waste biomass provides a sustain-


able, decentralized waste-to-energy option. Pyrolysis technology in rural areas can
improve resource efficiency and ecological sensitivity. Management of diverse waste
biomass in rural areas is difficult. Pyrolysis converts organic waste into valuable prod-
ucts, reducing landfill trash and pollution. The thermal breakdown of waste biomass
in pyrolysis can provide sustainable energy to rural areas. This could boost energy
security, rural development, and economic prosperity. Pyrolysis biochar can improve
soil fertility and water retention. The benefits include increased agricultural yield and
less chemical fertilizer use. Pyrolysis turns waste biomass into biochar, bio-oil, and
syngas. Circular economy strives to improve resource efficiency and reduce fossil
fuel use. Rural pyrolysis technology can help spread knowledge and skills. Innova-
tive waste management solutions could empower local populations and boost rural
development.
282 M. Anil Kumar et al.

Pyrolytic conversion and valorization of heterogeneous rural waste biomass


provides a gateway to a cleaner, self-sufficient future. Pyrolyzing waste biomass
can provide renewable energy, soil health, and environmental protection to rural
areas. Technical accessibility, infrastructure development, and community participa-
tion are unique challenges to implementation. Government, research, and business
must collaborate for innovation, customization, and sustainability. Pyrolysis tech-
nology can help rural communities promote environmental stewardship, socioeco-
nomic development, and resilient, self-sustaining communities that contribute to the
low carbon, circular economy, and global sustainability. The review covers feedstock
selection, pyrolysis process optimization, product characterization, environmental
impacts, and socioeconomic impacts. This literature analysis shows how pyrolytic
conversion can turn agricultural waste into a valuable resource while resolving envi-
ronmental concerns and fostering a sustainable future. Pyrolytic conversion and
valorization of waste biomass could improve rural communities’ environmental
sustainability and inclusion.

References

1. Abnisa, F., & Daud, W. M. A. W. (2014). A review on co-pyrolysis of biomass: An optional


technique to obtain a high-grade pyrolysis oil. Energy Conversion and Management, 87, 71–85.
https://doi.org/10.1016/j.enconman.2014.07.007
2. Abnisa, F., Daud, W. W., Ramalingam, S., Azemi, M. N. B. M., & Sahu, J. N. (2013). Co-
pyrolysis of palm shell and polystyrene waste mixtures to synthesis liquid fuel. Fuel, 108,
311–318. https://doi.org/10.1016/j.fuel.2013.02.013
3. Alabdrabalnabi, A., Gautam, R., & Sarathy, S. M. (2022). Machine learning to predict biochar
and bio-oil yields from co-pyrolysis of biomass and plastics. Fuel, 328, 125303. https://doi.
org/10.1016/j.fuel.2022.125303
4. Alam, S. S., Husain Khan, A., & Khan, N. A. (2022). Plastic waste management via ther-
mochemical conversion of plastics into fuel: A review. Energy Sources, Part A: Recovery,
Utilization, and Environmental Effects, 44(3), 1–20. https://doi.org/10.1080/15567036.2022.
2097750
5. Alazaiza, M. Y., Albahnasawi, A., Ali, G. A., Bashir, M. J., Nassani, D. E., Al Maskari, T., &
Abujazar, M. S. S. (2022). Application of natural coagulants for pharmaceutical removal from
water and wastewater: A review. Water, 14, 140. https://doi.org/10.3390/w14020140
6. Al-Rumaihi, A., Shahbaz, M., Mckay, G., Mackey, H., & Al-Ansari, T. (2022). A review of
pyrolysis technologies and feedstock: A blending approach for plastic and biomass towards
optimum biochar yield. Renewable and Sustainable Energy Reviews, 167, 112715. https://doi.
org/10.1016/j.rser.2022.112715
7. Anand, A., Kumar, V., & Kaushal, P. (2022). Biochar and its twin benefits: Crop residue manage-
ment and climate change mitigation in India. Renewable and Sustainable Energy Reviews, 156,
111959. https://doi.org/10.1016/j.rser.2021.111959
8. Anand, U., Reddy, B., Singh, V. K., Singh, A. K., Kesari, K. K., Tripathi, P., Kumar, P.,
Tripathi, V., & Simal-Gandara, J. (2021). Potential environmental and human health risks
caused by antibiotic-resistant bacteria (ARB), antibiotic resistance genes (ARGs) and emerging
contaminants (ECs) from municipal solid waste (MSW) landfill. Antibiotics, 10(4), 374. https://
doi.org/10.3390/antibiotics10040374
9. Armenise, S., SyieLuing, W., Ramírez-Velásquez, J. M., Launay, F., Wuebben, D., Ngadi, N., &
Muñoz, M. (2021). Plastic waste recycling via pyrolysis: A bibliometric survey and literature
11 Pyrolytic Conversion of Heterogenic Natural Waste Biomass from Rural … 283

review. Journal of Analytical and Applied Pyrolysis, 158, 105265. https://doi.org/10.1016/j.


jaap.2021.105265
10. Bandgar, P. S., Jain, S., & Panwar, N. L. (2022). A comprehensive review on optimization of
anaerobic digestion technologies for lignocellulosic biomass available in India. Biomass and
Bioenergy, 161, 106479. https://doi.org/10.1016/j.biombioe.2022.106479
11. Ben-Iwo, J., Manovic, V., & Longhurst, P. (2016). Biomass resources and biofuels potential for
the production of transportation fuels in Nigeria. Renewable and Sustainable Energy Reviews,
63, 172–192. https://doi.org/10.1016/j.rser.2016.05.050
12. Bhatt, M., Chakinala, A. G., Joshi, J. B., Sharma, A., Pant, K. K., Shah, K., & Sharma, A. (2021).
Valorization of solid waste using advanced thermo-chemical process: A review. Journal of
Environmental Chemical Engineering, 9, 105434. https://doi.org/10.1016/j.jece.2021.105434
13. Butler, E., Devlin, G., Meier, D., & McDonnell, K. (2011). A review of recent labora-
tory research and commercial developments in fast pyrolysis and upgrading. Renewable and
Sustainable Energy Reviews, 15, 4171–4186. https://doi.org/10.1016/j.rser.2011.07.035
14. Caputo, C., & Mašek, O. (2021). Spear (solar pyrolysis energy access reactor): Theoretical
design and evaluation of a small-scale low-cost pyrolysis unit for implementation in rural
communities. Energies, 14(8), 2189. https://doi.org/10.3390/en14082189
15. Çepelioğullar, Ö., & Pütün, A. E. (2013). Thermal and kinetic behaviors of biomass and plastic
wastes in co-pyrolysis. Energy Conversion and Management, 75, 263–270. https://doi.org/10.
1016/j.enconman.2013.06.036
16. Chen, H. L., Nath, T. K., Chong, S., Foo, V., Gibbins, C., & Lechner, A. M. (2021). The plastic
waste problem in Malaysia: Management, recycling and disposal of local and global plastic
waste. SN Applied Sciences, 3, 1–15. https://doi.org/10.1007/s42452-021-04234-y
17. Chen, M., Zhang, S., Su, Y., Niu, X., Zhu, S., & Liu, X. (2022). Catalytic co-pyrolysis of food
waste digestate and corn husk with CaO catalyst for upgrading bio-oil. Renewable Energy, 186,
105–114. https://doi.org/10.1016/j.renene.2021.12.139
18. Chen, W., Shi, S., Zhang, J., Chen, M., & Zhou, X. (2016). Co-pyrolysis of waste newspaper
with high-density polyethylene: Synergistic effect and oil characterization. Energy Conversion
and Management, 112, 41–48. https://doi.org/10.1016/j.enconman.2016.01.005
19. Chew, K. W., Chia, S. R., Chia, W. Y., Cheah, W. Y., Munawaroh, H. S. H., & Ong, W. J.
(2021). Abatement of hazardous materials and biomass waste via pyrolysis and co-pyrolysis
for environmental sustainability and circular economy. Environmental Pollution, 278, 116836.
https://doi.org/10.1016/j.envpol.2021.116836
20. Chormare, R., Sahoo, T. P., Chanchpara, A., Saravaia, H. T., & Madhava, A. K. (2023). Thermo-
chemical conversion and kinetic evaluation of Casuarina equisetifolia pines to biochar and
their utilization in sequestering toxic metal ions. Biomass Conversion and Biorefinery, 13,
9423–9434. https://doi.org/10.1007/s13399-022-03526-6
21. Costa, P., Pinto, F., Miranda, M., André, R., & Rodrigues, M. (2014). Study of the experi-
mental conditions of the co-pyrolysis of rice husk and plastic wastes. Chemical Engineering
Transactions, 39, 1639–1644. https://doi.org/10.3303/CET1439274
22. Demirbas, A., Pehlivan, E., & Altun, T. (2006). Potential evolution of Turkish agricultural
residues as bio-gas, bio-char and bio-oil sources. International Journal of Hydrogen Energy,
31(5), 613–620. https://doi.org/10.1016/j.ijhydene.2005.06.003
23. Dhanya, B. S., Mishra, A., Chandel, A. K., & Verma, M. L. (2020). Development of sustainable
approaches for converting the organic waste to bioenergy. Science of the Total Environment,
25, 138109. https://doi.org/10.1016/j.scitotenv.2020.138109
24. Duan, D., Wang, Y., Dai, L., Ruan, R., Zhao, Y., Fan, L., Tayier, M., & Liu, Y. (2017).
Ex-situ catalytic copyrolysis of lignin and polypropylene to upgrade bio-oil quality by
microwave heating. Bioresource Technology, 241, 207–213. https://doi.org/10.1016/j.biortech.
2017.04.104
25. Durak, H. (2023). Comprehensive assessment of thermochemical processes for sustainable
waste management and resource recovery. Processes, 11(7), 2092. https://doi.org/10.3390/pr1
1072092
284 M. Anil Kumar et al.

26. Esso, S. B. E., Xiong, Z., Chaiwat, W., Kamara, M. F., Longfei, X., Xu, J., Ebako, J., Jiang, L.,
Su, S., Hu, S., Wang, Y., & Xiang, J. (2022). Review on synergistic effects during co-pyrolysis
of biomass and plastic waste: Significance of operating conditions and interaction mechanism.
Biomass and Bioenergy, 159, 106415. https://doi.org/10.1016/j.biombioe.2022.106415
27. Foong, S. Y., Liew, R. K., Yang, Y., Cheng, Y. W., Yek, P. N. Y., Mahari, W. A. W., Lee, W.
Y., Han, C. S., Vo, D.-V. N., Le, Q. V., Aghbashlo, M., Tabatabaei, M., Sonne, C., Peng, W., &
Lam, S. S. (2020). Valorization of biomass waste to engineered activated biochar by microwave
pyrolysis: Progress, challenges, and future directions. Chemical Engineering Journal, 389,
124401. https://doi.org/10.1016/j.cej.2020.124401
28. Gin, A. W., Hassan, H., Ahmad, M. A., Hameed, B. H., & Din, A. M. (2021). Recent progress on
catalytic co-pyrolysis of plastic waste and lignocellulosic biomass to liquid fuel: The influence
of technical and reaction kinetic parameters. Arabian Journal of Chemistry, 14(4), 103035.
https://doi.org/10.1016/j.arabjc.2021.103035
29. Guo, M., & Bi, J. C. (2015). Characteristics and application of co-pyrolysis of coal/biomass
blends with solid heat carrier. Fuel Processing Technology, 138, 743–749. https://doi.org/10.
1016/j.fuproc.2015.07.018
30. Hunt, J. M. (2010). Shiga toxin–producing Escherichia coli (STEC). Clinics in Laboratory
Medicine, 30, 21–45. https://doi.org/10.1016/j.cll.2009.11.001
31. Ishangulyyev, R., Kim, S., & Lee, S. H. (2019). Understanding food loss and waste—Why are
we losing and wasting food? Foods, 8(8), 297. https://doi.org/10.3390/foods8080297
32. Kartik, S., Balsora, H. K., Sharma, M., Saptoro, A., Jain, R. K., Joshi, J. B., & Sharma, A.
(2022). Valorization of plastic wastes for production of fuels and value-added chemicals through
pyrolysis—A review. Thermal Science and Engineering Progress, 32, 101316. https://doi.org/
10.1016/j.tsep.2022.101316
33. Kassim, F. O., Thomas, C. P., & Afolabi, O. O. (2022). Integrated conversion technologies
for sustainable agri-food waste valorization: A critical review. Biomass and Bioenergy, 156,
106314. https://doi.org/10.1016/j.biombioe.2021.106314
34. Kibria, M. G., Masuk, N. I., Safayet, R., Nguyen, H. Q., & Mourshed, M. (2023). Plastic waste:
Challenges and opportunities to mitigate pollution and effective management. International
Journal of Environmental Research, 17(1), 20. https://doi.org/10.1007/s41742-023-00507-z
35. Kumar, S., Lohan, S. K., & Parihar, D. S. (2023). Biomass energy from agriculture. In A.
Rakshit, A. Biswas, D. Sarkar, V. S. Meena, & R. Datta (Eds.), Handbook of energy management
in agriculture. Springer. https://doi.org/10.1007/978-981-19-7736-7_10-1
36. Lebreton, L., & Andrady, A. (2019). Future scenarios of global plastic waste generation and
disposal. Palgrave Communications, 5(1), 1–11. https://doi.org/10.1057/s41599-018-0212-7
37. Lee, S. Y., Sankaran, R., Chew, K. W., Tan, C. H., Krishnamoorthy, R., Chu, D. T., & Show, P.
L. (2019). Waste to bioenergy: A review on the recent conversion technologies. BMC Energy,
1(1), 1–22. https://doi.org/10.1186/s42500-019-0004-7
38. Li, X., Li, J., Zhou, G., Feng, Y., Wang, Y., Yu, G., Deng, S., Huang, J., & Wang, B. (2014).
Enhancing the production of renewable petrochemicals by co-feeding of biomass with plastics
in catalytic fast pyrolysis with ZSM-5 zeolites. Applied Catalysis A: General, 481, 173–182.
https://doi.org/10.1016/j.apcata.2014.05.015
39. Lopes, T. F., Carvalheiro, F., Duarte, L. C., Gírio, F., Quintero, J. A., & Aroca, G. (2019).
Techno-economic and life-cycle assessments of small-scale biorefineries for isobutene and
xylo-oligosaccharides production: A comparative study in Portugal and Chile. Biofuels,
Bioproducts and Biorefining, 13(5), 1321–1332. https://doi.org/10.1002/bbb.2036
40. Maalouf, A., & Mavropoulos, A. (2023). Re-assessing global municipal solid waste genera-
tion. Waste Management & Research, 41(4), 936–947. https://doi.org/10.1177/0734242X2210
74116
41. Mahari, W. A. W., Tung, C., & Haur, W. (2018). Microwave copyrolysis of waste polyolefins and
waste cooking oil: Influence of N2 atmosphere versus vacuum environment. Energy Conversion
and Management, 171, 1292–1301. https://doi.org/10.1016/j.enconman.2018.06.073
42. Mahari, W. A. W., Azwar, E., Foong, S. Y., Ahmed, A., Peng, W., Tabatabaei, M., Aghbashlo,
M., Park, Y.-K., Sonne, C., & Lam, S. S. (2021). Valorization of municipal wastes using
11 Pyrolytic Conversion of Heterogenic Natural Waste Biomass from Rural … 285

co-pyrolysis for green energy production, energy security, and environmental sustainability:
A review. Chemical Engineering Journal, 421, 129749. https://doi.org/10.1016/j.cej.2021.
129749
43. Manikandan, S., Subbaiya, R., Biruntha, M., Krishnan, R. Y., Muthusamy, G., & Karmegam,
N. (2022). Recent development patterns, utilization and prospective of biofuel production:
Emerging nanotechnological intervention for environmental sustainability—A review. Fuel,
314, 122757. https://doi.org/10.1016/j.fuel.2021.122757
44. Maquart, P. O., Froehlich, Y., & Boyer, S. (2022). Plastic pollution and infectious diseases.
The Lancet Planetary Health, 6(10), e842–e845. https://doi.org/10.1016/S2542-5196(22)001
98-X
45. Melikoglu, M., Lin, C. S. K., & Webb, C. (2013). Analysing global food waste problem:
Pinpointing the facts and estimating the energy content. Central European Journal of
Engineering, 3, 157–164. https://doi.org/10.2478/s13531-012-0058-5
46. Mishra, R., Ong, H. C., & Lin, C. W. (2023). Progress on co-processing of biomass and plastic
waste for hydrogen production. Energy Conversion and Management, 284, 116983. https://doi.
org/10.1016/j.enconman.2023.116983
47. Mohamed, B. A., O’Boyle, M., & Li, L. Y. (2023). Co-pyrolysis of sewage sludge with ligno-
cellulosic and algal biomass for sustainable liquid and gaseous fuel production: A life cycle
assessment and techno-economic analysis. Applied Energy, 346, 121318. https://doi.org/10.
1016/j.apenergy.2023.121318
48. Mondal, P. (2021). Techno-economic and environmental performance assessment of a MSW to
energy plant for Indian urban sectors. Thermal Science and Engineering Progress, 21, 100777.
https://doi.org/10.1016/j.tsep.2020.100777
49. Neha, S., Ramesh, K. P. K., & Remya, N. (2022). Techno-economic analysis and life cycle
assessment of microwave co-pyrolysis of food waste and low-density polyethylene. Sustainable
Energy Technologies and Assessments, 52, 102356. https://doi.org/10.1016/j.seta.2022.102356
50. Ng, C. H., Mistoh, M. A., Teo, S. H., Galassi, A., Ibrahim, A., Sipaut, C. S., Foo, J., Seay, J.,
Taufiq-Yap, Y. H., & Janaun, J. (2023). Plastic waste and microplastic issues in Southeast Asia.
Frontiers in Environmental Science, 11, 427. https://doi.org/10.3389/fenvs.2023.1142071
51. Okolie, J. A., Epelle, E. I., Tabat, M. E., Orivri, U., Amenaghawon, A. N., Okoye, P. U., &
Gunes, B. (2022). Waste biomass valorization for the production of biofuels and value-added
products: A comprehensive review of thermochemical, biological and integrated processes.
Process Safety and Environmental Protection, 159, 323–344. https://doi.org/10.1016/j.psep.
2021.12.049
52. Omang, D. I., John, G. E., Inah, S. A., & Bisong, J. O. (2021). Public health implication of solid
waste generated by households in Bekwarra Local Government area. African Health Sciences,
21(3), 1467–1473. https://doi.org/10.4314/ahs.v21i3.58
53. Ong, H. C., Chen, W. H., Farooq, A., Gan, Y. Y., Lee, K. T., & Ashokkumar, V. (2019).
Catalytic thermochemical conversion of biomass for biofuel production: A comprehensive
review. Renewable and Sustainable Energy Reviews, 113, 109266. https://doi.org/10.1016/j.
rser.2019.109266
54. Rajkumar, P., & Murugavelh, S. (2022). Co-pyrolysis of wheat husk and residual tyre: Techno-
economic analysis, performance and emission characteristics of pyro oil in a diesel engine.
Bioresource Technology Reports, 19, 101164. https://doi.org/10.1016/j.biteb.2022.101164
55. Ray, R., & Singh, P. (2022). Prevalence and implications of shiga toxin-producing E. coli
in farm and wild ruminants. Pathogens, 11(11), 1332. https://doi.org/10.3390/pathogens111
11332
56. Rezaei, M., Ghobadian, B., Samadi, S. H., & Karimi, S. (2018). Electric power generation
from municipal solid waste: A techno-economical assessment under different scenarios in
Iran. Energy, 152, 46–56. https://doi.org/10.1016/j.energy.2017.10.109
57. Sanahuja-Parejo, O., Veses, A., Navarro, M. V., López, J. M., Murillo, R., Callén, M. S., &
García, T. (2019). Drop-in biofuels from the co-pyrolysis of grape seeds and polystyrene.
Chemical Engineering Journal, 377, 120246. https://doi.org/10.1016/j.cej.2018.10.183
286 M. Anil Kumar et al.

58. Scarlat, N., Dallemand, J. F., Monforti-Ferrario, F., & Nita, V. (2015). The role of biomass and
bioenergy in a future bioeconomy: Policies and facts. Environmental Development, 15, 3–34.
https://doi.org/10.1016/j.envdev.2015.03.006
59. Schanes, K., Dobernig, K., & Gözet, B. (2018). Food waste matters—A systematic review of
household food waste practices and their policy implications. Journal of Cleaner Production,
182, 978–991. https://doi.org/10.1016/j.jclepro.2018.02.030
60. Schipfer, F., Pfeiffer, A., & Hoefnagels, R. (2022). Strategies for the mobilization and deploy-
ment of local low-value, heterogeneous biomass resources for a circular bioeconomy. Energies,
15(2), 433. https://doi.org/10.3390/en15020433
61. Seah, C. C., Tan, C. H., Arifin, N. A., Hafriz, R. S. R. M., Salmiaton, A., Nomanbhay, S., &
Shamsuddin, A. H. (2023). Co-pyrolysis of biomass and plastic: Circularity of wastes and
comprehensive review of synergistic mechanism. Results in Engineering, 100989. https://doi.
org/10.1016/j.rineng.2023.100989
62. Sobamowo, G. M., & Ojolo, S. J. (2018). Techno-economic analysis of biomass energy
utilization through gasification technology for sustainable energy production and economic
development in Nigeria. Journal of Energy. https://doi.org/10.1155/2018/4860252
63. Somorin, T. O., Adesola, S., & Kolawole, A. (2017). State-level assessment of the waste-to-
energy potential (via incineration) of municipal solid wastes in Nigeria. Journal of Cleaner
Production, 164, 804–815. https://doi.org/10.1016/j.jclepro.2017.06.228
64. Soni, V. K., Singh, G., Vijayan, B. K., Chopra, A., Kapur, G. S., & Ramakumar, S. S. V. (2021).
Thermochemical recycling of waste plastics by pyrolysis: A review. Energy & Fuels, 35(16),
12763–12808. https://doi.org/10.1021/acs.energyfuels.1c01292
65. Stephenson, P. J., & Damerell, A. (2022). Bioeconomy and circular economy approaches need
to enhance the focus on biodiversity to achieve sustainability. Sustainability, 14(17), 10643.
https://doi.org/10.3390/su141710643
66. Suriapparao, D. V., Vinu, R., Shukla, A., & Haldar, S. (2020). Effective deoxygenation for the
production of liquid biofuels via microwave assisted co-pyrolysis of agro residues and waste
plastics combined with catalytic upgradation. Bioresource Technology, 302, 122775. https://
doi.org/10.1016/j.biortech.2020.122775
67. Tan, E. C., & Lamers, P. (2021). Circular bioeconomy concepts—A perspective. Frontiers in
Sustainability, 2, 701509. https://doi.org/10.3389/frsus.2021.701509
68. Tomita, A., Cuadros, D. F., Burns, J. K., Tanser, F., & Slotow, R. (2020). Exposure to waste
sites and their impact on health: A panel and geospatial analysis of nationally representative
data from South Africa, 2008–2015. The Lancet Planetary Health, 4(6), e223–e234. https://
doi.org/10.1016/S2542-5196(20)30101-7
69. Tripathi, N., Hills, C. D., Singh, R. S., & Atkinson, C. J. (2019). Biomass waste utilisation in
low-carbon products: Harnessing a major potential resource. NPJ Climate and Atmospheric
Science, 2(1), 35. https://doi.org/10.1038/s41612-019-0093-5
70. Uddin, M. N., Siddiki, S. Y. A., Mofijur, M., Djavanroodi, F., Hazrat, M. A., Show, P. L.,
Ahmed, S.F., & Chu, Y. M. (2021). Prospects of bioenergy production from organic waste
using anaerobic digestion technology: A mini review. Frontiers in Energy Research, 9, 627093.
https://doi.org/10.3389/fenrg.2021.627093
71. van Schalkwyk, D. L., Mandegari, M., Farzad, S., & Görgens, J. F. (2020). Techno-economic
and environmental analysis of bio-oil production from forest residues via non-catalytic and
catalytic pyrolysis processes. Energy Conversion and Management, 213, 112815. https://doi.
org/10.1016/j.enconman.2020.112815
72. Wang, Z., Burra, K. G., Lei, T., & Gupta, A. K. (2021). Co-pyrolysis of waste plastic and solid
biomass for synergistic production of biofuels and chemicals—A review. Progress in Energy
and Combustion Science, 84, 100899. https://doi.org/10.1016/j.pecs.2020.100899
73. Xue, Y., Zhou, S., Brown, R. C., Kelkar, A., & Bai, X. (2015). Fast pyrolysis of biomass and
waste plastic in a fluidized bed reactor. Fuel, 156, 40–46. https://doi.org/10.1016/j.fuel.2015.
04.033
74. Yaman, E., Ulusal, A., & Uzun, B. B. (2021). Co-pyrolysis of lignite and rapeseed cake: A
comparative study on the thermal decomposition behavior and pyrolysis kinetics. SN Applied
Sciences, 3, 1–15. https://doi.org/10.1007/s42452-020-04040-y
11 Pyrolytic Conversion of Heterogenic Natural Waste Biomass from Rural … 287

75. Ye, J. L., Cao, Q., & Zhao, Y. S. (2008). Co-pyrolysis of polypropylene and biomass. Energy
Sources, Part A, 30(18), 1689–1697. https://doi.org/10.1080/15567030701268419
76. Zhang, Y., Yu, Z., Zhang, X., & Ma, X. (2023). Comparative study on the synergistic co-
pyrolysis of Thlaspi arvense L. seed with different plastics: Thermal behaviors, product distri-
butions, and kinetics analysis. Biomass Conversion and Biorefinery, 13, 6197–6211. https://
doi.org/10.1007/s13399-021-01712-6
77. Zhao, Y., Wang, Y., Duan, D., Ruan, R., Fan, L., Zhou, Y., Dai, L., Lv, J., & Liu, Y. (2018).
Fast microwave-assisted ex-catalytic co-pyrolysis of bamboo and polypropylene for bio-
oil production. Bioresource Technology, 249, 69–75. https://doi.org/10.1016/j.biortech.2017.
09.184
78. Zhong, S., Zhang, B., Liu, C., Mwenya, S., & Zhang, H. (2022). A minireview on catalytic
fast co-pyrolysis of lignocellulosic biomass for bio-oil upgrading via enhancing monocyclic
aromatics. Journal of Analytical and Applied Pyrolysis, 164, 105544. https://doi.org/10.1016/
j.jaap.2022.105544
79. Zhou, G., Li, J., Yu, Y., Li, X., Wang, Y., Wang, W., & Komarneni, S. (2014). Optimizing the
distribution of aromatic products from catalytic fast pyrolysis of cellulose by ZSM-5 modifi-
cation with boron and co-feeding of low-density polyethylene. Applied Catalysis A: General,
487, 45–53. https://doi.org/10.1016/j.apcata.2014.09.009
80. Zhou, L., Wang, Y., Huang, Q., & Cai, J. (2006). Thermogravimetric characteristics and kinetic
of plastic and biomass blends co-pyrolysis. Fuel Processing Technology, 87(11), 963–969.
https://doi.org/10.1016/j.fuproc.2006.07.002
Chapter 12
Improving the Biogas Generation
Potential from Organic Wastes Using
Hydrochar as an Additive Lab-Scale
Case Study from Central Sweden: Part 1

Maria Kristoffersson, Maria Sandberg, and G. Venkatesh

Abstract At Biogasbolaget AB in Karlskoga in south-central Sweden, organic


wastes like food waste, manure, and silage are digested anaerobically to yield biogas,
which subsequently can be upgraded to biomethane, and used as a replacement for
fossil-diesel in public transport. The digesters at the firm are currently operating
below their maximum capacity. This chapter deals with the evaluation of the potential
of hydrochar to augment biogas production in a batch process. Hydrochar produced
from two sources—forestry sector and municipal organic wastes—were compared,
and using the Automatic Methane Potential Testing System (AMPTS II) in the lab
at Karlstad University, the optimal dosage was determined. Experiments were also
conducted with hydrochar alone, to verify if the hydrochar was being anaerobically
digested to yield biogas. The hydrochar sourced from municipal waste, when dosed
at 8 g/l, produced 841 Nml of biogas/gram of VS (volatile solids) in the substrate,
93% greater than the reference case of no addition of hydrochar. The forestry-sector-
sourced hydrochar on the other hand, at the same dosage, registered an increase of
just 16.6%. A streamlined environmental life cycle analysis showed that significant
climate-benefits can be availed of, implying environmental sustainability, when the
additional biogas is refined and used to replace fossil-diesel in public bus trans-
port. Hydrochar-assisted anaerobic digestion of organic wastes may be posited as
a technology which may entrench itself in the circular bio-economies of tomorrow,
and by doing so, contribute to a set of sustainable development goals. While these
were batch-digestion experiments, this part of the two-part series recommends more-
realistic continuous-digestion experiments which incidentally form the focus of Part
2.

M. Kristoffersson · M. Sandberg · G. Venkatesh (B)


Energy and Environmental Systems Group, Department of Engineering and Chemical Sciences,
Karlstad University, Karlstad, Sweden
e-mail: venkatesh.govindarajan@kau.se
M. Sandberg
e-mail: maria.sanberg@kau.se

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2024 289
G. Baskar et al. (eds.), Circular Bioeconomy Perspectives in Sustainable Bioenergy
Production, Energy, Environment, and Sustainability,
https://doi.org/10.1007/978-981-97-2523-6_12
290 M. Kristoffersson et al.

Keywords Anaerobic digestion · Automatic methane potential testing system ·


Batch process · Biogas · Circular bioeconomy · Hydrochar · Hydrothermal
carbonisation

12.1 Introduction

In the decade 2010–2019, the total anthropogenic greenhouse gas (GHG) emissions
amounted to approximately 54.4 billion ton-CO2 -equivalents (CO2 -e). It is by now
a clichéd fact that the annual GHG emissions hereafter must be halved (in 2030
with respect to 2019), if we intend and hope to keep the global average temper-
ature rise below 1.5 °C. If ‘net-zero’ is the target set, the GHG emissions which
will inevitably occur, will have to be compensated by ‘negative GHG emissions’
or in other words, carbon-negative solutions and alternatives. The United Nations
Environmental Programme (UNEP) estimates that by the end of this century, the
temperature would have risen by between 2.4 and 2.6 °C [1].
While that is an overview of the world, in Sweden (where this case study is based),
over the period 1990–2021, GHG emissions have decreased by 33% [1], from 71.4 to
47.8 million tons CO2 -eq., and Sweden has set for itself an ambitious target of ‘net-
zero GHG emissions’ in 2045 [2]. To reach there, among other things, the transport
sector which currently accounts for one-third of the country’s GHG emissions, ought
to be decarbonised. A 70% decrease in GHG emissions from the transport sector has
been targeted in 2030, with respect to 2019 [3]. In 2022, non-fossil fuels powered
about 93% of public transport in Sweden—39% hydrogenated vegetable oil, 31%
biogas, 14% rapeseed methyl ester, and 8% electricity (hydropower + nuclear energy)
[4]. Fossil-diesel accounted for approximately 7.2% of the total vehicle-kilometres
travelled by buses in 2022—equivalent to an energy use of 342.8 GWh.
Biogas (or biomethane when refined) is an output of anaerobic digestion of organic
materials (wastes, more often; and thereby a permanent fixture in the circular bio-
economies of the future). Compressed biogas finds use as fuel for cars, light trucks and
buses, while liquid biogas is more suitable for heavy vehicles, ships and industrial
applications, and as a substitute for fossil-based natural gas (which also contains
methane), it is poised to play a key role in combating climate change and truncating
the anthropogenic greenhouse gas footprint [5], and thereby contributing to nearing
the targets set under sustainable development goal (SDG) 13. Anaerobic digestion of
organic wastes lies at the confluence of sustainable waste management and renewable
energy generation (SDG 7), both of which are sine qua non is a circular bioeconomy.
Biochar and hydrochar, which will entrench themselves as useful and valuable
biorefinery outputs in future, can be produced from renewable organic wastes. They
will find use as fuels, adsorbents, soil conditioners, additives (catalysts if we may say
so) for composting and anaerobic digestion [6]. Hydrochar is produced through the
hydrothermal carbonisation (HTC) route. As opposed to biochar which is a product of
pyrolysis of organic wastes which need to be dried before being fed into the furnace,
moisture content is not a hindrance to the HTC process. Additives like activated
12 Improving the Biogas Generation Potential from Organic Wastes Using … 291

carbon, graphene, nanotubes, biochar, and hydrochar have merited attention of late;
the last two being posited as substitutes for other carbonaceous materials. Biochar
and hydrochar confer multiple benefits—low cost, diversity of sources from which
they can be produced, environmental sustainability, and the possibility of valorising
digestates.
The physico-chemical properties of hydrochar can be modified to suit different
end-use applications [7]. HTC is a suitable method for handling wastewater sludge,
but in that case, it has a tendency of being enriched in heavy metals (HM) like silver,
cadmium, chromium, copper, mercury, nickel, lead, and zinc [8]. They [8] have also
pointed out that if the starting material is sludge from the forestry sector, or digestate
from food waste anaerobic digester, the HM concentrations therein are much lower
vis-à-vis municipal wastewater sludge. The hydrochar may resist inhibitory agents in
a digester, enrich the microbial diversity in it, augment the activity of the acidogenic
bacteria, catalyse electron transfers among reactants, and promote the hydrolysis of
polysaccharides, proteins and lipids, and thus sustain biogas production [6].
Hydrochar addition to anaerobic digesters is a relatively nascent field of research,
with most of the peer-reviewed publications being less than five years old. All of them
are lab-scale studies conducted in a batch reactor, with the objective of estimating the
change in the biogas generation potential of the substrate being digested. Hydrochar
sourced from the digestate of an anaerobic digester [9], bamboo [10], rice straws [11],
and wastewater sludge [12], augmented biogas production by 49%, 131%, 127%,
and 18.4%, respectively, with respect to reference cases without hydrochar addition.
One of the partners in this research project—the Biogasbolaget (Biogas Company)
in Karlskoga in central Sweden—generates biogas by digesting a blend of silage,
slurry, and food waste gathered from the food processing sector and slaughterhouses
[13]. It supplies 48 GWh worth of transport fuel every year (equivalent to, and thereby
supplanting 5 million litres of fossil-diesel). In addition to producing fuel for the
transport sector, Biogasbolaget also supplies fertiliser in the form of the digestate
from the anaerobic digester. By augmenting the output of biogas, and thereby the
fuel it is refined to, more buses in the region can be powered by biomethane, bringing
down in the process, the consumption of fossil-diesel.
As mentioned earlier, the relative newness of this field of research warrants the
conduct of more lab-scale trials which would eventually lead to pilot-scale studies to
establish hydrochar in the circular bio-economies of the future, as a product which
would enable the valorisation of a range of organic wastes into biomethane. As there
are no studies conducted till date, with hydrochar produced from sludge sourced
from the forestry sector (which incidentally is a significant contributor to the Swedish
economy), this study—batch anaerobic digestion of organic substrate with hydrochar
from two provenances—HTC of forestry sludge and HTC of municipal sludge—can
be considered as a valuable addition to the knowledge base.
The specific goals set by the authors for this study are as under:
• To investigate if hydrochar augments biomethane production. If yes, to determine
the dosage at which the yield is the maximum (4, 8, or 10 g/l)
292 M. Kristoffersson et al.

• To compare the effect of hydrochar originating from two different provenances—


forestry sector sludge and municipal sludge—on the augmentation of the yield of
biomethane
• To carry out a simple environmental life cycle analysis (E-LCA) to study the
effects of hydrochar-assisted anaerobic digestion on the global warming potential,
and thereby study its potential to contribute to the sustainable development goals
7 and 13.
A brief background focusing on food waste (the substrate in the digester) and
the anaerobic digestion process itself, follows the introduction, before the method-
ology is briefly described. The results are presented and discussed succinctly before
concluding with some take-home messages in support of a circular bioeconomy, and
the interventions needed to advance towards one sooner than later.

12.2 Background

12.2.1 Food Waste as Substrate to the Digester

The global generation of food waste in 2025 would amount to 2.2 billion tons [14].
The quantities of food waste can be decreased, but cannot be entirely avoided. A
category called labelled as ‘inedible and unavoidable food waste’ which includes
items like fruit peels and skins, bone-bits (meat products), and spent coffee grounds
has been defined in literature [15]. There is also a huge fraction of avoidable food
waste, which can be decreased. While challenges posed by pollutants and pathogens
will loom larger, there will also be greater opportunities to extract value out of
the biodegradable fractions. While landfilling organic biodegradable wastes is a
sheer destruction of value, incineration for energy recovery and composting is better
approaches to adopt. However, the most environment-friendly method of handling
food waste is anaerobic digestion to yield biogas [16].
About 100,000 tons of food waste are collected annually in Sweden, for valori-
sation—corresponding to about 10 kg per capita. However, there is still a fraction of
food waste which, owing to the absence of proper source segregation, does not find
its way to the biological valorisation facilities—biogas generation and composting
[17]. When food waste is subjected to source segregation, pre-treatment in the
form of pulverisation, dilution, and separation from plastic and metal is manda-
tory. The biogas yield in the digester depends to a great extent on the segregation
and pre-treatment methods adopted.
12 Improving the Biogas Generation Potential from Organic Wastes Using … 293

12.2.2 Anaerobic Digestion and Biogas Production

From Hydrolysis to Methanogenesis: Biogas production in an anaerobic digester is


a series of steps which begins with hydrolysis of the substrates, proceeds to the forma-
tion of volatile fatty acids, therefrom to acetic acid and hydrogen, before yielding
methane and carbon dioxide (the mixture we refer to as biogas).
• Hydrolysis: Complex organic molecules like carbohydrates, fats, and proteins
which cannot serve as usable substrates are disintegrated enzymatically by
hydrolysing microorganisms to smaller ones like simpler sugars, fatty acids, and
alcohols. In general, when the substrate is particulate in nature, this turns out to
be the rate-limiting step in the anaerobic digestion process.
• Acidogenesis (Fermentation): The products of hydrolysis are acted upon by acido-
genic bacteria to yield volatile fatty acids (VFA) like acetic acid, propionic acid,
butyric acid, etc., and ammonia, carbon dioxide, and hydrogen.
• Acetogenesis: This is a reductive process in which carbon dioxide and the lower-
molecular weight organic VFAs are converted to acetate by anaerobic aceto-
genic bacteria. Autotrophic acetogens as opposed to the heterotrophic ones we
have been referring to heretofore can produce acetate from carbon dioxide and
hydrogen (both products of acidogenesis). It is noteworthy that acetate production
and methanogenesis take place simultaneously.
• Methanogenesis: Biogas can be formed either by the disintegration of acetate into
carbon dioxide and methane by acetoclastic (‘acetate breaking’) methanogens;
or by the reduction of carbon dioxide by hydrogen by the hydrogenotrophic
(‘hydrogen consuming’) methanogens. The former account for close to 70% of
the total biogas yield. When the substrate is food (and vegetable) waste, this
becomes the rate-limiting step, owing to the high activity of the acidogens which
results in rapid accumulation of VFAs, and a conspicuous drop in the pH which
damps down methanogenic activity. This is the rate-limiting step and is sensitive
to changes in operating conditions like temperature, pH, substrate composition,
feed rate, and leakage of oxygen into the chamber.
Role of Heavy Metals in the Biogas Generation Process: Heavy metals like cobalt
(Co), iron (Fe), and nickel (Ni)—micronutrients—are considered to play an enabling
role in the anaerobic digestion process by virtue of their reductive-oxidative prop-
erties. While a deficiency of Ni and Co depresses methane generation by 10% and
25%, respectively, heavy metals in very high concentrations—as is the case in indus-
trial wastewater and municipal sludge—are toxic to microorganisms. Copper (Cu),
Ni, and zinc (Zn) in the right amounts catalyse hydrolysis and sustain microbial
activity in the digester, in general. Acidogenesis is favoured by lower concentra-
tion of cadmium (Cd) and higher concentration of chromium (Cr) [18], while addi-
tions of Co, Zn, magnesium (Mg), niobium (Nb), calcium (Ca), Ni, Cd, and Fe in
the form of salts and metal oxides are common, to aid the four steps of anaerobic
digestion—described in the previous sub-section—by catalysing electron transfer
processes during enzymatic degradation [6].
294 M. Kristoffersson et al.

The Residual Digestate: The macronutrients—nitrogen and phosphorus—which


entered the anaerobic digester are left behind in the residual digestate, which thereby
has fertiliser value. As more and more of the organics are hydrolysed, and eventually
converted to methane and carbon dioxide, the nitrogen and phosphorus become more
easily accessible and available to the plants, when the digestate is transferred to the
soil on cropland. However, the presence of undesirables like heavy metals, pesticides,
pharmaceuticals, and disease-causing pathogens may make the use of this digestate
as fertiliser, questionable. If the digestate traces its origins to animal dung, source-
segregated food wastes, wastes from farms, slaughterhouses and the food processing
sector, and other types of easily biodegradable organic wastes, it is labelled as a ‘bio-
fertiliser’, and has to be certified by Avfall Sverige (Waste Management Sweden,
www.avfallsverige.se) as being fit for use as a fertiliser.

12.3 Methodology

12.3.1 Sourcing of Substrate, Graft, and Hydrochar

The study was conducted in the laboratory on the premises of Karlstad University
in early 2023. The substrate—a blend of food waste, slurry, and slaughterhouse
waste—was obtained from Biogasbolaget AB (one of the firms associated with this
study) and mixed with glycerol and refrigerated to ensure that it did not biodegrade.
The graft (active bacteria + organics) to be used along with hydrochar (sometimes
referred to as HC hereafter in the chapter) produced from forest-sector sludge was
placed in a warm compartment at 37 °C for 13 days, so that the organics could be
digested and converted to biogas which could be released out to the atmosphere,
and the residual active bacteria could be subsequently added on to the substrate. The
second graft to be used along with the HC sourced from municipal sludge, on the
other hand, was kept warm for just a day before being added on to the substrate. The
pulp and paper mill contributing sludge for hydrothermal carbonisation (HTC) was
Stora Enso’s Heinola-based (Finland) plant, and the HC-producer was C-Green.

12.3.2 Estimation of the Total Solids and Volatile Solids

Figure 12.1 is a self-explanatory simple sketch, which shows how the total solids
(TS) and volatile solids (VS) content of the different substrates and the two grafts
were calculated. While TS, as shown in Fig. 12.1, is the dry solids content of the
sample, the difference between TS and the Ashes (shown in the split bar on the far
right of Fig. 12.1) gives one the VS content (which can be expressed as a percentage
12 Improving the Biogas Generation Potential from Organic Wastes Using … 295

Fig. 12.1 Illustration of the TS and VS estimation tests

Table 12.1 Results of the TS and VS tests (values as % of total wet mass of sample)
Material Total solids (TS) Volatile solids (VS)
Substrate (average of all the types) 8.8 7.8
Graft (average)—used with HC from forest-sector 4.9 3.6
sludge
Graft (average)—used with HC from municipal 5.5 4.1
sludge
Silage (average) 26.5 5.2
Glycerol (average) 88 88

of TS, as well as a percentage of the total mass of the wet sample). The results—
averages calculated from multiple trials—are given in Table 12.1, in the Results and
Discussion section.

12.3.3 Setup of the Automatic Methane Potential Testing


System II

The laboratory-scale experiments in batch anaerobic digestion were conducted using


the Automatic Methane Potential Testing System (AMPTS II). This system utilises
fourteen 500-ml flasks (80% working volume), which are placed in a water bath
warmed and held at a mesophilic temperature of 37 °C. Mixing during the digestion
within each flask is accomplished by stirrers which operate continuously on a stir-
pause cycle (stirring for a minute and pausing for the next). The resultant biogas
is then ‘refined’ by dissolving the carbon dioxide in it, in sodium hydroxide. The
quantity of methane generated is measured continuously, by the AMPTS II software
programme, and the cumulative generation curve is plotted.
296 M. Kristoffersson et al.

12.3.4 Testing with Forest-Sector Hydrochar

Hydrochar having its provenance in paper mill sludge is pulverised and mixed with
substrate from Biogasbolaget. Studies done in the past have shown that the optimal
dosage of HC lies between 4 and 10 g/l. This motivated the selection of dosages 4,
8, and 10 g/l, for this set of experiments. As also mentioned earlier, the graft was
digested separately to determine its methane gas yield (this was done in duplicate).
The substrate mixed with the graft was digested without the addition of HC, and the
results thereof served as the reference case for comparison. All the tests (the reference
and those with the additions of different dosages of HC) were done in triplicate, over
a period of 18 days. For each of the four different dosages of HC (including the
reference case of 0 g/l), 54.8 g of graft and 11.86 g of substrate were employed.
About 57 g of graft were added to the two flasks designated for the determination of
the methane-yielding potential of the graft alone. De-ionised water was added to all
the flasks to bring the working volume in each of them to 400 ml, before measuring
the pH.
An additional test was conducted with graft + HC, without any substrate, to see if
the hydrochar itself yielded biogas. The leachability of VFAs from the HC was also
tested by mixing the HC (8 g/l) with de-ionised water, and letting it stay in the warm
water bath. This was necessitated to establish if the HC contributed VFAs to the
acetogens in the digester, to be converted eventually to methane and carbon dioxide.

12.3.5 Testing with Hydrochar Produced from Municipal


Sludge

All the tests described in Sect. 3.4 were repeated by simply changing the prove-
nance of the hydrochar from forest-sector sludge to municipal sludge. The digestion
happened over a period of 15 days, as opposed to 18 days for the previous set. C-
Green delivered the HC needed for this second set of tests in pulverised form, thus
obviating the need for an extra pulverisation step in the lab. The mass of graft needed
in each flask for the second set of tests was less by 5.53 g, vis-à-vis the first set
(49.27 g instead of 54.8 g).

12.3.6 Streamlined Environmental-LCA (E-LCA)

The hydrochar production through the HTC-route (at C-Green) uses electricity (500
kWh per ton of HC total solids). At the rate of 8 g/l HC dosage to anaerobic digestion
at Biogasbolaget, about 628 tons of HC would need to be produced and supplied by C-
Green, annually. It was deemed necessary and interesting to analyse the HC addition
process, and the far-reaching consequences thereof in the wider economy, from an
12 Improving the Biogas Generation Potential from Organic Wastes Using … 297

environmental perspective. The functional unit chosen was the excess refined biogas
resulting from the addition of 628 tons of HC per year, and the corresponding avoided
impacts from the production and use of an equivalent amount of diesel–fuel to power
buses was calculated. The HC production process using the HTC-route, anaerobic
digestion to produce biogas, refining of the biogas to biomethane, distribution of the
biomethane, and its combustion in the use phase was included within the system
boundary. The life cycle of diesel likewise—from rigs-to-road, so to say—was also
accounted for. Figure 12.2 shows the processes in the life cycles of diesel and HC-
assisted biogas production, refining, and use, and Table 12.2 summarises the relevant
data.

Fig. 12.2 Processes in the life cycles of biomethane and diesel, in the estimation of the net
greenhouse gas footprint

Table 12.2 Biogas-related data [22, 23]


Data Value Units
Low heating value of biogas (97% methane) 48.5 MJ/kg
Density of biogas (97% methane) 0.75 Kg/m3
Specific fuel usage (biogas bus) 19.4 MJ/km
Reference value of annual biogas production 5 million Nm3 /year
Biogas production when hydrochar produced from forest-sector 5.83 million Nm3 /year
sludge is added
Biogas production when hydrochar produced from municipal 9.67 million Nm3 /year
sludge is added
Life cycle GHG emissions of refined biogas 0.28 kg CO2 -eq/
km
Life cycle GHG emissions of fossil-diesel 1.1 kg CO2 -eq/
km
Additional energy usage for refining the extra biogas produced 845 GJ
(HPWS process)
298 M. Kristoffersson et al.

High Pressure Water Scrubbing (HPWS) is used at the Biogasbolaget to refine


biogas to biomethane, and electricity use amounts to 770 MJ/ton-CO2 -removed)
[19]. The value for the GHG footprint of electricity (Swedish mix) used for the HTC
and HPWS processes was obtained from the Ecoinvent database [20] in SimaPro
[21]. It was assumed that the marginal increase in the GHG footprint was equal to
that added on by the HPWS refining of the additional biogas generated by adding
HC. The net avoided GHG emissions in a year, when the extra biomethane produced
by adding hydrochar is used to substitute fossil-diesel as transport fuel for buses was
calculated.

12.4 Results and Discussion

The results for biomethane production, pH pre- and post-digestion, the VFA concen-
tration in the digestate, and the E-LCA are presented as compared between the first
and second sets of experiments—performed with HC from forest-sector sludge and
municipal sludge, respectively. Microscopy results pertain only to the second set.

12.4.1 Biomethane Production (BMP)

Figures 12.3 and 12.4 present the cumulative methane production from 12 and 14
flasks of the AMPTS II system, for municipal-sludge-sourced and forest-sludge-
sourced HC, respectively. The red line depicts the production of methane in the flask
with only the graft in it; the other values are adjusted by subtracting this value from
them. In Fig. 12.3, among the two 8 g/l yellow lines, the upper one depicts methane
production equal to 881 ml/g VS. The uppermost of the three blue lines (10 g/l
addition) corresponds to a production of 904 ml/g VS. These have to be analysed
vis-à-vis the orange reference line (no HC addition) which reaches 485 ml/g VS.
Also observed in Fig. 12.3 is the fact that the production tapers off after 5 days for
the reference case, while the others where HC is added to the substrate-graft blends
in the flasks, register a late rise to much higher values than the reference case at the
end of the 15-day period.
As seen in Fig. 12.4, the 8 g/l addition of HC from forest-sector sludge (yellow
line) yielded the greatest amount of methane (560 ml/gVS) over the 18-day period—
about 100 ml/gVS greater than the reference case (orange line). It is also clear that
4 mg/l HC addition yielded less methane than the reference case, the reason thereof
not being very clear.
12 Improving the Biogas Generation Potential from Organic Wastes Using … 299

Fig. 12.3 Biomethane potential (BMP) for reference case, graft, and different concentrations of
hydrochar from municipal sludge, over 15 days

Fig. 12.4 Biomethane potential (BMP) for reference case, graft, and different concentrations of
hydrochar from forest-sector sludge, over 18 days

12.4.2 PH Before and After Digestion

The average pH values recorded before and after the digestion are given in Table 12.3.
For the set of experiments done with HC from municipal sludge, the pH decreased
slightly for the reference and the HC addition cases, oscillating around neutrality for
the reference. As pointed out by authors earlier [24], the optimal pH should be in
the range of 6.5–7.2. The pH in the flask with only the graft decreased, but stayed
slightly alkaline all throughout. For the other set, with HC sourced from forest-sector
300 M. Kristoffersson et al.

Table 12.3 pH values pre- and post-digestion


Case Hydrochar (HC) from municipal sludge Hydrochar (HC) from forest-sector sludge
pH pH % pH pH %
pre-digestion post-digestion change pre-digestion post-digestion change
0 g/l 7.2 6.92 − 3.9 7.36 7.5 1.9
HC
4 g/l 7.31 6.88 − 5.9 7.44 7.46 0.3
HC
8 g/l 7.29 6.83 − 6.3 7.5 7.54 0.5
HC
10 g/l 7.25 6.82 − 5.9 7.5 7.49 − 0.1
HC
Only 7.95 7.22 − 9.2 8.25 8.59 4.1
graft

sludge, the pH showed a tendency to increase slightly, while staying always in the
slightly alkaline range. Markedly, the decreases in the pH for the experiments done
with HC from municipal sludge were higher in general than the increases in the same
for those done with HC from forest-sector sludge. Nutrient addition (for nitrogen)
along with the HC from the forest-sector sludge to the anaerobic digester is common
[8]. This results in a higher ammonium concentration in the digester and thereby an
increase in pH (as seen in Table 3) that there was a difference in the duration over
which the digestion proceeded—15 days where HC from municipal sludge was used,
and 18 days for the other—may also be a contributing factor to the increase in the
pH for the latter.

12.4.3 Volatile Fatty Acids (VFAs)

The VFA concentration in the digestate, as seen in Fig. 12.5, generally increased as
the dosage of the HC rose. However, it is unclear why there was a drop in the VFA
concentration with respect to the reference case, from 360 to 327 mg/l, when 4 mg/
l HC from forest-sector sludge was added. The span between the reference case and
10 mg/l addition is 100 mg/l and 77 mg/l, respectively, for the bars shown in blue
(HC from forest-sector sludge) and red (HC from municipal sludge), respectively.
The leaching test referred to earlier and showed that HC did indeed release VFAs to
the de-ionised water—an average of 77.5 mg/l and 84 mg/l, respectively, for 8 mg/l
HC from municipal sludge and forest-sector sludge.
12 Improving the Biogas Generation Potential from Organic Wastes Using … 301

Fig. 12.5 VFA concentrations in the digestate, for both sets of experiments

12.4.4 Observations from Microscopy

The digestate from the experiments conducted with HC from municipal sludge was
observed under the microscope. The dark blobs in Fig. 12.6 shown by red arrows
are the residual fragments of HC in the digestate. Around the blobs, and on their
surfaces, one can notice what presumably are bacterial colonies in biofilms. Quite
obviously, it was easier to locate HC blobs when HC was initially added at higher
concentrations (8 and 10 mg/l), as is also clear from Fig. 12.6.

12.4.5 E-LCA Results

The E-LCA results are presented in Fig. 12.7. The increase in biogas generation by
adding HC (produced from forest-sector sludge or municipal sludge) entails addi-
tional resource usage and thereby GHG emissions during the biogas to biomethane
refining step (refer Fig. 12.2). As seen from Table 12.2, the additional biogas gener-
ated by adding HC sourced from municipal sludge (9.67 million Nm3 ) is much
greater than that generated by adding HC from the other provenance (5.83 million
Nm3 ). This is reflected in the difference between the upper and central black bars in
Fig. 12.7 (3091 tons and 5164 tons of CO2 -eq.). Additional biomethane production,
as per the system expansion assumption made in the E-LCA, implies substitution
of a greater quantity of fossil-diesel as transport fuel for buses. This is reflected by
the two darker green bars (upper and central), and the difference between them. The
net avoided GHG emissions, as shown in Fig. 12.7, are 14,783 tons (for HC from
municipal sludge) and 8937 tons CO2 -eq (for HC from forest-sector sludge).
302 M. Kristoffersson et al.

Fig. 12.6 Microscopy of the digestate, clearly showing residual fragments of hydrochar and
biofilms in their vicinity (HC sourced from municipal sludge)

Fig. 12.7 Comparative E-LCA results for the GHG emissions and net avoided emissions
12 Improving the Biogas Generation Potential from Organic Wastes Using … 303

12.4.6 Overview of the Findings

It has been clearly established that addition of HC to the anaerobic digester yielded
more biogas, and that is an interesting finding for a circular bioeconomy which
is gradually taking shape. The effect of the provenance of the HC has also been
clearly demonstrated through the trials conducted in the laboratory. HC addition to
the substrate provides more VFA to the methanogenic bacteria, as noted in earlier
studies [6, 11], and thereby the possibility of producing more biogas. The fact that
VFAs (which could have been converted to biogas by the methanogens) were found
in the digestate after the 15-day and 18-day trials (see Fig. 12.5) can be attributed
to the fact that the trials could have been prolonged by a few more days to further
enhance the output of biogas.
The microscopy test revealed that bacterial biofilms find a habitat for themselves
on the surfaces of the HC chunks in the digestate. The E-LCA showed clearly that
there are huge environmental benefits to be availed of, by investing in the HTC process
for HC production, and subsequently using it to augment the anaerobic digestion of
substrates of different types. A clear contribution to the Sustainable Development
Goals 7 and 13 has been demonstrated.

12.5 Conclusions and Recommendations

The batch-digestion experiments presented in this chapter (the first of a two-part


series) augur well as a justification for the production of HC and its utilisation in
anaerobic digestion. Municipal sludge emerges as a better option vis-a-vis forest-
sector sludge, as it yields much more biogas.
The utilisation of the digestate from Karlskoga’s Biogasbolaget (thereby intro-
ducing circularity within the system) as the in-feed to the HTC equipment to produce
hydrochar is also something which may be of interest to the firm. One of the recom-
mendations of the first part—testing by mixing HC from municipal sludge with
substrate in a more-realistic continuous-digestion process—leads seamlessly to Part
2, which also presents results of a life cycle costing analysis to establish the economic
sustainability of this intervention which has now been proven to be environmentally
sustainable.

Acknowledgements The authors would like to thank everyone at C-Green, Biogasbolaget and
Karlstad University for their help and support with this study, at different points of time in the first
half of 2023.
304 M. Kristoffersson et al.

References

1. Naturvårdsverket [Nature Conservation Agency], n.d. (2023). Globala utsläpp av växthusgaser


[Global greenhouse gas emissions]. Accessed on March 30, 2023 at https://www.naturvard
sverket.se/amnesomraden/klimatomstallningen/det-globala-klimatarbetet/globala-utslapp-av-
vaxthusgaser/
2. Sverigesmiljomål [Sweden’s environmental targets]. (2023). Begränsad klimatpåverkan -
Sveriges miljömål [Limiting climate effects—Sweden’s environmental targets]. Accessed on
March 31, 2023 at https://sverigesmiljomal.se/miljomalen/begransad-klimatpaverkan/
3. Naturvårdsverket [Nature Conservation Agency], n.d. (2023). B. Sveriges utsläpp och upptag
av växthusgaser [Sweden’s GHG emissions and capture]. Accessed on the March 31, 2023,
at https://www.naturvardsverket.se/data-och-statistik/klimat/sveriges-utslapp-och-upptag-av-
vaxthusgaser/
4. Bussar-Miljöbarometern [Buses-The Eco-barometer]. (2023). Accessed at https://2030.miljob
arometern.se/nationella-indikatorer/bilen/andel-fossiloberoende-fordon-i-trafik-b1c/bussar/,
on April 27, 2023.
5. Norrman, J., Arnell, J., Belhaj, M., & Flodström, E. (2005). Biogas som drivmedel för bussar
i kollektivtrafik [Biogas as fuel in the public transportation sector]. IVL Svenska Miljöin-
stitutet. Accessible at https://www.ivl.se/publikationer/publikationer/Biogassomdrivmedelför
bussarikollektivtrafik
6. Cavali, M., Libardi Junior, N., Mohedano, R. A., Belli Filho, P., da Costa, R. H. R., & de
Castilhos Junior, A. B. (2022). Biochar and hydrochar in the context of anaerobic digestion for
a circular approach: An overview. Science of The Total Environment, 822, 153614. https://doi.
org/10.1016/j.scitotenv.2022.153614
7. Quintana-Najera, J., Blacker, A. J., Fletcher, L. A., & Ross, A. B. (2021). The effect of augmen-
tation of biochar and hydrochar in anaerobic digestion of a model substrate. Bioresource Tech-
nology, 321, 124494. https://doi.org/10.1016/j.biortech.2020.124494; Starr, K., Gabarrell, X.,
Villalba, G., Talens, L., & Lombardi, L. (2012). Life cycle assessment of biogas upgrading
technologies. Waste Management, 32, 991–999.
8. Baresel, C., Axegård, P., Lazic, A., Bornold, N., Yang, J.-J., & Malovanyy, A. (2023).
Framtidens slamhantering vid Roslagsvatten (The future of sludge handling at Roslagsvatten).
Accessible at https://ivl.se/publikationer/publikationer/framtidens-slamhantering-vid-roslag
svatten-behandling-av-kommunalt-orotat-slam-med-htc-teknik-oxypower-htctm-och-rening-
av-htc-vatten-med-sbr-och-mbbr.html. IVL svenska miljöinstitutet, Stockholm.
9. Xu, S., Wang, C., Duan, Y., & Wong, J.W.-C. (2020). Impact of pyrochar and hydrochar
derived from digestate on the co-digestion of sewage sludge and swine manure. Bioresource
Technology, 314, 123730. https://doi.org/10.1016/j.biortech.2020.123730
10. Choe, U., Mustafa, A. M., Lin, H., Xu, J., & Sheng, K. (2019). Effect of bamboo hydrochar on
anaerobic digestion of fish processing waste for biogas production. Bioresource Technology,
283, 340–349. https://doi.org/10.1016/j.biortech.2019.03.084
11. Xu, J., Mustafa, A. M., Lin, H., Choe, U. Y., & Sheng, K. (2018). Effect of hydrochar on
anaerobic digestion of dead pig carcass after hydrothermal pretreatment. Waste Management,
78, 849–856. https://doi.org/10.1016/j.wasman.2018.07.003
12. Wu, B., Yang, Q., Yao, F., Chen, S., He, L., Hou, K., Pi, Z., Yin, H., Fu, J., Wang, D., & Li, X.
(2019). Evaluating the effect of biochar on mesophilic anaerobic digestion of waste activated
sludge and microbial diversity. Bioresource Technology, 294, 122235. https://doi.org/10.1016/
j.biortech.2019.122235
13. Karlskoga Energi. (2018). Så gör vi biogas [This is how we produce biogas]. Accessed on the
January 2, 2023 at http://www.karlskogaenergi.se/om-oss/vara-verksamheter/biogas/
14. Xu, Q., Luo, L., Li, D., Johnravindar, D., Varjani, S., Wong, J. W. C., & Zhao, J. (2022).
Hydrochar prepared from digestate improves anaerobic co-digestion of food waste and sewage
sludge: Performance, mechanisms, and implication. Bioresource Technology, 362, 127765.
https://doi.org/10.1016/j.biortech.2022.127765
12 Improving the Biogas Generation Potential from Organic Wastes Using … 305

15. Teigiserova, D. A., Hamelin, L., & Thomsen, M. (2019). Review of high-value food waste
and food residues biorefineries with focus on unavoidable wastes from processing. Resources,
Conservation and Recycling, 149, 413–426. https://doi.org/10.1016/j.resconrec.2019.05.003
16. Zhou, Y., Engler, N., Li, Y., & Nelles, M. (2020). The influence of hydrothermal operation
on the surface properties of kitchen waste-derived hydrochar: Biogas upgrading. Journal of
Cleaner Production, 259, 121020. https://doi.org/10.1016/j.jclepro.2020.121020
17. Avfall Sverige. (2023). I mål med matavfallet [On track with food waste]. Accessed at https://
www.avfallsverige.se/fakta-statistik/insamling/matavfall/i-mal-med-matavfallet/, on April 3,
2023.
18. Kadam, R., Khanthong, K., Jang, H., Lee, J., & Park, J. (2022). Occurrence, fate, and impli-
cations of heavy metals during anaerobic digestion: A review. Energies, 15, 8618. https://doi.
org/10.3390/en15228618
19. Starr, K., Gabarrell, X., Villalba, G., Talens, L., & Lombardi, L. (2012). Life cycle assessment
of biogas upgrading technologies. Waste Management, 32, 991–999. https://doi.org/10.1016/j.
wasman.2011.12.016
20. Swiss Center for Life-Cycle Inventories. Ecoinvent database V3.4 (2019). Technical report.
Accessible at www.ecoinvent.org
21. PRé Consultants. (2023). SimaPro 9.4. https://doi.org/10.1016/j.wasman.2011.12.016
22. Biogasbolaget i Mellansverige AB. (2018). KarlskogaEnergi. Accessed on February 1, 2023
at http://www.karlskogaenergi.se/Vara-tjanster/biogas/biogasbolaget/
23. Lyng, K.-A., & Brekke, A. (2019). Environmental life cycle assessment of biogas as a fuel for
transport compared with alternative fuels. Energies, 12, 532. https://doi.org/10.3390/en1203
0532
24. Ayodele, O. O., Adekunle, A. E., Adesina, A. O., Pourianejad, S., Zentner, A., & Dornack,
C. (2021). Stabilization of anaerobic co-digestion of biowaste using activated carbon of
coffee ground biomass. Bioresource Technology, 319, 124247. https://doi.org/10.1016/j.bio
rtech.2020.124247
Chapter 13
Improving the Biogas Generation
Potential from Organic Wastes Using
Hydrochar as an Additive Lab-Scale
Case Study from Central Sweden: Part 2

Annette Liisa Kariis, Maria Sandberg, and G. Venkatesh

Abstract In Part 2 of the two-part series, single-stage anaerobic co-digestion in two


continuously fed reactors replaced the batch process of Part 1. A life-cycle costing
analysis was carried out to determine if, investing in a hydrothermal carbonization
(HTC) system to produce hydrochar in-plant to augment biogas production will be
economically feasible. Hydrochar addition resulted in a 59% rise in biogas yield (and
53.5% in methane yield). The study confirmed the techno-economic feasibility for
coupling an HTC plant with a digester supplying 25% of the digestate it produces,
to the former, as the raw material for hydrochar production. The rest of the digestate
(rich in carbon, nitrogen, and phosphorus) can be used as fertiliser. Investing in
an HTC plant contributing to a rise in methane production of 17% (or 53%) will
result in a net profit of 363 million SEK (or 1237 million SEK) over a 20-year
period. If the Karlskoga biogas plant decides to rely on purchasing hydrochar from
the external market instead, the corresponding net profit will be 177 million SEK
(or 1052 million SEK) over the same 20-year period, implying that a decision to
integrate and interconnect is likely to be economically more feasible, in a circular
bioeconomy in the future.

Keywords Anaerobic digestion · Continuous reactor · Biogas · Circular


bioeconomy · Life-cycle costing · Hydrochar · Hydrothermal carbonisation · Net
present value

A. L. Kariis · M. Sandberg · G. Venkatesh (B)


Energy and Environmental Systems Group, Department of Engineering and Chemical Sciences,
Karlstad University, Karlstad, Sweden
e-mail: venkatesh.govindarajan@kau.se
M. Sandberg
e-mail: maria.sanberg@kau.se

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2024 307
G. Baskar et al. (eds.), Circular Bioeconomy Perspectives in Sustainable Bioenergy
Production, Energy, Environment, and Sustainability,
https://doi.org/10.1007/978-981-97-2523-6_13
308 A. L. Kariis et al.

13.1 Introduction

In the second decade of this century, the average annual greenhouse gas (GHG)
emissions rose to a record high in human history [1]. As indicated in Part I of this
series, the realisation that a decoupling between socio-economic development and
dependence on fossil fuels is indispensable in the years to come has been a shot in
the arm for a circular bioeconomy, which is gradually entrenching itself.
Food waste is among the handful of substitutes for fossil fuels, one can avail
of. In 2019, 931 million tons of food waste were generated globally, of which 61%
emanated from households [2]. The Swedish Food Agency has repeatedly empha-
sised on the indispensability of both decreasing avoidable food waste, and valorising
the unavoidable fraction to the greatest extent possible, in biorefineries of a future
circular bioeconomy [3].
The most common approach to the valorisation of food wastes globally has been
anaerobically digesting (or co-digesting with wastewater sludge, and agricultural
wastes) them to yield biogas (or biomethane, after refining). Biogas is an energy
source for the generation of heat and electricity, as well as the production of transport
fuel. The resulting digestate which is rich in nutrients (phosphorus especially) finds
use as organic fertiliser. Substituting conventional fossil fuels hereby, and decreasing
the need for the production of synthetic fertilisers, uncovers significant environmental
benefits, while generating additional income to more people along the value chain—
in other words, being socially desirable, economically attractive, and environment-
friendly. Over the 10-year period between 2009 and 2018, the number of biogas plants
in Europe (micro-, small-, and industrial scales of production) increased rapidly,
tripling from 6200 to 18,202 [4]. As they have proliferated, new challenges have
emerged. These have spurred more research in a multitude of aspects related to anaer-
obic digestion of organic wastes—optimisation of the operational variables, config-
uration of the bioreactors, opportunities for and the effects of co-digesting, control
of nutrients, pretreatment of the substrates to improve biodegradability (digestability
in other words), and the effect of additives like hydrochar and biochar on the yield
of biogas.

13.1.1 Hydrochar as an Additive

Hydrochar—one of the additives referred to in the previous paragraph—is a carbona-


ceous material produced using biomass as the starting material, through a process
called hydrothermal carbonisation (HTC). The carbonisation occurs in an aqueous
medium (‘hydro’), which serves both as solvent and catalyst, at around 160–280 °C
(‘thermal’), at above the saturation pressure at the operating temperature/s. At a
supercritical state, water contributes to an increase in the solubility and reactivity of
the biomass. The carbonisation mechanisms include hydrolysis, dehydration, decar-
boxylation, condensation, polymerisation, and aromatisation [5, 6]. The products
13 Improving the Biogas Generation Potential from Organic Wastes Using … 309

comprise a gaseous fraction (mostly carbon dioxide), a liquid fraction which is a


mixture of organic compounds, and a solid fraction which is the hydrochar in ques-
tion. While the need for drying the biomass prior to the process is obviated (this
being a process which happens anyway in an aqueous medium), the resulting slurry
needs to be dewatered. However, the energy requirement for the dewatering of the
slurry is much less than what would have been required for pre-process-drying if
the latter had been necessary [7] (This has a favourable effect on the environmental
footprint of the HTC process, which has been dwelt upon in Part I).
An addition of 10 g/l of hydrochar (from rice straw) augmented the biogas yield,
in a study [8], by 90% when the substrate was pre-treated pork wastes. Two years
thereafter, a 19% rise in biogas production was observed when hydrochar produced
by hydrothermally carbonising food wastes was added to an anerobic digester which
was using the same type of food wastes, as substrates [9]. A majority of the studies
have resorted to the batch process (adopted in Part I of the current study). Among the
published studies which have adopted a 25-day continuous digestion process [10,
11], the former reported a 5% rise in biogas yield and a 0% rise in methane yield
[with municipal organic waste as substrate and hydrochar sourced from spent coffee
grounds] while the latter reported a 12% increase in methane yield [air-dried chicken
manure as substrate and pyrochar from fruit wastes].
Hydrochar is reported to have a strong buffering capacity and an ability to tone
down the effects of inhibitory agents in the digester [4, 12], a larger surface area,
greater pore sizes, and pore volume vis-à-vis pyrochar [13]. While being a carrier
for the substrate, it tends to promote the colony-formation (and sequestration or
sheltering) of bacteria on the aforesaid large surface area, facilitating quicker and
easier direct interspecies electron transfer (abbreviated as DIET by Cavali et al.,
Ayodele et al., [1, 14]), leading consequently to improved hydrolysis, acidogensis,
acetogenesis and subsequently, a rise in biogas production [12]. Colony formation
concentrates active microorganisms—all the different types described in Part I; an
intercellular (bacterial) distance of less than 1 µm is deemed to be necessary to ensure
oxidation of the volatile fatty acids and hydrogen production.
DIET is further enhanced by the presence of oxygenated functional groups (C–O,
C=O, and C–OH among others) which abound in hydrochar [1]. The fact that the
HTC process is conducted at a lower temperature than pyrolysis is the reason behind
the presence of many more oxygenated functional groups in hydrochar as compared
to pyrochar. These oxygenated functional groups have both reductive and oxidative
abilities and thereby are able to both accept electrons from bacteria and donate
electrons to methanogens [15]. The alkaline nature of hydrochar (pH remaining
close to 7.4) also contributes to optimising the pH in the digester, and thus favours
methanogenesis (methanogens are sensitive to lower pH values) [8]. Slightly alkaline
pH values promote the formation of more VFAs, thanks to greater activity on the
part of acidogens. If the hydrochar is produced from wastewater sludge, humus-like
constituents in it improve the oxidative ability of the hydrochar and augment the
production of short-chain fatty acids like acetic acid and propionic acid. Hydrochar
which remains in the digestate improves the fertiliser value of the same, by ensuring
that nutrients—nitrogen and phosphorus—are sequestered within the matrix. It also
310 A. L. Kariis et al.

immobilises heavy metals, contributing in its own way to reduced leaching of the
same to the water bodies [4].

13.1.2 Objectives

In that backdrop generated by referring to past studies conducted with hydrochar,


and as a continuation from Part I of this two-part series, the objectives of this study
are as under:
• To investigate how the addition of hydrochar influences biogas (and methane)
production, and the stability of a more-realistic continuous anaerobic digestion
process, as opposed to the batch process adopted in Part I.
• To determine if it would be possible to interconnect the anaerobic digester with
an HTC-reactor and study the associated material flows.
• To assess the economic feasibility of producing hydrochar in-plant vis-à-vis
purchasing it from an external source, with biogas being refined to biomethane
(as detailed in Part I) and sold to the transport sector.

13.2 Methodology

13.2.1 Characterisation of the Hydrochar Used in the Study

The hydrochar used in the experiments, similar to Part I, was sourced from C-Green.
Data about its characteristics [16] have been listed in Table 13.1.

Table 13.1 Characteristics


Specific area (BET) 35 m2 /g
of the hydrochar used in the
study Conductivity 947 µS/cm
Moisture content 0.9%
pH 5.3
Ash content 22.8%
Metal Concentration (mg/kg TS)
Silver 0.1
Cadmium 1.42
Chromium 19
Copper 27.5
Nickel 35.6
Zinc 392
Lead 4.2
13 Improving the Biogas Generation Potential from Organic Wastes Using … 311

13.2.2 Setup in the Laboratory

Two lab-scale single-stage continuous anaerobic digestion reactors (one of them,


the reference; the other for tests with hydrochar addition) were set up at Karlstad
University (refer to Fig. 13.1). Tests were performed between 24 February and 2 May
2023. The tanks were fed and emptied regularly, and the volume of biogas collected
was measured on a prescheduled daily basis.
The reactors shown in Fig. 13.1 (volume of 15 L each) had an inlet for the substrate
(which was injected in) and outlets for the digestate and the biogas. The digestate
outlet was coupled to a collection tank by a tube. The biogas which was generated
was collected at the backside of the reactors in a plastic bag as shown. The reactors
had their dedicated stirrers driven by motors placed atop them, and immersion heaters
maintained the temperature at 35 ± 1 °C required for mesophilic digestion.

Fig. 13.1 The laboratory setup for the continuous anaerobic digestion trials
312 A. L. Kariis et al.

13.2.3 Substrate and Graft Acquisition from Biogasbolaget

As with the experiments in Part I, Biogasbolaget provided the graft (digestate from
its anaerobic digester) and substrate (a mixture of food waste, glycerol, manure, and
silage). Glycerol, as is commonly known, aids biogas production. The Total Solids
(TS) content and Volatile Solids (VS) content were estimated (refer to Part I). As
gathered from a handbook online [17], a reasonable organic loading rate (OLR)
would be 3–5 g VS per litre graft per day. However, to adapt to the conditions in the
laboratory reactors, 10 g VS per litre graft was chosen as the OLR. The composition
of the mixture fed to the reactor was 90.9% food waste + manure, 8.5% silage, and
0.6% glycerol; equivalent to 25 L of substrate, 1.5 dl of glycerol, and 351 g of silage.
For the reference case (without addition of hydrochar), 118 ml of substrate +
silage + glycerol in the proportion stated in the previous paragraph were mixed
with 481 ml deionised water. For the test case, 4.8 g of finely ground hydrochar
(corresponding to 8 g/l) were added to the substrate-mixture first, before dilution
with deionised water, to ensure complete dissolution of the hydrochar. The diluted
substrate-mix was introduced into the reactors using a 100 ml plastic syringe.

13.2.4 Measurements in the Laboratory

The volume of biogas was measured every day using a 100 ml air-tight glass syringe.
In the last week of the 68-day long experiment, the methane content in the biogas was
also measured using a gas chromatograph with a flame ionising detector (Clarus 480;
non-polar capillary column; J&W Scientific, Elite-5, 30 m × 0.25 mm × 0.25 µm;
using helium at 7 bar pressure).
The TS in the filtrate from the digestate which was collected in the tank depicted
in Fig. 13.1 was measured and compared with that in the substrate-mix, to determine
the TS reduction during the digestion process. The pH of the same was measured on
a daily basis, in order to keep track of the degree of stability. The activity of microor-
ganisms is adversely affected at high (over 8) and low (below 7) pH values. The VFA
and TOC contents in the digestate were determined according to the Hach Lange
LCK365 and LCK 386 standards respectively. LCK365 is intended for concentra-
tions in the range of 75–3600 mg/l propionic acid, while LCK 386 is meant for
30–300 mg/l.

13.2.5 Life-Cycle Costing

Life-cycle costing (LCC) of a project or a business undertaking accounts for the


present value (PV, determined on the basis of a chosen discount rate) of all the
expenditure and revenue streams over its operating life-cycle and calculates the net
13 Improving the Biogas Generation Potential from Organic Wastes Using … 313

present value (NPV) as the difference between the PV of the revenue streams and that
of the expenditure streams [18]. An LCC was carried out for two different scenarios,
to determine which of the two would be more feasible/profitable economically.
• Scenario 1: Biogasbolaget will have its own captive hydrochar-producing HTC-
unit which will supply the hydrochar to be added to the anaerobic digester. A
greater portion of the hydrochar will be sold on the market, while biogas will be
sold to the public transport sector.
• Scenario 2: Biogasbolaget will purchase the amount of hydrochar (HC used as
acronym at some junctures in the text) needed to be added to the digester, from
an external supplier. The biogas produced, will be sold to the transport sector.
Data pertaining to C-Green [19] and Biogasbolaget [20] relevant for the LCC
calculations are summarised in Table 13.2.

Table 13.2 Data used in the LCC analysis


Data item Value Source/Reference
Capital investment for an 80 million SEK [19]
HTC-reactor
Functional lifetime of the 20 years
HTC-reactor
Operation and maintenance 1.6 million SEK per year
expenses (estimated in the year prior to
the start of the functional
lifetime, so-called year-zero in
LCC)
Energy demand during the 500 kWh electricity/ton HC
operation phase
Total annual output 5000 ton HC total solids
Electricity tariff 1.37 SEK/kWh [21]
Biomethane produced as 40 GWh [20]
transport fuel
Conversion 1 Nm3 biomethane fuel = [22]
factor—Nm3 biomethane fuel 9.67 kWh
to energy
Price of biomethane (as fuel) 17.5 SEK/Nm3 [20]
Estimated price of purchased 10,000 SEK/ton HC solids [19]
hydrochar (estimated in the year prior to
the start of the functional
lifetime, so-called year-zero in
LCC)
Total annual HC demand at 623 ton HC solids/year Calculated value
Biogasbolaget
Discount rate r = 5% [23]
Inflation rate i = 10% Swedish Statistics Agency
(www.scb.se)
314 A. L. Kariis et al.

Corresponding to a given demand for hydrochar at Biogasbolaget (623 tons HC


solids per year), the rise in biomethane yield in Part I (batch process) was 17%. This
study (as stated later on in the article) showed that a rise of 53% could be achieved
by availing of continuous anaerobic digestion.
The present value (PV) of all the costs associated with the building, operation, and
maintenance of an HTC-facility at Biogasbolaget, when apportioned to the 623 tons
of HC supplied to the digester can be calculated according to Eq. 13.1. Note that
623 tons account for 12.4% of the total estimated annual output of 5000 tons.
(( ( )) )
(1 + i)1 (1 + i)2 (1 + i)20
PVHTC = I + (O + M ) + + ... + /5000 ∗623
(1 + r)1 (1 + r)2 (1 + r)20
(13.1)

In Eq. 13.1 (pertaining to Scenario 1), ‘I’ stands for the initial capital investment,
‘O + M’ stands for the annual operation and maintenance expenses estimated in the
year prior to the start of operation (the so-called year-zero in LCC). This increases
from year to year, at the inflation rate. The values ‘i’ and ‘r’ represent the inflation
rate and the discount rate respectively. The former is assumed to be 10% (even though
the average in Sweden till before 2020 has been around 2%; inflation has been very
high in the third decade of this century following the pandemic and the Ukraine
crisis), and the latter 5%.
The PV of the costs associated with the purchase of hydrochar from an external
supplier over the 20-year period (the same as the functional lifetime of the HTC-
facility in Scenario 1) can be calculated using Eq. 13.2.
( )
(1 + i)1 (1 + i)2 (1 + i)20
PVpurchase, HC = Annual HC cost + + ··· +
(1 + r)1 (1 + r)2 (1 + r)20
(13.2)

‘Annual HC cost’ in Eq. 13.2, as listed in Table 13.2, is the cost of purchase as
determined in year-zero, similar to ‘O + M’ in Eq. 13.1. It increases from year to
year at the rate of inflation. The PV of the marginal increase in annual benefit owing
to the rise in the production of biomethane fuel (17% in Part I and 53% in this study)
can be calculated using Eq. 13.3.
( )
(1 + i)1 (1 + i)2 (1 + i)20
PVadditional benefit = Annual add. benefit + + · · · +
(1 + r)1 (1 + r)2 (1 + r)20
(13.3)

‘Annual additional benefit’ is determined on the basis of known price of


biomethane fuel in the year prior to the start of operation. This, like ‘O + M’ in
Eq. 13.1 and ‘Annual HC cost’ in Eq. 13.2, will increase from year to year at the rate
of inflation.
13 Improving the Biogas Generation Potential from Organic Wastes Using … 315

The Net Present Value (NPV) was subsequently determined for both scenarios. For
the first scenario, ‘NPVHTC ’ is the difference between PVadditional benefit (Eq. 13.2) and
PVHTC (Eq. 13.1). The other one ‘NPVHC ’ is the difference between PVadditional benefit
(Eq. 13.2) and PVpurchase, HC (Eq. 13.3).

NPVHTC = PVadditional benefit − PVHTC (13.4)

NPVHC = PVadditional benefit − PVpurchase, HC (13.5)

The capital investment listed in Table 13.2 corresponds to a total annual production
of 5000 tons of HC solids. At a concentration of 8 g HC/l, one would need 623 tons
of HC solids per year for the digester. If Biogasbolaget would design an HTC-facility
exactly similar to the one at C-green (including the capacity and the output), it would
have to deal with surplus hydrochar. It can very well scout for potential buyers in
Sweden. One can very well expect a rise in demand for hydrochar in the years to come,
as more biogas plants decide to augment biogas (and thereby biomethane) production
by adding hydrochar to their digesters. The LCC was extended to determine the effect
of the unit selling price of the surplus hydrochar [(5000–623) tons] on the NPV of
the ‘production and sale of 4377 tons of hydrochar’.
( )
(1 + i)1 (1 + i)2 (1 + i)20
PVsale, HC = Benefitsale, HC + + · · · + (13.6)
(1 + r)1 (1 + r)2 (1 + r)20

‘Benefit sale, HC ’ is the total benefit from selling 4377 tons of HC annually. This is
estimated in year-zero and increases from year to year at the rate of inflation.

NPVsale, HC = PVsale, HC − PVtotal production cost, surplus HC (13.7)

PVtotal production cost, surplus HC is the present value of the life-cycle cost of producing
4377 tons of HC solids per year, for 20 years. This is 7 times the value of PVHTC
calculated in Eq. 13.1, corresponding to 623 tons of HC solids per year. The unit
selling price of HC was set at 10,000 SEK/ton (the same as what Biogasbolaget would
have to pay if they decided to purchase the 623 tons they would need annually).
Sensitivity Analysis: As indicated earlier, inflation in Sweden has fluctuated over
time and has had an average value of 2%, as per the Swedish Statistics Central
Bureau, and at the time of writing, stands at 10%. The discount rate impacts the NPV
of business projects, and since that can also vary (quite like the inflation rate does),
a sensitivity analysis was deemed to be necessary, to bolster the LCC. In addition
to the values of ‘i’ and ‘r’, the unit purchase price of hydrochar (Scenario 2) which
was assumed to be 10,000 SEK/ton HC solids was varied, and the sensitivity of the
NPV to the variation was tested. Likewise, a sensitivity analysis of the unit selling
price of the surplus hydrochar (Scenario 1) was also carried out.
316 A. L. Kariis et al.

13.3 Results and Discussion

Some of the key results from this study have been presented hereunder.

13.3.1 Biogas and Methane Production

Figure 13.2 shows the results for daily biogas production, for a total period of 42 days.
On average, the daily production of biogas over these 6 weeks was 533 ml/g VS.
This was over 7.6 times more than the corresponding average value for the reference
case (without the addition of hydrochar in other words).
The reason why Fig. 13.2 starts off from day-27 was the detection of leakage
in the equipment, which took some time to be remedied. Once it was remedied,
the daily biogas production rose and remained at an average value of 533 ml/g VS.
While the problem was resolved for the reactor to which hydrochar was added, no
marked change in the biogas generation was recorded for the reference reactor. It is
seen from Fig. 13.2 that the yield dropped on day-33 before picking up again and
stabilising on day-41. The reasons for this behaviour are not very clear, as the pH
remained fairly stable throughout the experiment. Clearly, a 663% increase in biogas
production resulted thanks to the addition of hydrochar into the reactor.
Biogas can have varying compositions of carbon dioxide and methane, and from
an energetic point of view, it is the methane content which is of importance. The
reference case registered an average of 24.6% methane in the biogas, while addition
of hydrochar increased this share to an average of 60.9% in this study which was
equivalent to 367 ml/g VS. It is thus clear that the methane generation potential
is more than doubled by the addition of hydrochar to the substrate in the digester.

Fig. 13.2 Daily biogas production depicted in Fig. 13.1, between day-27 and day-68
13 Improving the Biogas Generation Potential from Organic Wastes Using … 317

There were some practical difficulties with availing of the gas chromatograph which
resulted in limiting the readings to 7 (within the said 9-day period towards the fag
end of the experiment).

13.3.2 Material Flow Analysis

Figure 13.3 presents a material flow analysis (MFA) diagram for Biogasbolaget with
an in-house HTC-facility. The substrate-mix (23.27 tons per day, with 11% total solids
content) flows into the digester along with deionised water (192.05 tons per day). On
the outlet side, there are two streams—190 tons per day of water and 10 tons per day
of dry solids. To maintain an HC concentration of 8 g/l in the digester, if the HTC-
reactor has a 70% efficiency of conversion of biomass to solid hydrochar, 2.46 tons
(per day) of digestate solids will have to be recirculated to the HTC-reactor. Biogas,
digestate as fertiliser, and the 190 tons per day of water are the outflows from the
system (the sale of surplus hydrochar which is not connected to the digestion process
within the firm has not been indicated as an outflow).
The fertiliser stream which leaves the system on the right has more carbon and
nutrients (nitrogen and phosphorus; as well as some essential micronutrients), when
compared to a similar stream from a digester which does not avail of hydrochar
addition. Evidently, this is also a saleable product for Biogasbolaget, in addition to
biogas (or refined biomethane) and surplus hydrochar. The fact that the digestate with
hydrochar is richer in carbon and nutrients also motivates the recirculation shown in
Fig. 13.3.

Fig. 13.3 Material flow diagram for the HTC-digester combination at Biogasbolaget
318 A. L. Kariis et al.

13.3.3 Life-Cycle Costing

Table 13.3 presents the PV and NPV values for Scenarios 1 and 2. Values corre-
sponding to rise in biomethane production of 17 and 53% are both represented
adjacent to each other.
If Biogasbolaget decides to invest in the HTC-facility yielding 5000 tons of HC
solids per year, and the captive consumption remains constant at 623 tons per year,
business sense would dictate finding a market for the surplus 4377 tons (which has a
total life-cycle cost associated with it). While the PV of 31 million SEK (Table 13.3)
corresponds to the 623 tons used in the digester. The PV of life-cycle costs associated
with the surplus would be 219 million SEK. Considering a unit sale price for the
surplus hydrochar (determined in and for year-zero), of 10,000 SEK per ton HC
solids, the PV of the total benefits over the 20-year period would equal 1479 million
SEK, giving a NPV for the ‘produce-and-sell-surplus-hydrochar’ part of the project
equal to 1260 million SEK. This NPV then would be added on to the NPVHTC
reported in Table 13.3, to obtain the total NPV of the investment in the HTC-facility,
intended to produce two saleable outputs.
In Fig. 13.4, the effect of variations in the discount rate (between 2 and 6%), and
the inflation rate (between 2 and 10%) has been depicted graphically, for Scenario
1. The NPVHTC in this case corresponds to the apportioned present value of the
life-cycle costs to the production of 623 tons of HC solids per year, and a rise in
biomethane production by 17%. In other words, the NPV of the ‘produce-and-sell-
surplus-hydrochar’ part of the project has not been included.
The NPV for Scenario 2 will depend on the purchase price of hydrochar and
there is an uncertainty surrounding that value as there is no market as such for
hydrochar. However, as mentioned earlier, one expects this to become a reality in the
years to come, as a circular bioeconomy entrenches itself around the world. With an
increase in this purchase price, the NPV will tend to decrease. As seen in Fig. 13.5,
as long as the purchase price is below 18,000 SEK/ton HC solids, the NPV will
be positive (suggesting that it would still not be unprofitable to purchase hydrochar

Table 13.3 PV and NPV values for Scenarios 1 and 2, (17 and 53% rise in the biomethane produc-
tion) [sale of surplus hydrochar in Scenario 1 is overlooked in this comparison, but becomes realistic
when there is a burgeoning market and demand for it]
Variables Value (million SEK) Equations used Scenarios
17% increase 53% increase
PVHTC 31 31 Equation 13.1 1
PVpurchase, HC 217 217 Equation 13.2 2
PVadditional benefit 394 1268 Equation 13.3 1 and 2
NPVHTC 363 1237 Equation 13.4 1
NPVHC 177 1051 Equation 13.5 2
NPVHTC − NPVHC 186 Scenario 1 economically more
attractive
13 Improving the Biogas Generation Potential from Organic Wastes Using … 319

Fig. 13.4 Variations in NPV with discount and inflation rates, for a 17% rise in biomethane
production for Scenario 1 (not considering any sale of surplus hydrochar)

from external suppliers—as, if and when they begin to exist—and use it to augment
biogas production).
One can now turn back to Scenario 1 and consider both the sale of surplus
hydrochar and the additional biomethane fuel. To exactly break even, the surplus
hydrochar would have to be sold at 1479 SEK/ton HC solids. This is the point at
which the value of NPVsale, HC (calculated using Eq. 13.7) drops to zero. At this
point, the total NPV, as seen in Fig. 13.6, is 363 million SEK (the same as NPVHTC
from Table 13.3). If there is no or less demand for hydrochar, and Biogasbolaget is
compelled to sell it at a lower price than the aforesaid breakeven price, NPVsale, HC

Fig. 13.5 The sensitivity of the NPV in Scenario 2 to the purchase price of hydrochar, for a 17%
rise in biomethane production
320 A. L. Kariis et al.

Fig. 13.6 The sensitivity of the NPV to the unit sale price of hydrochar

will be negative, and this will pull down the total NPV of the in-house HTC-facility
installation project. At a selling price of 10,000 SEK/ton, the total NPV stands at
1623 million SEK.
It must be mentioned, however, that most of the data obtained [19, 20] are studied
approximations made by experts in the business. There are very few HTC-facilities
in operation at the time of writing, and access to realistic data is difficult. This study
uncovers something which has the potential to attract interest among biogas plants in
the future. But, on date, no biogas plant purchases hydrochar from C-green. While it
would have been reasonable to envisage an HTC-facility at Biogasbolaget producing
precisely what the digester would need (623 tons of HC solids per year), the authors
have designed a scenario which is ideal, desirable and realistic, when viewed from the
perspective of a circular bioeconomy in which hydrochar would be a saleable product,
sought-after by biogas plants in Europe and beyond. The electricity tariff may not
increase at the general inflation rate considered in the analysis (10%); it is possible
that it may even drop slightly. The selling price of biomethane fuel (supplied to the
transport sector), determined at year-zero, is 17.5 SEK/Nm3 . This is also assumed
to increase at the general inflation rate, which may not be the case in reality.
Uncertainty surrounds the selling price of the surplus hydrochar in Scenario 1.
As gathered from Peter Axegård of C-Green [19], there is a demand for hydrochar at
combined heat and power plants, at the time of writing. If new demand at biogas plants
is uncovered, courtesy studies like this one, the value and thereby the selling price of
hydrochar will increase. But to meet the increasing demand, if more HTC-facilities
are set up, and supply tends to rise faster than demand (when all the facilities are
operating at 100% capacity), the price will drop down. The possibilities though are
difficult to predict. The sensitivity analysis conducted contributes a little to dealing
with this uncertainty.
13 Improving the Biogas Generation Potential from Organic Wastes Using … 321

13.4 Conclusions and Recommendations

The findings of this study, as well as those in Part I, augur well for a circular bioe-
conomy, like the one Sweden is keen on establishing in the years to come. By exten-
sion, it will motivate trials and experiments in other parts of the world, to verify,
confirm, and subsequently commercialise the utilisation of hydrochar in biogas
plants. Biogas plants like Biogasbolaget for instance, can diversify into the produc-
tion, use, and sale of hydrochar, and also deliver richer organic fertilisers to farmers
as one of the three saleable outputs, thereby contributing to phosphorus recovery.
The integration between the HTC-reactor and the anaerobic digester, supplying to
each other (as depicted in Fig. 13.4), uncovers environmental and techno-economic
advantages.
The LCC and the accompanying sensitivity analysis ‘speak out’ in support of
captive production of hydrochar by the HTC process, to be partly used in-house to
produce more biogas and thus sell more refined biomethane fuel, and largely sold to
potential users in the external marketplace (CHP plants, biogas plants, etc.)
Valorising what would otherwise be dealt with in non-profitable and undesirable
ways, to replace and substitute what is not desired in a circular bioeconomy in the
future, is indisputably a win–win situation, en route to the attainment of a clutch of
complementary sustainable development goals.

Acknowledgements The authors would like to thank everyone at C-Green, Biogasbolaget and
Karlstad University for their help and support with this study, at different points of time in the first
half of 2023.

References

1. Cavali, M., Junior, N. L., de Almeida Mohedano, R., Belli Filho, P., da Costa, R. H. R., & de
Castilhos Junior, A. B. (2022). Biochar and hydrochar in the context of anaerobic digestion for
a circular approach: An overview. Science of the Total Environment, 822, 153614. https://doi.
org/10.1016/j.scitotenv.2022.153614
2. European Commission. (n.d.). Food waste. Available at: https://food.ec.europa.eu/safety/food-
waste_en. Accessed on February 7, 2023.
3. Livsmedelsverket. (2022). Därför ska vi minska matsvinnet [Therefore, shall we reduce food
waste] Available at: https://www.livsmedelsverket.se/matvanor-halsa--miljo/maltider-i-vard-
skola-och-omsorg/matsvinn-i-storkok/handbok-for-minskat-matsvinn/darfor-ska-vi-minska-
matsvinnet. Accessed on February 7, 2023.
4. Chiappero, M., Norouzi, O., Hu, M., Demichelis, F., Berruti, F., Di Maria, F., Mašek, O., &
Fiore, S. (2020). Review of biochar role as additive in anaerobic digestion processes. Renewable
and Sustainable Energy Reviews, 131, 110037. https://doi.org/10.1016/j.rser.2020.110037
5. Merzari, F., Goldfarb, J., Andreottola, G., Mimmo, T., Volpe, M., & Fiori, L. (2020).
Hydrothermal carbonization as a strategy for sewage sludge management: Influence of process
withdrawal point on hydrochar properties. Energies, 13, 2890. https://doi.org/10.3390/en1311
2890
6. Assis, E. I. N. C., Gidudu, B., & Chirwa, E. M. N. (2022). Hydrothermal carbonisation of paper
sludge: Effect of process conditions on hydrochar fuel characteristics and energy recycling
322 A. L. Kariis et al.

efficiency. Journal of Cleaner Production, 373, 133775. https://doi.org/10.1016/j.jclepro.2022.


133775
7. Pecchi, M., & Baratieri, M. (2019). Coupling anaerobic digestion with gasification, pyrolysis
or hydrothermal carbonization: A review. Renewable and Sustainable Energy Reviews, 105,
462–475. https://doi.org/10.1016/j.rser.2019.02.003
8. Xu, J., Mustafa, A. M., Lin, H., Choe, U. Y., & Sheng, K. (2018). Effect of hydrochar on
anaerobic digestion of dead pig carcass after hydrothermal pretreatment. Waste Management,
78, 849–856. https://doi.org/10.1016/j.wasman.2018.07.003
9. Zhou, Y., Engler, N., Li, Y., & Nelles, M. (2020). The influence of hydrothermal operation
on the surface properties of kitchen waste-derived hydrochar: Biogas upgrading. Journal of
Cleaner Production, 259, 121020. https://doi.org/10.1016/j.jclepro.2020.121020
10. Ayodele, O. O., Adekunle, A. E., Adesina, A. O., Pourianejad, S., Zentner, A., & Dornack,
C. (2021). Stabilization of anaerobic co-digestion of biowaste using activated carbon of
coffee ground biomass. Bioresource Technology, 319, 124247. https://doi.org/10.1016/j.bio
rtech.2020.124247
11. Ma, J., Pan, J., Qiu, L., Wang, Q., & Zhang, Z. (2019). Biochar triggering multipath methano-
genesis and subdued propionic acid accumulation during semi-continuous anaerobic digestion.
Bioresource Technology, 293, 122026. https://doi.org/10.1016/j.biortech.2019.122026
12. Wang, R., Peng, P., Song, G., Zhao, Z., & Yin, Q. (2022). Effect of corn Stover hydrochar on
anaerobic digestion performance of its associated wastewater. Environmental Pollution, 315,
120430. https://doi.org/10.1016/j.envpol.2022.120430
13. Wu, B., Yang, Q., Yao, F., Chen, S., He, L., Hou, K., Pi, Z., Yin, H., Fu, J., Wang, D., & Li, X.
(2019). Evaluating the effect of biochar on mesophilic anaerobic digestion of waste activated
sludge and microbial diversity. Bioresource Technology, 294, 122235. https://doi.org/10.1016/
j.biortech.2019.122235
14. Ayodele, O. O., Adekunle, A. E., Alagbe, O. A., Anguruwa, G. T., Ademola, A. A., Odega,
C. A., & Dornack, C. (2022). Application of biomass-derived hydrochar in process stability
of anaerobic digestion. Bioresour. Technol. Rep., 17, 100903. https://doi.org/10.1016/j.biteb.
2021.100903
15. Xu, Q., Yang, G., Liu, X., Wong, J. W. C., & Zhao, J. (2023). Hydrochar mediated anaer-
obic digestion of bio-wastes: Advances, mechanisms and perspectives. Science of the Total
Environment, 884, 163829. https://doi.org/10.1016/j.scitotenv.2023.163829
16. Axegård, P., Baresel, C., Lazic, A., Bornold, N., Yang, J. -J., & Malovanyy, A. (2023). Framti-
dens slamhantering vid Roslagsvatten (Sludge handling in the future at Roslagsvattten). Ivl
Sven. Miljöinstitutet. Accessible at www.ivl.se
17. Schnürer, A., & Jarvis, Å. (2009). Microbiology handbook for biogas plants. Acces-
sible at: https://www.build-a-biogas-plant.com/PDF/Microbiology_handbook_for-Biogas_
plants. Published by Avfall Sverige and Svenskt Gasteknisk Center AB.
18. Venkatesh, G. (2019). Life-cycle costing: A primer (1st ed.). Bookboon. Available at: https://
bookboon.com/en/life-cycle-costing-ebook?mediaType=ebook
19. Axegård, P. (2023). C-Green. Personal Communication in March-May 2023.
20. Lagerblad, J. (2023). Biogasbolaget. Personal Communication in April 2023
21. Swedish Energy Agency (Energimyndigheten). (2023). Energiläget. Available at: https://www.
energimyndigheten.se/statistik/energilaget/?currentTab=1. Accessed on May 9, 2023.
22. Svenskt Gastekniskt Center AB. (2012). Basic data on biogas. Accessible at www.sgc.se/ckf
inder/userfiles/files/BasicDataonBiogas2012.pdf
23. Nordlöf, P. (2014). Summary of the Swedish debate on the discount rate. In Powerpoint presen-
tation on the 6th of November 2014 for Trafikverket (Swedish Transport Administration).
Accessed in May 2023 on www.trafikverket.se
Chapter 14
Heterogeneous Hydrochar-Based
Catalysts for Biodiesel Production

Muhammad Aliyu, Umer Rashid, Wan Azlina Wan Ab Karim Ghani,


Muhamad Amran bin Mohd Salleh, Balkis Hazmi, Ibrahim Garba Shitu,
and Ali Salisu

Abstract The potential of heterogeneous hydrochar-based catalysts as a desir-


able technology for biodiesel production is highlighted in this book chapter. These
hydrochar-based catalysts have renewable features, are cost-effective, and have effi-
cient catalytic properties, making them an excellent choice in the field due to their
high carbon content and unique physicochemical features. This chapter aims to
provide a comprehensive review of the feedstock of hydrochar from different biomass
sources including agricultural waste, food waste, and forestry waste. The hydrochar
production technologies and factors affecting the properties of hydrochar during
production were also discussed. The catalytic performance to produce biodiesel can
be considerably improved by the development of types of heterogeneous hydrochar-
based catalysts via introducing active species such as metal nanoparticles or acid- or
base-functional groups onto the hydrochar surface, and the parameters that influence

M. Aliyu · U. Rashid (B) · B. Hazmi


Institute of Nanoscience and Nanotechnology, Universiti Putra Malaysia (UPM), 43400 Serdang,
Selangor, Malaysia
e-mail: umer.rashid@upm.edu.my; dr.umer.rashid@gmail.com
M. Aliyu
Department of Material Engineering, Graduate School of Engineering, Kyushu Institute of
Technology, 1-1 Sensuicho, Tobata-ku, Kitakyushu, Fukuoka 804-8550, Japan
U. Rashid
Center of Excellence in Catalysis for Bioenergy and Renewable Chemical (CBRC), Faculty of
Science, Chulalongkorn University, Pathumwan, Bangkok 10330, Thailand
W. A. W. A. K. Ghani · M. A. M. Salleh
Department of Chemical and Environmental Engineering, Faculty of Engineering, Universiti
Putra Malaysia, 43400 Serdang, Selangor, Malaysia
I. G. Shitu
Department of Physics, Faculty of Natural and Applied Science, Sule Lamido University, Kafin
Hausa, Jigawa State, Nigeria
A. Salisu
Department of Chemistry, Faculty of Natural and Applied Science, Sule Lamido University, Kafin
Hausa, Jigawa State, Nigeria

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2024 323
G. Baskar et al. (eds.), Circular Bioeconomy Perspectives in Sustainable Bioenergy
Production, Energy, Environment, and Sustainability,
https://doi.org/10.1007/978-981-97-2523-6_14
324 M. Aliyu et al.

biodiesel yield and selectivity during biodiesel production utilizing heterogeneous


hydrochar-based catalysts were discussed. The multifaceted nature of hydrochar,
combined with its renewable origin and distinctive physicochemical properties,
positions it as a valuable resource for advancing sustainable and efficient catalytic
processes. Finally, this chapter advances knowledge on heterogeneous hydrochar-
based catalysts and encourages further research in this promising area while outlining
current research and future directions.

Keywords Hydrochar · Biomass · Hydrothermal carbonization · Heterogeneous


catalysts · Biodiesel production

14.1 Introduction

Energy has always been recognized as a fundamental driver for sustaining economic
growth in any nation, with coal, natural gas, and crude oil serving as major sources to
meet this energy demand [1]. However, the rapid expansion of industrialization and
urbanization is projected to lead to a 33% increase in global energy consumption by
2035 [2]. The depletion of non-renewable resources and their contribution to envi-
ronmental issues like global warming have fueled the desire for alternative energy
conservation and utilization approaches [3]. The world’s focus has shifted toward
sustainable and eco-friendly strategies for producing renewable energy sources. The
objective is to develop non-toxic and biodegradable alternatives that offer environ-
mentally friendly and economically viable solutions [4]. One highly preferred renew-
able energy source in the present era is biodiesel, with biodiesel standing out as a
non-toxic, biodegradable, and renewable energy option [5].
Biodiesel plays a crucial role in addressing the pressing need for sustainable and
renewable energy sources worldwide. Biodiesel is produced from renewable sources
including soybean oil, canola oil, palm oil, or waste cooking oil [6].
Biodiesel can be used in diesel engines in replacement of petroleum-based diesel
fuel without any modifications. It has various advantages over conventional diesel
fuel, including decreased carbon dioxide, particulate matter, and sulfur dioxide
emissions [7]. Biodiesel is also less hazardous and combustible than diesel fuel,
which makes it easier to handle and transport. The biodiesel production process
of biodiesel involves a chemical reaction called transesterification or esterification,
which converts vegetable oil or animal fat into biodiesel [8]. During this process, the
vegetable oil or animal fat is mixed with alcohol and a catalyst, which breaks down
the oil molecules and produces biodiesel glycerin and water [9]. With the increasing
demand for biodiesel, researchers are actively pursuing innovative low-cost cata-
lysts that can optimize the efficiency and sustainability of the biodiesel production
process.
In recent years, heterogeneous hydrochar-based catalysts have emerged as a
promising solution within this field of research [10]. These catalysts have garnered
14 Heterogeneous Hydrochar-Based Catalysts for Biodiesel Production 325

significant attention due to their inherent advantages, which make them highly attrac-
tive for biodiesel production. One of the key advantages of heterogeneous hydrochar-
based catalysts is their renewable nature [11]. The use of hydrochar as a precursor
for catalyst synthesis capitalizes on the abundant carbon content present in biomass,
effectively utilizing renewable resources [12]. This aligns perfectly with the goal
of sustainable energy production and contributes to reducing the reliance on non-
renewable fossil fuels [13]. Additionally, the cost-effectiveness of heterogeneous
hydrochar-based catalysts enhances their appeal for biodiesel production. The avail-
ability of biomass feedstock and the relatively simple synthesis process of hydrochar-
based catalysts result in lower production costs compared to conventional catalysts
[14].
The term “hydrochar” encompasses a carbon-rich substance that is derived
through the process of hydrothermal carbonization (HTC), often known as “wet
carbonization,” from various biomass sources ranging from agricultural residues
and food waste to forestry by-products, contributing to the utilization of diverse
organic materials [15]. HTC includes heating biomass at high pressures and temper-
atures in the presence of water, forming hydrochar, a carbon-rich solid material [16].
Hydrochar usually contains a high percentage of carbon range between 50 and 90%,
as well as small amounts of hydrogen, oxygen, and nitrogen. It might also have
ash, which contains minerals and trace elements from the original feedstock [16].
Hydrochar has several possible applications, including its use as a fuel source for
power generation, as a soil supplement in agriculture and as a catalyst. Hydrochar
serves as an exceptional precursor for catalyst synthesis, primarily due to its abundant
carbon content and distinctive physicochemical properties [2].
Hydrochar has a variety of properties and a porous structure with a large surface
area, making it an attractive choice for catalyst production [17]. The chemical compo-
sition of hydrochar can be modified through the introduction of active species onto its
surface [18]. This procedure involves the addition of materials such as metal nanopar-
ticles or acid–base functional groups to the hydrochar matrix, which improves its
catalytic capabilities for biodiesel production [3]. Therefore, the catalyst becomes
capable of enabling and accelerating the conversion of triglycerides or fatty free
acids (FFA) into biodiesel, leading to a significant improvement in the overall effi-
ciency of the biodiesel production process [19]. Furthermore, the incorporation of
active species onto the hydrochar matrix also contributes to the enhanced stability
of the catalyst, enabling it to maintain its catalytic activity over prolonged reaction
periods. Another significant advantage of this approach is its potential to mitigate
the environmental impact associated with biodiesel production [5].
This book chapter delves into the improvements of heterogeneous hydrochar-
based catalysts for biodiesel production. This study investigates multiple compo-
nents, such as their activation, modification, catalytic performance on biodiesel
production, and ongoing research efforts aimed at enhancing their performance and
usability in real-world biodiesel production scenarios. We hope to provide significant
insights and contribute to the improvement of biodiesel production methods using
hydrochar-based catalysts by investigating the most recent results and breakthroughs
in this field.
326 M. Aliyu et al.

14.2 Hydrochar

14.2.1 Feedstock, Composition, and Properties of Hydrochar

Biomass serves as the primary feedstock for hydrochar production, encompassing


various organic materials derived from agricultural waste, food waste, and forestry
waste as shown in Fig. 14.1 [20]. These diverse biomass feedstocks contribute to
the composition of hydrochar, which is a carbon-rich material obtained through
hydrothermal carbonization [21]. The composition of hydrochar depends on the
specific biomass source used and the process parameters employed during its produc-
tion. It typically contains carbon, hydrogen, oxygen, and small amounts of other
elements, such as nitrogen and sulfur [22]. The elemental composition and chemical
properties of hydrochar can vary, influencing its reactivity, surface area, and porosity.
Therefore, the selection of biomass feedstock and control of process parameters play
a crucial role in determining the composition and properties of hydrochar [23].

14.2.1.1 Agricultural Waste

Agricultural waste is the organic by-products that remain after the cultivation and
harvesting of crops [24]. These residues encompass a wide range of plant materials
such as stems, leaves, husks, shells, and straws, which are typically left behind
in the fields after the primary crop has been collected [25]. With the large-scale
agricultural practices prevalent in many regions, agricultural residues have become
a significant and easily accessible source of biomass [26]. Their abundance and
availability make them attractive feedstock options for various applications, including
hydrochar production. Rice straws, corn stalks, sugarcane bagasse, and wheat straw
are among the most utilized agricultural residues in hydrochar production processes
[27].
The potential of these agricultural wastes is maximized by using them as feed-
stock hydrochar production, converting what would otherwise be considered waste
into a lucrative resource [28]. The use of agricultural residues for hydrochar produc-
tion has various advantages. For instance, it provides a sustainable and renewable
alternative to existing fossil fuel-based resources, helping to establish a more ecolog-
ically friendly energy sector [29]. Using agricultural wastes reduces the demand for
non-renewable resources, promotes resource efficiency, and reduces the environ-
mental effect of fossil fuel consumption [30]. Furthermore, using agricultural waste
for producing hydrochar helps the efficient use of agricultural waste, and instead of
being dumped or left in the fields, these wastes are turned into valuable goods with
potential applications in a variety of industries. This minimizes waste and provides
potential for economic growth and resource efficiency in the agricultural sector.
14 Heterogeneous Hydrochar-Based Catalysts for Biodiesel Production 327

Fig. 14.1 The distribution of feedstock for hydrochar production

14.2.1.2 Food Waste

Food waste encompasses various organic materials that are no longer consumed or
used, originating from households, restaurants, food processing industries, and other
sources [31]. It poses a significant environmental challenge due to the substantial
quantities of food waste that end up in landfills, resulting in the generation of green-
house gas emissions, particularly methane (CH4 ). To address this issue, utilizing
food waste as a feedstock for hydrochar production offers an eco-friendly solution
to convert this waste into a valuable resource [32]. Food waste consists of a wide
range of materials, including fruit and vegetable peels, expired or spoiled food, and
discarded leftovers. These materials, if not properly managed, would contribute to
the increasing volume of waste and its associated environmental impacts [33]. On
the other hand, redirecting food waste to hydrochar production can successfully turn
its organic content into a carbon-rich material via the process of HTC. Using food
waste as a feedstock for hydrochar production has various advantages [34]. Firstly,
328 M. Aliyu et al.

it helps to address the environmental concerns associated with food waste disposal
in landfills. The use of food waste for hydrochar production reduces the emission
of methane, a potent greenhouse gas, lowering the carbon footprint and mitigating
climate change. Moreover, the conversion of food waste into hydrochar presents
an opportunity to generate a valuable carbonaceous material that can be utilized in
various applications [35].
Hydrochar derived from food waste possesses a high carbon content and unique
physicochemical properties, making it suitable for use as a precursor for catalyst
synthesis or as a soil amendment to improve soil quality and fertility [36]. Further-
more, the use of food waste for hydrochar production adds to the circular economy
concept by closing the loop in the food system [37]. To sum up, food waste is a
serious environmental issue, but it also has considerable promise as a feedstock
for hydrochar production. Not only may the environmental implications of food
waste disposal be reduced by converting it into hydrochar. The use of food waste for
hydrochar production adheres to circular economy principles and adds to resource
efficiency and environmental sustainability.

14.2.1.3 Forestry Waste

Forestry waste involves a wide range of products generated during forestry activities,
including branches, bark, sawdust, wood chips, and other woody waste. In traditional
forestry practices, these materials are often considered waste or by-products, and their
disposal may cause cost and ecological challenges [38]. However, through the utiliza-
tion of forestry residue as feedstocks for hydrochar production, these materials can
be converted into valuable resources with a wide range of applications [39]. Due to
activities such as tree harvesting, logging, and wood processing, the forestry industry
generates a substantial amount of wood waste. Previously, these wastes were either
left to disintegrate in the forest or disposed of by burning or landfilling, which can
add to air pollution, soil degradation, and misused resources. These materials can be
effectively exploited by channeling forestry waste toward hydrochar production [15].
The use of forestry wastes as feedstock for hydrochar production has various advan-
tages. For instance, it contributes to addressing environmental challenges linked with
forestry residue disposal [40]. The release of harmful pollutants from the burning or
decomposition of forestry waste is decreased by turning these wastes into hydrochar,
contributing to improved air quality and lower greenhouse gas emissions. Further-
more, hydrochar made from forestry residue can be used in a variety of applications
[41].

14.2.1.4 Composition and Properties of Hydrochar

The composition of hydrochar obtained from different biomass sources can vary
depending on the feedstock used and the specific hydrothermal carbonization process
employed [16]. Generally, the hydrochar is predominantly composed of carbon,
14 Heterogeneous Hydrochar-Based Catalysts for Biodiesel Production 329

with content of volatile matter, moisture, and other organic compounds compared
to the original biomass [2]. Hydrochar exhibits unique physicochemical properties,
including high carbon content, low ash content, and stability. Its characteristics make
it an excellent precursor for catalyst synthesis and various other applications, such
as soil amendment, carbon sequestration, and energy storage [17]. The composi-
tion and properties of hydrochar are influenced by several factors, including the
type of biomass feedstock, HTC process parameters, and post-treatment techniques
[42]. The composition of hydrochar varies depending on the biomass feedstock used.
Different biomass sources, such as agricultural waste, food waste, and forestry waste,
have distinct chemical compositions, including varying levels of carbon, hydrogen,
oxygen, nitrogen, and other trace elements [43]. These elements contribute to the
overall composition and can impact the properties and potential applications of
hydrochar. The carbon content of hydrochar is a crucial property that determines its
energy content and stability. Hydrochar typically has a high carbon content, ranging
from 50 to 90%, which makes it a potential source of renewable energy [44]. The
carbon concentration of hydrochar also determines its calorific value, making it a
desirable option for energy production [3].
The surface area and pore structure of hydrochar are also essential properties.
Hydrochar possesses a porous structure with a high surface area, which provides
numerous sites for adsorption and chemical reactions. The pore size distribution,
including micropores, mesopores, and macropores, affects the reactivity of the
hydrochar [19]. These properties make hydrochar suitable for applications such as
adsorbents, soil amendments, and catalyst support. The presence of functional groups
on the surface of hydrochar is another significant aspect of its composition. These
functional groups, including hydroxyl, carboxyl, and phenolic groups, contribute
to the reactivity and surface chemistry of hydrochar. They can serve as active sites
for various applications or be used to attach different functional groups, leading to
increased interactions with various reactants in various environmental applications
[5]. The thermal stability of hydrochar is also an important property. Hydrochar
exhibits good thermal stability, which enables its use in high-temperature processes
without significant degradation [45]. This property is advantageous for applications
such as energy conversion, where hydrochar can be utilized as a solid fuel or feed-
stock for thermochemical processes. Table 14.1 shows the feedstocks, elemental
compositions, and properties of hydrochar.
Other properties of hydrochar include its pH, ash content, and density. The pH
of hydrochar can influence its behavior in soil applications and its interaction with
other substances [46]. The ash content represents the inorganic components present
in hydrochar, which can impact its combustion characteristics and the release of
nutrients when used as a soil amendment. The density of hydrochar influences its
handling and transportation, with lower densities being preferable for easier handling
[11]. Finally, the composition and qualities of hydrochar are determined by the
biomass feedstock, HTC process parameters, and post-treatment procedures [47].
The current study focuses on modifying the composition and features of hydrochar
to optimize its potential as a sustainable and adaptable material in the pursuit of a
greener and more sustainable future.
330

Table 14.1 The feedstocks, elemental compositions, and properties of hydrochar


Biomass Feedstocks C (%) O (%) H (%) N (%) S (%) Ash content BET surface Pore volume Average pore Yield Refs.
(%) area (m2 /g) (cm2 /g) diameter (nm) (%)
Agricultural Walnut shell 58.32 11.85 3.17 2.53 1.22 1.49 18.89 0.09 5.37 54.34 [5]
waste Millet stalk 53.23 12.15 6.73 4.03 1.74 1.34 10.34 0.08 7.24 51.67 [48]
Wheat straw 57.76 11.45 6.21 1.43 0.87 1.67 20.43 0.05 5.43 66.34 [49]
Rice straw 62.32 13.45 4.87 2.03 1.02 1.09 11.89 0.07 4.67 59.43 [50]
Food waste Banana peels 66.13 11.35 5.23 4.03 1.94 1.84 11.34 0.07 5.24 52.43 [51]
Watermelon 56.43 14.35 7.43 3.23 1.34 2.34 13.34 0.03 8.34 51.23 [52]
peels
Grapefruit 67.1 16.5 5.2 2.8 1.8 2.4 18.23 0.08 7.98 53.67 [13]
peels
Cassava 59.83 17.15 9.43 5.33 1.74 3.24 16.54 0.06 9.14 49.67 [53]
rhizome
Forestry Bamboo 70.43 11.55 4.13 2.23 0.84 1.14 33.34 0.05 7.04 57.34 [54]
waste sawdust
Wood 68.22 15.14 5.53 2.73 2.04 2.04 26.34 0.02 4.14 55.76 [55]
sawdust
Pinus 62.31 13.12 7.36 3.83 1.71 1.94 17.14 0.07 5.30 58.54 [56]
Caribbean
Spent liquor 65.71 12.24 6.41 2.83 2.11 1.88 25.61 0.09 9.04 52.43 [57]
M. Aliyu et al.
14 Heterogeneous Hydrochar-Based Catalysts for Biodiesel Production 331

14.2.2 Hydrochar Production

Hydrochar production encompasses various techniques, including conventional


hydrothermal carbonization (HTC) and microwave-assisted hydrothermal carboniza-
tion (MA-HTC), which offer efficient and sustainable pathways for converting
biomass into a valuable carbon-rich material.

14.2.2.1 Conventional Hydrothermal Carbonization (HTC)

Hydrothermal carbonization (HTC) has recently gained recognition as a practical and


established conventional thermochemical technique for producing char [58]. In this
method, biomass materials are transformed into a carbon-rich solid substance referred
to as hydrochar. This transformation occurs within an aqueous environment, typically
at temperatures ranging from 175 to 250 °C, with specific timeframes and pressures
that vary based on the type of feedstock biomass used [46]. One of the potential
HTC’s significant benefits is its compatibility with wet biomass, which eliminates
the requirement to dry the biomass prior to carbonization. HTC achieves this by
processing moist biomass in hot, pressurized water, leading to significant energy
and cost savings that would otherwise be required for pre-drying the biomass before
undergoing thermochemical conversion [58]. The reaction conditions in conventional
HTC, including temperature, pressure, reaction time, and biomass-to-water ratio, are
crucial in influencing the properties and yield of hydrochar [45]. Higher temperatures
and longer reaction times generally lead to increased carbonization and the formation
of a more carbon-rich hydrochar. Adjusting the pressure and biomass-to-water ratio
can also impact the chemical reactions and the final characteristics of the hydrochar
produced [59]. As a process of producing hydrochar, conventional HTC has several
advantages. A significant advantage is its adaptability to a wide variety of biomass
feedstocks. This adaptability enables the use of a wide range of organic materials,
boosting the efficient use of biomass resources and reducing waste [60].

14.2.2.2 Microwave-Assisted Hydrothermal Carbonization (MA-HTC)

Microwave-assisted hydrothermal carbonization (MA-HTC) is an emerging tech-


nique that combines the principles of conventional HTC with the efficient heating
capabilities of microwaves [61]. In MA-HTC, biomass is mixed with water and
subjected to microwave irradiation, which rapidly heats the reaction mixture and
accelerates the carbonization process [62]. The localized and rapid heating provided
by microwaves promotes efficient biomass conversion and reduces reaction times
compared to conventional HTC. MA-HTC offers advantages such as energy effi-
ciency, shorter processing times, and improved control over reaction conditions [33].
Both HTC and MA-HTC present efficient and sustainable methods for hydrochar
production. They offer opportunities to convert various types of biomasses, including
332 M. Aliyu et al.

agricultural residues, food waste, and forestry by-products, into valuable carbon-
rich material. These techniques contribute to the utilization of biomass resources,
reducing waste and promoting a circular economy. Furthermore, they have the
potential for integration with existing waste management systems and may be
adapted for large-scale production, making them viable solutions for sustainable
biomass. However, to get the appropriate hydrochar qualities, temperature, biomass-
to-water ratio, and reaction time must be properly optimized. Table 14.2 illustrates
a comparison of microwave-assisted HTC against conventional HTC.

Table 14.2 A comparison between micro-assisted HTC and conventional HTC


MA-HTC Conventional HTC
Heating In MA-HTC, microwave radiation is HTC relies on external heating sources
method used to rapidly heat the reaction like electric heaters, ovens, or
mixture. It allows for precise and autoclaves to raise the temperature of
efficient heating of the reactants the reaction vessel. It is a slower and
less precise heating method
Reaction Microwave heating is generally faster, Conventional HTC usually requires
speed and reactions can be completed more longer reaction times due to the slower
quickly compared to conventional heating process
HTC. It is ideal for achieving rapid
temperature ramps
Temperature Microwave heating offers better Temperature control in conventional
control temperature control and uniformity HTC can be less precise, potentially
within the reaction vessel, resulting in leading to temperature gradients within
consistent reaction conditions the vessel
Energy Microwave heating is energy efficient Conventional heating methods may be
efficiency as it directly heats the reactants, less energy-efficient because they heat
reducing energy wastage the surroundings as well as the
reactants
Scalability Scaling up MA-HTC can be Conventional HTC is more easily
challenging and expensive due to the scalable and adaptable to larger
need for specialized equipment and reactors and industrial processes
potential issues with uniform heating
at larger scales
Reactor Reactor design for MA-HTC is more Conventional HTC can use standard
design complex and requires materials reactor vessels, making it simpler in
compatible with microwave radiation terms of equipment
Safety Safety precautions are necessary when Conventional HTC is considered safer
working with microwaves, including but still requires standard safety
the risk of thermal runaway reactions measures
14 Heterogeneous Hydrochar-Based Catalysts for Biodiesel Production 333

14.2.2.3 The Reaction Mechanism Occurs During Hydrothermal


Carbonization

The mechanism of carbonization and the synthesis route of hydrochar via the
hydrothermal carbonization (HTC) process of biomass involves a series of complex
thermochemical reactions [15]. These processes are as follows: Depolymerization
is the stage of HTC, and complex organic polymers within the biomass, such as
hemicellulose, cellulose, and lignin, undergo depolymerization. This means that
large and intricate molecules break down into smaller, simpler compounds [41].
Hydrolysis is a fundamental reaction in HTC. It involves the cleavage of chem-
ical bonds within the biomass due to the presence of water under high-temperature
and high-pressure conditions. This process produces simpler organic compounds,
including sugars and organic acids [18]. Dehydration reactions occur as the temper-
ature within the HTC reactor rises. These reactions lead to the removal of water
molecules from the biomass. The removal of water makes an environment conducive
to carbonization. Decarboxylation reactions are particularly important in the HTC
process. During this step, certain organic compounds within the biomass, such as
carboxylic acids, release carbon dioxide (CO2 ) when exposed to high temperatures.
This process reduces the oxygen content in the biomass and contributes to the forma-
tion of carbon-rich structures [19]. Direct carbonization is a key transformation step
in HTC. It involves the conversion of the remaining organic components or lignin
into carbon-rich solids. Complex organic molecules break down, and the resulting
carbonaceous materials resemble charcoal. Isomerization reactions involve the rear-
rangement of atoms within molecules, resulting in the formation of molecules with
the same molecular formula but different structural arrangements. Rearrangement
reactions involve the reshuffling of atoms and chemical groups within the biomass
components. These rearrangements lead to the formation of more stable carbon struc-
tures [5]. Polymerization is the process through which smaller organic molecules
formed during the previous reactions reassemble into larger, more complex carbon
structures. This step contributes to the development of the hydrochar’s solid structure.
Surface diffusion involves the movement of molecules on the surface of the devel-
oping hydrochar. This can lead to the rearrangement of atoms and functional groups,
further enhancing the hydrochar’s properties [63]. Figure 14.2 depicts the mecha-
nism of carbonization and the synthesis route of hydrochar via the HTC process of
biomass.

14.2.2.4 Parametric Factors Affecting Hydrochar Production

Hydrochar production is a complex process influenced by various parametric factors


that significantly impact the properties and yield of hydrochar. Proper understanding
and control of these factors are essential for optimizing the production process and
obtaining hydrochar with desired characteristics [64]. The following are some of the
key parametric factors that play a crucial role in hydrochar production. Figure 14.3
depicts the parametric factors influencing hydrochar production.
334 M. Aliyu et al.

Fig. 14.2 The mechanism of carbonization and the synthesis route of hydrochar via HTC process
of biomass

Temperatures

The reaction temperature during hydrochar production is a crucial parameter that


significantly influences the quality and properties of the resulting hydrochar [62].
Higher temperatures generally accelerate the reaction kinetics and promote the
carbonization process, leading to increased carbon content and enhanced thermal
stability of the hydrochar [33]. The higher temperatures promote the formation of
carbon-rich structures and accelerate the breakdown of complex organic molecules.
However, excessively high temperatures can have negative consequences, such as
the loss of vital organic components through volatilization and the beginning of
undesired side reactions that can compromise the quality of the hydrochar [65]. The
optimal temperature range for hydrochar production varies depending on the specific
feedstock being utilized and the desired properties of the hydrochar [2]. Different
biomass feedstocks have different thermal degradation characteristics and require
14 Heterogeneous Hydrochar-Based Catalysts for Biodiesel Production 335

Fig. 14.3 Illustration of the parametric factors affecting hydrochar production

specific temperature conditions to maximize carbonization while minimizing unde-


sirable reactions. Furthermore, the desired properties of the hydrochar, such as its
carbon content, porosity, and stability, also influence the optimal temperature range
[18]. To determine the optimal temperature for hydrochar production, a thorough
understanding of the feedstock characteristics and the desired hydrochar properties
is necessary. Experimental studies and optimization techniques are often employed
to identify the temperature range that yields hydrochar with the desired properties
[3].

Biomass-to-Liquid Ratio

The biomass-to-liquid ratio is a critical parameter that influences the HTC process
and ultimately affects the properties of the resulting hydrochar [19]. This ratio refers
to the proportion of biomass material relative to the liquid medium, typically water,
used in the process. The biomass-to-liquid ratio plays a crucial role in determining the
extent of carbonization and the characteristics of hydrochar [5]. A higher biomass-
to-liquid ratio generally leads to increased carbonization of the biomass and higher
carbon content in the hydrochar. This is because a higher proportion of biomass
material facilitates a more extensive reaction with the liquid medium, promoting the
conversion of organic compounds into carbon-rich structures [63]. Therefore, the
hydrochar produced at higher biomass-to-liquid ratios tends to have higher carbon
content, increased stability, and enhanced energy content.
However, it is important to strike a balance when selecting the biomass-to-liquid
ratio. The presence of a large amount of biomass can hinder the efficient transfer of
336 M. Aliyu et al.

heat, causing irregular temperature distribution and potentially resulting in incom-


plete carbonization. Incomplete carbonization can lead to lower carbon content,
reduced stability, and compromised properties of the hydrochar [58]. Therefore, it is
essential to optimize the biomass-to-liquid ratio to ensure effective carbonization and
the production of high-quality hydrochar [45]. The best ratio may change depending
on the biomass feedstock, the desired hydrochar features, and the processing condi-
tions. Experimental studies and optimization approaches are widely utilized to deter-
mine a suitable biomass-to-liquid ratio to produce hydrochar with the required
characteristics and maximize the overall efficiency of the HTC process.

Residence Time

The residence time is a key parameter that controls the duration of the HTC process
and has a considerable impact on the degree of carbonization as well as the qualities of
the resulting hydrochar, which is an essential variable in the production of hydrochar
[4]. A longer residence time provides more time for the biomass to undergo the
complex chemical reactions involved in carbonization. This extended duration allows
for a more thorough conversion of organic compounds into carbon-rich structures,
resulting in higher carbon content and increased stability of the hydrochar [66]. It
enables the formation of a well-developed and interconnected carbon matrix, which
enhances the overall quality and desired properties of hydrochar.
However, it is crucial to establish a balance when determining the time of resi-
dence. Excessive residence times can result in over-carbonization, whereby the
hydrochar turns overly carbonized and loses some of its desirable properties such as
hydrochar reactivity, increased brittleness, reduced porosity, and limiting its appli-
cation [43]. The optimal residence time is determined by several variables, including
the biomass feedstock, the desired properties of the hydrochar, and the process condi-
tions [4]. The suitable residence time for hydrochar production is determined through
experimental research and process optimization approaches. It is feasible to achieve
the necessary level of carbonization, stability, and other critical characteristics of the
hydrochar by carefully controlling the residence time.

Pressure

The pressure used during the hydrochar production process is an important parameter
that has a large impact on the reaction kinetics and the properties of the produced
hydrochar [63]. They improve the interaction of biomass and water by increasing
the solubility of biomass components in the liquid phase. This increased solu-
bility enhances reactant diffusion and accelerates reaction rates, resulting in faster
carbonization. Therefore, hydrochar generated at greater pressures has a higher
degree of carbonization and a denser structure [45]. However, the pressure levels
14 Heterogeneous Hydrochar-Based Catalysts for Biodiesel Production 337

employed in hydrochar production must be carefully considered. Excessive pres-


sure can result in energy-intensive activities and higher expenses for the equipment
needed to handle high-pressure conditions [60].
Finding the perfect pressure range will be critical for balancing the desired
carbonization level with the practical efficiency of the production process. The pres-
sure conditions required for optimal hydrochar production are determined by several
factors, including the kind of biomass feedstock, desired hydrochar qualities, and
overall process design [47]. It is feasible to achieve the appropriate level of carboniza-
tion and modify the properties of the hydrochar to match specific requirements by
carefully managing the pressure applied during hydrochar production. This enables
the production of hydrochar with desirable properties such as better stability, higher
energy content, improved porosity, and higher carbon yield [42]. These parametric
factors interact with each other, and their optimal values depend on the specific
feedstock, desired hydrochar properties, and process conditions.

14.2.3 Activation and Modification Hydrochar


as a Promising Catalysts

Hydrochar activations and modifications have sparked substantial interest in the


field of biodiesel production due to their ability to transform hydrochar into effective
catalysts. Hydrochar produced by the HTC of biomass is unique by its high carbon
content and porous structure, making it a desirable option for catalyst production.
Figure 14.4 depicts the activation and modification of hydrochar.

Fig. 14.4 Illustration of the activation and modification of hydrochar


338 M. Aliyu et al.

14.2.3.1 Activation of Hydrochar

Activation is a crucial post-treatment process employed to enhance the properties of


hydrochar obtained through HTC. The objective of activation is to improve the surface
area, porosity, and catalytic activity of hydrochar [67]. Activation methods include
physical activation (e.g., using carbon dioxide or steam) and chemical activation (e.g.,
using phosphoric acid or potassium hydroxide). These processes produce a highly
porous structure with a huge surface area known as activated carbon or activated
hydrochar, which can serve as a perfect surface for catalytic reactions [63].

Physical Activation

Physical activation processes involve the use of carbon dioxide or steam to produce
pores within the hydrochar structure [15]. Carbon dioxide activation, also known
as gas activation, involves subjecting hydrochar to a stream of carbon dioxide at
high temperatures, resulting in the formation of micropores and mesopores within
the hydrochar [40]. On the other hand, the physical activation process involves
heating the hydrochar for several hours in an inert atmosphere, such as nitrogen
or argon, or in the presence of steam or carbon dioxide [17]. This process reduces
the volatile components and generates micropores and mesopores on the surface of
the hydrochar, leading to a highly porous structure with a large surface area and
high catalytic activity. The product of this process is known as activated carbon or
activated hydrochar [44].
Activating hydrochar with steam is a physical process that is both more secure
and beneficial to the environment than chemical activation. After the HTC of the
biomass, an endothermic process of steam activation is performed [68]. This process
takes place at temperatures ranging from 200 to 800 °C, with a period of 30–60 min
and a heating rate of 5–10 °C/min. CO2 activation is an endothermic process in which
carbon dioxide is used as an activation gas rather than steam due to its less reactive
nature and increased oxidation reaction [3]. To make activated HC, CO2 is activated
at temperatures ranging from 200 to 900 °C. CO2 activation of hydrochar improves
porosity and overall surface area with oxygen-containing functional groups, which
boosts the catalytic activity of the hydrochar. The principle of gasification and the
removal of volatile compounds govern the physical activation process, leading to the
formation of a larger pore structure [19].

Chemical Activation

Chemical activation is the process of introducing functional groups that generate


porosity in hydrochar by using chemicals such as phosphoric acid or potassium
hydroxide. Phosphoric acid activation is a popular process that includes activating
hydrochar with phosphoric acid and then carbonizing it at high temperatures [69].
The phosphoric acid reacts with the hydrochar, causing pores to form and a porous
14 Heterogeneous Hydrochar-Based Catalysts for Biodiesel Production 339

structure to develop. Similarly, the activation of hydrochar with potassium hydroxide,


which includes the carbonaceous material, forms pores and increases surface area
[66]. Chemical activation methods allow for control over the pore size and distribution
of functional groups in the hydrochar, allowing for customization of its catalytic
capabilities [5].
Hydrochar can also be chemically activated by impregnating it with different
chemicals, and lower activation temperatures, increased carbon yield, greater surface
area, formation of diverse functional groups, and well-developed micro- and meso-
porous structure are the multiple benefits of the chemical activation of hydrochar
[70]. Activated hydrochar has various applications for catalysts and other applica-
tions, including water treatment, air pollution management, and energy storage [71].
The activation processes, whether physical or chemical, result in the formation of
activated carbon, which possesses a highly porous structure with a large surface area.
The increased surface area provides abundant active sites for reactant adsorption and
interaction, leading to enhanced catalytic performance [72]. The presence of pores
and functional groups within the activated carbon facilitates the diffusion of reactants
and promotes desirable catalytic reactions [68]. In general, activation methods are
critical in converting hydrochar into activated carbon with increased surface area and
reactivity. The resulting activated hydrochar serves as an appropriate framework for
catalytic processes, providing improved catalytic performance and efficiency [73].

14.2.3.2 Modification of Hydrochar

Modification of hydrochar involves the introduction of active species onto the surface
of hydrochar, which can significantly enhance its catalytic performance for various
applications, including biodiesel production [15]. Three common types of active
species used in hydrochar modification are acid- or base-functional groups and metal
nanoparticles.

Modification of Hydrochar with Acid

Modification of hydrochar with acid involves introducing acid-functional groups onto


the hydrochar surface. This modification process enhances the catalytic properties of
hydrochar and expands its potential applications in various fields, including biodiesel
production [43]. Acid modification of hydrochar can be achieved by treating it with
acid solutions, such as sulfuric acid (H2 SO4 ) or hydrochloric acid (HCl) [2]. The
acid reacts with the hydrochar surface, leading to the formation of acid-functional
groups, such as carboxylic (–COOH) or phenolic (–OH) groups. These acid groups
act as catalysts in heterogeneous acid-catalyzed reactions, promoting the necessary
chemical transformations in biodiesel production [17].
The modification process involves several steps, including the preparation of the
acid solution, the immersion or impregnation of hydrochar in the acid solution, and
subsequent washing and drying to remove excess acid [43]. The concentration of
340 M. Aliyu et al.

the acid solution and the duration of the modification process can be optimized
to achieve the desired level of acid-functional groups on the hydrochar surface.
These acid groups provide active sites for the esterification and transesterification
reactions involved in the conversion of feedstock into biodiesel. They facilitate the
necessary chemical reactions, improve reaction rates, and enhance the selectivity
of biodiesel production [3]. The modified hydrochar with acid-functional groups
offers several advantages. It can improve the efficiency of biodiesel production by
promoting the conversion of triglycerides or fatty free acids (FFA) into biodiesel
and reducing the formation of unwanted by-products [19]. Additionally, the acid
modification of hydrochar can enhance the stability of the catalyst, ensuring its long-
term performance during repeated use. The acidic modification of hydrochar enables
the development of more sustainable and effective catalysts for biodiesel production
[69].

Modification of Hydrochar with Base

Modification of hydrochar with base involves the introduction of base-functional


groups onto the hydrochar surface. This modification process enhances the catalytic
properties of hydrochar and broadens its applications in various fields, including
biodiesel production [45]. Base modification of hydrochar can be achieved by treating
it with base solutions, such as sodium hydroxide (NaOH) or potassium hydroxide
(KOH). The base reacts with the hydrochar surface, leading to the formation of base-
functional groups, such as hydroxyl (–OH) or amine (–NH2 ) groups [74]. These
base groups act as catalysts in heterogeneous base-catalyzed reactions, facilitating
the necessary chemical transformations in biodiesel production. The modification
process typically involves the preparation of the base solution, immersion, or impreg-
nation of hydrochar in the base solution, and subsequent washing and drying to
remove the excess base [46]. The concentration of the base solution and the dura-
tion of the modification process can be optimized to achieve the desired level of
base-functional groups on the hydrochar surface [11].
The incorporation of base-functional groups onto the hydrochar surface enhances
its catalytic activity in biodiesel production. These base groups provide active sites for
the transesterification reactions involved in converting feedstock into biodiesel [50].
They promote the conversion of triglycerides or FFA into biodiesel and contribute
to increased reaction rates and improved selectivity. Modified hydrochar with base-
functional groups offers several advantages. It enhances the efficiency of biodiesel
production by facilitating the desired chemical reactions and minimizing unwanted
side reactions [67]. Additionally, the base modification of hydrochar can improve the
stability and reusability of the catalyst, ensuring consistent performance over multiple
reaction cycles, to further improve the catalytic performance of base-modified
hydrochar in biodiesel production [20].
14 Heterogeneous Hydrochar-Based Catalysts for Biodiesel Production 341

Modification of Hydrochar with Metal Nanoparticles

The introduction of metal species into the hydrochar surface is essential for the modi-
fication of hydrochar with metal nanoparticles. This modification procedure improves
hydrochar’s catalytic features and improves its potential applications, including
biodiesel production [23]. The modification process begins with the preparation of
metal precursor solutions, which contain metal ions of interest, such as nickel (Ni),
palladium (Pd), or iron (Fe). These metal precursors are typically in the form of
salts, such as nickel nitrate or palladium chloride. The hydrochar is then immersed
or impregnated in the metal precursor solution, allowing the metal ions to adsorb
onto its surface [27]. After impregnation, the hydrochar is subjected to a reduc-
tion step, typically using a reducing agent such as sodium borohydride (NaBH4 ), to
convert the metal ions into metal nanoparticles or subjected to calcination for certain
temperatures. The reduction reaction results in the deposition of well-dispersed metal
nanoparticles on the hydrochar surface [30].
The incorporation of metal nanoparticles onto the hydrochar surface enhances its
catalytic activity in biodiesel production. These metal nanoparticles act as cata-
lysts, promoting the necessary chemical reactions involved in the conversion of
feedstock into biodiesel [75]. They provide active sites for the transesterification
or esterification reactions, facilitating the breakdown of triglycerides or FFA and
improving the efficiency and selectivity of the biodiesel production process. The
metal nanoparticles provide increased surface area and improve the accessibility of
reactants to catalytic sites, enhancing the overall catalytic performance [76]. Addi-
tionally, metal nanoparticles can exhibit specific catalytic properties, such as high
activity and selectivity, enabling more efficient and effective biodiesel production
[77], exploring various metal precursors, optimization strategies for nanoparticle size
and distribution, and the influence of different metals on the catalytic performance
of heterogeneous hydrochar-based catalysts.

14.2.4 Catalytic Performance of Heterogeneous


Hydrochar-Based Catalysts in Biodiesel Production

The catalytic performance of heterogeneous hydrochar-based catalysts in biodiesel


production has gained significant attention due to their potential to improve the
efficiency and sustainability of the process [78]. Here are some key aspects of their
catalytic performance in biodiesel production:

14.2.4.1 Transesterification/Esterification Activity

Heterogeneous hydrochar-based catalysts have shown remarkable catalytic activity in


the transesterification and esterification reactions, which are crucial steps in biodiesel
342 M. Aliyu et al.

production [2]. Transesterification involves the conversion of triglycerides or FFA


present in the feedstock, such as vegetable oils or animal fats with alcohol, typically
methanol or ethanol, to produce biodiesel or fatty acid methyl esters (FAME) or fatty
acid ethyl esters (FAEE) [79]. On the other hand, esterification refers to the reaction
between free fatty acids (FFA) and alcohol to form FAME. The surface functional
groups of hydrochar materials, such as hydroxyl (–OH) or carboxyl (–COOH) and
sulfonate (–SO3 H) groups, play a crucial role in facilitating the transesterification
and esterification reaction [80]. These functional groups can act as acid or base sites,
effectively catalyzing the conversion of triglycerides into FAMEs. The hydroxyl
groups can serve as acid sites by protonating the alcohol, while the carboxyl groups
can act as base sites by deprotonating the carbonyl group of the triglycerides or FFA
[81].
The presence of these functional groups on the hydrochar’s surface enhances
the catalytic activity by promoting the activation of reactants and facilitating the
desired reaction pathways. The acid–base catalysis promotes the breaking of the
ester bonds in the triglycerides and the subsequent esterification of the fatty acids
with the alcohol to form FAME [82]. Furthermore, the hydrochar’s surface functional
groups also interact with the reactants through adsorption and surface reactions. This
interaction enhances the contact between the reactants and the catalyst, improving the
reaction kinetics and increasing the conversion efficiency [83]. The catalytic activity
of hydrochar-based catalysts in transesterification or esterification reactions is closely
related to the concentration and accessibility of the surface functional groups. Factors
such as the hydrochar preparation conditions, precursor material, and post-treatment
methods can influence the density and distribution of these functional groups on the
hydrochar surface [84].
In general, hydroxyl and carboxyl groups on the surface of heterogeneous
hydrochar-based catalysts play an important role in accelerating the transesterifi-
cation reaction in biodiesel production [2]. Its acid–base properties promote the
activation of reactants and provide the essential catalytic sites for the conversion of
triglycerides and FFA into FAME. The catalytic activity of hydrochar-based catalysts
can be increased by modifying the hydrochar composition and surface properties,
leading to increased efficiency and yield in biodiesel production processes [85].
Figure 14.5 depicts an illustration of biodiesel production through esterification/
transesterification of triglycerides or free fatty acids.

14.2.4.2 Biodiesel Yield and Selectivity

The catalytic activity of heterogeneous hydrochar-based catalysts in biodiesel


production is influenced by several key factors, including the catalyst composi-
tion, surface area, porosity, and surface functional groups [86]. Each of these
factors contributes to the overall performance and effectiveness of the catalysts in
promoting the transesterification reaction and influencing the biodiesel yield and
selectivity [6]. To begin, the catalyst composition is critical in determining catalytic
14 Heterogeneous Hydrochar-Based Catalysts for Biodiesel Production 343

Fig. 14.5 Illustration of production of biodiesel via esterification/transesterification of triglyceride


or free fatty acids

activity. Hydrochar-based catalysts can be made from a variety of precursor mate-


rials, including biomass which is the main source [9]. The catalyst composition is
influenced by the precursor material used and the HTC treatment conditions. The
carbon concentration, elemental composition, and surface functional groups of the
hydrochar can all have an impact on its catalytic efficiency [87]. Catalysts with larger
areas of surface and porosity have more active sites for transesterification and ester-
ification reactions to take place. The increased availability of active sites improves
the interaction between the catalyst and the reactants, leading to higher biodiesel
yield [9].
Lastly, the presence of certain surface functional groups on hydrochar-based cata-
lysts, such as hydroxyl (–OH) or carboxyl (–COOH), influences reaction kinetics and
selectivity toward FAME production. These functional groups may serve as acid or
basic sites, allowing the transesterification of triglycerides and alcohol [8]. The acidic
or basic character of the functional groups facilitates reactant activation, allowing
for efficient esterification and enhancing selectivity toward FAME. To improve the
catalytic activity of hydrochar-based catalysts in biodiesel production, the composi-
tion, surface area, porosity, and functional groups of the catalyst need to be optimized
[88]. By adjusting these factors, the amount of biodiesel that can be made can be
increased by increasing the number of active sites, making the mass transfer as
efficient as possible, and promoting the desired reaction pathways [89].
Optimization of reaction conditions is crucial for achieving high biodiesel yields
and desired selectivity in biodiesel production using hydrochar-based catalysts.
Several key parameters, including catalyst loading, reaction temperature, alcohol-
to-oil molar ratio, and reaction time, play significant roles in determining the effi-
ciency of the process. Figure 14.6 depicts an illustration of the variables influencing
biodiesel yield.
344 M. Aliyu et al.

Fig. 14.6 Schematic of the factors affecting biodiesel yield

Catalyst Loading

The dosage of hydrochar-based catalyst used in the transesterification and esterifica-


tion reaction is referred to as catalyst loading. It is an important parameter to consider
achieving a balance between attaining the desired catalytic activity and reducing the
costs associated with excessive catalyst usage [90]. Catalyst loading plays a signifi-
cant role in determining the number of active sites available for the transesterification
or esterification reaction to occur. Increasing the catalyst loading generally leads to
a higher density of active sites on the catalyst’s surface, facilitating the conversion of
triglycerides into biodiesel and glycerol. This can result in improved reaction kinetics
and higher biodiesel yields [89].
Optimizing the catalyst loading is critical for maximizing biodiesel production and
ensuring optimal usage of the hydrochar-based catalyst. It requires finding an appro-
priate balance between providing enough active sites for the reaction and reducing
catalyst costs [91]. Biodiesel production can improve the efficiency of the transester-
ification and esterification processes by carefully determining the optimum catalyst
loading and optimizing the conversion of triglycerides or FFA into biodiesel while
keeping catalyst consumption at an economically sustainable amount. The suitable
catalyst loading depends on a variety of factors, including the type of hydrochar-
based catalyst utilized, the properties of the feedstock, and the required reaction
conditions [92]. Typically, experimental investigations and optimization approaches
are used to find the best catalyst loading for a certain biodiesel production system.
These studies entail evaluating catalytic performance at various catalyst loadings,
analyzing biodiesel yields, considering financial implications, and comparing entire
process efficiency.
14 Heterogeneous Hydrochar-Based Catalysts for Biodiesel Production 345

Reaction Temperature

The reaction temperature is a key factor that significantly impacts the efficiency and
selectivity of the transesterification and esterification reactions in biodiesel produc-
tion. It plays a crucial role in controlling the reaction kinetics and influencing the
distribution of products formed [93]. Typically, these reactions are carried out at
elevated temperatures within a range of 50–90 °C. Higher reaction temperatures
result in higher reaction rates due to higher molecular activity and collision frequency
of the reactant molecules. This faster reaction kinetics can be favorable in terms of
reducing reaction time and increasing the production of biodiesel. However, it is
vital to note that extremely high temperatures can also enhance the occurrence of
side reactions, resulting in the formation of unwanted by-products [94].
Optimizing the reaction temperature is essential for achieving the desired reaction
kinetics and product selectivity. Biodiesel production may optimize the conversion
of triglycerides or FFA into biodiesel while reducing the formation of undesirable by-
products by carefully managing the temperature [95]. Finding the optimum temper-
ature for the reaction requires considering several factors, including the nature of
the catalyst, the type of feedstock utilized, and the desired qualities of the biodiesel.
To determine the appropriate reaction temperature for biodiesel production via eval-
uating the reaction performance and product distribution at various temperatures,
consider both the biodiesel yield and the quality of the generated product [80]. It also
helps to ensure the economic viability and environmental sustainability of biodiesel
production by ensuring that the process operates under the best possible conditions
in terms of reaction kinetics and product selectivity.

Alcohol-to-Oil Molar Ratio

The alcohol-to-oil molar ratio is a crucial parameter that significantly influences


the transesterification and esterification reaction in biodiesel production. It refers to
the ratio of alcohol, typically methanol or ethanol, to triglyceride or FFA feedstock,
which is commonly derived from vegetable oils or animal fats [96]. This ratio plays a
vital role in determining the extent of transesterification or esterification and directly
affects the conversion efficiency and biodiesel yield. A higher alcohol-to-oil molar
ratio commonly improves the completion of the transesterification and esterification
reactions. The excess alcohol ensures that there are enough alcohol molecules avail-
able to react with the triglycerides or FFA, resulting in a higher conversion of triglyc-
erides or FFA into biodiesel. This increased conversion efficiency is advantageous
since it increases biodiesel yield, resulting in more efficient feedstock utilization
[97].
However, it is essential to realize that excessive alcohol use might have negative
effects. Initially, increasing the alcohol-to-oil molar ratio can result in increased costs
due to the reaction requiring more alcohol. Alcohol is frequently more expensive
than the feedstock itself, and utilizing too much can have a substantial impact on the
overall economics of biodiesel production [81]. Furthermore, a high alcohol amount
346 M. Aliyu et al.

in the reaction mixture can complicate the separation and purification processes
during the biodiesel product’s downstream processing. This can result in higher
energy usage and additional purification processes, increasing the entire process’s
complexity and cost. Finding the optimum alcohol-to-oil molar ratio is important for
balancing conversion efficiency, and yield [86]. Considerations include the specific
feedstock characteristics, the catalyst used, and the intended quality criteria for the
biodiesel yield. This optimization contributes to the overall economic feasibility and
sustainability of biodiesel production by providing effective resource utilization and
limiting the environmental impact associated with excessive alcohol consumption
and purification challenges.

Reaction Time

The reaction time in the production of biodiesel refers to the period that the transes-
terification or esterification reaction is allowed to proceed. It is an essential variable
that has a direct impact on the level of conversion and the overall yield of biodiesel.
The response time is essential in obtaining the target biodiesel production efficiency
while limiting unwanted side effects [98]. Prolonged reaction times regularly improve
triglyceride or FFA conversion into biodiesel by allowing more time for the trans-
esterification or esterification reaction to occur. This expanded duration allows for
better exploitation of the available reactants, perhaps leading to higher biodiesel
yield. On the other hand, the longer reaction times could raise the possibility of side
reactions such as hydrolysis and saponification. These side reactions might result in
the formation of undesired compounds, lowering the quality of the biodiesel [99].
Optimizing the reaction time is vital for generating high biodiesel yields while
reducing unwanted reactions. The optimum reaction time is determined by taking
numerous aspects into account, including the specific hydrochar-based catalyst
employed, the properties of the feedstock, and the desired quality criteria for
the biodiesel product. Experimentation and process optimization approaches are
frequently used to determine the best reaction time for a certain biodiesel production
process. Efforts to optimize the reaction time not only led to increased biodiesel
yields but also added to the economic viability and environmental sustainability of
biodiesel production.
Optimizing the reaction time, as well as other reaction variables like catalyst
loading, reaction temperature, and alcohol-to-oil molar ratio, aims to improve the
efficiency and selectivity of biodiesel production [100]. Biodiesel production can
enhance mass transfer efficiency, maximize the usage of active sites on the catalyst,
and achieve an appropriate acid–base balance by carefully adjusting these parame-
ters. This enhancement increases biodiesel production, selectivity for high-quality
FAME, and overall process economics and sustainability. Table 14.3 summarizes the
activation, modification, and parametric parameters influencing biodiesel yield.
Table 14.3 The activation, modification, and parametric factors affecting biodiesel yield
Hydrochar Activation/ Biodiesel Type of reaction Catalyst Temperature/ Reaction Alcohol-oil-molar Biodiesel Refs.
feedstock Modification feedstock loading time (°C) time ratio yield (%)
(wt.%) (min)
Musa K2 CO3 /CaO Jatropha Esterification 7 65 60 12:1 98.38 [101]
Champa oil
peduncle
waste
Coconut KOH Safflower Transesterification 10 70 120 19:1 99.00 [102]
endocarp oil
Orange H2 SO4 Oleic acid Esterification 3 80 180 20:1 92.80 [103]
peel
Ziziphus H2 SO4 Oleic acid Esterification 10 80 240 20:1 91.00 [104]
Mauritania
L
Corncob H2 SO4 Palm fatty Esterification 5 70 300 20:1 92.00 [105]
residue acid
distillate
Mesocarp K2 CO3 / Waste Transesterification 4 70 120 12:1 95.36 [98]
fibers Cu(NO3 )2 cooking
14 Heterogeneous Hydrochar-Based Catalysts for Biodiesel Production

oil
Water CH3 C6 H4 SO2 Oleic acid Esterification 3 75 90 15:1 95.56 [106]
hyacinth
Empty fruit K2 CO3 / Waste Esterification–transesterification 5 70 120 12:1 97.1 [96]
bunch Cu(NO3 )2 cooking
oil
347
348 M. Aliyu et al.

14.2.4.3 Catalyst Stability and Regeneration

Catalyst stability and regeneration are essential considerations in the practical appli-
cation of hydrochar-based catalysts in biodiesel production. The carbonaceous nature
of hydrochar provides inherent stability, allowing it to withstand the harsh reac-
tion conditions involved in the transesterification and esterification process [82].
The robust carbon matrix of hydrochar ensures its structural integrity and maintains
catalytic activity even at high temperatures and in the presence of active sites. The
carbon-based structure of hydrochar exhibits excellent thermal stability, preventing
catalyst degradation and minimizing catalyst deactivation mechanisms such as coke
formation or sintering of active sites. This stability allows hydrochar-based catalysts
to maintain their catalytic activity over prolonged reaction times, ensuring consis-
tent performance and reliable biodiesel production [107]. Table 14.4 summarizes
the comparison of hydrochar-based catalysts with conventional catalysts in terms of
biodiesel yields, productivity, and recyclability.
Furthermore, the reusability of heterogeneous hydrochar-based catalysts is advan-
tageous for their practical application. Hydrochar-based catalysts can be easily regen-
erated and reused, reducing the need for frequent catalyst replacement. Regeneration
methods typically involve cleaning or rejuvenating the catalyst to restore its initial
activity [108]. Techniques such as washing with solvents, thermal treatment, or chem-
ical reactivation can effectively remove contaminants or reactant residues and restore
the catalyst’s surface properties. The reusability of hydrochar-based catalysts not only
reduces overall process costs but contributes to the economic viability and sustain-
ability of biodiesel production. The negative environmental effect of the process is
decreased waste formation and increased catalyst usage [87].
It is important to note that the stability and reusability of hydrochar-based cata-
lysts can depend on various factors, including the catalyst composition, preparation
method, reaction conditions, and feedstock characteristics [7]. Optimization of these
parameters is crucial to maximize catalyst stability and ensure efficient regeneration,

Table 14.4 The comparison of hydrochar-based catalysts with conventional catalysts in terms of
biodiesel yields, productivity, and recyclability
Heterogeneous hydrochar-based Conventional (homogeneous) catalysts:
catalysts
Biodiesel Hydrochar-based catalysts often Yields can vary among different
yield offer high yields in biodiesel homogeneous catalysts; some may offer
production, contributing to efficient high yields, while others may not be as
processes effective
Productivity They can enhance productivity due Productivity depends on the specific
to their porous structure and active catalyst used; it can range from highly
sites, promoting faster reactions efficient to less effective
Recyclability Hydrochar-based catalysts are The recyclability of homogeneous
generally recyclable, reducing costs catalysts varies widely; some may
and waste in the long term degrade over time, potentially increasing
expenses and waste
14 Heterogeneous Hydrochar-Based Catalysts for Biodiesel Production 349

enabling extended catalyst lifetimes and reducing the overall operating costs. In a
nutshell, the carbonaceous characteristics of hydrochar-based catalysts permit them
to withstand strong reaction conditions in biodiesel production [9]. The reusability
of hydrochar-based catalysts reduces the need for frequent replacement and helps
the process’s economic feasibility. Hydrochar-based catalysts can provide long-
lasting catalytic activity by optimizing catalyst stability and regeneration techniques,
ensuring constant and sustainable biodiesel production.

14.2.4.4 Reaction Mechanisms

The production of biodiesel involves the conversion of triglycerides and FFA, which
are commonly found in vegetable oils or animal fats, into biodiesel and glycerol or
water through transesterification and esterification reactions. This process is typically
carried out in the presence of a catalyst to enhance the reaction rate and efficiency
[8]. Figure 14.7 depicts a schematic of the reaction mechanism of the active site on
the surface hydrochar-based catalysts.
In the adsorption step, the triglycerides or FFA and alcohol molecules are adsorbed
onto the surface of the heterogeneous hydrochar-based catalyst. Hydrochar possesses
a porous structure that provides a large surface area with abundant adsorption sites.
This porous nature facilitates the interaction between the reactants and the catalyst,
promoting their adsorption and subsequent reaction [88]. Once the reactants are
adsorbed, the transesterification and esterification reactions take place. The alcohol
molecules, such as methanol or ethanol, react with the triglycerides bound to the
catalyst’s surface. This results in the formation of fatty acid alkyl esters (biodiesel)
and glycerol or water as a by-product. The hydrochar-based catalyst plays a crucial
role by providing active sites or functional groups on its surface, such as hydroxyl
(–OH) or carboxyl (–COOH) groups [90]. These active sites act as acid-based cata-
lysts or base-based catalysts, facilitating the breaking ester bonds in the triglyceride
molecules and forming ester bonds in the biodiesel molecules.
After the transesterification and esterification reactions occur, the biodiesel and
glycerol molecules need to be desorbed from the catalyst’s surface. The porous
structure of the hydrochar-based catalyst aids in the desorption process, allowing
the products to be released from the catalyst [85]. After the reaction, the biodiesel
and glycerol must be separated. Separation can be performed in a variety of ways
due to their various physical and chemical properties, such as boiling points and
densities. Distillation and centrifugation are two methods widely used to separate
biodiesel from glycerol. It should be noted that the composition and properties of the
heterogeneous hydrochar-based catalyst utilized can have an impact on the reaction
mechanism. Different catalysts may have different active sites or functions, leading
to differences in the reaction route and kinetics [109].
350 M. Aliyu et al.

Fig. 14.7 Schematic of the reaction mechanism of the active site on the surface hydrochar-based
catalysts

14.3 Present Research Status and Future Research Targets

Research in hydrochar production is currently focused on utilizing various biomass


feedstocks, such as agricultural waste, food waste, and forestry waste, to produce
carbon-rich hydrochar through hydrothermal carbonization (HTC) and microwave-
assisted hydrothermal carbonization (MA-HTC) processes. The composition and
properties of hydrochar are influenced by the specific feedstock used and the process
parameters employed. Key factors affecting the quality and yield of hydrochar include
temperature, biomass-to-liquid ratio, residence time, and pressure. Optimization of
14 Heterogeneous Hydrochar-Based Catalysts for Biodiesel Production 351

these parameters is being studied to obtain hydrochar with desirable characteristics,


including high carbon content, stability, and porosity. The research aims to optimize
the hydrochar production process and explore its potential application as a catalyst for
biodiesel production. The current study focuses on enhancing the catalytic activity of
hydrochar through activation and modification methods. Physical and chemical acti-
vation processes are explored to increase the surface area and porosity of hydrochar,
while modification involves introducing active species or metal nanoparticles onto
its surface.
In future, research in the field should concentrate on many different areas.
Initially, there is a need to investigate and evaluate various biomass feedstocks for
hydrochar production in terms of availability, sustainability, and required hydrochar
qualities. This includes investigating new and unconventional biomass sources.
Process optimization is another important target, involving the fine-tuning of param-
eters such as temperature, biomass-to-liquid ratio, residence time, and pressure in
the HTC process. The goal is to enhance carbonization efficiency, increase carbon
content, and optimize the properties of hydrochar. Advanced characterization tech-
niques, including surface analysis, should be further developed to gain a deeper under-
standing of hydrochar’s composition, structure, and properties, as well as its interac-
tions with other substances. Additionally, there is a need to explore and expand the
potential applications of hydrochar in various fields as a part of catalyst support such
as soil amendment, carbon sequestration, and wastewater treatment. This involves
investigating its performance in different scenarios and optimizing its properties
for specific applications. Comprehensive life cycle assessments and environmental
impact studies are essential to evaluate the environmental footprint and sustain-
ability of hydrochar production processes, including assessing its potential benefits
in terms of reducing greenhouse gas emissions, managing waste, and improving
resource efficiency. To bridge the gap between laboratory-scale research and large-
scale production, efforts should be directed toward developing scalable and cost-
effective hydrochar production technologies and overcoming technical, economic,
and logistical challenges. Furthermore, the optimization of activation and modifi-
cation processes for hydrochar to enhance catalytic activity is important, including
fine-tuning physical and chemical activation methods and exploring novel strategies
for more efficient and sustainable catalyst production. Understanding the structure,
morphology, and surface chemistry of activated and modified hydrochar-based cata-
lysts through advanced analytical techniques will contribute to optimizing catalytic
performance and designing more effective catalysts. The stability and regenerability
of hydrochar-based catalysts should also be studied to develop long-lasting catalysts
with minimal performance decline over multiple reaction cycles, ensuring economic
viability and sustainability. Process optimization and scale-up studies are needed to
optimize reaction conditions and assess the feasibility of hydrochar-based catalysts
in large-scale biodiesel production, aiming for high yields, improved selectivity, and
energy-efficient processes.
352 M. Aliyu et al.

14.4 Conclusions

In conclusion, the choice of biomass feedstock and process parameters signifi-


cantly influence the composition and properties of hydrochar produced through
hydrothermal carbonization, while hydrochar activations and modifications hold
promise for transforming hydrochar into effective catalysts for biodiesel produc-
tion. The utilization of agricultural waste, food waste, and forestry waste as feed-
stock offers a sustainable solution for waste management and resource utiliza-
tion. Optimizing temperature, biomass-to-liquid ratio, residence time, and pres-
sure is crucial for obtaining hydrochar with desired characteristics. Hydrothermal
carbonization and microwave-assisted hydrothermal carbonization provide efficient
and environmentally friendly methods for hydrochar production, contributing to
biomass conversion and waste reduction. On the other hand, physical and chem-
ical activation methods, as well as modifications with acid/base-functional groups
or metal nanoparticles, enhance the catalytic performance of hydrochar. Heteroge-
neous hydrochar-based catalysts exhibit remarkable activity in biodiesel production
reactions, with catalyst composition, surface area, porosity, and functional groups
influencing the overall performance. The stability and reusability of hydrochar-based
catalysts further contribute to their long-term effectiveness and cost-effectiveness.
Ultimately, continued research and optimization of hydrochar activation, modifica-
tion, and catalytic performance are necessary for advancing the field and facilitating
the commercialization of hydrochar-based catalysts in the biodiesel industry.

References

1. Rattanaya, T., Kongjan, P., Cheewasedtham, C., Bunyakan, C., Yuso, P., Cheirsilp, B., &
Jariyaboon, R. (2022). Application of palm oil mill waste to enhance biogas upgrading and
hornwort cultivation. Journal of Environmental Management, 309, 114678. https://doi.org/
10.1016/j.jenvman.2022.114678
2. Alfredo Quevedo-Amador, R., Elizabeth Reynel-Avila, H., Ileana Mendoza-Castillo, D.,
Badawi, M., & Bonilla-Petriciolet, A. (2022). Functionalized hydrochar-based catalysts
for biodiesel production via oil transesterification: optimum preparation conditions and
performance assessment. Fuel, 312, 122731. https://doi.org/10.1016/j.fuel.2021.122731
3. Munir, M. T., Ul Saqib, N., Li, B., & Naqvi, M. (2023). Food waste hydrochar: an alternate
clean fuel for steel industry. Fuel, 346, 128395. https://doi.org/10.1016/j.fuel.2023.128395
4. Liu, F., Yu, R., Ji, X., & Guo, M. (2018). Hydrothermal carbonization of holocellulose
into hydrochar: Structural, chemical characteristics, and combustion behavior. Bioresource
Technology, 263, 508–516. https://doi.org/10.1016/j.biortech.2018.05.019
5. Başakçılardan Kabakcı, S., & Baran, S. S. (2019). Hydrothermal carbonization of various
lignocellulosics: Fuel characteristics of hydrochars and surface characteristics of acti-
vated hydrochars. Waste Management, 100, 259–268. https://doi.org/10.1016/j.wasman.2019.
09.021
6. Yaashikaa, P. R., Kumar, P. S., & Karishma, S. (2022). Bio-derived catalysts for produc-
tion of biodiesel: a review on feedstock, oil extraction methodologies, reactors and lifecycle
assessment of biodiesel. Fuel, 316, 123379. https://doi.org/10.1016/j.fuel.2022.123379
14 Heterogeneous Hydrochar-Based Catalysts for Biodiesel Production 353

7. Fonseca, J. M., Spessato, L., Cazetta, A. L., da Silva, C., & Almeida, V. D. C. (2022).
Sulfonated carbon: synthesis, properties and production of biodiesel. Chemical Engineering
and Processing-Process Intensification, 170, 108668.
8. Ravikumar, R., Kiran, K., Hebbar, G. S., Naresh, H., & Panigrahi, S. P. (2022). An overview
of nano-catalysts in biodiesel production. Journal of Mines, Metals and Fuels. https://doi.org/
10.18311/jmmf/2022/31998
9. Adhikesavan, C., Ganesh, D., & Charles Augustin, V. (2022). Effect of quality of waste
cooking oil on the properties of biodiesel, engine performance and emissions. Cleaner
Chemical Engineering, 4, 100070. https://doi.org/10.1016/j.clce.2022.100070
10. Sztancs, G., Juhasz, L., Nagy, B. J., Nemeth, A., Selim, A., Andre, A., Toth, A. J., Mizsey,
P., & Fozer, D. (2020). Co-hydrothermal gasification of Chlorella vulgaris and hydrochar: The
effects of waste-to-solid biofuel production and blending concentration on biogas generation.
Bioresource Technology, 302, 122793. https://doi.org/10.1016/j.biortech.2020.122793
11. Nzediegwu, C., Naeth, M. A., & Chang, S. X. (2021). Carbonization temperature and feedstock
type interactively affect chemical, fuel, and surface properties of hydrochars. Bioresource
Technology, 330, 124976. https://doi.org/10.1016/j.biortech.2021.124976
12. Safari, F., Javani, N., & Yumurtaci, Z. (2018). Hydrogen production via supercritical water
gasification of almond shell over algal and agricultural hydrochars as catalysts. International
Journal of Hydrogen Energy, 43, 1071–1080. https://doi.org/10.1016/j.ijhydene.2017.05.102
13. Inkoua, S., Li, C., Zhang, L., Li, X., Zhang, S., Xiang, J., Hu, S., Wang, Y., & Hu, X. (2023).
Hydrothermal carbonization of grapefruit peels in swine manure-derived liquid: Property of
resulting hydrochars and activated carbons. Fuel, 351, 128889. https://doi.org/10.1016/j.fuel.
2023.128889
14. Marrakchi, F., Sohail Toor, S., Haaning Nielsen, A., Helmer Pedersen, T., & Aistrup
Rosendahl, L. (2023). Bio-crude oils production from wheat stem under subcritical water
conditions and batch adsorption of post-hydrothermal liquefaction aqueous phase onto acti-
vated hydrochars. Chemical Engineering Journal, 452, 139293. https://doi.org/10.1016/j.cej.
2022.139293
15. Luo, X., Huang, Z., Lin, J., Li, X., Qiu, J., Liu, J., & Mao, X. (2020). Hydrothermal carboniza-
tion of sewage sludge and in-situ preparation of hydrochar/MgAl-layered double hydroxides
composites for adsorption of Pb(II). Journal of Cleaner Production, 258, 120991. https://doi.
org/10.1016/j.jclepro.2020.120991
16. Pereira, G. R., Lopes, R. P., Wang, W., Guimarães, T., Teixeira, R. R., & Astruc, D.
(2022). Triazole-functionalized hydrochar-stabilized Pd nanocatalyst for ullmann coupling.
Chemosphere, 308, 136250. https://doi.org/10.1016/j.chemosphere.2022.136250
17. Zhuang, X., Liu, J., & Ma, L. (2022). Facile synthesis of hydrochar-supported catalysts from
glucose and its catalytic activity towards the production of functional amines. Green Energy
and Environment. https://doi.org/10.1016/j.gee.2022.01.012
18. Vega, M. F., Florentino-Madiedo, L., Díaz-Faes, E., & Barriocanal, C. (2021). Influence of
feedwater pH on the CO2 reactivity of hydrochars: Co-carbonisation with a bituminous coal.
Renewable Energy, 170, 824–831. https://doi.org/10.1016/j.renene.2021.01.100
19. Zhuang, X., Song, Y., Zhan, H., Bi, X. T., Yin, X., & Wu, C. (2020). Pyrolytic conversion of
biowaste-derived hydrochar: Decomposition mechanism of specific components. Fuel, 266,
117106. https://doi.org/10.1016/j.fuel.2020.117106
20. Bona, D., Scrinzi, D., Tonon, G., Ventura, M., Nardin, T., Zottele, F., Andreis, D., Andreottola,
G., Fiori, L., & Silvestri, S. (2022). Hydrochar and hydrochar co-compost from OFMSW
digestate for soil application: 2. Agro-environmental properties. Journal of Environmental
Management, 312, 114894. https://doi.org/10.1016/j.jenvman.2022.114894
21. Scrinzi, D., Bona, D., Denaro, A., Silvestri, S., Andreottola, G., & Fiori, L. (2022). Hydrochar
and hydrochar co-compost from OFMSW digestate for soil application: 1. Production and
chemical characterization. Journal of Environmental Management, 309, 114688. https://doi.
org/10.1016/j.jenvman.2022.114688
22. Wortmann, M., Keil, W., Brockhagen, B., Biedinger, J., Westphal, M., Weinberger, C., Diestel-
horst, E., Hachmann, W., Zhao, Y., Tiemann, M., Reiss, G., Hüsgen, B., Schmidt, C., Sattler,
354 M. Aliyu et al.

K., & Frese, N. (2022). Pyrolysis of sucrose-derived hydrochar. Journal of Analytical and
Applied Pyrolysis, 161, 105404. https://doi.org/10.1016/j.jaap.2021.105404
23. Wang, Z., Zhai, Y., Wang, T., Wang, B., Peng, C., & Li, C. (2020). Pelletizing of hydrochar
biofuels with organic binders. Fuel, 280, 118659. https://doi.org/10.1016/j.fuel.2020.118659
24. Magdziarz, A., Wilk, M., & Wądrzyk, M. (2020). Pyrolysis of hydrochar derived from
biomass—Experimental investigation. Fuel, 267, 117246. https://doi.org/10.1016/j.fuel.
2020.117246
25. Shao, Y., Long, Y., Wang, H., Liu, D., Shen, D., & Chen, T. (2019). Hydrochar derived from
green waste by microwave hydrothermal carbonization. Renewable Energy, 135, 1327–1334.
https://doi.org/10.1016/j.renene.2018.09.041
26. Zhang, Y. N., Guo, J. Z., Wu, C., Huan, W. W., Chen, L., & Li, B. (2022). Enhanced removal
of Cr(VI) by cation functionalized bamboo hydrochar. Bioresource Technology, 347, 126703.
https://doi.org/10.1016/j.biortech.2022.126703
27. Huezo, L., Vasco-Correa, J., & Shah, A. (2021). Hydrothermal carbonization of anaerobi-
cally digested sewage sludge for hydrochar production. Bioresource Technology Reports, 15,
100795. https://doi.org/10.1016/j.biteb.2021.100795
28. Al Afif, R., Tondl, G., & Pfeifer, C. (2023). Experimental and simulation study of hydrochar
production from cotton stalks. Energy, 276, 127573. https://doi.org/10.1016/j.energy.2023.
127573
29. Luo, Y., Lan, Y., Liu, X., Xue, M., Zhang, L., Yin, Z., He, X., Li, X., Yang, J., Hong, Z.,
Naushad, M., & Gao, B. (2023). Hydrochar effectively removes aqueous Cr(VI) through syner-
gistic adsorption and photoreduction. Separation and Purification Technology, 317, 123926.
https://doi.org/10.1016/j.seppur.2023.123926
30. He, J., Ren, S., Zhang, S., & Luo, G. (2021). Modification of hydrochar increased the capacity
to promote anaerobic digestion. Bioresource Technology, 341, 125856. https://doi.org/10.
1016/j.biortech.2021.125856
31. Sridhar, A., Kapoor, A., Senthil Kumar, P., Ponnuchamy, M., Balasubramanian, S., & Prab-
hakar, S. (2021). Conversion of food waste to energy: A focus on sustainability and life cycle
assessment. Fuel, 302, 121069. https://doi.org/10.1016/j.fuel.2021.121069
32. Sharma, P., Gaur, V. K., Sirohi, R., Varjani, S., Hyoun Kim, S., & Wong, J. W. C. (2021).
Sustainable processing of food waste for production of bio-based products for circular
bioeconomy. Bioresource Technology, 325, 124684.
33. Xie, X., Peng, C., Song, X., Peng, N., & Gai, C. (2022). Pyrolysis kinetics of the hydrothermal
carbons derived from microwave-assisted hydrothermal carbonization of food waste digestate.
Energy, 245, 123269. https://doi.org/10.1016/j.energy.2022.123269
34. Kumar, M., Dutta, S., You, S., Luo, G., Zhang, S., Show, P. L., Sawarkar, A. D., Singh, L., &
Tsang, D. C. W. (2021). A critical review on biochar for enhancing biogas production from
anaerobic digestion of food waste and sludge. Journal of Cleaner Production, 305, 127143.
35. Ul Saqib, N., Sarmah, A. K., & Baroutian, S. (2019). Effect of temperature on the fuel
properties of food waste and coal blend treated under co-hydrothermal carbonization. Waste
Management, 89, 236–246. https://doi.org/10.1016/j.wasman.2019.04.005
36. Usmani, Z., Sharma, M., Awasthi, A. K., Sharma, G. D., Cysneiros, D., Nayak, S. C., Thakur,
V. K., Naidu, R., Pandey, A., & Gupta, V. K. (2021). Minimizing hazardous impact of food
waste in a circular economy—Advances in resource recovery through green strategies. Journal
of Hazardous Materials, 416, 126154. https://doi.org/10.1016/j.jhazmat.2021.126154
37. Wang, L., Chi, Y., Du, K., Zhou, Z., Wang, F., & Huang, Q. (2022). Hydrothermal treatment
of food waste for bio-fertilizer production: Formation and regulation of humus substances
in hydrochar. Science of the Total Environment, 838, 155900. https://doi.org/10.1016/j.scitot
env.2022.155900
38. Cao, L., Zhang, C., Chen, H., Tsang, D. C. W., Luo, G., Zhang, S., & Chen, J. (2017).
Hydrothermal liquefaction of agricultural and forestry wastes: State-of-the-art review and
future prospects. Bioresource Technology, 245, 1184–1193.
39. Deng, F., Jiang, H., Xie, Z., Chen, Y., Zhou, P., Liu, X., & Li, D. (2022). Nutrient recovery
from biogas slurry via hydrothermal carbonization with different agricultural and forestry
14 Heterogeneous Hydrochar-Based Catalysts for Biodiesel Production 355

residue. Industrial Crops and Products, 189, 115891. https://doi.org/10.1016/j.indcrop.2022.


115891
40. Liu, Z., Quek, A., Kent Hoekman, S., Srinivasan, M. P., & Balasubramanian, R. (2012). Ther-
mogravimetric investigation of hydrochar-lignite co-combustion. Bioresource Technology,
123, 646–652. https://doi.org/10.1016/j.biortech.2012.06.063
41. Li, S., Ma, Q., Chen, L., Yang, Z., Aqeel Kamran, M., & Chen, B. (2022). Hydrochar-
mediated photocatalyst Fe3 O4 /BiOBr@HC for highly efficient carbamazepine degradation
under visible LED light irradiation. Chemical Engineering Journal, 433, 134492. https://doi.
org/10.1016/j.cej.2021.134492
42. Donar, Y. O., Çaǧlar, E., & Sinaǧ, A. (2016). Preparation and characterization of agricultural
waste biomass based hydrochars. Fuel, 183, 366–372. https://doi.org/10.1016/j.fuel.2016.
06.108
43. Huangfu, X., Xu, N., Yang, J., Yang, H., Zhang, M., Ye, Z., Wang, S., & Chen, J. (2020). Trans-
port and retention of hydrochar-diatomite nanoaggregates in water-saturated porous sand:
Effect of montmorillonite and phosphate at different ionic strengths and solution pH. Science
of the Total Environment, 703, 134487. https://doi.org/10.1016/j.scitotenv.2019.134487
44. Fakudze, S., Zhang, Y., Wei, Y., Li, Y. H., Chen, J., Wang, J., & Han, J. (2023).
Taguchi-optimized oxy-combustion of hydrochar/coal blends for CO2 capture and maxi-
mized combustion performance. Energy., 267, 126602. https://doi.org/10.1016/j.energy.2022.
126602
45. Santos, M. M., Marques Sierra, A. L., Amado-Fierro, Á., Suárez, M., Blanco, F., La Fuente,
J. M. G., Diez, M. A., & Centeno, T. A. (2023). Reducing cement consumption in mortars by
waste-derived hydrochars. Journal of Building Engineering, 75, 106987. https://doi.org/10.
1016/j.jobe.2023.106987
46. Celletti, S., Bergamo, A., Benedetti, V., Pecchi, M., Patuzzi, F., Basso, D., Baratieri, M., Cesco,
S., & Mimmo, T. (2021). Phytotoxicity of hydrochars obtained by hydrothermal carbonization
of manure-based digestate. Journal of Environmental Management, 280, 111635. https://doi.
org/10.1016/j.jenvman.2020.111635
47. Roman, F. F., Diaz De Tuesta, J. L., Praça, P., Silva, A. M. T., Faria, J. L., & Gomes, H. T.
(2021). Hydrochars from compost derived from municipal solid waste: Production process
optimization and catalytic applications. Journal of Environmental Chemical Engineering, 9,
104888. https://doi.org/10.1016/j.jece.2020.104888
48. Dong, X., Guo, S., Ma, M., Zheng, H., Gao, X., Wang, S., & Zhang, H. (2020). Hydrothermal
carbonization of millet stalk and dilute-acid-impregnated millet stalk: combustion behaviors
of hydrochars by thermogravimetric analysis and a novel mixed-function fitting method. Fuel,
273, 117734. https://doi.org/10.1016/j.fuel.2020.117734
49. Reza, M. T., Rottler, E., Herklotz, L., & Wirth, B. (2015). Hydrothermal carbonization (HTC)
of wheat straw: Influence of feedwater pH prepared by acetic acid and potassium hydroxide.
Bioresource Technology, 182, 336–344. https://doi.org/10.1016/j.biortech.2015.02.024
50. Li, Y., Tsend, N., Li, T. K., Liu, H., Yang, R., Gai, X., Wang, H., & Shan, S. (2019). Microwave
assisted hydrothermal preparation of rice straw hydrochars for adsorption of organics and
heavy metals. Bioresource Technology, 273, 136–143. https://doi.org/10.1016/j.biortech.2018.
10.056
51. Yusuf, I., Flagiello, F., Ward, N. I., Arellano-García, H., Avignone-Rossa, C., & Felipe-Sotelo,
M. (2020). Valorisation of banana peels by hydrothermal carbonisation: Potential use of the
hydrochar and liquid by-product for water purification and energy conversion. Bioresource
Technology Reports, 12, 100582. https://doi.org/10.1016/j.biteb.2020.100582
52. Azaare, L., Commeh, M. K., Smith, A. M., & Kemausuor, F. (2021). Co-hydrothermal
carbonization of pineapple and watermelon peels: Effects of process parameters on hydrochar
yield and energy content. Bioresource Technology Reports, 15, 100720. https://doi.org/10.
1016/j.biteb.2021.100720
53. Nakason, K., Panyapinyopol, B., Kanokkantapong, V., Viriya-empikul, N., Kraithong, W., &
Pavasant, P. (2018). Characteristics of hydrochar and liquid fraction from hydrothermal
carbonization of cassava rhizome. Journal of the Energy Institute, 91, 184–193. https://doi.
org/10.1016/j.joei.2017.01.002
356 M. Aliyu et al.

54. Li, Y., Meas, A., Shan, S., Yang, R., & Gai, X. (2016). Production and optimization of
bamboo hydrochars for adsorption of Congo red and 2-naphthol. Bioresource Technology,
207, 379–386. https://doi.org/10.1016/j.biortech.2016.02.012
55. Wang, C., Zhang, S., Wu, S., Sun, M., & Lyu, J. (2020). Multi-purpose production with
valorization of wood vinegar and briquette fuels from wood sawdust by hydrothermal process.
Fuel, 282, 118775. https://doi.org/10.1016/j.fuel.2020.118775
56. Andrade, J. G. S., Porto, C. E., Moreira, W. M., Batistela, V. R., & Scaliante, M. H. N. O.
(2023). Production of hydrochars from Pinus caribaea for biosorption of methylene blue and
tartrazine yellow dyes. Cleaner Chemical Engineering, 5, 100092. https://doi.org/10.1016/j.
clce.2022.100092
57. Zhao, K., Li, Y., Zhou, Y., Guo, W., Jiang, H., & Xu, Q. (2018). Characterization of
hydrothermal carbonization products (hydrochars and spent liquor) and their biomethane
production performance. Bioresource Technology, 267, 9–16. https://doi.org/10.1016/j.bio
rtech.2018.07.006
58. Khosravi, A., Zheng, H., Liu, Q., Hashemi, M., Tang, Y., & Xing, B. (2022). Production and
characterization of hydrochars and their application in soil improvement and environmental
remediation. Chemical Engineering Journal, 430, 133142. https://doi.org/10.1016/j.cej.2021.
133142
59. Cheng, H., Ji, R., Yao, S., Song, Y., Sun, Q., Bian, Y., Wang, Z., Zhang, L., Jiang, X., &
Han, J. (2021). Potential release of dissolved organic matter from agricultural residue-derived
hydrochar: Insight from excitation emission matrix and parallel factor analysis. Science of the
Total Environment, 781, 146712. https://doi.org/10.1016/j.scitotenv.2021.146712
60. Karatas, O., Khataee, A., & Kalderis, D. (2022). Recent progress on the phytotoxic effects of
hydrochars and toxicity reduction approaches. Chemosphere, 298, 134357.
61. Ma, X. Q., Liao, J. J., Chen, D. B., & Xu, Z. X. (2021). Hydrothermal carbonization of
sewage sludge: Catalytic effect of Cl− on hydrochars physicochemical properties. Molecular
Catalysis, 513, 111789. https://doi.org/10.1016/j.mcat.2021.111789
62. Li, C., Feng, Y., Zhong, F., Deng, J., Yu, T., Cao, H., & Niu, W. (2022). Optimization of
microwave-assisted hydrothermal carbonization and potassium bicarbonate activation on the
structure and electrochemical characteristics of crop straw-derived biochar. Journal of Energy
Storage, 55, 105838. https://doi.org/10.1016/j.est.2022.105838
63. Zulkornain, M. F., Shamsuddin, A. H., Normanbhay, S., Md Saad, J., & Ahmad Zamri,
M. F. M. (2022). Optimization of rice husk hydrochar via microwave-assisted hydrothermal
carbonization: Fuel properties and combustion kinetics. Bioresource Technology Reports, 17,
100888. https://doi.org/10.1016/j.biteb.2021.100888
64. Gajera, Z. R., Mungray, A. A., Rene, E. R., & Mungray, A. K. (2023). Hydrothermal carboniza-
tion of cow dung with human urine as a solvent for hydrochar: An experimental and kinetic
study. Journal of Environmental Management, 327, 116854. https://doi.org/10.1016/j.jen
vman.2022.116854
65. Cruz, O. F., Silvestre-Albero, J., Casco, M. E., Hotza, D., & Rambo, C. R. (2018). Activated
nanocarbons produced by microwave-assisted hydrothermal carbonization of Amazonian fruit
waste for methane storage. Materials Chemistry and Physics, 216, 42–46. https://doi.org/10.
1016/j.matchemphys.2018.05.079
66. Xu, J., Zhang, J., Huang, J., He, W., & Li, G. (2020). Conversion of phoenix tree leaves
into hydro-char by microwave-assisted hydrothermal carbonization. Bioresource Technology
Reports, 9, 100353. https://doi.org/10.1016/j.biteb.2019.100353
67. Khan, M. A., Hameed, B. H., Siddiqui, M. R., Alothman, Z. A., & Alsohaimi, I. H. (2022).
Physicochemical properties and combustion kinetics of food waste derived hydrochars.
Journal of King Saud University-Science, 34(4), 101941. https://doi.org/10.1016/j.jksus.2022.
101941
68. Wang, F., Guo, C., Liu, X., Sun, H., Zhang, C., Sun, Y., & Zhu, H. (2022). Revealing carbon-
iron interaction characteristics in sludge-derived hydrochars under different hydrothermal
conditions. Chemosphere, 300, 134572. https://doi.org/10.1016/j.chemosphere.2022.134572
14 Heterogeneous Hydrochar-Based Catalysts for Biodiesel Production 357

69. Peng, H., Li, Y., Wen, J., & Zheng, X. (2021). Synthesis of ZnFe2 O4 /B, N-codoped biochar via
microwave-assisted pyrolysis for enhancing adsorption-photocatalytic elimination of tetracy-
cline hydrochloride. Industrial Crops and Products, 172, 114066. https://doi.org/10.1016/j.
indcrop.2021.114066
70. Yu, J., Xie, S., & Zhang, T. (2022). Influences of hydrothermal carbonization on phosphorus
availability of swine manure-derived hydrochar: Insights into reaction time and temperature.
Materials Science for Energy Technologies, 5, 416–423. https://doi.org/10.1016/j.mset.2022.
10.001
71. Roman, F. F., Diaz, J. L., Tuesta, D., Praça, P., Faria, J. L., & Gomes, H. T. (2021). Hydrochars
from compost derived from municipal solid waste: Production process optimization and
catalytic applications. Journal of Environmental Chemical Engineering, 9, 1–9. https://doi.
org/10.1016/j.jece.2020.104888
72. Xiong, J., Pan, Z., Xiao, X., Huang, H., Lai, F., Wang, J., & Chen, S. (2019). Study on the
hydrothermal carbonization of swine manure: The effect of process parameters on the yield/
properties of hydrochar and process water. Journal of Analytical and Applied Pyrolysis, 144,
104692. https://doi.org/10.1016/j.jaap.2019.104692
73. Aliyu, M., Abdullah, A. H., & Tahir, M. I. B. M. (2022). A potential approach for converting
rubber waste into a low-cost polymeric adsorbent for removing methylene blue from aqueous
solutions. Indonesian Journal of Chemistry, 22, 653–665. https://doi.org/10.22146/ijc.69674
74. Aliyu, M., Abdullah, A. H., & Tahir, M. I. B. M. (2022). Adsorption tetracycline from aqueous
solution using a novel polymeric adsorbent derived from the rubber waste. Journal of the
Taiwan Institute of Chemical Engineers, 136, 104333. https://doi.org/10.1016/j.jtice.2022.
104333
75. McIntosh, S., Padilla, R. V., Rose, T., Rose, A. L., Boukaka, E., & Erler, D. (2022). Crop
fertilisation potential of phosphorus in hydrochars produced from sewage sludge. Science of
the Total Environment., 817, 153023. https://doi.org/10.1016/j.scitotenv.2022.153023
76. Zhang, S., Sheng, K., Yan, W., Liu, J., Shuang, E., Yang, M., & Zhang, X. (2021). Bamboo
derived hydrochar microspheres fabricated by acid-assisted hydrothermal carbonization.
Chemosphere, 263, 128093. https://doi.org/10.1016/j.chemosphere.2020.128093
77. Hurst, G., Ruiz-Lopez, S., Rivett, D., & Tedesco, S. (2022). Effect of hydrochar from acid
hydrolysis on anaerobic digestion of chicken manure. Journal of Environmental Chemical
Engineering, 10, 108343. https://doi.org/10.1016/j.jece.2022.108343
78. Yu, J., Zhu, Z., Zhang, H., Chen, T., Qiu, Y., Xu, Z., & Yin, D. (2019). Efficient removal of
several estrogens in water by Fe-hydrochar composite and related interactive effect mechanism
of H2 O2 and iron with persistent free radicals from hydrochar of pinewood. Science of the
Total Environment, 658, 1013–1022. https://doi.org/10.1016/j.scitotenv.2018.12.183
79. Li, H., Cao, M., Watson, J., Zhang, Y., & Liu, Z. (2021). In Situ hydrochar regulates Cu fate
and speciation: Insights into transformation mechanism. Journal of Hazard Materials, 410,
124616. https://doi.org/10.1016/j.jhazmat.2020.124616
80. Coral, N., Brasil, H., Rodrigues, E., da Costa, C. E. F., & Rumjanek, V. (2019). Microwave-
modified hydrotalcites for the transesterification of soybean oil. Sustainable Chemistry and
Pharmacy, 11, 49–53. https://doi.org/10.1016/j.scp.2019.01.002
81. Woranuch, W., Ngaosuwan, K., Kiatkittipong, W., Wongsawaeng, D., Appamana, W., Powell,
J., Lalthazuala Rokhum, S., & Assabumrungrat, S. (2022). Fine-tuned fabrication parameters
of CaO catalyst pellets for transesterification of palm oil to biodiesel. Fuel, 323, 124356.
https://doi.org/10.1016/j.fuel.2022.124356
82. Cao, Y., Dhahad, H. A., Esmaeili, H., & Razavi, M. (2022). MgO@CNT@K2 CO3 as a supe-
rior catalyst for biodiesel production from waste edible oil using two-step transesterification
process. Process Safety and Environmental Protection, 161, 136–146. https://doi.org/10.1016/
j.psep.2022.03.026
83. Xia, S., Hu, Y., Chen, C., Tao, J., Yan, B., Li, W., Zhu, G., Cheng, Z., & Chen, G. (2022).
Electrolytic transesterification of waste cooking oil using magnetic Co/Fe–Ca based catalyst
derived from waste shells: A promising approach towards sustainable biodiesel production.
Renewable Energy, 200, 1286–1299. https://doi.org/10.1016/j.renene.2022.10.071
358 M. Aliyu et al.

84. Takase, M., Bryant, I. M., Essandoh, P. K., & Amankwa, A. E. K. (2023). A comparative study
on performance of KOH and 32%KOH/ZrO 2–7 catalysts for biodiesel via transesterification
of waste Adansonia digitata oil. Green Technologies and Sustainability, 1, 100004. https://
doi.org/10.1016/j.grets.2022.100004
85. Zhang, W., Wang, C., Luo, B., He, P., Zhang, L., & Wu, G. (2022). Efficient and economic
transesterification of waste cooking soybean oil to biodiesel catalyzed by outer surface of
ZSM-22 supported different Mo catalyst. Biomass and Bioenergy, 167, 106646. https://doi.
org/10.1016/j.biombioe.2022.106646
86. Sukpancharoen, S., Katongtung, T., Rattanachoung, N., & Tippayawong, N. (2023).
Unlocking the potential of transesterification catalysts for biodiesel production through
machine learning approach. Bioresource Technology, 378, 128961. https://doi.org/10.1016/j.
biortech.2023.128961
87. Khanian-Najaf-Abadi, M., Ghobadian, B., Dehghani-Soufi, M., & Heydari, A. (2023). Exper-
imental evaluation of simultaneous variations in biodiesel yield and color using choline
hydroxide catalyst in an ultrasonic reactor. Journal of Cleaner Production, 382, 134767.
https://doi.org/10.1016/j.jclepro.2022.134767
88. Ningaraju, C., Yatish, K. V., & Sakar, M. (2022). Simultaneous refining of biodiesel-derived
crude glycerol and synthesis of value-added powdered catalysts for biodiesel production: A
green chemistry approach for sustainable biodiesel industries. Journal of Cleaner Production,
363, 132448. https://doi.org/10.1016/j.jclepro.2022.132448
89. Parida, S., Singh, M., & Pradhan, S. (2022). Biomass wastes: A potential catalyst source for
biodiesel production. Bioresource Technology Reports, 18, 101081.
90. Ashine, F., Kiflie, Z., Prabhu, S. V., Tizazu, B. Z., Varadharajan, V., Rajasimman, M., Joo, S.
W., Vasseghian, Y., & Jayakumar, M. (2023). Biodiesel production from Argemone mexicana
oil using chicken eggshell derived CaO catalyst. Fuel, 332, 126166. https://doi.org/10.1016/
j.fuel.2022.126166
91. Kedir, W. M., & Asere, T. G. (2022). Biodiesel production from waste frying oil using catalysts
derived from waste materials. Journal of the Turkish Chemical Society Section A: Chemistry,
9, 939–952. https://doi.org/10.18596/jotcsa.997456
92. Sami, H. Q. A., Hanif, M. A., Rashid, U., Nisar, S., Bhatti, I. A., Rokhum, S. L., Tsubota,
T., & Alsalme, A. (2022). Environmentally safe magnetic nanocatalyst for the production of
biodiesel from Pongamia pinnata oil. Catalysts, 12, 1266. https://doi.org/10.3390/catal1210
1266
93. Jaruwat, D., Udomsap, P., Chollacoop, N., & Eiad-ua, A. (2019). Preparation of carbon
supported catalyst from cattail leaves for biodiesel fuel upgrading application. Materials
Today: Proceedings, 17, 1319–1325.
94. Amenaghawon, A. N., Obahiagbon, K., Isesele, V., & Usman, F. (2022). Optimized biodiesel
production from waste cooking oil using a functionalized bio-based heterogeneous catalyst.
Cleaner Engineering and Technology, 8, 100501. https://doi.org/10.1016/j.clet.2022.100501
95. Tosun, Z., & Aydin, H. (2022). Combustion, performance and emission analysis of propanol
addition on safflower oil biodiesel in a diesel engine. Cleaner Chemical Engineering., 3,
100041. https://doi.org/10.1016/j.clce.2022.100041
96. Yusuff, A. S., & Owolabi, J. O. (2019). Synthesis and characterization of alumina supported
coconut chaff catalyst for biodiesel production from waste frying oil. South African Journal
of Chemical Engineering, 30, 42–49. https://doi.org/10.1016/j.sajce.2019.09.001
97. Abdullah, R. F., Rashid, U., Hazmi, B., Ibrahim, M. L., Tsubota, T., & Alharthi, F. A.
(2022). Potential heterogeneous nano-catalyst via integrating hydrothermal carbonization for
biodiesel production using waste cooking oil. Chemosphere, 286, 131913. https://doi.org/10.
1016/j.chemosphere.2021.131913
98. Parandi, E., Safaripour, M., Mosleh, N., Saidi, M., Rashidi Nodeh, H., Oryani, B., & Rezania,
S. (2023). Lipase enzyme immobilized over magnetic titanium graphene oxide as catalyst for
biodiesel synthesis from waste cooking oil. Biomass and Bioenergy, 173, 106794. https://doi.
org/10.1016/j.biombioe.2023.106794
14 Heterogeneous Hydrochar-Based Catalysts for Biodiesel Production 359

99. Abdullah, R. F., Rashid, U., Ibrahim, M. L., NolHakim, M. A. H. L., Moser, B. R., &
Alharthi, F. A. (2021). Bifunctional biomass-based catalyst for biodiesel production via
hydrothermal carbonization (HTC) pretreatment—Synthesis, characterization and optimiza-
tion. Process Safety and Environmental Protection, 156, 219–230. https://doi.org/10.1016/j.
psep.2021.10.007
100. Rajendran, N., Kang, D., Han, J., & Gurunathan, B. (2022). Process optimization, economic
and environmental analysis of biodiesel production from food waste using a citrus fruit peel
biochar catalyst. Journal of Cleaner Production, 365, 132712. https://doi.org/10.1016/j.jcl
epro.2022.132712
101. Naidu, B. R., & Venkateswarlu, K. (2022). Dried water extract of pomegranate peel ash
(DWEPA) as novel and biorenewable heterogeneous catalyst for biodiesel production and
biopotent quinoxalines synthesis. Bioresource Technology Reports, 18, 101107. https://doi.
org/10.1016/j.biteb.2022.101107
102. Nath, B., Basumatary, B., Brahma, S., Das, B., Kalita, P., Rokhum, S. L., & Basumatary, S.
(2023). Musa champa peduncle waste-derived efficient catalyst: Studies of biodiesel synthesis,
reaction kinetics and thermodynamics. Energy, 270, 126976. https://doi.org/10.1016/j.energy.
2023.126976
103. Tu, R., Sun, Y., Wu, Y., Fan, X., Wang, J., Shen, X., He, Z., Jiang, E., & Xu, X. (2019). Effect
of surfactant on hydrothermal carbonization of coconut shell. Bioresource Technology, 284,
214–221. https://doi.org/10.1016/j.biortech.2019.03.120
104. Espro, C., Satira, A., Mauriello, F., Anajafi, Z., Moulaee, K., Iannazzo, D., & Neri, G. (2021).
Orange peels-derived hydrochar for chemical sensing applications. Sensors and Actuators B:
Chemical, 341, 130016. https://doi.org/10.1016/j.snb.2021.130016
105. Zhang, C., Ma, X., Huang, T., Zhou, Y., & Tian, Y. (2020). Co-hydrothermal carbonization of
water hyacinth and polyvinyl chloride: Optimization of process parameters and characteriza-
tion of hydrochar. Bioresource Technology, 314, 123676. https://doi.org/10.1016/j.biortech.
2020.123676
106. Siregar, A. G. A., Manurung, R., & Taslim, T. (2021). Synthesis and characterization of
sodium silicate produced from corncobs as a heterogeneous catalyst in biodiesel production.
Indonesian Journal of Chemistry, 21, 88–96. https://doi.org/10.22146/ijc.53057
107. Yani, F. T., Ulhaqi, R., Pratiwi, W. P., Pontas, K., & Husin, H. (2019). Utilization of water
hyacinth-based biomass as a potential heterogeneous catalyst for biodiesel production. Journal
of Physics: Conference Series, 1402(5), 055005.
108. Nahas, L., Dahdah, E., Aouad, S., El Khoury, B., Gennequin, C., Abi Aad, E., & Estephane,
J. (2023). Highly efficient scallop seashell-derived catalyst for biodiesel production from
sunflower and waste cooking oils: Reaction kinetics and effect of calcination temperature
studies. Renewable Energy, 202, 1086–1095. https://doi.org/10.1016/j.renene.2022.12.020
109. Adepoju, T. F., Victor, E., Ekop, E. I., Emberru, R. E., Balogun, T. A., & Adeniyi, A. D. (2022).
Residual wood ash powder: A predecessor for the synthesis of CaO–K2 O–SiO2 base catalyst
employed for the production of biodiesel from Asimina triloba oil seed. Case Studies in
Chemical and Environmental Engineering, 6, 100252. https://doi.org/10.1016/j.cscee.2022.
100252
Chapter 15
Circular Bioeconomy Approaches
for Valorizing Waste Streams into Bio-jet
Fuel

Louella Concepta Goveas, S. M. Vidya, Ramesh Vinayagam,


and Raja Selvaraj

Abstract As the aviation sector is experiencing phenomenal growth, the substan-


tial demand for aviation fuel is also increasing. Synthesis of sustainable aviation fuel
would help curb this enhanced demand while reducing greenhouse gas (GHGs) emis-
sions. In this regard, bio-jet fuels have shown promising potential as they are renew-
able, release less CO2 on combustion, highly compatible with aviation engines, and
can be used without any modifications. Several researchers have exploited sustain-
able waste sources for bio-jet fuel synthesis by various conversion technologies,
which utilize the principles of circular bioeconomy. This book chapter particu-
lars on the aspects of circular bioeconomy along with the conversion technologies,
advantages, and disadvantages of bio-jet fuel. Additionally, circular economy-based
production of bio-jet fuel from various waste streams, as demonstrated by researchers
are thoroughly reviewed. The various challenges to bio-jet fuel production and their
possible solutions are assessed to facilitate the suitable implementation of a circular
bioeconomy.

Keywords Circular bioeconomy · Bio-jet fuel · Bioenergy · Waste stream ·


Valorization

L. C. Goveas (B) · S. M. Vidya


Department of Biotechnology Engineering, NMAM Institute of Technology, Nitte (Deemed to be
University), Mangalore, Karnataka 574110, India
e-mail: goveas.louella@nitte.edu.in
R. Vinayagam · R. Selvaraj (B)
Department of Chemical Engineering, Manipal Institute of Technology, Manipal Academy of
Higher Education, Manipal, Karnataka 576104, India
e-mail: raja.s@manipal.edu

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2024 361
G. Baskar et al. (eds.), Circular Bioeconomy Perspectives in Sustainable Bioenergy
Production, Energy, Environment, and Sustainability,
https://doi.org/10.1007/978-981-97-2523-6_15
362 L. C. Goveas et al.

15.1 Introduction

The aviation sector is budding rapidly and is anticipated to continue to grow. The
demand for air transport is rising at an average of 4.3% per year and by 2030 at least
200,000 flights are predicted to fly around the world daily [1]. Subsequently, a 5%
increase per annum in the fossil fuel requirement for the aviation sector by 2030 is
projected [2]. Depletion of fossil fuels is presently the cause for alarm and addition-
ally, using non-renewable energy sources results in the emission of greenhouse gases
(GHGs), mostly CO2, which contributes to global warming. The CO2 emissions from
the worldwide aviation sector accounted for 2.3% of global energy-related emissions
in 2020 and are expected to reach 4.3–4.8% by 2030 [3]. Therefore, the develop-
ment of renewable aviation fuel which can aid in the facilitation of continued energy
supply with reduced environmental load is vital.
Bio-jet fuel is the most promising renewable substitute and has been considered
to replace up to 30% of the fuel for the aviation industry by 2030. The utilization of
this fuel offers many advantages such as reduced CO2 emissions, non-requirement
of engine modifications, and lesser aromatic components [4]. Raw materials such
as corn grains, oil seeds, and non-edible crops were primarily used for bio-jet fuel
production by various conversion technologies, however, recent times have seen
utilization of waste resources such as agricultural and forest residues, municipal
solid waste streams, oil industry wastewaters as well as wet wastes [5]. Not only
does this contribute to the generation of a sustainable energy source, but also helps
implement the concept of circular bioeconomy [6]. Circular bioeconomy refers to
the utilization of waste into useful products including energy, thereby creating a
zero-waste environment. This consequentially reduces natural energy and resource
depletion and thereby enforces the concept of recycling and reuse [7].
Although the need for bio-jet fuel is vast and its utilization is highly advanta-
geous, many challenges affect its large-scale manufacture and subsequent marketing.
A constant supply of feedstock, price difference, well-researched manufacturing
process, and availability of cheap hydrogen are some of the numerous challenges
that must be worked upon to facilitate the full-fledged use of bio-jet fuel as an
alternative [8].
The present review discusses the significance of bio-jet fuels as an alternate energy
source for the aviation industry. It also substantiates the principles of circular bioe-
conomy and its emphasis on recycling wastes and reuses into useful products. The
main focus of this review is on the circular economy-based generation of bio-jet fuels
from various waste streams, while addressing the challenges faced for substantial
bio-jet fuel production.
15 Circular Bioeconomy Approaches for Valorizing Waste Streams … 363

15.2 Circular Bioeconomy—The Concept

Circular bioeconomy is the merger of bioeconomy and circular economy, two promi-
nent approaches with several intersecting topics (Fig. 15.1) [9]. A thorough under-
standing of circular bioeconomy requires the reader to comprehend the basic concepts
of these two approaches.
The idea of bioeconomy was introduced in the 2000s by European Union to
propagate the implementation of biotechnology for valuable product generation.
Bioeconomy considers the utilization and subsequent conversion of renewable bio-
based resources and their waste streams into valuable products. Bioeconomy also
aims at the development of economy by generating employment opportunities, and
by responding to social and environmental challenges [10]. It majorly depends on
sustainable techniques for the utilization of biological sources, scientific technolo-
gies, and advancements in all economic sectors leading to a sustainable economy
[7]. It provides solutions to worldwide problems by delivering safe and healthy food,
bio-energy as a substitute to fossil fuels, novel processing techniques and chemicals,
new product characteristics, sustainable bio-farming practices, and green chemistry,
thereby resulting in a reduction of GHG emissions leading to minimization of global
warming [11]. The sustainable development goals (SDGs) are a set of 17 specific
objectives implemented by the United Nations to achieve prosperity while promoting
a safe and sustainable environment [12]. Bioeconomy is considered highly impor-
tant as it can help establish several SDGs by aiding in achieving the goal of zero
hunger by providing safe and healthy food, enabling clean water for consumption
by wastewater treatment, helping to generate bioenergy resulting in the reduction of
climate change, and also promoting sustainability in all the aspects of living [10].
The circular economy is the concept of manufacture and utilization, wherein the
prevailing constituents and goods are reused, restored, revamped, and recycled for
prolonged periods. In this way, the life cycle of products is extended [13]. It implies
minimizing waste, by recycling the materials from a product at the end of its life cycle.
These materials are further reused to generate new value. Circular economy hence

Fig. 15.1 Concepts of bioeconomy and circular economy


364 L. C. Goveas et al.

lessens the utilization of raw materials, energy, and water by conversion of generated
debris into useful products [14]. The basic principles of the circular economy include
reuse, recycling, and repair thereby minimizing the total amount of debris that is
generated, circulation of goods and resources at their highest value, and regenerating
the environment. The major focus of circular economy is the continuous utilization
of wastes thereby limiting the input of energy and raw materials [15]. This concept
aims to offer solutions to two issues, i.e., conservation of resources, and substantially
minimizes the adverse influence on environment and climate. The circular economy
is distinct from the concept of linear economy, wherein the goods are discarded at the
end of their life cycle, which results in a lot of expenditure of energy and resources
[9]. The conversion from a linear to a circular economy would result in environmental
protection by reduction of waste eliminating pollution, reducing the requirement of
resources in this era of population crisis and boosting pecuniary growth.
The merger of bioeconomy and circular economy substantiates the sustainable
use of biological sources, to generate goods, improving their yield, with a simulta-
neous decrease in detrimental effects on the environment. This incorporation, thereby
assists in the development of the present economic strategy [13]. If the implemen-
tation of a linear economy is persistent, it results in the depletion of existing natural
sources, unraveling into an ecological crisis. Therefore, a bioeconomy-dependent
circular economy will enable us to restrict the expenditure of natural resources, by
recycling and reuse of generated biological-based goods and their wastes [16]. Hence,
circular bioeconomy emphasizes that the utilization of biological resources does not
adversely affect the global climate, biodiversity, and the environment. Circularity and
sustainable development are both promoted by circular bioeconomy, which ensures
that the biological resources are utilized to their maximum strength [17]. This results
in the complete minimization of waste products and restricts the utilization of fossil
fuels.

15.3 Bio-jet Fuel—Synthesis from Waste Streams

Commercial aviation fuel is synthesized by distilling crude oil at 205 °C and


comprises n-paraffins, branched iso-alkanes, naphthenes (cyclic alkanes), and
aromatics. Naphthene content reduces the freezing point of the fuel, while the
presence of aromatics helps in the prevention of leakages. The number of alkanes
contributes to the energy density. The major problem with the utilization of this fuel is
its carbon emission contributing to climate change [18]. Additionally, the depleting
sources of non-renewable reserves make it extremely challenging to fend for the
world in these times of booming economy. A suitable substitute is the conversion of
waste biomasses into sustainable aviation fuel. The present section deliberates on the
basics of bio-jet fuel, its production routes, and circular bioeconomy-based synthesis
through various waste streams.
15 Circular Bioeconomy Approaches for Valorizing Waste Streams … 365

15.3.1 Bio-jet Fuel—Composition and Advantages

Bio-jet fuel is also called bio-kerosene or synthetic paraffinic kerosene and consists
of n-paraffins, cyclo-alkanes, and branched iso-alkanes. Aromatics may or may not
be present and is dependent on the synthesis route and raw materials. Hence, bio-
jet fuel has to be blended up to a maximum of 50% with conventional jet fuel [4].
The utilization of bio-jet fuel substantially reduces CO2 release and helps achieve
zero-carbon emissions. As this fuel is manufactured from biomass, it reduces CO2
emissions between 20 and 90% when compared with jet fuels. Air-craft engines do not
require any modifications to sustain this fuel and hence could be blended with fossil
fuels, without any adaptation. Additionally, these fuels possess fewer sulfur amounts,
exhibit lesser emissions from the tailpipe, have exceptional cold flow characteristics,
and have greater heat and oxidation stability [19]. One major drawback of bio-jet
fuels is their absence of aromatics which can result in fuel leaking, due to shrinkage
and hardening of seals. The addition of aromatics could help solve this problem, but
combustion causes soot formation [20].

15.3.2 Production Routes

Bio-based raw materials are converted into bio-jet fuel via several conversion tech-
nologies which predominantly depend on the composition of raw materials. Four
classes of conversion technologies include oil-to-jet, gas-to-jet, sugar-to-jet, and
alcohol-to-jet which multiple techniques available under each class [21] (Fig. 15.2).

Fig. 15.2 Classification of bio-jet fuel production routes


366 L. C. Goveas et al.

Oil-to-jet: Hydrogenated esters and fatty acids (HEFA), catalytic hydro-thermolysis


(CHT), and hydro-treated depolymerized cellulosic jet (HDCJ) are the three
techniques involved under oil-to-jet technologies [21].
Triglycerides, unsaturated and saturated fats usually present in vegetable and
animal-based oils are treated via HEFA to produce bio-jet fuels. HEFA is a two-
step process, the saturated fats, obtained by the conversion of unsaturated fats and
triglycerides via catalytic hydrogenation, are further converted into straight-chain
alkanes and propane along with water and CO2 . The straight alkanes are cracked and
isomerized into mixed alkane liquid fuel, later fractionated to generate bio-jet fuel
along with naphtha, paraffinic diesel, and light gases. The first step is catalyzed by
Ni, Co, Mo, or zeolites of noble metals while, metals such as platinum, and nickel
associated with zeolite and activated carbon catalyze the second step [8].
CHT or hydrothermal liquefaction is utilized for the conversion of plant-based
or algal oils into bio-jet fuel with aromatics. Processing steps, such as conjuga-
tion, cyclization, and cross-linking are carried out on triglycerides, which are then
subjected to hydrolysis and cracking in the presence of catalysts and water. A mixture
of hydrocarbons is obtained by decarboxylation and dehydration, later fractionated
into naphtha, diesel, and bio-jet fuel [21, 22].
HDCJ is a two-step technique where the bio-oil obtained from pyrolysis of ligno-
cellulosic biomass along with biochar and biogas is converted into bio-jet fuel.
Catalytically hydro-treated bio-oil is subjected to hydrogenation at high tempera-
tures to achieve hydrocarbon fuel. This mixture is then fractionated to separate the
bio-jet fuel [23].
Although the techniques of production of hydrocarbon mixtures are different in
these three techniques, the downstream processing step of fractionating into bio-jet
fuel is the same.
Gas-to-jet: Several types of gas such as syngas, natural gas, or bio-gas can be
converted into bio-jet fuel via Fischer Tropsch and gas fermentation techniques.
Syngas and biogas are produced by gasification and pyrolysis of lignocellulosic
biomass.
Fischer Tropsch process involves the synthesis of liquid bio-jet fuel from syngas
and is achieved in six steps. Dried lignocellulosic biomass is ground and subjected to
gasification under steam and pure oxygen at high temperatures. Ash and tar present in
the syngas is removed via a cooling system, beside the gasifier. An acid gas removal
system removes impurities such as hydrogen sulfide, acid gas, carbon dioxide, and
sulfides. The purified syngas then is mixed with hydrogen and carbon monoxide,
which then enters the bio-jet oil production process. A series of catalytic processes
convert the syngas to liquid fuels in the Fischer Tropsch system, which can either
be operated at low temperatures (200–250 °C) or high temperatures (300–350 °C)
with iron-based and cobalt-based oxide nanoparticle-based catalysts, respectively.
The liquid fuel is then subjected to hydrocracking, and distilled to achieve bio-jet
fuel. Excess gas is used for the generation of electricity and is utilized to operate the
system [24].
15 Circular Bioeconomy Approaches for Valorizing Waste Streams … 367

The syngas produced from lignocellulosic biomass could be microbially


fermented to produce bio-jet fuel instead of catalytic conversion. This process, known
as gas fermentation, usually involves the fermentation of cold syngas into butanol/
acetone by Clostridium species, an acetogenic bacteria. These alcohols could be then
converted into bio-jet fuel via an alcohol-to-jet-based process involving several steps
including oligomerization, distillation, and hydrogenation. Raw materials such as
organic waste arising from households and industries could also be refurbished into
bio-jet fuel via gas fermentation [21].
Sugar-to-jet: The conversion of lignocellulosic biomass containing multiple compo-
nents is processed to sugars which are then converted to bio-jet oil by two promi-
nent processes, i.e., aqueous phase reforming of sugars to hydrocarbons and direct
fermentative conversion of sugars to bio-jet fuel.
The aqueous phase reforming process includes pre-treatment of lignocellulosic
biomass via enzymes to obtain cellulose and hemicellulose, which are further acid or
enzymatically hydrolyzed to obtain short-length sugars (5–6 carbon). Hydrogenation
or hydrogenolysis is then performed to convert these sugars to polyhydric alcohols
and short-chain oxygenated compounds, respectively [21]. These intermediates are
then converted to oxygenates via aqueous phase reforming which includes several
steps such as dehydrogenation, deoxygenation, hydrogenation, hydrogenolysis, and
cyclization. The oxygenates formed are then processed into alkanes via any of three
reactions, i.e., direct catalytic condensation, acid condensation, and hydrogenation/
dehydrogenation. The generated alkanes are then converted to bio-jet fuel via a series
of steps [25].
Genetically engineered microbial strains ferment sugars directly to alkane kind
fuels, bypassing the alcohol step. Raw materials such as lignocellulosic biomass
or crops are primarily pre-treated to produce small-chain sugars enzymatically or
mechanically [26]. These sugars are then converted to acetyl CoA via glycolysis
which are then converted to isopentyl pyrophosphate, a liquid fuel precursor, via
either of two pathways, i.e., mevalonic acid or deoxyxylulose-5-phosphate pathway.
This precursor is then converted to bio-jet fuel post-hydro-processing [21].
Alcohol-to-jet: Several alcohols such as methanol, ethanol, and higher alcohols are
processed into bio-jet fuels by catalytic conversion or fermentation. The lignocel-
lulosic biomass is pre-treated by acid or enzymes to obtain sugars which are then
subsequently fermented to alcohols by microbes or yeasts. During catalytic conver-
sion, the olefins produced from the catalytic dehydration of alcohols are oligomerized
to produce an intermediate distillate, which is then converted to hydrocarbons via
hydrogenation. In the final step, these hydrocarbons are then distilled to obtain the
bio-jet fuel [8].
368 L. C. Goveas et al.

15.3.3 Conversion of Waste Streams to Bio-jet Fuels

The raw materials used for the production of bio-jet fuels are starch, sugars, oils,
and triglycerides. The first-generation jet fuel was produced from food crops, but
due to overwhelming demand, the present focus is on agricultural wastes, forest, and
woody by-products, organic matter from municipal and industrial wastes, animal
fats, and used vegetable oils (Table 15.1). In the present times, the focus is on the
use of algal oils, a promising raw material for biofuel production, and also aids in
carbon sequestration.
Lignocellulosic Wastes: The cellulose and hemicellulose present in lignocellu-
losic wastes can be essentially converted into bio-jet fuel by techniques such
as HDCJ, Fischer Tropsch process, direct fermentation, and catalytic conversion.
Agricultural by-products, forest residues, and several waste biomasses have been
exploited by several researchers to produce bio-jet fuel with detailed economic anal-
ysis. Oxygenates were obtained from hemicellulose and cellulose present in corn-
cobs, which were subsequently converted into bio-jet fuel by hydrogenation and
hydrodeoxygenation at a yield of 0.125 kg/kg of corncob, with an operating cost of
1540 dollars per ton [37]. Cellulose and hemicellulose from corn stalks were utilized
as a source for bio-jet fuel production by aqueous conversion, wherein the lignin
which usually goes as waste was also used for H2 production. This hydrogen was
subsequently involved in hydrogenation during the fuel production process [38]. Rice
husk, a major agricultural waste byproduct in Taiwan was pyrolyzed and converted
into liquid bio-jet fuel by the Fischer Tropsch process at a final yield of 52%. Anal-
ysis of the fuel by gas chromatography–mass spectrometry revealed the presence of
iso- and cyclo-alkanes which made it appropriate for usage as a bio-jet fuel [39]. In
a recent study, poplar biomass was treated by enzymes and dilute acids, the released
sugars were fermented by Clostridium thermoaceticum, to acetic acid. Acetic acid

Table 15.1 Bio-jet fuel from various lignocellulosic and oil-based biomass
Raw material Production technique Yield References
Chicken fat Hydro-treating 46.47% [27]
Microalgae, Pongamia seeds, and HEFA, fermentation – [28]
sugarcane
Waste cooking oil Hydro-processing – [29]
Grindelia squarrosa Biphasic tandem catalytic process 90% carbon [30]
Jatropha curcas fruit Fermentation, alcohol to jet 40.37% [31]
Rice husk Fischer Tropsch process 42.50% [32]
Douglas fir sawdust pellets Tandem catalytic conversion – [33]
Vegetable oil, sugarcane juice HEFA, alcohol to jet – [34]
Sawdust Fischer Tropsch process – [35]
Poplar biomass Fermentation, alcohol to jet 330 L/ton [36]
15 Circular Bioeconomy Approaches for Valorizing Waste Streams … 369

thus produced was converted into hydrocarbon fractions via reactive distillation,
hydrotreatment, dehydration, and oligomerization. The unused lignin content may
be burned for electricity generation or hydrogen production leading to the reusing of
by-products generated during this process [36]. Waste woody biomass obtained post-
forest harvesting was pre-treated to release sugars, which were then consequently
converted to iso-butanol by fermentation. The produced iso-butanol was oligomer-
ized and hydrogenized to generate iso-paraffinic kerosene. The residue obtained post
this process was further utilized for biogas production, whereas the wastewater was
treated and recycled back for the pre-treatment step [40]. Similarly, another study
in British Columbia utilized wood pellets and forest residues that were pyrolyzed to
produce bio-oil which was upgraded and converted into bio-jet fuel. Additionally,
bio-gas and bio-char were produced which were utilized for other purposes. Lifecycle
analysis revealed that about 300 million liter per year of bio-jet fuel could be produced
per year from forest residue which could reduce the carbon intensity by 110% [41].
When potato by-products and sugar beets from processing industries were used as
raw materials for bio-jet fuel production, the GHG emissions were reduced by 52%
and 44% correspondingly. The production proceeded by alcohol-to-jet conversion,
where the pre-treated sugars were fermented to acetone, butanol, and ethanol by
Clostridium strains. These alcohols were then catalytically converted to bio-jet fuel
by condensation and hydro-treatment [42]. Eucalyptus wood chips were converted
to bio-jet oil by pre-treatment, fermentation into alcohols, catalytic hydrogenation,
and oligomerization. Fermentation was performed by Saccharomyces cerevisiae and
Clostridium strains. It was observed that the eucalyptus-based fuel showed 93% less
GHG emissions than the fossil fuel-based jet fuel [43].
Oil and Triglyceride-Based Wastes: Processes such as HEFA and CHT convert
the triglycerides and esters present in oil-based wastes into sustainable bio-jet fuel.
The most popular oil-based waste streams utilized by researchers are non-edible oil
seeds, waste cooking oils, and animal fats.
Waste cooking oils contain free fatty acids which are a valuable reserve for conver-
sion into bio-based fuels. As it is generated worldwide on a large-scale basis, its
repeated consumption and further, its disposal results in severe health and envi-
ronmental hazards [44]. Waste cooking oil obtained from Egyptian restaurants was
subjected to hydrothermal catalytic cracking in the presence of a nano zinc aluminate
catalyst. Bio-jet fuel was then obtained by fractional distillation and blended at 5%
with Jet A1 fuel [45]. Waste cooking oil was pre-treated and converted into bio-jet
fuel by a one-step hydrogenation reaction. The carbon intensity of this fuel was 63%
lesser than that of conventional fuel, while the GHG emissions were a mere 18%
[46].
Non-edible oil seeds, oily biomass, and residues depict maximum potential to
be converted into biofuels as their production is cost-effective without competition
to food security. Jatropha and castor oil were converted into bio-jet fuel by HEFA
and resulting in reduced environmental impacts in China [47]. Oil, extracted from
Jatropha curcas dehulled fruits, was subjected to hydrogenation and converted to
propane. This propane was then converted into bio-jet fuel via a series of reactions
370 L. C. Goveas et al.

through the HEFA process [47]. Bio-jet fuel was produced from non-edible date palm
seed oil by a one-step process catalyzed by tantalum phosphate. The hydrocarbon
mixture obtained as a product contained 53.6% bio-jet fuel, while the rest was green
diesel fraction [48]. Sunflower oil was extracted from sunflower residual wastes by
the cold press and treated by catalytic cracking under a ZSM-5 catalyst. The upgraded
oil was then distilled to obtain hydrocarbons, which on hydrogenation yielded bio-
jet fuel. The physical and chemical properties were comparable with conventional
aviation fuel [49]. Similarly, waste soybean oil and palm oil residues were converted
via a single step into good quality bio-jet fuel-based hydrocarbons through a Pd-based
zeolite catalyst without hydrogen addition [50].
Animal carcasses are generated in large amounts owing to increased consumption
of meat and poultry, due to the global population. About 25% of the carcass contains
fat, which could be easily utilized for the production of biofuels. The fuels obtained
from these triglycerides release lower emissions when compared with oil seed-based
crops. In a recent study, chicken fat was hydro-processed by a single step into bio-jet
fuel with 94% conversion and reduced carbon emissions [27]. Bio-oil produced by
pyrolysis of lignocellulosic waste was blended with poultry fat, to give better quality.
Subsequent conversion and hydro-treatment by biochar-based catalyst resulted in the
generation of bio-jet fuel [51].

15.4 Challenges—Emphasis on Economy and Environment

Although the generation of bio-jet fuel is valuable for circular bioeconomy-based


sustainable development directing to reduced carbon emissions, several problems
affect its large-scale production and utilization. The primary challenge is the unin-
terrupted supply and availability of feedstock. If non-edible crops are utilized as raw
materials, their growth would lead to food and water scarcity [52]. Additionally, the
application of fertilizers and pesticides causes groundwater and soil contamination.
If agricultural by-products, forest residues, industrial processing wastes, or waste
oils are to be employed as a source, it must be ensured that a continuous supply
with consistent quality must always be accessible which is rather challenging [53].
Feedstocks such as weeds, algae, and halophytes having the potential to grow in
wastes and waste regions could be used as sources for bio-jet fuel production. This
will solve the difficulty of competition for food crops and water usage, also, they are
not harmful to the environment. On the contrary, they sequester atmospheric CO2
and help in the prevention of climate change [54].
Production of bio-jet fuel requires a large water footprint, which is dependent
on the raw material. The water footprint was considerably low when microalgae
were utilized and drastically increased when crops, such as sugarcane, rapeseed
oil, jatropha, and soybean oils were used as raw materials, as irrigation contributed
to the water usage. On the other hand, the energy requirement for the growth of
microalgae is high contributing to the increased costs of bio-jet fuel [55]. Another
challenge is the increased production costs which make bio-jet fuel expensive when
15 Circular Bioeconomy Approaches for Valorizing Waste Streams … 371

compared with conventional aviation fuels. The increased cost is mostly due to
the present availability of feedstocks and the continued development of production
processes. Additionally, factors, such as feedstock pre-treatment, hydrogen and cata-
lyst consumption, and operational costs also contribute to the high costs. A ton of
synthetic jet fuel costs 317 dollars per ton in 2017, while the cost of raw mate-
rials for bio-jet fuel such as animal waste, and algal and vegetable oils is higher
(550–1200 dollars per ton) [56]. Lignocellulosic waste requires processing by pre-
treatment, fermentation, or pyrolysis to be made suitable for bio-jet fuel production.
The residual waste generated consists of soil contaminants and micro-organisms that
must be treated before disposal which may contribute to economic costs. Hence the
production costs of bio-jet fuel exceed 7–8 times more than those of conventional
jet fuels [23].
Renewable jet fuels have to be blended with synthetic fuels, as utilization of pure
bio-jet fuel may freeze at flight operational conditions. Bio-jet fuels also have less
thermal stability properties at high temperatures and may be non-stable at continued
storage conditions. These fuels may contain small fractions of metals, impurities, and
micronutrients from the raw materials, requiring rigorous processing [56]. Subse-
quent quality check is performed to ensure its usage in aircraft. The produced fuel
should be resistant to a wide range of operational conditions and offer good perfor-
mance. Presently, bio-jet fuels blend with Jet A-1 at 1:1 without any issues and most
aircrafts have been certified to fly on 50% alternate fuels [23].

15.5 Conclusions

Production of bio-jet fuels from lignocellulosic and triglyceride-based waste streams


serves to solve multiple issues of non-renewable energy scarcity, meeting the
demands of the aviation industry, environmental pollution, and global climate change.
The utilization of wastes to produce a useful product includes aspects of reuse, recy-
cling, refurbishment, and reduction leading to the implementation of a circular bioe-
conomy. This also contributes to the United Nation’s Sustainable Development Goals
and helps build a secure future for the upcoming generations. The large-scale produc-
tion and utilization of bio-jet fuel is yet to be achieved owing to several challenges
such as the non-availability of stable feedstocks and environmental and economic
issues. Efficient technology and feedstock could be achieved through research with
support from the government, leading to sustainable production of bio-jet fuels.

References

1. Fageda, X., & Teixidó, J. J. (2022). Pricing carbon in the aviation sector: Evidence from the
European emissions trading system. Journal of Environmental Economics and Management,
111, 102591. https://doi.org/10.1016/J.JEEM.2021.102591
372 L. C. Goveas et al.

2. Timmis, A. J., et al. (2015). Environmental impact assessment of aviation emission reduc-
tion through the implementation of composite materials. International Journal of Life Cycle
Assessment, 20(2), 233–243. https://doi.org/10.1007/S11367-014-0824-0/METRICS
3. Börjesson, P., Björnsson, L., Ericsson, K., & Lantz, M. (2023). Systems perspectives on
combined production of advanced biojet fuel and biofuels in existing industrial infrastruc-
ture in Sweden. Energy Conversion and Management X, 19, 100404. https://doi.org/10.1016/
J.ECMX.2023.100404
4. Lim, J. H. K., et al. (2021). Utilization of microalgae for bio-jet fuel production in the aviation
sector: Challenges and perspective. Renewable and Sustainable Energy Reviews, 149, 111396.
https://doi.org/10.1016/J.RSER.2021.111396
5. Moreno-Gómez, A. L., Gutiérrez-Antonio, C., Gómez-Castro, F. I., & Hernández, S. (2020).
Production of biojet fuel from waste raw materials. Process System Engineering for Biofuels
Development. https://doi.org/10.1002/9781119582694.CH6
6. Koytsoumpa, E. I., Magiri-Skouloudi, D., Karellas, S., & Kakaras, E. (2021). Bioenergy with
carbon capture and utilization: A review on the potential deployment towards a European
circular bioeconomy. Renewable and Sustainable Energy Reviews, 152, 111641. https://doi.
org/10.1016/J.RSER.2021.111641
7. Goveas, L. C., Nayak, S., Vinayagam, R., Loke Show, P., & Selvaraj, R. (2022). Microalgal
remediation and valorisation of polluted wastewaters for zero-carbon circular bioeconomy.
Bioresource Technology, 365, 128169. https://doi.org/10.1016/J.BIORTECH.2022.128169
8. Tiwari, R., et al. (2023). Environmental and economic issues for renewable production of bio-jet
fuel: A global prospective. Fuel, 332, 125978. https://doi.org/10.1016/J.FUEL.2022.125978
9. Tan, E. C. D., & Lamers, P. (2021). Circular bioeconomy concepts—A perspective. Frontiers
in Sustainability. https://doi.org/10.3389/FRSUS.2021.701509
10. Eversberg, D., & Fritz, M. (2022). Bioeconomy as a societal transformation: Mentalities,
conflicts and social practices. Sustainable Production and Consumption, 30, 973–987. https://
doi.org/10.1016/J.SPC.2022.01.021
11. Wei, X., et al. (2022). From biotechnology to bioeconomy: A review of development dynamics
and pathways. Sustainability, 14(16), 10413. https://doi.org/10.3390/SU141610413
12. Wang, X., Khurshid, A., Qayyum, S., & Calin, A. C. (2022). The role of green innovations,
environmental policies and carbon taxes in achieving the sustainable development goals of
carbon neutrality. Environmental Science and Pollution Research, 29(6), 8393–8407. https://
doi.org/10.1007/S11356-021-16208-Z/TABLES/9
13. D’Amato, D., & Korhonen, J. (2021). Integrating the green economy, circular economy and
bioeconomy in a strategic sustainability framework. Ecological Economics, 188, 107143.
https://doi.org/10.1016/J.ECOLECON.2021.107143
14. Škrinjarí, T. (2020). Empirical assessment of the circular economy of selected European coun-
tries. Journal of Cleaner Production, 255, 120246. https://doi.org/10.1016/J.JCLEPRO.2020.
120246
15. D’Urzo, M., & Campagnaro, C. (2023). Design-led repair and reuse: An approach for an
equitable, bottom-up, innovation-driven circular economy. Journal of Cleaner Production,
387, 135724. https://doi.org/10.1016/J.JCLEPRO.2022.135724
16. Carus, M., & Dammer, L. (2018). The circular bioeconomy—Concepts, opportunities, and
limitations. Industrial Biotechnology, 14(2), 83–91. https://doi.org/10.1089/IND.2018.291
21.MCA
17. Sharma, R., & Malaviya, P. (2023). Ecosystem services and climate action from a circular
bioeconomy perspective. Renewable and Sustainable Energy Reviews, 175, 113164. https://
doi.org/10.1016/J.RSER.2023.113164
18. Goh, B. H. H., et al. (2022). Recent advancements in catalytic conversion pathways for synthetic
jet fuel produced from bioresources. Energy Conversion and Management, 251, 114974. https://
doi.org/10.1016/J.ENCONMAN.2021.114974
19. Lahijani, P., Mohammadi, M., Mohamed, A. R., Ismail, F., Lee, K. T., & Amini, G. (2022).
Upgrading biomass-derived pyrolysis bio-oil to bio-jet fuel through catalytic cracking and
hydrodeoxygenation: A review of recent progress. Energy Conversion and Management, 268,
115956. https://doi.org/10.1016/J.ENCONMAN.2022.115956
15 Circular Bioeconomy Approaches for Valorizing Waste Streams … 373

20. Lin, C. H., & Wang, W. C. (2020). Direct conversion of glyceride-based oil into renewable
jet fuels. Renewable and Sustainable Energy Reviews, 132, 110109. https://doi.org/10.1016/J.
RSER.2020.110109
21. Wang, W. C., & Tao, L. (2016). Bio-jet fuel conversion technologies. Renewable and
Sustainable Energy Reviews, 53, 801–822. https://doi.org/10.1016/J.RSER.2015.09.016
22. Wicker, R. J., Kwon, E., Khan, E., Kumar, V., & Bhatnagar, A. (2023). The potential of mixed-
species biofilms to address remaining challenges for economically-feasible microalgal biore-
fineries: A review. Chemical Engineering Journal, 451, 138481. https://doi.org/10.1016/J.CEJ.
2022.138481
23. Hussain, I., Ganiyu, S. A., Alasiri, H., & Alhooshani, K. (2022). A state-of-the-art review on
waste plastics-derived aviation fuel: Unveiling the heterogeneous catalytic systems and techno-
economy feasibility of catalytic pyrolysis. Energy Conversion and Management, 274, 116433.
https://doi.org/10.1016/J.ENCONMAN.2022.116433
24. Collis, J., Duch, K., & Schomäcker, R. (2022). Techno-economic assessment of jet fuel produc-
tion using the Fischer-Tropsch process from steel mill gas. Frontiers in Energy Research, 10,
1049229. https://doi.org/10.3389/FENRG.2022.1049229/BIBTEX
25. Zhu, X., Xiao, J., Wang, C., Zhu, L., & Wang, S. (2022). Global warming potential analysis of
bio-jet fuel based on life cycle assessment. Carbon Neutrality, 1(1), 1–13. https://doi.org/10.
1007/S43979-022-00026-4/TABLES/8
26. Majidian, P., Tabatabaei, M., Zeinolabedini, M., Naghshbandi, M. P., & Chisti, Y. (2018).
Metabolic engineering of microorganisms for biofuel production. Renewable and Sustainable
Energy Reviews, 82, 3863–3885. https://doi.org/10.1016/J.RSER.2017.10.085
27. Moreno-Gómez, A. L., Gutiérrez-Antonio, C., Gómez-Castro, F. I., & Hernández, S. (2021).
Modelling, simulation and intensification of the hydroprocessing of chicken fat to produce
renewable aviation fuel. Chemical Engineering and Processing—Process Intensification, 159,
108250. https://doi.org/10.1016/J.CEP.2020.108250
28. Klein-Marcuschamer, D., et al. (2013). Technoeconomic analysis of renewable aviation fuel
from microalgae, Pongamia pinnata, and sugarcane. Biofuels, Bioproducts and Biorefining,
7(4), 416–428. https://doi.org/10.1002/BBB.1404
29. Chen, R. X., & Wang, W. C. (2019). The production of renewable aviation fuel from waste
cooking oil. Part I: Bio-alkane conversion through hydro-processing of oil. Renewable Energy,
135, 819–835. https://doi.org/10.1016/J.RENENE.2018.12.048
30. Yang, X., et al. (2018). Production of high-density renewable aviation fuel from arid land crop.
ACS Sustainable Chemistry and Engineering, 6(8), 10108–10119. https://doi.org/10.1021/ACS
SUSCHEMENG.8B01433/SUPPL_FILE/SC8B01433_SI_001.PDF
31. Romero-Izquierdo, A. G., Gutiérrez-Antonio, C., Gómez-Castro, F. I., & Hernández, S. (2022).
Synthesis and intensification of a biorefinery to produce renewable aviation fuel, biofuels,
bioenergy and chemical products from Jatropha curcas fruit. IET Renewable Power Generation,
16(14), 2988–3008. https://doi.org/10.1049/RPG2.12388
32. Wang, W. C., Liu, Y. C., & Nugroho, R. A. A. (2022). Techno-economic analysis of renewable
jet fuel production: The comparison between Fischer-Tropsch synthesis and pyrolysis. Energy,
239, 121970. https://doi.org/10.1016/J.ENERGY.2021.121970
33. Zhang, X., et al. (2016). Enhancement of jet fuel range alkanes from co-feeding of lignocel-
lulosic biomass with plastics via tandem catalytic conversions. Applied Energy, 173, 418–430.
https://doi.org/10.1016/J.APENERGY.2016.04.071
34. Diederichs, G. W., Ali Mandegari, M., Farzad, S., & Görgens, J. F. (2016). Techno-economic
comparison of biojet fuel production from lignocellulose, vegetable oil and sugar cane juice.
Bioresource Technology, 216, 331–339. https://doi.org/10.1016/J.BIORTECH.2016.05.090
35. Zhang, Y., et al. (2015). Production of jet and diesel biofuels from renewable lignocellulosic
biomass. Applied Energy, 150, 128–137. https://doi.org/10.1016/J.APENERGY.2015.04.023
36. Crawford, J. T., Shan, C. W., Budsberg, E., Morgan, H., Bura, R., & Gustafson, R. (2016).
Hydrocarbon bio-jet fuel from bioconversion of poplar biomass: techno-economic assessment.
Biotechnology and Biofuels. https://doi.org/10.1186/s13068-016-0545-7
374 L. C. Goveas et al.

37. Li, Y., et al. (2017). Process and techno-economic analysis of bio-jet fuel-range hydro-
carbon production from lignocellulosic biomass via aqueous phase deconstruction and catalytic
conversion. Energy Procedia, 105, 675–680. https://doi.org/10.1016/J.EGYPRO.2017.03.374
38. Hao, J., Xiao, J., Song, G., & Zhang, Q. (2021). Energy and exergy analysis of bio-jet fuel
production from lignocellulosic biomass via aqueous conversion. Case Studies in Thermal
Engineering, 26, 101006. https://doi.org/10.1016/J.CSITE.2021.101006
39. Chen, Y. K., Lin, C. H., & Wang, W. C. (2020). The conversion of biomass into renewable jet
fuel. Energy, 201, 117655. https://doi.org/10.1016/J.ENERGY.2020.117655
40. Ganguly, I., Pierobon, F., Bowers, T. C., Huisenga, M., Johnston, G., & Eastin, I. L. (2018).
‘Woods-to-Wake’ Life Cycle Assessment of residual woody biomass based jet-fuel using mild
bisulfite pretreatment. Biomass and Bioenergy, 108, 207–216. https://doi.org/10.1016/J.BIO
MBIOE.2017.10.041
41. Ringsred, A., van Dyk, S., & Saddler, J. (2021). Life-cycle analysis of drop-in biojet fuel
produced from British Columbia forest residues and wood pellets via fast-pyrolysis. Applied
Energy, 287, 116587. https://doi.org/10.1016/J.APENERGY.2021.116587
42. Moretti, C., López-Contreras, A., de Vrije, T., Kraft, A., Junginger, M., & Shen, L. (2021). From
agricultural (by-)products to jet fuels: Carbon footprint and economic performance. Science of
the Total Environment, 775, 145848. https://doi.org/10.1016/J.SCITOTENV.2021.145848
43. Silva Braz, D., & Pinto Mariano, A. (2018). Jet fuel production in eucalyptus pulp mills:
Economics and carbon footprint of ethanol vs. butanol pathway. Bioresource Technology, 268,
9–19. https://doi.org/10.1016/J.BIORTECH.2018.07.102
44. Goh, B. H. H., et al. (2020). Progress in utilisation of waste cooking oil for sustainable biodiesel
and biojet fuel production. Energy Conversion and Management, 223, 113296. https://doi.org/
10.1016/J.ENCONMAN.2020.113296
45. El-Araby, R., Abdelkader, E., El Diwani, G., & Hawash, S. I. (2020). Bio-aviation fuel via
catalytic hydrocracking of waste cooking oils. Bulletin of the National Research Centre, 44,
1–9. https://doi.org/10.1186/S42269-020-00425-6
46. Zhang, Z., Wei, K., Li, J., & Wang, Z. (2022). Life-cycle assessment of bio-jet fuel production
from waste cooking oil via hydroconversion. Energies, 15(18), 6612. https://doi.org/10.3390/
EN15186612/S1
47. Liu, H., Zhang, C., Tian, H., Li, L., Wang, X., & Qiu, T. (2021). Environmental and techno-
economic analyses of bio-jet fuel produced from Jatropha and castor oilseeds in China. Inter-
national Journal of Life Cycle Assessment, 26(6), 1071–1084. https://doi.org/10.1007/S11367-
021-01914-0/METRICS
48. Rambabu, K., et al. (2023). Sustainable production of bio-jet fuel and green gasoline from date
palm seed oil via hydroprocessing over tantalum phosphate. Fuel, 331, 125688. https://doi.org/
10.1016/J.FUEL.2022.125688
49. Zhao, X., Wei, L., Julson, J., Qiao, Q., Dubey, A., & Anderson, G. (2015). Catalytic cracking of
non-edible sunflower oil over ZSM-5 for hydrocarbon bio-jet fuel. New Biotechnology, 32(2),
300–312. https://doi.org/10.1016/J.NBT.2015.01.004
50. Choi, I. H., Hwang, K. R., Han, J. S., Lee, K. H., Yun, J. S., & Lee, J. S. (2015). The direct
production of jet-fuel from non-edible oil in a single-step process. Fuel, 158, 98–104. https://
doi.org/10.1016/J.FUEL.2015.05.020
51. Roy, P., et al. (2023). Hydrotreatment of pyrolysis bio-oil with non-edible carinata oil and
poultry fat for producing transportation fuels. Fuel Processing Technology, 245, 107753. https://
doi.org/10.1016/J.FUPROC.2023.107753
52. Zhang, L., Butler, T. L., & Yang, B. (2020). Recent trends, opportunities and challenges of
sustainable aviation fuel. In Green energy to sustainability: Strategies for global industries
(pp. 85–110). https://doi.org/10.1002/9781119152057.CH5
53. Why, E. S. K., Ong, H. C., Lee, H. V., Gan, Y. Y., Chen, W. H., & Chong, C. T. (2019).
Renewable aviation fuel by advanced hydroprocessing of biomass: Challenges and perspective.
Energy Convers. Manag., 199, 112015. https://doi.org/10.1016/J.ENCONMAN.2019.112015
54. Dolah, R., Zafar, S., & Hassan, M. Z. (2022) Alternative jet fuels: biojet fuels’ challenges
and opportunities. In Value-Chain Biofuels Fundamentals, Technology and Standardization
(pp. 181–194). https://doi.org/10.1016/B978-0-12-824388-6.00022-1
15 Circular Bioeconomy Approaches for Valorizing Waste Streams … 375

55. Deane, J. P., & Pye, S. (2018). Europe’s ambition for biofuels in aviation—A strategic review
of challenges and opportunities. Energy Strategy Reviews, 20, 1–5. https://doi.org/10.1016/J.
ESR.2017.12.008
56. Zhang, C., Hui, X., Lin, Y., & Sung, C. J. (2016). Recent development in studies of alternative jet
fuel combustion: Progress, challenges, and opportunities. Renewable and Sustainable Energy
Reviews, 54, 120–138. https://doi.org/10.1016/J.RSER.2015.09.056
Chapter 16
Sustainable Bioethanol Production
from the Pretreated Waste
Lignocellulosic Feedstocks

Belete Tessema Asfaw, Meroda Tesfaye Gari, Mani Jayakumar,


and Gurunathan Baskar

Abstract Due to its production from renewable feedstocks, usage as an alternative


fuel, and favorable effects on the environment, bioethanol is one of the most intriguing
biofuels currently on the market. It is currently mostly made using basic ingredients
that contain sugar and starch. However, a variety of readily available lignocellulosic
biomass feedstocks are used to produce bioethanol, including waste from forestry,
agriculture, and the processing of cereals. The complex mixture of carbohydrates
known as lignocellulose needs to undergo an efficient pretreatment practice in order
to generate the pathways necessary for the synthesis of fermentable sugars, which are
subsequently hydrolyzed and fermented into bioethanol. Renewable lignocellulosic
raw materials are cheap feedstocks that do not take away market share from the
food and feed sector, promoting sustainability despite financial and technological
barriers. The procedures for bioethanol extraction and purification, as well as the
creation of bioethanol from renewable raw materials, have also been covered. This
paper examines the developments in the manufacturing of bioethanol as a fuel from

B. T. Asfaw · M. Jayakumar (B)


Department of Chemical Engineering, Haramaya University, Haramaya Institute of Technology,
Haramaya, Ethiopia
e-mail: drjayakumarmani@haramaya.edu.et
B. T. Asfaw · M. T. Gari
College of Biological and Chemical Engineering, Addis Ababa Science and Technology
University, Addis Ababa, Ethiopia
M. Jayakumar
Department of Biotechnology, Faculty of Engineering, Karpagam Academy of Higher Education,
Coimbatore 641 021, Tamilnadu, India
Centre for Natural Products and Functional Foods, Karpagam Academy of Higher Education,
Coimbatore 641 021, Tamilnadu, India
G. Baskar
Department of Biotechnology, St. Joseph’s College of Engineering, Chennai 600119, India
School of Engineering, Lebanese American University, Byblos 1102 2801, Lebanon

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2024 377
G. Baskar et al. (eds.), Circular Bioeconomy Perspectives in Sustainable Bioenergy
Production, Energy, Environment, and Sustainability,
https://doi.org/10.1007/978-981-97-2523-6_16
378 B. T. Asfaw et al.

a variety of renewable feedstocks, as well as the methodologies for purification and


separation, prospects, and challenges.

Keywords Bioethanol · Biofuels · Pretreatment · Sustainable feedstocks · Waste


biomass

16.1 Introduction

Only a small portion of fossil fuels is used to develop new chemicals and materials,
while the majority—roughly 75%—are still used to generate heat and energy. Fossil
fuels are naturally replenished by the carbon cycle at a rate that is significantly
slower than the rate at which they are now used [29]. Fossil fuel production is
also unsustainable because the majority of its reserves are concentrated in a small
number of countries. Additionally, burning fossil fuels and altering land use brought
on by human activities result in higher greenhouse gas emissions, which worsen the
crisis caused by global warming [6, 28, 34]. Much work has gone into creating other
energy sources, such as biofuels like biodiesel and bioethanol, in anticipation of their
unavoidable deaths. The usage of both as liquid fuels, particularly for transportation,
makes them both enticing, however, bioethanol is used more frequently than the other
[54]. Other potential uses for bioethanol include industrial processes and heating.
Finding alternative fuels that are practical from a commercial and environmental
standpoint is a hot topic right now [3]. Biogas, bioethanol, biodiesel, and biobu-
tanol are the most ideal biofuels to replace fossil fuels [66]. Bioethanol is one of the
potential fuel alternatives and is produced from a range of sources [57]. By substi-
tuting bioethanol for gasoline, more than 80% will reduce carbon emissions and
overall gasoline consumption will go down by 30%. One technique for studying the
generation of bioethanol is continuous fermentation [57]. The output of bioethanol
can increase by up to four times when it is produced continuously rather than using
centrifuges and settling tanks [71].
Because they are better sources of the readily available starch and sugar needed
for fermentation, food crops like wheat, corn, potatoes, beets, and sugarcane have
historically been utilized as feedstocks for alcoholic fermentation [11]. But as the
population of the world increases and the amount of arable land decreases, worries
about the production of fuel from food crops have risen [14]. Researchers are looking
into non-edible biomass sources such as wheat straw, grass, crop leftovers, lignocel-
lulosic materials, and algae in order to produce eco-friendly bioethanol [26, 49]. As a
result, a greater variety of feedstock materials can now be used to produce bioethanol.
Technological breakthroughs in the field have made possible the ability to create
ethanol derived from a wider variety of lignocellulosic biomass feedstocks. The
use of fermentation practice now allows for the exploitation of a hitherto untapped
biomass resource to produce bioethanol [69]. This chapter focused to summarize
the current state of knowledge on biomass raw materials feedstocks, bioethanol
synthesis, various by-products and future prospects and challenges were addressed.
16 Sustainable Bioethanol Production from the Pretreated Waste … 379

16.2 Potential Biomass and Its Sources

The preparation of bioethanol is a possibility for various biomass types. Some of


the chosen biomasses can be categorized based on raw materials that comprise
sugar, starch, and lignocellulosic because of the chemical makeup of the carbohy-
drate sources [58]—materials with sugar, including sugar beet, sugarcane, molasses,
cheese processing waste (whey), and sweet sorghum, materials with starch, including
corn, wheat, and cassava; and materials with lignocellulosic biomass, including straw
and leftover crop and agricultural waste [6].

16.2.1 Sugar

The two major sugar-producing plants within the world are sugar beet and sugar
cane. Sugarcane and sugar beet are used to manufacture around 40% of the bioethanol
consumed worldwide [6]. The vast majority of Saccharomyces species produce Inver-
tase enzyme, which is readily capable to hydrolyze them. The chemical Invertase
cleaves the glycosidic bond between the glucose and fructose monomers amid the
primary arrange of alcoholic maturation. In a method known as glycolysis, each
glucose particle is broken into two pyruvate particles [60]. This bioprocess is more
feasible than when producing bioethanol from feedstocks containing starch since
pretreatment is not required when producing bioethanol from feedstocks containing
sugar (sucrose). To extract sugars from sugar crops and turn them into a fermentation
medium, all that is required is a milling process. In this situation, juice or molasses
can be used to directly generate ethanol [73].
Sugarcane is collected in the fields used in the initial stage of ethanol manufac-
ture—the roots, stalk, crowns, and leaves of the sugarcane plant [37]. Since the stalks
contain the majority of the sugars, they are valuable for commercial processing. The
majority of the leaves and tops are cut off the stalks and left on the fields during
mechanical harvest, safeguarding the soil and inhibiting the growth of weeds and
other plant species. A stem plus a straw make up sugarcane (or trash). Before the
cane is ground into juice, the sugarcane stem is removed to produce sugar (sucrose) or
alcohol (ethanol). The waste material produced when the juice from sugarcane stalks
is extracted is known as bagasse. Straw is made up of accessible tops, dry leaves, and
fresh leaves that have not yet been harvested (or garbage) [43]. The tops of the cane
plant have desiccated, often brownish-colored leaves that are located between the top
end and the last stalk node. Fresh leaves are colored green and golden. The leaves
may be utilized for the following purposes: (1) as a fuel, (2) transformation into oil,
gas, and char; and (3) gasification-based conversion to methanol. The tops could
be fed to ruminants dry or fresh, utilized as a raw material to manufacture methane
through anaerobic fermentation, or utilized for energy after having the water content
reduced. Normally, when the crop is harvested, SB and SS are burned in the open
380 B. T. Asfaw et al.

countryside or used as an untapped source of simple sugars that might be converted


into alcohol [6].
Sugar beet is the primary raw material used to make sugar in Europe. Both the
by-products of sugar beet processing, high-purity crystal sugar and thin and thick
juice, can be transformed into bioethanol or other bio-based goods [36]. Raw sugar
beet cassettes are also effective building blocks for the synthesis of bioethanol. The
configuration of the bioprocess, microbiological stability, and transport properties
are all impacted by the employment of intermediates in the sugar processing process.
Granulated sugar and sugar syrup can both be used all year as the primary ingredients
in the synthesis of bioethanol. Additionally, they can act as a base for a number of
chemical end products or intermediates [6].
When fresh juice is turned into sugar, molasses is produced. It is a thick liquid that
can be mechanically divided. It contains larger concentrations of the microorganisms
needed to make alcohol, ethanol, or ethanol fuel. Sugarcane typically yields 23 L of
molasses. The classification process, cane variety, and soil type all affect the molasses
composition [10].
Using the juice created by the extraction process, sugar and ethanol are created.
Molasses, a leftover from the manufacturing of sugar, is combined with the ethanol
juice when preparing the mash for ethanol production [73]. The cogeneration system
uses bagasse from the extraction system to produce steam and power [15]. Significant
amounts of fermentable sugars can be found in molasses that has been diluted (thrice)
with high-quality water and permitted to age while being within the nearness of a
yeast culture (Saccharomyces cerevisiae), either in batches or continuously [52].
Lower orders of sugar (such as pentoses, tetroses, trioses, and dioses, etc.), water
solvent amino acids, lignin, and other natural divisions, etc., remain unfermented
and are left in spent wash after the fermentable sugars (FS) have been synthesized by
the action of yeast as an alcohol rectified spirit (RS) or ethanol. Greater amounts of
organic components undergo a reduction process that results in an unpleasant odor
[10].

16.2.2 Starch

Starch was a high-yield feedstock for ethanol manufacturing that was a polysac-
charide made up only of glucose molecules and used to store energy in plants
((C6 H10 O5 )n, n = 300–360) [75]. Currently, starch from cereal crops and juice or
molasses from sugarcane plants are the two main feedstock types used in industrial
bioethanol production facilities. Sugarcane accounts for about 60% of the world’s
ethanol production, with starchy cereals accounting for the other 40% [4].
Making bioethanol from starch required the following three steps: fermenta-
tion of glucose to produce ethanol and carbon dioxide; hydrolysis of higher sugars
to monosaccharides (like glucose); and product purification. Around 60% of the
bioethanol generated worldwide is made from starch-containing feedstocks [27].
16 Sustainable Bioethanol Production from the Pretreated Waste … 381

16.2.3 Lignocellulose

Due to the renewable nature of lignocellulosic biomass and its lack of competition
with food crops, producing bioethanol from these raw materials is desirable and
sustainable [40]. Additionally, using bioethanol made from lignocellulosic biomass
can significantly reduce greenhouse gas emissions. Agricultural wastes like corn
stover, crop straws, sugar cane bagasse, agricultural and forestry residues, herbal
plants (alfalfa, switchgrass), short-rotation woody crops instead of grass, pulp and
paper mill waste, forestry wastes, cornstalks, waste paper, and paper products, as well
as food industry trash and municipal solid waste incorporate waste [21]. The produc-
tion of bioethanol from these feedstocks may be a preferable and useful alternative
to the disposal of these by-products. Cellulose, hemicellulose, lignin, protein, ash,
and trace amounts of extractives make up lignocellulosic biomass. Due to its avail-
ability, accessibility, and sustainable supply, lignocellulosic biomass is being taken
into consideration as a source of feedstock for the manufacture of bioethanol [13].
Typically, lignin, hemicellulose, and cellulose make up around 10–25%, 20–30%,
and 40–50%, respectively, of the agricultural lignocellulosic biomass [2]. Investiga-
tion of several sugar juices used as feedstocks in ethanol synthesis from sugar crops
are illustrated in Table 16.1.
Lignocellulose is a complicated mixture of carbohydrates that requires effective
pretreatment in order to allow enzymes to produce fermentable sugars, which are
then hydrolyzed and fermented into ethanol [6]. Lignocellulosic resources like wood
and agricultural waste must be converted into sugars by mineral acids [59]. The
four steps that followed one another to create bioethanol from lignocellulose were
pretreatment, hydrolysis, fermentation, and separation/purification [1].

16.3 Pretreatment

The complicated holocellulose polymers have been broken down into straightforward
fermentable sugars using a variety of pretreatment techniques [19]. Pretreatment
practice should obtain lignin for the creation of valuable final products, utilize as
little heat and power as feasible, reduce inhibitor formation, prevent the breakdown
of pentose sugars, and make the process more economical [45]. To get rid of lignin, an
assortment of pretreatment strategies is being utilized, counting physical, chemical,
physicochemical, and organic (biological) ones [22]. The size and crystallinity of
the biomass are reduced by physical pretreatment techniques, including milling and
grinding [61].
382 B. T. Asfaw et al.

Table 16.1 Investigation of several sugar juices used as feedstocks in ethanol synthesis from sugar
crops
Name of the Significant investigation Major accomplishments References
crops
Sweet Estimates of yield, yield Extraction of 38.89 gallons of [25]
sorghum components, and quality were ethanol per ton of stalk
taken into account, and they
varied significantly between
the two seasons
Sugarcane A thermotolerant yeast was Isolation and selection of a [32]
isolated using an enrichment thermotolerant yeast strain from
approach from cane juice for the juice that produced large
the synthesis of ethanol at a amounts of ethanol at both 40
high temperature and 45°
Cells were acclimated to With adapted cells, there was [9]
galactose medium, and juices observed to be more ethanol
were added (30%) than with unadapted cells
Sugar beet Four sugar beet substrates’ Ethanol production (> 40 g/l) [48]
potential is being investigated
Dates Date genetic variations’ juices All cultivars generated ethanol at [72]
were extracted, and a maximum volumetric
fermentation was tested at 300 concentration of 25.0%
˚C and pH 7
Watermelon For maximum sugar recovery, After the concentration [5]
the trials were run at a variety procedure, the final concentrate’s
of transmembrane pressures sugar level dramatically rose
(100–300 kPa)

16.3.1 Physical Pretreatment

By chipping, grinding, shearing, or milling, lignocellulosic biomass can be reduced


to smaller fragments [46]. This increases surface area and reduces particle size,
increasing cellulose conversion by making it easier for cellulases to access the
surface of the biomass [76]. To produce particles with a three- to five-mm-diameter
sieve, primary size reduction is done using hammer mills or Wiley mills. Sonication,
microwave, infrared, gamma radiation, pyrolysis, and radiation with gamma radia-
tion are additional efficient physical therapeutic methods [55]. In order to liberate the
starch that is bound to the corn fiber, achieve homogeneity in the material’s disper-
sion, and achieve a dramatic rise in the rate of enzymatic hydrolysis, mechanical
instruments and processes are commonly employed to cut or grind the corn fiber.
Physical preparation, such as grinding, however, consumes a lot of energy and is not
economically viable [22].
16 Sustainable Bioethanol Production from the Pretreated Waste … 383

16.3.2 Chemical Pretreatment

Among the several pretreatment categories that were initially developed, chemical
pretreatment has received the most research to date. In order to delignify cellu-
losic materials, it has been widely employed [63]. Cellulose and hemicellulose poly-
mers are subjected to chemical hydrolysis, which recovers sugar monomers from
the polymers. This is a vital treatment method for lignocellulosic biomass. The most
common used chemical pretreatments include acid and alkali-based hydrolysis tech-
niques and peroxide [2]. Although it has been found that chemical pretreatments
have greater efficiency compared to non-chemical pretreatments, they do have draw-
backs, including the inability to synthesize new compounds, the need to neutralize
by-products of downstream processes, and environmental concerns. Acid pretreat-
ment is the most widely used chemical procedure for handling lignocellulosic mate-
rials, and sulfuric acid is one of the best solvents for hemicellulose dissolution [79].
During the acid pretreatment, hemicelluloses are broken down into soluble sugars,
which improves cellulose accessibility [22].

16.3.3 Physico-chemical Pretreatments

Physical–chemical preparation techniques are noticeably more efficient than physical


ones. Ozone, acids, alkalis, peroxide, and organic solvents are a few chemical agents
used in these processes [42]. Pretreatment methods used to prepare biomass for
saccharification include autohydrolysis (steam explosion), ammonia fiber explosion
(AFEX), liquid hot water (LHW), steam explosion, steam explosion of SO2 , acid
treatment, and alkali treatment [35, 55].
Regularly utilized to treat lignocellulosic biomass, steam explosion pretreatment
has a reasonable energy need and cost. 177 °C, 5 min, and the best steam explosion
pretreatment conditions for pseudo stem were discovered to be 2.2% H2 SO4 (v/v),
which produced a high total glucose yield of 91.0%. The optimal conditions for
rachis were 198 °C, 5 min, 1.5% H2 SO4 (v/v), and a total glucose yield of 87.1%
[20]. The steam explosion uses high-temperature and high-pressure steam (between
160 and 260 °C) to heat the biomass for a brief period of time (a few seconds to a
few minutes) [22].
By making cellulose more accessible to enzymes than other pretreatment tech-
niques, LHW pretreatment limits the generation of breakdown products that restrict
the growth of microorganisms that cause aeration [41]. It also lowers downstream
pressure. After being processed by liquid hot water and enzyme hydrolysis, the
majority of people think that lignocellulose should be thrown away, burned [80].
384 B. T. Asfaw et al.

16.3.4 Biological Pretreatment

Biological pretreatment requires less energy and chemical input, produces fewer
inhibitors, and uses microorganisms and/or enzymes in a milder environment (around
ambient temperature and pressure) [53]. To be a useful alternative to thermochem-
ical pretreatment, biological pretreatment still needs to be improved in a number
of areas, including incubation time and overall performance (i.e., sugar production)
[74]. By utilizing ligninolytic enzymes or microorganisms that release these enzymes,
biological pretreatment can break down some lignin and alter its structure. In order
to release sugars from holocellulose, hydrolytic enzymes are frequently utilized in
the biorefinery process [31].

16.4 Cereal By-Products Processing for Bioethanol


Production

The following are examples of cereal grains: wheat (Triticum spp.), rye (Secale spp.),
barley (Hordeum spp.), oats, millet (Pennisetum spp.), corn (Zea spp.), sorghum
(Avena spp.), and rice, all members of the Gramineae plant family (Sorghum spp.).
In 2018, the three most popular cereals produced globally were maize (38.7%), rice
and potatoes (26.4%), and wheat (24.8%), together making up 90% of all grain output.
The myriad nutrient-rich chemicals or fractions found in cereal by-products, which
can be exploited as novel materials to produce both food and non-food products, are
still an untapped resource. Beyond the food business, the bioethanol industry provides
a significant alternative to the use of grain by-products [62]. To extract starch from
the substrate, the grains are either placed through dry grinding or wet grinding as the
initial stage in the manufacturing of ethanol. The process of “gelatinization” involves
heating the starch to a high temperature [30].
The main source of cereal by-products is the milling sector. The two most common
techniques for making cereal are dry and wet grinding [18]. The waste products from
the starchy endosperm known as bran, germ, and polish are also eliminated during
dry milling. Wet milling produces bran, steep solids, germ, and gluten (used for oil
extraction), as well as the most intact starch granules conceivable [8]. The procedures
carried out prior to milling produce a by-product known as “grain screens,” which
mostly include all the cereal seeds that don’t meet grading specifications. Along with
the refined cereal flours, additional, cereal germ and bran are produced during the
milling process, with the latter being used to create cereal germ oil. Most by-products
of grain milling are thought to be cereal bran. After the endosperm has been sieved
out of the crushed cereal grains, it comprises the coat seed and aleurone layers that
are still there [47]. Before breaking down into smaller bits and sifting to form distinct
fractions based on size-exclusion separation in sieves, cereals must first be graded
and washed. The grain sizes of the cereals vary depending on the mill used [62].
During the gelatinization process, the starch is heated to a high temperature after
16 Sustainable Bioethanol Production from the Pretreated Waste … 385

Fig. 16.1 Schematic representation of manufacturing ethanol from lignocellulosic feedstocks

milling. After gelatinization, amylolytic enzymes liquefy and saccharify the thick
slurry. Then, they release simple sugars that yeast or other microorganisms use to
perform anaerobic fermentation to create ethanol.
The processes of distillation, rectification, and dehydration separate and concen-
trate the CO2 and ethanol produced. The amount of starch in the substrate, the oper-
ating conditions, and the ethanol production method all affect how much ethanol
is created. An integrated process is being used to boost ethanol output while
reducing production costs and process time, alongside simultaneous saccharifica-
tion and co-fermentation, separate hydrolysis and fermentation, and simultaneous
saccharification and co-fermentation [30]. Overall process depicted in Fig. 16.1.

16.4.1 Corn By-Products

There are two milling processes used for corn: dry and wet grinding. The process of
wet milling softens the grain’s kernels and makes it easier to separate their constituent
parts by soaking them in water and SO2 . Wet milling produces a mixture of gluten,
bran, step particles, and germ as a by-product after removing as many intact starch
granules as possible (for oil extraction) [77]. The most by-products of corn grain
are corn soak alcohol, corn germ supper, maize bran, and corn gluten supper. The
most conclusion items of corn grain are starch (endosperm divisions) and oil (germ)
[8]. Corn steep liquor is a suitable option for fermentation since it includes vitamins,
minerals, and nitrogen (especially amino nitrogen) [67].
Usually, the pericarp is where maize fiber and bran are found (bran), one way
the two differ from one another is that corn bran’s cell walls originate from the
386 B. T. Asfaw et al.

endosperm component as opposed to maize fiber [65]. About 60–70 g/kg and 80–
110 g/kg, respectively, of corn bran and fiber are present in each kilo of maize grain
[8].

16.4.2 Rice By-Products

It is one of the cereal crops that are grown and consumed most commonly around
the globe [70]. Because of its remarkable environmental adaptability and human
success in changing the surrounding agroecosystem, rice can currently be produced
in a broad range of environments and climates. Because of its variety of nutritional
benefits and ability to enhance calorie intake, rice is most frequently ingested in its
cooked form [68].
The majority of the paddy rice’s rice bran, which makes up about 10% of the
total, is used as feedstock, with the remainder going toward the extraction of bran
oil. When brown rice is turned into white rice during the milling process, endosperm
and germ are created in quantities that are mixed with the pericarp and aleurone of
the grain’s outer layers [17]. Rice bran, which is created, contains various levels of
endosperm and germ in the outer layers of brown rice milled into white rice. In terms
of composition, rice bran includes sizeable amounts of oil, dietary fibers, proteins,
and extremely valuable bioactive phytochemicals [16].

16.4.3 Wheat By-Products

Because it performs better theologically than other crops, wheat is the most crucial
crop for manufacturing bread [50]. The most popular industrial method of processing
wheat is known as multiple grinding processes, which are used in roller milling and
sifting procedures to gradually fracture the wheat kernel and extract the starchiest
endosperm possible so as to produce the most flour with a high degree of purity.
The by-products of this process include bran, germ, and aleurone, the outside kernel
layers. Since the commonplace flour extraction rate is lies between the range of 73
and 77%, the amount of milling by-products (23–27%) represents a crucial economic
factor in the production of wheat flour [51].
Wheat bran, which accounts for approximately 25% of the total weight of all
wheat kernels, is a crucial by-product of milling [24]. When making breakfast cereals,
nibble suppers, child nourishment, chilled nourishment, dairy nourishment, sauces,
dressings, and condiments, wheat bran can be utilized as a fixing or as creature
nourish. Wheat bran is widely utilized in food as a result of its technological char-
acteristics, which make it straightforward to experience innovative forms that will
make strides its functionalization and utilize in expansion to its dietary benefits.
Bran is frequently added to cereal-based dishes, which diminishes the final product’s
quality and sensory appeal [23]. Wheat bran can benefit from enzymatic treatment,
16 Sustainable Bioethanol Production from the Pretreated Waste … 387

extrusion, extraction, fermentation, milling, and thermal treatments to improve its


nutritive, sensory, and physical qualities for culinary uses. Wheat bran, however, is a
one of the most significant feedstock for fermentation techniques that result in pure
chemicals because of its composition and accessibility [78] (Table 16.2).

16.5 Applications of Cereal By-Products

The grains’ cereal bran is a reservoir of nutrients. Cereal bran’s chemical makeup
is extremely complex, and by including it in your diet on a regular basis, you may
take advantage of all of its health benefits [56]. As found in people consuming diets
based on cereal grains, it contains a variety of bioactive components, such as dietary
fiber, phytosterols, polyphenols, and phenolic acids, which may provide a range of
biological activity and other health advantages in addition to the typical nutrients
like protein, vitamins, minerals, and fats [33].
Because of its plentiful reserves and low cost of collection and transportation,
maize fiber, a by-product of the corn processing industry, is a feasible raw material for
the production of cellulosic ethanol and value-added commodities [22]. Worldwide,
large amounts of maize are used to make ethanol, but particularly in the USA, where
over 95% of the country’s ethanol is generated along with being used as animal feed
and food for humans. Both wet milling and dry grinding methods can be used to turn
corn into ethanol [39]. It was discovered that rice bran oil, an important by-product of
the extraction of rice bran, is an excellent source of a number of bioactive substances,
including sterols, tocopherols, and tocotrienols. Numerous bioactive compounds with
anti-inflammatory and anti-cancer characteristics may be found in rice bran oil, which
is a great source of this oil [8].

16.6 Future Prospects and Challenges

The parts that came before them have amply shown that bioethanol made from
biomass wastes, such as rice, corn, and sugarcane; sweet sorghum; cassava; sugar
beets and wheat; witch grass; short-rotation woody crops; pulp and paper mill waste;
forestry wastes; cornstalks; waste paper; and so forth, has served as a reliable supplier
for numerous essential applications. However, there is a need for further research
on the synthesis routes of bioethanol from different potential biomasses that require
low-cost production and maximum results.
For the synthesis of bioethanol, there are several pretreatments available, among
them are biological, chemical, physicochemical, and physical techniques. However,
selecting the best one from the ones mentioned is difficult based on different criteria,
such as economic and environmental viewpoints, and selecting the maximum conver-
sion method is a problem. Therefore, future research is mandatory to fulfill the above
388 B. T. Asfaw et al.

Table 16.2 Processing sugar crops and grain by-products’ useful components
Cereal By-product Functional Health benefits Examples References
compounds
Rice Bran and Vitamins, Proteins exhibit 179–389 mg/kg [18]
husk proteins, hypoallergenic of tocopherols
dietary fiber qualities, vitamins and tocotrienols
and oil have antioxidant (compounds
characteristics, made up of
and fiber lowers vitamin E) are
the risk of found in rice
cardiovascular bran
disease
Sorghum Bran Phenolic Policosanoids and Policosanol [7]
and millet compounds, phytosterols, content, hexane
phytosterols which are present extracts
and in phenolic significantly
policosanons compounds, have reduced
the ability to cholesterol
lower cholesterol absorption by up
and act as to 17%
antioxidants
Corn Dietary fiber, Dietary fiber Dietary fiber [22]
Corn fiber helps to improve produced at its
gum, Corn health by highest rate at
fiber oil and preventing and 86.84%,
Xylitol controlling increased the
disease. As a corn fiber gum
result, gum can be yield by 12%
replaced with and results
corn fiber gum in showed that
products after treatment
including with sulfuric
adhesives, acid at 1210 C
thickeners, in comparison to
stabilizers, and untreated maize
emulsifiers. fiber, the oil
Phytosterols concentration in
included in corn the residual
fiber oil reduce solids increased
serum cholesterol from 1.4 to
levels. Xylitol is a 12.2% and the
pricey natural total phytosterol
sweetener with level increased
sweetness from 176.9 to
comparable to 1433.1 mg/g and
sucrose, yet it is the production
safe for diabetics of xylitol was
and does not minimal (0.43 g/
promote tooth g xylose)
damage
(continued)
16 Sustainable Bioethanol Production from the Pretreated Waste … 389

Table 16.2 (continued)


Cereal By-product Functional Health benefits Examples References
compounds
Wheat Bran and Fiber, Bran fiber aids in The prebiotic [44, 64]
germ arabinoxylans the speeding of effects of
intestinal transit arabinoxylans
and the expansion for obesity and
of feces. other metabolic
Arabinoxylans disorders, as
aid in the blood’s well as their
glucose level capacity to
being reduced reduce blood
cholesterol and
the post-prandial
glycemic
response, are
credited with
providing health
benefits
Oat Bran and oat Food fiber that Dietary fiber Oat bran has a [18]
mill scraps is soluble, increases fecal total dietary
β-glucan volume, and fiber content of
studies have 16.0% dry
shown that matter and at
-glucan lowers least 5.5% of
blood cholesterol -glucan per dry
matter
Barley Spent grains β-glucan and Dietary fiber Phytosterols in [38]
dietary fiber lowers blood barley oils are
sugar spikes after able to at safe
meals, which is a dosages,
result of -glucan dramatically
reduced levels of
low-density
lipoprotein
cholesterol
Rye Bran Flavonoids, The addition of In rye grain has [12]
coumarins, grain flour and been found to be
lignans, bran to food between 65 and
stilbenes, and products 300 mg/100 g of
condensed enhances the grains
tannins health-promoting
properties of
those goods as
they are sources
of numerous
beneficial
bioactive
components
390 B. T. Asfaw et al.

criteria. Additionally, a thorough analysis of the two or more pretreatment tech-


niques combined is advised for maximal biomass breakdown. Additionally, focused
efforts are required to address the need for low-cost, economically viable biomass
pretreatment techniques. It is recommended to regulate energy-intensive preparation
procedures.

16.7 Conclusions

It is clear from the supplied data that bioethanol can be a different approach to the
current fuel problem. In recent decades, significant advancements have been made
in the pretreatment of renewable biomass, the production of cellulase, the process of
separating and purifying bioethanol, and the co-fermentation of sugars (pentose and
hexose). However, depending on the cost of manufacture alone, bioethanol is still
more expensive than fossil fuels, excluding sugar cane-based bioethanol synthesis.
The major issue is still how to make bioethanol cheaper to produce. The biore-
finery concept is necessary to reduce the cost of producing bioethanol by more
completely utilizing renewable feedstocks and producing extra value-added by-
products (such as bio-based components from lignin). As a result, bioethanol will be
more economically competitive.

References

1. Amornraksa, S., et al. (2020). Systematic design of separation process for bioethanol production
from corn Stover. BMC Chemical Engineering, 2(1), 1–16.
2. Anwar, Z., Gulfraz, M., & Irshad, M. (2014). Agro-industrial lignocellulosic biomass a key
to unlock the future bio-energy: A brief review. Journal of Radiation Research and Applied
Sciences, 7(2), 163–173.
3. Asghari, M., & Al-e, S. M. J. M. (2021). Green vehicle routing problem: A state-of-the-art
review. International Journal of Production Economics, 231, 107899.
4. Balat, M., & Balat, H. (2009). Recent trends in global production and utilization of bio-ethanol
fuel. Applied Energy, 86(11), 2273–2282.
5. Bhattacharjee, C., Saxena, V. K., & Dutta, S. (2017). Watermelon juice concentration using
ultrafiltration: Analysis of sugar and ascorbic acid. Food Science and Technology International,
23(7), 637–645.
6. Bušić, A., et al. (2018). Bioethanol production from renewable raw materials and its separation
and purification: A review. Food Technology and Biotechnology, 56(3), 289–311.
7. Carr, S. N., et al. (2005). Effects of different cereal grains and ractopamine hydrochloride on
performance, carcass characteristics, and fat quality in late-finishing pigs. Journal of Animal
Science, 83(1), 223–230.
8. Dapčević-Hadnađev, T., Hadnađev, M., & Pojić, M. (2018). The healthy components of cereal
by-products and their functional properties. In Sustainable recovery and reutilization of cereal
processing by-products (pp. 27–61). Elsevier.
9. Dhaliwal, S. S., et al. (2011). Enhanced ethanol production from sugarcane juice by galac-
tose adaptation of a newly isolated thermotolerant strain of Pichia kudriavzevii. Bioresource
Technology, 102(10), 5968–5975.
16 Sustainable Bioethanol Production from the Pretreated Waste … 391

10. Dotaniya, M. L., et al. (2016). Use of sugarcane industrial by-products for improving sugarcane
productivity and soil health. International Journal of Recycling of Organic Waste in Agriculture,
5, 185–194.
11. Duraisam, R., Salelgn, K., & Berekete, A. K. (2017). Production of beet sugar and bio-ethanol
from sugar beet and it bagasse: A review. International Journal of Engineering Trends and
Technology (IJETT), 43(4), 222–233.
12. Dziki, D. (2022). Rye flour and rye bran: New perspectives for use.
13. Edeh, I. (2021). Bioethanol production: An overview. In Bioethanol technologies (p. 1).
14. Edgerton, M. D. (2009). Increasing crop productivity to meet global needs for feed, food, and
fuel. Plant Physiology, 149(1), 7–13.
15. Ensinas, A. V., et al. (2009). Reduction of irreversibility generation in sugar and ethanol
production from sugarcane. Energy, 34(5), 680–688.
16. Fărcaș, A., et al. (2021). Cereal processing by-products as rich sources of phenolic compounds
and their potential bioactivities. Nutrients, 13(11), 3934.
17. Ferreira, S. C., et al. (2018). New functional ingredients from agroindustrial by-products for
the development of healthy foods. In Encyclopedia of food security and sustainability. Elsevier.
18. Galanakis, C. M. (2022). Sustainable applications for the valorization of cereal processing
by-products. Foods, 11(2), 1–15. https://doi.org/10.3390/foods11020241
19. Gómez Millán, G., et al. (2019). Recent advances in the catalytic production of platform
chemicals from holocellulosic biomass. ChemCatChem, 11(8), 2022–2042.
20. Guerrero, A. B., Ballesteros, I., & Ballesteros, M. (2017). Optimal conditions of acid-
catalysed steam explosion pretreatment of banana lignocellulosic biomass for fermentable
sugar production. Journal of Chemical Technology & Biotechnology, 92(9), 2351–2359.
21. Gundupalli, M. P., et al. (2022). Hydrothermal liquefaction of lignocellulosic biomass for
production of biooil and by-products: current state of the art and challenges. In Biofuels and
bioenergy (pp. 61–84).
22. Guo, Y., et al. (2022). Production of cellulosic ethanol and value-added products from corn
fiber. Bioresources and Bioprocessing, 9(1), 1–18.
23. Heiniö, R. L., et al. (2016). Sensory characteristics of wholegrain and bran-rich cereal foods—
A review. Trends in Food Science and Technology, 47, 25–38. https://doi.org/10.1016/j.tifs.
2015.11.002
24. Higuchi, M. (2014). Antioxidant properties of wheat bran against oxidative stress. In Wheat
and rice in disease prevention and health (pp. 181–199). Elsevier.
25. Hunsigi, G., Yekkeli, N. R., & Kongawad, B. Y. (2010). Sweet stalk sorghum: An alternative
sugar crop for ethanol production. Sugar Tech, 12(1), 79–80.
26. Jayakumar, M., Kuppusamy Vaithilingam, S., Karmegam, N., & Jabesa, A. (2022). Bioethanol
production from green biomass resources: Emerging technologies. In Encyclopedia of green
materials (pp. 1–12).
27. Jayakumar, M., Kuppusamy Vaithilingam, S., Karmegam, N., Gebeyehu, K. B., et al.
(2022). Fermentation technology for ethanol production: current trends and challenges. In B.
Gurunathan & R. B. T.-B. Sahadevan (Eds.), Biofuels and bioenergy (pp. 105–131). Elsevier.
https://doi.org/10.1016/b978-0-323-90040-9.00015-1
28. Jayakumar, M., Gebeyehu, K. B., et al. (2022). Waste ox bone based heterogeneous cata-
lyst synthesis, characterization, utilization and reaction kinetics of biodiesel generation from
Jatropha curcas oil. Chemosphere, 288, 132534.
29. Jayakumar, M., et al. (2023). Bioethanol production from agricultural residues as lignocellu-
losic biomass feedstock’s waste valorization approach: A comprehensive review. In Science of
the total environment (p. 163158).
30. Jeevan Kumar, S. P., Sampath Kumar, N. S., & Chintagunta, A. D. (2020). Bioethanol produc-
tion from cereal crops and lignocelluloses rich agro-residues: Prospects and challenges. SN
Applied Sciences, 2(10), 1673.
31. Keshav, P. K., et al. (2021). Lignocellulosic ethanol production from cotton stalk: an overview
on pretreatment, saccharification and fermentation methods for improved bioconversion
process. In Biomass conversion and biorefinery (pp. 1–17).
392 B. T. Asfaw et al.

32. Limtong, S., Sringiew, C., & Yongmanitchai, W. (2007). Production of fuel ethanol at high
temperature from sugar cane juice by a newly isolated Kluyveromyces marxianus. Bioresource
Technology, 98(17), 3367–3374.
33. Luithui, Y., Baghya Nisha, R., & Meera, M. S. (2019). Cereal by-products as an important
functional ingredient: Effect of processing. Journal of Food Science and Technology, 56, 1–11.
34. Machado, C. F. R., et al. (2018). Carbon dioxide and ethanol from sugarcane biorefinery as
renewable feedstocks to environment-oriented integrated chemical plants. Journal of Cleaner
Production, 172, 1232–1242.
35. Mankar, A. R., et al. (2021). Pretreatment of lignocellulosic biomass: A review on recent
advances. Bioresource Technology, 334, 125235.
36. Marzo, C., et al. (2019). Status and perspectives in bioethanol production from sugar beet. In
Bioethanol production from food crops (pp. 61–79). Elsevier.
37. de Matos, M., Santos, F., & Eichler, P. (2020). Sugarcane world scenario. In Sugarcane
biorefinery, technology and perspectives (pp. 1–19). Elsevier.
38. Moreau, R. A., Flores, R. A., & Hicks, K. B. (2007). Composition of functional lipids in hulled
and hulless barley in fractions obtained by scarification and in barley oil. Cereal Chemistry,
84(1), 1–5. https://doi.org/10.1094/CCHEM-84-1-0001
39. Mosier, N. S., & Ileleji, K. E. (2020). How fuel ethanol is made from corn. In Bioenergy
(pp. 539–544). Elsevier.
40. Mujtaba, M., et al. (2023). Lignocellulosic biomass from agricultural waste to the circular
economy: A review with focus on biofuels, biocomposites and bioplastics. Journal of Cleaner
Production, 136815.
41. Nahak, B. K., et al. (2022). Advancements in net-zero pertinency of lignocellulosic biomass for
climate neutral energy production. Renewable and Sustainable Energy Reviews, 161, 112393.
42. Naji, S. Z., & Tye, C. T. (2022). A review of the synthesis of activated carbon for biodiesel
production: Precursor, preparation, and modification. Energy Conversion and Management: X,
13, 100152.
43. Negrão, D. R., et al. (2021). Inorganics in sugarcane bagasse and straw and their impacts for
bioenergy and biorefining: A review. Renewable and Sustainable Energy Reviews, 148, 111268.
44. Onipe, O. O., Jideani, A. I. O., & Beswa, D. (2015). Composition and functionality of wheat
bran and its application in some cereal food products. International Journal of Food Science
and Technology, 50(12), 2509–2518. https://doi.org/10.1111/ijfs.12935
45. Owusu, P. A., & Asumadu-Sarkodie, S. (2016). A review of renewable energy sources,
sustainability issues and climate change mitigation. Cogent Engineering, 3(1), 1167990.
46. Oyedeji, O., et al. (2020). Understanding the impact of lignocellulosic biomass variability
on the size reduction process: A review. ACS Sustainable Chemistry & Engineering, 8(6),
2327–2343.
47. Papageorgiou, M., & Skendi, A. (2018). Introduction to cereal processing and by-products. In
Sustainable recovery and reutilization of cereal processing by-products (pp. 1–25). https://doi.
org/10.1016/B978-0-08-102162-0.00001-0
48. Park, S. C., & Baratti, J. (1991). Comparison of ethanol production by Zymomonas mobilis
from sugar beet substrates. Applied Microbiology and Biotechnology, 35, 283–291.
49. Pensri, B., et al. (2016). Potential of fermentable sugar production from Napier cv. Pakchong
1 grass residue as a substrate to produce bioethanol. Energy Procedia, 89, 428–436.
50. Podolska, G., Aleksandrowicz, E., & Szafrańska, A. (2020). Bread making potential of Triticum
aestivum and Triticum spelta species. Open Life Sciences, 15(1), 30–40.
51. Prueckler, M., et al. (2014). Wheat bran-based biorefinery 1: Composition of wheat bran and
strategies of functionalization. LWT-Food Science and Technology, 56(2), 211–221.
52. Raspor, P., & Goranovič, D. (2008). Biotechnological applications of acetic acid bacteria.
Critical Reviews in Biotechnology, 28(2), 101–124.
53. Rastogi, M., & Shrivastava, S. (2017). Recent advances in second generation bioethanol produc-
tion: An insight to pretreatment, saccharification and fermentation processes. Renewable and
Sustainable Energy Reviews, 80, 330–340.
16 Sustainable Bioethanol Production from the Pretreated Waste … 393

54. Rony, Z. I., et al. (2023). Alternative fuels to reduce greenhouse gas emissions from marine
transport and promote UN sustainable development goals. Fuel, 338, 127220.
55. Saini, J. K., Saini, R., & Tewari, L. (2015). Lignocellulosic agriculture wastes as biomass
feedstocks for second-generation bioethanol production: Concepts and recent developments.
3 Biotech, 5, 337–353.
56. Seal, C. J., et al. (2021). Health benefits of whole grain: Effects on dietary carbohydrate quality,
the gut microbiome, and consequences of processing. Comprehensive Reviews in Food Science
and Food Safety, 20(3), 2742–2768.
57. Sebayang, A. H., et al. (2016). A perspective on bioethanol production from biomass as
alternative fuel for spark ignition engine. RSC Advances, 6(18), 14964–14992.
58. Selvakumar, P., et al. (2022). Optimization of binary acids pretreatment of corncob biomass
for enhanced recovery of cellulose to produce bioethanol. Fuel, 321, 124060. https://doi.org/
10.1016/j.fuel.2022.124060
59. Shrestha, S., et al. (2021). Different facets of lignocellulosic biomass including pectin and its
perspectives. Waste and Biomass Valorization, 12, 4805–4823.
60. Siddiqui, A. J., et al. (2022). Enzymes in food fermentations. In African fermented food
products-new trends (pp. 101–133). Springer.
61. Sitotaw, Y. W., et al. (2021). Ball milling as an important pretreatment technique in
lignocellulose biorefineries: a review (pp. 1–24). In Biomass conversion and biorefinery.
62. Skendi, A., et al. (2020). Advances on the valorisation and functionalization of by-products
and wastes from cereal-based processing industry. Foods, 9(9), 1243.
63. Stanley, J. T., et al. (2022). Potential pre-treatment of lignocellulosic biomass for the
enhancement of biomethane production through anaerobic digestion—A review. Fuel, 318,
123593.
64. Stevenson, L., et al. (2012). Wheat bran: Its composition and benefits to health, a European
perspective. International Journal of Food Sciences and Nutrition, 63(8), 1001–1013. https://
doi.org/10.3109/09637486.2012.687366
65. Sui, W., et al. (2018). Effect of wheat bran modification by steam explosion on structural
characteristics and rheological properties of wheat flour dough. Food Hydrocolloids, 84, 571–
580.
66. Tabatabaei, M., et al. (2015). Renewable energy and alternative fuel technologies. In BioMed
research international.
67. Tan, J. P., et al. (2016). Use of corn steep liquor as an economical nitrogen source for biosuc-
cinic acid production by Actinobacillus succinogenes. In IOP conference series: earth and
environmental science (p. 12058). IOP Publishing.
68. Tegegne, B., Belay, A., & Gashaw, T. (2020). Nutritional potential and mineral profiling of
selected rice variety available in Ethiopia. Chemistry International, 6(1), 21–29.
69. Tse, T. J., Wiens, D. J., & Reaney, M. J. T. (2021). Production of bioethanol—A review of
factors affecting ethanol yield. Fermentation, 7(4), 268. https://doi.org/10.3390/fermentation
7040268
70. Wang, J., et al. (2018). Effect of climate change on the yield of cereal crops: A review. Climate,
6(2), 41.
71. Watt, S., et al. (2010). Analysis of a model for ethanol production through continuous
fermentation: Ethanol productivity. International Journal of Chemical Reactor Engineering,
8(1).
72. Zabed, H., et al. (2014). Bioethanol production from fermentable sugar juice. The Scientific
World Journal.
73. Zabed, H., et al. (2017). Bioethanol production from renewable sources: Current perspectives
and technological progress. Renewable and Sustainable Energy Reviews, 71, 475–501.
74. Zabed, H. M., et al. (2019). Recent advances in biological pretreatment of microalgae and
lignocellulosic biomass for biofuel production. Renewable and Sustainable Energy Reviews,
105, 105–128.
75. Zabochnicka-Świątek, M., & Sławik, L. (2010). Bioethanol-production and utilization.
Archivum Combustionis, 30(3), 237–246.
394 B. T. Asfaw et al.

76. Zhang, H., Han, L., & Dong, H. (2021). An insight to pretreatment, enzyme adsorption and enzy-
matic hydrolysis of lignocellulosic biomass: Experimental and modeling studies. Renewable
and Sustainable Energy Reviews, 140, 110758. https://doi.org/10.1016/j.rser.2021.110758
77. Zhang, R., et al. (2021). Comprehensive utilization of corn starch processing by-products: A
review. Grain & Oil Science and Technology, 4(3), 89–107.
78. Zhao, H.-M., Guo, X.-N., & Zhu, K.-X. (2017). Impact of solid state fermentation on nutritional,
physical and flavor properties of wheat bran. Food Chemistry, 217, 28–36.
79. Zhou, Z., et al. (2018). Lignocellulosic biomass to biofuels and biochemicals: A comprehen-
sive review with a focus on ethanol organosolv pretreatment technology. Biotechnology and
Bioengineering, 115(11), 2683–2702.
80. Zhuang, X., et al. (2016). Liquid hot water pretreatment of lignocellulosic biomass for
bioethanol production accompanying with high valuable products. Bioresource Technology,
199, 68–75.
Chapter 17
Waste Biomass Supply Chain
for Sustainable Bioenergy Production

C. Nirmala, M. Sridevi, P. Loganathan, Mani Jayakumar,


and Gurunathan Baskar

Abstract The advancement of ecologically benign, renewable, and sustainable alter-


native energy sources is vital for the current global economic theories. The largest
source of renewable energy, bioenergy podia advance the economy and the circular
economy by serving as a strategic component. To fulfill the expanding demands
of humanity, biomass materials such as agricultural products, alcohol fuels indus-
trial waste, landfill gas, solid waste, and wood have attracted interest for use in the
synthesis of biofuels. The energy generated from biomass is less harmful to the envi-
ronment, inexpensive, readily accessible locally in great quantities, and offers job
opportunities for employees in diverse global boundaries. Sustainability, the decline

C. Nirmala
Department of Biotechnology, Paavai Engineering College, Paavai Institutions, Namakkal, Tamil
Nadu, India
M. Sridevi
Department of Biotechnology, Vinayaka Missions Kirupananda Variyar Engineering College,
Vinayaka Missions Research Foundation (Deemed to be University), Salem, Tamil Nadu, India
P. Loganathan
Department of Electrical and Electronics Engineering, Vinayaka Missions Kirupananda Variyar
Engineering College, Vinayaka Missions Research Foundation (Deemed to be University), Salem,
Tamil Nadu, India
M. Jayakumar (B)
Department of Chemical Engineering, Haramaya Institute of Technology, Haramaya University,
Dire Dawa, Ethiopia
e-mail: drjayakumarmani@haramaya.edu.et
Department of Biotechnology, Faculty of Engineering, Karpagam Academy of Higher Education,
Coimbatore 641 021, Tamilnadu, India
Centre for Natural Products and Functional Foods, Karpagam Academy of Higher Education,
Coimbatore 641 021, Tamilnadu, India
G. Baskar
Department of Biotechnology, St. Joseph’s College of Engineering, Chennai 600119, India
School of Engineering, Lebanese American University, Byblos 1102 2801, Lebanon

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2024 395
G. Baskar et al. (eds.), Circular Bioeconomy Perspectives in Sustainable Bioenergy
Production, Energy, Environment, and Sustainability,
https://doi.org/10.1007/978-981-97-2523-6_17
396 C. Nirmala et al.

in emissions of greenhouse gases and an upsurge in manufacturing jobs are the bene-
fits of using renewable sources to manufacture energy, chemicals, and fuels. These
elements promote the development of local economic and social systems. The current
chapter focuses on the availability of numerous waste biomass materials and their
characteristics as viable renewable energy sources. The analysis also includes the
current state of the world’s energy supply, the need for sustainable alternative energy
sources, established and developing technologies for biomass conversion, and the
circular economy of waste biomass management. From the review, it is possible
to infer that utilizing cost-effective sources for implementation in energy platforms
and biofuels will increase the competitiveness of bio-based companies and boost the
economics of existing value chains. Understanding the characteristics of biomass
materials enables us to determine the best approach to use them for energy in the
future and cut down on our reliance on fossil fuels.

Keywords Biomass · Waste · Bioenergy · Pretreatment · Conversion ·


Technologies

17.1 Introduction

Around the world, there is a continually growing demand for bioenergy (biofuels,
heat, and power), which makes up about one-tenth of the total primary energy supply
[1]. At the global scale, there is a steady increase in the population, and by 2050, it is
foreseen that it reaches a maximum of 9.7 billion. The sinewy rise in population leads
to rise in the global energy demand with a yearly surge of 1.5%. Currently, the fuels
from fossil origin provide exceptionally 84% of the world’s energy needs, which has
an adverse milieu effect [2]. The majority of nations promote the use of renewable
energy such as hydro, wind, solar, and bioenergy to meet energy requirement with
lower greenhouse gas discharge. Such issues consequently prompted the prudent use
of accessible earthy and waste materials for bioenergy production [3]. The current
global energy system depends extensively on energy derived from biomass, such as
living beings, wood, food crops like maize, energy crops, and agricultural residues,
tangential residuals such as manure, wastes from food process, forests, yards, and
farms [4, 5].
Utilizing diverse forms of biomass, biorefinery operate in an environmentally
responsible manner and create a variety of biofuels, chemicals, and bioproducts [6].
The general cycle of biomass energy is shown in Fig. 17.1. Based on the feedstocks
utilized in their manufacture, the biofuels are categorized as first, second, and third
generation. The biofuels at initial phases (first generation) are produced using the
plants edible sections as a raw material. The manufacturing of second phase biofuels
uses lignocellulosic resources. Microalgae are utilized as a feedstock to manufacture
the third phase biofuels. First-generation biofuel manufacture is well-developed and
known technology. However, because of the economic and environmental constraints,
their manufacture has come under scrutiny. Plantations with a high rate of biomass
17 Waste Biomass Supply Chain for Sustainable Bioenergy Production 397

Fig. 17.1 Cycle of biomass energy

growth on non-arable land use agro-industrial waste products such as group 1 terres-
trial biomass feedstock’s (sugarcane, corn grain, oil seed, soy bean, etc.,) and group
2 feedstocks that are neither starchy, edible, or food-related like cellulosic biomass
have less impact on the availability of food and can significantly reduce carbon
emissions [7].
In order to attain environmental sustainability for the production of bioenergy, the
biomass supply chain would need to be thoughtfully and sensibly designed in order
to accommodate the still-evolving but promising idea of lignocellulosic and algal
biorefineries. Although there is enough biomass potential in the world to meet the
expanding demand, the distribution of different biomass resources is uneven. The
potential annual supply of biomass on a global scale is anticipated between 97 and
147 oxyjoules (EJ) by 2030 that accounts for about 40% agricultural trash and 37–
66 EJ of debris. The potential supply still available for various vegetation products
varies, including 24–43 EJ of forest residues, and 33–39 EJ of energy crops [8].

17.1.1 Present Global Bioenergy Framework

The biomass supply chain encompasses multiple stages in the bioenergy production
process with numerous alternate approaches [9]. The logistics of biomass, the gener-
ation of biomass as a feedstock for bioenergy, the transition of biomass to energy, and
398 C. Nirmala et al.

the distribution of bioenergy or its carriers for use are all different parts of the biomass
supply chain. Rational design and comprehensive assimilation of various biomass
supply chain activities are likely to increase the amount of energy returned, amelio-
rate the greenhouse gas equilibrium, and decrease the water footprint of the bioenergy
producing facility [10]. Current biomass policy, which is for the most part persuaded
by divided commitments to overarching segment goals, with climate alter and the
vitality industry ordinarily being the driving ones, falls flat to completely capture
the benefits of biomass esteem chains inside their territorial or national setting [11,
12]. Suboptimal interventions as a result of this process have prevented the market
from adopting solutions that are well-unified, efficiently used resources, and domes-
tically sourced with regional differences. Modern, sustainable biomass technologies
are expected to play a significant part in efforts to bipartite the amount of renewable
energies, according to REmap 2030, the worldwide strategy created by the Inter-
national Renewable Energy Agency (IRENA). Supply and demand forecasts for
bioenergy globally through 2030 looks at the potentials for many technologies and
fields, as well as the renewable energy resource to be scaled up. By 2030, biomass
use may double from its current level and account for 108 EJ, or 60% of the total
net consumption of renewable energy and 20% of all original energy supply, if all
technology options envisioned in the REmap analysis are implemented [13].
Currently, two-thirds (35 EJ) of biomass consumption are accounted by conven-
tional space heating and cooking techniques like burning logs [14]. This would be
replaced by contemporary biomass which would include significantly higher shares
for power and transportation applications. Up to 41 EJ of heat might be used by
constructions and commercial enterprise, with only 6 EJ coming from conventional,
less eco-friendly uses, accounting for 36 EJ, or thirty percent of all biomass used in
2030, 36 EJ for power and district heating, and 31 EJ (nearly 29%) for transportation.
Policies pertaining to biofuels are crucial for the growth of the energy sector, partic-
ularly in emerging nations. A number of countries have also conducted important
policy-related actions in the last few decades to support the adoption and develop-
ment of bioenergy. Policies affecting a variety of industries, including agriculture,
research, industry, and commerce, have a substantial impact on how profitable is the
bioenergy and biofuel production [15]. Given the variety of tools available for imple-
menting policy such as subsidies, taxes, etc. and how it is employed, it is challenging
to recognize germane approaches and measure their exact impacts [16].
The biomass policy framework combines modeling, analytical, and participatory
methods and requires substantial involvement from all parties engaged at every level
of the value chain [17]. In order to build and maintain market acceptance, it is
necessary to evaluate difficulties and gaps in the availability of feedstock, resource
competency, and legitimacy across the biomass supply chain. Their participation
diligently pertains in the managerial process and makes it easier for particular actions
and planned interventions to be communicated [18].
Creating a reliable and effective supply network governance system for the
production of resilient bioenergy and biofuels is one of the primary problems.
Although there are several research projects on the production of bioenergy and
biofuels, there are few commercial cases for their adoption in developing nations like
17 Waste Biomass Supply Chain for Sustainable Bioenergy Production 399

India. Within the current investigation, the preferences for pretreatment of lignocel-
lulosic biomass (LB) for the creation of bioenergy are assessed. The technological
viability and challenges carried out for improving the biorefinery processing from
biomass and biowaste supply chain are discussed.

17.2 Sources and Classification of Waste Biomass

Waste biomasses, which incorporate rural waste, wood squander, food squander,
metropolitan strong waste, sewage slop squander, lignocellulosic particulates
includes rural and backwoods buildups, seaweeds, crops residues, fertilizer, food
misspend, and natural part of metropolitan strong waste had an enormous flair to give
energy and worth add upped items by means of various transformation innovations
(Fig. 17.2). These squanders are inescapable on the grounds that they are produced
straightforwardly or by implication by giving the human culture food, energy, and
water. One of the foremost imperative viewpoints of accomplishing an economical
future is viable administration of these unavoidable biomass squanders. Numerous
non-edible oil seed crops are utilized for bioenergy production [19, 20]. Carboniza-
tion, catalytic forms, gasification, liquefaction, pyrolysis, transesterification, and
torrefaction are cases of thermochemical forms. On the other hand, biomethanation,
fermentation, and enzymatic forms are common natural advances for changing over
biomass [20]. Chemical and enzymatic methods are used for hydrolysis of biomass
residues to release fermentable sugars, which are then converted by microorganisms
into of high-value goods such as biofuels and biochemicals. To make biofuels and
biochemical structural blocks, the feedstocks go through a heated breakdown of the
natural components, paying little attention to microbial fermentable materials (i.e.,
saccharides) and non-fermentable portions (i.e., lignin) [21]. In comparison to natural
change processes, thermochemical strategies react quicker due to the joining of tall
temperatures, weights, and impetuses [22].
400 C. Nirmala et al.

Fig. 17.2 Categorization of


biogenic materials and waste
biomass

The synthesis and features of biomass find out the interaction boundaries, response
rate, yields, and character of the transformation items in general. The primary features
of biomass are not completely resolved by general along with extreme arrangement
[23].

17.2.1 Lignocellulosic Biomass

A feedstock that exists as carbon–neutral, viable, and economical is LB, which


typically contains lignin (10.24 wt%), hemicellulose (20.40 wt%), and cellulose
(40.60 wt%) [24]. Crop residues from agriculture, biomass from forestry, and energy
crops are all examples of LB sources. Rural squanders are the byproducts of the
collecting and preparing of rural vegetative crops. Crop wastes are typically fall into
two distinct categories: field residues and residues from agro-food processing. These
agro-deposits do not contain food; thus, they do not compete with the food supply or
ripe agricultural land. Despite its favorable benefits on society, the environment, and
profitability in a range of valuable products for the domestic, industrial, building, and
power industries, the bulk of forestry biomass is still poorly used. The availability
and expense of some of the barriers that restrict the use of agricultural and forestry
biomass for the production of bioenergy are correlated with their seasonal, regional,
and climatic fluctuations. Corn stover, sugarcane bagasse, rice straw, and wheat straw
are the most common crop wastes, accounting for 731.3 million tons, 354.3 million
tons, 203.6 million tons, and 180.7 million tons, respectively [25]. In the agricultural
waste (AW) category, these agricultural residues are the most common LBs. As a rule,
17 Waste Biomass Supply Chain for Sustainable Bioenergy Production 401

scientists would utilize the two terms “horticultural buildup” and “farming waste”
reciprocally [26–28].
Energy crops have prompted a rise in research interest globally due to their diver-
sity, quick development, high rate of production, ability to fix CO2 during photo-
synthesis, cost-effectiveness, and capability to thrive on unfavorable soils [29]. As
a result, their cultivation may have a significant impact on the biofuel industry’s
ability to meet the demand for clean energy. Most common energy crops incorporate
miscanthus, elephant grass, crossover poplar, switch grass, and timothy grass.

17.2.2 Oilseed Crops

Plant derived oils have historically been a major agricultural product. Oil can
be found in the form of triglycerides, triacylglycerol, lipids, and fatty acids in a
number of plant species [30]. As stores of energy and carbon for the improvement
of seedlings, these elements are stored in cells and plant seeds. Triacylglycerol’s
structural resemblance to long-chain hydrocarbons serves as the basis for a prac-
tical substitute for goods that use hydrocarbons as their primary ingredient. Addi-
tionally, due to the physicochemical characteristics of fatty acids, non-edible plant
derived oils are utilized to create biodiesel and are essential constituents of paints,
varnishes, lubricants, and inks [31]. Plant oils are in high demand across a variety
of sectors, including agriculture, nutraceuticals, cosmetics, pharmaceuticals, food,
and biorefining. Mustard oil, olive oil, almond oil, corn oil, canola oil, coconut oil,
grapeseed oil, sunflower oil, soybean oil, and vegetable oil are just a few of the
plant derived oils used in cooking and food preparation [32]. The following plants
are used to make non-edible plant derived oils: microalgae, silk cotton tree, rubber
seed, Azadirachta indica (Neem), Ailanthus altissima (heaven tree), Madhuca longi-
folia (Mahua), Jatropha, Pongamia pinnata (Karanja), Linum usitatissimum (flax),
Ricinus communis (Castor), Sapindus mukorossi (Soapnut), Toona sinensis (Juss),
Vernicia fordii (Tung), etc. [33]. Different oilseed crops exhibit noticeable variations
in the oil percentage. The oilseed plants might potentially be genetically modified to
produce oils with a higher concentration, it can be deduced [34]. Oil seeds may be
used to produce biofuels and biodiesel, which is a great alternative to petroleum-based
energy [35].

17.2.3 Municipal Solid Waste

Municipal solid waste (MSW) refers to waste products collected from homes and
municipalities (such as rubbish, recyclable and non-recyclable leftovers). MSW
might also contain industrial, commercial, and institutional (ICI) garbage, which
includes leftovers from enterprises, big industries, hospitals, and educational institu-
tions. It ought to be referenced that the organization of MSW and ICI shifts relying
402 C. Nirmala et al.

upon the beginning, creation designs, family pay and geological area [36]. MSW
includes items such as cooking and home equipment, rubber and plastic compo-
nents, metallic substances, cardboard and paper scraps, inert materials, digital trash,
and so on [37]. There are also biodegradable and non-biodegradable fractions in
MSW. It is important to note that approximately 15% of municipal waste is recycled,
while the remaining waste is dumped in open areas or landfills [38].
The key drivers of the global increase in MSW creation are population growth,
gentrification, and industry. By 2025, it is anticipated that the amount of MSW
produced worldwide will exceed 2 billion tons annually [39]. It is projected that in
a few emerging nations and different areas of the planet, the MSW age could reach
or surpass that of created countries without legitimate guidelines and arrangements
for landfilling and squander reusing offices [40]. Food preferences, consumption
patterns, behavior among consumers, and living conditions in both rural and urban
locations are all changing as the rate of MSW generation rises.
MSW age makes serious ecological contamination when unmanaged. Besides, its
change into esteem added items could give an answer for the difficulties of energy
lack and maintainable waste administration. MSW can be discarded, repurposed, or
hailed into fossil fuels in a variety of ways, including garbage dumps, soil fertiliza-
tion, anaerobic assimilation, pyrolysis, and gasification [37]. The cremation of 1 ton
of MSW could emanate 1.3 lots of CO2 comparable outflows, which is like how
much CO2 discharges from oil based power plants [41]. In addition, burning MSW
releases a significant amount of pollutants into the atmosphere, including fly ash and
particulate matter, refuse management strategy becoming unsustainable. However,
it has been demonstrated that MSW fly and bottom ash contains heavy metals that
pose a threat to ecosystems [42]
For the burial of non-recyclable waste, many municipalities worldwide prefer
to dispose of MSW in landfills. Despite its potential, MSW disposal in landfills is
fraught with difficulties, including groundwater contamination caused by methane
gas emissions and landfill leachate. MSW landfill leachate is both acute and chroni-
cally toxic, and it frequently enters groundwater biomagnification. Besides, leachate
could likewise sully the progression of water streams [43]. The generation of energy
from MSW contributes to the reduction of pollution and has the potential to support
a nation’s economic development by enhancing energy security and the management
of waste.

17.2.4 Food Waste

The term “food waste” refers to the organic and biodegradable garbage that is created
from a range of sources, such as food processing facilities, eateries, kitchens, and
residences [44]. This material is a part of the “organic fraction” of MSW. Due to
improper handling and usage of food, a lot of food waste is sent every year. Food
waste can also be caused by overproduction, contamination of produce by bacteria,
pests, and insects, excessive shopping, and postponed eating [45]. Every year, the
17 Waste Biomass Supply Chain for Sustainable Bioenergy Production 403

food retail network loses over 1.3 billion tons of food, including handled meat, dairy
products, organic foods, and vegetables [46].
Management of waste by numerous strategies for food waste includes landfilling
and burning. Food waste disposal can result in methane emissions, a more potent
GHG than CO2 [47]. The dry biomass is best disposed of by burning or cremation,
on the other side. Therefore, high-dampness food waste may result in the incinerator
using more energy, which would result in high operating costs. An effective substitute
for methanogenic bacteria producing biogas (or biomethane) through biomethanation
is anaerobic digestion of food waste [48].
Food waste comprises of cellulose, carbohydrates, proteins, lipids, fatty acids, and
organic acids. Also, the carbs available in food waste could go through hydrolysis
to deliver monosaccharides and oligosaccharides reasonable for organic transfor-
mation [46]. Food waste can also be a valuable resource for the fermentation of
bioethanol [49], biobutanol [50], and biohydrogen [51] due to its organic compo-
sition. As biohydrogen is acquiring prominence, food waste can end up being an
eco-accommodating and financially savvy feedstock for its creation through photo-
graph/dim maturation and gasification. By pyrolysis [52] and hydrothermal gasi-
fication [53], discarded waste food can also be transformed thermochemical into
bio-oil, biochar, and hydrogen-rich syngas. Activated carbon and biochar production
from waste food has also demonstrated promising results [54]. The boundaries, for
example, the arrangement of food squander, pretreatment techniques, and handling
boundaries impact the development of biofuels.

17.2.5 Cattle Manure

Cattle excrement refers to the metabolic and waste byproducts of poultry and animal
farming. Fertilizer is an important ingredient that contains natural matter and nutrients
that are essential for crop development [55]. Furthermore, animal faces may include a
number of diseases that pose environmental concerns. Waste feed, waste feed water,
and metabolic waste or faces are all common components of animal manure. In spite
of being a significant wellspring of horticultural supplements, domesticated animals
compost can likewise add to the outflow of GHGs, for example, CH4 by microbial
decay [56]. It is important to note that animal manure emissions account for 10%
of all emissions from agricultural production [57]. The United Nation defined AW
as waste from various agricultural operations, such as manure and other wastes
from farms, poultry houses, and slaughterhouses; collect waste; compost overflow
from fields; pesticides that enter the water, air, or soils; and fields were drained of
salt and silt. The aforementioned definition includes waste streams produced by all
modern agricultural processes, including animal and animal processing waste. To put
it plainly, AW contains four important parts: animal waste, crop waste, hazardous
waste (like pesticides), and processing waste [58].
404 C. Nirmala et al.

Composting and vermicomposting are two other methods of traditional manure


treatment [57]. These medicines are extremely normal in agricultural nations on
account of their straightforwardness and cost-adequacy. Composting also ensures
that plants will have access to nutrients. Clay soils also get more aeration and water
infiltration from composting. Creature fertilizer could be valorized by a few tech-
niques, for example, anaerobic processing [59], dull maturation [60], maturation [61],
pyrolysis [62], aqueous liquefaction [63], aqueous gasification and torrefaction [64]
to deliver bio-oil, bioethanol, biomethane, bio-raw petroleum, biohydrogen, syngas
and roasted biomass, separately. After the anaerobic digestion of compost, the diges-
tate left over might be utilized as a raw material for the production of biochar, bio-oil,
and syngas by pyrolysis, liquefaction, and gasification, individually.
Cattle manure has proven tremendous promise for the generation of biochemicals
and biofuels through biological conversion and thermochemical practices. Horse
dung has been shown to be an effective fuel for hydrothermal gasification-based
hydrogen generation by [45]. In a different research, ethanol plant wastewater and
chicken manure were digested together anaerobically to produce feedstock [65].
Catalytic hydrothermal gasification was used to produce hydrogen using a mixture
of poultry manure and eucalyptus wood [66]. The chemical composition of different
waste biomass was illustrated in Table 17.1.

Table 17.1 Chemical composition of waste biomass


S. No. Biomass sources (%) Cellulose (%) Hemicellulose (%) Lignin References
1 Live-oak-leaves 19.2 10.9 23.7 [67]
2 Wire-grass 34.8 22.8 18.7 [67]
3 Corn-stover 37.5 22.4 17.6 [68]
4 Paddy straw 32.0 24.0 13.0 [68]
5 Food waste 18.30 7.55 2.16 [69]
(household)
6 Dairy cow 17.1 18.1 5.17 [70]
manure
7 Hybrid poplar 43.8 14.7 25.7 [71]
8 Whole tree 51.2 23.4 25.4 [72]
9 Pine 42 21 30 [73]
10 Swine manure 12.2 34.0 5.38 [70]
11 Beech wood 44.2 33.5 21.8 [74]
12 Water hyacinth 24.5 34.1 8.6 [75]
13 Napier grass 47 31 22 [76]
17 Waste Biomass Supply Chain for Sustainable Bioenergy Production 405

According to [77], cellulose, a polymer of glucose (C6 H10 O5 ) linked by 1,4-


glycosidic bonds, is one of the most abundant organic compound present on earth.
The structure of cellulose molecule is crystalline and linear, with a couple of nebulous
zones and completely breaks down at 300–400 °C [78]. Hemicellulose, in contrast
to cellulose, has a structure that is composed of several monomers like mannose,
xylose, galactose, glucose, and others. ranging from 500 to 3000° of polymerization
and highly branched [79]. Coniferyl alcohol (G), sinapyl alcohol (S), and P-coumaryl
alcohol (H) are the three aromatic subunits (monolignols) that make up lignin, which
may be a complex cross-linked polymer of arbitrary structure in differentiate to
hemicellulose and cellulose. The fundamental distinction among the three is the
quantity of methoxy gatherings.

17.3 Waste Biomass Pretreatment

Agriculture wastes and other biomass can be turned into valuable goods that are
economical, sustainable, and good for the environment [80]. One of the crucial
components of integrated conversion processes of biorefineries is the pretreatment,
a standalone procedure that comes before a specific process. Figure 17.3 illustrates
the overall process of waste biomass pretreatment for bioenergy conversion.

Fig. 17.3 Overall process of waste biomass pretreatment for bioenergy conversion
406 C. Nirmala et al.

Effectiveness, technical, profitable, and environmental viability of pretreatments


are the critical factors that make the primordial biomass structure more simple and
accessible to the enzymes and reagents needed to create high added-value products
thus increase the biodegradability and digestibility of agricultural LB. Significant
lignocellulosic material size reduction under mesophilic conditions also boosts the
production of volatile fatty acids during anaerobic digestion. In addition, structural
changes such as a decrease in cellulose’s crystallinity and rise in surface area result
from the shrinking of lignocellulosic material. The amount of energy needed to shrink
the size of lignocellulosic contents varies on the biomass type and humidity content
[81, 82]. The study emphasizes the widely used pretreatment techniques, including
physical, chemical, and biological techniques.

17.3.1 Physical Pretreatment

Physical pretreatment processes enlarge the surface area and decrease the particle
size of waste biomass that results in improved production yield, reduced cellu-
lose structure, improved transition effectiveness, and higher microbial and enzyme
accessibility during processing. The common physical pretreatment methods are
mechanical, ultrasonic, microwave, and thermal pretreatment technique.
Cutting, shearing, milling, agitation, and chipping are some of the other mechan-
ical pretreatment techniques that lowers the amount of lignin, the level of crosslinking
and the dimension of the particulates thus enables the easy degradation of cellu-
lose and hemicellulose [83]. Comparatively mechanical pretreatment is simple and
has no negative impact on the environment. In general, mechanical pretreatment
also includes the use of vibrating, hammering, two-rolling, milling, chipping, and
grinding procedures.
The movement of a screw inside a small container, pretreatment of ligno-
cellulose materials is accomplished through a combination of heat and mechan-
ical activities known as Extrusion pretreatment. Heat can be applied to lignocel-
lulosic substances utilizing irradiation pretreatment techniques such an electron
beam, microwave, ultrasonic, and gamma rays, which dissolve hemicellulose, reduce
cellulose crystallinity, and depolymerize lignin [84, 85].

17.3.2 Chemical Pretreatment

To debilitate and break up the strong lignocellulose gem structure, increment the
surface region, and upgrade the biodegradability of farm land residues, special
chemicals are utilized in chemical pretreatments. To reduce the handling, corrosion,
and risk of toxicity, diluted version of acids such as HCl, H2 SO4 , H3 PO4 , HNO3 ,
Na2 CO3 , CH3 COOH, CH2 O2 , C4 H4 O4, caustic agents such as NaOH, NH4 OH,
KOH, NH3 ·H2 O, Ca(OH)2 are used to break the glycosidic bonds, disrupt the
17 Waste Biomass Supply Chain for Sustainable Bioenergy Production 407

cellulose and hemicellulose in biomass, and consequently promote the release of


fermentable sugar [86]. Oxidative pretreatment is made easier by the hemicellulose
and lignin in the biomass being broken down by ozone gas and hydrogen peroxide
(H2 O2 ). The chemical breakdown of H2 O2 into –OH and O2 helps to destroy the
lignin’s structure while preventing the release of any inhibiting byproducts [87].
To ensure efficient transition of cellulose, lignocellulosic material is first
pretreated with organic solvents such as acetone, ethanol, methanol, ethylene glycol,
and tetrahydrofuranol to disrupt the internal linkages of hemicelluloses and lignin.
Occasionally, catalysts are utilized to speed up the chemical process, such as HCl,
H2 SO4 , or bases like NaOH, NH3 , and CaCO3 [88].
The exothermic wet oxidation pretreatment is carried out at lower temperatures.
This process involves mixing water with lignocellulosic material before adding an
oxidizing substance, such as air, oxygen, or hydrogen peroxide [89]. Following, the
feedstock is heated to a temperature of 125–300 °C and a pressure of 0.5–20 MPa. The
three key variables influencing the pretreatment methods are pressure, temperature,
and residence time. The inclusion of oxidizers in pretreatment results in the synthesis
of free radicals and affects the reaction rate. The rate of reaction is proportional to
the oxygen levels, and the cost of pretreatment might increase depending on the
oxygen purity used. Alkyl aryl ether bond cleavage, electrophilic replacements, and
oxidative cleavage of aromatic nuclei are the main processes [90].
The lignocellulosic material is oxidized using a sophisticated wet explosion
pretreatment, which is followed by physical disruption and depends on tempera-
ture, pressure, and residence duration [91]. The pretreatment takes place at 140 and
220 °C temperature and a 0 and 3.5 MPa pressure. In ozonolysis, lignocellulosic
material has been pretreated with ozone, a potent oxidizing agent that influences
the lignin fraction during the process. The cellulose remained unaltered after the
pretreatment; although it was observed that the hemicellulose was partly changed.
The ozone concentrations, biomass particle size, and water content of the material
also affect the pretreatment process [92].

17.3.3 Biological Pretreatment

Utilizing enzymes, aerobic bacteria, and fungi to degrade lignin and facilitate the
conversion of lignocellulose into usable products is known as biological pretreatment
[93]. This technique is environmentally friendly, creates little to no inhibitor, uses
less energy, and discharges no harmful substances into the environment [94]. Fungi
belonging to white-rot fungus, soft-rot, and brown-rot classes are known as fungal-
based biological pretreatment approach to break down lignin. For the biodegradation
of lignin, white-rot fungi from the class Basidiomycetes Phanerochaete chrysospo-
rium are frequently utilized. Termites are another organism utilized to break down
lignocellulose-containing materials. According to reports, Termes hospes, Nasu-
titermes ephratae, and Microcerotermes parvus release lignocellulose-degrading
enzymes that help wheat straw break down before being digested anaerobically.
408 C. Nirmala et al.

Organisms from soil, broiler, wastewater, etc., are utilized as microbial consortium
pretreatment specialists for farming wastes [95].
Agricultural wastes and other LB can be pretreated using enzymes such as cellu-
lases and hemicelluloses. To improve their conversion efficiency for the manufac-
ture of biobutanol, farming wastes like sugarcane bagasse, Napier grass, corn stover,
cassava pulp, rice straw, etc. have undergone pretreatment with enzymes. By using
bacteria that produce enzymes, bacterial pretreatment slows down the polymerization
of lignocellulose [96].

17.3.4 Hybrid Pretreatment

The bioenergy researchers are currently working to create new pretreatment tech-
niques that will effectively separate LB into streams of cellulose, hemicellulose, and
lignin [97]. Numerous researchers have employed hybrid pretreatment techniques
in this area, and the outcomes are encouraging. Various pretreatment techniques are
combined in order to reduce pretreatment costs and increase its efficacy. Liquid hot
water is processed with chemical-free in thermochemical pretreatments at higher
pressures and temperatures. When pressurized water enters the biomass, it causes
cellulose to hydrate, hemicellulose to be hydrolyzed, and the lignin component of the
biomass to be removed. Since the pretreatment increases the lignocellulosic mate-
rials’ methane output, it is frequently employed in anaerobic digestion processes [24].
Pretreatment with alkaline solution at pH 11 and 13 increased biodegradability and
biogas production. Another combination that increases methane yield while reducing
pretreatment energy use is physical treatment mixed with biological process. The
formation of inhibitors is decreased when wet oxidation and alkali pretreatment are
combined. Besides, steam explosion combined with wet oxidation overcomes the
huge particle size of biomass. Steam explosion and wet oxidation are used to handle
hard biomass as well as biomass with greater lignin levels. The enzymatic hydrol-
ysis of cellulose is improved by pretreatment with the ionic liquids combined with
ultrasonic and microwave pretreatment [91].
Numerous studies with combination pretreatment strategies are rather low than
mono pretreatment strategies. In order to minimize pretreatment costs and improve
biogas yields, it is important to further investigate the applicability of combined
pretreatment practices for numerous class of biomass [98]. The overall waste biomass
pretreatment practices are shown in Fig. 17.4. Merits and demerits of various
pretreatment practices are listed in Table 17.2.
17 Waste Biomass Supply Chain for Sustainable Bioenergy Production 409

Pretreatment

Physical Chemical Biological

Mechanical Acid Fungal

Ultrasonic Bacterial
Alkali

Thermal Oxidation Termite

Organic solvent Microbial consortium

Enzymes

Fig. 17.4 Waste biomass pretreatment practices functions, merits, and demerits

17.4 Traditional Approaches of Waste Biomass Conversion


to Bioenergy

The biological origin is one of several variables that affect the composition of
biomass. The abundance of biomass as trash in many forms, including agricultural
waste, forest residue, industrial waste, MSW and wood waste, has made its acqui-
sition simpler, more affordable, and environmentally benign [99]. If biomass is not
processed, it may cause a number of social and environmental issues. The socioe-
conomic and environmental effects of first-generation biofuels derived from edible
crops on food security, land use, and biodiversity have also been challenged [100].
These concerns are allayed by the use of waste biomass for energy harvesting [39].
However, due to its intrinsic chemical and physical characteristics, without pretreat-
ment, biomass is insufficiently valuable and less effective for direct use. Biomass
has some characteristics that make it unsuitable for direct use as fuel, including high
oxygen concentration, hydrophilic nature, uneven composition, moisture content,
low lower grindability, calorific value, and fibrous structure [101]. Because of this,
converting biomass into higher-quality biofuel and other products with added value
is essential. To this end, extensive research has been done on a variety of biomass
conversion processes, which are broadly split into two classes: biochemical conver-
sion and thermochemical conversion [102]. Inherently slower and less effective
than thermochemical processes, biochemical conversion techniques use bacteria and
enzymes to break down biomass into smaller molecules [103]. Thermochemical
conversion techniques, on the other hand, use heat to break down biomass into low-
molecular compounds through a sequence of regulated physicochemical reactions to
produce the required results. Although the thermochemical reactions require outside
heat at least initially, they are quite simple to regulate [104]. Gasification, combustion,
pyrolysis, hydrothermal liquefaction (HTL), hydrothermal carbonization (HTC), and
Table 17.2 Merits and demerits of various pretreatment practices
410

S. No. Pretreatment practices Merits Demerits


1 Mechanical • No inhibitors • High operational cost
• Significant reduction in the size • High energy consumption
• Easy operation and simple • Cannot degrade lignin
• More CH4 production • Required additional pretreatments
• Minimized crystallinity • Low sugar yield
• Enzyme digestibility induced
• Higher porosity
• Biomass size reduction
• Low investment
2 Ultrasonic • Low energy utilization • Low selectivity
• Diminish the degree of polymerization • High energy consumption
• No formation of inhibitors • Loss of energy in diluted media
• Increased hydrogen fermentation • Large-scale reactor design challenges
• Minimal waste production
• Multiple product formation
• Intensified processing
3 Thermal • Non-toxic • Generation of toxins
• Environment friendly • Generation of inhibitors
• Highly reliable method • High temperature
4 Acid • Short residence time • High toxicity and corrosiveness
• High enzymatic deconstruction reaction rates • Formation of poisonous substances and inhibitors
• Efficient lignin removal • Corrosion in equipment
• Decomposition of hemicellulose • Difficult in recycling
• Low cost
(continued)
C. Nirmala et al.
Table 17.2 (continued)
S. No. Pretreatment practices Merits Demerits
5 Alkali • Low inhibitors • Environmental pollution
• Increased surface area • Residence time is long
• Decrease in the crystallinity of biomass • Consumption-condensation if high energy
• Delignification rate is high • Redistribution of lignin
• Utilized milder conditions • Economically non-viable
• Operating temperature is less
• Diminished polymerization
• Low degree of cellulose crystallinity
6 Oxidation • Ecologically friendly practice • Formation of toxic and inhibitor
• Highly selective • High oxidant required
• Can operate and room temperature • Expensive process
• Hemicellulose and lignin elimination • Special instrument required
• The structure of cellulose is not changed
7 Organic solvent • Reaction time is short • High corrosive
• High toxicity
• Environment pollution
8 Fungal • Mild conditions • Long reaction period
• Low energy consumption • Low cellulose recovery
• Energy efficient • Low efficiency
• No pollution • Long residence time
17 Waste Biomass Supply Chain for Sustainable Bioenergy Production

9 Bacterial • Low energy consumption • Increased incubation period


• Less production of hazardous chemicals • Lengthy preparation period
• Eco friendly
10 Termite • Mild action circumstances • Long pretreatment time
• Low energy consumption
• Environment friendly
(continued)
411
Table 17.2 (continued)
412

S. No. Pretreatment practices Merits Demerits


11 Microbial consortium • Environment friendly • Needs further research and investigation
• Dependable operability
12 Enzymes • Higher cellulose digestibility • Long incubation time
• Low cost • Require more investigations
• Rapid response time
• Extremely effective
• No corrosive problems
• A minimal energy footprint
C. Nirmala et al.
17 Waste Biomass Supply Chain for Sustainable Bioenergy Production 413

torrefaction are the main thermochemical conversion processes. These processes


produce a variety of energy outputs, including bio-oil, biochar, combustible gases,
and syngas. Additionally, hydrothermal processes are best suited for biomass with
high moisture content because they use subcritical or supercritical conditions of water
at high pressures, whereas other thermal processes are better suited for biomass with
relatively lower moisture content [105]. Advantages, limitations and challenges of
traditional approaches used in waste biomass conversion to bioenergy are detailed
in Table 17.3.

17.4.1 Innovative Techniques for Waste Biomass Conversion

Rapid industrialization and population growth and increased awareness in global


warming prioritize emerging technologies and procedures that quicken the switch
from fossil fuels to renewable energy sources. It is essential to meet future global
energy demand and to safeguard and/or restoring the environment. The innovative
techniques for waste biomass conversion are listed in Table. 17.4.

17.5 Bioenergy Applications

From the stoves used in homes for cooking to huge power plants are used to generate
electricity [127]. New woodstove designs increase the effectiveness of the heating
or cooling system and reduce the fuel usage. Biomass is used for businesses and in
industry for a range of functions, including energy production, space heating, and
hot water heating [128]. Many industrial environments, including lumber factory,
naturally produce organic waste [129].

17.5.1 Biodiesel

Biodiesel is an effective replacement for diesel; it is a sustainable, fueled source


that can improve the environment, replace petroleum thus develop job opportunities.
Biodiesel is produced from a various feedstocks mixture that includes reprocessed
soybean oil, cooking oil, and animal fats [1, 81]. Biodiesel initially concerned with
specialized quality and engine operation conditions and may be used in diesel engines
without alteration in a mixes of 5% or 20% biodiesel [130].
Table 17.3 Traditional approaches of waste biomass conversion to bioenergy, advantages, limitations, and challenges
414

S. No. Name of approach Advantages Limitations Challenges References


1 Combustion A typical combustion process has Combustion is not suitable The combustion cannot initiate [106]
four steps with increasing for wet biomass wastes such as because the temperature cannot
temperature, i.e., animal manure, sewage, raw increase high enough to enter the
drying, devolatilization, char digestates, etc subsequent combustion stages
oxidation, and gaseous product
combustion which can be
expressed in terms of biomass
weight loss as a function of
temperature
2 Gasification Compared to combustion, for This method has the intrinsic Biomass gasification is [107, 108]
electricity generation can achieve limitation in obtaining reasonable comparatively a more complex
higher efficiency using biomass information about dispersed process since it involves
wastes particles, such as particle size multi-scale and multi-physics
The primary step in the distribution and particle processes, and it is influenced by
production of a variety of shrinkage. In contrast, the various process parameters. The
bio-based chemicals, such as Eulerian–Lagrangian method majority of these determining
green synthesis, which can begin tracks the trajectory of each factors are interdependent, and as
with syngas, is also provided by particle and features the inherent a result, their effects may vary
biomass gasification information of dispersed depending on the stage of a
particles, thus has emerged as a reaction. Consequently, it is
promising way to simulate dense challenging to have a pure
gas–solid reacting flow discussion of the effects of
process parameters separately
(continued)
C. Nirmala et al.
Table 17.3 (continued)
S. No. Name of approach Advantages Limitations Challenges References
3 Pyrolysis The merit of the pyrolysis process The demerit of the process The removal of char is [109, 110]
is that it is a zero waste process includes high energy challenging. Due to scale-up
The final purpose is to achieve consumption due to its constraints, large-scale systems
energy-efficient biomass wastes endothermic nature, and further must be thoroughly researched
conversion to produce tailored research is needed before its
products effectively (i.e., to industrial-scale implementation
enhance product selectivity). Due
to their inherent renewable
qualities, unconventional heating
sources like solar-thermal-driven
biomass pyrolysis are also
gaining popularity
4 Torrefaction Torrefied biomass, which is The main demerit of torrefaction The system’s drawback is the [111, 112]
composed of a modified is Volume of torrefied biomass is substantial quantity of heat that is
polymeric structure compared to reduced only slightly applied to the product or material
the raw biomass, has better fuel throughout the entire process,
properties such as low moisture which causes flavor, color, and
content, higher heating value, clouding degradation upon
lower volatile content, and better storage and shortens the product’s
grindability, making it suitable overall shelf life
for direct usage as a fuel.
17 Waste Biomass Supply Chain for Sustainable Bioenergy Production

Torrefaction has the advantage of


producing no greenhouse gases.
Instead, it fixed the carbon as
densified and dehydrated biomass
for further thermochemical
processing
(continued)
415
Table 17.3 (continued)
416

S. No. Name of approach Advantages Limitations Challenges References


5 Hydrothermal processing Water participates in the process It has a number of drawbacks, The main difficulties with [113–115]
as a reactant as well as a solvent, including expensive initial hydrothermal processing include
eliminating the need for the equipment costs because of the a lack of fundamental
time-consuming pre-drying. specialized machinery needed and thermodynamic data under
Since the water is still liquid, the severe corrosion that requires extreme settings, a lack of
latent heat is preserved, maintenance because of the understanding of reaction
increasing the thermal efficiency extreme reaction conditions kinetics, and the impact of mass
of the process. Furthermore, transfer on the process
compared to most thermal
processes, the hydrothermal
conversion uses low
temperatures. Additionally,
because gaseous SOx and NOx
easily dissolve in water, there is a
relative decrease in the
production of hazardous gases
C. Nirmala et al.
17 Waste Biomass Supply Chain for Sustainable Bioenergy Production 417

Table 17.4 Innovative techniques for waste biomass conversion


S. No. Innovative technology References
1 Self-energized pine dust pyrolysis system for reduction in external [116]
electrical energy supply and to improve the yield of dry bio-oil
2 Used a rectification column as the pyrolysis reactor and a series of [117]
differently arranged condensers to model a pyrolysis system for
waste plastics. The condensation and fractionation units were then
simulated using ChemCad software using this model oil. The
separation of near-boiling components is improved for only up to
10 theoretical stages
3 Advanced analytical methods like Fourier transform ion cyclotron [118]
resonance mass spectrometry (FT-ICR MS) and 1H and 13C
nuclear magnetic resonance (NMR) spectroscopy were used to
characterize tire pyrolysis oil produced in a lab-scale twin-auger
reactor. This research will aid in developing the upgrading tactics
4 To recover benzene, toluene, ethylbenzene, xylene chemicals, and [119]
limonene from tire pyrolysis oil produced by an industrial-scale
pyrolysis plant, a pilot-scale distillation plant was created, built,
and put into operation
5 Plastic waste can be catalytically pyrolyzed to produce liquid oil [120]
and other products with added value using natural zeolite that has
undergone modification. The catalytic properties of new thermal
(TA) and acidic (AA) activation are improved
6 The amount of biomass production was first estimated for 20 years [121]
using the LandGEM and EViews software in order to determine the
amount of electrical energy needed (in the specific study area), and
the economic, environmental, and energy analysis was completed
using the Homer software. The estimated payback period was
10 years, and the annual electricity production was estimated to be
11,658.265 MWh
7 Temperature, residence time, catalyst loading, and feed ratio are [122]
examples of operational parameters and are the only factors that
affect the quality and kind of product forms throughout the
pyrolysis process. The thermal behavior caused by the structural
makeup of hemicellulose showed that it has a less impact on the
generation of bio-oils during the pyrolysis process
8 Hydrothermal liquefaction (HTL) is a method of producing bio-oil [123]
from an aqueous slurry of biomass or organic materials at
temperatures between 280 and 370 °C (lower than the pyrolysis
process), high pressures of 10–25 MPa, and water contentment
9 Bioelectricity is mostly generated by burning lignocellulose [124]
feedstock derived from biomass sources such as farm waste,
plantation forests, sawdust from sawmills, and natural forests. The
use of lignocellulose from forestry and agricultural biomass to
generate electricity by direct burning was reported in an Australian
study on the viability of producing bioelectricity from biomass to
reduce greenhouse gas emissions
(continued)
418 C. Nirmala et al.

Table 17.4 (continued)


S. No. Innovative technology References
10 Fourth-Generation of biofuels is a general term for the application [125]
of genetic engineering, molecular biology, and transdisciplinary
physicochemical techniques to optimize and boost the yield in the
process of biofuel production. In order to increase biofuel output,
the fourth generation of biofuels uses genetically engineered algae
that accumulate large levels of lipid and carbohydrates
11 The biomass of Graesiella emersonii microalga serves as feedstock [126]
for biodiesel production as well as producing valuable ω-6 and
carotenoids

17.5.2 Biofuels

Using conventional technologies, biofuels are produced from vegetable oils and
sugars [131]. Advanced biofuels, in contrast, are produced from agricultural waste,
cellulosic biomass, and woody crops, that makes it challenging to excerpt the neces-
sary fuel [1]. Due to these significant limitations in first-generation biofuel production
and threatening in food supply and biodiversity, advanced biofuel technologies are
developed. Many first-generation biofuels depend on subsidies, have high production
and transportation emissions, are expensive and only partially reduce greenhouse gas
emissions. Advanced biofuels can resolve these issues and provide a larger share of
the world’s fuel supply in a cheaper, sustainable, and environmentally friendly way
[132].

17.5.3 Energy for Transportation

Animal waste and plant biomass are recycled to produce fuel for transportation
and energy, heat, power, fuel and steam. Biomass is also used in the production of
chemicals such as detergents, bio-fertilizers, and erosion control products, as well
as the food processing, animal feed, and wood products industries, which include
construction and fiber products like paper and derivatives, all use [133].

17.5.4 Biogas

Biogas is a mixture of gases, also known as landfill gas (LFG) produced by the
breakdown of organic matter by anaerobic digestion using anaerobic bacteria in a
closed system. Biogas includes methane (CH4 ), trace amounts of hydrogen sulfide
(H2 S), carbon dioxide (CO2 ), siloxanes and moisture [134]. It is an inexhaustible
source and exerts a very small carbon footprint [135].
17 Waste Biomass Supply Chain for Sustainable Bioenergy Production 419

17.6 Circular Economy of Waste Biomass Management

A global movement that offers hope for the sustainable utilization of natural resources
is the circular economy. The biomass ramifications in utility must be understood
by contributors from the initial design of products up to waste governance that
includes all the stages in value chain, in order to develop a circular bioeconomy
[136]. A circular economy depends on resources being used to their fullest poten-
tial forever and nearly eliminating unrecoverable trash [137]. An efficient response
to the combined problem of rising waste production and rising energy demand is
waste-to-energy (Fig. 17.5).
However, the variety of waste poses a significant management difficulty. Because
of its affordability and accessibility of raw materials, biomass waste valorization is a
desirable energy alternative [138]. The following issues must be addressed: (1) Using
renewable fertilizers to decouple the petrochemical industry from biomass produc-
tion; (2) increasing stakeholder cooperation across value chains; (3) minimizing
waste from food and agriculture; and (4) providing an abundant supply of biomass
for bio-based products. Based on the final analysis utilizing linear regression and
artificial neural network, new models are created to predict biomass wastes higher
heating value [138]. Biomass is essential for the creation of energy and material
items in a circular economy.
The bioeconomy includes primary production, industrial sectors specifically
biological-oriented forestry, agricultural, aquaculture and fisheries, and economic

Fig. 17.5 Waste-to-energy conversion process


420 C. Nirmala et al.

sectors, to create feed, food, energy, bio-based goods, and services [11]. Primary
production includes carbon farming and separation and is the cost-efficient and
sustainable use of natural resources management activity. It includes biodiversity,
water, land, and other environmental resources organization. Industrial processing
and distribution encompass a product’s design and manufacturing procedure in terms
of both the product’s primary purpose and the possibility for waste reduction and
recycling. Additionally, cascading has the prospective to upgrade the coherence of
processing and utilizing waste and residue from forestry and agriculture. Food waste
can be reduced by increasing circularity in packaging and product delivery, as well
as by assuring recyclable materials and reducing environmental effect. Consump-
tion includes the distribution of food and bio-based products as well as their reuse,
recycling, and disposal. End-of-life refers to the point at which the waste produced
from the biomass production, processing, consumption, and bioenergy production
that are. For trash generated from biomass and bio-based goods to be recycled and
reused, waste sorting must be improved that leads to better recycling technologies and
procedures as well as the extraction of valuable compounds as processing byprod-
ucts. Furthermore, essential inputs for the generation of bioenergy include biomass
and organic waste. Energy recovery, however, should only be used when lower-level
options in the waste hierarchy are not viable [139].

17.7 Challenges, Futuristic Concerns, and Motivation

Energy efficiency and renewable energy are referred to as the “twin pillars” of energy
policy. To manage and cut back on carbon dioxide emissions, each resource should be
enhanced. The exploration, production, and use of energy are governed by numerous
distinct energy policies on a global scale. These regulations are the result of a variety
of sources, including trade associations, commodity companies, automobiles, and
producers of wind and solar energy. The study of energy economic science has
recently focused on a number of topics, including property, energy markets and
economic processes, energy and environmental law and regulations, global warming,
and addressing many challenges associated to the rapid spread of renewable energy
technology in developing countries. The bulk of agricultural facilities in the devel-
oped world are mechanized because rural regions have been electrified. Rural elec-
trification has considerably increased production, but it has also increased energy
consumption.
Biomass supply chains are challenged in six key ways. Among the difficulties
and issues include the availability of supplies, digitization and automation, supply
chain alliances, community-based energy generation, social opportunities, and laws.
Agriculture in developing countries generates a lot of biomass, which has an impact
on the economic stability. The geographical and unpredictable availability, where
some biomass can only be harvested during specific seasons in certain places, is a
key factor influencing changes in the supply of biomass. The limited availability of
biomass will have an adverse effect on the project’s financial performance. On the
17 Waste Biomass Supply Chain for Sustainable Bioenergy Production 421

other hand, developing countries continue to use biomass for low-value purposes,
which is frequently a cheap replacement for investing a sizable sum of money in
establishing a local biomass sector. Unfortunately, if substantial volumes of biomass
were locked in these low-value activities, its availability would become extremely
rarer.
Effects of climate change on how some crops are harvested. Changes in rainfall,
average temperature, and humidity, for example, might affect crop dispersion and
output as a result of climate change. As a result, the supply chain for biomass will
become more vulnerable and its availability of biomass will be impacted. In fact,
supply uncertainty is one of the main issues hindering the growth of the bioeconomy
[140]. In order to lessen the vulnerability of the biomass supply chain, future crops
are expected to breed climate-resilient crops that are resilient to heat stress, biotic
stress, and salinity stress. In the near future, meticulous planning will be necessary
to make the transition from conventional crops to productively modified ones. By
increasing the efficiency of conversion through chemical addition, the product yield
of biological and thermochemical conversion processes to turn biomass into value-
added products can be increased, particularly the addition of nanoparticles or the
introduction of various novel catalysts [for instance, single-atom catalysts (SACs),
transition metal dichalcogenides (TMDC) catalysts, and metal–organic framework
(MOF)] catalysts [140].
Second, industry actors can lower the supply risk by emphasizing suitable manage-
ment and scheduling strategies. In general, additional biomass feedstocks and prod-
ucts can be stored for future use, especially when there is a lack of biomass available
that enables the ongoing production of biomass [141]. Energy storage mechanisms,
such as thermal or battery storage, can be used, for instance, in biomass-to-power
pathways. Finally, diversifying the biomass utilized in each biomass conversion
process is yet another possible tactic to lower supply risk. For downstream oper-
ations in biomass conversion plants, process automation utilizing big data, cloud
computing, and IoT is essential. In addition to monitoring and controlling the process,
the acquired data can be used for scheduling and storage management [141].
A virtual platform or data storage area that links all supply chain participants and
offers a transparent information source about the supply chain might be imagined as
block chain technology. It is possible for it to function without outside assistance.
Block chain technology can provide suppliers for biomass supply chains with real-
time information on the quality and origin of the biomass. Centralized power gener-
ation systems could be helpful for efficient management. However, they typically
cost too much to scale up. Labor shortages are the key challenges in biomass supply
networks. The day-to-day operations of biomass supply chains have been consider-
ably impacted by limitations on movement and other steps to curb the virus’s spread.
This is a result of the labor-intensive tasks involved in the biomass supply chain,
particularly those involving biomass gathering and harvesting [141].
422 C. Nirmala et al.

17.8 Conclusions

Utilizing waste is undoubtedly the most economical method for producing renew-
able energy, which also benefits the environmental cleaning. The present review has
uncovered the conventional and innovative techniques for waste-to-bioenergy conver-
sion from waste substrates/feedstock. Nevertheless, when compared to conven-
tional food crop-based biofuels, the production of biofuels from biomass waste is
still regarded as having more robust transportation and conversion technology, etc.
Converting wastes might assist strengthen a circular economy model by lowering
waste burden and the negative environmental effect of waste disposal while creating
many bioproducts in a clean and sustainable way. The use of bioenergy is an intriguing
application, but it should not be employed at the expense of waste avoidance and/
or reduction. Although bioenergy cannot be viewed as a saving solution, it is an
alternative, a supplement to a set of waste management initiatives that governments
must do. Research being done now aims to address the shortcomings of currently
used technologies and enhance their effectiveness and economics.
Still, more techniques are needed
• To improve the digestibility and biodegradability of LB into biofuel and other
valuable products and to decrease wastewater, animal waste, etc., innovative
pretreatment strategies should be developed.
• It is advised that advanced technologies like artificial intelligence, robots, smart
meters, and machine learning be widely used to reduce energy usage, reaction
times, and monitoring of the numerous pretreatment operations’ activities and
outputs.
• For biomass utilization, further research on the techno-economic, environmental,
and life cycle assessments of various pretreatment techniques is required.
• To exhibit the potential of the biorefinery concepts, focus is required.
Nevertheless, constant research is being conducted in order to address the flaws
in the most recent technological advancements and increase their usefulness and
economic viability.

References

1. Jayakumar, M., Gebeyehu, K. B., Abo, L. D., et al. (2023). A comprehensive outlook on topical
processing methods for biofuel production and its thermal applications: Current advances,
sustainability and challenges. Fuel, 349, 128690.
2. Neupane, D., Adhikari, P., Bhattarai, D., et al. (2022). Does climate change affect the yield
of the top three cereals and food security in the world? Earth, 3, 45–71.
3. Jayakumar, M., Gindaba, G. T., Gebeyehu, K. B., et al. (2023). Bioethanol production from
agricultural residues as lignocellulosic biomass feedstock’s waste valorization approach: A
comprehensive review. Science of The Total Environment, 163158.
4. Tesfaye Gari, M., Tessema Asfaw, B., Arumugasamy, S. K., et al. (2023). Natural resources-
based activated carbon synthesis. In Encyclopedia of green materials (pp. 1–11). Springer.
17 Waste Biomass Supply Chain for Sustainable Bioenergy Production 423

5. Jayakumar, M., Vaithilingam, S. K., Karmegam, N., et al. (2022). Bioethanol production from
green biomass resources: Emerging technologies.
6. Perlack. R. D. (2005). Biomass as feedstock for a bioenergy and bioproducts industry: The
technical feasibility of a billion-ton annual supply. Oak Ridge National Laboratory.
7. Mahapatra, S., Kumar, D., Singh, B., & Sachan, P. K. (2021). Biofuels and their sources
of production: A review on cleaner sustainable alternative against conventional fuel, in the
framework of the food and energy nexus. Energy Nexus, 4, 100036.
8. Bioenergy IEA. (2009). Bioenergy—A sustainable and reliable energy source. International
Energy Agency Bioenergy.
9. Hiloidhari, M., Sharno, M. A., Baruah, D. C., & Bezbaruah, A. N. (2023). Green and sustain-
able biomass supply chain for environmental, social and economic benefits. Biomass and
Bioenergy, 175, 106893.
10. Panoutsou, C., & Singh, A. (2020). A value chain approach to improve biomass policy
formation. GCB Bioenergy, 12, 464–475.
11. Ludlow, P. (2018). The European Commission. In The New European Community (pp. 85–
132). Routledge.
12. Havelaar, A. H., Ivarsson, S., Löfdahl, M., & Nauta, M. J. (2013). Estimating the true inci-
dence of campylobacteriosis and salmonellosis in the European Union, 2009. Epidemiology &
Infection, 141, 293–302.
13. Corbelle-Rico, E., & Crecente-Maseda, R. (2014). Evaluating IRENA indicator “risk of farm-
land abandonment” on a low spatial scale level: The case of Galicia (Spain). Land Use Policy,
38, 9–15.
14. Rosillo-Calle, F., & Woods, J. (2012). The biomass assessment handbook. Routledge.
15. Ghosh, S. K. (2016). Biomass & bio-waste supply chain sustainability for bio-energy and
bio-fuel production. Procedia Environmental Sciences, 31, 31–39.
16. Edler, J. (2013). Review of policy measures to stimulate private demand for innovation.
Concepts and effects. Compendium of Evidence on the Effectiveness of Innovation Policy
Intervention, 13, 44.
17. Karlsson, I., Rootzén, J., & Johnsson, F. (2020). Reaching net-zero carbon emissions in
construction supply chains—Analysis of a Swedish road construction project. Renewable
and Sustainable Energy Reviews, 120, 109651.
18. Bellù, L. G. (2013). Value chain analysis for policy making. Methodological Guidelines and
country cases for a quantitative approach Roma. Food and Agriculture Organization.
19. Wang, K., & Tester, J. W. (2023). Sustainable management of unavoidable biomass wastes.
Green Energy and Resources, 100005.
20. Lee, S. Y., Sankaran, R., Chew, K. W., et al. (2019). Waste to bioenergy: A review on the
recent conversion technologies. BMC Energy, 1, 1–22.
21. Kushwaha, D., Srivastava, N., Mishra, I., et al. (2019). Recent trends in biobutanol production.
Reviews in Chemical Engineering, 35, 475–504.
22. Council, N. R. (2012). Renewable fuel standard: Potential economic and environmental effects
of US biofuel policy. National Academies Press.
23. Xing, J., Luo, K., Wang, H., et al. (2019). A comprehensive study on estimating higher heating
value of biomass from proximate and ultimate analysis with machine learning approaches.
Energy, 188, 116077.
24. Gundupalli, M. P., Gundupalli, S. P., & Thomas, A. S. S., et al. (2022). Hydrothermal lique-
faction of lignocellulosic biomass for production of biooil and by-products: current state of
the art and challenges. Biofuels and Bioenergy, 61–84.
25. Sarkar, N., Ghosh, S. K., Bannerjee, S., & Aikat, K. (2012). Bioethanol production from
agricultural wastes: An overview. Renewable Energy, 37, 19–27.
26. Jayakumar, M., Kuppusamy Vaithilingam, S., Karmegam, N., et al. (2022). Fermentation
technology for ethanol production: current trends and challenges. In B. Gurunathan, & R. B.
T.-B. Sahadevan (Eds.), Biofuels and bioenergy (pp. 105–131). Elsevier.
27. Duque-Acevedo, M., Belmonte-Ureña, L. J., Cortés-García, F. J., & Camacho-Ferre, F. (2020).
Agricultural waste: Review of the evolution, approaches and perspectives on alternative uses.
Global Ecology and Conservation, 22, e00902.
424 C. Nirmala et al.

28. Obi, F. O., Ugwuishiwu, B. O., & Nwakaire, J. N. (2016). Agricultural waste concept, genera-
tion, utilization and management. Nigerian Journal of Technology (NIJOTECH), 35, 957–964.
https://doi.org/10.4314/njt.v35i4.34
29. Singh, A., Nanda, S., Guayaquil-Sosa, J. F., & Berruti, F. (2021). Pyrolysis of Miscanthus
and characterization of value-added bio-oil and biochar products. The Canadian Journal of
Chemical Engineering, 99, S55–S68.
30. Singh, B., Guldhe, A., Rawat, I., & Bux, F. (2014). Towards a sustainable approach for
development of biodiesel from plant and microalgae. Renewable and Sustainable Energy
Reviews, 29, 216–245.
31. Kumar, A., Sharma, A., & Upadhyaya, K. C. (2016). Vegetable oil: nutritional and industrial
perspective. Current Genomics, 17, 230–240.
32. Lavenburg, V. M., Rosentrater, K. A., & Jung, S. (2021). Extraction methods of oils and
phytochemicals from seeds and their environmental and economic impacts. Processes, 9,
1839.
33. Khan, I. U., Chen, H., Yan, Z., & Chen, J. (2021). Extraction and quality evaluation of biodiesel
from six familiar non-edible plants seeds. Processes, 9, 840.
34. Blackshaw, R., Johnson, E., Gan, Y., et al. (2011). Alternative oilseed crops for biodiesel
feedstock on the Canadian prairies. Canadian Journal of Plant Science, 91, 889–896.
35. Ajala, O. E., Aberuagba, F., Odetoye, T. E., & Ajala, A. M. (2015). Biodiesel: Sustainable
energy replacement to petroleum-based diesel fuel—A review. ChemBioEng Reviews, 2, 145–
156.
36. Kumar, A., & Samadder, S. R. (2017). An empirical model for prediction of household solid
waste generation rate—A case study of Dhanbad, India. Waste Management, 68, 3–15.
37. Nanda, S., & Berruti, F. (2021). Municipal solid waste management and landfilling technolo-
gies: A review. Environmental Chemistry Letters, 19, 1433–1456. https://doi.org/10.1007/s10
311-020-01100-y
38. Rajendran, N., Gurunathan, B., Han, J., et al. (2021). Recent advances in valorization of
organic municipal waste into energy using biorefinery approach, environment and economic
analysis. Bioresource Technology, 337, 125498.
39. Gunarathne, V., Ashiq, A., Ramanayaka, S., et al. (2019). Biochar from municipal solid
waste for resource recovery and pollution remediation. Environmental Chemistry Letters, 17,
1225–1235.
40. Fazeli, A., Bakhtvar, F., Jahanshaloo, L., et al. (2016). Malaysia’s stand on municipal solid
waste conversion to energy: A review. Renewable and Sustainable Energy Reviews, 58, 1007–
1016.
41. Ouda, O. K. M., Cekirge, H. M., & Raza, S. A. R. (2013). An assessment of the potential
contribution from waste-to-energy facilities to electricity demand in Saudi Arabia. Energy
Conversion and Management, 75, 402–406.
42. Clavier, K. A., Paris, J. M., Ferraro, C. C., & Townsend, T. G. (2020). Opportunities and
challenges associated with using municipal waste incineration ash as a raw ingredient in
cement production—A review. Resources, Conservation and Recycling, 160, 104888.
43. Mishra, S., Tiwary, D., Ohri, A., & Agnihotri, A. K. (2019). Impact of Municipal Solid Waste
Landfill leachate on groundwater quality in Varanasi, India. Groundwater for Sustainable
Development, 9, 100230.
44. Tong, H., Shen, Y., Zhang, J., et al. (2018). A comparative life cycle assessment on four waste-
to-energy scenarios for food waste generated in eateries. Applied Energy, 225, 1143–1157.
45. Nanda, S., Dalai, A. K., Gökalp, I., & Kozinski, J. A. (2016). Valorization of horse manure
through catalytic supercritical water gasification. Waste Management, 52, 147–158.
46. Kiran, E. U., Trzcinski, A. P., Ng, W. J., & Liu, Y. (2014). Bioconversion of food waste to
energy: A review. Fuel, 134, 389–399.
47. Thapa, P., Hasnine, M. D. T., Zoungrana, A., et al. (2022). Food waste treatments and the
impact of composting on carbon footprint in Canada. Fermentation, 8, 566.
48. Ren, Y., Wang, C., He, Z., et al. (2022). Enhanced biomethanation of lipids by high-solid
co-digestion with food waste: Biogas production and lipids degradation demonstrated by
long-term continuous operation. Bioresource Technology, 348, 126750.
17 Waste Biomass Supply Chain for Sustainable Bioenergy Production 425

49. Kiran, E. U., & Liu, Y. (2015). Bioethanol production from mixed food waste by an effective
enzymatic pretreatment. Fuel, 159, 463–469.
50. Qin, Z., Duns, G. J., Pan, T., & Xin, F. (2018). Consolidated processing of biobutanol produc-
tion from food wastes by solventogenic Clostridium sp. strain HN4. Bioresource Technology,
264, 148–153.
51. Yasin, N. H. M., Mumtaz, T., & Hassan, M. A. (2013). Food waste and food processing waste
for biohydrogen production: A review. Journal of Environmental Management, 130, 375–385.
52. Patra, B. R., Nanda, S., Dalai, A. K., & Meda, V. (2021). Slow pyrolysis of agro-food wastes
and physicochemical characterization of biofuel products. Chemosphere, 285, 131431.
53. Nanda, S., Isen, J., Dalai, A. K., & Kozinski, J. A. (2016). Gasification of fruit wastes and
agro-food residues in supercritical water. Energy Conversion and Management, 110, 296–306.
54. Patra, B. R., Nanda, S., Dalai, A. K., & Meda, V. (2021). Taguchi-based process optimiza-
tion for activation of agro-food waste biochar and performance test for dye adsorption.
Chemosphere, 285, 131531.
55. Chen, J.-H. (2006). The combined use of chemical and organic fertilizers and/or biofertilizer
for crop growth and soil fertility. In International workshop on sustained management of the
soil-rhizosphere system for efficient crop production and fertilizer use (pp. 1–11). Citeseer.
56. Nguyen, B. T., Trinh, N. N., & Bach, Q.-V. (2020). Methane emissions and associated micro-
bial activities from paddy salt-affected soil as influenced by biochar and cow manure addition.
Applied Soil Ecology, 152, 103531.
57. Khoshnevisan, B., Duan, N., Tsapekos, P., et al. (2021). A critical review on livestock manure
biorefinery technologies: Sustainability, challenges, and future perspectives. Renewable and
Sustainable Energy Reviews, 135, 110033.
58. Nagendran, R. (2011). Agricultural waste and pollution. In Waste (pp. 341–355). Elsevier.
59. Yao, Y., Huang, G., An, C., et al. (2020). Anaerobic digestion of livestock manure in cold
regions: Technological advancements and global impacts. Renewable and Sustainable Energy
Reviews, 119, 109494.
60. Dareioti, M. A., Vavouraki, A. I., Tsigkou, K., et al. (2021). Dark fermentation of sweet
sorghum stalks, cheese whey and cow manure mixture: Effect of pH, pretreatment and organic
load. Processes, 9, 1017.
61. Yan, Q., Liu, X., Wang, Y., et al. (2018). Cow manure as a lignocellulosic substrate for fungal
cellulase expression and bioethanol production. AMB Express, 8, 1–12.
62. Su, G., Ong, H. C., Zulkifli, N. W. M., et al. (2022). Valorization of animal manure via
pyrolysis for bioenergy: A review. Journal of Cleaner Production, 343, 130965.
63. Liu, Q., Xu, R., Yan, C., et al. (2021). Fast hydrothermal co-liquefaction of corn stover and
cow manure for biocrude and hydrochar production. Bioresource Technology, 340, 125630.
64. Itoh, T., Iwabuchi, K., Maemoku, N., et al. (2019). A new torrefaction system employing
spontaneous self-heating of livestock manure under elevated pressure. Waste Management,
85, 66–72.
65. Cheong, D.-Y., Harvey, J. T., Kim, J., & Lee, C. (2019). Improving biomethanation of chicken
manure by co-digestion with ethanol plant effluent. International Journal of Environmental
Research and Public Health, 16, 5023.
66. Yong, T.L.-K., & Matsumura, Y. (2012). Catalytic gasification of poultry manure and euca-
lyptus wood mixture in supercritical water. Industrial & Engineering Chemistry Research,
51, 5685–5690.
67. Matt, F. J., Dietenberger, M. A., & Weise, D. R. (2020). Summative and ultimate analysis
of live leaves from southern US forest plants for use in fire modeling. Energy & Fuels, 34,
4703–4720.
68. Sawatdeenarunat, C., Surendra, K. C., Takara, D., et al. (2015). Anaerobic digestion
of lignocellulosic biomass: Challenges and opportunities. Bioresource Technology, 178,
178–186.
69. Matsakas, L., Kekos, D., Loizidou, M., & Christakopoulos, P. (2014). Utilization of household
food waste for the production of ethanol at high dry material content. Biotechnology for
Biofuels, 7, 1–9.
426 C. Nirmala et al.

70. Lu, J., Li, H., Zhang, Y., & Liu, Z. (2018). Nitrogen migration and transformation during
hydrothermal liquefaction of livestock manures. ACS Sustainable Chemistry & Engineering,
6, 13570–13578.
71. Wang, K., Zhang, J., Shanks, B. H., & Brown, R. C. (2015). The deleterious effect of inor-
ganic salts on hydrocarbon yields from catalytic pyrolysis of lignocellulosic biomass and its
mitigation. Applied Energy, 148, 115–120.
72. Vassilev, S. V., Baxter, D., Andersen, L. K., & Vassileva, C. G. (2010). An overview of the
chemical composition of biomass. Fuel, 89, 913–933.
73. Sannigrahi, P., Miller, S. J., & Ragauskas, A. J. (2010). Effects of organosolv pretreatment and
enzymatic hydrolysis on cellulose structure and crystallinity in Loblolly pine. Carbohydrate
Research, 345, 965–970.
74. Demirbaş, A. (2005). Thermochemical conversion of biomass to liquid products in the aqueous
medium. Energy Sources, 27, 1235–1243.
75. Ruan, T., Zeng, R., Yin, X.-Y., et al. (2016). Water hyacinth (Eichhornia crassipes) biomass
as a biofuel feedstock by enzymatic hydrolysis. BioResources, 11, 2372–2380.
76. Reddy, K. O., Maheswari, C. U., Dhlamini, M. S., et al. (2018). Extraction and characterization
of cellulose single fibers from native african napier grass. Carbohydrate Polymers, 188, 85–91.
77. Heinze, T. (2016). Cellulose: structure and properties. In Cellulose chemistry and properties:
Fibers, nanocelluloses and advanced materials, 1–52.
78. Acharya, B., Sule, I., & Dutta, A. (2012). A review on advances of torrefaction technologies
for biomass processing. Biomass Conversion and Biorefinery, 2, 349–369.
79. Phan, P. T., Nguyen, B.-S., Nguyen, T.-A., et al. (2023). Lignocellulose-derived mono-
sugars: A review of biomass pre-treating techniques and post-methods to produce sustainable
biohydrogen. Biomass Conversion and Biorefinery, 13, 8425–8439.
80. Koul, B., Yakoob, M., & Shah, M. P. (2022). Agricultural waste management strategies for
environmental sustainability. Environmental Research, 206, 112285.
81. Jayakumar, M., Gebeyehu, K. B., Selvakumar, K. V., et al. (2022). Waste Ox bone based hetero-
geneous catalyst synthesis, characterization, utilization and reaction kinetics of biodiesel
generation from Jatropha curcas oil. Chemosphere, 288, 22–24. https://doi.org/10.1016/j.che
mosphere.2021.132534
82. Awogbemi, O., & Von Kallon, D. V. (2022). Pretreatment techniques for agricultural waste.
Case Studies in Chemical and Environmental Engineering, 100229.
83. Zhang, H., Han, L., & Dong, H. (2021). An insight to pretreatment, enzyme adsorption
and enzymatic hydrolysis of lignocellulosic biomass: Experimental and modeling studies.
Renewable and Sustainable Energy Reviews, 140, 110758. https://doi.org/10.1016/j.rser.2021.
110758
84. Tsapekos, P., Kougias, P. G., Frison, A., et al. (2016). Improving methane production from
digested manure biofibers by mechanical and thermal alkaline pretreatment. Bioresource
Technology, 216, 545–552.
85. Niu, P., Zhang, L., Liu, G., & Cheng, H. (2012). Graphene-like carbon nitride nanosheets for
improved photocatalytic activities. Advanced Functional Materials, 22, 4763–4770.
86. Rezania, S., Oryani, B., Cho, J., et al. (2020). Different pretreatment technologies of ligno-
cellulosic biomass for bioethanol production: An overview. https://doi.org/10.1016/j.energy.
2020.117457
87. Tan, H., Li, J., He, M., et al. (2021). Global evolution of research on green energy and
environmental technologies: A bibliometric study. Journal of Environmental Management,
297, 113382.
88. Kumari, D., & Singh, R. (2018). Pretreatment of lignocellulosic wastes for biofuel production:
A critical review. Renewable and Sustainable Energy Reviews, 90, 877–891. https://doi.org/
10.1016/J.RSER.2018.03.111
89. Chaturvedi, V., & Verma, P. (2013). An overview of key pretreatment processes employed for
bioconversion of lignocellulosic biomass into biofuels and value added products. 3 Biotech,
3, 415–431.
17 Waste Biomass Supply Chain for Sustainable Bioenergy Production 427

90. Bhaskar, T., Bhavya, B., Singh, R., et al. (2011). Thermochemical conversion of biomass to
biofuels. In Biofuels (pp. 51–77). Elsevier.
91. Khan, M. U., Usman, M., Ashraf, M. A., et al. (2022). A review of recent advancements in
pretreatment techniques of lignocellulosic materials for biogas production: Opportunities and
Limitations. Chemical Engineering Journal Advances, 10, 100263. https://doi.org/10.1016/j.
ceja.2022.100263
92. Brodeur, G., Yau, E., Badal, K., et al. (2011). Chemical and physicochemical pretreatment of
lignocellulosic biomass: A review. Enzyme Research, 2011.
93. Iram, A., Berenjian, A., & Demirci, A. (2021). A review on the utilization of lignin as a
fermentation substrate to produce lignin-modifying enzymes and other value-added products.
Molecules, 26, 2960.
94. Hasson, D., Shemer, H., & Sher, A. (2011). State of the art of friendly “green” scale control
inhibitors: A review article. Industrial & Engineering Chemistry Research, 50, 7601–7607.
95. Chaurasia, B. (2019). Biological pretreatment of lignocellulosic biomass (water hyacinth) with
different fungus for enzymatic hydrolysis and bio-ethanol production resource: Advantages,
future work and prospects. Acta Science Agriculture, 5.
96. He, C.-R., Kuo, Y.-Y., & Li, S.-Y. (2017). Lignocellulosic butanol production from Napier
grass using semi-simultaneous saccharification fermentation. Bioresource Technology, 231,
101–108.
97. Jayakumar, M., Kuppusamy Vaithilingam, S., Karmegam, N., et al. (2022). Fermentation
technology for ethanol production: Current trends and challenges. Biofuels and Bioenergy,
105–131. https://doi.org/10.1016/b978-0-323-90040-9.00015-1
98. Zheng, Y., Zhao, J., Xu, F., & Li, Y. (2014). Pretreatment of lignocellulosic biomass for
enhanced biogas production. Progress in Energy and Combustion Science, 42, 35–53.
99. Yang, H., Zhou, Y., & Liu, J. (2009). Land and water requirements of biofuel and implications
for food supply and the environment in China. Energy Policy, 37, 1876–1885.
100. Kurowska, K., Marks-Bielska, R., Bielski, S., et al. (2020). Food security in the context of
liquid biofuels production. Energies, 13, 6247.
101. van der Stelt, M. J. C. (2011). Chemistry and reaction kinetics of biowaste torrefaction.
102. Porpatham, E., Ramesh, A., & Nagalingam, B. (2012). Effect of compression ratio on the
performance and combustion of a biogas fuelled spark ignition engine. Fuel, 95, 247–256.
103. Fatehi, H., Weng, W., Li, Z., et al. (2021). Recent development in numerical simulations and
experimental studies of biomass thermochemical conversion. Energy & Fuels, 35, 6940–6963.
104. Anca-Couce, A. (2016). Reaction mechanisms and multi-scale modelling of lignocellulosic
biomass pyrolysis. Progress in Energy and Combustion Science, 53, 41–79.
105. Wijekoon, P., Wickramasinghe, C., Athapattu, B. C. L., et al. (2021). Biomass valorization
and phytoremediation as integrated technology for municipal solid waste management for
developing economic context. Biomass Conversion and Biorefinery, 11, 363–382.
106. Koppejan, J., & Van Loo, S. (2012). The handbook of biomass combustion and co-firing.
Routledge.
107. Wang, S., Hu, C., Luo, K., et al. (2023). Multi-scale numerical simulation of fluidized beds:
Model applicability assessment. Particuology, 80, 11–41.
108. Wang, S., & Shen, Y. (2020). CFD-DEM study of biomass gasification in a fluidized bed
reactor: Effects of key operating parameters. Renewable Energy, 159, 1146–1164.
109. Scott, D. S., Majerski, P., Piskorz, J., & Radlein, D. (1999). A second look at fast pyrolysis
of biomass—The RTI process. Journal of Analytical and Applied Pyrolysis, 51, 23–37.
110. Jahirul, M. I., Rasul, M. G., Chowdhury, A. A., & Ashwath, N. (2012). Biofuels production
through biomass pyrolysis—A technological review. Energies, 5, 4952–5001.
111. Sarker, T. R., Nanda, S., Dalai, A. K., & Meda, V. (2021). A review of torrefaction technology
for upgrading lignocellulosic biomass to solid biofuels. BioEnergy Research, 14, 645–669.
112. Chiou, B.-S., Cao, T., Valenzuela-Medina, D., et al. (2018). Torrefaction kinetics of almond
and walnut shells: Effects of inorganic species. Journal of Thermal Analysis and Calorimetry,
131, 3065–3075.
428 C. Nirmala et al.

113. Rabenau, A. (1985). The role of hydrothermal synthesis in preparative chemistry. Angewandte
Chemie International Edition in English, 24, 1026–1040.
114. Gao, Y., Yu, B., Wu, K., et al. (2016). Physicochemical, pyrolytic, and combustion charac-
teristics of hydrochar obtained by hydrothermal carbonization of biomass. BioResources, 11,
4113–4133.
115. Munir, M. T., Mansouri, S. S., Udugama, I. A., et al. (2018). Resource recovery from organic
solid waste using hydrothermal processing: Opportunities and challenges. Renewable and
Sustainable Energy Reviews, 96, 64–75.
116. Charis, G., Danha, G., Muzenda, E., & Nhubu, T. (2021). Modeling a sustainable, self-
energized pine dust pyrolysis system with staged condensation for optimal recovery of bio-oil.
Frontiers in Energy Research, 8, 594073.
117. Krzywda, R., & Wrzesińska, B. (2021). Simulation of the condensation and fractionation unit
in waste plastics pyrolysis plant. Waste and Biomass Valorization, 12, 91–104.
118. Campuzano, F., Abdul Jameel, A. G., Zhang, W., et al. (2020). Fuel and chemical properties
of waste tire pyrolysis oil derived from a continuous twin-auger reactor. Energy & Fuels, 34,
12688–12702.
119. Martínez, J. D., Veses, A., Callén, M. S., et al. (2023). On fractioning the tire pyrolysis oil
in a pilot-scale distillation plant under industrially relevant conditions. Energy & Fuels, 37,
2886–2896.
120. Miandad, R., Rehan, M., Barakat, M. A., et al. (2019). Catalytic pyrolysis of plastic waste:
Moving toward pyrolysis based biorefineries. Frontiers in Energy Research, 7, 27.
121. Alayi, R., Sadeghzadeh, M., Shahbazi, M., et al. (2022). Analysis of 3E (energy-economical-
environmental) for biogas production from landfill: A case study. International Journal of
Chemical Engineering, 2022.
122. Varma, A. K., Shankar, R., & Mondal, P. (2018). A review on pyrolysis of biomass and
the impacts of operating conditions on product yield, quality, and upgradation. In Recent
advancements in biofuels and bioenergy utilization (227–259).
123. Biller, P., & Ross, A. B. (2016). Production of biofuels via hydrothermal conversion. In
Handbook of biofuels production (pp. 509–547). Elsevier.
124. Farine, D. R., O’Connell, D. A., John Raison, R., et al. (2012). An assessment of biomass
for bioelectricity and biofuel, and for greenhouse gas emission reduction in Australia. Gcb
Bioenergy, 4, 148–175.
125. Mat Aron, N. S., Khoo, K. S., Chew, K. W., et al. (2020). Sustainability of the four generations
of biofuels—A review. International Journal of Energy Research, 44, 9266–9282.
126. Kang, N. S., Cho, K., An, S. M., et al. (2022). Taxonomic and biochemical characterization
of Microalga Graesiella emersonii GEGS21 for its potential to become feedstock for biofuels
and bioproducts. Energies, 15, 8725.
127. Hossain, M. F., Hossain, S., & Uddin, M. J. (2017). Renewable energy: Prospects and trends
in Bangladesh. Renewable and Sustainable Energy Reviews, 70, 44–49.
128. Zhang, X., Ruoshui, W., Molin, H., & Martinot, E. (2010). A study of the role played by
renewable energies in China’s sustainable energy supply. Energy, 35, 4392–4399.
129. Väisänen, T., Haapala, A., Lappalainen, R., & Tomppo, L. (2016). Utilization of agricultural
and forest industry waste and residues in natural fiber-polymer composites: A review. Waste
Management, 54, 62–73.
130. Degfie, T. A., Mamo, T. T., & Mekonnen, Y. S. (2019). Optimized biodiesel production from
waste cooking oil (WCO) using calcium oxide (CaO) nano-catalyst. Scientific Reports, 9,
1–8. https://doi.org/10.1038/s41598-019-55403-4
131. Demirbas, A. (2007). Progress and recent trends in biofuels. Progress in Energy and
Combustion Science, 33, 1–18.
132. Zhang, W., Elaine, A. Y., Rozelle, S., et al. (2013). The impact of biofuel growth on agriculture:
Why is the range of estimates so wide? Food Policy, 38, 227–239.
133. (US) EIA. (2012). Annual energy review 2011. Government Printing Office.
134. Chozhavendhan, S., Karthigadevi, G., Bharathiraja, B., et al. (2023). Current and prog-
nostic overview on the strategic exploitation of anaerobic digestion and digestate: A review.
Environmental Research, 216, 114526.
17 Waste Biomass Supply Chain for Sustainable Bioenergy Production 429

135. Kafarov, V., & Rosso-Cerón, A. M. (2021). Biomass as a source for heat, power and chemicals.
Advances in Carbon Management Technologies: Biomass Utilization, Manufacturing, and
Electricity Management, 2, 1.
136. Romero-Perdomo, F., & González-Curbelo, M. Á. (2023). Integrating multi-criteria tech-
niques in life-cycle tools for the circular bioeconomy transition of agri-food waste biomass:
A systematic review. Sustainability, 15, 5026.
137. Vela, I. C., Vilches, T. B., Berndes, G., et al. (2022). Co-recycling of natural and synthetic
carbon materials for a sustainable circular economy. Journal of Cleaner Production, 365,
132674.
138. Awogbemi, O., & Von Kallon, D. V. (2022). Valorization of agricultural wastes for biofuel
applications.
139. Hailemariam, A., & Erdiaw-Kwasie, M. O. (2023). Towards a circular economy: Implica-
tions for emission reduction and environmental sustainability. Business Strategy and the
Environment, 32, 1951–1965.
140. Andiappan, V., How, B. S., & Ngan, S. L. (2021). A perspective on post-pandemic biomass
supply chains: Opportunities and challenges for the new norm. Process Integration and
Optimization for Sustainability, 5, 1003–1010.
141. Lim, C. H., Lim, S., How, B. S., et al. (2021). A review of industry 4.0 revolution potential in a
sustainable and renewable palm oil industry: HAZOP approach. Renewable and Sustainable
Energy Reviews, 135, 110223.
Chapter 18
Manifesting Sustainability Toward Food
Waste into Bioenergy: Biorefinery
in a Circular Economic Approach

Devi Sri Rajendran, Swethaa Venkatraman, R. Rahul, M. Afrrin, P. Karthik,


and Vinoth Kumar Vaidyanathan

Abstract The global issue of excessive food waste has raised significant concerns,
underscoring the urgent need for sustainable interventions incorporating eco-friendly
technologies. Around 1.3 billion tons of food are discarded every year, representing a
one-third of the world total food production. This food wastage comes at a consider-
able cost to the world economy, estimated at around $750 billion. The escalating
generation of food waste highlights the pressing need for effective measures to
address this issue with a dual purpose. Utilization of food waste includes distillery
waste, feed and poultry waste, oil processing waste, cooking oil waste, seafood waste,
dairy waste, vegetable and fruit waste for production of various biobased products
reducing economic losses and promoting sustainability by harnessing. This encom-
passes the utilization of diverse bioprocesses, including acidogenesis, fermentation,
photosynthesis, solventogenesis, methanogenesis, bio-electrogenesis, and oleagi-
nous processes, each capable of yielding a wide spectrum of valuable products.
Within this book chapter, the primary emphasis on products, specifically bioelec-
tricity, biodiesel, and biofuels, as crucial elements for achieving sustainability and
advancing the objectives outlined in sustainable development goals (SDG)-12. SDG-
12 leads to responsible consumption and production, particularly SDG 12.3, which
aims to halve global per capita food waste. To ensure the economic viability of these
bioprocesses and align with sustainable goals, it is imperative to adopt a biorefinery
strategy that optimizes the utilization of residual organic waste for the recovery of a
diverse range of green products within a circular economy framework.

D. S. Rajendran · S. Venkatraman · V. K. Vaidyanathan (B)


Integrated Bioprocessing Laboratory, Department of Biotechnology, Faculty of Engineering and
Technology, School of Bioengineering, SRM Institute of Science and Technology (SRMIST),
Tamil Nadu, Kattankulathur, Chengalpattu District 603203, India
e-mail: vinothkv@srmist.edu.in
R. Rahul · M. Afrrin · P. Karthik
Department of Food Technology, Faculty of Engineering, Karpagam Academy of Higher
Education, Coimbatore, India

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2024 431
G. Baskar et al. (eds.), Circular Bioeconomy Perspectives in Sustainable Bioenergy
Production, Energy, Environment, and Sustainability,
https://doi.org/10.1007/978-981-97-2523-6_18
432 D. S. Rajendran et al.

Keywords Food waste · Bioprocess · Circular economy · Sustainable


development goals · Biofuels

18.1 Introduction

Fossil fuels are being the primary energy source, leading to the emission of substan-
tial quantities of greenhouse gases (GHG) and contributing to approximately 65% of
the elevated mortality rate attributed to toxic air pollutants [1]. The carbon dioxide
liberated during the burning of fossil fuels accounts for approximately two-thirds of
GHG emissions. The anthropogenic activity since the beginning of industrialization
has elevated atmospheric carbon dioxide levels from 280 ppm to about 417 ppm in
the last 151 years. In order to conform to the Paris Agreement by 2030, countries are
required to cut back 30 gigatons of GHG emissions per year (UNEP 2021 data). The
world faces a climate emergency with impending disasters such as high-intensity
hurricanes, heat waves, and drought. In order to substantially derail global temper-
ature trends and reduce the burden on the environment, a shift toward sustainable
biofuels such as bio-oil and bioethanol as an energy source is imperative [2].
Bio-oils are renewable substitutes for fuel oil and diesel, which can be utilized
as combustion material for power and heat generation as a feedstock for various
chemicals production such as polyols, carboxylic acids, phenolics, and levulinates
[3]. Due to increasingly strict regulatory standards and rising global awareness of
environmental concerns, the market of bio-lubricants is projected to grow from $
2.0 billion in 2020 to $ 2.4 billion by 2025, reflecting a 4.1% compound annual
growth rate during this period [4]. Recent developments with bio-oil have opened up
its application in polyurethane foam formation, asphalt binder production, plastics,
carbonaceous materials, pesticides, and fertilizers [5].
On the other hand, the bioethanol market is growing exponentially and is extrapo-
lated to reach $ 64.8 billion by 2025 with 14% of compound annual growth rate. Many
countries have shifted to bioethanol blended fuels for power generation and transport
fuel to reduce their GHG emissions. The last three decades have recorded extensive
research on the utilization of lignocellulosic biomass for ethanol production [6]. The
United Nations (UN) established a comprehensive framework called the Sustain-
able Development Goals (SDGs) 2030, consisting of 17 overarching goals and 169
specific targets in September 2015 [7]. These SDGs collectively represent a globally
recognized reflection of the needs and aspirations of stakeholders, encompassing
social, economic, and environmental dimensions [8].
SDG target 12.3 aims to lower the food waste by 50% at consumption and retail
stages. This SDG 12.3 also the minimize loss of food during production, such as
post-harvest waste, by 2030 through the promotion of 3R (reduce, reuse, recycle)
practices [9]. It is important to note that the global’s population is projected to reach
approximately nine billion people by 2050, which will inevitably lead to an increased
demand for food. This rise in demand will necessitate increased food production,
potentially resulting in even more food waste [10]. To provide context, it is worth
18 Manifesting Sustainability Toward Food Waste into Bioenergy … 433

noting that around 1.3 billion metric tons of food are discarded annually on a global
scale, equivalent to one-third of the production of total food (FAO 2019).
A little more than a quarter, specifically 27% of food waste mainly occurs during
the consumption phase [11]. The worldwide energy consumption associated with the
food system is estimated to exceed 70 exajoules (1 exajoule = 1018 J) [12]. Notably,
food production contributes towards the emission of greenhouse gas upto 20 to 30%,
a significant driver of global warming and a key factor in climate change [13].
The Paris Agreement from COP 2015 aimed to increase earth temperature
reduced below 2 °C, preferably 1.5°, to reduce environmental risks. Food waste
contributes towards the production of 4.4 gigatons of CO2 and 8% of global green-
house gas emissions. Shifting to plant-based foods from animal-based ones, like red
meat, can cut the food sector’s carbon footprint. Reducing food waste conserves not
only resources but also energy and water. Sustainability throughout the supply chain
is crucial in minimizing the environmental impact of food waste [14–16].
The fourth industrial revolution and technological progress have improved global
living standards but also pose potential harm [17]. Thanks to its diverse climate and
soil resources, India ranks second in global fresh fruit and vegetable production after
China. This study focuses on the entire fresh produce supply chain, from farming to
consumption. The waste generated in this chain, including land, water, and energy
misuse, has adverse environmental impacts [18].
Every stage of the food supply chain contributes to greenhouse gas emissions,
causing environmental pollution that persists even after food waste disposal [19, 20].
Addressing food waste, especially in the realm of fresh produce, presents a distinct
challenge due to factors such as its perishable nature, technological constraints,
communication gaps among supply chain stakeholders, and mismanagement [21].
The current disorderly patterns of consumption and production contribute to a world-
wide crisis. However, reshaping these patterns has the potential to boost economic
growth, improve human well-being, and mitigate detrimental environmental effects
[22]. From 2017 to 2020, 83 countries, the European Union and territories, docu-
mented their efforts to transition toward sustainable consumption and production
(SCP) practices. Additionally, 136 policies and 27 activities were registered in 2020
aimed at implementing SCP practices [9].
Despite progress, there is a lack of specific area-wise data, highlighting the need
for effective policies to enhance resource efficiency [23]. On the basis of available
information as of 2016, an average of food (14%) went to waste before reaching the
retail market. It is important to note that achieving SDG targets often present syner-
gies, reducing food loss can contribute to addressing multiple goals such as poverty
eradication (target 1.5), zero hunger (target 2.4), promoting healthy lifestyles (target
3.9), and improving freshwater resources (target 6.4) within the UN Sustainable
Development Goals (UN-SDGs 2019) [24].
This study primarily centers around the 2030 SDG 12, which promotes sustainable
production and consumption practices. It is worth noting that Goal 12 has been
identified as having the highest number of trade-offs [8]. The study chiefly delves
into the difficulties linked to the monitoring of S12.3 SDG, which involves reducing
food waste by half, and it explores potential strategies to attain this target by 2030.
434 D. S. Rajendran et al.

18.2 Food Waste

Food waste is a pressing global problem affecting nearly every country, squandering
approximately more than a billion tons of food annually, which includes household
waste of 61%, food service sector of 26%, and retail of 13% [25]. At present, malnu-
trition affects two billion individuals across the globe, and it is estimated that this
number will continue to rise by 2050. To fulfill the requirements of the expanding
global population, a 60% increase in food production will be necessary. With food
insecurity affecting millions worldwide, addressing food waste becomes a crucial
endeavor in sustainable, healthy, and resilient is a crucial goal [26]. One of the key
benefits of reducing food waste is its positive impact on individuals, the environment,
and economic well-being. It enhances food security and tackles the global challenges,
including biodiversity loss, climate change, and pollution [27]. Additionally, it helps
alleviate the burden on waste management systems.
The UN-SDGs have set a target to reduce food waste by half by 2030. In March
2021, the Food Waste Index 2021 report was published by the United Nations Envi-
ronment Program (UNEP), presenting a standardized approach to measure and report
food waste, in accordance with SDG target 12.3 [9]. This groundbreaking index
reveals that global food waste is more than double according to the previous esti-
mates. As per the report, an astounding one billion tons of food are wasted annually at
the retail and consumer levels. This excessive food waste generates 8–10% of global
GHG emissions, causing adverse impacts on biodiversity, water and land resources
and imposing high financial costs on governments, businesses, and households,
totaling nearly a trillion US dollars [28].

18.2.1 The Food Waste Index and Sustainable Development


Goal 12.3

Accurate and standardized method indicating fundamental starting point for effec-
tively reducing food waste. However, there are research gaps, a notable one being
the lack of information regarding the proportion of inedible parts in food waste.
The Food Waste Index, as it is currently measured, encompasses the food meant for
human consumption along with edible components. The food waste is divided into
specific sectors between what’s edible and what’s not is essential. This knowledge is
crucial for stakeholders as it aids in comprehending the issue and crafting solutions
[29]. Assessing food wastage within the retail and food service sectors will enhance
estimates in many countries and contribute to the development of national strategies
for preventing food waste. This report plays a vital role in tracking national progress
toward the 2030 goal and the achievement of SDG 12.3 [30].
SDG 12.3 focuses on reducing global food waste. Its primary objective is to
minimize global per capita food waste by half at both the retail and consumer levels,
and to reduce food waste in production and supply chains by 2030. SDG 12.3 aims
18 Manifesting Sustainability Toward Food Waste into Bioenergy … 435

to promote efficient food management, improve resource sustainability, and address


hunger and food security issues by minimizing waste along with food supply chain. It
contributes to the broader sustainable development agenda by advancing responsible
consumption and production practices, with the aim of creating a more sustainable
and equitable future.
Improving the collection, analysis, and modeling of food waste data provides a
good understanding of the global food waste problem and its significant potential for
prevention in lower, middle, and high-income countries [31]. They are promoting
a standardized methods for nations to gauge food waste across households, food
service establishments, and retail outlets. Nations that embrace this method will
produce comprehensive data to inform their domestic food waste reduction plans
and enable meaningful international comparisons [32]. Requests for data related to
the food waste index will be made biennially, aligning with data requisitions from
the United Nations statistics division [33].

18.2.2 Types of Food Waste

18.2.2.1 Fruit and Vegetable Waste

The production and processing of fruits and vegetables result in biomass losses that
are both unavoidable and potential environmental hazard [34]. These losses pose
environmental risks and lead to the loss of valuable nutrients. Extracting juices and
its derivatives from fruits and vegetables results in a significant volume of waste,
which includes peels from fruits such as bananas, apples, carrots and oranges [35].
Numerous research studies have explored biogas production, primarily focusing on
food waste alone or its co-digestion with other substrates [36]. For instance, Mu et al.
[37] investigated the anaerobic co-digestion of cabbage waste and potato waste, and
the methane production potential of pumpkin peels was investigated, along with an
analysis of the energy and economic aspects of anaerobic co-digestion involving
fluted pumpkin and poultry manure. It is worth noting that there is a scarcity of
research dedicated to the sole digestion of individual fruits or vegetables [37]. Yan
et al. [38] conducted a study in which they compared methane generation from
the waste of 20 leafy plant species, particularly emphasizing herbaceous plants
commonly used in China.

18.2.2.2 Seafood Waste

The fisheries industry, once viewed as a source of waste, now holds significant
economic potential. The production of fish has surged from 37 million tons in 1961
to 200 million tons in 2016, with per capita fish consumption of fish rising from 9 to
20 kg [39]. However, a concerning practice in this industry is fish discarding, where
unwanted catches are returned to the sea, dead or alive, with severe consequences for
436 D. S. Rajendran et al.

fish species and marine ecosystems [40]. Efficient utilization of fish offers various
benefits. Fish protein can be used for collagen, gelatin, protein hydrolysates, peptides,
and bioactive peptides [41]. Lipids can be processed into fish oil, and fish waste
can yield enzymes, pigments, calcium, and flavor compounds [42]. This promotes
a circular economy and reduces environmental burdens from fish waste. However,
proper regulation is vital due to food safety-associated risks in valorizing fish by-
products.

18.2.2.3 Dairy Waste

A fully operational dairy production facilityhas potential of processing around


500,000 liters of milk daily, may yield an estimated 200 to 350 Kg of waste [43].
Beyond this, dairy waste offers potential for the production of biosurfactants [44].
The dairy sector generates various by-products, such as buttermilk, whey, and their
derivatives. Whey is a major environmental contaminant because of its high organic
matter content [45]. It poses a significant threat to the environment because of its
elevated levels of both chemical and biological oxygen demand. Whey typically
contains components such as 0.6–0.7% protein, 4.5–5.0% lactose, and 0.4–0.5%
lipids, along with minerals and vitamins [46]. Utilizing whey sustainably for the
production of natural pigments has the potential to enhance the bioeconomy across
various industries and, consequently, boost the national bioeconomy [47].

18.2.2.4 Cooking Oil Waste

The total waste frying oils is produced globally, particularly in densely populated
nations like China produces approximately 5.6 million tons of waste frying oils.
In comparison, the USA contributes 1.2 million tons, India generates 1.1 million
tons, and the European Union (EU) generates around 0.9 million tons. The prevalent
approach for managing waste vegetable oil is its conversion into biodiesel, a process
that enhances the chemical and physical properties of raw waste vegetable oil for
practical use [48]. However, it is essential to note that this conversion process involves
substantial energy consumption and carries environmental impacts [49].

18.2.2.5 Oil Processing Waste

Globally, an estimated oil waste disposed each year was 200 million gallons. Intensive
oil processing at a significant scale led to the production of substantial waste streams.
These waste streams include tallow, soap stock, marine oils, lard and fatty acids
obtained during seed oil extraction [50]. The economics of biodiesel production can
be enhanced by utilizing residual seedcakes as a primary co-product. Although some
oilcakes contain toxic compounds that render them unfit for consumption, they hold
potential for energy production [51]. However, it is important to note that the high
18 Manifesting Sustainability Toward Food Waste into Bioenergy … 437

protein content in oilseed cakes can present challenges during the pyrolysis process,
potentially slowing down the degradation rate [52].

18.2.2.6 Feed and Poultry Waste

Organic solid waste is a problem with negative externalities, and it is imperative for
the agroindustry, especially in nations heavily engaged in chicken meat production,
consumption, and exports, to assume responsibility for its proper management [53].
Within the poultry production chain (PPC), four key stages contribute to generating
organic waste: reproductive poultry, hatchery operations, and broiler growth [54].
This waste comprises a variety of elements, including reproductive hatchery waste,
poultry litter, feed wastage, solid residues from flotation sludge, carcasses, truck
cleaning, charcoal, ash, and sausage casings [55]. These residues exhibit diverse char-
acteristics, including total organic carbon ranging from 4.4 to 56.22% and nitrogen
content ranging from 0.2 to 9.33% [56]. It is evident that in the PPC, there is a
significant mismanagement of raw materials, energy, and nutrient resources. This is
a concern, especially in developing countries, as decreasing these valuable resources
is unsustainable [56].

18.2.2.7 Distillery Waste

Distillery waste comprises the assortment of by-products and residues that stem from
the distillation process employed in producing alcoholic beverages like whiskey,
vodka, rum, or gin [57]. These waste materials manifest in various forms, each
possessing its distinct composition and characteristics. Firstly, there exist wash
waters, which contain dissolved sugars, alcohols, and assorted organic compounds.
These typically originate during the initial fermentation of grains or fruits. Secondly,
there is the presence of stillage, alternatively referred to as pot ale or slop. This
constitutes the liquid residue left in the still following the distillation process, char-
acterized by an abundance of non-fermentable solids, organic matter, and, at times,
residual alcohol [58].
Additionally, the distillation procedure frequently yields dried distillers’ grains,
the residual grains left over from fermentation [59]. These grains are often dehy-
drated and subsequently repurposed as animal feed, owing to their enriched protein
and nutrient content. Spent lees represent sediment or residue that settles at the bottom
of fermentation tanks or barrels during aging [60]. Efficiently managing distillery
waste assumes paramount importance in mitigating environmental pollution and opti-
mizing the utilization of valuable resources. Various techniques, such as anaerobic
digestion, composting, and biogas or biofuels production, can be employed to treat
and repurpose distillery waste ecologically and sustainably [61].
438 D. S. Rajendran et al.

18.3 Necessary Action Taken by Various Countries


and Organization

Several countries and organizations have taken proactive steps to address the issue of
food waste. Norway, the United States and Australia have committed to reducing food
waste generation by 50% by 2030 in alignment with SDG 12.3 [62]. Then, France
and Italy also enacted regulations prohibiting supermarkets from discarding unsold
food and mandating that they donate excess food to charitable organizations and
food banks [63]. Collaborative efforts are essential in combating unnecessary food
waste and safeguarding the environment. To encourage global cooperation, UNEP
has initiated regional task forces dedicated to addressing the issue of food waste in
regions such as the Pacific and Asia, Western Asia, Latin America, Africa, and the
Caribbean [64]. These groups aim to support countries with limited data on food
waste by enhancing their capacity to measure and reduce it. This initiative fosters
peer-to-peer collaboration and facilitates the establishment of reports on SDG 12.3
progress in 2022, ultimately helping nations develop national strategies to prevent
food waste [65].

18.4 Sustainable Food Future

Food waste substantially squanders precious resources, time, and finances. Effecting
substantial transformation necessitates collaborative actions aimed at establishing a
more comprehensive, sustainable, and robust food system [66]. Reducing food waste
across all tiers of society presents noteworthy advantages in terms of environmental
preservation, social well-being, and economic gains [67]. Accurate measurement of
food waste, including non-edible portions, and monitoring per capita food waste
generation nationally are crucial for tracking progress toward SDG 12.3 [68]. The
food waste index provides a methodology for countries to measure food waste,
spanning from households to retail sectors and food service. The FAO recom-
mends various practices to combat food waste at the individual or household level,
such as adopting healthier diets, buying what’s necessary, and using effective food
storage methods [69]. Figure 18.1 explains different methods used for the biological
valorization of food waste into different bioenergy products.

18.4.1 Electro-fermentation

18.4.1.1 Bioelectricity

Microbial fuel cells (MFCs) are an innovative and sustainable technology that directly
converts chemical energy from waste into bioelectricity [70, 71]. Microorganisms
18 Manifesting Sustainability Toward Food Waste into Bioenergy … 439

Fig. 18.1 Various methods of biological valorization of food waste into different bioenergy
products

present an anode that acts as a biocatalyst and generates electrons and protons
to facilitate the organic matter electrochemical oxidation. The cathode contains a
terminal electron acceptor, driving electron transfer through an external circuit for
bioelectricity generation, while protons pass through a separating membrane [72]. In
the cathode chamber, electrons and protons undergo reduction processes, producing
valuable products and treating wastewater [73]. Similar technologies, MFCs have the
potential to become hybrid systems, offering not only bioelectricity but also efficient
value-added products and waste treatment [6].

18.4.1.2 Biohydrogen

Biohydrogen (H2 ) exhibits sustainable and green fuel, drawing global research atten-
tion for its production from waste sources. H2 can be biologically generated through
methods like photo-fermentation, anaerobic fermentation, and bio-photolysis, either
individually or in combination. Acetogenic bacteria are essential for H2 produc-
tion in these processes. Various waste materials, including vegetable market waste,
440 D. S. Rajendran et al.

food waste, wastewater, and more, have been explored for biological H2 production
(Table 18.1) [6]. Food waste, rich in easily degradable organic matter, is particularly
promising and has been studied at semi-pilot and pilot scales [74].
Several factors influence H2 production in anaerobic fermentation, such as
inoculum type, pretreatment, pH, co-substrates, reactor setup, temperature, and
feed composition [104]. Integrating anaerobic fermentation with photo-fermentation
enhances H2 production, with the metabolites from anaerobic fermentation serving
as valuable inputs performed in two-phase as well as single-stage operations [105].

18.4.2 Anaerobic Fermentation/Acidogenesis

18.4.2.1 Biomethane

Biogas primarily contains carbon dioxide (CO2 ), methane (CH4 ), nitrogen (N2 ),
hydrogen sulfide (H2 S) and moisture. Among these components, methane production
is a well-explored technology for generating bioenergy, particularly using food waste
as a substrate in anaerobic fermentation (Table 18.2). The production of CH4 from
food waste is influenced by various factors, including temperature, pH, carbon-to-
nitrogen ratio [101], inoculum type [106], reactor design [107], organic loading rate,
volatile fatty acids [108], and co-digestion [109]. Recent progress in biogas produc-
tion has spurred the creation of high-rate biomethanation technologies, subsequently
enhancing process efficiency [72]. Methane production, which forms biogas, from
food waste has progressed to a commercial level in India and other countries, with
various industries adopting and implementing full-scale technologies.

18.4.2.2 Biohythane

Methane (CH4 ) and hydrogen (H2 ), each have their own limitations. H2 is reac-
tive and has flammability characteristics, posing storage challenges, while CH4 has
low flammability. Combining CH4 and H2 in a 1:4 ratio results in biohythane, an
emerging fuel with clean attributes and high calorific efficiency [128, 129]. This
blend enhances combustion rates, significantly improving the efficacy of CH4 -fueled
vehicles. The specific hydrogen (H2 ) production rate of 66.7 L/kg of total volatile
solid (TVS) feed was reported, along with biogas production rate of 0.72 m3 /kg
of TVS fed during the second phase of standard biohytane production [130]. In a
reactor scale study involving distillery spent wash, the highest recorded biohythane
production was found to be 147.5 ± 2.4 L [74]. The highest H2 yield was reported
in the continuous stirred tank reactor, and the highest CH4 yield was observed in the
anaerobic fluidized bed reactor [131].
Table 18.1 Various food wastes used for the production of different biodiesel and biohydrogen products
System Waste Yield Reference
Biodiesel
Solvent extraction Citrus peel 96.30% [75]
Chlorella vulgaris Fruit and vegetable waste [76]
Auxenochlorella protothecoides Food waste digestate 49.89 ± 0.31% [77]
Solvent extraction Waste cooking oil 90% [78]
Cryptococcus curvatus Food waste and municipal wastewater 28.6 ± 2.2% [79]
Rhodotorula glutinis 19.6 ± 0.2%
Candida lipolytica Corn cobs hydrolysate 19.4 g/L [80]
Yarrowia lipolytica 11.3 g/L
Cryptococcus curvatus Sweet sorghum bagasse 63.98–73.26% [81]
Chlorella vulgaris Agro-industrial coproducts, ethanol thin stillage and 43% [79]
soy whey
Chlorella sorokiniana Food waste and municipal wastewater 20–22.9% [79]
Schizochytrium mangrovei and Chlorella pyrenoidosa Rice, noodles, meat, and vegetables 300 mg/g [82]
Torulaspora maleeae Sugarcane molasses 0.31 g/L [83]
Torulaspora globosa 0.2 g/L
Acid and base catalyzed transesterification and oil Spent coffee grounds 82% [84]
18 Manifesting Sustainability Toward Food Waste into Bioenergy …

extraction
Cryptococcus curvatus Cheese whey 30–35% [85]
(continued)
441
Table 18.1 (continued)
442

System Waste Yield Reference


Aspergillus niger Waste cooking olive oil 0.49 g/g [86]
Yarrowia lipolytica Waste oil from chicken products, frying fish, frying 34.02–57.89% [87]
vegetables
Hydrolysis and esterification Waste cooking oil 91.8% fatty acid [88]
Ultrasonication oil extraction, hydrolysis and Spent coffee grounds 97% [89]
fermentation 0.5 g/g
Biohydrogen
Clostridium Vegetable waste 7.09 L/L-d [90]
Hydrothermal treatment Vegetable waste 38.1 mL/gVS [91]
Steam gasification Household food waste 73.6% [92]
Microwave co-pyrolysis Kitchen food waste 43.02% [93]
Purple non-sulfur bacteria Sugar cane bagasse 1.96 mol of H2 /mol sugar [94]
Sewage sludge Fallen leaves 37.8 mL/g of VS [95]
Anaerobic activated sludge Vegetable waste 85.65 mL-H2 g−1 VS [96]
Sludge Kitchen waste and white rice 1.27 mmol/g of COD [97]
Pre-treated anaerobic activated sludge Glucose 8.65 mol/kg CODR [98]
Anaerobic activated sludge Municipal food waste, kitchen 245 mL/g-COD [99]
Anaerobic activated sludge Food waste 9.67 L/h [100]
Scenedesmus sp. R-16 Starch-rich food waste 1643.5 mL/L [100]
Anaerobic activated sludge Vegetable waste 23.96 mmol/day [101]
(continued)
D. S. Rajendran et al.
Table 18.1 (continued)
System Waste Yield Reference
Anaerobic activated sludge Municipal food waste 370 mL/g of VS [99]
Clostridium, Prevotella, Paludibacter, Ensifer, and Rice straw 1.19 mol H2 /mol glucose [99]
Petrimonas
Digested sludge POME 2.45 mol/mol-sugar [102]
Anaerobic mixed culture Waste wheat 654.7 L/kg [103]
18 Manifesting Sustainability Toward Food Waste into Bioenergy …
443
444 D. S. Rajendran et al.

Table 18.2 Various food wastes used for the production of biohythane and biomethane products
System Waste Yield Reference
Biohythane
Dark fermentation and Food waste CH4 (70–90%, v/v) + H2 [110]
second-stage anaerobic (10–30%, v/v)
digestion
Alkali extraction Vegetable waste Hydrogen and methane at [111]
31.08 and 66.17%
Anaerobic digestion Food waste 90% methane [112]
Anaerobic digestion Pineapple waste Hydrogen: 1240 mL/L-d [113]
Methane: 812 mL/L-d
Phragmites australis Municipal food waste Hydrogen: 184.7 ± [114]
1.83 mL/g-VS added
Methane: 31.2 kJ/g-VS
added
Hydrogenogenic method Organic waste Hydrogen: 215 mL/g VS [115]
added
Methane: 676 mL/g VS
added
Anaerobic sludge Food waste [6]
Digested sludge Food waste 115.2 L of H2 /kg VS and [116]
334.7 L of CH4 /kg COD
Anaerobic consortium Distillery spent wash 147.5 ± 2.4 L [74]
Biomethane
Anaerobic digestion Spent coffee grounds 36.0 mL-CH4 /g-VS added [117]
Anaerobic digestion Vegetable and meat waste 439.6 mL/gVS [91]
Anaerobic digestion Food and plastic waste 520.8 mL CH4 /gVS [118]
Spontaneous Canteen food waste 80% [119]
acidification
Anaerobic digestion Municipal solid waste 429.16 mL/g VS [120]
Sargassum spp. Food waste 615.5 L of CH4 /kg [121]
Anaerobic digester Sugarcane waste 36.24 [122]
Anaerobic digester Vegetable waste 247.3 mL/g of VS [123]
Anaerobic activated Food waste 18.4 g/L (as acetic acid) [124]
sludge
(continued)
18 Manifesting Sustainability Toward Food Waste into Bioenergy … 445

Table 18.2 (continued)


System Waste Yield Reference
Anaerobic pretreated Food waste 11.1 g/L [101]
sludge
Anaerobic activated Food waste 29.1 g-COD/L [125]
sludge
Anaerobic activated Food waste 0.918 g/g [126]
sludge
Anaerobic activated Food waste 6.3 g/L [127]
sludge

18.4.3 Solventogenesis

18.4.3.1 Bioethanol

Global demand for ethanol production reached 100 billion L in 2015 and continues
to rise, driven by increasing needs. To address this increasing demand, a critical
transition occurs from fossil-derived ethanol production to renewable resource-
based methods [132]. Presently, pilot-scale bioethanol production is predominantly
centered around sugarcane and corn. However, a significant challenge arises because
sugarcane and corn are used for bioethanol and serve as food and animal feed, creating
competition between food and fuel resources. Research has focused on utilizing agri-
cultural waste such as rice straw, sugar cane bagasse, wheat straw, and other biomass
(Table 18.3) [133]. One of the major hurdles in this transition is the interference
of lignin converting cellulose and hemicelluloses into monosaccharides, posing a
significant limitation to ethanol production. The implementation of simultaneous
saccharification and fermentation in a single stage resulted in diminished ethanol
yields, specifically amounting to 0.31 g of ethanol per gram of total solids [134]. The
high organic content found in food waste makes it a promising option for bioethanol
production, and it offers a potential solution to the food-versus-fuel challenge.

18.4.3.2 Biobutanol

Biobutanol (butyl alcohol), serves various purposes as a extractant, solvent, supple-


ment, eluent, and precursor for other compounds. The biobutanol blended with gaso-
line used as a transportation fuel. Biobutanol possesses advantageous properties such
as low vapor pressure, insensitivity to water, lower toxicity, reduced volatility, and
flammability. Notably, biobutanol can replace gasoline, making it a drop-in fuel of
significant interest. Biobutanol production primarily relies on starch-rich crops like
wheat, rice, maize, energy crops, cassava, trees, grasses, and algae (Table 18.3).
Therefore, bio-butanol is categorized as first, second, or third-generation depending
on the substrate utilized [157]. The main production process involves extracting
446 D. S. Rajendran et al.

Table 18.3 Various food wastes used for the production of different biobutanol and bioethanol
products
System Waste Yield Reference
Biobutanol
Clostridium beijerinckii Carrot waste 6.4 g/L [135]
Clostridium Starchy food waste 12.1 g/L [136]
saccharoperbutylacetonicum N1-4
Clostridium sp. strain BOH3 Orange peel 17.7 g/L [137]
Clostridium beijerinckii P260 Vegetable waste 59% [138]
Clostridium sp. strain BOH3 Starchy food waste 14.1 g/L [139]
Clostridium acetobutylicum Vegetable waste 1.5 g/L [140]
C. acetobutylicum CICC 8012 Meat waste 12.8 g/L [123]
Clostridium acetobutylicum YM1 Cooked rice 6.01 g/L [141]
C. acetobutylicum DSM 792 Apple peel waste 14 g/L [142]
C. acetobutylicum NCIM 2877 Orange peel waste 19.5 g/L [143]
C. acetobutylicum ATCC 824 Sweet potato vines 6.4 g/L [144]
Bioethanol
Saccharomyces cerevisiae Municipal food 77% [145]
waste
Acid extraction Municipal food 1.29 kg CO2 equiv. [146]
waste per 1 gal
Enzymatic hydrolysis Vegetable food 73.2 g/L [147]
waste
Saccharomyces cerevisiae Potato and wheat 7.52% (v/v) [148]
bran
Enzymatic saccharification Carbohydrate food 75.70% [149]
waste
Batch fermentation Carrot discard juices 11.98 g/L [150]
Ensiling + centrifugation Orange peel waste 120 g/kg TS [151]
Saccharomyces cerevisiae Spent coffee 8.3%v/v [152]
grounds
Saccharomyces cerevisiae 424A Corn stover 68.30% [153]
Saccharomyces cerevisiae Pe-2 Sugarcane bagasse 85.41–88.12 [154]
Hydrolysis and anaerobic Waste hamburger 0.271 g ethanol/g [155]
fermentation WH
Fast pyrolysis Spent coffee 29% [156]
grounds
18 Manifesting Sustainability Toward Food Waste into Bioenergy … 447

sugars from the substrate through various pretreatment methods and then utilizing
these extracts with strains from the Clostridiaceae family. Interestingly, domestic
food waste, which is rich in carbon, can also be utilized for producing biobutanol
[158].

18.4.4 Oleaginous Metabolism

Oleaginous metabolism holds substantial importance within the realm of biore-


finery operations. Biorefineries are establishments dedicated to the transformation
of diverse biomass sources, including agricultural and forestry residues, as well as
algae and other organic materials like food waste. The objective is to convert these
resources into a broad spectrum of valuable end products, encompassing biofuels,
biochemicals, and various other bioproducts. Oleaginous microorganisms, which
include specific yeasts, algae, and bacteria, have garnered notable attention in the
biorefinery context due to their remarkable capacity to accumulate significant quanti-
ties of lipids, which are essentially fats and oils. This lipid accumulation serves as an
effective energy storage mechanism. This process offers the potential for lipid produc-
tion, biofuel generation (such as biodiesel), waste reduction, resource recovery, and
environmental benefits, aligning with sustainability goals by efficiently reusing and
converting organic waste into useful products, thus promoting a more circular and
eco-friendly approach to resource utilization.

18.4.4.1 Biodiesel

Biodiesel gives a sustainable energy source derived from natural renewable materials.
Chemically, it comprises long-chain fatty acids of methyl esters. In acid-catalyzed
transesterification, acid initiates the reaction by donating a proton to the oxygen
atom of the ester functional group, which then undergoes nucleophilic attack by an
alcohol molecule. This leads to the formation of the desired ester products (biodiesel)
and glycerol as a by-product. Base-catalyzed transesterification, the base removes
a proton from the alcohol, creating an alkoxide ion that is more nucleophilic. This
ion then attacks the ester, leading to the formation of biodiesel and glycerol. Tradi-
tionally, biodiesel is manufactured using various feedstocks such as waste cooking
oils, vegetable oils, and animal fats. However, the limited availability of these feed-
stocks hampers the further growth of the biodiesel industry. A range of cost-effective
carbon sources, including molasses [144], flour extracts, grape must, radish brine,
hydrolyzed agricultural residues [52], municipal wastewater [79], distillery wastew-
ater, cheese whey [159], food and feed waste [160], and municipal wastewater have
been successfully employed as primary feedstock [79].
448 D. S. Rajendran et al.

18.5 Circular Bioeconomy and Biorefinery

The increasing global demand for energy and resources is driving a shift from a
fossil-based linear economy to a sustainable circular bioeconomy. The concept of
a circular economy aims to modernize the traditional recycling of materials within
natural ecosystems. The bioeconomy relies on renewable feedstocks that can produce
a wide array of biobased products, effectively functioning as a self-contained refinery.
It introduces the notion of cascading, involving the diversified utilization of materials
through a series of production processes. Cascading is achieved through intercon-
nected networks of factories, where organic by-products, instead of being disposed
of in landfills or returned directly to the soil, are repurposed as inputs for producing
value-added products. This process transforms food waste into a valuable resource for
biorefineries, where biological materials are progressively converted into bio-based
chemicals, plastics, pharmaceuticals, and fuel. Also, this transition encompasses a
broad spectrum of scientific, managerial, and engineering disciplines [161]. One
intriguing approach to maximize the value derived from food waste is to reintegrate
it into the food supply chain. This concept modernizes the traditional practice of soil
restoration by repurposing food residues as inputs for the agri-food industry.
The biogenic waste emerges as a promising feedstock for fostering a bioeconomy
[44]. The approach involves the integration of biogenic waste into an integrated
biorefinery framework to produce bio-based products, ultimately leading to the
establishment of a sustainable waste-based bioeconomy (Fig. 18.2). This integrated
approach combines various enabling bioprocesses within a cascading loop, including
acidogenesis, fermentation, methanogenesis, oleaginous processes, solventogenesis,
photosynthesis, bio-electrogenesis, and more. Such integration facilitates the efficient
conversion of waste into economically valuable products.
Food waste is generated throughout the entire food supply chain (FSC), with
a significant portion originating from the consumption stage (56%) and the food

Fig. 18.2 Sustainability toward food waste into bioenergy in a circular economic approach
18 Manifesting Sustainability Toward Food Waste into Bioenergy … 449

manufacturing stage (around 38%). While food waste from the consumption stage
varies widely in composition, food waste from food processing is notably consistent
in composition, making it a valuable resource for extracting high-value products.
Regrettably, this type of food waste is often directed to landfills, incurring economic
and environmental costs, or at best, repurposed for animal feed or energy recovery
processes like anaerobic digestion. To prevent this loss or underutilization of valu-
able resources, the food waste can serve as a feedstock in biorefineries specializing
in value-added products and chemicals. The biorefinery concept can yield a diverse
array of products, including biofuels, platform chemicals, biomass, bioelectricity
biomaterials, animal feed, biofertilizers and more through food waste. Importantly,
it enhances the efficiency of waste treatment processes. The food waste biorefinery
concept can serve as a valuable complement to fossil-based refineries, addressing crit-
ical drivers for the bioeconomy, including climate concerns, ecosystem services, and
resource security [161]. In the near future, the food waste biorefinery approach can
potentially establish a sustainable, environmentally friendly pathway with minimal
environmental impact.

18.6 Techno-economic Analysis

A techno-economic analysis of food waste entails an evaluation that combines tech-


nical and economic factors to assess the feasibility and sustainability of various
processes and technologies related to the management and utilization of food waste.
Research assesses the environmental and economic consequences of extracting
energy and material resources from food waste. The findings reveal that, per ton
of waste treated, anaerobic digestion exhibits the lowest environmental impacts in
13 of the 19 categories analyzed, including a net-negative global warming potential.
In-vessel composting is found to be the least environmentally sustainable option,
despite being favored over incineration in alignment with circular economy princi-
ples. Incineration emerges as the most cost-effective option (£71/t), while landfilling
proves to be the most expensive (£123/t). Managing the annual collection of food
waste around 4.9 million tons from UK households through these methods results
in the generation of 340,000 tons of CO2 equivalent and costs £452 million, accom-
panied by various other environmental problems. This conserves energy around 1.9
PJ, primarily through electricity generation via incineration. They were incinerated,
savings of £103 million and 360,000 tons of CO2 equivalent per year could be
achieved compared to current waste management, rendering incineration a carbon-
negative technology. The environmental benefits of incineration would be exceeded
if all food waste was processed through anaerobic digestion. This method could save
490,000 tons of CO2 and generate 50% more electricity per ton of waste compared
to incineration [162].
Another study conducted a techno-economic analysis of an integrated
system for the biofuels and polyhydroxyalkanoates (PHAs) production, including
450 D. S. Rajendran et al.

bioethanol, biohydrogen„ and 2,3-butanediol, using FW as the feedstock. The inte-


grated process was developed, resulting in a daily production plan of 2.01 million tons
(MT) of PHAs, biohydrogen (0.29 MT), bioethanol (4.79 MT), and 2,3-butanediol
(6.79 MT), all from food waste of 50 MT. The economic analysis revealed a positive
net present value (NPV) of $4.47 million, a return on investment (ROI) of 13.68%, a
payback period of 7.31 years, and an internal rate of return (IRR) of 11.95%. Sensi-
tivity analysis was employed to assess economic viability. The minimum selling
price (MSP) of PHAs was found to be 4.83 dollars per kilogram, with the lowest
achievable MSP (2.41 dollars per kilogram) achieved with a 30% solid loading [75].

18.7 Conclusions

The global issue of excessive food waste is a pressing concern, not only in terms
of economic losses but also for its environmental impact. The annual discard of
1.3 billion tons of food worldwide emphasizes the urgency of sustainable interven-
tions. Eco-friendly technologies and bioprocesses, such as acidogenesis, fermenta-
tion, and bio-electrogenesis, offer promising solutions for transforming food waste
into valuable bio-based products like bioelectricity, biodiesel, and biofuels. These
products align with SDG, particularly SDG 12.3 and SDG 8. Adopting a biorefinery
strategy that embraces a circular economy framework is imperative to maximize
economic viability and sustainability. By converting the food waste as a potential
resource, one can reduce waste and contribute to a more sustainable and prosperous
future for all.
The following recommendations have been put forth in relation to the implementa-
tion of SDG 12.3. Firstly, the integration of advanced technologies like data analytics
and blockchain in the food supply chain should be encouraged to enhance trace-
ability and reduce losses. Second, comprehensive consumer education and awareness
campaigns are vital to promote responsible food consumption and minimize waste
at the household level. Third, governments and international organizations should
establish and enforce regulations targeting food waste reduction, including the setting
of targets and policies. Embracing the principles of a circular economy, optimizing
food systems, and repurposing food waste should be a priority. Collaborative part-
nerships among governments, businesses, and civil society, knowledge sharing, and
investments in innovative solutions for waste repurposing are necessary. Sustain-
able packaging materials and cross-border cooperation should also be promoted.
Furthermore, standardizing metrics and data collection for food waste is essential,
and support for local and community initiatives addressing food waste is crucial.
By implementing these recommendations, SDG 12.3 can progress toward a more
sustainable and responsible approach to food management, significantly reducing
food waste and its global consequences.
The integration of biorefinery and bioeconomy principles can be effectively real-
ized, thereby fostering a closed-loop, zero-waste manufacturing system centered
around the high-value products generation from food waste. To achieve this, further
18 Manifesting Sustainability Toward Food Waste into Bioenergy … 451

research is imperative to fine-tune process efficiency and explore avenues for scal-
ability and commercialization, ensuring that forthcoming ventures are economi-
cally viable, socially acceptable, and environmentally sustainable. Additionally, it
would be advantageous to investigate the interconnectedness between food waste-
to-bioenergy conversion, sustainable food waste management, and the integrated
biorefinery and bioeconomy. This holistic approach offers a sustainable means of
valorizing food waste with minimal long-term environmental repercussions. Further-
more, addressing the issue necessitates a concerted effort to enhance public aware-
ness through training, socialization, and education. This is pivotal in shifting societal
perspectives and attitudes toward food waste and its potential for valorization.

Declaration of Competing Interest The authors declare that they have no known competing finan-
cial interests or personal relationships that could have appeared to influence the work reported in
this paper.

Funding This research did not receive any specific grant from funding agencies in the public,
commercial, or not-for-profit sectors.

References

1. Lelieveld, J., Evans, J. S., Fnais, M., Giannadaki, D., & Pozzer, A. (2015). The contribution of
outdoor air pollution sources to premature mortality on a global scale. Nature, 525, 367–371.
2. Osman, A. I., Qasim, U., Jamil, F., Ala’a, H., Jrai, A. A., Al-Riyami, M., Al-Maawali, S., Al-
Haj, L., Al-Hinai, A., & Al-Abri, M. (2021). Bioethanol and biodiesel: Bibliometric mapping,
policies and future needs. Renewable and Sustainable Energy Reviews 152, 111677.
3. Jiang, W., Kumar, A., & Adamopoulos, S. (2018). Liquefaction of lignocellulosic materials
and its applications in wood adhesives—A review. Industrial Crops and Products, 124, 325–
342.
4. Spekreijse, J., Lammens, T., Parisi, C., Ronzon, T., & Vis, M. (2019). Insights into the
European market for bio-based chemicals. Analysis Based on Ten Key Product Categories.
5. Hu, K., Peris, A., Torán, J., Eljarrat, E., Sarrà, M., Blánquez, P., & Caminal, G. (2020).
Exploring the degradation capability of Trametes versicolor on selected hydrophobic
pesticides through setting sights simultaneously on culture broth and biological matrix.
Chemosphere, 250, 126293.
6. Sarkar, O., & Venkata Mohan, S. (2016). Deciphering acidogenic process towards biohy-
drogen, biohythane, and short chain fatty acids production: Multi-output optimization strategy.
Biofuel Research Journal, 3, 458–469.
7. Anastas, P., Nolasco, M., Kerton, F., Kirchhoff, M., Licence, P., Pradeep, T., Subramaniam,
B., & Moores, A. (2021). The power of the United Nations sustainable development goals in
sustainable chemistry and engineering research. ACS Sustainable Chemistry & Engineering,
9, 8015–8017.
8. Fonseca, L. M., Domingues, J. P., & Dima, A. M. (2020). Mapping the sustainable
development goals relationships. Sustainability, 12, 3359.
9. Ardra, S., & Barua, M. K. (2022). Halving food waste generation by 2030: The challenges
and strategies of monitoring UN sustainable development goal target 12.3. Journal of Cleaner
Production, 380, 135042.
10. Rohr, J. R., Barrett, C. B., Civitello, D. J., Craft, M. E., Delius, B., DeLeo, G. A., Hudson, P.
J., Jouanard, N., Nguyen, K. H., & Ostfeld, R. S. (2019). Emerging human infectious diseases
and the links to global food production. Nature Sustainability, 2, 445–456.
452 D. S. Rajendran et al.

11. Coudard, A., Corbin, E., de Koning, J., Tukker, A., & Mogollón, J. M. (2021). Global water
and energy losses from consumer avoidable food waste. Journal of Cleaner Production, 326,
129342.
12. Usubiaga-Liaño, A., Behrens, P., & Daioglou, V. (2020). Energy use in the global food system.
Journal of Industrial Ecology, 24, 830–840.
13. Binns, C. W., Lee, M. K., Maycock, B., Torheim, L. E., Nanishi, K., & Duong, D. T. T. (2021).
Climate change, food supply, and dietary guidelines. Annual Review of Public Health, 42,
233–255.
14. Bechthold, A., Boeing, H., Tetens, I., Schwingshackl, L., & Nöthlings, U. (2018). Perspec-
tive: Food-based dietary guidelines in Europe—Scientific concepts, current status, and
perspectives. Advances in Nutrition, 9, 544–560.
15. Al-Ansari, T., Korre, A., Nie, Z., & Shah, N. (2015). Development of a life cycle assessment
tool for the assessment of food production systems within the energy, water and food nexus.
Sustainable Production and Consumption, 2, 52–66.
16. Joshi, P., & Visvanathan, C. (2019). Sustainable management practices of food waste in Asia:
Technological and policy drivers. Journal of Environmental Management, 247, 538–550.
17. Stanujkic, D., Popovic, G., Zavadskas, E. K., Karabasevic, D., & Binkyte-Veliene, A. (2020).
Assessment of progress towards achieving Sustainable Development Goals of the “Agenda
2030” by using the CoCoSo and the Shannon Entropy methods: The case of the EU Countries.
Sustainability, 12, 5717.
18. Gokarn, S., & Kuthambalayan, T. S. (2017). Analysis of challenges inhibiting the reduction
of waste in food supply chain. Journal of Cleaner Production, 168, 595–604.
19. Gokarn, S., & Choudhary, A. (2021). Modeling the key factors influencing the reduction of
food loss and waste in fresh produce supply chains. Journal of Environmental Management,
294, 113063.
20. Porter, L. (2018). From an urban country to urban country: Confronting the cult of denial in
Australian cities. Australian Geographer., 49, 239–246.
21. Gardas, B. B., Raut, R. D., & Narkhede, B. (2018). Evaluating critical causal factors for post-
harvest losses (PHL) in the fruit and vegetables supply chain in India using the DEMATEL
approach. Journal of Cleaner Production, 199, 47–61.
22. Schulz, C., & Bailey, I. (2014). The green economy and post-growth regimes: Opportunities
and challenges for economic geography. Geografiska Annaler: Series A, Physical Geography,
96, 277–291.
23. Pal, M. S., & Bhatia, M. (2022). Current status, topographical constraints, and implementation
strategy of municipal solid waste in India: A review. Arabian Journal of Geosciences, 15, 1176.
24. Alawneh, R., Ghazali, F., Ali, H., & Sadullah, A. F. (2019). A novel framework for integrating
United Nations Sustainable Development Goals into sustainable non-residential building
assessment and management in Jordan. Sustainable Cities and Society, 49, 101612.
25. Capanoglu, E., Nemli, E., & Tomas-Barberan, F. (2022). Novel approaches in the valorization
of agricultural wastes and their applications. Journal of Agriculture and Food Chemistry, 70,
6787–6804.
26. Movilla-Pateiro, L., Mahou-Lago, X. M., Doval, M. I., & Simal-Gandara, J. (2021). Toward
a sustainable metric and indicators for the goal of sustainability in agricultural and food
production. Critical Reviews in Food Science and Nutrition, 61, 1108–1129.
27. Johnston, J. L., Fanzo, J. C., & Cogill, B. (2014). Understanding sustainable diets: A descrip-
tive analysis of the determinants and processes that influence diets and their impact on health,
food security, and environmental sustainability. Advances in Nutrition, 5, 418–429.
28. Yang, N., Zhang, H., Chen, M., Shao, L.-M., & He, P.-J. (2012). Greenhouse gas emissions
from MSW incineration in China: Impacts of waste characteristics and energy recovery. Waste
Management, 32, 2552–2560.
29. Thamagasorn, M., & Pharino, C. (2019). An analysis of food waste from a flight catering
business for sustainable food waste management: A case study of halal food production
process. Journal of Cleaner Production, 228, 845–855.
18 Manifesting Sustainability Toward Food Waste into Bioenergy … 453

30. Fattibene, D., Recanati, F., Dembska, K., & Antonelli, M. (2020). Urban food waste: A
framework to analyse policies and initiatives. Resources, 9, 99.
31. Bond, M., Meacham, T., Bhunnoo, R., & Benton, T. (2013). Food waste within global food
systems. Global Food Security Swindon, UK.
32. Annosi, M. C., Brunetta, F., Bimbo, F., & Kostoula, M. (2021). Digitalization within food
supply chains to prevent food waste. Drivers, barriers and collaboration practices. Industrial
Marketing Management, 93, 208–220.
33. van den Bos Verma, M., de Vreede, L., Achterbosch, T., & Rutten, M. M. (2020). Consumers
discard a lot more food than widely believed: Estimates of global food waste using an energy
gap approach and affluence elasticity of food waste. PLoS One, 15, e0228369.
34. Babu, S., Rathore, S. S., Singh, R., Kumar, S., Singh, V. K., Yadav, S. K., Yadav, V., Raj, R.,
Yadav, D., & Shekhawat, K. (2022). Exploring agricultural waste biomass for energy, food
and feed production and pollution mitigation: A review. Bioresource Technology, 127566.
35. Sánchez, M., Laca, A., Laca, A., & Díaz, M. (2021). Value-added products from fruit and
vegetable wastes: A review. CLEAN–Soil, Air, Water, 49, 2000376.
36. Kunatsa, T., & Xia, X. (2022). A review on anaerobic digestion with focus on the role of
biomass co-digestion, modelling and optimisation on biogas production and enhancement.
Bioresource Technology, 344, 126311.
37. Mu, H., Zhao, C., Zhao, Y., Hua, D., Li, Y., Zhang, X., & Cao, Y. (2019). Optimizing methane
production from anaerobic batch co-digestion of apple pomace and vegetable waste. Journal
of Biobased Materials and Bioenergy, 13, 508–516.
38. Yan, F., Hu, H., Lu, L., & Zheng, X. (2016). Rhamnolipids induce oxidative stress responses in
cherry tomato fruit to Alternaria alternate. Pest Management Science, 72, 1500–1507. https://
doi.org/10.1002/ps.4177
39. Tlusty, M. F., Tyedmers, P., Bailey, M., Ziegler, F., Henriksson, P. J. G., Béné, C., Bush, S.,
Newton, R., Asche, F., & Little, D. C. (2019). Reframing the sustainable seafood narrative.
Global Environmental Change, 59, 101991.
40. Calderwood, J., Pedreschi, D., & Reid, D. G. (2021). Technical and tactical measures to reduce
unwanted catches in mixed fisheries: Do the opinions of Irish fishers align with management
advice? Marine Policy, 123, 104290.
41. Marti-Quijal, F. J., Remize, F., Meca, G., Ferrer, E., Ruiz, M.-J., & Barba, F. J. (2020).
Fermentation in fish and by-products processing: An overview of current research and future
prospects. Current Opinion in Food Science, 31, 9–16.
42. Venugopal, V. (2021). Valorization of seafood processing discards: Bioconversion and bio-
refinery approaches. Frontiers in Sustainable Food Systems, 5, 611835.
43. Pérez-Marroquín, X. A., Estrada-Fernández, A. G., García-Ceja, A., Aguirre-Álvarez, G., &
León-López, A. (2023). Agro-food waste as an ingredient in functional beverage processing:
Sources. Functionality, Market and Regulation, Foods, 12, 1583.
44. Kumar, P. S., Mohanakrishna, G., Hemavathy, R. V., Rangasamy, G., & Aminabhavi, T.
M. (2022). Sustainable production of biosurfactants via valorisation of industrial wastes as
alternate feedstocks. Chemosphere, 137326.
45. Hameed, A., Anwar, M. J., Perveen, S., Amir, M., Naeem, I., Imran, M., Hussain, M., Ahmad,
I., Afzal, M. I., & Inayat, S. (2023). Functional, industrial and therapeutic applications of dairy
waste materials. International Journal of Food Properties, 26, 1470–1496.
46. Gutierrez-Hernandez, C. A., Hernandez-Almanza, A., Hernandez-Beltran, J. U., Balagu-
rusamy, N., & Hernandez-Teran, F. (2022). Cheese whey valorization to obtain single-cell
oils of industrial interest: An overview. Food Bioscience, 102086.
47. Poonia, A., & Pandey, S. (2023). Production of microbial pigments from whey and their
applications: A review. Nutrition & Food Science, 53, 265–284.
48. Hosseinzadeh-Bandbafha, H., Li, C., Chen, X., Peng, W., Aghbashlo, M., Lam, S. S., &
Tabatabaei, M. (2022). Managing the hazardous waste cooking oil by conversion into bioen-
ergy through the application of waste-derived green catalysts: A review. Journal of Hazardous
Materials, 424, 127636.
454 D. S. Rajendran et al.

49. Di Fraia, S., Massarotti, N., Prati, M. V., & Vanoli, L. (2020). A new example of circular
economy: Waste vegetable oil for cogeneration in wastewater treatment plants. Energy
Conversion and Management, 211, 112763. https://doi.org/10.1016/j.enconman.2020.112763
50. Sen Siow, H., Sudesh, K., Murugan, P., & Ganesan, S. (2021). Mealworm (Tenebrio molitor)
oil characterization and optimization of the free fatty acid pretreatment via acid-catalyzed
esterification. Fuel, 299, 120905.
51. Cao, Y., Doustgani, A., Salehi, A., Nemati, M., Ghasemi, A., & Koohshekan, O. (2020). The
economic evaluation of establishing a plant for producing biodiesel from edible oil wastes in
oil-rich countries: Case study Iran. Energy, 213, 118760.
52. Yu, X., Zheng, Y., Dorgan, K. M., & Chen, S. (2011). Oil production by oleaginous yeasts
using the hydrolysate from pretreatment of wheat straw with dilute sulfuric acid. Bioresource
Technology, 102, 6134–6140.
53. Kee, S. H., Chiongson, J. B. V., Saludes, J. P., Vigneswari, S., Ramakrishna, S., & Bhubalan,
K. (2021). Bioconversion of agro-industry sourced biowaste into biomaterials via microbial
factories—A viable domain of circular economy. Environmental Pollution, 271, 116311.
54. Chiarelotto, M., Restrepo, J. C. P. S., Lorin, H. E. F., & Damaceno, F. M. (2021). Composting
organic waste from the broiler production chain: A perspective for the circular economy.
Journal of Cleaner Production, 329, 129717.
55. Tabler, T. (2019). Environment-friendly management of livestock and poultry mortalities. In
Animal environment and welfare.
56. Chiarelotto, M., Restrepo, J. C. P. S., Lorin, H. E. F., & Damaceno, F. M. (2021). Composting
organic waste from the broiler production chain: A perspective for the circular economy.
Journal of Cleaner Production, 329, 129717. https://doi.org/10.1016/j.jclepro.2021.129717
57. Piyani, H. O., & Chandrarin, G. (2023). Analysis of the influence of financial literacy on
business sustainability through the utilization of e-commerce: A study of MSMEs in the
food and beverage industry sector in Balikpapan City. European Journal of Business and
Management Research, 8, 306–314.
58. Salanță, L. C., Coldea, T. E., Ignat, M. V., Pop, C. R., Tofană, M., Mudura, E., Borșa, A.,
Pasqualone, A., & Zhao, H. (2020). Non-alcoholic and craft beer production and challenges.
Processes, 8, 1382.
59. Iram, A., Cekmecelioglu, D., & Demirci, A. (2020). Distillers’ dried grains with solubles
(DDGS) and its potential as fermentation feedstock. Applied Microbiology and Biotechnology,
104, 6115–6128.
60. Portarapillo, M., Danzi, E., Sanchirico, R., Marmo, L., & Di Benedetto, A. (2021). Energy
recovery from vinery waste: Dust explosion issues. Applied Sciences, 11, 11188.
61. Bhatnagar, N., Ryan, D., Murphy, R., & Enright, A. M. (2022). A comprehensive review of
green policy, anaerobic digestion of animal manure and chicken litter feedstock potential—
Global and Irish perspective. Renewable and Sustainable Energy Reviews, 154, 111884.
62. Orr, L., Goossens, Y., Heinrich, M., & Brüggemann, N. (2023). Jointly reducing food waste—
The experiences of the German discussion forum for wholesale and retail. Sustainability, 15,
12289.
63. Wikurendra, E. A., Abdeljawad, N. S., & Nagy, I. (2023). A review of municipal waste manage-
ment with zero waste concept: Strategies, potential and challenge in Indonesia. International
Journal of Environmental Science and Development, 14.
64. Ögel, İY., Ecer, F., & Özgöz, A. A. (2023). Identifying the leading retailer-based food waste
causes in different perishable fast-moving consumer goods’ categories: Application of the
F-LBWA methodology. Environmental Science and Pollution Research, 30, 32656–32672.
65. Bos-Brouwers, H., Achterbosch, T. J., Castelein, B., & Cloutier, J. M. I. (2023). Theory of
change: Acceleration agenda for reducing food waste 2022–2025: Recommendations to the
Ministry of Agriculture, Nature and Food Quality (LNV). Wageningen Food & Biobased
Research.
66. Redman, E., & Redman, A. (2014). Transforming sustainable food and waste behaviors by
realigning domains of knowledge in our education system. Journal of Cleaner Production,
64, 147–157.
18 Manifesting Sustainability Toward Food Waste into Bioenergy … 455

67. Onoh, C. E., & Ogbuagu, T. C. (2023). The impact of social work intervention on sustainable
consumption through food waste reduction in Gävle Sweden: A qualitative study on the
environmental and socio-economic benefits.
68. Munir, K. (2022). Sustainable food waste management strategies by applying practice
theory in hospitality and food services—A systematic literature review. Journal of Cleaner
Production, 331, 129991.
69. Graham-Rowe, E., Jessop, D. C., & Sparks, P. (2014). Identifying motivations and barriers to
minimising household food waste. Resources, Conservation and Recycling, 84, 15–23.
70. Goud, R. K., Arunasri, K., Yeruva, D. K., Krishna, K. V., Dahiya, S., & Mohan, S. V. (2017).
Impact of selectively enriched microbial communities on long-term fermentative biohydrogen
production. Bioresource Technology, 242, 253–264.
71. Ishii, S., Suzuki, S., Norden-Krichmar, T. M., Wu, A., Yamanaka, Y., Nealson, K. H., &
Bretschger, O. (2013). Identifying the microbial communities and operational conditions for
optimized wastewater treatment in microbial fuel cells. Water Research, 47, 7120–7130.
72. Dahiya, S., Kumar, A. N., Sravan, J. S., Chatterjee, S., Sarkar, O., & Mohan, S. V. (2018). Food
waste biorefinery: Sustainable strategy for circular bioeconomy. Bioresource Technology, 248,
2–12.
73. Xia, X., Tokash, J. C., Zhang, F., Liang, P., Huang, X., & Logan, B. E. (2013). Oxygen-reducing
biocathodes operating with passive oxygen transfer in microbial fuel cells. Environmental
Science and Technology, 47, 2085–2091.
74. Pasupuleti, S. B., & Mohan, S. V. (2015). Single-stage fermentation process for high-value
biohythane production with the treatment of distillery spent-wash. Bioresource Technology,
189, 177–185.
75. Rajendran, N., & Han, J. (2022). Techno-economic analysis of food waste valorization for
integrated production of polyhydroxyalkanoates and biofuels. Bioresource Technology, 348,
126796.
76. Mahmoud, A. H., Hussein, M. Y., Ibrahim, H. M., Hanafy, M. H., Salah, S. M., El-Bassiony, G.
M., & Abdelfattah, E. A. (2022). Mixed microalgae-food waste cake for feeding of Hermetia
illucens larvae in characterizing the produced biodiesel. Biomass and Bioenergy, 165, 106586.
https://doi.org/10.1016/j.biombioe.2022.106586
77. Talapatra, N., & Ghosh, U. K. (2022). New concept of biodiesel production using food waste
digestate powder: Co-culturing algae-activated sludge symbiotic system in low N and P paper
mill wastewater. Science of The Total Environment, 844, 157207
78. Suzihaque, M. U. H., Syazwina, N., Alwi, H., Ibrahim, U. K., Abdullah, S., & Haron, N.
(2022). A sustainability study of the processing of kitchen waste as a potential source of
biofuel: Biodiesel production from waste cooking oil (WCO). Materials Today: Proceedings,
63, S484–S489.
79. Chi, Z., Zheng, Y., Jiang, A., & Chen, S. (2011b). Lipid production by culturing oleaginous
yeast and algae with food waste and municipal wastewater in an integrated process. Applied
Biochemistry and Biotechnology, 165, 442–453.
80. Kahr, H., Pointner, M., Krennhuber, K., Wallner, B., & Jäger, A. (2015). Lipid production
from diverse oleaginous yeasts from steam exploded corn cobs. Agronomy Research, 13(2).
81. Liang, Y., Tang, T., Siddaramu, T., Choudhary, R., & Umagiliyage, A. L. (2012). Lipid produc-
tion from sweet sorghum bagasse through yeast fermentation. Renewable Energy, 40(1),
130–136.
82. Pleissner, D., Lam, W. C., Sun, Z., & Lin, C. S. K. (2013). Food waste as nutrient source in
heterotrophic microalgae cultivation. Bioresource Technology, 137, 139–146.
83. Papone, T., Kookkhunthod, S., Paungbut, M., & Leesing, R. (2016). Producing of microbial
oil by mixed culture of microalgae and oleaginous yeast using sugarcane molasses as carbon
substrate. J. Clean Energy Technol, 4(4), 253–256.
84. Haile, M. (2014). Integrated volarization of spent coffee grounds to biofuels. Biofuel Research
Journal, 1, 65–69.
85. Takakuwa, N., & Saito, K. (2010). Conversion of beet molasses and cheese whey into fatty acid
methyl esters by the yeast Cryptococcus curvatus. Journal of Oleo Science, 59(5), 255–260.
456 D. S. Rajendran et al.

86. Papanikolaou, S., Dimou, A., Fakas, S., Diamantopoulou, P., Philippoussis, A., Galiotou-
Panayotou, M., & Aggelis, G. (2011). Biotechnological conversion of waste cooking olive
oil into lipid-rich biomass using Aspergillus and Penicillium strains. Journal of Applied
Microbiology, 110(5), 1138–1150.
87. Bialy, H. El, Gomaa, O. M., & Azab, K. S. (2011). Conversion of oil waste to valuable fatty
acids using oleaginous yeast. World Journal of Microbiology and Biotechnology, 27(12),
2791–2798.
88. Vescovi, V., Rojas, M. J., Baraldo, A., Botta, D. C., Santana, F. A. M., Costa, J. P., Machado, M.
S., Honda, V. K., de Lima Camargo Giordano, R., & Tardioli, P. W. (2016). Lipase-catalyzed
production of biodiesel by hydrolysis of waste cooking oil followed by esterification of free
fatty acids. Journal of the American Oil Chemists’ Society, 93, 1615–1624.
89. Rocha, M. V. P., de Matos, L. J. B. L., de Lima, L. P., da Silva Figueiredo, P. M., Lucena,
I. L., Fernandes, F. A. N., & Gonçalves, L. R. B. (2014). Ultrasound-assisted production of
biodiesel and ethanol from spent coffee grounds. Bioresource Technology, 167, 343–348
90. Jung, J.-H., Sim, Y.-B., Ko, J., Park, S. Y., Kim, G.-B., & Kim, S.-H. (2022). Biohydrogen
and biomethane production from food waste using a two-stage dynamic membrane bioreactor
(DMBR) system. Bioresource Technology, 352, 127094.
91. Ding, L., Cheng, J., Qiao, D., Li, H., & Zhang, Z. (2022a). Continuous co-generation of biohy-
drogen and biomethane through two-stage anaerobic digestion of hydrothermally pretreated
food waste. Energy Conversion and Management, 268, 116000.
92. Moogi, S., Lam, S. S., Chen, W.-H., Ko, C. H., Jung, S.-C., & Park, Y.-K. (2022). Household
food waste conversion to biohydrogen via steam gasification over copper and nickel-loaded
SBA-15 catalysts. Bioresource Technology, 366, 128209
93. Nyambura, S. M., Jufei, W., Hua, L., Xuebin, F., Xingjia, P., Bohong, L., Ahmad, R., Jialiang,
X., Bertrand, G. V, Ndiithi, J., & Xuhui, L. (2022). Microwave co-pyrolysis of kitchen food
waste and rice straw for waste reduction and sustainable biohydrogen production: Thermo-
kinetic analysis and evolved gas analysis. Sustainable Energy Technologies and Assessments,
52, 102072
94. Mirza, S. S., Qazi, J. I., Liang, Y., & Chen, S. (2019). Growth characteristics and photofer-
mentative biohydrogen production potential of purple non sulfur bacteria from sugar cane
bagasse. Fuel, 255, 115805.
95. Yang, G., Hu, Y., & Wang, J. (2019). Biohydrogen production from co-fermentation of fallen
leaves and sewage sludge. Bioresource Technology, 285, 121342.
96. Marone, A., Massini, G., Patriarca, C., Signorini, A., Varrone, C., & Izzo, G. (2012).
Hydrogen production from vegetable waste by bioaugmentation of indigenous fermentative
communities. International Journal of Hydrogen Energy, 37(7), 5612–5622.
97. Wang, Y.-H., Li, S.-L., Chen, I.-C., & Cheng, S.-S. (2009). Starch hydrolysis characteristics
of hydrogen producing sludge in thermophilic hydrogen fermentor fed with kitchen waste.
International Journal of Hydrogen Energy, 34(17), 7435–7440.
98. Goud, R. K., & Mohan, S. V. (2011). Pre-fermentation of waste as a strategy to enhance
the performance of single chambered microbial fuel cell (MFC). International Journal of
Hydrogen Energy, 36(21), 13753–13762.
99. El-Bery, H., Tawfik, A., Kumari, S., & Bux, F. (2013). Effect of thermal pre-treatment on
inoculum sludge to enhance bio-hydrogen production from alkali hydrolysed rice straw in a
mesophilic anaerobic baffled reactor. Environmental Technology, 34(13–14), 1965–1972.
100. Ren, H.-Y., Kong, F., Cui, Z., Zhao, L., Ma, J., Ren, N.-Q., & Liu, B.-F. (2019). Cogeneration
of hydrogen and lipid from stimulated food waste in an integrated dark fermentative and
microalgal bioreactor. Bioresource Technology, 287, 121468.
101. Karthikeyan, O. P., & Visvanathan, C. (2013). Bio-energy recovery from high-solid organic
substrates by dry anaerobic bio-conversion processes: A review. Reviews in Environmental
Science & Biotechnology, 12, 257–284.
102. Mahmod, S. S., Azahar, A. M., Tan, J. P., Jahim, J. M., Abdul, P. M., Mastar, M. S., Anuar,
N., Yunus, M. F. M., Asis, A. J., & Wu, S.-Y. (2019). Operation performance of up-flow
anaerobic sludge blanket (UASB) bioreactor for biohydrogen production by self-granulated
18 Manifesting Sustainability Toward Food Waste into Bioenergy … 457

sludge using pre-treated palm oil mill effluent (POME) as carbon source. Renewable Energy,
134, 1262–1272.
103. Gorgeç, F. K., & Karapinar, I. (2019). Production of biohydrogen from waste wheat in contin-
uously operated UPBR: the effect of influent substrate concentration. International Journal
of Hydrogen Energy, 44(32), 17323–17333.
104. Mohan, S. V., Mohanakrishna, G., Goud, R. K., & Sarma, P. N. (2009). Acidogenic fermenta-
tion of vegetable based market waste to harness biohydrogen with simultaneous stabilization.
Bioresource Technology, 100, 3061–3068.
105. Mohan, S. V., Chiranjeevi, P., Chandrasekhar, K., Babu, P. S., & Sarkar, O. (2019). Acidogenic
biohydrogen production from wastewater. In Biohydrogen (pp. 279–320). Elsevier.
106. Deepanraj, B., Sivasubramanian, V., & Jayaraj, S. (2017). Multi-response optimization of
process parameters in biogas production from food waste using Taguchi-Grey relational
analysis. Energy Conversion and Management, 141, 429–438.
107. Kondusamy, D., & Kalamdhad, A. S. (2014). Pre-treatment and anaerobic digestion of food
waste for high rate methane production—A review. Journal of Environmental Chemical
Engineering, 2, 1821–1830.
108. Xu, Z., Zhao, M., Miao, H., Huang, Z., Gao, S., & Ruan, W. (2014). In situ volatile fatty
acids influence biogas generation from kitchen wastes by anaerobic digestion. Bioresource
Technology, 163, 186–192.
109. Agyeman, F. O., & Tao, W. (2014). Anaerobic co-digestion of food waste and dairy manure:
Effects of food waste particle size and organic loading rate. Journal of Environmental
Management, 133, 268–274.
110. Bolzonella, D., Battista, F., Cavinato, C., Gottardo, M., Micolucci, F., Lyberatos, G., & Pavan,
P. (2018). Recent developments in biohythane production from household food wastes: a
review. Bioresource Technology, 257, 311–319.
111. Deheri, C., & Acharya, S. K. (2022). Purified biohythane (biohydrogen+biomethane) produc-
tion from food waste using CaO2+CaCO3 and NaOH as additives. International Journal of
Hydrogen Energy, 47(5), 2862–2873.
112. Chen, G., Li, H., Song, Y., Mu, L., Pei, L., Zhou, T., Qing, Z., & Zeng, Y. (2023). Optimiza-
tion of food-to-microorganism ratio and addition of Tween 80 to enhance biohythane produc-
tion via two-stage anaerobic digestion of food waste. Journal of Chemical Technology &
Biotechnology, 98(10), 2477–2488.
113. Nguyen, T.-T., Ta, D.-T., Lin, C.-Y., Chu, C.-Y., & Ta, T.-M.-N. (2022). Biohythane production
from swine manure and pineapple waste in a single-stage two-chamber digester using gel-
entrapped anaerobic microorganisms. International Journal of Hydrogen Energy, 47(60),
25245–25255.
114. Alavi-Borazjani, S. A., Tarelho, L. A. da C., & Capela, M. I. (2022). Addition of biomass
ash as a promising strategy for high-value biohythane production from the organic fraction
of municipal solid waste. Biomass and Bioenergy, 159, 106392.
115. Alavi-Borazjani, S. A., da Cruz Tarelho, L. A., Marques Ruivo, L. C., Paula da Silva Seabra,
M., & Capela, M. I. (2023). Effects of adding micro- and nano-sized biomass fly ash on
two-stage biohythane production from the urban organic solid waste. International Journal
of Hydrogen Energy.
116. Yeshanew, M. M., Frunzo, L., Pirozzi, F., Lens, P. N. L., & Esposito, G. (2016b). Production
of biohythane from food waste via an integrated system of continuously stirred tank and
anaerobic fixed bed reactors. Bioresource Technology, 220, 312–322.
117. Lee, M., Yang, M., Choi, S., Shin, J., Park, C., Cho, S.-K., & Kim, Y. M. (2019). Sequential
production of lignin, fatty acid methyl esters and biogas from spent coffee grounds via an
integrated physicochemical and biological process. Energies, 12(12), 2360.
118. Zhang, L., Yao, D., Tsui, T.-H., Loh, K.-C., Wang, C.-H., Dai, Y., & Tong, Y. W. (2022).
Plastic-containing food waste conversion to biomethane, syngas, and biochar via anaerobic
digestion and gasification: Focusing on reactor performance, microbial community analysis,
and energy balance assessment. Journal of Environmental Management, 306
458 D. S. Rajendran et al.

119. Wehner, M., Kleidorfer, I., Whittle, I., Bischof, D., Bockreis, A., Insam, H., Mueller, W., &
Hupfauf, S. (2023). Decentralised system for demand-oriented collection of food waste—
Assessment of biomethane potential, pathogen development and microbial community
structure. Bioresource Technology, 376, 128894.
120. Agrawal, A. V., Chaudhari, P. K., & Ghosh, P. (2023). Effect of mixing ratio on biomethane
potential of anaerobic co-digestion of fruit and vegetable waste and food waste. Biomass
Conversion and Biorefinery.
121. Castro, Y. A., Rodríguez, A., & Rivera, E. (2022). Biomethane production kinetics during the
anaerobic co-digestion of Sargassum spp. and food waste using batch and fed-batch systems
in Punta Cana, Dominican Republic. Materials for Renewable and Sustainable Energy, 11(3),
287–297.
122. Alruqi, M., & Sharma, P. (2023). Biomethane Production from the Mixture of Sugarcane
Vinasse, Solid Waste and Spent Tea Waste: A Bayesian Approach for Hyperparameter
Optimization for Gaussian Process Regression. Fermentation, 9(2), 120.
123. Liu, J., Fan, S., Bai, K., & Xiao, Z. (2021). Combining acetone-butanol-ethanol production
and methyl orange decolorization in wastewater by fermentation with solid food waste as
substrate. Renewable Energy, 179, 2246–2255
124. Li, Y., Su, D., Feng, H., Yan, F., Liu, H., Feng, L., & Liu, G. (2017). Anaerobic acidogenic
fermentation of food waste for mixed-acid production. Energy Sources, Part A: Recovery,
Utilization, and Environmental Effects, 39(7), 631–635.
125. Hong, C., & Haiyun, W. (2010). Optimization of volatile fatty acid production with co-
substrate of food wastes and dewatered excess sludge using response surface methodology.
Bioresource Technology, 101(14), 5487–5493
126. Wang, K., Yin, J., Shen, D., & Li, N. (2014). Anaerobic digestion of food waste for volatile
fatty acids (VFAs) production with different types of inoculum: effect of pH. Bioresource
Technology, 161, 395–401.
127. Dahiya, S., & Joseph, J. (2015). High rate biomethanation technology for solid waste manage-
ment and rapid biogas production: an emphasis on reactor design parameters. Bioresource
Technology, 188, 73–78.
128. Pasupuleti, S. B., Sarkar, O., & Mohan, S. V. (2014). Upscaling of biohydrogen production
process in semi-pilot scale biofilm reactor: Evaluation with food waste at variable organic
loads. International Journal of Hydrogen Energy, 39, 7587–7596.
129. Sen, B., Aravind, J., Kanmani, P., & Lay, C.-H. (2016). State of the art and future concept
of food waste fermentation to bioenergy. Renewable and Sustainable Energy Reviews, 53,
547–557.
130. Cavinato, C., Giuliano, A., Bolzonella, D., Pavan, P., & Cecchi, F. (2012). Bio-hythane produc-
tion from food waste by dark fermentation coupled with anaerobic digestion process: A long-
term pilot scale experience. International Journal of Hydrogen Energy, 37, 11549–11555.
131. Yeshanew, M. M., Frunzo, L., Pirozzi, F., Lens, P. N. L., & Esposito, G. (2016). Production
of biohythane from food waste via an integrated system of continuously stirred tank and
anaerobic fixed bed reactors. Bioresource Technology, 220, 312–322.
132. Sarkar, N., Ghosh, S. K., Bannerjee, S., & Aikat, K. (2012). Bioethanol production from
agricultural wastes: An overview. Renewable Energy, 37, 19–27.
133. Humbird, D., Davis, R., Tao, L., Kinchin, C., Hsu, D., Aden, A., Schoen, P., Lukas, J.,
Olthof, B., & Worley, M. (2011). Process design and economics for biochemical conversion
of lignocellulosic biomass to ethanol: Dilute-acid pretreatment and enzymatic hydrolysis of
corn stover. National Renewable Energy Laboratory (NREL), Golden, CO (United States).
134. Walker, K., Vadlani, P., Madl, R., Ugorowski, P., & Hohn, K. L. (2013). Ethanol fermentation
from food processing waste. Environmental Progress & Sustainable Energy, 32, 1280–1283.
135. López-Linares, J. C., Coca, M., Plaza, P. E., Lucas, S., & García-Cubero, M. T. (2023).
Waste-to-fuel technologies for the bioconversion of carrot discards into biobutanol. Renewable
Energy, 202, 362–369.
136. Wang, X., Sun, H., Wang, Y., Wang, F., Zhu, W., Wu, C., Wang, Q., & Gao, M. (2023). Feasi-
bility of Efficient, Direct, Butanol Production from Food Waste without Nutrient Supplement
by Clostridium saccharoperbutylacetonicum N1-4. Sustainability, 15(7), 6061
18 Manifesting Sustainability Toward Food Waste into Bioenergy … 459

137. Su, G., Chan, C., & He, J. (2022). Enhanced biobutanol production from starch waste via
orange peel doping. Renewable Energy, 193, 576–583.
138. Teresa Sponza, D., & Öztekin, R. (2022). 1-Butanol Production from the Food Wastes in
Dokuz Eylül University Campus Buca-Izmir, Turkey. Open J Case Reports.
139. Zhang, C., Li, T., Su, G., & He, J. (2020). Enhanced direct fermentation from food waste to
butanol and hydrogen by an amylolytic Clostridium. Renewable Energy, 153, 522–529
140. Zhang, C., Ling, Z., & Huo, S. (2021). Anaerobic fermentation of pretreated food waste for
butanol production by co-cultures assisted with in-situ extraction. Bioresource Technology
Reports, 16, 100852. https://doi.org/10.1016/j.biteb.2021.100852
141. Ozturk, A. B., Arasoglu, T., Gulen, J., Cheng, S., Al-Shorgani, N. K. N., Habaki, H., Egashira,
R., Kalil, M. S., Yusoff, W. M. W., & Cross, J. S. (2021). Techno-economic analysis of a two-
step fermentation process for bio-butanol production from cooked rice. Sustainable Energy &
Fuels, 5(14), 3705–3718.
142. Raganati, F., Procentese, A., Olivieri, G., Russo, M. E., & Marzocchella, A. (2016). Butanol
production by fermentation of fruit residues. Chemical Engineering Transactions, 49, 229–
234.
143. Joshi, S. M., Waghmare, J. S., Sonawane, K. D., & Waghmare, S. R. (2015). Bio-ethanol and
bio-butanol production from orange peel waste. Biofuels, 6(1–2), 55–61.
144. Alvarez, R. M., Rodriguez, B., Romano, J. M., Diaz, A. O., Gomez, E., Miro, D., Navarro, L.,
Saura, G., & Garcia, J. L. (1992). Lipid accumulation in Rhodotorula glutinis on sugar cane
molasses in single-stage continuous culture. World Journal of Microbiology & Biotechnology,
8, 214–215.
145. Passadis, K., Christianides, D., Malamis, D., Barampouti, E. M., & Mai, S. (2022). Valorisation
of source-separated food waste to bioethanol: pilot-scale demonstration. Biomass Conversion
and Biorefinery, 12(10), 4599–4609. https://doi.org/10.1007/s13399-022-02732-6
146. Byun, J., Kwon, O., Kim, J., & Han, J. (2022). Carbon-Negative Food Waste-Derived
Bioethanol: A Hybrid Model of Life Cycle Assessment and Optimization. ACS Sustainable
Chemistry & Engineering, 10(14), 4512–4521
147. Zhou, H., Jiang, J., Zhao, Q., Wang, Z., Li, L., Gao, Q., & Wang, K. (2023). Performance
of high solids enzymatic hydrolysis and bioethanol fermentation of food waste under the
regulation of saponin. Bioresource Technology, 387, 129486.
148. Gherbi, Y., Boudjema, K., Djeziri, M., & Fazouane–Naimi, F. (2023). Isolation and identifi-
cation of thermotolerant yeast strains producing bioethanol from agro-food wastes. Biomass
Conversion and Biorefinery.
149. Lee, J., Kim, S., Lee, K. H., Lee, S. K., Chun, Y., Kim, S. W., Park, C., & Yoo, H. Y.
(2022). Improvement of bioethanol production from waste chestnut shells via evaluation of
mass balance-based pretreatment and glucose recovery process. Environmental Technology &
Innovation, 28, 102955
150. Clementz, A. L., Manuale, D., Sanchez, E., Vera, C., & Yori, J. C. (2019). Use of discards
of bovine bone, yeast and carrots for producing second generation bio-ethanol. Biocatalysis
and Agricultural Biotechnology, 22, 101392.
151. Fazzino, F., Mauriello, F., Paone, E., Sidari, R., & Calabrò, P. S. (2021). Integral valorization
of orange peel waste through optimized ensiling: Lactic acid and bioethanol production.
Chemosphere, 271, 129602.
152. Karmee, S. K., Niemeijer, B., Casiraghi, L., Mlambo, B., Lapkin, A., & Greiner, L. (2014).
Facile biocatalytic synthesis of a macrocyclic lactone in sub-and supercritical solvents.
Biocatalysis and Biotransformation, 32(2), 125–131.
153. Jin, M., Gunawan, C., Uppugundla, N., Balan, V., & Dale, B. E. (2012). A novel inte-
grated biological process for cellulosic ethanol production featuring high ethanol productivity,
enzyme recycling and yeast cells reuse. Energy & Environmental Science, 5(5), 7168–7175.
154. Silva, C. O. G., Vaz, R. P., & Filho, E. X. F. (2018). Bringing plant cell wall-degrading
enzymes into the lignocellulosic biorefinery concept. Biofuels, Bioproducts and Biorefining,
12(2), 277–289.
460 D. S. Rajendran et al.

155. Han, W., Liu, Y., Xu, X., Huang, J., He, H., Chen, L., Qiu, S., Tang, J., & Hou, P. (2020).
Bioethanol production from waste hamburger by enzymatic hydrolysis and fermentation.
Journal of Cleaner Production, 264, 121658.
156. Ktori, R., Kamaterou, P., & Zabaniotou, A. (2018). Spent coffee grounds valorization through
pyrolysis for energy and materials production in the concept of circular economy. Materials
Today: Proceedings, 5(14), 27582–27588.
157. Stoeberl, M., Werkmeister, R., Faulstich, M., & Russ, W. (2011). Biobutanol from food
wastes—Fermentative production, use as biofuel an the influence on the emissions. Procedia
Food Science, 1, 1867–1874.
158. Huang, H., Singh, V., & Qureshi, N. (2015). Butanol production from food waste: A
novel process for producing sustainable energy and reducing environmental pollution.
Biotechnology for Biofuels, 8, 1–12.
159. Castanha, R. F., Mariano, A. P., de Morais, L. A. S., Scramin, S., & Monteiro, R. T. R.
(2014). Optimization of lipids production by Cryptococcus laurentii 11 using cheese whey
with molasses. Brazilian Journal of Microbiology, 45, 379–387.
160. Schneider, T., Graeff-Hönninger, S., French, W. T., Hernandez, R., Claupein, W., Holmes, W.
E., & Merkt, N. (2012). Screening of industrial wastewaters as feedstock for the microbial
production of oils for biodiesel production and high-quality pigments. Journal of Combustion.
161. Amulya, K., Jukuri, S., & Mohan, S. V. (2015). Sustainable multistage process for enhanced
productivity of bioplastics from waste remediation through aerobic dynamic feeding strategy:
Process integration for up-scaling. Bioresource Technology, 188, 231–239.
162. Slorach, P. C., Jeswani, H. K., Cuéllar-Franca, R., & Azapagic, A. (2019). Environmental and
economic implications of recovering resources from food waste in a circular economy. Science
of the Total Environment, 693, 133516. https://doi.org/10.1016/j.scitotenv.2019.07.322
163. Rajendran, N., Kang, D., Han, J., & Gurunathan, B. (2022). Process optimization, economic
and environmental analysis of biodiesel production from food waste using a citrus fruit peel
biochar catalyst. Journal of Cleaner Production, 365, 132712.
164. Mitra, D., van Leeuwen, J. H., & Lamsal, B. (2012). Heterotrophic/mixotrophic cultivation
of oleaginous Chlorella vulgaris on industrial co-products. Algal Research, 1(1), 40–48.

You might also like