You are on page 1of 244

Biofuel and Biorefinery Technologies 8

Meisam Tabatabaei
Mortaza Aghbashlo   Editors

Biodiesel
From Production to Combustion
Biofuel and Biorefinery Technologies

Volume 8

Series editors
Vijai Kumar Gupta, Department of Chemistry and Biotechnology, Tallinn
University of Technology, Tallinn, Estonia
Maria G. Tuohy, School of Natural Sciences, National University of Ireland
Galway, Galway, Ireland
This book series provides detailed information on recent developments in biofuels &
bioenergy and related research. The individual volumes highlight all relevant
biofuel production technologies and integrated biorefinery methods, describing the
merits and shortcomings of each, including cost-efficiency. All volumes are written
and edited by international experts, academics and researchers in the respective
research areas.
Biofuel and Biorefinery Technologies will appeal to researchers and post-
graduates in the fields of biofuels & bioenergy technology and applications,
offering not only an overview of these specific fields of research, but also a wealth
of detailed information.

More information about this series at http://www.springer.com/series/11833


Meisam Tabatabaei Mortaza Aghbashlo

Editors

Biodiesel
From Production to Combustion

123
Editors
Meisam Tabatabaei Mortaza Aghbashlo
Biofuel Research Team (BRTeam) Department of Mechanical Engineering of
Karaj, Iran Agricultural Machinery, Faculty of
Agricultural Engineering and Technology,
and College of Agriculture and Natural
Resources
Microbial Biotechnology Department University of Tehran
Agricultural Biotechnology Research Karaj, Iran
Institute of Iran (ABRII), Agricultural
Research, Education and Extension
Organization (AREEO)
Karaj, Iran

ISSN 2363-7609 ISSN 2363-7617 (electronic)


Biofuel and Biorefinery Technologies
ISBN 978-3-030-00984-7 ISBN 978-3-030-00985-4 (eBook)
https://doi.org/10.1007/978-3-030-00985-4

Library of Congress Control Number: 2018955161

© Springer Nature Switzerland AG 2019


This work is subject to copyright. All rights are reserved 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, express 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 Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface

This book is about biodiesel production presenting in-depth information on the state
of the art of global biodiesel production and investigates its impact on climate
change. Biodiesel is arguably the most commercialized type of petrodiesel alter-
native. A number of parameters including increasing energy demands and wors-
ening environmental conditions on one hand and similar physicochemical
properties of biodiesel to those of petrodiesel, on the other hand, are among the
main driving factors of the growing interest in biodiesel.
The present book, which is the eighth book in the series on Biofuel and
Biorefinery Technologies, offers a comprehensive reference guide to biodiesel
production by internationally recognized experts in the field of biodiesel production
from both academia and industry. The 10 chapters cover various aspects of bio-
diesel production technology from the basics, i.e., major principles of operation,
process control, and troubleshooting to production systems (reactor technologies) as
well as biodiesel purification and upgrading technologies. In addition, conventional
and emerging applications of biodiesel by-products with a view to further econo-
mize biodiesel production, economic risk analysis, and critical comparison of
biodiesel production systems as well as techno-economical aspects of biodiesel
plants are also comprehensively reviewed and discussed. Providing in-depth and
cutting-edge information on central developments in the field, “Biodiesel: From
Production to Combustion” also thoroughly investigates the important aspects of
biodiesel production and combustion by taking advantage of advanced sustain-
ability analysis tools including life cycle assessment (LCA) and exergy approaches.
In closing, the application of Omics technologies in biodiesel production is pre-
sented and discussed. The book is intended for all researchers, practitioners, and
students who are interested in the current trends and future prospects of biodiesel
production technologies.
It is expected that the present volume on biodiesel would assist both the sci-
entific and industrial communities in further developing this industry worldwide.
We are thankful to the authors of all the chapters for their efficient cooperation and
also for their readiness in revising the manuscripts. We also would like to extend
our appreciation to the reviewers who in spite of their busy schedule assisted us by

v
vi Preface

evaluating the manuscripts and provided their critical comments to improve the
manuscripts. We would like to sincerely thank Dr. Vijai Kumar Gupta and Dr.
Maria G. Tuohy and the team of Springer Nature, in particular, Dr. Andrea
Schlitzberger, Mr. Arumugam Deivasigamani, and Mr. Viju Falgon Jayabalan for
their cooperation and efforts in producing this book.

Karaj, Iran Meisam Tabatabaei


October 2018 Mortaza Aghbashlo
Contents

1 Global Biodiesel Production: The State of the Art


and Impact on Climate Change . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
Mahbod Rouhany and Hugh Montgomery
2 Biodiesel Production Systems: Reactor Technologies . . . . . . . . . . . 15
Thomas Ernst Müller
3 Biodiesel Production Systems: Operation, Process Control
and Troubleshooting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
Nídia S. Caetano, Vera Ribeiro, Leonardo Ribeiro, Andresa Baptista
and Joaquim Monteiro
4 Biodiesel Purification and Upgrading Technologies . . . . . . . . . . . . . 57
Hamed Bateni, Alireza Saraeian, Chad Able and Keikhosro Karimi
5 Applications of Biodiesel By-products . . . . . . . . . . . . . . . . . . . . . . . 101
Hajar Rastegari, Hossein Jazini, Hassan S. Ghaziaskar
and Mohammad Yalpani
6 Economic Risk Analysis and Critical Comparison
of Biodiesel Production Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
Seyed Soheil Mansouri, Carina L. Gargalo, Isuru A. Udugama,
Pedram Ramin, Mauricio Sales-Cruz, Gürkan Sin
and Krist V. Gernaey
7 Techno-economical Aspects of Biodiesel Plants . . . . . . . . . . . . . . . . 149
Syed Taqvi, Mohamed Elsholkami and Ali Elkamel
8 Biodiesel Production and Consumption: Life Cycle
Assessment (LCA) Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
Mohammad Ali Rajaeifar, Meisam Tabatabaei, Mortaza Aghbashlo,
Saeed Sadeghzadeh Hemayati and Reinout Heijungs

vii
viii Contents

9 Exergy-Based Sustainability Analysis of Biodiesel Production


and Combustion Processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193
Mortaza Aghbashlo, Meisam Tabatabaei, Mohammad Ali Rajaeifar
and Marc A. Rosen
10 “Omics Technologies” and Biodiesel Production . . . . . . . . . . . . . . . 219
Reza Sharafi and Gholamreza Salehi Jouzani
Chapter 1
Global Biodiesel Production: The State
of the Art and Impact on Climate
Change

Mahbod Rouhany and Hugh Montgomery

Abstract Biodiesel is a diesel-equivalent alternative fuel derived from biological


sources such as edible and nonedible oils, animal fats, and waste cooking oils
through processing. In addition to being a transportation fuel, biodiesel is also used
in some jurisdictions for electricity generation in engines and turbines. The world’s
biodiesel supply grew from 3.9 billion liters in 2005 to 18.1 billion liters in 2010
and is expected to exceed 33 billion liters in 2016 and reach 41.4 billion liters in
2025, a 25% increase over 2016 levels. Biodiesel prices have been facing down-
ward pressure due to low global petro-diesel prices, however, blending mandates
have largely sheltered the biodiesel market by lending consistency to demand.
International prices of biodiesel are expected to increase in nominal terms over the
next 10 years driven by the recovery of crude oil markets and prices of biofuel
feedstock. It should be mentioned that the majority of countries producing biodiesel
feedstock also have a vibrant domestic market and most or all of their supply is
used to meet domestic mandate-driven demand. This dual role, as both producer
and consumer, partially explains the limited international trade in biodiesel feed-
stocks. Most of the limited biodiesel trade over the next 10 years is expected to be
composed of Argentina’s exports to the US. While there is a debate on the sus-
tainability of biodiesel, many studies using lifecycle assessment (LCA) have
demonstrated that biodiesel results in 20–80% less greenhouse emissions when
compared to petro-diesel. As crude oil becomes more energy intensive to extract
and refine, expected efficiency gains in biodiesel feedstock production and refining,
the commercialization of second-generation biodiesel using nonfood feedstocks,
combined with the growing market share of biodiesel will result in further reduction
of harmful climate-impacting emissions by replacing petro-diesel with biodiesel.

M. Rouhany (&)
Strategic Carbon Management Inc., Vancouver, BC V5Z 1Z1, Canada
e-mail: mrouhany@gmail.com
H. Montgomery
Division of Medicine, Centre for Human Health and Performance, University College
London, London, UK

© Springer Nature Switzerland AG 2019 1


M. Tabatabaei and M. Aghbashlo (eds.), Biodiesel,
Biofuel and Biorefinery Technologies 8,
https://doi.org/10.1007/978-3-030-00985-4_1
2 M. Rouhany and H. Montgomery

1.1 Introduction

Biodiesel, derived from processing biological sources such as edible and nonedible
oils, animal fats, and waste cooking oils, has similar properties to petro-diesel. It
can be used to enhance certain characteristics of petro-diesel, such as lubricity,
aiding fuel performance, and extending engine life (Traviss 2012; Pacini et al.
2014). Compared to petro-diesel, it has a higher cetane number (and thus better
ignition quality) but a lower heating value, higher density, and higher viscosity
(Taher and Al-Zuhair 2017) and is thus less suitable for colder climates due to
gelling, clouding, and overall reduced cold weather performance (Traviss 2012).
Biodiesel can be blended in all ratios and many jurisdictions use these, from farm
level to industrial scale, in preference to pure biodiesel. The quality of biodiesel is
determined by the quality of feedstock oil, the processing technology used, and the
process parameters (Knothe et al. 2010; Rathore et al. 2016). Biodiesel and ethanol
make up the majority of the renewable share of the world road and marine trans-
portation sector’s energy demand (REN21 2016). Biodiesel is also utilized in sta-
tionary machinery and in some jurisdictions for heat and electricity generation
(Rathore et al. 2016).
The net environmental benefit of biodiesel is a topic of continuing debate.
Biodiesel is biodegradable. Whether used pure or as a petro-diesel blend, it can
provide air quality benefits namely lower loads of carbon monoxide, sulfur oxides,
and volatile organic compounds (Pacini et al. 2014). In many cases, net greenhouse
gas emissions are reduced. The majority of criticism targets the negative impacts
that biodiesel sourced from agriculture-based biomass feedstock farming have on
forests and grasslands, food and animal feed prices, loss of biodiversity due to
mono-cropped fields, water resource management, food security, and air quality. To
date, edible oilseeds such as soybean and rapeseed have been the dominant bio-
diesel feedstocks. Biofuels developed from food or animal feed crops are referred to
as “first generation” or conventional biofuels. Developing biodiesel from crops that
can be grown on land that is not suitable for growing food, from biomass sources
that are less dependant on the availability of land, or from nonedible feedstocks or
by-products, can alleviate many of the sustainability concerns. Biofuels that are
developed from nonedible biomass except algae are known as “second generation”
or advanced biofuels. Biodiesel produced from microalgae would be considered a
“third generation” biofuel. Algae, municipal and industrial organic waste, sugar
cane bagasse, corn stover, perennial grasses, cereal straw, as well as forestry and
agricultural waste are examples of more sustainable feedstock. These sources, while
not yet produced at commercial scale, are receiving considerable attention due to
their smaller environmental footprint (Rathore et al. 2016; Anuar and Abdullah
2016; Royal Academy of Engineering 2017).
Important challenges for the biodiesel industry come from low petro-diesel
prices, fuel–food competition resulting in reciprocal price increases and destabi-
lization of the feedstock market, as well as negative socio-environmental impacts of
the feedstock oilseeds (Anuar and Abdullah 2016). The implementation of biofuel
1 Global Biodiesel Production: The State of the Art … 3

supporting policies and legislation, selection of low-cost sustainable nonedible


feedstock, and production process improvements for better quality and cheaper
production costs could eventually lead to a worldwide replacement of petro-diesel
with biodiesel.

1.2 Brief History

The first known transesterification of a vegetable oil was conducted by E. Duffy and
J. Patrick in 1853. This was four decades before Rudolf Diesel’s engine first ran
independently in Augsburg Germany on August 10, 1893 (Abdalla and Oshaik
2013). The diesel engine, since its inception, could run on a variety of fuels
including vegetable oils. One of the first publicly demonstrated uses of biodiesel in
a diesel engine was in the year 1900 when, during the Paris exposition, the French
company Otto operated a small diesel engine on peanut oil. According to Rudolf
Diesel’s papers, published in 1912 and 1913, in addition to research by the French
on peanut oil, experiments were being conducted in St. Petersburg using castor oil
and train-oil (oil obtained from the blubber of marine animals) with excellent
results. France, Belgium, Germany, Italy, and the UK had varying interests in fuels
from vegetable oils during the first half of the twentieth century (Knothe et al.
2010). Triglycerides from easily available oil-rich feedstocks were contenders for
being the main fuel source for the diesel engine in its early years. However, natural
oils are viscose with relatively low cetane numbers compared to petro-diesel, which
resulted in them gradually being replaced by petroleum oil (Taher and Al-Zuhair
2017).
The petroleum industry has commonly dominated the global fuel market with its
cheaper production and price. Generally, when petroleum fuel supplies are plentiful
and inexpensive, interest in bio-sourced oils has been low. Disruption of petroleum
fuel supplies during World War II drove countries like Argentina, Brazil, India, and
China to use vegetable oil as fuel (Van Gerpen et al. 2007).
The petroleum oil embargo of the 1970s led to a renewed interest by the United
States, Austria, and South Africa in vegetable oils and their direct use in diesel
engines as fuel. Since the 1920s, diesel engine manufacturers had altered their
designs to match the lower viscosity of petroleum diesels (Van Gerpen et al. 2007;
Abdalla and Oshaik 2013). Thermal cracking, pyrolysis, transesterification, the
formation of microemulsions, and dilution of oils with solvent were, thus, experi-
mented with to address the viscosity limitations of vegetable oils. With the emer-
gence of suitable catalysts, the transesterification with short-chain alcohols, such as
methanol and ethanol, became the preferred and most commonly used method to
convert bio-oils to biodiesel (Taher and Al-Zuhair 2017). The term biodiesel was
most likely first used around 1984. The commercial production of biodiesel started
in the early 1990s and the first standard for biodiesel was published in 2001, the
ASTM D6751.
4 M. Rouhany and H. Montgomery

1.3 State and Future of Biodiesel Demand and Supply

The world’s market share of diesel in transportation fuels has been increasing in
comparison to gasoline and this share is expected to continue to grow globally at
varied rates mainly driven by non-OECD countries. Biodiesel production growth
has been following this trend and is increasing faster than that of ethanol.
International trade in biodiesel has also been considerably higher than the trade in
ethanol and, despite its small share compared to production, the international
biodiesel trade has been paramount in the development of the biodiesel industry in
developing economies. Pro-biodiesel policies in the EU and USA have driven the
development and expansion of biodiesel industries for export in agricultural
countries with established oilseed industries, namely palm-based biodiesel in
Indonesia and Argentina’s soy-based biodiesel (Naylor and Higgins 2017). Global
fuel demand in conjunction with domestic policies and trade interactions are the
main drivers for the global biodiesel sector.
Between 2005 and 2015, global biodiesel production expanded by more than
20% per year, which resulted in a sevenfold expansion in a single decade. This
occurred parallel to a rise in petro-diesel prices during the same period. Diesel and
oil prices have been in decline since mid-2014 and lower petroleum prices stimulate
petro-diesel use. However, despite the downward pressure from recent low oil
prices and policy uncertainty in some markets, biofuel production and demand
continued to increase in 2016, and ethanol and biodiesel still comprised the
majority of the renewable share of global energy demand for transportation with
roughly 4% of the world road transport fuel (REN21 2017; Naylor and Higgins
2017).
According to the Organization for Economic Cooperation and Development and
the UN’s Food and Agriculture Organization 2016 Agricultural Outlook, global
biodiesel use is expected to gradually increase over the next 10 years. The largest
demand increase will be from developing countries, mainly Indonesia, Brazil, and
Argentina, with an estimated 68% increase in 2025 compared to 2015 (OECD/FAO
2016).
The European Union and the United States are, together, the largest influencers
of biofuel demand. Implementation of biofuel mandates has led to an increase in
biofuel use in the United States. The current maize-based ethanol mandate is
expected to decline after 2018 and be replaced by an increase in the advanced
mandate covering biofuels from sources other than maize. This would result in
lower ethanol use and an increase in biodiesel use in the United States. In the
European Union, the Renewable Energy Directive target has to be met by 2020
which is expected to sustain an expansion of ethanol and biodiesel fuel use until
then. Thereafter, a decrease is expected in line with lower gasoline and diesel use
prospects. Palm oil is expected to decline as a feedstock in European biodiesel.
In developing countries, biodiesel use is also expected to expand steadily with
Indonesia, Brazil, and Argentina leading the way due to their domestic mandates.
Biofuel demand is expected to remain low in Central Asia and Eastern Europe as
1 Global Biodiesel Production: The State of the Art … 5

these regions are either oil and gas producers or lack biofuel incentive policies for
producers or blending mandates for consumers (OECD-FAO 2015). Global bio-
diesel supply grew from 3.9 billion liters in 2005 to 30.8 billion liters in 2016 and is
expected to reach 41.4 billion liters in 2025, a 34% increase over 2016 levels
(Onguglo et al. 2016; REN21 2016; OECD/FAO 2016). An estimated 72% of
biofuel production (in energy terms) was fuel ethanol, 23% was biodiesel, and 4%
was hydrotreated vegetable oil (REN21 2017). More than 80% of the world’s
biodiesel production is from vegetable oils, with the majority produced from
European canola and soybeans from the United States, Brazil, and Argentina.
Indonesian palm oil and other sources such as jatropha and coconut make up a
small share of vegetable-based biodiesel. Waste-based biodiesel accounted for 8%
of the global supply in 2015 (OECD/FAO 2016; REN21 2016). In 2015, biodiesel
was responsible for 162,600 direct and indirect jobs in Brazil while in the same year
the U.S. biodiesel sector provided 49,486 direct and indirect jobs (REN21 2016).
Whilst spread across many countries, biodiesel production is dominated by only
a few In 2016, the EU was the largest producer (with a 26% share of global
production), and 76% of the world’s fatty acid methyl ester (FAME) biodiesel was
produced by the EU, United States, Brazil, Argentina, and Indonesia. No other
country outside of this group had a share larger than 5% (REN21 2016, 2017)
(Fig. 1.1).
The domestic policy incentives in the United States, Argentina, Brazil, and
Indonesia and, to a lesser extent, the fulfillment of the Renewable Energy Directive
(RED) target in the European Union, are the main drivers for global biodiesel
production (OECD/FAO 2016).
The EU is experiencing a decline in investment in new biodiesel capacity mainly
due to a continuing decrease in policy and public support for first-generation

Other countries, United States,


19.8% 17.9%

Brazil, 12.3%

EU-28, 26.0%
Argentina, 9.7%

Thailand, 4.5%
Indonesia, 9.7%

Fig. 1.1 Major biodiesel-producing countries in 2016 (REN21 2017)


6 M. Rouhany and H. Montgomery

biofuels, including biodiesel, as a result of environmental concerns and an


increasing interest in electric mobility. Among individual countries, the US remains
the world leader in biodiesel production supported by its agricultural policy and by
the federal renewable fuel standard. Brazil is solidifying its place as the second
largest producer of biodiesel with 13% of the global share in 2016 (REN21 2017).
Figure 1.2 demonstrates the trends in US biodiesel production, consumption,
imports, and exports from 2001 to 2015 (U.S. EIA 2017). The peak in 2008 was
largely due to a biodiesel tax credit in the European Union, which drove up US
exports and production. Exports dropped after the tax credit was phased out. The
increase in production and consumption from 2010 onward was largely to meet the
requirements of the second phase of the Renewable Fuel Standard (RFS). The RFS
is a federal mandate that requires a minimum volume of renewable fuels to be
blended in the transportation fuel sold in the United States. Its second phase
required the use of 34 billion liters of renewables in 2008 increasing to 136 billion
liters in 2022 with a cap on the share of corn-starch ethanol and a minimum
requirement for the share of cellulosic biofuels.
In 2013, the consumption of biodiesel in the US surpassed its production, and
the volume of biodiesel imported by the US exceeded exports and has continued to
increase. The growth in consumption and imports since then is likely due to the
favorable regulatory framework and increased efforts to reduce greenhouse gas
emissions.
With its substantial biodiesel production capacity, Argentina has been a leading
supplier of imported biodiesel for the EU, the United States, and other countries
since 2010. In 2013, the EU imposed a heavy anti-dumping import tax on
Argentinian biodiesel which resulted in Argentina’s biodiesel manufacturing
capacity being underutilized despite growing domestic demand (REN21 2017).
PRODUCTION AND CONSUMPTION = MILLION GALLONS

6000

U.S. Biodiesel Production, Exports,


IMPORT AND EXPORTS = THOUSAND BARRELS

5000
Imports, and Consumption
4000

3000

2000

1000

0
2001 2002 2003 2004 2005 2006 2007 2008 2009 2010 2011 2012 2013 2014 2015
Production Exports Consumption Imports

Fig. 1.2 U.S. biodiesel 10-year production, consumption, imports, and exports from 2001 to 2015
(U.S. EIA 2017)
1 Global Biodiesel Production: The State of the Art … 7

Most biodiesel-feedstock-producing countries also have an active domestic


market and most or all of their supply is used to meet domestic mandate-driven
demand. This dual role, as both producer and consumer, partially explains the
limited international trade in biodiesel feedstocks. The European Union mostly
imports vegetable oil-based biodiesel from countries such as Argentina, Indonesia,
and Malaysia (Onguglo et al. 2016). Most of the limited biodiesel trade is com-
posed of Argentina’s exports to the US (Pacini et al. 2014). Figure 1.3 provides a
10-year overview of global biodiesel production, consumption, and exports from
2007 to 2016.
Biodiesel price is influenced by the type of feedstock, production volume,
production process, government incentives, food prices, and research and devel-
opment costs. As edible oils comprise more than 80% of the world’s biodiesel
feedstock, biodiesel prices closely follow vegetable oil prices. Policies which
support prices of vegetable oil also influence the demand for biodiesel (OECD/FAO
2016).
Biodiesel prices have been facing downward pressure due to low global
petro-diesel prices; however, blending mandates have largely sheltered the biodiesel
market by lending consistency to demand (REN21 2016). Figure 1.4 provides an
overview of the average U.S. Diesel and B99/B100 Biodiesel price over the last
10 years. International prices of biodiesel are expected to increase in nominal terms
over the next 10 years driven by the recovery of crude oil markets and prices of
biofuel feedstock (OECD/FAO 2016).

35,000.00
MILLIONS OF LITRES BIODIESEL

30,000.00
25,000.00
20,000.00
15,000.00
10,000.00
5,000.00
0.00
2007 2008 2009 2010 2011 2012 2013 2014 2015 2016

Exports consumpƟon ProducƟon

Fig. 1.3 Global biodiesel production, consumption, and exports 10-year overview (OECD/FAO
2016)
8 M. Rouhany and H. Montgomery

USD COST PER GASOLINE GALLON EQUIVALENT (GGE) $5.00

$4.50

$4.00

$3.50

$3.00

$2.50

$2.00

$1.50

$1.00
B99/B100 Diesel
$0.50

$0.00

Fig. 1.4 Average U.S. diesel and B99/B100 biodiesel 10-year price overview (USDoE 2017)

1.4 The Biodiesel Policy Landscape

The impressive growth of the global biodiesel market and industry during the last
decade at rates exceeding 20% per year despite downward pressure from low fossil
fuel prices is primarily driven by policies enhancing production and demand at the
national and regional level. Blending mandates, tax exemptions, subsidies, fuel
quality standards, import tariffs, and investment backings are examples of such
supportive regulations. Such policies are, in turn, driven and influenced by a
combination of factors, such as a desire for increased energy security, environ-
mental concerns and climate-related targets, lobby groups, feedstock availability,
effective use of co-products, enhancing rural development, and increasing the
demand and price for vegetable oils (REN21 2016; Cadham 2015; Naylor and
Higgins 2017).
Examples where biodiesel production has been profitable in the absence of
additional financial incentives are very few. Studies show that this has only been
achieved with palm oil as the feedstock and during times when feedstock prices
were low and oil prices were high. To improve the overall financial and opportunity
costs, governments often accompany quantitative targets with other policies such as
blending mandates, subsidies, and tax credits (Naylor and Higgins 2017).
The policy instrument that is most commonly used across various countries and
regions is the blend mandate. A blend mandate specifies a share or volume of
biodiesel to be blended with petro-diesel. Blending mandates lead to consistency in
demand which is instrumental in protecting biodiesel markets from the effects of
1 Global Biodiesel Production: The State of the Art … 9

low global petro-diesel prices. By the end of 2015, biodiesel blend mandates were
in place at the national level in 18 countries (REN21 2016).
The majority of mandates are in place in the EU, where its Renewable Energy
Directive requires a 10% renewable content in fuel by 2020. RED establishes
sustainability requirements for liquid biofuels, including greenhouse gas
(GHG) reductions, land use changes, and other environmental, social and economic
criteria. A 7% limit on the share of food-crop-based transportation biofuels to the
EU’s 10% renewable mandate and the exclusion of biofuels grown on land with
peat or high carbon stocks was introduced in the amendments to RED. Adoption of
second-generation biofuels is further incentivised through a 0.5% voluntary target
and by allowing the contribution of nonfood crop-based biofuels to be
double-counted toward meeting the overall EU target (Naylor and Higgins 2017;
Araújo et al. 2017).
Agricultural support and the expansion of renewable fuels and climate mitigation
have been the main motivations driving biodiesel policies in the EU. Recently, a
stronger focus on sustainability and reducing GHG emissions has resulted in
changes in EU policies regarding feedstock sourcing. The current regulations
require that all biofuels from existing plants must result in a 50% reduction in
lifecycle greenhouse gas emissions in comparison to fossil fuels, beginning in 2018.
New plants should demonstrate a 60% reduction in GHG emissions in their biofuel
product considering emissions from cultivation, processing, and transport. A 2015
amendment to RED requires that calculations of indirect land use change (iLUC)
emissions associated with biodiesel feedstock be incorporated in GHG emission
calculations by fuel suppliers. iLUC emissions do not officially count in the GHG
reduction targets. As the majority of current biodiesel feedstocks will not meet the
50% reduction in GHG emissions target, EU member states are increasingly con-
sidering alternative feedstocks such as waste oils which provide significant GHG
emission reductions compared to fossil fuels and do not have land-use change
impacts. It is expected that the legislation that will replace the RED after it expires
in 2020 will have more stringent sustainability criteria, namely further limits on
GHG emissions, on the use of food crop feedstocks, and on land-use change
impacts. A reduction in the food crop share from the current 7 to 3.8% in 2030 and
raising the minimum greenhouse gas savings over fossil fuel alternatives to 70% by
2021 was proposed in the European commission in late 2016 (Naylor and Higgins
2017).
In the EU, petro-diesel is the primary fuel used for road transportation which
accounted for roughly 75% of the energy used in transportation in the EU in 2016.
The share of diesel fuel in the EU’s road transport grew from 52% in 2000 to 70%
in 2014. Historically, the European biodiesel industry was developed in order to
provide a substitute for petro-diesel. The EU introduced the Renewable Energy
Directive in 2009, which required 10% of all transportation energy to come from
renewable resources by 2020. RED allows member states flexibility in selecting
their own policies for meeting the target. Between 2005 and 2015, the EU’s bio-
diesel production tripled, and its production capacity expanded more than fivefold.
10 M. Rouhany and H. Montgomery

In 2016, 80% of the EU biofuels market was composed by biodiesel and ethanol
held the remaining 20% (EEA 2016; Naylor and Higgins 2017).
The EU has implemented a 3.5% import duty on biodiesel blends of B30 (30%
biodiesel content) and under, and a 6.5% import duty on B30–B100 (pure biodiesel
with no blending) fuels to protect its domestic rapeseed and biodiesel production.
Other EU trade policies include anti-dumping tariffs on biodiesel imports from the
USA, Canada, Argentina, and Indonesia. In September 2016, the EU terminated its
anti-dumping duties against Argentina and Indonesia (Naylor and Higgins 2017).
In the United States under the Renewable Fuel Standard, the Environmental
Protection Agency releases annual biomass-based diesel volume requirements. By
the end of 2015, biodiesel blend mandates were in place in 27 jurisdictions (REN21
2016). For 2017, the volume requirement for biomass-based diesel was 7.6 billion
liters (2.0 billion gallons). The RFS places a cap on the share of corn-starch ethanol
and a minimum requirement for the share of cellulosic biofuels. A $1-per-gallon
biodiesel blending tax credit was implemented in 2005, which expired at the end of
2016. Furthermore, the American Renewable Fuel and Job Creation Act of 2017
was introduced in the US Senate on April 26, 2017 to replace the Biodiesel
blending credit. The bill modifies and extends the income tax credit for biodiesel
and renewable diesel used as fuel, and the excise tax credit for biodiesel fuel
mixtures. The Act proposes a $1-per-gallon production credit for biodiesel pro-
duced in the United States from December 2016 until December 2020 and an
additional 10 cent-per-gallon credit for small US biodiesel producers (under 15
million gallons/year). The small producer credit would be available to biodiesel
produced from all feedstocks (Library of Congress 2017).
The political context within each nation forms its policy priorities, goals,
instruments, and methods. While national biodiesel policy implementation in major
producing countries seems to address a wide range of interests across several
objectives, in reality, the support of specific sectors and interests, such as farm
lobbies and energy groups, often determines policy design and implementation.
Large agricultural economies often install policies that indirectly support local
agriculture by enhancing the use of domestic oil crops for biodiesel feedstock to
support farm revenues throughout their agricultural supply chain. Consequently, all
large biodiesel producing nations are using their domestic agricultural products as
the main feedstock for biodiesel production, resulting in a complex interaction of
energy and agricultural interests. These interests provide the drive for governments
to maintain and even enhance their support for the biodiesel sector during the
current era of low crude oil prices (Naylor and Higgins 2017). In addition, there are
national and international interests in reducing fossil fuel use so as to reduce GHG
emissions and meet climate targets. Fossil fuel lobbies and political forces working
to expand fossil fuel use, as is currently seen in play in the US, are opposing and
complicating factors. This creates a state of affairs in which uncertainties exist that
could significantly change the projections for biofuel markets over the next decade.
US and EU policies on climate mitigation, feedstock sourcing, blending mandates,
and trade barriers together with fuel prices and the biodiesel sector’s ability to
1 Global Biodiesel Production: The State of the Art … 11

commercialize nonfood-based biodiesel will be the main factors determining the


future of biodiesel (OECD/FAO 2016; Naylor and Higgins 2017).

1.5 Biodiesel, the Environment, and Climate Change

At the United Nations Framework Convention on Climate Change’s (UNFCCC)


22nd Conference of the Parties (COP22) in Marrakesh, Morocco in late 2016, more
than 100 countries had officially agreed to limit global warming to below 2 °C
under the Paris agreement. Additionally, leaders of 48 developing countries com-
mitted jointly to work toward achieving 100% renewable energy in their respective
nations under the Climate Vulnerable Forum (CVF) (REN21 2016, 2017).
Given the real and imminent threat of climate change, it is now an issue high on
the agenda of governments and citizens around the globe. To meet the challenge of
increasing energy access and reducing poverty while reducing GHG emissions
enough to meet the COP22 target of limiting global temperature increase, extraction
of remaining fossil fuel reserves will have to stop altogether and the use of
renewable energy and energy efficiency instruments will have to be significantly
increased.
There are many drivers and advantages for the use of biofuels but due to the
increased global focus on biofuels’ environmental threats and social impacts, the
sustainability of biodiesel is more carefully considered and assessed today than was
the case when biofuels first became commercially available. There are a large
number of studies assessing the sustainability of biodiesel that come to a wide range
of conclusions, which is fueling the debate on biodiesel’s sustainability. The
diversity and sometimes conflict in results arise from differences in methodologies,
feedstock sources, land use and land use change impacts, selection of system
boundaries, and functional units, as well as allocation methods.
The controversy over the environmental and social impacts of first-generation
biodiesels commonly centers around the food versus fuel debate and the negative
climate impacts of land-use change (REN21 2017). There is little agreement on the
magnitude of the impact of biodiesel on food security. Using edible oils as biodiesel
feedstock could act as a buffer on the impact of food crop production variations in
different years (Naylor and Higgins 2017).
Greenhouse gas emissions from biodiesel are commonly assessed using a life-
cycle assessment (LCA). Such assessments calculate the amount of greenhouse
gases that are emitted per unit of fuel over its lifecycle from production to use. For
biodiesel, this includes emissions and/or carbon sequestration, in addition to
land-use changes from the growing of feedstock and allocation of by-products,
when applicable (Pacini et al. 2014).
The potential impact of biodiesel feedstock sources on indirect land use changes
(iLUC), such as deforestation, is a cause of concern for the sustainability and more
specifically the GHG emission savings of biodiesel. This could even, in some cases,
result in biodiesel generating more lifecycle GHG emissions than petro-diesel.
12 M. Rouhany and H. Montgomery

In one of the most extensive studies to date, the UK’s Royal Academy of
Engineering conducted an assessment of over 250 separate studies on the GHG
emission reductions of biofuels versus fossil fuels (Naylor and Higgins 2017).
In the UK study, the GHG emissions per unit of energy generated for
first-generation biodiesels produced from common feedstocks displayed a large
variation ranging from 4 to 505 grams of CO2 equivalent-per-Mega Joule (gCO2e/
MJ) across different LCA studies. As a point of comparison, it should be noted that
the carbon intensity of EU petro-diesel is around 84 gCO2e/MJ. However, the
average biodiesel GHG emissions from all the feedstocks considered were lower
than emissions from fossil diesel if no land use change (LUC) was involved. The
only type of first-generation biodiesel that would meet the EU RED requirement for
50% less GHG emissions compared to conventional diesel was palm oil biodiesel
without LUC (Naylor and Higgins 2017).
Where land-use change-related carbon emissions are included in the calcula-
tions, all varieties of first-generation biodiesels considered in the study had a higher
average carbon footprint than petro-diesel. Soybeans had the largest negative GHG
emission impact, which could be due to soybean cultivation in South and Central
America actuating both direct and indirect land use change (iLUC). Biodiesel
produced from palm oil harvested from peat and forest lands in Indonesia and
Malaysia demonstrated 3–40 times higher GHG emissions per unit of energy
compared to petro-diesel. A large variability was observed in results of the assessed
studies including LUC-related GHG emissions. This is due to the differences in
LUC GHG estimation methods and emission factors and the fact that some studies
included either direct or indirect LUC-related emissions and others included both
(Naylor and Higgins 2017).
The average GHG emissions per unit of energy for second-generation biodiesels
from nonedible feedstocks are considerably lower than petro-diesel, with the values
ranging from −88 to 80 gCO2e/MJ. Negative values are a result of credits for
co-products. The three feedstocks evaluated were Jatropha, Camelina and used
cooking oil/tallow. The average carbon intensity of Jatropha, used cooling oil/
tallow, and Camelina are, respectively, 26, 27, and 33 gCO2e/MJ. Similar to
first-generation biodiesels, the range of these results varied broadly due to regional
differences in yield and different estimation methods particularly in regard to
co-product allocation. In most of the studies assessed by the Royal Academy of
Engineering biodiesel from tallow and used cooking oil showed 60–90% lower
carbon intensity than petro-diesel. The average GHG intensity value for
third-generation microalgae biodiesel was 3.5 times higher than conventional diesel
also with a large variation in the individual results. Due to costly and
energy-intensive production, biodiesel produced from algae at its current phase of
development results in more GHG emissions than its petroleum counterpart and is
not yet a viable choice (Naylor and Higgins 2017).
Agriculture phase LUC is the major contributor to biodiesel GHG emissions
followed by the transesterification process. The EU is intent on a continuous
reduction in the share of first-generation biofuels in transport fuel and increasing the
share of climate-friendly advanced biofuels (REN21 2017). As crude oil becomes
1 Global Biodiesel Production: The State of the Art … 13

more energy intensive to extract and refine, the commercialization of


second-generation biodiesel using nonedible feedstocks, paired with efficiency
gains in production and refining techniques, will potentially result in further
reduction of harmful climate impacting emissions by replacing petro-diesel with
biodiesel (Pacini et al. 2014).
In terms of other pollutants, biodiesel can favorably reduce particulate matter by
nearly 88% relative to petro-diesel while in terms of NOx there are varied results,
with some claiming biodiesel emits greater amounts of nitrogen oxides than
petro-diesel. Using 100% biodiesel in heavy-duty highway engines produces on
average almost 70% less hydrocarbons, 50% less particulates and carbon monoxide,
and 10% more NOx emissions. Biodiesel has negligible sulfur oxide emissions and
half the ozone-forming potential of petro-diesel (Araújo et al. 2017).

References

Abdalla BK, Oshaik FOA (2013) Base-transesterification process for biodiesel fuel production
from spent frying oils. Agric Sci 4:85–88
Anuar MR, Abdullah AZ (2016) Challenges in biodiesel industry with regards to feedstock,
environmental, social and sustainability issues: a critical review. Renew Sustain Energy Rev
58:208–223
Araújo K, Mahajan D, Kerr R, Silva MD (2017) Global biofuels at the crossroads: an overview of
technical, policy, and investment complexities in the sustainability of biofuel development.
MDPI AG, Basel
Cadham WJ (2015) Biomass for bioenergy and/or transportation biofuels: exploration of key
drivers influencing biomass allocation, Master of Science edn. University of British Columbia,
Vancouver BC
EEA (2016-last update) Transport in Europe: key facts and trends, European Environment
Agency. https://www.eea.europa.eu/signals/signals-2016/articles/transport-in-europe-key-
facts-trends. Accessed August 2017
Knothe G, Krahl J, van Gerpen J (2010) The biodiesel handbook, 2nd edn. AOCS Press, Urbana
Illinois
Library of Congress CG (2017-last update) S.944—American renewable fuel and job creation act
of 2017. https://www.congress.gov/bill/115th-congress/senate-bill/944. Accessed 28 July 2017
Naylor RL, Higgins MM (2017) The political economy of biodiesel in an era of low oil prices.
Renew Sustain Energy Rev 77:695–705
OECD/FAO (2016) OECD-FAO agricultural outlook 2016–2025. Data for tables extracted on 04
Jul 2017 from OECD. Stat. OECD Publishing, Paris
OECD-FAO (2015) OECD-FAO agricultural outlook 2015–2024. OECD Publishing, Paris
Onguglo B, Pacini H, Kane M, Lleander l, Paredes I, Payosova T, Dent R, Czinar M, Grisoli R,
Bartocci Liboni L, Elias dos Santos M, Rodrigues Alves MF, Pacheco LM, Garcia de
Oliveira B, Horta LA, Rogan R, Kennedy H, Golden J, McDonald J, Poirrier A, Strapasson A,
Dupont J, Desplechin E, Maniatis K, Aguiar R, Thomaz LF, Hardjakusumah C, Ostheimer G,
Ramakrishna YB, Søgaard J, Coleman B, Ayuso M, Boshell F, Nakada S, Guardabassi P,
Neeft J (2016) Second-generation biofuel markets: state of play, trade and developing country
perspectives. UNCTAD/DITC/TED/2015/8. United Nations—UNCTAD
Pacini H, Sanches-Pereira A, Durleva M, Kane M, Bhutani A (2014) The state of the biofuels
market: regulatory, trade and development perspectives. Trade Environment, Climate Change
and Sustainable Development Branch, DITC, UNCTAD
14 M. Rouhany and H. Montgomery

Rathore D, Nizami A, Singh A, Pant D (2016) Key issues in estimating energy and greenhouse gas
savings of biofuels: challenges and perspectives. Biofuel Res J 3(2):380–393
REN21 (2017) Renewables 2017 global status report (Paris: REN21 Secretariat).
REN21-Renewable energy policy network for the 21st century
REN21 (2016) Renewables 2016 global status report (Paris: REN21 Secretariat).
REN21-renewable energy policy network for the 21st century
Royal Academy of Engineering (2017) Sustainability of liquid biofuels. Royal Academy of
Engineering, London UK
Taher H, Al-Zuhair S (2017) Emerging green technologies for biodiesel production. Front
Bioenergy Biofuels
Traviss N (2012) Breathing easier? The known impacts of biodiesel on air quality. Biofuels 3
(3):285–291
U.S. EIA (2017-last update) U.S. energy information administration monthly energy review. http://
www.eia.gov/totalenergy/data/monthly/#renewable. Accessed 28 July 2017
USDoE A (2017-last update) U.S. Department of Energy Alternative Fuels Data Center (AFDC),
Clean cities alternative fuel price reports. http://www.afdc.energy.gov/fuels/prices.html.
Accessed 07 May 2017
Van Gerpen JH, Peterson CL, Goering CE (2007) Biodiesel: an alternative fuel for compression
ignition engines. ASAE Distinguished Lecture No. 31 2007, pp 1–22
Chapter 2
Biodiesel Production Systems: Reactor
Technologies

Thomas Ernst Müller

Abstract The dwindling of fossil resources has prompted producers of fuels, fine
chemicals, and polymers to switch from fossil carbon sources and search for
renewable feedstock. Biomass holds one of the keys to this transition to a circular
economy. In this context, biodiesel obtained by transesterification of natural oils
with alcohols is gaining importance in the fuel sector. Various reactor concepts
have been developed for the transesterification reaction. Depending on the scale of
the biodiesel production plant, reactors with varying designs are operated in the
batch, semi-batch mode, or continuously. In this chapter, the optimal reactor
technologies are analyzed with respect to the stages the chemical conversion runs
through. The initial reaction mixture of natural oil and methanol, the most common
alcohol in biodiesel production, is characterized by a liquid–liquid two-phase
system. The high polarity difference of natural oil and methanol leads to a mixa-
bility gap and formation of a natural oil-rich phase and a methanol-rich phase. The
mass transfer of the reagents across the phase boundary is slow relative to the
chemical reaction, thereby resulting in diffusion limitations. Various mixing tech-
nologies, such as sonication, and the use of microreactors are explored to overcome
these diffusion limitations. Once the reaction is 15–20% complete, the reaction
mixture becomes homogeneous, reducing the need for intensive mixing. As the
reaction continues and higher conversions are obtained, the fatty acid methyl ester
separates from glycerin. The two phases are separated and purified. Recent tech-
nologies for process intensification aim at enhancing mass and heat transfer at all
stages of the reaction.

T. E. Müller (&)
Chemical and Process Engineering, Rheinische Fachhochschule Köln,
Schaevenstraße 1 a-b, Köln 50676, NRW, Germany
e-mail: thomas.mueller@rfh-koeln.de

© Springer Nature Switzerland AG 2019 15


M. Tabatabaei and M. Aghbashlo (eds.), Biodiesel,
Biofuel and Biorefinery Technologies 8,
https://doi.org/10.1007/978-3-030-00985-4_2
16 T. E. Müller

2.1 Introduction

The dwindling of fossil resources has spurred the need to switch the production of
fuels, chemicals, and polymers from fossil carbon sources to renewable feedstock.
Biomass holds one of the keys to the transition of a fossil-based unidirectional to a
circular economy. It is a renewable and widely available resource. In the fuel sector,
biodiesel has been on the market for some time and is gaining further in importance.
It is an environmentally benign, biodegradable, nontoxic fuel associated with
comparably low emissions (Chuah et al. 2017; Kumar et al. 2013). Currently, two
main types of biodiesel fuels are produced on a large scale—fatty acid methyl esters
and hydroprocessed esters and fatty acids (Dimian and Rothenberg 2016). The term
biodiesel refers primarily to fatty acid methyl esters. For biodiesel production,
natural oils, i.e., the glycerol ester of fatty acids are used as raw material. The
primary feedstock can be vegetable oil crops derived from rapeseed, sunflower, soy
or palm (Dimian and Rothenberg 2016). Moreover, residual lipid materials and
nonedible vegetable oils derived from jatropha, camelina, and ricinus are suitable
(Dimian and Rothenberg 2016; Bhuiya et al. 2016; Demirbas et al. 2016). Future
raw material (Perego and Ricci 2012) may include algae biomass with a certain oil
contents (Brennan and Owende 2010) or lipids obtained by the conversion of
carbohydrates with special yeasts (Arous et al. 2016). The fatty acid moieties
comprise alkyl chains differing in the number of carbon atoms (R′, R″, R‴ = C14H29
to C20H41) (Feasibility report small-scale biodiesel production 2006). Besides sat-
urated alkyl chains, unsaturated chains are also present to a smaller extent. For
biodiesel production, natural oil is reacted with alcohol to the corresponding fatty
acid ester and glycerol in an equilibrium reaction (Eq. 2.1). The fatty acid ester is
then purified to yield biodiesel within the legal specifications (International A
2013).
O
O
O R' OH
RO R'
O [Cat.]
3 R OH + O R''
RO
O
R''
+ OH ð2:1Þ
O
O
O R''' OH
RO R'''

Methanol (R = H) is used typically as the alcohol in biodiesel production. The high


polarity of methanol, however, results in a mixability gap with natural oil.
Consequently, the initial reaction mixture of natural oil and methanol is a liquid–liquid
two-phase system comprising a natural oil-rich phase and a methanol-rich phase.
Once the reaction is 15–20% complete, the reaction mixture becomes homogeneous,
reducing the need for intensive mixing. As the reaction continues and higher con-
versions are obtained, the fatty acid methyl ester separates from glycerol.
Subsequently, the fatty acid methyl ester is separated from the glycerol and purified to
biodiesel. The initial stage of the reaction is characterized by severe diffusion limi-
tations caused by the limited mutual solubility of natural oil and methanol. The low
mutual solubility reduces the rate of mass transfer of the reagents across the phase
boundary. Altogether, the rate of biodiesel production is limited by the slowest step in
2 Biodiesel Production Systems: Reactor Technologies 17

the sequence of physical and chemical steps. When diffusion limitations are present,
the overall reaction rate is reduced. To counter this effect the interphase area is
increased by technical means, and various mixing technologies, such as mechanical
stirring, static mixers, sonication, or microreactors, have been explored. A larger
interphase area leads to a higher flux of reagents across the phase boundary and helps
overcome the diffusion limitations. More recent technologies for process intensifi-
cation are aimed at overlaying chemical reaction with physical separation of products
and excess reagents.
Diverse reactor concepts have been developed for the transesterification of
natural oils with alcohols to biodiesel (Dimian and Bildea 2008). Depending on the
scale of the biodiesel production plant, the reactor designs comprise batch,
semi-batch, or continuous operation (He and Gerpen 2016).
Biodiesel production plants based on batch reactors require a vessel that is filled
with the reagents natural oil, methanol and catalyst as well as equipment for
work-up. This results in comparably small initial capital and infrastructure invest-
ment. The operation of batch reactors is flexible and allows accommodating vari-
ations in feedstock type, composition, and quantity. Even so, the major drawbacks
of biodiesel batch processes include low productivity due to the time needed to
charge and empty the reactor, a certain variation in product quality as every batch is
unique, and more intensive labor and energy requirements compared to continuous
operation.
Biodiesel processes based on reactors operated in the semi-batch mode are
similar to the batch process. The production commences with a smaller volume of
the reaction mixture than the vessel will hold. Reactants are then added until the
vessel is filled. This process, however, is relatively labor-intensive and rarely used.
Biodiesel processes based on continuous-flow reactors are preferred over batch
processes in large-capacity commercial production. The most common type of
continuous-flow reactor is the continuous stirred-tank reactor. Yet mixing of the two
phases present at the initial stage of the reaction can be a challenge. Conventionally,
mechanical stirrers ensure macro-mixing of the reaction mixture. The use of static
mixers can enhance the mixing. Micro-mixing can be improved by ultrasound or by
operation at supercritical conditions. Fixed bed tubular reactors have been intro-
duced to overcome intrinsic limitations in conversion caused by the use of
back-mixed reactors. Even though reactive distillation has been explored to intensify
the process, it has not yet been used to produce biodiesel on a commercial scale.
Compared to biodiesel production in batch processes, continuous operation
results in a more constant product quality. In addition, lower operating costs are
obtained per unit of product. Certain capital investment, however, is needed to build
the plant. In general, continuous-flow processes require intricate process controls
and online monitoring of product quality. Pumps for natural oil and methanol and
the dosing system for the catalyst are operated continuously. Moreover, the pumps
used for product removal and the equipment in downstream processing are operated
continuously. Feedback loops of process analytics to the operation parameters
conform to the principles of Green Chemistry (Gupta et al. 2010; Jessop et al. 2009;
Anastas and Eghbali 2010). Within certain constraints, the production capacity can
18 T. E. Müller

be adjusted to the availability of natural oil by adjusting the feed rate. Since the
vessel size is usually fixed, such changes in the feed rate lead to a different resi-
dence time of the reaction mixture in the vessel. Adjusting the ratio of
methanol-to-natural oil as well as the catalyst concentration helps to compensate for
the resulting changes in conversion.

2.1.1 Mixing in Biodiesel Production

In biodiesel production, natural oil and methanol form a two-phase system. The
limited mutual solubility of natural oil and methanol results in diffusion limitations
across the phase boundary restricting the overall rate of reaction (Gerpen et al.
2005). The difference in solubility is particularly relevant at the onset of the
reaction. Creating a large interphase between the natural oil-rich phase and the
alcohol-rich phase enhances mass transfer across the phase boundary. Thus, good
reaction engineering is required in biodiesel production. If the interphase area is too
small, the chemical reaction rate is slowed down by the limited availability of one
of the reactants. The reaction rate over conventional catalysts depends on the
concentration of natural oil and methanol (Wei et al. 2014), and the highest reaction
rates are observed when the concentrations of both reagents are approximately
equal. This holds for homogeneous catalysts, such as sodium or potassium
hydroxide. With potassium hydroxide, the conversion of triglyceride to diglyceride
is the rate-determining step with an activation energy of 30.2 kJ/mol and 26.8 kJ/
mol for palm oil and mustard oil, respectively (Issariyakul and Dalai 2012). In the
case of heterogeneous catalysts, the concentration of methanol and natural oil on the
surface of the catalyst ought to be similar (Ilgen and Akin 2012).

2.1.2 Batch Reactors

The batch reactor is typically a vessel that is equipped with some type of agitation.
The main characteristics of a batch reactor are that the vessel is first filled with
unreacted material, the reaction then proceeds, and the reaction mixture is removed
sometime later on. Consequently, the vessel holds a reaction mixture with different
compositions depending on which time one happens to look at it. For biodiesel
production, the tank is filled with the reactants, i.e., natural oil, alcohol, and cata-
lyst. The reaction mixture is then heated and agitated for a certain period. After the
required time has elapsed, the contents of the vessel are drained out, fatty acid ester
and glycerol are separated, and the two products are further processed. Batch
reactors are generally used in small biodiesel production plants, but they are rela-
tively inflexible in terms of productivity. To increase production, it may be nec-
essary to reduce the cycle time, set up further vessels or replace the vessel with a
larger vessel.
2 Biodiesel Production Systems: Reactor Technologies 19

2.1.3 Continuous-Flow Reactors

The most common continuous-flow system in biodiesel production is the continuous


stirred-tank reactor (CSTR). As for the batch process, the reactor is conventionally a
vessel that is equipped with some type of agitation. The reactor is set up in a
continuous-flow system. The reactants are added continuously, and an equal mass
flow of the product mixture is continuously withdrawn. Adequate agitation is required
to increase the interphase area between the two phases as well as to ensure uniform
chemical composition and temperature in all volume elements of the reaction mixture.
Characteristic for the operation of a CSTR is that the incoming stream of reactants
becomes mixed with the reaction mixture contained in the vessel. Due to the resulting
low concentration of the reactants in the reaction mixture, the productivity of a CSTR
for biodiesel production is low compared to that of a vessel operated in batch or in a
plug flow reactor. When a CSTR is operated at steady state, the concentration of
reactants, intermediates, and products is even in all volume elements of the vessel and
with time. The chemical composition of the reaction mixture at the reactor outlet is
equal to the composition in the reaction mixture. Because of this so-called
back-mixing of the reaction mixture, there is always a certain concentration of unre-
acted reactants and intermediates at the outlet of the reactor. To address this issue, the
percentage of conversion can be raised by increasing the reactor size and, hence, the
residence time of the reaction mixture inside the vessel (Fig. 2.1).
To enhance conversion, more than one reactor can be used in a cascade. In
biodiesel production, the process often involves an arrangement of two consecutive
CSTRs (He and Gerpen 2016). In the first reactor, the natural oil is reacted with
approximately 80% of the alcohol. Then, the outlet stream goes through a glycerol
removal step before entering the second reactor. The remaining 20% of the alcohol
is then added to this reactor. As a result, this system generates higher conversions of
the natural oil. A lower excess of alcohol is needed compared to a process involving
a single CSTR (Fig. 2.2).

Fig. 2.1 Relative size of a 1000


continuous-flow reactor in
Volume VR,CSTR / Volume VR,Batch

400
comparison to the size of a
batch reactor to achieve a
certain conversion (adopted 100
from Emig and Klemm 2005).
The analysis assumes a 40
first-order reaction A ! B
10

1
99.9 99.6 99 96 90 60 0
Conversion [%]
20 T. E. Müller

Dried oil

Methanol

Catalyst

Biodiesel

Phase Phase
separaƟon separaƟon

Glycerol
First reactor Second reactor

Fig. 2.2 Biodiesel production in a process with a cascade of two continuous stirred-tank reactors
(adopted from He and Gerpen 2016)

Tubular reactors provide higher conversions than back-mixed reactors in


continuous-flow operation, resulting in a higher efficiency than reactions in CSTR
at comparable residence times (Emig and Klemm 2005). Furthermore, the use of a
back-mixed reactor followed by a tubular reactor has been suggested (Suryanto
et al. 2015). Oscillatory flow provides enhanced mixing of the phases (Suryanto
et al. 2015; Harvey et al. 2001). In combination with a heterogeneous catalyst, a
fixed bed reactor is frequently used. Adiabatic reactors with upstream or down-
stream heat exchangers make the individual apparatus simpler. As with a CSTR, a
two-stage reactor concept is advantageous. In typical operation, methanol and
natural oil are fed into a first transesterification reactor. Favorable conditions are,
e.g., a temperature range of 200–220 °C, a pressure range of 40–70 bar at a liquid
hourly space velocity of 0.5–1 h−1 and a methanol-to-oil ratio of 1:2 (Dimian and
Rothenberg 2016). To limit the pressure drop, relatively large catalyst particles are
employed. Typical catalyst beds comprise 3 mm extrudates. The exit stream of the
first reactor is decompressed thereby removing a large fraction of the unconverted
methanol. To shift the equilibrium, the phases are separated and glycerol is
removed by decantation. The fatty acid ester phase is mixed with a feed of methanol
and enters a second fixed bed reactor operated at similar conditions as the first
reactor, and the conversion is increased from 90–93% to up to 99.5%. The product
mixture then undergoes work-up in a separation train, thereby removing excess
methanol and glycerol from the biodiesel. An advantage of using heterogeneous
catalysts for biodiesel refining is that the neutralization step can be omitted. This
reduces the need to wash the biodiesel and glycerol and thus limits waste salt
production.
2 Biodiesel Production Systems: Reactor Technologies 21

2.1.4 Static Mixers in Biodiesel Reactors

Static mixers are devices consisting of spiral-shaped internal parts within an


enclosure, such as a tube or pipe that promotes turbulent flow. Static mixers have no
moving parts and can be very effective at mixing liquids that are not readily
miscible under normal conditions. In biodiesel production, static mixers can
improve mixing or the transesterification reaction is performed entirely in such
static mixers. Favorable conditions to achieve close to complete transesterification
are, e.g., a temperature of 60 °C at a residence time of 30 min using a catalyst
concentration of 1.5% (Thompson and He 2007).

2.1.5 Reactors Operated at Supercritical Conditions

Conventionally, homogeneous catalysts such as sodium or potassium hydroxide or


sulfonic acids are used in biodiesel production to increase the rate of the transes-
terification reaction. The catalyst has to be removed after the reaction to ensure fuel
quality. Applying supercritical conditions reduces the need to use large amounts of
catalyst in biodiesel production (Wen et al. 2009; Cao et al. 2016). In the super-
critical state, the differences between liquid and vapor disappear, and the mixture
assumes the qualities of both a liquid and a vapor. The lower limit of the super-
critical conditions is defined by the critical point of a fluid that, in turn, is defined by
its critical temperature and critical pressure. For methanol, the supercritical region is
above 240 °C and 7.95 MPa (Sykioti et al. 2013). At temperatures and/or pressures
above the critical point, the natural oil dissolves in methanol to form a single phase.
Due to decreased diffusion limitations, the transesterification occurs much faster to
reach completion in the supercritical state. Even without a catalyst, the reaction is
completed in a few minutes. This allows for short residence times in comparatively
small reactors. Noteworthy, a certain amount of water and free fatty acids is tol-
erated in the system. Soap formation, common in the traditional process, is elim-
inated (Demirbas 2006; van Kasteren and Nisworo 2007; Saka and Isayama 2009).
A drawback is the high temperatures and pressures needed to achieve the super-
critical state. To take the reaction mixture to supercritical conditions, a certain
amount of energy is required. Part of this energy can be recovered during
decompression and through heat exchange between inlet and outlet streams. Even
though implementing pressure reactors and heaters leads to higher investment costs,
large biodiesel producers may find this process advantageous.
22 T. E. Müller

2.1.6 Use of Ultrasound in Biodiesel Reactors

Mixing in biodiesel reactors can be enhanced by employing ultrasound (Veljkovic


et al. 2012). Generally, ultrasound can be a useful tool to improve the mixing of
liquids that tend to separate. The ultrasound energy that is transferred into the fluid
creates intense vibrations and cavitation bubbles. When the bubbles burst, sudden
contraction of the fluid leads to intense mixing of the phases (Rooze et al. 2013;
Hihn et al. 2012; Suslick and Flannigan 2008). The enhanced mixing by ultrasound
can considerably increase the overall rate of the transesterification reaction (Sarma
et al. 2008; Hanh et al. 2008; Teixeira et al. 2009). As an alternative to ultrasound,
hydrodynamic cavitation has also been suggested (Chuah et al. 2017). Higher
conversions are obtained or the reactor can be operated at shorter residence times.
Alternatively, the process can be run at a lower temperature, thereby reducing the
need to heat the reaction mixture and saving considerable energy. Especially for
small producers, ultrasound can be a good choice (He and Gerpen 2016). Here,
small producers refer to operators with a capacity of fewer than 2 million gallons
per annum.

2.1.7 Microreactors in Biodiesel Production

The use of microreactors has also been explored for biodiesel production (Šalić and
Zelić 2011). The two phases are mixed in a micromixer followed by a residence
time unit. The micromixer gives rise to the formation of tiny droplets. Transport
over the phase boundary is enhanced by slug flow in the residence time unit, where
a series of liquids slugs of one phase are separated by the second phase. Using this
approach, natural oil has been fully converted with a residence time of 2 min at a
temperature of 60 °C (Guan et al. 2008). A zigzag configuration of the
microchannels is beneficial, because it reduces droplet size. At a residence time of
0.5 min and a temperature of 56 °C, 99.5% yield of the methyl ester was obtained
using a 9:1 molar ratio of methanol to natural oil and a catalyst concentration of 1.2
wt% (Wen et al. 2009).

2.1.8 Reactive Distillation for Biodiesel Production

In reactive distillation, chemical reaction and product separation occur concurrently


in one unit. Reactive distillation is commonly employed as a unit operation for
reversible chemical reactions. In reactive distillation, the reaction products are
progressively removed from the reaction zone, thus reducing the extent of backward
reaction. The result is an enhanced overall rate in the forward direction. A reactive
distillation column consists of a series of stages, where a down-flowing liquid is
2 Biodiesel Production Systems: Reactor Technologies 23

brought into contact with an up-flowing gas stream. In parallel to the mass
exchange between liquid and gas phase, the chemical reaction takes place inside the
column. The reaction progresses as the mixture passes through successive stages.
Tray columns with discrete stages or packed columns where stages are defined by a
certain part of the column height may be used. In general, tray columns are pre-
ferred for homogeneous reaction systems due to the higher liquid holdup and the
resulting longer retention times.
In biodiesel production, reversible transesterification of natural oil and methanol
to fatty acid methyl esters can be promoted by reactive rectification. The large
difference in the boiling temperature of fatty acid methyl esters and methanol
facilitates the separation. When the reactive rectification column is operated at
ambient pressures, the temperature the reaction mixture assumes is determined by
the temperature methanol boils from the reaction mixture. The transesterification
reaction occurs only in the liquid phase. Consequently, the conversion is controlled
by the residence time of the liquid phase and the catalyst concentration. The resi-
dence time is established by the ratio of liquid holdup and the feed rate of the
reactants.
For a reactive distillation reactor system, a reaction time of 10–15 min and
productivity of 7–9 gallons per gallon reactor volume per hour have been reported
at a relatively low excess of alcohol of approximately 3.5:1 mol/mol (He et al.
2005; He et al. 2006, 2007). Another study reports 99.8% conversion at a residence

Condenser
Vapor stream enriched with
low-boiling component

Ascending
vapor stream
Methanol

Catalyst

Dried oil

Descending
Reboiler liquid stream
Reflux enriched with
high-boiling component
Biodiesel
Glycerol

Fig. 2.3 Reactive distillation setup for biodiesel production (adapted from He and Gerpen 2016)
24 T. E. Müller

time of 6 min (Silva et al. 2013). The smaller excess of alcohol reduces downstream
separation steps and energy demand. The generally smaller size of the reactor and
distillation train may reduce the investment costs compared to conventional
continuous-flow plants. So far, reactive distillation has not been introduced com-
mercially in biodiesel production, because reactive distillation columns tend to be
more complex to operate (Fig. 2.3).

2.2 Concluding Remarks

As fossil resources are dwindling, biodiesel is gaining further importance as a


renewable resource in the fuel sector. In biodiesel production, natural oils are
reacted with methanol to the corresponding fatty acid methyl ester. Diverse reactor
concepts have been developed for the transesterification reaction. Depending on the
scale of the biodiesel production plant, the reactor designs comprise batch,
semi-batch or continuous operation. Natural oil and methanol have a low mutual
solubility, giving rise to a two-phase system characterized by severe diffusion
limitations. The low rate of mass transfer of the reagents across the phase boundary
limits the rate of the transesterification reaction especially at the onset of the
reaction. Increasing the interphase area leads to a higher flux of reagents across the
phase boundary and helps in overcoming the diffusion limitations. Various mixing
technologies, such as mechanical stirring, static mixers, sonication, and microre-
actors, have been introduced to ensure adequate mass transfer. Process intensifi-
cation is aimed at overlaying chemical reaction with separation.

References

Anastas P, Eghbali N (2010) Chem Soc Rev 39:301–312


Arous F, Frikha F, Triantaphyllidou I-E, Aggelis G, Nasri M, Mechichi T (2016) J Clean Prod
133:899–909
Bhuiya MMK, Rasul MG, Khan MMK, Ashwath N, Azad AK, Hazrat MA (2016) Renew Sustain
Energy Rev 55:1129–1146
Brennan L, Owende P (2010) Renew Sustain Energy Rev 14:557–577
Cao N, Zhang Y, Yang B, Wang Y, Zhang G (2016) Energy Sources Part A: Recovery Util
Environ Eff 38:3354–3359
Chuah LF, Klemes JJ, Yusup S, Bokhari A, Akbar MM (2017) J Clean Prod 146:181–193
Demirbas A (2006) Energy Convers Manag 47:2271–2282
Demirbas A, Bafail A, Ahmad W, Sheikh M (2016) Energy Explor Exploit 34:290–318
Dimian AC, Bildea S (eds) (2008) Computer-aided design case studies. Wiley, Weinheim,
pp 399–428
Dimian AC, Rothenberg G (2016) Catal Sci Technol 6:6097–6108
Emig and Klemm (2005) Technische Chemie, Einführung in die Chemische Reaktionstechnik.
Springer, Berlin
Feasibility report small scale biodiesel production (2006) Illinois Waste Management and
Research Center Champaign, IL
2 Biodiesel Production Systems: Reactor Technologies 25

Gerpen JV, Knothe G, Krahl J (eds) (2005) The biodiesel handbook. AOCS Press, Champaign, Ill,
pp. 26–41
Guan G, Kusakabe K, Moriyama K, Sakurai N (2008) Chem Eng Trans 14:237–244
Gupta M, Paul S, Gupta R (2010) Curr Sci 99:1341–1360
Hanh HD, Dong NT, Starvarache C, Okitsu K, Maeda Y, Nishimura R (2008) Energy Convers
Manag 49:276–280
Harvey AP, Mackley MR, Stonestreet P (2001) Ind Eng Chem Res 40:5371–5377
He B, Gerpen JV (2016) Reactors for biodiesel production, eXtension Issues—Innovation—
Impact
He B, Singh A, Thompson J (2005) Trans ASABE 48:2237–2243
He B, Singh A, Thompson J (2006) Trans ASABE 49:107–112
He B, Singh A, Thompson J (2007) Trans ASABE 50:123–128
Hihn J-Y, Doche M-L, Mandroyan A, Hallez L, Pollet BG (2012) Ultrasound for better reactor
design. In: Chen D, Sharma SK, Mudhoo A (eds) Handbook on applications of ultrasound,
pp 599–622
Ilgen O, Akin AN (2012) Appl Catal B: Environ 126:342–346
International A (2013) Standard test method for determination of total monoglycerides, total
diglycerides, total triglycerides, and free and total glycerin in B-100 biodiesel methyl esters by
gas chromatography, West Conshohocken, vol ASTM D6584
Issariyakul T, Dalai AK (2012) Can J Chem Eng 90:342–350
Jessop PG, Trakhtenberg S, Warner J (2009) ACS symposium series, 1000 (Innovations in
industrial and engineering chemistry), pp 401–436
van Kasteren JMN, Nisworo AP (2007) Resources Conserv Recycl 50:442–458
Kumar N, Varun, Chauhan SR (2013) Renew Sustain Energy Rev 21
Perego C, Ricci M (2012) Catal Sci Technol 2:1776–1786
Rooze J, Rebrov EV, Schouten JC, Keurentjes JTF (2013) Ultrason Sonochem 20:1–11
Saka S, Isayama Y (2009) Fuel 88:1307–1313
Šalić A, Zelić B (2011) goriva i maziva 50:85–110
Sarma AK, Sarmah JK, Barbora L, Kalita P, Chatterjee S, Mahanta P, Goswami P (2008) Recent
Pat Eng 2:47–58
Silva NLD, Rios LF, Maciel MRW, Filho RM (2013) Materials and processes for energy:
communicating current research and technological developments. In: Méndez-Vilas A
(ed) Formatex, pp 244–251
Suryanto, Utomo WB, Marwan (2015) Int J Sci Res (IJSR) 4:103–106
Suslick KS, Flannigan DJ (2008) Annu Rev Phys Chem 59:659–683
Sykioti EA, Assael MJ, Huber ML, Perkins RA (2013) J Phys Chem Ref Data 42:043101/043101–
043101/043110
Teixeira LSG, Assis JCR, Mendonça DR, Santos ITV, Guimarães PRB, Pontes LAM,
Teixeira JSR (2009) Fuel Process Technol 90:1164–1166
Thompson J, He B (2007) Trans ASABE 50:161–165
Veljkovic VB, Avramovic JM, Stamenkovic OS (2012) Renew Sustain Energy Rev 16:1193–1209
Wei Y, Zhang J, Zhang M, Zhang Y (2014) Advances in materials and materials processing IV. In:
Advanced materials research, Durnten-Zurich, Switzerland, vol 887–888, pp 501–504
Wen D, Jiang H, Zhang K (2009a) Prog Nat Sci 19:273–284
Wen Z, Yu X, Tu ST, Yan J, Dahlquist E (2009b) Biores Technol 100:3054–3060
Chapter 3
Biodiesel Production Systems:
Operation, Process Control
and Troubleshooting

Nídia S. Caetano, Vera Ribeiro, Leonardo Ribeiro, Andresa Baptista


and Joaquim Monteiro

Abstract Biodiesel is a renewable fuel, produced from waste cooking oils, animal
fats, vegetable and algae oils. Its use is intended to replace diesel in conventional
diesel engines, causing lower polluting emissions. To produce biodiesel, certain
details must be carefully considered, namely feedstock composition, reaction
parameters, process conditions, process equipment, purification processes, analysis
of biodiesel properties, troubleshooting and storage. In what concerns feedstock
composition, parameters such as acidity, insolubles, moisture, phospholipids, sul-
phur, polymerized triglycerides, impurities, etc., must be determined to decide
about the pretreatment steps (washing, degumming, filtration, bleaching, deodor-
ization, among others) to be implemented, and the need for esterification prior to
transesterification. In what concerns the selection of process equipment some
questions arise, namely the materials, heating methods and thermal insulation to
use, alternatives to enhance the reaction, need for neutralization and process control
system. The purification process includes biodiesel purification, methanol recovery
and glycerine valorisation. The excess methanol must be recovered from biodiesel
and glycerine by distillation and reused in the process while glycerine can be further
purified and sold for application from the chemical to the pharmaceutical industry.
The quality of biodiesel must be certified by the analyses performed according to
the standards (e.g. EN 14214, ASTM D6751). Troubleshooting is needed in bio-
diesel production during start-up and under steady production of a facility; prob-
lems may arise regarding quality and appearance of biodiesel, reaction conditions,
methanol removal, stirring in reactors, glycerine and biodiesel separation, as well as
excess of water and other feedstock impurities. Biodiesel can be stored for up to 6
months; its storage poses challenges concerning degradation by contact with air and

N. S. Caetano (&)  V. Ribeiro  L. Ribeiro  A. Baptista  J. Monteiro


ISEP – School of Engineering, P.Porto – Polytechnic of Porto,
R. Dr. António Bernardino de Almeida 431, 4249-015 Porto, Portugal
e-mail: NSC@isep.ipp.pt
N. S. Caetano
LEPABE – Laboratory for Process Engineering, Environment, Biotechnology and Energy,
Faculty of Engineering, University of Porto (FEUP), R. Dr. Roberto Frias S/N,
4200-465 Porto, Portugal

© Springer Nature Switzerland AG 2019 27


M. Tabatabaei and M. Aghbashlo (eds.), Biodiesel,
Biofuel and Biorefinery Technologies 8,
https://doi.org/10.1007/978-3-030-00985-4_3
28 N. S. Caetano et al.

light, which cause oxidation. Some additives could extend the lifespan of biodiesel
by increasing oxidation stability; other technique is the fractionation to remove the
undesired fatty acid methyl ester (FAME).

Nomenclature
ACEA European Automobile Manufacturers’ Association
ASTM American Society for Testing and Materials
BXX XX% (v/v) Biodiesel
CO2 Carbon Dioxide
CO Carbon Monoxide
DIN German Institute for Standardization
EN European Normalization
EU European Union
FAME Fatty Acid Methyl Ester
FFA Free Fatty Acid
HC Hydrocarbon
HFRR High-Frequency Reciprocating Rig
ISO International Organization for Standardization
NaOH Sodium Hydroxide
NOx Nitrogen Oxides
RME Rapeseed Oil Methyl Ester
SO2 Sulphur Dioxide

3.1 Introduction

The main purpose of biodiesel production is to use it as a renewable fuel instead of


diesel in diesel engines for automobiles, trucks, farm equipment, marine vessels,
planes, etc. Most of the diesel engines already in the market need little or no
modification to burn even pure biodiesel (B100). However, it is convenient to be
aware of two sorts of problems that may appear after switching from diesel to
biodiesel in a diesel engine. First, in a heavy-duty engine, already in service for a
long time, biodiesel can cause clogging in some parts since biodiesel is a better
lubricant and solvent than diesel, and so can cause the release of deposits from
metallic walls of the engine. Second, old engines produced before 1993 may have
seals and hoses made of rubbers not compatible with biodiesel. Due to its solvent
capacity, these rubbers can swell or degrade. Table 3.1 lists the models of vehicles,
whose engines are compatible with biodiesel.
According to Haseeb et al. (2011), the materials used in the construction of
engines can be grouped mainly into ferrous alloys, non-ferrous alloys and elas-
tomers. Common elastomers like natural rubber, nitrile, chloroprene/neoprene, etc.,
are not suitable for use with biodiesel (Table 3.2). Between the other two groups of
3 Biodiesel Production Systems: Operation … 29

Table 3.1 Models of vehicles compatible with biodiesel mixed with diesel
Manufacture Model Remarks Source
Case All recent models Must use Viton and Teflon *
Cummins All recent models – *
Caterpillar All recent models Biodiesel must meet Standard ASTM D *
6751
Ford All recent models Only B5 FAME *
GM All recent models Since 2004 only B5 *
John Deere All recent models Biodiesel must meet Standard ASTM *
PS 121 and DIN 51606
Kubota All models Only B5 *
Mack All models Only B5 *
Mercedes-Benz All models with Only B5 USA (must meet *
Common Rail Injection Standard ASTM D 6751) **
Only B7 EU (must meet Standard EN
14112)
New Holland All models Only B20 *
Nissan All models Only B5 must meet Standard ASTM D *
6751
Volvo All models Only B5 RME
Volkswagen All recent models Only B5 USA (must meet *
Standard ASTM D 6751) **
Only B7 EU (must meet Standard EN ***
14112)
B100 RME in vehicles denoted with
PR code 2G0
* www.officialbiodiesel.com
** ACEA (2014)
*** Volkswagen (2010)

materials, viz., ferrous alloys and non-ferrous alloys, the former are more resistant
to biodiesel attack. Among non-ferrous alloys, copper alloys and lead alloys are the
most vulnerable to biodiesel attack, followed by aluminium (Bhardwaj et al. 2014).
Biodiesel contains about 10% (w/w) of oxygen—this oxygen contributes to
lowering the air requirements for the combustion of biodiesel. Biodiesel is
biodegradable; its spillage will not harm the environment, and hence, biodiesel is
very suitable for marine vessels. The lubricant properties of biodiesel are more
pronounced than those of diesel, increasing the lifespan of the engine.
The heating value of biodiesel is about 10% lower than that of diesel, which
means that the performance of a diesel engine is not significantly affected by
switching fuel from diesel to biodiesel. According to Xue et al. (2011), the
reductions in emissions were accompanied with negligible power loss and increase
in fuel consumption. Regarding safety aspects, biodiesel is not harmful neither to
individuals nor to the environment (Demirbas 2008a) and it is safe to handle and
30 N. S. Caetano et al.

Table 3.2 Compatibility of biodiesel and diesel with materials commonly used in seals (DuPont
2017)
Material Type of biodiesel Comparison with diesel
Teflon B100 Slight difference
Nylon 6/6 B100 Slight difference
Nitrile B100 Hardness lowers 20%; swelling 18%
Viton A401-C B100 Slight difference
Viton GFLT B100 Slight difference
Fluorosilicone B100 Hardness is the same; swelling 7%
Fluoroetane B100 Hardness is the same; swelling 6%
Polypropylene B100 Hardness drops 10%; swelling 8–15%
Polyvinyl B100 Much worse
Polyvinyl B50 Worse
Polyvinyl B40 Worse
Polyvinyl B30 Worse
Polyvinyl B20 Equal
Polyvinyl B10 Equal
Tygon B100 Worse

store, as it is unlikely to explode owing to its high flash point. However, the
production process of biodiesel can be hazardous because lye (NaOH) is often used
as a catalyst, and methanol, the most frequently used alcohol, is highly flammable
(Demirbas 2008b).
The pollutants emissions from biodiesel combustion, namely CO2, CO, hydro-
carbons (HC) and particulates are lower than those from diesel combustion; besides,
since biodiesel has no sulphur, its burning does not generate SO2. Wang et al.
(2000), based on field test results, showed that heavy-duty trucks, fuelled with B35,
emitted significantly lower particulate matter (PM) and moderately lower CO and
hydrocarbon (HC) than the same trucks fuelled with conventional diesel. It should
be noted that the negative effects of particulates on human health, such as triggering
asthma and allergies, have been well documented (Rajagopal and Zilberman 2007).
The amount of CO2 generated by burning a certain quantity of biodiesel is slightly
above the amount of CO2 absorbed during the growth of the feedstock crops from
which the same quantity of biodiesel is produced—hence, biodiesel does not
contribute significantly to global warming.
Table 3.3 shows the average percentage variation of some pollutants emissions
generated after switching from diesel to biodiesel (B20 and B100) in engines.
Polycyclic Aromatic Hydrocarbons (PAH) and Nitrated Polycyclic Aromatic
Hydrocarbons (nPAH) are among the major health-threatening emissions and can
induce mutagenicity and consequently cancers.
As mentioned earlier, many studies have shown that the use of biodiesel as
replacement of diesel results in lower emissions of CO, particulates and soot
(Pushparaj and Ramabalan 2013; Miri et al. 2017; Putrasari and Lim 2017).
3 Biodiesel Production Systems: Operation … 31

Table 3.3 Average percentage variation of some pollutants emissions generated after switching
from diesel to biodiesel (B20 and B100) in engines (NBB 2018)
Pollutant B100 B20
CO2 (%) −76.4 −15.3
CO (%) −48.1 −12.3
Hydrocarbons (%) −67.4 −20.1
Particles (%) −47.2 −12.0
NOx (%) +10.3 0
SO2 (%) −100.0 −20.0
Toxic gases (%) −60 to −90a −20a
Polycyclic Aromatic Hydrocarbons (PAH) (%) −80b −13b
Nitrated Polycyclic Aromatic Hydrocarbons (nPAH) (%) −90b −50b
Speciated Hydrocarbons Ozone-Forming Potential (%) −50b −10b
a
Biodiesel Fact Sheet
b
Biodiesel Emissions

However, increase in NOx emissions has been observed on many occasions (Dincer
2008). For instance, it has been reported that B20 and B30 blends would generally
result in statistically significantly higher NOx emissions (Bakeas et al. 2011). The
emission of aromatic and polyaromatic compounds, as well as their toxic and
mutagenic effect, have been generally considered to be reduced with biodiesel use
(Lapuerta et al. 2008).
Table 3.4 lists important properties required for both diesel and biodiesel
according to some of the international standards, enabling the comparison of bio-
diesel with diesel.
Table 3.5 tabulates the required specifications of diesel and biodiesel in terms of
the composition and thermophysical properties of the fuel, according to the
National Renewable Energy Laboratory (NREL) of the USA.

3.2 Feedstock for Biodiesel Production

Biodiesel can be produced from different types of feedstock such as waste oils,
animal fats, edible and non-edible vegetable or algae oils (Mata et al. 2014). These
types of feedstock are very different from each other in terms of composition,
physicochemical properties but also in terms of full life cycle analysis and prices
(Upham et al. 2009; Jeschke 2009).
In Europe, biodiesel is most commonly produced from rapeseed oil; in the USA,
biodiesel is predominantly produced from soybean oil; in Malaysia and Indonesia,
palm oil is the most significant source of biodiesel; while in India and Southeast
Asia, Jatropha tree is the most important source for biodiesel production (Demirbas
2009). The most wanted vegetable oil sources are soybean, canola, palm and
32 N. S. Caetano et al.

Table 3.4 Properties required for diesel and biodiesel as set forth by international standards
Property Diesel Biodiesel Biodiesel
(EN 590) (DIN 51606) (EN 14214)
Density at 15 °C (kg/m3) 820–845 875–900 860–900
Kinematic viscosity at 40 °C (mm2/s) 2.0–4.5 3.5–5.0 3.5–5.0
Flash point (°C) >55 >110 >101
Sulphur content (mg/kg) <10 <10.0 <10.0
Sulfated ash content (% weight) <0.01 <0.03 <0.02
Water content (mg/kg) <200 <300 <500
Carbon residue (% w/w) <0.3 <0.3 <0.3
Total contamination (mg/kg) <24 <20 <24
Corrosion by copper (3 h at 50 °C) Class I Class I Class I
Cetane number >51.0 >49.0 >51.0
Methanol (% weight) – <0.3 <0.20
Ester content (% weight) – – >96.5
Monoglyceride content (% weight) – <0.8 <0.8
Diglyceride content (% weight) – <0.4 <0.2
Triglyceride content (% weight) – <0.4 <0.2
Free glycerine (% weight) – <0.02 <0.02
Total glycerine (% weight) – <0.25 <0.25
Iodine value (g I2/100 g) – <115 <120
Phosphorous content (mg/kg) – <10 <4.0
Alkali metals as Na and K (mg/kg) – <5.0 <5.0

Table 3.5 Required specifications for diesel and biodiesel as set forth by the NREL
Feature Biodiesel Diesel
Standard ASTM D6751 ASTM D975
Composition C12–C22 FAME C10–C21 HC
Lower heating value (kJ/m3) 32,636 36,594
Kinematic viscosity at 40 °C (mm2/s) 2–6 2–4
Specific heat at 300 K (J/kg/K) 1909 1850
Density at 15 °C (kg/m3) 878 848
Water (% volume) 0.05 0.05
Carbon (% mass) 77 87
Hydrogen (% mass) 12 13
Oxygen (% mass) 11 0
Sulphur (% mass) 0.05 0.05
Temperature of solidification (°C) −15 to 16 −35 to −15
Cetane number 48–60 40–55
BOCLE Scuff (g) >7000 3600
HFRR (µm) 314 685
3 Biodiesel Production Systems: Operation … 33

rapeseed. The principal animal fat sources are beef tallow, lard, poultry fat and fish
oils (Demirbas 2008b).
Taking into account only the chemical reaction, the best feedstock for biodiesel
production is refined vegetable oil because with such raw material the transesteri-
fication reaction is the most complete and the shortest (Demirbas 2008b). However,
at the feedstock selection stage for a biodiesel production plant, social, economic,
geographical and engineering aspects must also be taken into account. These aspects
are mainly related to the life cycle of each feedstock, including availability of land,
logistics, costs (transportation and storage), energy supply and balance, greenhouse
gas (GHG) emissions, use of pesticides, soil erosion and fertility, contribution to
biodiversity losses, creation and maintenance of jobs, as well as water availability
and effects on air quality. Other important features have to do with taxes, policies,
grants and legislation in force in each country (Mata et al. 2014).
Biodiesel production costs are mainly influenced by the feedstock price, which
can represent 70–80% of the total biodiesel production cost when using vegetable
oils. However, biodiesel production cost can be reduced by up to 50–70% by using
waste oils or animal fats (Hajjari et al. 2017).
Engineering of a biodiesel production plant is complex, multidisciplinary and
integrates a set of stages such as definition of the processing system, plant design,
plant and equipment sizing, heat and mass balances, chemical products require-
ments, utilities and waste management, construction materials, among others. All
these stages of an engineering project depend on the feedstock composition and its
physicochemical properties (Mittelbach 2009).
Table 3.6 presents the average areal productivity of raw oil obtained through
pressing or solvent extraction, yielded from several oleaginous plants commonly
used in the production of biodiesel.

3.2.1 Novel Feedstocks-Microalgae

Biodiesel oil feedstocks can be obtained from oleaginous plants as well as from
algae, through either pressing or solvent extraction. Microalgae grow in salt, fresh
and brackish water, depending on the species. There are 20,000–30,000 species of
microalgae (Mata et al. 2010); some common species include Ptatymonas,
Botryococcus braunii, Cyclotella, Chlorella protothecoides, etc. The conditions for
successful microalgae production vary, i.e., water salinity from 0.1 to 10%; water
temperature from 2 to 30 °C; water pH of over 5.8. The main advantage of
microalgae is the fact that they do not compete with food crops.
There are installations to produce microalgae consisting essentially of a shallow
depth channel, in a closed circuit, where water flows, moved by pumps, with
microalgae in suspension (open ponds). There are also installations, as the previous
one, except that the open channel is replaced by tubes of polymeric and transparent
materials. The productivity of the process can be boosted by adding CO2 and/or
residual waters to the referred water circuit. CO2 helps with photosynthesis
34 N. S. Caetano et al.

Table 3.6 Average areal Oleaginous crop kg oil/ha year


productivity of raw oil
obtained from some Avocado 2217
oleaginous plants (Tickell Cocoa 863
et al. 1999; Karmakar et al. Coconut 2260
2010; Mata et al. 2010) Corn 145
Flax (Linseed) 402
Jatropha 1590
Jojoba 1528
Macadamia 1887
Microalgae 47,000–94,000
Olive 1019
Opium poppy 978
Palm 5000
Peanut 890
Rapeseed 1000
Rice 696
Soy 420
Sunflower 800
Walnut pecam 1505

(and can be provided by an industrial unit operating in the vicinity of the microalgae
production facility) whereas residual waters contain nutrients, such as NO−3 or NH+4 ,
promoting the growth of microalgae (Show et al. 2017). The average productivity, in
open ponds without special care is from 10 to 30 g/m2/d (dry weight) while with CO2
and nutrients, the productivity can reach 170 g/m2/d (dry weight) (Pulz 2007; Zittelli
et al. 2013). It is anticipated that in the near future, extensive production of
microalgae in photobioreactors can make the whole process more sustainable.

3.3 Production Process

In Europe, the production of biodiesel has increased in the last decades. The pro-
duction of biodiesel for several European countries in 2013 is listed in Table 3.7.
According to the literature, several processes can be used to reduce the viscosity
of oils for its use in diesel engines. These processes include blending, esterification,
transesterification, micro-emulsification and pyrolysis (Knothe et al. 1997;
Demirbas 2008c; Vijayaraj and Sathiyagnanam 2016). Transesterification is widely
used in biodiesel production since it is more efficient to produce a fuel with
properties that comply with biodiesel standards. The transesterification process
occurs in three consecutive reversible reactions between the oil or fat and an
alcohol, in which triglycerides are converted, stepwise, to diglycerides, mono-
glycerides, and finally glycerol, as shown in Fig. 3.1 (Mahmudul et al. 2017).
3 Biodiesel Production Systems: Operation … 35

Table 3.7 Biodiesel Country Ton/year


production in several
European countries in 2013 Austria 239,000
(European Biodiesel Board Belgium 565,000
2017) Bulgaria 13,000
Croatia 33,000
Cyprus 1000
Czech Republic 210,000
Denmark/Sweden 334,000
Estonia 0
Finlanda 320,000
France 1,885,000
Germany 2,516,000
Greece 220,000
Hungary 150,000
Irelanda 24,000
Italy 387,000
Latvia 61,000
Lithuania 118,000
Luxembourg 0
Malta 1000
Netherlandsa 1,248,000
Poland 648,000
Portugal 314,000
Romania 137,000
Slovakia 105,000
Slovenia 2000
Spain 618,000
United Kingdom 277,000
a
Data include hydro-diesel production

The formation of alkyl esters and glycerol through the reaction between mono-
glycerides and methanol determines the overall reaction rate since monoglycerides
are the most stable intermediates (Demirbas 2005).

Fig. 3.1 The reactions Triglyceride (TG) + ROH ↔ Diglyceride (DG) + RCOOR1
involved in the
transesterification process Diglyceride (DG) + ROH ↔ Monoglyceride (MG) + RCOOR2
leading to biodiesel and
glycerol production Monoglyceride (MG) + ROH ↔ Glycerol + RCOOR3
Or in summary:
RCOOR1
Triglyceride (TG) + 3 ROH ↔ Glycerol + RCOOR2
RCOOR3
36 N. S. Caetano et al.

Stoichiometrically, the reaction requires 3 mols of alcohol for each mol of


triglyceride. However, an excess of alcohol is usually used to increase the reaction
yield. Several aspects such as the type of catalyst (if it is used), alcohol/oil molar
ratio, temperature, pressure, water and free fatty acid (FFA) content can influence
the transesterification reaction efficiency. The transesterification can be performed
via non-catalytic or catalytic processes, while the latter can involve using homo-
geneous catalysts, heterogeneous catalysts, or biocatalysts. Regardless of the use of
catalyst, the efficiency of the transesterification reaction can be enhanced using
microwaves, supercritical conditions, ultrasounds and/or membranes (Vyas et al.
2010).

3.3.1 Non-catalytic and Catalytic Transesterification

The non-catalytic transesterification is a thermochemical process performed in


tubular or bubble column reactors at high temperature (250–500 °C). Biodiesel
production by non-catalytic transesterification can achieve efficiencies of up to 95–
99%. However, it has some disadvantages such us biodiesel degradation prone to
occur above 270 °C and high production costs related to the high temperatures used
(Kwon et al. 2013).
The catalytic transesterification can be carried out using homogeneous catalysts,
heterogeneous catalysts or biocatalysts. Moreover, both homogeneous and hetero-
geneous catalysts can be acid or alkali.
The use of biocatalysts in transesterification is advantageous since it overcomes
disadvantages of chemical catalyst and requires moderate temperatures (<70 °C),
which leads to lower operation costs. However, the application of biocatalysts in
transesterification reaction is limited owing to the inhibitory effect of alcohol and
their sensitivity to the reactions parameters, such as temperature, pH, alcohol/oil
molar ratio among others. Biocatalysts can be deactivated by alcohol when the
molar ratio of alcohol to triglycerides exceeds the stoichiometric ratio for the
transesterification reaction. Some studies have been carried out to develop tech-
nologies aimed at avoiding biocatalyst inhibition (Mesiano et al. 1999; Ejikeme
et al. 2010; Ghaly et al. 2010; Liu et al. 2012).

3.3.1.1 Supercritical Conditions

The transesterification reaction carried out under supercritical conditions can be


non-catalytic or catalytic, by using chemical catalysts or biocatalyst. Biodiesel
production under supercritical conditions is carried out by using the alcohol at
pressure and temperature above its critical pressure and temperature. For methanol,
which is the most commonly used alcohol in biodiesel production, the critical
3 Biodiesel Production Systems: Operation … 37

temperature is 512.2 K and the critical pressure is 8.1 MPa, which means that the
methanol used for transesterification reaction must be at above 512.2 K and
8.1 MPa (Demirbas 2005).

3.3.1.2 Microwave-Assisted Process

The use of microwave technology accelerates the transesterification reaction


through direct energy delivery by microwave irradiations (in the wavelengths
ranging from 0.01 to 1 m and the correspondent frequency range of 0.3–300 GHz)
(Gude et al. 2013). The efficiency of microwave technology is related to the shorter
time needed for heat transfer compared with conventional heating; this is due to
heterogeneous heating and its dependency on the thermal conductivity of materials
(Tran et al. 2017).

3.3.1.3 Ultrasound-Assisted Process

The ultrasound technology is based on the ultrasonic irradiations at high frequen-


cies (2–10 MHz) or low frequencies (20–100 kHz). The ultrasonic irradiation, in
reactant mixture, creates molecular cavitation (formation, growth and implosive
collapse of bubbles) which promotes local heat generation in a shorter time com-
pared with the conventional heating (Tran et al. 2017).

3.3.1.4 Membrane-Assisted Process

Biodiesel production assisted by membrane technology occurs in a membrane


reactor, which can work in batch (Cheng et al. 2010), semi-batch (Dubé et al. 2007)
or continuous flow (Baroutian et al. 2011). Membranes can be manufactured with
several types of materials, i.e., organic or inorganic. Membrane reactors are used,
simultaneously, for chemical reaction and separation of products (Tran et al. 2017).
When emulsified oil droplets are dispersed in alcohol, the transesterification
reaction occurs at the interface between oil and methanol phases. Since the reaction
products (biodiesel and glycerol) are soluble in alcohol and alcohol molecule size is
smaller than the pore size of the membrane, the mixture of alcohol/biodiesel passes
through the membrane but the membrane retains the unreacted oil droplets. The
chemical reaction taking place in membrane reactors can also be catalyzed by
homogeneous catalyst, heterogeneous catalyst, or biocatalyst but recent studies
indicate that heterogeneous catalyst shows better efficiency compared with the
homogenous catalysts and biocatalysts (Xu et al. 2015). Biodiesel production using
membranes technology has gained more attention over the past years but its
application is still limited (Tran et al. 2017).
38 N. S. Caetano et al.

3.3.2 Free Fatty Acids and Water Content

The presence of FFAs and water allows the existence of undesired parallel reac-
tions, which cause soaps formation and catalyst consumption, entailing the decrease
of the transesterification yield (Demirbas 2005). To avoid these undesired parallel
reactions, esterification can be carried out prior to transesterification. In the ester-
ification reaction, the FFAs react with an alcohol to produce biodiesel and water
according to the reaction shown in the following equation (Eq. 3.1). The water
content can be removed by flash vacuum evaporation after the esterification reac-
tion (Kombe et al. 2013).

RCOOH ðFree Fatty AcidÞ þ R0 OH $ RCOOR0 þ Water ð3:1Þ

Esterification reaction can be non-catalytic or catalytic by using homogeneous


catalysts, heterogeneous catalysts or biocatalysts. Moreover, similar to the trans-
esterification reaction, the efficiency of esterification can also be enhanced using
microwaves, supercritical conditions, ultrasounds or membranes.
Overall, despite the investigations concerning new technologies for biodiesel
production, their application in industrial biodiesel production plants is still very
limited. At the industrial level, biodiesel is generally produced by homogeneous
base-catalysed transesterification and using methanol as alcohol. Methanol is pre-
ferred over the other alcohols because it has a smaller molecule, which favours the
reaction rate, and because it is more economic. The layout of a typical industrial
facility to produce biodiesel through homogeneous base-catalysed transesterifica-
tion is shown in Fig. 3.2.

Fig. 3.2 Layout of a typical industrial facility to produce biodiesel


3 Biodiesel Production Systems: Operation … 39

In biodiesel production, two alkali catalysts are mostly used, NaOH or KOH.
The former is cheaper and easier to find for household production of biodiesel,
whereas the latter generates a valuable effluent that can be used as agricultural
fertilizer (Caetano et al. 2012). Mixing NaOH with methanol (CH3OH) results in
sodium methoxide (CH3NaO); while mixing KOH with methanol results in
potassium methoxide (CH3OK). Instead of methanol, sometimes and less often,
ethanol is used (Caetano et al. 2017). Figure 3.3 shows a reactor for production of
methoxide species.

3.3.3 Reactions and Biodiesel Production

To produce sodium or potassium methoxide (i.e. CH3NaO or CH3KO, respec-


tively), NaOH or KOH should be slowly poured into methanol while the mixture is
stirred by a mechanical stirrer to ensure that the alkali is completely dissolved.
Then, the sodium methoxide or the potassium methoxide must rest for about 8 h.
Both methoxide species are corrosive and flammable. Thus, the electric motor of
the mechanical stirrer must be insulated so that it will not generate sparks, as most
explosions would eventually originate here. Both methoxide species must be

Fig. 3.3 Reactor for


production of methoxide
species
40 N. S. Caetano et al.

produced in airy places, away from ignition sources while the operator must wear
suitable clothing, mask and safety glasses; and the tanks for the preparation of either
of the methoxide species must be made of stainless steel or of high-density
polyethylene.
To convert 1 m3 of raw oil into biodiesel, it is common to use from 0.120 up to
0.230 m3 of methanol, using the upper limit is preferable to ensure a complete
transesterification. Titration is a test performed to determine the pH of the oil to be
converted into biodiesel; using which, it is possible to estimate the amount of alkali
(NaOH or KOH) to be mixed with methanol. In the case of raw oil, titration is
usually not necessary because the acidity of raw oil is predictable, i.e. it can safely
be assumed that with 3.5 kg of alkali per 1 m3 of raw oil, good quality biodiesel
will be produced. However, titration is required in the case of waste cooking oils,
and the main laboratorial means needed are as follows:
• Isopropyl alcohol 99% (v/v);
• Aqueous solution of alkali, NaOH or KOH, with 1 kg of alkali per 1 m3 of
water;
• Phenolphthalein, strips or electrodes to measure pH;
• Laboratory scale with high precision (0.1 g);
• Pipette or syringe with 2 ml volume;
• Two pipettes or syringes with 20 ml volume;
• Glass beaker with 100 ml;
• Thermometer;
• Waste cooking oil.
The procedure to perform titration with NaOH is as follows:
1. Add 2 ml of waste cooking oil to 20 ml of isopropyl alcohol;
2. Add 2 drops of phenolphthalein to the mixture prepared in the previous step;
3. Stir the mixture produced in step 2 till a transparent or translucent liquid is
obtained;
4. Add an aqueous solution of NaOH (1 kg of NaOH per 1000 m3 of water)
dropwise, and simultaneously stir the mixture until it turns pink in colour and
remains as such over 15 s, or till its pH reaches 8.5;
5. Determine the volume of the aqueous NaOH solution used in step 4;
6. Divide the volume determined in step 5 by 2;
7. Add 3.5 to the value determined in step 6;
8. The value determined in step 7 is the weight (expressed in kg) of NaOH which
has to be used per 1 m3 of waste cooking oils to be converted into biodiesel.
It is recommended to filtre the raw oil upstream the reactor; the strainer should
be metallic, preferably of stainless steel, with a mesh of 100 lm. The inflow of raw
oil in the reactor should have a FFAs content of below 0.5% (w/w) and water
content of less than 0.5% (w/w) (Van Gerpen 2005). If raw oil contains water,
saponification will occur during the production of biodiesel in the reactor. To check
if the raw oil contains water, it is usual to take a sample of such raw oil
3 Biodiesel Production Systems: Operation … 41

(about 0.5 dm3) and heat up this sample to 60 °C—if there is water in the sample,
there will be crackling sounds. To extract this water, it is necessary to heat up the
raw oil to around 105 °C and keep it at this temperature until the aforesaid crackling
sounds stop.
To produce biodiesel in the reactor, the temperature of the raw oil at the inlet of
the reactor must be around 54 °C; and when the reactor is full of raw oil, methoxide
should be added, while the mixture is being stirred continuously. The conversion of
raw oil into biodiesel can take from 15 to 60 min, depending on the stirring of the
mixture—less time needed for better agitation. It is possible to stir the mixture with
a centrifugal pump instead of using a mechanical stirrer, in which case, the pump
extracts the mixture from the reactor to reintroduce it into the same reactor.
There are reactors with one or two vessels. In reactors with one vessel, all the
processes involved in the conversion of raw oil into biodiesel, i.e. heating, trans-
esterification, phase separation and the extraction of glycerin, as well as biodiesel
drying, must occur within just one vessel (Fig. 3.4) and only the production of
methoxide is performed within another reactor (Fig. 3.3). Systems with two vessels
consist of one vessel used for the reaction while the second vessel is used for
methanol evaporation.
The vessels should have conical bottoms with an angle of 45°. The outflow of
gases from the reactor must be done through ducts. The heating of the reactor
content can be achieved with using heating coils with the heat power required
standing at about 15 kW/m3 of the content of the reactor.
The thorough conversion of raw oil into biodiesel is followed by a stage of
sedimentation that lasts between 4 and 8 h. During sedimentation, the temperature
of biodiesel contained within the reactor should not fall below 40 °C to prevent the
solidification of glycerin. After sedimentation, follows the drainage of glycerin
which is bright red in colour and the drain valve has to be closed as soon as
biodiesel which is amber yellow in colour enters the drainage duct. During the time
elapsing between the end of the outflow of glycerin and the beginning of the
outflow of biodiesel through the drainage duct, it is usual to have the outflow of
soap flakes—the amount of these flakes increases with higher water contents either
in raw oil or in the methoxide used. The biodiesel that unduly leaves the reactor
during the drainage of glycerin can be recovered from the drain tank because
biodiesel floats over glycerin. The glycerin resulting from the biodiesel production
contains impurities, namely traces of catalyst, methanol, etc. but after purification,
this glycerin can be used as lubricant, fuel, or even as raw material for the cosmetic
and pharmaceutical industries (Garlapati et al. 2016). More specifically, glycerine
can be used in soap, food, paint, toothpaste, pharmaceutical and cosmetics indus-
tries (Garlapati et al. 2016). A study revealed the feasibility of using glycerine as an
ingredient in poultry feed (Jung and Batal 2011), concluding that even in the
presence of excess methanol, there is no negative impact on broiler performance.
Overall, if produced in a sustainable way, biodiesel can help countries dependent on
petroleum with enhancing energy security and the overall livelihood of their nations
including farmers (Stoetaert and Vandamme 2009; Datta and Mandal 2014;
Ben-Iwo et al. 2016).
42 N. S. Caetano et al.

V1 - inlet for raw oil;


V2, V3, V6, V7, V8, V9, V11 - process valves;
Recircula on of washing through V2, B, Rotameters 1 and 2, V9 and sprinkler;
V4 - valve for glycerin outlet;
V5 – valve for methoxide inlet;
V10 – valve for methanol inlet to the condenser;
V12 – pressuriza on valve;
V13 - inlet valve for washing water;
V14 - inlet valve for phosphoric acid (H3PO4) to control the pH of the washing water;
SC - sensor control;
B - pump;
PI - pressure indicator.

Fig. 3.4 A single vessel reactor for biodiesel production


3 Biodiesel Production Systems: Operation … 43

Fig. 3.5 Reactor used for


washing of biodiesel with
bubbling air

The purpose of washing is to remove impurities from biodiesel; usually, three


rounds of washing are performed. The most common washing methods are (i) with
water shower and (ii) with bubbling air. The first washing method requires water
with pH  7 (but pH should never exceeds 7) and phosphoric acid (H3PO4), citric
acid (C6H8O7), and antifoam. The water should have a low salinity and be free of
chlorine and limestone (distilled and demineralized water is preferred). There is the
possibility of using acetic acid (CH3COOH) instead of phosphoric acid, but the
latter is preferred because the volume of acetic acid needed is roughly three times
the volume of phosphoric acid needed. While if acetic acid is used, the microbial
content of biodiesel will be lower which will allow higher lifespans of biodiesel
during storage. If phosphoric acid is used, the effluents (water enriched with
phosphate chemical species) can be used as farm irrigation water. Citric acid is an
antioxidant used to prevent oxidation of biodiesel caused by the washing process.
The antifoam is only necessary in the case of excessive production of foam—it is
recommended at 0.5 kg of antifoam per 1 m3 of water. In this washing method, the
water is circulated by a centrifugal pump, between the conical bottom and the top of
the reactor from where it is sprinkled over the biodiesel and as the water is denser
than biodiesel, it accumulates in the bottom of the reactor and then follows, moved
by the pump, to the top of the reactor. During washing, the flow rate of water should
be around 0.07 m3/m3 of biodiesel/min; the temperature of the water must be in the
range of 35–45 °C.
It is important to control the pH of the washing water, which is done through a
sensor located at the bottom of the reactor (Fig. 3.4). This sensor is connected with
44 N. S. Caetano et al.

a dosing pump which will be activated if necessary, pump an aqueous solution of


phosphoric acid (H3PO4) into the water accumulated in the bottom of the reactor in
order to neutralize it (pH = 7). As mentioned earlier, a full washing process gen-
erally encompasses three rounds of washing. After each of these washing rounds,
the water in the bottom of the reactor must be decanted during at least 2 h and will
eventually be removed by the drainage tube. The water used in the first washing is
discarded, but the water used either in the second or the third rounds of washing can
be reused, respectively, for the first and the second washing rounds of the next
batch of biodiesel. In fact, in the first washing step, the aim is to neutralize the alkali
(NaOH or KOH) that was used to make methoxide; this washing lasts between 4
and 12 h. The second washing is done with water, so it is needed to set the pH of
the water used at 7 and this round also lasts from 4 to 12 h. The third round of
washing is similar to the second one lasting from 4 to 12 h as well.
The second washing method (Fig. 3.5), uses a multi-perforated plate, drilled
with orifices with diameters less than 1.5 mm, to diffuse air into biodiesel. The air
flow is generated with a typical air-compression facility; the flow rate of air blown
into the reactor should be around 0.07 m3 air/1 m3 biodiesel/min, and this flow
must be maintained for at least 30 min. This method is prone to cause formation of
foam and oxidation of biodiesel, which is averted with, respectively, antifoam and
citric acid. The noise caused by the compressor is the most significant nuisance of
this washing method.
The final steps in the production of biodiesel include dewatering, filtration and
storage.
Dewatering is achieved by evaporation, decantation or centrifugation. In the case
of evaporation, the reactor is either at atmospheric pressure or below atmospheric
pressure; vacuum is created by a vacuum pump. The lower the pressure in the
reactor is, the lower the evaporation pressure of water will be. In general, the
evaporation lasts for about 15 min. Decantation is less energy-intensive compared
with evaporation, but it is slower and must occur in a reactor with a conical bottom
for over 3 weeks, after which the water should be drained. Centrifugation is fast but
centrifuges are expensive and use energy intensively.
Filtration of biodiesel must be done with a metallic filter with a mesh below
5 lm. Storage tanks for biodiesel should be made of stainless steel or aluminium
alloys. It should be noted that copper, copper alloys, zinc and plumb are chemically
attacked by biodiesel. Knowledge about the behaviour of biodiesel concerning its
chemical stability is yet to further advanced.

3.4 Biodiesel Purification Processes

The mixture of biodiesel and glycerol is usually separated by gravity (decantation)


or centrifugal separation and the excess alcohol is distributed among biodiesel and
glycerol phases during separation. Since alcohol is more soluble in polar phases,
most of the unreacted alcohol will be removed with the glycerol phase. Therefore,
3 Biodiesel Production Systems: Operation … 45

after separation, biodiesel contains 3–4% of alcohol while glycerol contains around
15% of alcohol. One of the most used purification processes in biodiesel production
is the alcohol removal from biodiesel and glycerol (Atadashi 2015). The surplus of
alcohol used in the reaction can be recovered in order to lower the production costs
(the recovered methanol can be reused). Moreover, the emissions of alcohols into
the atmosphere (alcohols are toxic and flammable) can be prevented. On the other
hand, this process is vital in order to comply with biodiesel standards such as EN
14214 and ASTM D6751, in terms of alcohol content (Tabatabaei et al. 2015). The
allowable alcohol level in biodiesel, according to EN 14214 standard, is 0.2%, since
the alcohol content has a great influence on biodiesel flashpoint. Alcohol recovery
usually occurs after biodiesel and glycerol separation; otherwise, the quality of
biodiesel would decrease owing to chemical degradation.
Alcohol removal from biodiesel and glycerol can be performed using vacuum
flash evaporation, distillation, or water washing (Atadashi et al. 2011a; Amelio
et al. 2016). The most used process for alcohol removal is vacuum flash evaporation
as this method allows achieving more favourable results with lower costs compared
with the other methods (Stojković et al. 2014). A schematic overview of alcohol
recovery by vacuum flash evaporation can be observed in Fig. 3.6. Vacuum flash
evaporation could be enhanced by increasing the temperature and decreasing the
vacuum pressure. In industrial plants (working in continuous or batch modes)
designed to work at high temperature and high vacuum levels, such as 110 °C and
250 mbar, respectively, the methanol content of crude biodiesel, can be reduced to
0.1–0.2%, using a single step vacuum flash evaporation. However, it is important to
take into consideration that high temperatures have some disadvantages such as the
evaporation of water with methanol, biodiesel chemical degradation and foaming
formation. For these reasons, it is important to operate vacuum flash evaporation
under proper temperatures, depending on crude biodiesel properties.
Distillation can be performed in a single column, but this has great disadvantages
such as complex design, control and operation and is more expensive than the other
methods (Dunn 2011). Water washing for alcohol removal requires cheap equip-
ment; however, the removed methanol drags high amounts of water and the sub-
sequent methanol drying results in high costs (Demirbas 2007).
After alcohol removal, glycerol can be commercialized as technical grade
glycerol, being reused as lubricant or fuel. Also, it can be submitted to additional
treatment (acidulation to free fatty acids, followed by purification in an adsorption
column) and finally, it can be commercialized as pharmaceutical grade glycerol, in
which case, it may be used as raw material for cosmetic and pharmaceutical
industries, among others (Demirbas 2009).
After the production process and alcohol removal, biodiesel still contains some
impurities that must be removed by purification processes, in order to fulfill the
requirements set forth by biodiesel standards such as EN14214 and ASTM D6751
(Amelio et al. 2016). Those impurities, such as free glycerol, FFAs, catalyst, water,
glycerides, salts and soaps, reduce the quality of the produced biodiesel and could
consequently reduce the engine performance and cause several damages such as
corrosion, fouling, plugging and weakening in the engine system. Table 3.8 shows
46 N. S. Caetano et al.

Fig. 3.6 Recovery of methanol by vacuum flash evaporation

the effects of different types of impurities on biodiesel and engines systems (Berrios
and Skelton 2008; Stojković et al. 2014). To minimize or avoid the unfavourable
effects of impurities, standards such as EN14214 and ASTM D6751 establish the
quality requirements for biodiesel. In order to fulfill the purity requirements of these
standards, the impurities present in crude biodiesel must be removed by purification
technologies. The costs associated with biodiesel purification represent 60–80% of
total processing cost (Atadashi et al. 2011c). In an attempt to reduce costs related to
biodiesel purification, several purification technologies have been studied and used.
These technologies are generally divided as conventional or novel technologies.
The conventional technologies include wet and dry washing while the novel
technologies include membrane extraction (Veljković et al. 2015).
Biodiesel wet washing can be performed using: water washing; acid washing
followed by water washing; or organic solvents washing followed by water
washing. Water washing has been traditionally used, and involves high quantities of
water. Acids such as phosphoric, sulfuric or hydrochloric acid followed by the
addition of distilled water can be used to remove impurities from biodiesel. The
acids are added to neutralize the catalyst and decompose the soaps. Wet washing
using organic solvents can be performed using petroleum ether or n-hexane, which
is usually followed by demineralized water addition (Atadashi et al. 2011b, c). All
technologies used in wet washing are characterised by high water consumption,
3 Biodiesel Production Systems: Operation … 47

Table 3.8 Effects of impurities on biodiesel and engines systems (Berrios and Skelton 2008;
Atadashi et al. 2011c)
Impurity Effect
Water – Parallel reactions (FFA formation through hydrolysis)
– Corrosion
– Bacteriological growth (plugging)
– Reduces heat combustion
Glycerol – Settling and fouling problems
Soaps and catalyst – Deposits and plugging in injectors
– Corrosion
– Filter blockage
– Engine weakening
Alcohol – Low density and viscosity
– Low flash point (problems with storage, transportation and utilization)
– Corrosion
– Deterioration of rubber seals and gaskets
Glycerides – High viscosity
– Deposits and plugging in pistons, valves and injectors
– Crystallization
Free fatty acids – Corrosion
– Low oxidation stability

high costs related to wastewater treatment, and high energy and equipment costs,
associated with the required residual water evaporation after washing (Atadashi
et al. 2011a). Some studies show that the amount of the contaminated wastewater
produced is 20–30% (v/v) of the purified biodiesel (Atadashi et al. 2011c).
Environmental and economical disadvantages associated with wet washing have
led to the development of dry washing technologies. Biodiesel dry washing is
performed using (acidic and basic) adsorbents, which have strong affinity for polar
substances such as biodiesel impurities. This technology can use several types of
adsorbents such as silicates, ion exchange resins, cellulosic activated carbon,
activated fibre and activated clay, among others (Atadashi et al. 2011b; Atadashi
2015). Dry washing technologies save water and energy, and so, are more
favourable from the environmental perspective than wet washing ones. Moreover,
several investigations have shown that dry washing technologies have led to higher
purity biodiesel in comparison with wet washing technologies (Atadashi et al.
2011a).
The development of novel technologies to avoid the disadvantages of conven-
tional technologies and to improve biodiesel purity even further have resulted in the
introduction of membrane technology for biodiesel purification. This technology
applies membranes, which are considered as the most effective ways for biodiesel
purification and present economic and environmental advantages over the con-
ventional technologies (Atadashi et al. 2011a, b). Membranes are semipermeable
barriers that separate different species in solution while allowing restricted passage
of some components. They are selective either by pore size (porous membrane)
48 N. S. Caetano et al.

or because of their chemical affinity for permeating specific components.


Membranes used in this technology can be organic or ceramic. The organic
membranes can be applied for ultra-filtration, micro-filtration and reverse osmosis,
and are generally composed of polyamide, polysulphone, polycarbonate, or other
advanced polymers, which have chemical stability and better resistance to microbial
degradation. The ceramic or inorganic membranes have also attracted a great deal
of attention because of their superiority in terms of thermal, chemical and
mechanical stability, high porosity, high flux, long lifetime, resistance to microbial
degradation, increased resistance to fouling, and a narrower pore size distribution,
in comparison with organic membranes (Atadashi et al. 2011c).
Additionally and following all the above-mentioned purification steps, a final
vacuum distillation step can be performed to increase the purity of the produced
biodiesel by removing mono, di and triglycerides, as well as sulphur and other
heavy species in biodiesel. This final purification step is usually performed with
wiped film evaporators or thin film evaporators, in which distilled biodiesel reaches
a very high purity, typically >99.5%. The vacuum distillation step involves high
boiling points, high viscosity and heat sensitivity to separate the heavier compo-
nents from biodiesel (lighter phase).

3.5 Quality Control of Biodiesel

As stated before, biodiesel quality must be assured in order to prevent damages in


engines, as well as to prevent excessive production cost and environmental and
health hazards. Different standards have been adopted in Europe (EN 14214) and in
the USA (ASTM D6751), in Brazil, and many other countries. Table 3.9 lists usual
laboratorial tests, respective standards and test methods set forth for biodiesel
characterization in Europe.

3.6 Troubleshooting in Biodiesel Production

Troubleshooting is defined as a problem-solving methodology, which is used to


identify and solve a problem as fast and easy as possible and can be applied in case
of equipment/machines, systems or processes. The ultimate goal of troubleshooting
is to get the equipment/machine, system or process back into operation.
Troubleshooting is generally carried out by following a set of steps that can vary
depending on an application and its complexity. For example, the problem-solving
methodologies applicable to software development processes are different from the
methodologies applicable to industrial processes. Regardless of the application and
its complexity, there are a few main guidelines that are generally applied for
troubleshooting. Following these guidelines, the most common methodology for
3 Biodiesel Production Systems: Operation … 49

Table 3.9 Laboratorial tests, respective Standards and test methods for biodiesel in Europe (EN
14214)
Parameter Standard Method
Ester content EN 14103 Gas chromatography
Density at 15 °C EN ISO 12185 Oscillating U-tube
Viscosity at 40 °C EN ISO 3104 Viscometer
Flammability EN ISO 3679 Rapid equilibrium
Sulphur content EN ISO 20846 XRF EDX
Cetane index EN ISO 5165 Cetane engine
Coal residue EN ISO 10370 Micro-coal
Sulfated ash ISO 3987 Burning-gravimetry
Water content EN ISO 12937 Titration Karl Fischer
Total contamination EN 12662 Laboratorial filtration
Copper corrosion EN ISO 2160 Copper strip
Oxidation stability at 110 °C EN 14112 Rancidity meter
Linoleic acid methyl ester EN 14103 Gas chromatography
Polyunsaturated methyl esters EN 14103 Gas chromatography
Methanol content EN 14110 Gas chromatography
Monoglycerides content EN 14105 Gas chromatography
Diglycerides content EN 14105 Gas chromatography
Triglycerides content EN 14105 Gas chromatography
Free glycerin EN 14105 Gas chromatography
Total glycerin EN 14105 Gas chromatography
Sodium content EN 14108 AAS
Potassium content EN 14109 AAS
Calcium content EN 14538 ICP
Magnesium content EN 14538 ICP
Phosphor content EN 14107 ICP
Acidity EN 14104 Titration
Iodine value EN 14111 Titration

troubleshooting is the so-called five-step methodology, which consists of the fol-


lowing steps:
1. Verify that a problem actually exists—symptom recognition;
2. Isolate the cause of the problem;
3. Correct the cause of the problem;
4. Verify that the problem has been corrected;
5. Follow up to prevent future problems.
The general five-step methodology could be adopted differently depending on
the application. For example, in manufacturing industry, troubleshooting can be
performed according to the following methodologies: PDCA (Plan, Do, Check,
Act) cycle, cause and effect diagrams, TOPS/8D (Team Oriented Problem Solving),
50 N. S. Caetano et al.

Sigma Six, Kaizen, Matrix A3, among others. Regardless of the different names, the
application and compliance of these methodologies are very similar. The referred
methodologies can also be applied in chemical industry. However, specifically for
chemical industry, in which biodiesel production industry can be integrated, trou-
bleshooting is commonly based on a hazardous and operability study, known as
HAZOP.
The HAZOP methodology was developed in the 1960s for risk analysis in
chemical process systems (Kletz 1971), but has been extended to several types of
process systems, such as mining operations, nuclear power plant operations, soft-
ware development, etc. The first paper on HAZOP was published in 1974 (Lawley
1974). HAZOP is a structured and systemic examination of a process (planned or
existing) and/or operation, performed to identify and evaluate hazard scenarios that
can cause failures/problems/accidents and represent risks to personnel, process
operation, equipment, or environment (Rausand and Høyland 2004).
HAZOP study focuses on investigating deviations from design specifications
during operation. These deviations can cause failures/problems/accidents. In a
HAZOP study, the process system is divided into nodes, which can be specific
points of the process, equipment, process sections or operating steps. The HAZOP
study is performed by a team, which reviews the process system by checking the
proposed process design and applying appropriate guide words for appropriate
parameters for each defined node. A standard list of guide words is: no, more, less,
as well, part of, reverse, and other than (Baybutt 2015). Examples of parameters
applicable in each node are flow, pressure, temperature, composition, addition,
level, etc. The use of guide words for each parameter and each node provides the
opportunity to identify the following:
• Possible deviations from the design intent;
• Possible consequences of each deviation;
• Safeguards for each deviation;
• Actions or recommendations.
An example of HAZOP study development in which the parameter—FLOW and
the guide word—NO in a pump (node) is applied, is described in Fig. 3.7 and
Table 3.10.
Additionally, the analysis can continue with the risk evaluation through the
combination between frequency and severity in case of failures/problems/accidents.
The HAZOP analysis is usually carried out when detailed process design has been
complete (prior to plant construction) but also can be carried out during operation of
an existing process plant when failures/problems/accidents occur. In fact, the
HAZOP study can be used for troubleshooting purposes because this analysis
allows quick isolation of the cause of failure/problem/accident and the application
of procedures to correct the cause of the problem and/or minimize its consequences.
For that reason, the HAZOP study development must be performed prior to its
application in a problem-solving.
3 Biodiesel Production Systems: Operation … 51

NODE 1

Fig. 3.7 Node identification in a process flow diagram

Table 3.10 HAZOP study for NODE 1, parameter—FLOW and guide word—NO
Guide Parameter Deviation Possible causes Possible Safeguards Actions
word consequences required
NO FLOW NO – No level in tank; – Pump – Instrument – Proper
FLOW – Valve after tank damage; for maintenance
closed wrongly; – Motor over minimum – Low
– Valve after tank heating; tank level pressure
damage; – Motor – Instrument alarm
– Pipe obstruction; damage; for pressure – Low tank
– Pump malfunction transmitting level alarm
before
pump

Once located and repaired, the problem area should be tested to ensure proper
function. The repair process may require technical service, parts replacement or
redemption of a warranty.
Different methodologies may have slightly different names, but the similarities
are obvious.
• Investigation
– Problem Statement: Create a clear, concise statement of the problem.
– Problem Description: Identify the symptoms. What works? What does not?
– Identify Differences and Changes: What has changed recently? What is
unique about this system?
52 N. S. Caetano et al.

• Analysis
– Brainstorm: Gather Hypotheses: What might have caused the problem?
– Identify Likely Causes: Which hypotheses are most likely?
– Test Possible Causes: Schedule the testing for the most likely hypotheses.
Perform any non-disruptive testing immediately.
• Implementation
– Implement the Fix: Complete the repair.
– Verify the Fix: Is the problem really fixed?
– Document the Resolution: What did we do? Get a sign-off from the system
owner.
Overall, it should be noted that troubleshooting starts with collecting technical
records from relevant sources. These include the datasheets, suppliers, contractors,
operators and maintenance departments. Additionally, proper test instruments make
the process smoother and make it possible to more easily identify secondary
problems where they exist.

3.7 Conclusions

There are different biodiesel production systems, already being used at industrial
scale. The choice of production system depends not only on the available feedstock
quality, but also on the size of the production plants. Regardless of the type of
technology implemented (batch, semi-batch or continuous), different processes
might be implemented to produce quality biodiesel. Upon the production of bio-
diesel, it is mandatory that the product undergoes necessary purification steps and
then be analyzed to confirm that its quality specifications are in accordance with the
standards such as EN 14214 or ASTM D6751. On the other hand, the by-product,
i.e. glycerol must also be purified in order to make the whole process more eco-
nomically sustainable. Different purification technologies are available, both for
biodiesel and for glycerol and the right choice of technology is generally based on
the required products specifications, as well as on the feedstock characteristics.
Finally, regardless the technology used for biodiesel production, the risk of hazards
in this industry is high (e.g. fire, explosion, poisoning by methanol, etc.), and
therefore, a HAZOP study should be prepared and implemented throughout the
project.

Acknowledgements This work was partially supported by the project POCI-01-0145-


FEDER-006939 (Laboratory for Process Engineering, Environment, Biotechnology and Energy
—UID/EQU/00511/2013) funded by the European Regional Development Fund (ERDF), through
COMPETE2020—Programa Operacional Competitividade e Internacionalização (POCI) and by
the national funds, through FCT—Fundação para a Ciência e a Tecnologia.
3 Biodiesel Production Systems: Operation … 53

References

ACEA (2014) ACEA position concerning diesel that might contain more than 7% FAME. www.
acea.be/uploads/publications/140422_ACEA_position_on_B7_diesel.pdf. Accessed 28 Mar 2017
Amelio A, Loise L, Azhandeh R, Darvishmanesh S, Calabró V, Degrève J, Luis P, Van der
Bruggen B (2016) Purification of biodiesel using a membrane contactor: liquid–liquid
extraction. Fuel Process Technol 142:352–360
ASTM D6751 (2008) Standard specification for biodiesel fuel blend stock (B100) for middle
distillate fuels
ASTM D975 (2008) Standard specification for diesel fuel oils
Atadashi IM (2015) Purification of crude biodiesel using dry washing and membrane technologies.
Alexandria Eng J 54(4):1265–1272
Atadashi IM, Aroua MK Abdul, Aziz AR, Sulaiman NMN (2011a) Refining technologies for the
purification of crude biodiesel. Appl Energy 88(12):4239–4251
Atadashi IM, Aroua MK, Abdul Aziz A (2011b) Biodiesel separation and purification: a review.
Renew Energy 36(2):437–443
Atadashi IM, Aroua MK, Abdul Aziz AR, Sulaiman NMN (2011c) Membrane biodiesel
production and refining technology: a critical review. Renew Sustain Energy Rev 15(9):5051–
5062
Bakeas E, Karavalakis G, Stournas S (2011) Biodiesel emissions profile in modern diesel vehicles.
Part 1: Effect of biodiesel origin on the criteria emissions. Sci Total Environ 409(9):1670–1676
Baroutian S, Aroua MK, Raman AAA, Sulaiman NMN (2011) A packed bed membrane reactor
for production of biodiesel using activated carbon supported catalyst. Biores Technol 102
(2):1095–1102
Baybutt P (2015) A critique of the hazard and operability (HAZOP) study. J Loss Prev Process Ind
33:52–58
Ben-Iwo J, Manovic V, Longhurst P (2016) Biomass resources and biofuels potential for the
production of transportation fuels in Nigeria. Renew Sustain Energy Rev 63:172–192
Berrios M, Skelton RL (2008) Comparison of purification methods for biodiesel. Chem Eng J 144
(3):459–465
Bhardwaj M, Gupta P, Kumar N (2014) Compatibility of metals and elastomers in biodiesel: a
review. Int J Res 1(7):376–391
Biodiesel Emissions. National Biodiesel Board (nd). https://www.me.utexas.edu/*challengex/
family/emissions.pdf. Accessed 31 July 2018
Biodiesel Fact Sheet (nd) Biodiesel Co-ops. http://greenwayrides.com/site/assets/pdfs/Biodiesel-
fact-sheet.pdf. Accessed 31 July 2018
Caetano NS, Caldeira D, Martins AA, Mata TM (2017) Valorisation of spent coffee grounds:
production of biodiesel via enzymatic catalysis with ethanol and a co-solvent. Waste Biomass
Valoriz 8(6):1981–1994
Caetano NS, Silva VFM, Mata TM (2012) Valorization of coffee grounds for biodiesel production.
Chem Eng Trans 26:267–272
Cheng L-H, Yen S-Y, Su L-S, Chen J (2010) Study on membrane reactors for biodiesel production
by phase behaviors of canola oil methanolysis in batch reactors. Biores Technol 101(17):6663–
6668
Datta A, Mandal BK (2014) Use of Jatropha biodiesel as a future sustainable fuel. Energy Technol
Policy 1:8–14
Demirbas A (2005) Biodiesel production from vegetable oils via catalytic and non-catalytic
supercritical methanol transesterification methods. Prog Energy Combust Sci 31(5–6):466–487
Demirbas A (2007) Biofuels. Springer. ISBN 978-1-84628-995-8
Demirbas A (2008a) Relationships derived from physical properties of vegetable oil and biodiesel
fuels. Fuel 87(8–9):1743–1748
Demirbas A (2008b) Biodiesel: a realistic fuel for alternative diesel engines. Springer, London
54 N. S. Caetano et al.

Demirbas A (2008c) Biofuels sources, biofuel policy, biofuel economy and global biofuel
projections. Energy Convers Manag 49(8):2106–2116
Demirbas A (2009) Progress and recent trends in biodiesel fuels. Energy Convers Manag 50(1):
14–34
DIN 51606 (1997) Liquid fuels—diesel fuel of fatty acid methyl ester (FAME)—specifications.
German Institute for Standardization
Dincer K (2008) Lower emissions from biodiesel combustion. Energy Sources Part A: Recovery
Util Environ Eff 30(10):963–968
Dubé MA, Tremblay AY, Liu J (2007) Biodiesel production using a membrane reactor. Biores
Technol 98:639–647
Dunn RO (2011) Improving the cold flow properties of biodiesel by fractionation. INTECH Open
Access Publisher
Dupont (2017) https://dupont.secure.force.com/CRG_TlargiGuide. Accessed 15 Apr 2017
Ejikeme PM, Anyaogu ID, Ejikeme CL, Nwafor NP, Egbuonu CAC, Ukogu K, Ibemesi JA (2010)
Catalysis in biodiesel production by transesterification processes—an insight. E-J Chem 7
(4):1120–1132
EN 14214 (2012) Liquid petroleum products—fatty acid methyl esters (FAME) for use in diesel
engines and heating applications—requirements and test methods. European Committee for
Standardization
EN 590 (2009) Automotive fuels—diesel—requirements and test methods. European Committee
for Standardization
European Biodiesel Board (2017) Statistics—the EU biodiesel industry. http://www.ebb-eu.org/
stats.php. Accessed 9 Jan 2017
Garlapati VK, Shankar U, Budhiraja A (2016) Bioconversion technologies of crude glycerol to
value added industrial products. Biotechnol Rep 9:9–14
Ghaly AE, Dave D, Brooks MS, Budge S (2010) Production of biodiesel by enzymatic
transesterification: a review. Am J Biochem Biotechnol 6(2):54–76
Gude VG, Patil P, Martinez-Guerra E, Deng S, Nirmalakhandan N (2013) Microwave energy
potential for biodiesel production. Sustain Chem Process 1:5
Hajjari M, Tabatabaei M, Aghbashlo M, Ghanavati H (2017) A review on the prospects of
sustainable biodiesel production: a global scenario with an emphasis on waste-oil biodiesel
utilization. Renew Sustain Energy Rev 72:445–464
Haseeb A, Fazal M, Jahirul M, Masjuki H (2011) Compatibility of automotive materials in
biodiesel: a review. Fuel 90(3):922–931
Jeschke B (2009) Plant oil biofuel: rationale, production and application. In: Soetaert W,
Vandamme EJ (eds) Biofuels. Wiley, Chichester, UK
Jung B, Batal AB (2011) Nutritional and feeding value of crude glycerin for poultry. 2. Evaluation
of feeding crude glycerin to broilers. J Appl Poult Res 20:514–527
Karmakar A, Karmakar S, Mukherjee S (2010) Properties of various plants and animals feedstocks
for biodiesel production. Biores Technol 101(19):7201–7210
Kletz TA (1971) Hazard analysis—a quantitative approach to safety. In: Major loss prevention in
the process industries, Symposium series No. 34. Institution of Chemical Engineers, Rugby,
UK, pp 75–81
Knothe G, Dunn RO, Bagby MO (1997) Biodiesel: the use of vegetable oils and their derivatives
as alternative diesel fuels. ACS Symp Ser 666:172–208
Kombe GG, Temu AK, Rajabu HM, Mrema GD, Kansedo J, Lee KT (2013) Pre-treatment of high
free fatty acids oils by chemical re-esterification for biodiesel production—a review. Adv
Chem Eng Sci 3:242–247
Kwon EE, Yi H, Jeon YJ (2013) Transforming rapeseed oil into fatty acid ethyl ester (FAEE) via
the noncatalytic transesterification reaction. AIChE J 59:1468–1471
Lapuerta M, Armas O, Rodríguez-Fernández J (2008) Effect of biodiesel fuels on diesel engine
emissions. Prog Energy Combust Sci 34(2):198–223
Lawley HG (1974) Operability studies and hazard analysis. Chem Eng Prog 70(4):45–56
3 Biodiesel Production Systems: Operation … 55

Liu CH, Huang CC, Wang Y-W, Lee D-J, Chang J-S (2012) Biodiesel production by enzymatic
transesterification catalyzed by Burkholderia lipase immobilized on hydrophobic magnetic
particles. Appl Energy 100:41–46
Mahmudul HM, Hagos FY, Mamat R, Adam AA, Ishak WFW, Alenezi R (2017) Production,
characterization and performance of biodiesel as an alternative fuel in diesel engines—a
review. Renew Sustain Energy Rev 72:497–509
Mata T, Mendes A, Caetano N, Martins AA (2014) Properties and sustainability of biodiesel from
animal fats and fish oil. Chem Eng Trans 38:175–180
Mata TM, Martins AA, Caetano NS (2010) Microalgae for biodiesel production and other
applications: a review. Renew Sustain Energy Rev 14(1):217–232
Mesiano AJ, Beckman EJ, Russell AJ (1999) Supercritical biocatalysis. Chem Rev 99(2):623–634
Miri SM, Mousavi Seyedim SR, Ghobadian B (2017) Effects of biodiesel fuel synthesized from
non-edible rapeseed oil on performance and emission variables of diesel engines. J Clean Prod
142(4):3798–3808
Mittelbach M (2009) Process technologies for biodiesel production. In: Soetaert W, Vandamme EJ
(eds) Biofuels. Wiley, Chichester, UK
NBB (National Biodiesel Board) (2018). Emissions calculator. http://biodiesel.org/using-biodiesel/
handling-use/emissions-calculator. Accessed 31 July 2018
Pulz O (2007) Evaluation of GreenFuel’s 3D matrix algae growth engineering scale unit.
Performance summary report. IGV Institut für Getreideverarbeitung GmbH, Germany
Pushparaj T, Ramabalan S (2013) Green fuel design for diesel engine, combustion, performance
and emission analysis. Procedia Eng 64:701–709
Putrasari Y, Lim O (2017) A study on combustion and emission of GCI engines fueled with
gasoline-biodiesel blends. Fuel 189:141–154
Rajagopal D, Zilberman D (2007) Review of environmental, economic and policy aspects of
biofuels. Policy Research Working Paper; No. 4341. World Bank, Washington, DC. © World
Bank. https://openknowledge.worldbank.org/handle/10986/7337. Accessed 20 Mar 2017
Rausand M, Høyland A (2004) System reliability theory: models, statistical methods, and
applications, 2nd edn. Wiley, Hoboken, New Jersey
Show PL, Tang MSY, Nagarajan D, Ling TC, Ooi CW, Chang JS (2017) A holistic approach to
managing microalgae for biofuel applications. Int J Mol Sci 18(1):215–248
Soetaert W, Vandamme EJ (2009) Biofuels in perspective. In: Soetaert W, Vandamme EJ
(eds) Biofuels. Wiley, Chichester, UK
Stojković IJ, Stamenković OS, Povrenović DS, Veljković VB (2014) Purification technologies for
crude biodiesel obtained by alkali-catalyzed transesterification. Renew Sustain Energy Rev
32:1–15
Tabatabaei M, Karimi K, Horváth IS, Kumar R (2015) Recent trends in biodiesel production.
Biofuel Res J 7:258–267
Tickell J, Roman K (ed), Tickell K (1999) From the fryer to the fuel tank: the complete guide to
using vegetable oil as an alternative fuel. Green Teach Publishing, Florida
Tran D-T, Chang J-S, Lee D-J (2017) Recent insights into continuous-flow biodiesel production
via catalytic and non-catalytic transesterification processes. Appl Energy 185(1):376–409
Upham P, Thornley P, Tomei J, Boucher P (2009) Substitutable biodiesel feedstocks for the UK: a
review of sustainability issues with reference to the UK RTFO. J Clean Prod 17(S1):S37–S45
Van Gerpen J (2005) Biodiesel processing and production. Fuel Proc Technol 86(10):1097–1107
Veljković VB, Banković-Ilić IB, Stamenković OS (2015) Purification of crude biodiesel obtained
by heterogeneously-catalyzed transesterification. Renew Sustain Energy Rev 49:500–516
Vijayaraj K, Sathiyagnanam AP (2016) Experimental investigation of a diesel engine with methyl
ester of mango seed oil and diesel blends. Alexandria Eng J 55(1):215–221
Volkswagen (2010) Biodiesel statement. www.volkswagen.co.uk/assets/common/pdf/general/
biodiesel.pdf. Accessed 28 Mar 2017
Vyas AP, Verma JL, Subrahmanyam N (2010) A review on FAME production processes. Fuel
89(1):1–9
56 N. S. Caetano et al.

Wang W, Lyons D, Clark N, Gautam M, Norton P (2000) Emissions from nine heavy trucks
fueled by diesel and biodiesel blend without engine modification. Environ Sci Technol 34(6):
933–939
Xu W, Lijing G, Guomin X (2015) Biodiesel production optimization using monolithic catalyst in
a fixed-bed membrane reactor. Fuel 159:484–490
Xue J, Grift T, Hansen A (2011) Effect of biodiesel on engine performances and emissions. Renew
Sustain Energy Rev 15(2):1098–1116
Zittelli GC, Rodolfi L, Bassi N, Biondi N, Tredici MR (2013) Photobioreactors for microalgal
biofuel production. Algae for biofuels and energy. Springer, Berlin, pp 115–131
Chapter 4
Biodiesel Purification and Upgrading
Technologies

Hamed Bateni, Alireza Saraeian, Chad Able and Keikhosro Karimi

Abstract Biodiesel purification is a crucial process in meeting fuel grade standard


specifications. Inadequate purification results in a low-quality fuel and hampers
engine performance. Conventional wet and dry washing along with membrane
refining technologies are the most discussed methods in the literature for biodiesel
purification. The conventional wet washing is performed using organic solvents,
deionized water, or an acid solution. However, these methods result in a large
quantity of wastewater, which creates a significant cost for wastewater treatment
besides environmental impacts. Dry washing techniques were introduced to address
this deficiency. In these methods, an appropriate adsorbent media such as
Magnesol, an ion exchange resin, or active carbon is utilized to remove the
impurities. The challenges associated with wet and dry techniques motivated sci-
entists to seek more innovative techniques. Regarding biodiesel purification,
organic and ceramic membrane technologies have received increasing attention.
Biodiesel upgrading is an alternative route for producing diesel-like fuels with
properties that can exceed conventional petroleum diesel. Since oxygenated moi-
eties of biodiesel are the main reason for its poor fuel properties (e.g., high vis-
cosity, low energy density, low chemical stability, and poor cold flow behavior),
most of these upgrading techniques have focused on deoxygenation pathways.
Hydrodeoxygenation using various catalytic systems have also been studied
extensively to produce renewable diesel with high yields and high carbon

H. Bateni (&)  A. Saraeian


Department of Chemical and Biological Engineering, Iowa State University, Ames,
IA 50011, USA
e-mail: hbateni@iastate.edu
C. Able
Department of Chemical and Biomolecular Engineering, Ohio University,
Athens, OH 45701, USA
K. Karimi
Department of Chemical Engineering, Isfahan University of Technology,
Isfahan 84156-83111, Iran

© Springer Nature Switzerland AG 2019 57


M. Tabatabaei and M. Aghbashlo (eds.), Biodiesel,
Biofuel and Biorefinery Technologies 8,
https://doi.org/10.1007/978-3-030-00985-4_4
58 H. Bateni et al.

efficiency. This chapter presents the basics and applied aspects of biodiesel
purification and upgrading along with an overview on different techniques, chal-
lenges, and the overall trend of research.

4.1 Biodiesel Separation and Purification

After transesterification reactions, the biodiesel layer needs to be separated from the
glycerol and then subjected to further purification (Bateni and Karimi 2016a, b).
Uncomplicated and easy to handle separation techniques, e.g., gravitational settling
and centrifugation, can be used due to a sufficient difference in the density of
biodiesel (about 880 kg/m3) and glycerol (1050 kg/m3) along with a low mutual
solubility of these compounds (Van Gerpen et al. 2004; Atadashi et al. 2011a). The
biodiesel layer is typically contaminated with traces of catalyst, glycerol, oil and its
impurities, and alcohol (Bateni et al. 2014). The presence of impurities affects the
properties of biodiesel and its engine performance as reported in Table 4.1.
Inadequate purification leads to a low-quality fuel, resulting in severe engine
problems including plugging of filters, coking on injectors, excessive engine wear,
high carbon deposits, engine knocking, and lubricating oil thickening and gelling
(Demirbas 2009). Therefore, biodiesel purification is an important step in the
production of biofuel grade biodiesel meeting the standard specifications. Table 4.2
shows the standard specification of biodiesel proposed by the American Society for
Testing Materials (ASTM) and European Standard (EN).

Table 4.1 The effects of impurities on biodiesel and engines. Reproduce from reference (Berrios
and Skelton 2008) with permission from Elsevier
Impurity Effect
Free fatty acids (FFA) Corrosion
Low oxidation stability
Water Hydrolysis (FFA formation)
Corrosion
Bacteriological growth (filter blockage)
Methanol Low values of density and viscosity
Low flash point (transport, storage, and use problems)
Corrosion of Al and Zn pieces
Glycerides High viscosity
Deposits into injectors (carbon residue)
Crystallization
Metals (soap and catalyst) Deposits into injectors (carbon residues)
Filter blockage (sulfated ashes)
Engine weakening
Glycerol Setting problems
Increases aldehydes and acrolein emissions
4 Biodiesel Purification and Upgrading Technologies 59

Table 4.2 Biodiesel standard specification based on ASTM D6751 and EN 14214
ASTM D6751 EN 14214
Test Limits Test Limits
method method
Flash point D93 130.0 °C min EN ISO 120.0 °C min
3679
Water D2709 0.050 vol.% EN ISO 500 mg/kg
max 12937 max
Kinematic viscosity D445 1.9–6.0 mm2/s EN ISO 3.5–5.0 mm2/s
3104
Density – – EN ISO 860–900 kg/
3675 m3
Ester content – – EN 14103 96.6 mol%
Methanol content – – EN 14110 0.20 mol%
max
Sulfated ash D874 0.020 wt% ISO 3987 0.02 mol%
max max
Sulfur D5483 0.0015% max EN ISO 10.0 mg/kg
(S15) 20846 max
0.05 max
(S500)
Copper corrosion D130 No. 3 max EN ISO 1 degree of
2160 corrosion
Cetane number D613 47 min EN ISO 51 min
5165
Cloud point D2500 – – –
Carbon residue D4530 0.050 wt% EN ISO 0.30 mol%
max 10370 max
Acid number D664 0.50 mg KOH/ EN 14104 0.50 mg KOH/
g max g max
Free glycerin D6584 0.02 wt% max EN 14105 0.02 mol%
max
Total glycerin D6584 0.240 wt% EN 14105 0.25 mol%
max max
Phosphorous content D4951 0.001 wt% EN 14108 10 mg/kg max
max
Sodium/Potassium UOP391 5 ppm max. EN 14107 5 mg/kg max
Distillation temperature D1160 360 °C max – –
(90% recovery)

Numerous techniques have been studied for biodiesel purification, commonly


categorized as wet washing, dry washing, and membrane technologies (Jaber et al.
2015). Wet washing is performed using deionized/distilled water, acidified water, or
organic solvents (Atadashi et al. 2011a, c). The dry washing takes advantage of
proper adsorbents to filter out the impurities (Gomes et al. 2015). Despite the
60 H. Bateni et al.

relatively satisfying performance of the dry washing method, this technique suffers
from the lack of adsorbent reusability and limited knowledge about their chemistry
(Atadashi et al. 2012). Membrane technology is another well-established separation
method in the purification of bio-based products being explored for biodiesel
purification as well (Atadashi et al. 2012; Atadashi 2015). The separation and
purification techniques can be also classified based on the nature of their process as
equilibrium-based, affinity-based, membrane, solid–liquid, and reaction-based
separation processes (Huang and Ramaswamy 2013; Dechow 1989) as shown in
Fig. 4.1.
The major production and operating costs in biorefineries are determined by the
separation and purification processes. Hence, robust separation technologies are
required to achieve an economically viable plant. A conventional standalone
purification process is not usually enough to meet standard specifications.
Therefore, a proper combination of the aforementioned methods is required to
achieve an effective purification strategy in order to produce high-quality biodiesel
(Venkatesan 2013).

Fig. 4.1 Overview of the separation processes for biodiesel purification. Reprint from reference
(Bateni et al. 2017) with permission from Biofuel Research Journal
4 Biodiesel Purification and Upgrading Technologies 61

4.1.1 Equilibrium-Based Separation Processes in Biodiesel


Purification

Distillation and liquid–liquid extraction (LLE) are the most common equilibrium-
based biodiesel purification processes. Additionally, there has been a report on the
successful use of supercritical fluid extraction using carbon dioxide at 40 °C and
30 MPa with a biodiesel separation yield of 99.94% (Wei et al. 2014).

4.1.1.1 Distillation

Distillation refers to the separation of the components from a liquid mixture based
on their relative volatility in a unit (column) with selective heating and cooling.
Among different types of distillations, conventional distillation and evaporation are
the most common techniques usually used for alcohol and water removal from
crude biodiesel prior to further purification (Atadashi et al. 2011a, c; Stojković et al.
2014; Bateni et al. 2014; Bateni and Karimi 2016b; Gomes et al. 2011; Moser
2012). Molecular distillation is another type of distillation under high-vacuum used
for biodiesel purification. This method facilitates the transmission of the evaporated
molecules to the condensing surface. In fact, the distance that molecules need to
travel to reach the condensing surface is shorter than the free path of the molecules
which eliminates the collision with foreign gas molecules (Erich 1982). Therefore,
this technique can potentially result in a higher separation yield (Erich 1982), as it
was confirmed by Wang et al. (2010) for purification of biodiesel obtained from
waste cooking oil with about 98% separation yield at 120 °C.

4.1.1.2 Liquid–Liquid Extraction

Liquid–liquid extraction, also known as solvent extraction, is perhaps the most


common method for biodiesel purification which includes all the wet washing
techniques (Berrios and Skelton 2008; Veljković et al. 2015). In this process, an
appropriate solvent is used to extract the desirable component(s) from the liquid
feed (Huang and Ramaswamy 2013; Dechow 1989; Hanson 1971; Kertes 1971).
Water is the most common solvent in this process, used in water washing or
acidified water washing techniques (Atadashi et al. 2011c; Serrano et al. 2013;
Veljković et al. 2015; Abbaszadeh et al. 2014). Water temperature, volume, and pH
along with the insertion method are among the most effective factors on the per-
formance of this technique (Atadashi et al. 2011a; Muniyappa et al. 1996; Veljković
et al. 2015; Abbaszadeh et al. 2014; Coêlho et al. 2011; Saifuddin and Chua 2004).
It is expected that an increase in temperature and volume of water improves the
glycerol diffusivity (from the biodiesel phase to the water phase) and increases the
mass transfer surface area, resulting in a higher volumetric mass transfer coefficient
and consequently a better biodiesel purification (Atadashi et al. 2011a). However,
62 H. Bateni et al.

the water solubility in biodiesel is also increased at higher temperatures, which may
result in a higher water content in the final biodiesel (Stojković et al. 2014). From
the design perspective, providing the required volume with a multistage washing
process may be more favorable than an extraordinarily large washing unit. It was
repeatedly reported that the glycerol content of biodiesel decreased to the required
standard after multistage wet washing (Atadashi et al. 2011a, c; Ma et al. 1998;
Rahayu and Mindaryani 2007; Canakci and Van Gerpen 2003). Even though
mixing can improve the mass transfer coefficient, there is a concern about the
formation of emulsions and subsequent biodiesel losses at high agitation speeds
(Atadashi et al. 2011c). A water spraying technique at low velocity was introduced
to decline this concern while still providing a high mass transfer area (Saifuddin and
Chua 2004). High water consumption is the most concerning challenge of biodiesel
purification via water washing. It was reported that 3–10 L water is required to
purify one liter of biodiesel which raises additional concerns about handling a large
quantity of wastewater along with a high production cost (Jaber et al. 2015;
Veljković et al. 2015; Serrano et al. 2013). It is worth mentioning that the amount of
required water for biodiesel purification is still a function of vegetable oil quality
and reaction conditions (e.g., catalyst and alcohol contents) (Saifuddin and Chua
2004). There has been a great effort to seek a better purification process or a
superior solvent to overcome the challenges associated with high water consump-
tion in water-based washing methods. Jaber et al. (2015) introduced a novel
purification method with an emphasis on reusing the wastewater by way of
microfiltration and adsorption processes (using activated carbon) for recycling the
water stream. Such a process can decrease the overall water consumption by 15%
(Jaber et al. 2015).
On the other hand, the use of acidified water has also received attention since
less water is required to neutralize the homogeneous alkali catalyst remaining in the
biodiesel. Phosphoric acid, sulfuric acid, and hydrochloric acid are among the most
common acids used for this purpose (Atadashi et al. 2011c). It is noteworthy that
even though the hydrolysis of soaps to free fatty acids in the presence of acidified
water may decrease the degree of emulsification associated with the presence of
soaps (Atadashi et al. 2011c), it may also increase the acidity of the final product
(Huerga et al. 2014).
Organic solvents, e.g., petroleum ether and n-hexane, have also been used in
purification of biodiesel, where the crude biodiesel is dissolved in the organic
solvent and then subjected to further water washing steps. Figure 4.2 shows the
schematic of the biodiesel purification process used by Karaosmanoglu et al.
(1996), in which the crude biodiesel was subjected to a wet washing process after
methanol recovery using a rotary evaporator under vacuum.
The use of water in the aforementioned methods results in relatively high water
content in the purified biodiesel which needs to be removed prior to storage. This
process is energy and time intensive and increases the cost of the purification
process (Dugan 2007). Moreover, the water-based washing methods may not be as
successful for purification of biodiesel produced via heterogeneous transesterifi-
cation, especially using calcium-based catalysts which results in very stable calcium
4 Biodiesel Purification and Upgrading Technologies 63

Fig. 4.2 Biodiesel refining via dissolving in petroleum ether and washing with distilled water.
Reproduced from reference (Karaosmanoǧlu et al. 1996) with permission from American
Chemical Society

soaps dissolved in the biodiesel (Veljković et al. 2015; Alba-Rubio et al. 2012).
Consequently, further research has been done to develop a proper waterless
purification method.
Solvent extraction using ionic liquids (ILs) is another separation method intro-
duced in biodiesel purification. ILs are green, nonflammable, nonvolatile, reusable
but often expensive solvents which can dissolve a wide range of organic, inorganic,
and organometallic materials (Han and Row 2010; Zhao and Baker 2013; Shahbaz
et al. 2010). The Deep Eutectic Solvents (DES) are biodegradable, nontoxic,
nonreactive, and relatively cheap ILs with promising potential in biodiesel purifi-
cation (Shahbaz et al. 2010, 2011). Abbott et al. (2007) evaluated the potential of a
series of quaternary ammonium salts to form eutectic solvents with glycerol during
biodiesel purification where [EtNH4]Cl, [ClEtMe3N]Cl, and choline chloride
showed the most promising results for glycerol removal. Hayyan et al. (2010)
reported over 51% glycerol removal when a 1:1 molar ratio of choline chloride salt/
glycerol was used (in the case of a 1:1 molar ratio of DES/biodiesel). Shahbaz et al.
(2011) used promoting agents, i.e., ethylene glycol, 2,2,2-trifluoroacetamide, and
methyltriphenylphosphunium bromide, along with choline chloride to improve the
performance of the mixture for glycerol removal. Figure 4.3 provides an overview
for glycerol removal via this method (Abbott et al. 2007).

4.1.2 Affinity-Based Separation Processes in Biodiesel


Purification

Affinity-based separation is a method of separating a mixture based on a strong


specific interaction between the species and a solid phase introduced into the system
64 H. Bateni et al.

Fig. 4.3 Schematic diagram for biodiesel production and purification using a sub-eutectic salt.
Reproduced from reference (Abbott et al. 2007) with permission from the Royal Society of
Chemistry

(Manesiotis and Theodoridis 2016). Adsorption and ion exchange are among the
most common affinity-based techniques for biodiesel purification, often known as
dry washing methods (Atadashi 2015; Huang and Ramaswamy 2013). The
downstream wastewater treatment is avoided in dry washing methods due to the
absence of water. These methods can also be easily integrated into an existing plant
and require shorter purification times and less space compared to wet washing
processes (Atadashi et al. 2011c). Moreover, the water content of dry washed
biodiesel is typically within an acceptable range (less than 500 ppm) (Dugan 2007).

4.1.2.1 Adsorption

Adsorption refers to the “sorption” of the adsorbates (atoms, ions, or molecules)


from the bulk of a fluid to the surface of a polar amorphous or microcrystalline
solid, known as the adsorbent (Venkatesan 2013; LeVan and Carta 2008).
Therefore, the selection of an appropriate adsorbent is crucial in selectively
removing impurities (Atadashi 2015; Huang and Ramaswamy 2013). Silicon-based,
biomass, and activated carbon are the most common adsorbents for biodiesel
purification (Atadashi 2015; Atadashi et al. 2011c).
Silicon-Based Adsorbent
Silica gel, zeolites, and molecular sieves are some of the most common
silicon-based adsorbents greatly used in industrial adsorption (Venkatesan 2013;
Dechow 1989). Silica gel is a proper adsorbent for polar species (e.g., water,
methanol, glycerol, etc.) due to the presence of hydroxyl groups on the surface
(Ruthven 1984); therefore, it is often used for glycerol removal from biodiesel at
room temperature (Yori et al. 2007; Mazzieri et al. 2008; Manuale et al. 2011;
Predojević 2008; Faccini et al. 2011). The presence of multiple polar compounds
may affect the potential of silica gel in adsorbing a specific species. For example,
the adsorption of glycerol from the crude biodiesel decreases in the presence of
4 Biodiesel Purification and Upgrading Technologies 65

methanol due to the stronger affinity of alcohol with the surface. However, a limited
amount of water and soap does not affect the glycerol adsorption (Mazzieri et al.
2008). Silica gel also has great potential in adsorption of free fatty acids from
biodiesel (Manuale et al. 2011).
Naturally available silicon-based adsorbents such as bentonite can also be used
as an adsorbent for dry washing of biodiesel. However, a pretreatment step is likely
required to activate the substrate. For instance, Leeruang and Pengprecha (2012)
treated low silica content bentonite with 0.1 H2SO4 at 100 °C prior to use in
biodiesel purification.
Commercially available silicon-based adsorbents were also utilized in biodiesel
purification. Magnesol, an inorganic matrix of magnesium silicate and anhydrous
sodium sulfate, displayed promising results in removing soap, methanol, glycerol,
and water after being thoroughly mixed with crude biodiesel (Berrios and Skelton
2008; Faccini et al. 2011; Atadashi et al. 2011c).
The particle size is an important factor in the performance of adsorbents. In fact,
superior performance is expected from smaller particles due to a higher surface area
and more accessibility to the pores (Leeruang and Pengprecha 2012). Due to
technical recommendations, larger particles (typically 1/8 in. beads) are used in
industrial scale units, resulting in lower adsorption capacity due to mass transfer
limitations (Yori et al. 2007).
Biomass-Based Adsorbent
Lignocellulosic and cellulosic substrates are inexpensive, abundant, renewable, and
biocompatible materials (Bateni et al. 2016; Gomes et al. 2015; Noori and Karimi
2016a, b) with great potential for biodiesel purification (Gomes et al. 2015;
Manique et al. 2012).
Gomes et al. (2015) evaluated the potential of corn starch, rice starch, potato
starch, cassava starch, and eucalyptus bleached krafted cellulose in biodiesel
purification. The glycerol content was decreased to an acceptable range using any
of the following biomasses after 10 min of gentle agitation (150 rpm) at room
temperature: 5% potato starch, 2% corn starch, 1–2% cassava starch, and 1% rice
starch (Gomes et al. 2015). The presence of high silicon content in the porous
structure (mainly meso- and macropores) of rice husk ash (RHA) made it a proper
adsorbent for biodiesel purification. The use of 1–5% RHA could effectively
decrease the glycerol content, even though the water content was still beyond the
typically acceptable range (Manique et al. 2012).
Activated Carbon
Activated carbon (AC) is a promising porous adsorbent with a high surface area and
large specific volume usually manufactured via hydrothermal (steam) or chemical
activation of organic compounds such as sawdust, charcoal, coal, etc. (Venkatesan
2013; Yang 2003). Fadhil and Dheyab (2015) used activated carbon to purify the
biodiesel produced from waste cooking and frying oils. Moreover, sulfuric and
hydrochloric acids were used to pretreat the AC and enhance its performance. It was
reported that treated AC with sulfuric acid had slightly better performance in the
66 H. Bateni et al.

purification of biodiesel produced from waste fish frying oil compared to untreated
AC. However, the untreated AC was more successful in purifying the waste
cooking oil (Fadhil and Dheyab 2015).

4.1.2.2 Ion Exchange

Ion exchange is defined as the stoichiometric electrostatic adsorption process due to


Coulomb-attractive forces between the ions and the charged functional groups on a
solid-phase exchanger (Berrios et al. 2013). Therefore, ions with equivalent charge
are exchanged between the solution and the exchanger (Treybal 1980; Dechow
1989; LeVan and Carta 2008). This process is commonly performed through the
diffusion of ions between the phases; however, a chemical reaction can also drive
the process (e.g., the neutralization process) (Berrios et al. 2013). The ion
exchangers consist of a matrix with excess charges localized in specific sites of the
structure (Berrios et al. 2013; Grandison 1996). The synthetic ion exchanger matrix
is obtained via copolymerization of styrene crosslinked with divinylbenzene
(Dechow 1989). The matrix is then subjected to functionalization to incorporate
acid or base functional groups into the structure. The resulting ion exchange resins
can be classified based on functionality and density of their charges as strongly
acidic cation and basic anion exchangers as well as weakly acidic cation and basic
anion exchangers (Berrios et al. 2013; Dechow 1989). The strong acidity of the
resins is generally provided using sulfonic acid groups (sulfonated polystyrene,
R–SO−3 H+) with the ability to exchange its proton with other cations. Figure 4.4
shows the structure of strong acid cation resins (Berrios et al. 2013).

Fig. 4.4 Structure of a strong acid cation resin (sulfonated polystyrene resin). Reprint from
reference (Berrios et al. 2013) with permission from John Wiley and Sons
4 Biodiesel Purification and Upgrading Technologies 67

In the case of weakly acidic cation resins, carboxylic groups (R–CO−2 H+)
provide the interchangeable protons in the resins with an acid strength similar to
acetic acid (Berrios et al. 2013). On the other hand, the presence of quaternary
ammonium exchange sites provides a strong basic functionality for the anion
exchange resins, while the weak base anion resins derive their unique ion exchange
properties from the presence of radicals of secondary or tertiary amines (Berrios
et al. 2013; LeVan and Carta 2008).
The presence of counter-ions that migrate within the free space of the matrix
compensates the charge of the exchanger (Berrios et al. 2013). Besides the type and
density of the charges, structural characterization including the degree of
cross-linking, porosity, and particle size along with moisture content, exchange
capacity, and stability of the structure are the important factors for selecting a
proper ion exchange resin (Dechow 1989; Grandison 1996; Berrios et al. 2013).
The industrially available strong acid cation resins such as Dowex DR-G8 (Dow
Chemical), Lewatit S7968 (LANXESS), BD10 Dry (Rohm and Haas), and PD206
(Purolite) are among the ion exchange resins used for biodiesel purification
(Atadashi et al. 2011c; Faccini et al. 2011; Berrios et al. 2013). The presence of
acidic and basic adsorption sites on these adsorbents facilitates the removal of polar
compounds, including glycerol, glycerides, methanol, metals, and soap. Berrios and
Skelton (2008) showed that PD206 and BD10 dry cation exchange resins cannot
remove methanol from the crude biodiesel, even though they successfully decreased
the amount of soap and glycerol. However, Berrios et al. (2013) found that Lewatit
GF202 has the potential to remove all of the aforementioned impurities. It was also
reported that Lewatit GF202 can effectively decrease the acidity and viscosity of
biodiesel even though it did not show promising results in terms of removing the
metallic species (sodium and potassium from the alkaline catalysts) (Mata et al.
2011). Wall et al. (2011) studied the mechanism of biodiesel purification using
industrially available ion exchange resins, i.e. T45BD and T45BDMP (Thermax)
along with BD10 Dry (Dow Chemical). They concluded that filtration, physical
adsorption, ion exchange, and soap removal by glycerol affinity are the most
important factors during this process. It is noteworthy that the resins had a better
performance for sodium soap removal compared to potassium soap and decreasing
the particle size further improved the process in the case of sodium soaps (Wall
et al. 2011). Chen et al. (2012a) discovered glycerol’s great affinity to sodium form
sulfonated resins, even though the hydrogen form resins own a higher adsorption
capacity (Chen et al. 2012b). Dias et al. (2014) investigated the effect of resin
(PD206, Purolite) content (2–40%) on the ester and water content of the final
biodiesel samples from soybean oil and waste frying oil, where the purification was
performed at room temperature for 1 h. Purification of the samples using 40% resin
provided the best quality products, even though the ethyl ester contents were still
slightly lower than the standard value reported in EN 14214 (Dias et al. 2014).
Given the results of soybean oil biodiesel purification using 2% Purolite at room
temperature (Dias et al. 2014) and 65 °C (Faccini et al. 2011), it can be concluded
that purification at lower temperature resulted in lower water content in the final
biodiesel.
68 H. Bateni et al.

4.1.3 Solid–Liquid Separation Processes in Biodiesel


Purification

Conventional filtration, solid–liquid extraction (SLE), precipitation, and crystal-


lization are among the most common solid–liquid separation processes (Bhayani
and Ramarao 2013; Lewis 1996). Solid–liquid separation is probably the least
discussed concept for biodiesel purification in the literature, as homogeneous
catalysis transesterification and water washing are the most common reaction and
purification strategies, respectively. However, in the case of using heterogeneous
catalysts or dry washing methods, solid–liquid separation is a necessity where
conventional filtration is the most mature technology used for biodiesel purification.
It is worth mentioning that SLE technologies including conventional extraction,
ultrasound-assisted extraction, and microwave-assisted extraction are often used for
extraction of vegetable oil from the seeds prior to transesterification rather than
biodiesel purification (Bhayani and Ramarao 2013).

4.1.4 Membrane-Based Separation Technologies


in Biodiesel Purification

Recently, membrane technology has been introduced as an alternative to the typical


washing methods commonly employed in the purification of crude biodiesel.
Membrane filtration can directly address the challenges commonly associated with
conventional methods of biodiesel refining, condensing multiple step processes,
such as the removal of glycerol, the removal of leftover methanol/ethanol, and the
removal of unreacted mono-, di- and triglycerides into one or two-step processes
(Atadashi et al. 2011b). The use of a membrane can also avoid the high water
content involved in traditional water washing methods (Jaber et al. 2015). In
addition, the avoidance of water reduces the likelihood of unwanted emulsion in the
crude product, allowing for ease of separation based on two-phase behavior. The
membrane acts as a semipermeable filter, restricting the continued flow of unwanted
components whose removal is necessary for biodiesel purification. These methods
are often limited by temperature, pressure, and concentration differences as well as
phase behavior (Alves et al. 2013; Cheng et al. 2009). Terminology such as
microfiltration, ultrafiltration, and nanofiltration are often used to describe the
membrane technologies employed in these processes—the terms are dependent on
the pore size of the membrane, and are somewhat arbitrary. As a general rule, a
microfiltration membrane will retain particles in the 0.1–10 lm range; ultrafiltration
membranes retain particles sizes between 1 and 20 nm (Jönsson 2013). The pore
sizes can also be defined in the literature as nominal molecular weight cutoffs
(NMWCOs); that is, the lowest molecular weight for a 90% retention rate by the
membrane in Daltons (Da). This terminology is not common for microfiltration; for
ultrafiltration, the NMWCO is between 1 and 1000 kDa (Jönsson 2013).
4 Biodiesel Purification and Upgrading Technologies 69

Fig. 4.5 The designation of a FAME-rich phase from a crude biodiesel. Reprint from reference
(Alves et al. 2013) with permission from Elsevier

Nanofiltration is a more recent term, filling in between ultrafiltration and reverse


osmosis membranes with pores no larger than 1 nm or NMWCO between 200 and
1000 Da (Mänttäri et al. 2013).
Whether discussing organic or inorganic membranes, the separation itself
functions via size-exclusion or chemical permeability. Thus, it becomes paramount
to discuss a membrane, regardless of composition, that can selectively filter bio-
diesel (or fatty acid methyl esters—FAMEs) from a crude product, as seen in
Fig. 4.5.
Free glycerol exists in a crude biodiesel phase as a suspension of droplets whose
size distribution is dependent on the composition (Gomes et al. 2010; Wang et al.
2009) and temperature of the biodiesel flow (Saleh et al. 2011). In the case of
composition dependence, additional alcohol used in excess for transesterification can
decrease glycerol droplet size, as the alcohol can act as a surfactant which reduces
droplet surface tension, allowing for ease of permeability of unwanted glycerol
(Gomes et al. 2010). Saponification products had a similar effect in preventing the
glycerol droplet sizes from reaching desired levels (Saleh et al. 2010a, b). In tests
with crude biodiesel, the glycerol was found to form glycerol-alkali bonds, creating a
“reverse micelle” structure for the droplets in the crude biodiesel suspension (Wang
et al. 2009). Additionally, unreacted oils present in a semicontinuous membrane
reactor can form droplets of a bimodal size distribution dependent on the oil used. In
Falahati and Tremblay (2012) it was found that waste cooking oil formed larger
droplets than canola oil, leading to lower permeability of di- and triglycerides in the
case of the waste oil.
Intertwined with the size distribution, the separation is additionally limited by
the phase behavior of the crude biodiesel; thus, the temperature of the crude
retentate as well as the composition of the crude must be chosen such that two
distinct phases can be formed. In the case of temperature and composition,
70 H. Bateni et al.

Fig. 4.6 Ternary phase behavior of an example oil–FAME–methanol system, both simulated and
from experimental data at varying temperatures. Reprint from reference (Cheng et al. 2009) with
permission from Elsevier

Cheng et al. (2009) posited the separation effectiveness of a ternary oil–FAME–


methanol system; the results of simulation which can be seen in Fig. 4.6.
Once the phase behavior crosses into the single-phase region, the permeate flux
increases dramatically—however, the permeate composition is not fundamentally
different from the feedstock, implying that the crude biodiesel passes through the
membrane undeterred. Based on calculations involving the octanol–water coeffi-
cients, Cao et al. (2007) utilized an empirical model (Ho et al. 1990) to calculate the
point at which a heterogeneous two-phase flow of methanol and biodiesel would
exist in a membrane reactor based on its volume fractions and viscosities:
 0:29
/1 l
¼ 1:22 1
/2 l2

Based on this semiempirical model developed for components which differ


widely in viscosity, it was noted that phase inversion would occur at a methanol
volume fraction lower than 0.31; testing this with varying oil loadings, Cao et al.
(2007) found that the permeate flux dropped to near zero at a methanol volume
fraction of 0.28. Gomes et al. (2013) provided further insight into the effect of
acidified water on phase behavior; increasing emulsion in the crude biodiesel will
generally yield higher unwanted glycerol content in the permeate flow.
Demonstrated across a variety of porous membranes and pore sizes, the importance
4 Biodiesel Purification and Upgrading Technologies 71

of two-phase flow in membrane separation is readily apparent—FAMEs are much


more miscible in methanol than triglycerides, and so a FAME-rich methanol phase
facilitates separation of a pure biodiesel product (Dubé et al. 2007; Cheng et al.
2009).
The research around membrane technologies is still somewhat sparse and can be
sorted into two distinct categories: organic and inorganic membranes. Organic
membranes are polymer-based, including polysulfone (Alves et al. 2013; Alicieo
et al. 2002; He et al. 2006; Giorno et al. 2013), polyamide (Jiang et al. 2009), and
polycarbonate (Cao et al. 2007) among other highly developed polymers, including
regenerated celluloses and polyvinylidenefluoride (Giorno et al. 2013; Mah et al.
2012). Other polymers have been attempted in use for biodiesel purification
including polyacrylonitrile (He et al. 2006; Saleh et al. 2010b); however, due to the
hydrophobic behavior of the polymeric structure the likelihood of glycerol trans-
mitting into the purified permeate phase increases. Inorganic polymers focus gen-
erally on an alpha-alumina (Al2O3) support structure with titanium oxide (Cao et al.
2008a; Alicieo et al. 2002; Gomes et al. 2010, 2011, 2013; Baroutian et al. 2011;
Basso et al. 2006; Cao et al. 2008b) or zirconium oxide (Cheng et al. 2009).

4.1.4.1 Organic/Polymeric Membranes

The use of an organic membrane can bypass the need for water washing endemic to
traditional methods of biodiesel purification, thereby minimizing ecological prob-
lems that result from the further treatment of a waste stream. However, organic
membranes are susceptible to swelling and conformation changes due to their
structure, which can result in long-term pore structure alterations and partial fouling
of the membrane at more basic pH contents, as seen in Fig. 4.7 (Salahi et al. 2010).

Fig. 4.7 A polyacrylonitrile membrane before and after cleaning with a pH 10 solvent. Reprint
from reference (Salahi et al. 2010) with permission from Elsevier
72 H. Bateni et al.

The wide variety of organic membranes available can be sorted by their


hydrophilicity or hydrophobicity. Hydrophilic membranes are less susceptible to
fouling mechanisms when dealing with aqueous solutions, whereas a hydrophobic
material is useful when separating oils. The hydrophobicity or hydrophilicity
directly correlates with sessile drop contact angles; lower contact angles imply
higher hydrophilicities (Mänttäri et al. 2013).
Polysulfone is demonstrably hydrophilic; for low water content feeds this can be
beneficial for reduction of glycerol agglomeration typically associated with pore
fouling/blockage. In direct comparison with polyacrylonitrile membranes, as well
as water and acid washing methods, He et al. (2006) found large successes in the
use of polysulfone to decrease the ester losses involved in purifying biodiesel.
Further, it was observed that polyacrylonitrile permitted higher water content into
the permeate flow, demonstrating that polyacrylonitrile may be unsuitable for
biodiesel refining. The polysulfone membrane, by contrast, met the ASTM stan-
dards for water content without additional steps. Additionally, the polysulfone
membrane experienced the lowest ester losses across all of the methods attempted,
due to issues with emulsification in water washing methods at low temperatures (He
et al. 2006). Further comparisons between poly(ether-sulfone) membranes and
cellulose ester membranes demonstrated a successful separation of glycerol without
additional water for poly(ether-sulfone) alone (less than 0.02% by mass); however,
the pore sizes were only 10 and 30 kDa for the UF poly(ether-sulfone) membranes
as compared to 0.22 and 0.3 lm for the cellulose ester membranes (Alves et al.
2013).
As discussed in the beginning of this chapter, glycerol separation from biodiesel
remains a paramount issue, both in the potential damage to engines if present and
the difficulty of separation (Saleh et al. 2010b). Glycerol will tend to agglomerate
into droplets of various sizes; Saleh et al. (2010b) found these droplets varied from
250 nm in FAME to more than 2500 nm in 0.06% water by mass. Thus, a properly
tuned organic membrane with smaller pores than this could aid significantly in
glycerol separation. In one case, a modified hydrophilic polyacrylonitrile membrane
(Ultrafilic) in the 100 kDa range was tested for glycerol separation and led to an
ease of separation with increasing water content, from 3% in 180 min (FAME only)
to up to 63% in 180 min (FAME with 0.2% water by mass). This was reflected in
work from Alves et al. (2013), where a small addition of water (0.1 and 0.2% water
by mass) drastically improved glycerol removal for a 10 kDa poly(ether-sulfone)
membrane (from 0.02 to 0.009% glycerol in permeate). The increased separation
and droplet sizes demonstrate the glycerol/water miscibility that contributes to
two-phase flow (Gomes et al. 2010; Saleh et al. 2011; Alves et al. 2013; Wang et al.
2009). Adding soap and methanol to the mixture drastically lowers the separation of
FAME and glycerol; in the case of the aforementioned polyacrylonitrile membrane,
Saleh et al. (2010b) found negligible separation with 1% methanol and soap and
0.06% water content by mass (2% after 180 min). Further analysis was conducted
regarding glycerol droplet sizes with a three-level Box–Behnken factorial design
and dispersive light scattering; Saleh et al. (2010a) found a positive correlation with
4 Biodiesel Purification and Upgrading Technologies 73

glycerol separation and predicted glycerol droplet size based on this model, even in
homogeneous flow regimes (i.e., when there are no glycerol droplets).
Biodiesel production from microalgae cells appears a lucrative prospect due to
its relative lack of competition for agriculture-suitable land. It is limited by a few
bottlenecks, however, among other issues, harvesting efficiency and triglyceride
recovery remain substantial challenges (Wijffels et al. 2010). Membrane technology
could potentially solve these challenges without additional additives and at mod-
erate temperatures and pressures, so long as the said membrane is chosen based on
its non-fouling ability (Rossi et al. 2004). To test fouling across various organic
membranes, Giorno et al. (2013) utilized prior work conducted for algal cell sep-
aration (Rossi et al. 2008) and separation of proteins and lipids from algal
wastewater streams (Dumay et al. 2008) among others. Regenerated cellulose,
polysulfone, and polyvinylidenefluoride membranes in the ultrafiltration range
(nominal weight cutoffs of between 100 and 150 kDa) were tested for 20 min with
40 mL of fresh Nannochloropsis s.; the flux was measured with water before and
after contact with the membrane. The 100 kDa RC membrane maintained the
highest flux here and did not lose flux even in tests with partially sonicated cells
where fouling would be more readily apparent. Issues were apparent with the
polyvinylidenefluoride membrane (PVDF) as it alone lost flux after the tests; these
membranes are susceptible to flux declines dependent on small pH and oil fluctu-
ations (Mah et al. 2012). Additional tests were conducted in this paper (Giorno et al.
2013) with a 30 kDa RC membrane, as protein removal remained an issue at
100 kDa (only 61% removal). A much higher degree of purification was achieved
(89% protein removal). Sonication provided higher fluxes and purities due to the
prevention of agglomeration between triglycerides and proteins.
Other membranes could be potentially employed for the purification of biodiesel,
such as polyimide (PI) or polydimethylsiloxane (PDMS), and polyimide mem-
branes have been studied in recent years (Jiang et al. 2009) and as a precursor in
newer membranes (Freeman et al. 2012) for organic separation via pervaporation;
however, these membranes have yet to be studied for biodiesel purification and
their hydrophobicity would be highly susceptible to fouling without the proper
modifications (Mänttäri et al. 2013).

4.1.4.2 Inorganic/Ceramic Membranes

In contrast to their organic counterparts, inorganic membranes could prove superior


due to a narrower pore size distribution, increased fouling resistance, a longer
time-on-stream, and increased temperature and pH fluctuation resistance; this
becomes paramount when dealing with typical base catalysts used to facilitate
transesterification (Atadashi 2015; Barredo-Damas et al. 2010). Because of these
factors leading to an increased separation effectiveness, an inorganic membrane can
be effectively used in a continuous process to separate a FAME-rich permeate from
the resultant retentate as the transesterification process is taking place (Cao et al.
2008a; Baroutian et al. 2011). This is further discussed in Sect. 4.1.5.2.
74 H. Bateni et al.

Increasing the transmembrane pressure correlates positively with increasing


permeate flux, as expected (Atadashi et al. 2012; Gomes et al. 2010; Hua et al.
2007); however, at these higher transmembrane pressures, free glycerol content
increases in the permeate flow (Atadashi et al. 2012) and could lead to partial
fouling or blocking of the membrane pores over time (Gomes et al. 2010; Hua et al.
2007). Thus, a modest transmembrane pressure of 1–2 bar is recommended for
ultrafiltration ceramic membranes (Atadashi et al. 2012; Gomes et al. 2010; Hua
et al. 2007; Wang et al. 2009).
As with polymeric membranes, the importance of glycerol separation in bio-
diesel necessitates its research with ceramic membranes. Glycerol separation in
ceramic membranes has been tested with crude biodiesel produced from refined
palm oil (Wang et al. 2009) and canola oil (Saleh et al. 2011). Pore sizes between
0.02 and 0.6 lm have been tested (Wang et al. 2009; Saleh et al. 2011; Atadashi
et al. 2012), from the ultrafiltration to the microfiltration range; permeate flux
increases with increasing pore size (Atadashi et al. 2012). However, the glycerol
that permeates through the membrane also increases with the increasing pore size,
in spite of the glycerol-soap “reverse micelle” droplets that form at 2.21 lm, well
above any of the pore sizes (Wang et al. 2009). Temperature has a positive effect on
both the permeate flux and the separation up to 40 °C; Saleh et al. (2011) found
success in meeting ASTM standards with a 0.05 lm membrane at 25 °C for
glycerol (less than 0.02% by mass), whereas Atadashi et al. (2012) found up to
99.2% glycerol retention for a temperature of 40 °C. Wang et al. (2009) found
additional success with a 0.1 lm membrane at 60 °C; these tests were conducted
after removing methanol from the crude biodiesel via a heated vacuum process, and
may not reflect processes where methanol interferes with two-phase behavior (Saleh
et al. 2011).
The relationship between the separation of various cations and pore size is
complex; in Wang et al. (2009), the potassium and magnesium content of the
permeate decreased with decreasing pore sizes (from 4.25 mg/kg and 0.25 mg/kg,
respectively, at 0.6 lm to 1.70 mg/kg and 0.15 mg/kg at 0.1 lm, respectively).
However, the sodium and calcium content was unaffected by shifting pore sizes; the
sodium was mostly reduced regardless of membrane pore sizes, whereas calcium
was unhindered (Wang et al. 2009). This is further confirmed in Ferrero et al.
(2014) who tested 0.05 and 0.1 lm ceramic membranes in the removal of calcium
from pretreated soybean oil. Only 30% and 22% removal of calcium occurred
initially with the 0.05 and 0.1 lm membranes, respectively (Ferrero et al. 2014).
The addition of sodium carbonate greatly assisted with calcium removal, producing
less than 5 ppm calcium upon subsequent membrane filtration which meets the EN
14214 standard (Ferrero et al. 2014).
The presence of ethanol in the membrane separation process can further com-
plicate the mechanics of separation. The size of droplets in a double-emulsion flow
(i.e., water–oil droplets in a larger water phase, in this case) is drastically modified
by the presence of a hydrophilic surfactant (van der Graaf et al. 2005). Decreasing
the droplet size is a direct result of the decrease in the interfacial tension (Yılmaz
et al. 1999); the presence of increasing ethanol content in an oil–water suspension
4 Biodiesel Purification and Upgrading Technologies 75

decreases this interfacial tension (Pittia et al. 2005). For the purposes of membrane
separation of biodiesel with an alpha-alumina/titanium oxide membrane, Gomes
et al. (2010, 2011) found that increasing ethanol content in a biodiesel–glycerol
mixture from 5% ethanol to 20% ethanol had complex effects on permeate flux
(63.1 kg/h/m2 at 5% ethanol by mass, 78.4 kg/h/m2 at 10% ethanol by mass and
59.5 kg/h/m2 at 20% ethanol by mass) (Gomes et al. 2010). Increasing ethanol from
5 to 20% corresponded to a decrease in glycerol retention, from 99.6% at 5%
ethanol by mass to 98.1% at 20% ethanol by mass (Gomes et al. 2010).
In crude biodiesel, where small amounts of leftover catalyst and soaps are
present, glycerol purification is further hindered; Gomes et al. (2011) found drastic
decreases in permeate flux with crude biodiesel versus synthetic mixtures of bio-
diesel–glycerol–ethanol (12.9 kg/h/m2 for a crude mixture vs. 78.4 kg/h/m2 for a
synthetic mixture as seen above). The presence of soaps led to a greater tendency to
foul the membrane, as reflected in research in polymeric membranes (Saleh et al.
2010b); additionally, the binding of soaps to glycerol droplets allow for ease of
passage of glycerol through a biodiesel suspension (Wang et al. 2009). For ceramic
membranes, these agglomerates can typically be removed with a cleaning step with
an organic solvent (such as hexane) employed at high cross flow velocities and low
transmembrane pressures (Basso et al. 2006).
The phase behavior is substantially altered by the addition of water (Gomes et al.
2011, 2013) which reflects similar results for organic membranes (Saleh et al.
2010b). Gomes et al. (2011) found that the addition of 20% acidified water by mass
(0.5% HCl content) drastically increased the separation of glycerol with less than
0.02% glycerol by mass in the permeate flow, meeting ANP biodiesel specifica-
tions. However, this also produced a much lower steady-state permeate flux of
6.9 kg/h/m2 (Gomes et al. 2011). Higher amounts of acidified water (30% by mass)
correlated with much higher permeate fluxes (up to 57.6 kg/h/m2) across different
membrane sizes (from 0.05 to 0.2 lm); however, the glycerol in the permeate
ranged from 0.006 to 40% with no discernible pattern (Gomes et al. 2013). As the
water, ethanol and glycerol in this case comprise the dispersed phase (Gomes et al.
2013), increasing the water content can lead to coalescence, which inverts a
once-stable emulsion (Groeneweg et al. 1998); this would lead to unpredictable
results in glycerol separation due to a lack of agglomeration (Saleh et al. 2010b).
The concept of phase inversion is an important one, as two-phase flow becomes
an imperative in the separation of oil/unreacted triglycerides from FAME products
as well (Cao et al. 2007; Cheng et al. 2009; Dubé et al. 2007). Reiterating from
Fig. 4.6 above, differing compositions of an oil–FAME–methanol system will lead
to homogeneous or heterogeneous phase behavior that is temperature dependent
(Cheng et al. 2009). With porous zirconia oxide/carbon-supported ceramic disk
membranes of a 300 kDa molecular weight cutoff, Cheng et al. (2009) tested
varying compositions of oil–FAME–methanol for membrane filtration to assess the
role played by phase behavior. By testing a crucial point of 26:54:20 (oil:FAME:
methanol) by weight, it was found that raising the temperature from 20 to 60 °C
altered the phase behavior significantly, by crossing the LLE boundary line seen in
Figure A. This shift in the phase behavior corresponded to the appearance of
76 H. Bateni et al.

triglycerides in the permeate flow, from 0% by weight at 20 °C to roughly 25% by


weight at 60 °C; in this manner, the permeate concentration of triglycerides is
identical to the feed concentration, indicating no rejection (Cheng et al. 2009).

4.1.5 Reaction-Based Separation Processes in Biodiesel


Purification

Hybrid reaction-membrane separation (membrane bioreactor) and reactive distil-


lation are the most common reaction-based separation methods used for biodiesel
purification. In these methods, the simultaneous reaction and separation improve
the overall process yield (Huang and Ramaswamy 2013; Kolah et al. 2013; Lei
et al. 2005; Kiss 2013).

4.1.5.1 Reactive Distillation

Reactive distillation, an integrated process for simultaneous chemical reaction and


distillation in a single unit, is used to improve the selectivity and productivity
toward the desired product while reducing the energy consumption (Kolah et al.
2013; Lei et al. 2005; Kiss 2013). Wang et al. (2001) explained that the use of
reactive distillation in methyl acetate hydrolysis can decrease the energy con-
sumption by 10% and increase the productivity by 50% compared to a conventional
fixed-bed reactor.
However, this process also suffers from its own limitations. The temperature and
pressure of the reaction must be compatible with the desired distillation to expe-
rience an effective reactive distillation process. This process cannot be applied for
gas–liquid reactions requiring severe operating conditions. Moreover, the reaction
rate can also limit the application of reactive distillation in the case of very slow
reactions which need prolonged contact time, resulting in large column sizes (Kolah
et al. 2013; Kiss et al. 2006a, b, 2007; Kiss 2010; Dimian et al. 2009; Qiu et al.
2010; da Silva et al. 2010; Mueanmas et al. 2010; Gomez-Castro et al. 2010;
Machado et al. 2011; Poddar et al. 2015; Pérez-Cisneros et al. 2016).
The potential of reactive distillation using acid catalysts for biodiesel production
has been widely discussed in the literature (Simasatitkul et al. 2011). Different
designs have been proposed to achieve a high reaction rate and biodiesel yield in a
reasonably sized distillation column (Kolah et al. 2013). Reactive distillation is
potentially a more economically advanced process for biodiesel production and
purification, as it can decline the necessity of a downstream alcohol recovery unit to
separate excess alcohol from biodiesel and glycerol streams (Qiu et al. 2010).
Reactive distillation can also be considered as an effective pretreatment method to
esterify the feedstocks with a high free fatty acid content prior to alkali-catalyzed
transesterification (Kolah et al. 2013; Russbueldt and Hoelderich 2009).
4 Biodiesel Purification and Upgrading Technologies 77

4.1.5.2 Membrane Bioreactors for Biodiesel Processing

The characteristics of polysulfone membranes, such as high temperature resistance,


mechanical strength, and inert chemical behavior have made it lucrative for a
variety of applications, including as an anion exchange membrane (AEM) for
methanol fuel cells (Li and Wang 2005; Li et al. 2006; Hao et al. 2000; Slade et al.
2012). To enhance catalytic activity in an AEM environment, quaternized ammo-
nium functional groups are grafted to the polysulfone surface to enhance catalytic
activity (Slade et al. 2012); this strategy was, very recently, applied to the two-step
process of biodiesel production and refining (Shi et al. 2016a, b). After quaternizing
a chloromethylated polysulfone membrane, active functional groups for biodiesel
production were added via alkalization with potassium hydroxide. Shi et al. (2016b)
determined an optimum reaction synthesis time of 24 h and an optimum ratio of
chloromethyl ether:PSF of 10:1 based on 1H NMR (Hao et al. 2000). The
alkali-catalyzed polysulfone (APSF) membrane performed well with THF as a
cosolvent (conversion of 98.6% for soybean oil). However, n-hexane is recom-
mended due to its lower toxicity and ease of separation due to its similarity in
solubility with methanol (Shi et al. 2016a). Water content beyond 5% caused a
sharp decline in conversion rate, due to the affinity of water molecules for the
catalytic active sites; small amounts of water had only a very small negative effect.
The catalytic membrane was additionally stable (at optimum conditions of 60 °C
and a 4 h reaction time) after five runs with only a 2.1% loss of conversion. Further
studies by this group (Shi et al. 2016a) elucidated a kinetic model for the reaction
on the surface of the membrane:

dCRCOOH
c¼ ¼ g0  gx  k  CRCOOH
dt

Here, c is the reaction rate of oleic acid (here used as a model molecule for
soybean oil), and η0 and ηx are the internal and external diffusion efficiencies,
respectively. Upon converting the differentials to those inherent to the membrane,
the following model was used with a root mean square (RMS) error of 3.76 for the
experimental data:
 
Sm e
X ¼ 1  1=EXP g0  gx  kL 
R

In this equation, L, Snv, and e are membrane-specific parameters, L being


membrane thickness in mm, Snv, the column cross-sectional area in mm2, and e,
membrane porosity (Shi et al. 2016a).
Aside from this work, the vast majority of research regarding membrane reactors
has been conducted using inorganic membranes (Baroutian et al. 2011; Falahati and
Tremblay 2012; Cao et al. 2007, 2008a, b; Tremblay et al. 2008; Dubé et al. 2007).
This is owed to their stability, narrow pore distribution and resistance to chemical
attack relative to polymeric membranes of most types (Dubé et al. 2007; Baroutian
78 H. Bateni et al.

et al. 2011). Of this, a large portion of the work done regarding ceramic membrane
reactors for biodiesel purification has emerged out of a research group at the
University of Ottawa (Falahati and Tremblay 2012; Cao et al. 2007, 2008a, b;
Tremblay et al. 2008; Dubé et al. 2007).
Membrane reactors for refinery processing have been envisioned for some time
for applications including steam reformation and hydrogenation (Armor 1998); this
process was applied to biodiesel production and purification by Dube et al. (2007),
utilizing an inorganic carbon membrane of pore size 0.05 lm at temperatures
between 60 and 70 °C with a sulfuric acid catalyst at 6 wt%. Greater success was
found using a base catalyst at 1 wt% (95–96% conversion vs. 35–64% conversion
for sulfuric acid at various flow rates); however, this method produced soap as a
byproduct, indicating that it may have limited success in high-waste feedstocks
(Dubé et al. 2007). This membrane technology was further elucidated and improved
upon by Cao et al. (2007, 2008a, b). In earlier work (Cao et al. 2007), the droplet
sizes of a methanol–oil emulsion were modeled to determine optimum pore sizes
for a membrane bioreactor. Based on a model seen above in Sect. 4.1.4 (Ho et al.
1990), optimum amounts of methanol and oil loadings could be calculated by
maintaining a methanol volume fraction above 0.31 (Cao et al. 2007). Additionally,
work conducted by DeRoussel et al. (2001) produced a model for calculating
droplet sizes based on interfacial tension, shear rate, and viscosity ratio—it was
found that the minimum droplet size for an oil/methanol emulsion was 12 lm,
prompting pore sizes of 1.4 lm at the largest (Cao et al. 2007).
In attempting to bring the membrane bioreactor technology in line with con-
ventional biodiesel production processes, it may be beneficial to recycle the polar
methanol-rich phase in order to lower the methanol-to-oil ratio charged into the
bioreactor (Cao et al. 2008a, b). Cao et al. (2008b) employed a cooler immediately
after the permeate flow from an alpha-alumina/titanium oxide membrane, which can
be seen in Fig. 4.8, in order to separate the organic FAME phase from the
methanol-rich polar phase; even with a 100% recycle rate, no adverse effects on
FAME purity were seen regardless of the buildup of glycerol in the system.
However, glycerol buildup could potentially cause fouling through agglomeration
(Gomes et al. 2011; Saleh et al. 2010b), so a more stable recycle rate of 50% is
recommended in tandem with appropriate purging (Cao et al. 2008b). The mem-
brane bioreactor technology was applied to a number of feedstocks, including
soybean oil, canola oil, and yellow and brown grease (Cao et al. 2008a). Even with
high FFA feedstocks, no emulsion was experienced in the permeate flows, and the
distinct two-phase flow allowed for collection of a FAME-rich phase that far out-
performed a conventional batch reactor at the same conditions, as seen in Table 4.3
(Cao et al. 2008a). The soybean and canola oil FAME produced by the membrane
reactor could meet ASTM standards without additional water washing, demon-
strating significant successes in membrane bioreactor technology (Cao et al. 2008a).
The amount of catalyst utilized in the membrane reactor is significant and must
be carefully tuned so as to convert all of the feedstock (Tremblay et al. 2008)
without excess saponification (Baroutian et al. 2011). In the case of a packed bed of
potassium hydroxide on activated carbon, Baroutian et al. (2011) tested catalyst
4 Biodiesel Purification and Upgrading Technologies 79

Fig. 4.8 Membrane reactor design for continuous biodiesel transesterification and separation.
Note the “Polar phase” used to recycle methanol from a FAME-rich phase. Reprint from reference
(Cao et al. 2008a) with permission from Elsevier

Table 4.3 Comparisons in glycerol content between membrane and batch reactions at 65 °C.
Reprint from reference (Cao et al. 2008a) with permission from Elsevier
Lipid feedstock Biodiesel from membrane Biodiesel from batch
reactor (wt%) reaction (wt%)
Total Free Total Free
glycerine glycerine glycerine glycerine
Soybean 0.0685 0.00763 – –
Canola 0.0712 0.00654 0.131 0.0124
Palm 0.124 0.0117 – –
Yellow grease 0.0989 0.00735 0.685a 0.0254a
Brown grease 0.104 0.0138 0.797a 0.0171
Canola oil with methanol 0.0929 0.00749 – –
recycle
a
Does not meet the ASTM standard

loadings between 37.5 and 250 mg/cm3; increasing the catalyst loading from 37.5
to 143.75 mg/cm3 had a positive effect on conversion. Further increases beyond this
point had slight negative effects due to the creation of excess soaps that inhibited
the transesterification reaction (Baroutian et al. 2011). In the case of feeding catalyst
alongside an oil/methanol loading, Tremblay et al. (2008) conducted experiments
80 H. Bateni et al.

with a canola oil feedstock and ultralow catalyst loadings, down to 0.01% NaOH by
weight. Loadings below 0.05 wt% NaOH were unable to complete the transester-
ification reaction in a 1 h residence time, prohibiting continuous reactor operation;
further increases to the catalyst loading up to 1 wt% had insignificant effects on
FAME concentration in the permeate (from 56.6% at 0.05 wt% to 59.8% at 1 wt%)
(Tremblay et al. 2008). These results stress the importance of carefully tuning both
external and internal catalyst loadings to ensure economically viable and continu-
ous reactor operation.
As well as considering the catalyst loading, the residence time must be chosen
carefully so as to ensure that the reaction is carried to completion; if unreacted oil is
permitted to build up inside of the reactor, uncontrollable increases in transmem-
brane pressure can occur (Tremblay et al. 2008; Falahati and Tremblay 2012). In
the case of canola oil, Falahati and Tremblay (2012) found dramatic transmembrane
pressure increases at 35 min residence times; no such phenomenon occurred for 60
and 82 min residence times, indicating the necessity of at least 60 min for canola
oil. In tuning the catalyst loading, Tremblay et al. (2008) experienced transmem-
brane pressure buildup for catalyst loadings of 0.03 wt% NaOH at a 1 h residence
time for canola oil feed; increasing the residence time to 2 h prevented this phe-
nomenon. These increases in transmembrane pressure will directly lead to fouling
of the membrane if not properly handled (Hua et al. 2007).

4.2 Biodiesel Upgrading

Producing transportation fuels from biorenewable sources is an attractive option for


replacing nonrenewable resources and mitigating their environmental ramifications.
Among biorenewable materials, nonedible oils and fats are promising feedstocks
primarily because they do not raise the “food versus fuel” debate (Chhetri et al.
2008; Srinivasan 2009; Escobar et al. 2009; Greenwell et al. 2010; Zhao et al.
2013). Moreover, producing diesel fuel (C12–C18) is feasible from these materials
due to high availability, low price, and their appropriate hydrocarbon chain length.
Thus, several techniques such as thermal and catalytic cracking, hydrocracking,
emulsification, and transesterification have been employed to obtain diesel fuels
from these feedstocks (Parvizsedghy et al. 2015; Twaiq et al. 1999; Ma and Hanna
1999; Snåre et al. 2009; Schuchardt et al. 1998; Kim et al. 2004; Demirbas 2008).
Most of these techniques suffer from low energy efficiency and products with poor
cold flow and combustion properties, chemical instability, and corrosiveness.
Among the above-mentioned methods, transesterification of triglycerides with
alcohols produces fatty acid alkyl esters (FAEE) (i.e., “biodiesel”) that could be
used in modified engines by blending with petroleum diesel or as a pure biofuel
(Schuchardt et al. 1998; Kim et al. 2004; Demirbas 2008). Although, this fuel still
has many problems including moisture absorption and low power density mainly
rooting from its high oxygen content (Srifa et al. 2015). Alternatively, catalytic
upgrading of nonedible oils and fats could produce long-chain aliphatic
4 Biodiesel Purification and Upgrading Technologies 81

Fig. 4.9 The main possible reaction pathways for converting triglycerides to diesel fuels and
chemicals. Reprint from reference (Bateni et al. 2017) with permission from Biofuel Research
Journal

hydrocarbons (i.e., “renewable diesel”), enabling direct substitution of petroleum


diesel with biorenewable fuels (Knothe 2010). In this chapter, biodiesel upgrading
refers to the conversion of oils, fats, fatty acids, and triglycerides into hydrocarbon
products using thermochemical and catalytic processes.
To date, several processing techniques at various severities have been proposed
for upgrading bio-based oils and fats. Figure 4.9 represents some of the potential
routes for converting triglycerides into biorenewable diesel fuel and chemicals. This
scheme, however, does not provide any mechanistic insights into deoxygenation
pathways as scrutinized in other reviews (Gosselink et al. 2013a; Rogers and Zheng
2016). Moreover, it should be noted that the oxygen content of bio-feeds (e.g.,
triglycerides) is much higher than their nitrogen or sulfur content. These heteroa-
toms can be removed during deoxygenation processes (Bezergianni et al. 2010a).
As a result, an effective deoxygenation process would guarantee denitrogenation
and desulfurization of biorenewable diesel fuels. The focus of this chapter is pri-
marily deoxygenation processes, whereas nitrogen and sulfur removal processes
have been reviewed elsewhere (Yang et al. 2016; Hanafi and Mohamed 2011).
Among upgrading routes shown in Fig. 4.9, cracking is an unappealing method
due to low carbon efficiency. A significant amount of carbon is lost due to the
production of non-condensable gases and solid residues (coke) as typically
observed in solid acid catalysis processes (Sotelo-Boyas et al. 2010; Dupain et al.
2007; Vinh et al. 2011). Deoxygenation, on the other hand, is a promising pathway
in which a product with high carbon efficiency, maintaining a similar carbon
number. The product is a long-chain aliphatic hydrocarbon with fuel properties
similar to or better than petroleum diesel fuels (Snåre et al. 2009; Kiatkittipong et al.
82 H. Bateni et al.

2013; Phimsen et al. 2016). Also, maximizing unsaturated hydrocarbons may


enable the production of higher value chemicals in the future (Hollak et al. 2013).
Deoxygenation can be conducted under an inert atmosphere (e.g., Ar, He, or N2)
or in the presence of H2. Supported metallic catalysts are typically able to
deoxygenate bio-based oils and fats under inert environments (Snåre et al. 2006,
2008; Morgan et al. 2012). However, this process suffers from low conversion, low
selectivity to target molecules, quick catalyst coking and deactivation, and leaching
or sintering of the active species (Lestari et al. 2009a; Snåre et al. 2008).
Deoxygenation under these conditions occurs predominantly via decarbonylation
(DCO) and decarboxylation (DCO2). Introducing H2 into the reactor not only
suppresses undesired side reactions but also promotes hydrodeoxygenation
(HDO) as a competing reaction pathway (Madsen et al. 2011). HDO does not
release CO and CO2 and yields a product with the same carbon number as the
starting material, which is advantageous from both environmental and carbon atom
efficiency perspectives. Of course, these benefits come at the expense of costly
hydrogen. This is even more highlighted when triglyceride molecules contain
several C=C double bonds in their structure, e.g., rapeseed oil (Sotelo-Boyas et al.
2010). In this case, deoxygenation under H2 would lead to saturation of all double
bonds prior to removing oxygen atoms, causing excessive hydrogen consumption
(Han et al. 2011a; Coumans and Hensen 2017). On the other hand, deoxygenating
under inert atmospheres could yield hydrocarbons with unsaturated bonds (Snåre
et al. 2008).
In addition to the nature of the feedstock, several other parameters could affect
the deoxygenation pathway (Kiatkittipong et al. 2013). Operating conditions such
as temperature, pressure, and residence time as well as the presence or absence of
H2 are clearly able to determine many aspects of reactions. The choice of catalyst
can also significantly influence deoxygenation by promoting a certain route. For
example, transition metal sulfides typically favor the HDO pathway, whereas
carbon-supported noble metals generally prefer the DCO and DCO2 routes. The
addition of H2, on the other hand, has been found to reduce catalyst deactivation by
suppressing undesired side reactions that can lead to coking, and improve the
kinetics of deoxygenation (Immer et al. 2010; Rozmysłowicz et al. 2012).
Therefore, deoxygenation appears to proceed more sustainably and efficiently under
H2 environments, although the price of hydrogen might be a constraint for
commercialization.

4.2.1 Catalysts

Multiple categories of catalysts have been used for upgrading bio-oils, comprised of
conventional hydrotreating catalysts (transition metal sulfides), supported mono-
and bi-metallic catalysts, and novel deoxygenation catalysts including transition
metal carbides, nitrides, and phosphides. Hydrotreating catalysts are interesting
materials due to their industrial success in removing heteroatoms (mainly sulfur,
4 Biodiesel Purification and Upgrading Technologies 83

nitrogen, and oxygen) from petroleum fractions. Supported metals, in particular,


carbon-supported noble metals, are also active catalysts for a variety of reactions
including hydrogenation, dehydrogenation, and deoxygenation. Recently, transition
metals in various forms have gained a lot of attention as cheap catalysts that are able
to promote HDO reactions. Furthermore, the importance of support materials in
enhancing catalytic activity, selectivity, and stability is discussed in this section.

4.2.1.1 Transition Metal Sulfides

Hydrotreating catalysts are comprised of supported and promoted transition metal


sulfides. Molybdenum and tungsten are the two main transition metals used as
active phases. Alumina is the most widely used support material and nickel and
cobalt can be added as promoters. Hydrotreating is typically done at high hydrogen
pressures and temperatures to remove sulfur, nitrogen, oxygen, and trace metals
from petroleum cuts and improve fuel properties (Parkash 2003). Desulfurization
and denitrogenation are also very important from an environmental viewpoint
because of the detrimental effects of SOx and NOx emissions (Parkash 2003;
Rostrup-Nielsen 2008).
Transition metal sulfides have been also studied for deoxygenating bio-based
oils and fats with oxygen content as high as 40 wt%, as well as various model
compounds (Mortensen et al. 2011). The mechanism by which these catalysts
operate is depicted in Fig. 4.10 (Romero et al. 2010). Essentially, H2 is required to

Fig. 4.10 HDO mechanism of 2-ethylphenol, a model compound, on a MoS2 surface.


Reproduced from (Romero et al. 2010) with permission from Elsevier
84 H. Bateni et al.

react with a surface sulfur atom, yielding H2S and a vacant site. Oxygenates can
then chemisorb on this vacant site and react to produce deoxygenated molecules
and an oxidized catalyst surface. The Mo–O species on the surface are not as active
as Mo–S for hydrotreating (Şenol et al. 2005a). Therefore, these catalysts need the
continuous co-feeding of a sulfur source in order to maintain their active sulfide
phase (Furimsky 2000). However, this deprives the final biofuel product from its
sulfur-free nature (Badawi et al. 2011). The presence of H2 enhances the possibility
of HDO over DCO/DCO2 pathways when using these catalysts, but can also
hydrogenate the unsaturated bonds, which results in hydrogen overconsumption.
One of the key benefits of transition metal sulfides is their robustness under
hydrotreating conditions, which is the main reason for their vast industrial use since
their realization decades ago. These catalysts are cheap, well-characterized, easy to
handle, and require minimal maintenance (Ruddy et al. 2014). In addition, they
have been found to be active for deoxygenation reactions, particularly via the HDO
pathway (Şenol et al. 2005b). Thus, they can produce hydrocarbon products with
high carbon atom efficiency. However, these catalysts suffer from high sensitivity to
water, which exists basically in all bio-based feedstocks (Şenol et al. 2005a). They
are active only under high hydrogen pressures and elevated temperatures, needing
expensive compressors and reactors. In addition, coking issues associated with
hydrotreating catalysts require frequent regeneration cycles, which—similar to the
previous point—increases the costs related to producing fuels and chemicals from
biorenewable feedstocks (Furimsky and Massoth 1999). Furthermore, biofuels
obtained by hydrotreating could have unacceptable sulfur levels because a sulfur
source must be used for maintaining the catalyst’s active phase (Stanislaus et al.
2010). All of these problems have encouraged researchers to seek alternative cat-
alytic systems for producing diesel biofuels from bio-based oils and fats.

4.2.1.2 Supported Mono- and Bi-metallic Catalysts

Supported metallic catalysts are promising alternatives for producing biofuels from
renewable feedstocks because they are active for hydrogenation and deoxygenation
reactions at low to moderate temperatures and do not require the constant presence of
a sulfur-containing compound. These catalysts primarily include supported noble
metals (e.g., Pt, Pd, and Rh) and base metals (e.g., Ni). Murzin and coworkers have
studied the deoxygenation of various model compounds and some bio-based oils
over supported metallic catalysts (Kubičková et al. 2005; Snåre et al. 2006, 2007,
2008; Mäki-Arvela et al. 2007, 2008; Lestari et al. 2008, 2009b, c, 2010; Simakova
et al. 2009, 2011; Bernas et al. 2010). These studies clearly indicate the effectiveness
of supported metals for deoxygenation reactions. In particular, carbon-supported Pd
has been shown to produce high yields of diesel-like fuel from practical biore-
newable resources such as palm oil (Kiatkittipong et al. 2013).
Depending on the nature of support, these catalysts can proceed through the
DCO/DCO2 or HDO pathways. Inert supports such as carbon typically promote
DCO/DCO2 and produce hydrocarbon products with one less carbon atom. In the
4 Biodiesel Purification and Upgrading Technologies 85

presence of acidic supports, noble metals act as hydrogenation or hydrogenolysis


catalysts, whereas the support performs dehydration reactions, eventually removing
oxygen atoms via the HDO route. Regardless of the deoxygenation mechanism, the
addition of hydrogen significantly decreases catalyst deactivation by promoting
hydrogenation of coke forming material (Zhu et al. 2011). Supported metallic
catalysts have good hydrothermal stability that is an important feature for pro-
cessing biomaterials (Ruddy et al. 2014). However, they are sensitive to heteroa-
toms such as nitrogen, sulfur, and metals present in bio-based feedstocks (Nagy
et al. 2009; Choudhary and Phillips 2011; Ardiyanti et al. 2011). They can also lose
their activity due to leaching of the active phase into the reaction medium and/or
sintering of smaller metal particles into large particles and consequent reduction of
the active species (Han et al. 2011b). Some methods have been proposed to mitigate
these issues, but eventually, the paucity of most of these metallic species could
increase the cost of processing, making renewable diesel production economically
infeasible (Nagy et al. 2009).
Model compound studies with fatty acids and fatty acid esters showed that
supported noble metals had higher activity and selectivity to target molecules than
transition metal sulfide catalysts. However, their activity and selectivity were sig-
nificantly lowered in the case of real feedstocks, i.e., bio-oils (Boda et al. 2010;
Madsen et al. 2011; Phimsen et al. 2016; Kiatkittipong et al. 2013). The rationale
for such findings cannot be found in the literature and requires more attention.
Among various metals, Pd and Co are promising candidates because of their high
intrinsic activity for deoxygenating fatty acids (Srifa et al. 2015). Similar to pre-
vious findings, deoxygenation reactions occur mainly via DCO/DCO2 and HDO
over Pd and Co catalysts, respectively (Srifa et al. 2015; Phimsen et al. 2016). The
promising results for deoxygenating bio-based oils and fats over supported mono-
and bi-metallic catalysts call for further research to overcome problems such as
stability, longevity, and leaching/sintering. Solving these issues would enable the
engineering of profitable systems for producing cost-effective biorenewable diesel
fuels.

4.2.1.3 Novel Deoxygenation Catalysts

Recently, a new category of catalysts composed of carbides, nitrides, and phos-


phides of transition metals have appeared promising for converting bio-oils into
diesel-like hydrocarbons. Molybdenum and tungsten carbides were initially pre-
pared as inexpensive substitutes for noble metals because of their similar catalytic
behavior. These transition metal carbides are active hydrodeoxygenation catalysts
with great robustness and stability and high selectivity to the HDO pathway (Qin
et al. 2013; Han et al. 2011a, b, 2012; Hollak et al. 2013; Stellwagen and Bitter
2015; Gosselink et al. 2013b; Wang et al. 2013; Sousa et al. 2012). In particular, in
comparison with W2C, Mo2C was found to be more active and stable for the
conversion of oleic acid into n-octadecane (Hollak et al. 2013). Mo2C’s higher
stability could be rooted in its resistance to oxidation and sintering, as well as its
86 H. Bateni et al.

hydrogenation ability. Mo2C yielded hydrogenation products, i.e., stearic acid,


followed by deoxygenation to n-octadecane, whereas W2C produced octadecene
(probably due to the presence of WOx domains) followed by slow saturation to n-
octadecane (Hollak et al. 2013; Stellwagen and Bitter 2015). Both of these catalysts
were found to be more stable and robust than typical noble metal catalysts (Han
et al. 2011b).
Transition metal nitrides are also interesting materials for deoxygenating
bio-oils. Similar to metal carbides, they can perform hydrogenation reactions, but
they also exhibit acidic properties (Miga et al. 1999). Therefore, metal nitrides have
been utilized as hydrotreating, hydrogenation, dehydrogenation, and isomerization
catalysts for decades (Schlatter et al. 1988; Neylon et al. 1999). However, these
catalysts are rarely studied for deoxygenation reactions in a systematic manner. For
instance, ZSM-5-supported NiMoN and NiMoC were used for deoxygenation of
soybean oil (Wang et al. 2012). Clearly, the use of an acidic support such as ZSM-5
could make drawing any conclusions about the performance of the metal nitride
catalyst very difficult. In addition, Mo2N and Mo2C were not tested as a baseline to
see the effect of Ni addition on catalytic performance. Nonetheless, Mo2C exhibited
better catalytic properties under the conditions tested in this study (Wang et al.
2012). Among nitrides of tungsten, vanadium, and molybdenum, the latter had the
highest activity for conversion of canola oil to diesel-like hydrocarbons (Monnier
et al. 2010). Alumina-supported molybdenum nitride was extremely selective to
fully saturated n-alkanes and active for deoxygenation even after 450 h under the
reaction conditions.
Lately, transition metal phosphides have been found to be active catalysts for
conversion of bio-oils into fuels at conditions similar to hydrotreating reactions
(Yang et al. 2012, 2013a, 2015; Chen et al. 2014; Zarchin et al. 2015; Peroni et al.
2016; Gong et al. 2012). In a comparison between Ni2P/SBA-15 and its Ni
counterpart in a fixed-bed reactor, both catalysts were able to effectively deoxy-
genate methyl oleate (Yang et al. 2012). However, a Ni2P catalyst yielded over 50%
HDO products (i.e., n-octadecane), while a Ni catalyst almost exclusively yielded
C17 products (Yang et al. 2012). Hydrotreatment of jatropha oil was studied in a
fixed-bed reactor using PtPd/Al2O3 and NiMoP/Al2O3 at 330–390 °C, 3 MPa, and
a weight hourly space velocity of 2 h−1. The transition metal nitride behaved very
similar to the noble metal catalyst in that they both had high yields of liquid
hydrocarbons and excellent fuel properties. NiMoP proceeded through the HDO
pathway, whereas PtPd deoxygenated the oil via DCO/DCO2. The nitride catalyst
started losing its activity after 120 h on-stream, although deactivation was not
investigated in the case of the PtPd catalyst.
In general, the findings of the literature suggest the potential of these catalysts in
deoxygenation reactions. However, the field clearly lacks studies with direct
comparison of transition metal sulfides, supported metallic catalysts, and these
newly proposed catalysts. Additionally, several other factors including leaching/
sintering, stability, and robustness have not been addressed.
4 Biodiesel Purification and Upgrading Technologies 87

4.2.1.4 Catalytic Supports

Alumina materials are the most widely used supports in chemical industries owing
to their robustness, shapeable nature, tunable surface area, and Lewis acidity. They
offer great versatility for a wide range of applications including hydrotreating
processes. However, these supports are sensitive to hydrothermal conditions under
which they undergo recrystallization and experience significant surface area losses
(Laurent and Delmon 1994). Other metal oxides such as titania, zirconia, and ceria
are also considered interesting support materials due to their reducibility. Titania
and zirconia are oxophilic supports, whereas ceria is an electronegative material.
Zeolites are another class of acidic supports used for a variety of reactions such
as dehydration, cracking, and isomerization. Zeolites are shape selective and con-
tain both types of acid sites, i.e., Brønsted and Lewis (Liu et al. 2011). Silica
materials have weak acid sites that are essentially inept for deoxygenation reactions
and can be used in a variety of forms as catalytic supports. SiO2 is typically used to
study the effect of active materials without interference from a support (Liu et al.
2011). Mesoporous silicates are high surface area materials with pores large enough
for diffusion of bulky molecules such as triglycerides and fatty acids (Stöcker
2008). The catalytic properties of silica supports could be tuned by the type and
amount of metals incorporated in their structure (Nava et al. 2009). For example, Al
incorporation can introduce acid sites into the silica structure, whereas Ti could add
reducibility to the support. Several studies have shown the positive effect of
mesoporous silicates on the catalytic activity of deoxygenation reactions for model
compounds and real feedstocks (Wang et al. 2013; Kandel et al. 2014; Lestari et al.
2010).
Another class of supports that have proven useful in deoxygenation studies is
carbon supports. These supports are relatively inexpensive and have interesting
properties such as tunable functionality, surface area, and porosity
(Rodríguez-Reinoso 1998). Functionalized carbon, nanotubes and nanofibers of
carbon and mesoporous or microporous carbon have all been used in various
studies (Lam and Luong 2014). Carbon supports have shown better properties than
conventional alumina and silica supports for deoxygenation of fatty acids and fatty
acid esters (Han et al. 2011a). For instance, a high resistance to coking and sintering
was observed in the case of Pd/C as compared to Pd/Al2O3 for conversion of stearic
acid at 6 bar pressure and 300 °C (Snåre et al. 2006).

4.2.2 Operating Conditions

Operating conditions including temperature, pressure, and contact time have a


substantial effect on the performance of the upgrading process and consequently the
economy of the process.
Temperature needs to be carefully controlled due to its controversial effects on
the reaction. Increasing the temperature can often increase the rate of the reaction;
88 H. Bateni et al.

however, it may also decrease the selectivity of the process. In the case of
deoxygenation, increasing the temperature is favorable for the process as it
improves both the kinetics and degree of deoxygenation (Bezergianni et al. 2010a;
Şenol et al. 2005b). However, at temperatures above 400 °C, the yield may
decrease due to the formation of undesired by-products via cracking or isomer-
ization (Bezergianni et al. 2010b; Kiatkittipong et al. 2013; Phimsen et al. 2016).
Even though isomerization improves the cold flow properties of the final diesel/
biodiesel, it can decrease its cetane number (Bezergianni et al. 2010b). Higher
temperatures may also compromise the carbon efficiency due to the superior
influence on DCO compared to HDO (Yang et al. 2013b; Kubička and Kaluža
2010; Itthibenchapong et al. 2017).
Pressure (i.e., hydrogen pressure) positively influences the biodiesel upgrading
process due to facilitating hydrogenolysis and hydrogenation reactions (Zhao et al.
2013; Madsen et al. 2011), improving the stability and activity of the catalysts by
preventing coke formation on the catalyst via hydrogenation of coke precursors
(Immer et al. 2010; Rozmysłowicz et al. 2012), suppressing side reactions such as
cracking and decarboxylation leading to higher carbon yield (Kandel et al. 2014; Do
et al. 2009), and improving the kinetics of deoxygenation in general (Snåre et al. 2007;
Lestari et al. 2009b). However, handling high-pressure hydrogen in a large-scale
system is still a significant challenge for industrial sectors in terms of reactor, com-
pressor, and storage design leading to high operating and capital costs.
Contact time is another variable which may have controversial effects as longer
reaction time improves the conversion (Foraita et al. 2017; Immer et al. 2010;
Hollak et al. 2013) but cannot improve selectivity due to potential cracking and
DCO reactions (Phimsen et al. 2016; Kiatkittipong et al. 2013). Studying the effect
of time on the reaction performance in a batch or semi-batch system can provide
valuable insights into the reaction pathway along with great potential for optimizing
the process (Foraita et al. 2017; Stellwagen and Bitter 2015).
Sulfidation of conventional hydrotreating catalysts, usually containing transition
metal sulfides, is another important subject to discuss in biodiesel upgrading due to
the significant effect of sulfidation on the activity, selectivity, and stability of the
catalyst (Kubička and Horáček 2011; Coumans and Hensen 2017). In fact,
co-feeding a sulfiding agent (such as H2S) into the system can prevent the oxidation
of the active sites available in sulfide form (Kubička and Horáček 2011; Furimsky
and Massoth 1999).
It is worth mentioning that the reactor design, feedstock properties, and the
choice of solvents in a batch system can also affect the biodiesel upgrading process;
even though the effect may not be as significant.

4.2.3 Challenges and Potential

Finding nonedible biomass materials with a proper hydrocarbon chain length (C12–
C18) is indeed the greatest challenge in producing biorenewable diesel (Srinivasan
4 Biodiesel Purification and Upgrading Technologies 89

2009; Escobar et al. 2009). Therefore, waste oils, animal fats, nonedible vegetable
oils, and energy crops (e.g., algae) are among the most appropriate feedstocks. In
addition, the choice of biomass and optimal utilization of water and nutrients are
extremely important for growing high yields of biomass, specifically in countries
with severe climate conditions and limited water resources.
From a processing perspective, the most difficult challenge is in designing cheap,
stable, and robust catalysts active for deoxygenation of triglyceride molecules
present in bio-based oils and fats. The choice of appropriate model compounds can
also help in determining the effectiveness of catalysts (Gosselink et al. 2013a).
Ideally, other components in these feedstocks, such as proteins and mineral com-
pounds, should not deactivate the catalyst in order to avoid any pretreatment
requirements. In addition, the catalyst’s ability in producing renewable diesel with
excellent fuel properties (including low sulfur levels, high cetane number, and good
cold flow properties) in a single processing step is critical for commercial imple-
mentation. While normal long-chain alkane hydrocarbons could be used for cetane
enhancers in petroleum diesel fuels, isomerization would enhance cold flow
properties and enable the production of diesel substitutes.
Fortunately, the production of renewable diesel with excellent fuel properties is
already commercialized (Table 4.4), which can pave the way for future endeavors
on optimizing these systems and designing better catalysts (Phimsen et al. 2016;
Satyarthi et al. 2013). Several companies including Neste Oil, British Petroleum,
Conoco Phillips and Petrobras have invested in this area (Rantanen et al. 2005;
Aslam et al. 2015). However, current commercial processes are based on cascading
reactions, i.e., deoxygenation followed by isomerization. In order to produce
high-quality renewable diesel economically, it might be beneficial to design cata-
lysts capable of deoxygenation and isomerization in a one-step process. The
addition of hydrogen might also be advantageous for preserving catalytic activity
and obtaining higher yields of liquid products. Furthermore, utilizing by-products
such as unsaturated long-chain hydrocarbons or gases might help the economics
and carbon atom efficiency, making renewable diesel even more attractive.

Table 4.4 The fuel properties of Honeywell Renewable diesel versus petroleum diesel and
biodiesel (UOP 2017)
Petroleum ultralow sulfur Biodiesel fatty acid Renewable
diesel methyl ester diesel
Oxygen (%) 0 11 0
Cetane number 40–55 50–65 75–90
Energy density 43 38 44
(MJ/kg)
Sulfur (ppm) <10 <2 <2
Cold flow Baseline Poor Excellent
properties
Oxidative stability Baseline Poor Excellent
90 H. Bateni et al.

4.3 Conclusions

Biodiesel purification is a crucial step prior to storage and marketing to meet


standard specifications. The biodiesel purification methods can be categorized as
equilibrium-based, affinity-based, membrane technology, reaction-based, and solid–
liquid separation processes. Wet washing techniques using water or acidified water
is an equilibrium-based method which has been frequently used for biodiesel
purification. However, the challenges associated with this method have motivated
scientists to develop the purification techniques which mitigate water consumption.
Current problems with biodiesel fuels have stimulated researchers to explore
efficient upgrading methods, such as deoxygenation and hydrodeoxygenation, as
well as catalytic systems that are able to effectively convert oxygenated oils and fats
into diesel-like hydrocarbons. However, other than finding feedstocks that do not
interfere with the food chain, the greatest challenge is to design cheap and robust
catalysts capable of producing high-quality diesel fuel in a one-step process.
Nonetheless, the economic viability of diesel biofuel greatly depends on fossil fuel
prices and incentives from governments.

References

ASTM D6751 (2015) Standard specification for biodiesel fuel blend stock (B100) for middle
distillate fuels. ASTM International, West Conshohocken, PA
EN 14214 (2008) Automotive fuels—fatty acid methyl esters (fame) for diesel engines—
requirements and test methods. European Committee for Standardization, Brussels, Belgium
Abbaszadeh A, Ghobadian B, Najafi G, Yusaf T (2014) An experimental investigation of the
effective parameters on wet washing of biodiesel purification. Int J Autom Mech Eng 9:1525–
1537
Abbott AP, Cullis PM, Gibson MJ, Harris RC, Raven E (2007) Extraction of glycerol from
biodiesel into a eutectic based ionic liquid. Green Chem 9:868–872
Alba-Rubio A, Castillo MA, Albuquerque M, Mariscal R, Cavalcante C, Granados ML (2012) A
new and efficient procedure for removing calcium soaps in biodiesel obtained using CaO as a
heterogeneous catalyst. Fuel 95:464–470
Alicieo T, Mendes E, Pereira N, Lima OM (2002) Membrane ultrafiltration of crude soybean oil.
Desalination 148:99–102
Alves MJ, Nascimento SM, Pereira IG, Martins MI, Cardoso VL, Reis M (2013) Biodiesel
purification using micro and ultrafiltration membranes. Renew Energy 58:15–20
Ardiyanti A, Gutierrez A, Honkela M, Krause A, Heeres H (2011) Hydrotreatment of wood-based
pyrolysis oil using zirconia-supported mono-and bimetallic (Pt, Pd, Rh) catalysts. Appl Catal A
407:56–66
Armor J (1998) Applications of catalytic inorganic membrane reactors to refinery products.
J Membr Sci 147:217–233
Aslam M, Kothiyal N, Sarma A (2015) True boiling point distillation and product quality
assessment of biocrude obtained from Mesua ferrea L. seed oil via hydroprocessing. Clean
Technol Environ Policy 17:175–185
Atadashi I (2015) Purification of crude biodiesel using dry washing and membrane technologies.
Alexandria Engineering Journal 54:1265–1272
4 Biodiesel Purification and Upgrading Technologies 91

Atadashi I, Aroua M, Aziz AA (2011a) Biodiesel separation and purification: a review. Renew
Energy 36:437–443
Atadashi I, Aroua M, Aziz AA, Sulaiman N (2011b) Membrane biodiesel production and refining
technology: a critical review. Renew Sustain Energy Rev 15:5051–5062
Atadashi I, Aroua M, Aziz AA, Sulaiman N (2011c) Refining technologies for the purification of
crude biodiesel. Appl Energy 88:4239–4251
Atadashi I, Aroua M, Aziz AA, Sulaiman N (2012) High quality biodiesel obtained through
membrane technology. J Membr Sci 421:154–164
Badawi M, Paul J, Cristol S, Payen E, Romero Y, Richard F, Brunet S, Lambert D, Portier X,
Popov A (2011) Effect of water on the stability of Mo and CoMo hydrodeoxygenation
catalysts: a combined experimental and DFT study. J Catal 282:155–164
Baroutian S, Aroua MK, Raman AAA, Sulaiman NM (2011) A packed bed membrane reactor for
production of biodiesel using activated carbon supported catalyst. Biores Technol
102:1095–1102
Barredo-Damas S, Alcaina-Miranda M, Bes-Piá A, Iborra-Clar M, Iborra-Clar A, Mendoza-Roca J
(2010) Ceramic membrane behavior in textile wastewater ultrafiltration. Desalination
250:623–628
Basso RC, Viotto LA, Gonçalves LAG (2006) Cleaning process in ceramic membrane used for the
ultrafiltration of crude soybean oil. Desalination 200:85–86
Bateni H, Bateni F, Karimi K (2016) Effects of oil extraction on ethanol and biogas production
from Eruca sativa seed cake. In: Waste and biomass valorization. https://doi.org/10.1007/
s12649-016-9731-x
Bateni H, Karimi K (2016a) Biodiesel production from castor plant integrating ethanol production
via a biorefinery approach. Chem Eng Res Des 107:4–12
Bateni H, Karimi K (2016b) Biorefining of Eruca sativa plant for efficient biofuel production. RSC
Adv 6:34492–34500
Bateni H, Karimi K, Zamani A, Benakashani F (2014) Castor plant for biodiesel, biogas, and
ethanol production with a biorefinery processing perspective. Appl Energy 136:14–22
Bateni H, Saraeian A, Able C (2017) A comprehensive review on biodiesel purification and
upgrading. Biofuel Res J 15:559–612. https://doi.org/10.18331/BRJ2017.4.3.4
Bernas H, Eränen K, Simakova I, Leino A-R, Kordás K, Myllyoja J, Mäki-Arvela P, Salmi T,
Murzin DY (2010) Deoxygenation of dodecanoic acid under inert atmosphere. Fuel
89:2033–2039
Berrios M, Siles J, Martín M, Martín A (2013) Ion exchange. In: Ramaswamy S, Huang H-J,
Ramarao BV (eds) Separation and purification technologies in biorefineries. Wiley, Chichester,
United Kingdom
Berrios M, Skelton R (2008) Comparison of purification methods for biodiesel. Chem Eng J
144:459–465
Bezergianni S, Dimitriadis A, Kalogianni A, Pilavachi PA (2010a) Hydrotreating of waste cooking
oil for biodiesel production. Part I: Effect of temperature on product yields and heteroatom
removal. Biores Technol 101:6651–6656
Bezergianni S, Dimitriadis A, Sfetsas T, Kalogianni A (2010b) Hydrotreating of waste cooking oil
for biodiesel production. Part II: Effect of temperature on hydrocarbon composition. Biores
Technol 101:7658–7660
Bhayani BV, Ramarao BV (2013) Filtration-based separations in the biorefinery. In:
Ramaswamy S, Huang H-J, Ramarao BV (eds) Separation and purification technologies in
biorefineries. Wiley, Chichester, United Kingdom
Boda L, Onyestyák G, Solt H, Lónyi F, Valyon J, Thernesz A (2010) Catalytic hydroconversion of
tricaprylin and caprylic acid as model reaction for biofuel production from triglycerides. Appl
Catal A 374:158–169
Canakci M, van Gerpen J (2003) A pilot plant to produce biodiesel from high free fatty acid
feedstocks. Trans ASAE 46:945–954
92 H. Bateni et al.

Cao P, Dubé MA, Tremblay AY (2008a) High-purity fatty acid methyl ester production from
canola, soybean, palm, and yellow grease lipids by means of a membrane reactor. Biomass
Bioenerg 32:1028–1036
Cao P, Dubé MA, Tremblay AY (2008b) Methanol recycling in the production of biodiesel in a
membrane reactor. Fuel 87:825–833
Cao P, Tremblay AY, Dubé MA, Morse K (2007) Effect of membrane pore size on the
performance of a membrane reactor for biodiesel production. Ind Eng Chem Res 46:52–58
Chen B, Wang W, Liu X, Xue W, Ma X, Chen G, Yu Q, Li R (2012a) Adsorption study of
glycerol in biodiesel on the sulfonated adsorbent. Ind Eng Chem Res 51:12933–12939
Chen B, Wang W, Ma X, Wang C, Li R (2012b) Adsorption behaviors of glycerol from biodiesel
on sulfonated polystyrene-divinylbenzene resins in different forms. Energy Fuels
26:7060–7067
Chen J, Yang Y, Shi H, Li M, Chu Y, Pan Z, Yu X (2014) Regulating product distribution in
deoxygenation of methyl laurate on silica-supported Ni–Mo phosphides: Effect of Ni/Mo ratio.
Fuel 129:1–10
Cheng L-H, Cheng Y-F, Yen S-Y, Chen J (2009) Ultrafiltration of triglyceride from biodiesel
using the phase diagram of oil–FAME–MeOH. J Membr Sci 330:156–165
Chhetri AB, Watts KC, Islam MR (2008) Waste cooking oil as an alternate feedstock for biodiesel
production. Energies 1:3–18
Choudhary T, Phillips C (2011) Renewable fuels via catalytic hydrodeoxygenation. Appl Catal A
397:1–12
Coêlho DG, Almeida AP, Soletti JI, de Carvalho SH (2011) Influence of variables in the
purification process of castor oil biodiesel. Chem Eng Trans 24:829–834
Coumans A, Hensen E (2017) A model compound (methyl oleate, oleic acid, triolein) study of
triglycerides hydrodeoxygenation over alumina-supported NiMo sulfide. Appl Catal B
201:290–301
da Silva NDL, Santander CMG, Batistella CB, Maciel Filho R, Maciel MRW (2010) Biodiesel
production from integration between reaction and separation system: reactive distillation
process. Appl Biochem Biotechnol 161:245–254
Dechow FJ (1989) Separation and purification techniques in biotechnology. Noyes Publications,
Park Ridge, United States
Demirbas A (2008) Comparison of transesterification methods for production of biodiesel from
vegetable oils and fats. Energy Convers Manag 49:125–130
Demirbas A (2009) Progress and recent trends in biodiesel fuels. Energy Convers Manag 50:14–34
Deroussel P, Khakhar D, Ottino J (2001) Mixing of viscous immiscible liquids. Part 2:
Overemulsification—interpretation and use. Chem Eng Sci 56:5531–5537
Dias J, Santos E, Santo F, Carvalho F, Alvim-Ferraz M, Almeida M (2014) Study of an ethylic
biodiesel integrated process: raw-materials, reaction optimization and purification methods.
Fuel Process Technol 124:198–205
Dimian AC, Bildea CS, Omota F, Kiss AA (2009) Innovative process for fatty acid esters by dual
reactive distillation. Comput Chem Eng 33:743–750
Do PT, Chiappero M, Lobban LL, Resasco DE (2009) Catalytic deoxygenation of
methyl-octanoate and methyl-stearate on Pt/Al2O3. Catal Lett 130:9–18
Dubé M, Tremblay A, Liu J (2007) Biodiesel production using a membrane reactor. Biores
Technol 98:639–647
Dugan J (2007) A dry wash approach to biodiesel purification. http://www.biodieselmagazine.
com/article.jsp?article_id=1918. Assessed 15 Mar 2017
Dumay J, Radier S, Barnathan G, Berge J-P, Jaouen P (2008) Recovery of valuable soluble
compounds from washing waters generated during small fatty pelagic surimi processing by
membrane processes. Environ Technol 29:451–461
Dupain X, Costa DJ, Schaverien CJ, Makkee M, Moulijn JA (2007) Cracking of a rapeseed
vegetable oil under realistic FCC conditions. Appl Catal B 72:44–61
Erich K (1982) Separating processes. In: Erich K (ed) Handbook of laboratory distillation with an
introduction to pilot plant distillation. Elsevier, New York, United States
4 Biodiesel Purification and Upgrading Technologies 93

Escobar JC, Lora ES, Venturini OJ, Yáñez EE, Castillo EF, Almazan O (2009) Biofuels:
environment, technology and food security. Renew Sustain Energy Rev 13:1275–1287
Faccini CS, Cunha MED, Moraes MSA, Krause LC, Manique MC, Rodrigues MRA,
Benvenutti EV, Caramão EB (2011) Dry washing in biodiesel purification: a comparative
study of adsorbents. J Braz Chem Soc 22:558–563
Fadhil A, Dheyab M (2015) Purification of biodiesel fuels produced from spent frying oils over
activated carbons. Energy Sources Part A: Recovery Util Environ Eff 37:149–155
Falahati H, Tremblay A (2012) The effect of flux and residence time in the production of biodiesel
from various feedstocks using a membrane reactor. Fuel 91:126–133
Ferrero G, Almeida M, Alvim-Ferraz M, Dias J (2014) Water-free process for eco-friendly
purification of biodiesel obtained using a heterogeneous Ca-based catalyst. Fuel Process
Technol 121:114–118
Foraita S, Liu Y, Haller GL, Baráth E, Zhao C, Lercher JA (2017) Controlling hydrodeoxygena-
tion of stearic acid to n-heptadecane and n-octadecane by adjusting the chemical properties of
Ni/SiO2–ZrO2 catalyst. ChemCatChem 9:195–203
Freeman BD, Paul DR, Czenkusch K, Ribeiro CP, Ba C (2012) Thermally Rearranged
(TR) polymers as membranes for ethanol dehydration
Furimsky E (2000) Catalytic hydrodeoxygenation. Appl Catal A 199:147–190
Furimsky E, Massoth FE (1999) Deactivation of hydroprocessing catalysts. Catal Today 52:381–
495
Giorno F, Mazzei R, Giorno L (2013) Purification of triacylglycerols for biodiesel production from
Nannochloropsis microalgae by membrane technology. Biores Technol 140:172–178
Gomes MCS, Arroyo PA, Pereira NC (2011) Biodiesel production from degummed soybean oil
and glycerol removal using ceramic membrane. J Membr Sci 378:453–461
Gomes MCS, Arroyo PA, Pereira NC (2013) Influence of acidified water addition on the biodiesel
and glycerol separation through membrane technology. J Membr Sci 431:28–36
Gomes MCS, Pereira NC, de Barros STD (2010) Separation of biodiesel and glycerol using
ceramic membranes. J Membr Sci 352:271–276
Gomes MG, Santos DQ, de Morais LC, Pasquini D (2015) Purification of biodiesel by dry
washing, employing starch and cellulose as natural adsorbents. Fuel 155:1–6
Gomez-Castro FI, Rico-Ramirez V, Segovia-Hernandez JG, Hernandez S (2010) Feasibility study
of a thermally coupled reactive distillation process for biodiesel production. Chem Eng Process
49:262–269
Gong S, Shinozaki A, Shi M, Qian EW (2012) Hydrotreating of jatropha oil over alumina based
catalysts. Energy Fuels 26:2394–2399
Gosselink RW, Hollak SA, Chang SW, van Haveren J, de Jong KP, Bitter JH, van Es DS (2013a)
Reaction pathways for the deoxygenation of vegetable oils and related model compounds.
Chemsuschem 6:1576–1594
Gosselink RW, Stellwagen DR, Bitter JH (2013b) Tungsten-based catalysts for selective
deoxygenation. Angew Chem 125:5193–5196
Grandison AS (1996) Ion-exchange and electrodialysis. In: Grandison AS, Lewis MJ
(eds) Separation processes in the food and biotechnology industries. Woodhead Publishing,
Cambridge, United Kingdom
Greenwell H, Laurens L, Shields R, Lovitt R, Flynn K (2010) Placing microalgae on the biofuels
priority list: a review of the technological challenges. J R Soc Interface 7:703–726
Groeneweg F, Agterof W, Jaeger P, Janssen J, Wieringa J, Klahn J (1998) On the mechanism of
the inversion of emulsions. Chem Eng Res Des 76:55–63
Han D, Row KH (2010) Recent applications of ionic liquids in separation technology. Molecules
15:2405–2426
Han J, Duan J, Chen P, Lou H, Zheng X (2011a) Molybdenum carbide-catalyzed conversion of
renewable oils into diesel-like hydrocarbons. Adv Synth Catal 353:2577–2583
Han J, Duan J, Chen P, Lou H, Zheng X, Hong H (2011b) Nanostructured molybdenum carbides
supported on carbon nanotubes as efficient catalysts for one-step hydrodeoxygenation and
isomerization of vegetable oils. Green Chem 13:2561–2568
94 H. Bateni et al.

Han J, Duan J, Chen P, Lou H, Zheng X, Hong H (2012) Carbon-supported molybdenum carbide
catalysts for the conversion of vegetable oils. Chemsuschem 5:727–733
Hanafi SA, Mohamed MS (2011) Recent trends in the cleaning of diesel fuels via desulfurization
processes. Energy Sources Part A: Recovery Util Environ Eff 33:495–511
Hanson C (1971) Solvent extraction: the current position. In: Hanson C (ed) Recent advances in
liquid–liquid extraction. Pergamon Press, Oxford, United Kingdom
Hao JH, Chen C, Li L, Yu L, Jiang W (2000) Preparation of solvent-resistant anion-exchange
membranes. Desalination 129:15–22
Hayyan M, Mjalli FS, Hashim MA, Alnashef IM (2010) A novel technique for separating
glycerine from palm oil-based biodiesel using ionic liquids. Fuel Process Technol 91:116–120
He H, Guo X, Zhu S (2006) Comparison of membrane extraction with traditional extraction
methods for biodiesel production. J Am Oil Chem Soc 83:457–460
Ho RM, Wu CH, Su AC (1990) Morphology of plastic/rubber blends. Polym Eng Sci 30:511–518
Hollak SA, Gosselink RW, van Es DS, Bitter JH (2013) Comparison of tungsten and molybdenum
carbide catalysts for the hydrodeoxygenation of oleic acid. ACS Catal 3:2837–2844
Hua F, Tsang Y, Wang Y, Chan S, Chua H, Sin S (2007) Performance study of ceramic
microfiltration membrane for oily wastewater treatment. Chem Eng J 128:169–175
Huang H-J, Ramaswamy S (2013) Overview of biomass conversion processes and separation and
purification technologies in biorefineries. In: Ramaswamy S, Huang H-J, Ramarao BV
(eds) Separation and purification technologies in biorefineries. Wiley, Chichester, United
Kingdom
Huerga IR, Zanuttini MS, Gross MS, Querini CA (2014) Biodiesel production from Jatropha
curcas: Integrated process optimization. Energy Convers Manag 80:1–9
Immer JG, Kelly MJ, Lamb HH (2010) Catalytic reaction pathways in liquid-phase deoxygenation
of C18 free fatty acids. Appl Catal A 375:134–139
Itthibenchapong V, Srifa A, Kaewmeesri R, Kidkhunthod P, Faungnawakij K (2017)
Deoxygenation of palm kernel oil to jet fuel-like hydrocarbons using Ni-MoS2/c-Al2O3
catalysts. Energy Convers Manag 134:188–196
Jaber R, Shirazi M, Toufaily J, Hamieh A, Noureddin A, Ghanavati H, Ghaffari A, Zenouzi A,
Karout A, Ismail A (2015) Biodiesel wash-water reuse using microfiltration: toward
zero-discharge strategy for cleaner and economized biodiesel production. Biofuel Res J
2:148–151
Jiang LY, Wang Y, Chung T-S, Qiao XY, Lai J-Y (2009) Polyimides membranes for
pervaporation and biofuels separation. Prog Polym Sci 34:1135–1160
Jönsson A-S (2013) Microfiltration, ultrafiltration and diafiltration. In: Ramaswamy S, Huang H-J,
Ramarao BV (eds) Separation and purification technologies in biorefineries. Wiley, Chichester,
United Kingdom
Kandel K, Anderegg JW, Nelson NC, Chaudhary U, Slowing II (2014) Supported iron
nanoparticles for the hydrodeoxygenation of microalgal oil to green diesel. J Catal 314:142–
148
Karaosmanoǧlu F, Cıǧızoǧlu KB, Tüter M, Ertekin S (1996) Investigation of the refining step of
biodiesel production. Energy Fuels 10:890–895
Kertes AS (1971) The chemistry of solvent extraction. In: Hanson C (ed) Recent advances in
liquid–liquid extraction. Pergamon Press, Oxford, United Kingdom
Kiatkittipong W, Phimsen S, Kiatkittipong K, Wongsakulphasatch S, Laosiripojana N,
Assabumrungrat S (2013) Diesel-like hydrocarbon production from hydroprocessing of
relevant refining palm oil. Fuel Process Technol 116:16–26
Kim H-J, Kang B-S, Kim M-J, Park YM, Kim D-K, Lee J-S, Lee K-Y (2004) Transesterification
of vegetable oil to biodiesel using heterogeneous base catalyst. Catal Today 93:315–320
Kiss AA (2010) Separative reactors for integrated production of bioethanol and biodiesel. Comput
Chem Eng 34:812–820
Kiss AA (2013) Advanced distillation technologies: design, control and applications. Wiley,
Chichester, United Kingdom
4 Biodiesel Purification and Upgrading Technologies 95

Kiss AA, Dimian AC, Rothenberg G (2006a) Solid acid catalysts for biodiesel production—
towards sustainable energy. Adv Synth Catal 348:75–81
Kiss AA, Omota F, Dimian AC, Rothenberg G (2006b) The heterogeneous advantage: biodiesel
by catalytic reactive distillation. Top Catal 40:141–150
Kiss AA, Dimian AC, Rothenberg G (2007) Biodiesel by catalytic reactive distillation powered by
metal oxides. Energy Fuels 22:598–604
Knothe G (2010) Biodiesel and renewable diesel: a comparison. Prog Energy Combust Sci
36:364–373
Kolah AK, Lira CT, Miller DJ (2013) Reactive distillation for the biorefinery. In: Ramaswamy S,
Huang H-J, Ramarao BV (eds) Separation and purification technologies in biorefineries. Wiley,
Chichester, United Kingdom
Kubička D, Horáček J (2011) Deactivation of HDS catalysts in deoxygenation of vegetable oils.
Appl Catal A 394:9–17
Kubička D, Kaluža L (2010) Deoxygenation of vegetable oils over sulfided Ni, Mo and NiMo
catalysts. Appl Catal A 372:199–208
Kubičková I, Snåre M, Eränen K, Mäki-Arvela P, Murzin DY (2005) Hydrocarbons for diesel fuel
via decarboxylation of vegetable oils. Catal Today 106:197–200
Lam E, Luong JH (2014) Carbon materials as catalyst supports and catalysts in the transformation
of biomass to fuels and chemicals. ACS Catal 4:3393–3410
Laurent E, Delmon B (1994) Influence of water in the deactivation of a sulfided NiMo c-Al2O3
catalyst during hydrodeoxygenation. J Catal 146:281288–285291
Leeruang U, Pengprecha S (2012) Purification of biodiesel by adsorption with activated low silica
bentonite. In: International conference on chemical processes and environmental issues,
Singapore
Lei Z, Chen B, Ding Z (2005) Special distillation processes. Elsevier, Amsterdam, Netherlands
Lestari S, Maki-Arvela P, Bernas H, Simakova O, Sjöholm R, Beltramini J, Lu GM, Myllyoja J,
Simakova I, Murzin DY (2009a) Catalytic deoxygenation of stearic acid in a continuous
reactor over a mesoporous carbon-supported Pd catalyst. Energy Fuels 23:3842–3845
Lestari S, Mäki-Arvela P, Simakova I, Beltramini J, Lu GM, Murzin DY (2009b) Catalytic
deoxygenation of stearic acid and palmitic acid in semibatch mode. Catal Lett 130:48–51
Lestari S, Mäki-Arvela P, Eränen K, Beltramini J, Lu GM, Murzin DY (2010) Diesel-like
hydrocarbons from catalytic deoxygenation of stearic acid over supported Pd nanoparticles on
SBA-15 catalysts. Catal Lett 134:250–257
Lestari S, Mäki-Arvela P, Beltramini J, Lu G, Murzin DY (2009c) Transforming triglycerides and
fatty acids into biofuels. Chemsuschem 2:1109–1119
Lestari S, Simakova I, Tokarev A, Mäki-Arvela P, Eränen K, Murzin DY (2008) Synthesis of
biodiesel via deoxygenation of stearic acid over supported Pd/C catalyst. Catal Lett 122:247–
251
Levan MD, Carta G (2008) Section 16. Adsorption and ion exchange. In: Green DW, Perry RH
(eds) Perry’s chemical engineers’ handbook, 8th edn. McGraw-Hill, New York, United States
Lewis MJ (1996) Solids separation processes. In: Separation processes in the food and
biotechnology industries. Woodhead Publishing, Cambridge, United Kingdom
Li L, Wang Y (2005) Quaternized polyethersulfone Cardo anion exchange membranes for direct
methanol alkaline fuel cells. J Membr Sci 262:1–4
Li M, Zhang H, Shao Z-G (2006) Quaternized poly (phthalazinone ether sulfone ketone)
membrane doped with H3PO4 for high-temperature PEMFC operation. Electrochem
Solid-State Lett 9:A60–A63
Liu Y, Sotelo-Boyás R, Murata K, Minowa T, Sakanishi K (2011) Hydrotreatment of vegetable
oils to produce bio-hydrogenated diesel and liquefied petroleum gas fuel over catalysts
containing sulfided Ni–Mo and solid acids. Energy Fuels 25:4675–4685
Ma F, Clements LD, Hanna MA (1998) Biodiesel fuel from animal fat. Ancillary studies on
transesterification of beef tallow. Ind Eng Chem Res 37:3768–3771
Ma F, Hanna MA (1999) Biodiesel production: a review. Bioresource Technol 70:1–15
96 H. Bateni et al.

Machado GD, Aranda DA, Castier M, Cabral VF, Cardozo-Filho LC (2011) Computer simulation
of fatty acid esterification in reactive distillation columns. Ind Eng Chem Res 50:10176–10184
Madsen AT, Christensen CH, Fehrmann R, Riisager A (2011) Hydrodeoxygenation of waste fat
for diesel production: Study on model feed with Pt/alumina catalyst. Fuel 90:3433–3438
Mah S-K, Leo C, Wu TY, Chai S-P (2012) A feasibility investigation on ultrafiltration of palm oil
and oleic acid removal from glycerin solutions: flux decline, fouling pattern, rejection and
membrane characterizations. J Membr Sci 389:245–256
Mäki-Arvela P, Kubickova I, Snåre M, Eränen K, Murzin DY (2007) Catalytic deoxygenation of
fatty acids and their derivatives. Energy Fuels 21:30–41
Mäki-Arvela P, Snåre M, Eränen K, Myllyoja J, Murzin DY (2008) Continuous decarboxylation
of lauric acid over Pd/C catalyst. Fuel 87:3543–3549
Manesiotis P, Theodoridis G (2016) Affinity-based separations in bioanalysis. J Chromatogr B
1021:1–2
Manique MC, Faccini CS, Onorevoli B, Benvenutti EV, Caramão EB (2012) Rice husk ash as an
adsorbent for purifying biodiesel from waste frying oil. Fuel 92:56–61
Mänttäri M, van der Bruggen B, Nyström M (2013) Nanofiltration. In: Ramaswamy S, Huang H-J,
Ramarao BV (eds) Separation and purification technologies in biorefineries. Wiley, Chichester,
United Kingdom
Manuale D, Mazzieri V, Torres G, Vera C, Yori J (2011) Non-catalytic biodiesel process with
adsorption-based refining. Fuel 90:1188–1196
Mata TM, Cardoso N, Ornelas M, Neves S, Caetano NS (2011) Evaluation of two purification
methods of biodiesel from beef tallow, pork lard, and chicken fat. Energy Fuels 25:4756–4762
Mazzieri V, Vera C, Yori J (2008) Adsorptive properties of silica gel for biodiesel refining. Energy
Fuels 22:4281–4284
Miga K, Stanczyk K, Sayag C, Brodzki D, Djéga-Mariadassou G (1999) Bifunctional behavior of
bulk MoOxNy and nitrided supported NiMo catalyst in hydrodenitrogenation of indole. J Catal
183:63–68
Monnier J, Sulimma H, Dalai A, Caravaggio G (2010) Hydrodeoxygenation of oleic acid and
canola oil over alumina-supported metal nitrides. Appl Catal A: General 382:176–180
Morgan T, Santillan-Jimenez E, Harman-Ware AE, Ji Y, Grubb D, Crocker M (2012) Catalytic
deoxygenation of triglycerides to hydrocarbons over supported nickel catalysts. Chem Eng J
189:346–355
Mortensen PM, Grunwaldt J-D, Jensen PA, Knudsen K, Jensen AD (2011) A review of catalytic
upgrading of bio-oil to engine fuels. Appl Catal A: General 407:1–19
Moser BR (2012) Preparation of fatty acid methyl esters from hazelnut, high-oleic peanut and
walnut oils and evaluation as biodiesel. Fuel 92:231–238
Mueanmas C, Prasertsit K, Tongurai C (2010) Transesterification of triolein with methanol in
reactive distillation column: simulation studies. Int J Chem React Eng 8
Muniyappa PR, Brammer SC, Noureddini H (1996) Improved conversion of plant oils and animal
fats into biodiesel and co-product. Bioresource Technol 56:19–24
Nagy G, Pölczmann G, Kalló D, Hancsók J (2009) Investigation of hydrodearomatization of gas
oils on noble metal/support catalysts. Chem Eng J 154:307–314
Nava R, Pawelec B, Castaño P, Álvarez-Galván M, Loricera C, Fierro J (2009) Upgrading of
bio-liquids on different mesoporous silica-supported CoMo catalysts. Appl Catal B: Environ
92:154–167
Neylon M, Choi S, Kwon H, Curry K, Thompson L (1999) Catalytic properties of early transition
metal nitrides and carbides: n-butane hydrogenolysis, dehydrogenation and isomerization.
Appl Catal A: General 183:253–263
Noori MS, Karimi K (2016a) Chemical and structural analysis of alkali pretreated pinewood for
efficient ethanol production. RSC Adv 6:65683–65690
Noori MS, Karimi K (2016b) Detailed study of efficient ethanol production from elmwood by
alkali pretreatment. Biochem Eng J 105:197–204
Parkash S (2003) Refining processes handbook. Gulf Professional Publishing, Burlington, United
States
4 Biodiesel Purification and Upgrading Technologies 97

Parvizsedghy R, Sadrameli SM, Towfighi Darian J (2015) Upgraded biofuel diesel production by
thermal cracking of castor biodiesel. Energy Fuels 30:326–333
Pérez-Cisneros ES, Mena-Espino X, Rodríguez-López V, Sales-Cruz M, Viveros-García T,
Lobo-Oehmichen R (2016) An integrated reactive distillation process for biodiesel production.
Comput Chem Eng 91:233–246
Peroni M, Mancino G, Baráth E, Gutiérrez OY, Lercher JA (2016) Bulk and c-Al2O3-supported
Ni2P and MoP for hydrodeoxygenation of palmitic acid. Appl Catal B: Environ 180:301–311
Phimsen S, Kiatkittipong W, Yamada H, Tagawa T, Kiatkittipong K, Laosiripojana N,
Assabumrungrat S (2016) Oil extracted from spent coffee grounds for bio-hydrotreated diesel
production. Energy Convers Manag 126:1028–1036
Pittia P, Mastrocola D, Nicoli M (2005) Effect of colloidal properties of oil-in-water emulsions on
ethanol liquid–vapour partition. Food Res Int 38:585–595
Poddar T, Jagannath A, Almansoori A (2015) Biodiesel production using reactive distillation: a
comparative simulation study. Energy Procedia 75:17–22
Predojević ZJ (2008) The production of biodiesel from waste frying oils: a comparison of different
purification steps. Fuel 87:3522–3528
Qin Y, Chen P, Duan J, Han J, Lou H, Zheng X, Hong H (2013) Carbon nanofibers supported
molybdenum carbide catalysts for hydrodeoxygenation of vegetable oils. RSC Adv 3:17485–
17491
Qiu Z, Zhao L, Weatherley L (2010) Process intensification technologies in continuous biodiesel
production. Chem Eng Process: Process Intensif 49:323–330
Rahayu SS, Mindaryani A (2007) Optimization of biodiesel washing by water extraction. In:
Proceedings of the world congress on engineering and computer science, San Francisco, United
States
Rantanen L, Linnaila R, Aakko P, Harju T (2005) NExBTL-biodiesel fuel of the second
generation. SAE technical paper
Rodríguez-Reinoso F (1998) The role of carbon materials in heterogeneous catalysis. Carbon
36:159–175
Rogers KA, Zheng Y (2016) Selective deoxygenation of biomass-derived bio-oils within
hydrogen-modest environments: a review and new insights. Chemsuschem 9:1750–1772
Romero Y, Richard F, Brunet S (2010) Hydrodeoxygenation of 2-ethylphenol as a model
compound of bio-crude over sulfided Mo-based catalysts: promoting effect and reaction
mechanism. Appl Catal B: Environ 98:213–223
Rossi N, Derouiniot-Chaplain M, Jaouen P, Legentilhomme P, Petit I (2008) Arthrospira platensis
harvesting with membranes: fouling phenomenon with limiting and critical flux. Bioresource
Technol 99:6162–6167
Rossi N, Jaouen P, Legentilhomme P, Petit I (2004) Harvesting of cyanobacterium Arthrospira
platensis using organic filtration membranes. Food Bioprod Process 82:244–250
Rostrup-Nielsen JR (2008) 13.11 steam reforming. In: Ertl G, Knozinger H, Schuth F, Weitkamp J
(eds) Handbook of heterogeneous catalysis. Wiley-VCH Verlag GmbH & Co. KGaA,
Weinheim, Germany
Rozmysłowicz B, Maki-Arvela P, Tokarev A, Leino A-R, Eränen K, Murzin DY (2012) Influence
of hydrogen in catalytic deoxygenation of fatty acids and their derivatives over Pd/C. Ind Eng
Chem Res 51:8922–8927
Ruddy DA, Schaidle JA, Ferrell III, JR, Wang J, Moens L, Hensley JE (2014) Recent advances in
heterogeneous catalysts for bio-oil upgrading via “ex situ catalytic fast pyrolysis”: catalyst
development through the study of model compounds. Green Chem 16:454–490
Russbueldt BM, Hoelderich WF (2009) New sulfonic acid ion-exchange resins for the
preesterification of different oils and fats with high content of free fatty acids. Appl Catal A:
General 362:47–57
Ruthven DM (1984) Principles of adsorption and adsorption processes. Wiley, New York, United
States
98 H. Bateni et al.

Saifuddin N, Chua K (2004) Production of ethyl ester (biodiesel) from used frying oil:
optimization of transesterification process using microwave irradiation. Malays J Chem
6:77–82
Salahi A, Gheshlaghi A, Mohammadi T, Madaeni SS (2010) Experimental performance evaluation
of polymeric membranes for treatment of an industrial oily wastewater. Desalination 262:235–
242
Saleh J, Dubé MA, Tremblay AY (2010a) Effect of soap, methanol, and water on glycerol particle
size in biodiesel purification. Energy Fuels 24:6179–6186
Saleh J, Dubé MA, Tremblay AY (2011) Separation of glycerol from FAME using ceramic
membranes. Fuel Process Technol 92:1305–1310
Saleh J, Tremblay AY, Dubé MA (2010b) Glycerol removal from biodiesel using membrane
separation technology. Fuel 89:2260–2266
Satyarthi J, Chiranjeevi T, Gokak D, Viswanathan P (2013) An overview of catalytic conversion
of vegetable oils/fats into middle distillates. Catal Sci Technol 3:70–80
Schlatter JC, Oyama ST, Metcalfe III, JE, Lambert Jr, JM (1988) Catalytic behavior of selected
transition metal carbides, nitrides, and borides in the hydrodenitrogenation of quinoline. Ind
Eng Chem Res 27:1648–1653
Schuchardt U, Sercheli R, Vargas RM (1998) Transesterification of vegetable oils: a review. J Braz
Chem Soc 9:199–210
Şenol O, Viljava T-R, Krause A (2005a) Hydrodeoxygenation of aliphatic esters on sulphided
NiMo/c-Al2O3 and CoMo/c-Al2O3 catalyst: the effect of water. Catal Today 106:186–189
Şenol O, Viljava T-R, Krause A (2005b) Hydrodeoxygenation of methyl esters on sulphided
NiMo/c-Al2O3 and CoMo/c-Al2O3 catalysts. Catal Today 100:331–335
Serrano M, Bouaid A, Martínez M, Aracil J (2013) Oxidation stability of biodiesel from different
feedstocks: influence of commercial additives and purification step. Fuel 113:50–58
Shahbaz K, Mjalli F, Hashim M, Alnashef I (2011) Eutectic solvents for the removal of residual
palm oil-based biodiesel catalyst. Sep Purif Technol 81:216–222
Shahbaz K, Mjalli FS, Hashim M, Al-Nashef IM (2010) Using deep eutectic solvents for the
removal of glycerol from palm oil-based biodiesel. J Appl Sci 10:3349–3354
Shi W, Li H, Su Y, Liu J (2016a) Biodiesel production by quaternized polysulfone membrane:
experimental and kinetics model. Energy Procedia 104:402–406
Shi W, Li H, Zhou R, Zhang H, Du Q (2016b) Biodiesel production from soybean oil by
quaternized polysulfone alkali-catalyzed membrane. Bioresource Technol 210:43–48
Simakova I, Rozmysłowicz B, Simakova O, Mäki-Arvela P, Simakov A, Murzin DY (2011)
Catalytic deoxygenation of C18 fatty acids over mesoporous Pd/C catalyst for synthesis of
biofuels. Top Catal 54:460–466
Simakova I, Simakova O, Mäki-Arvela P, Simakov A, Estrada M, Murzin DY (2009)
Deoxygenation of palmitic and stearic acid over supported Pd catalysts: effect of metal
dispersion. Appl Catal A: General 355:100–108
Simasatitkul L, Siricharnsakunchai P, Patcharavorachot Y, Assabumrungrat S, Arpornwichanop A
(2011) Reactive distillation for biodiesel production from soybean oil. Korean J Chem Eng
28:649–655
Slade RC, Kizewski JP, Poynton SD, Zeng R, Varcoe JR (2012) Alkaline membrane fuel cells. In:
Meyers RA (ed) Encyclopedia of sustainability science and technology. Springer, New York,
United States
Snåre M, Kubičková I, Mäki-Arvela P, Chichova D, Eränen K, Murzin DY (2008) Catalytic
deoxygenation of unsaturated renewable feedstocks for production of diesel fuel hydrocarbons.
Fuel 87:933–945
Snåre M, Kubickova I, Mäki-Arvela P, Eränen K, Murzin DY (2006) Heterogeneous catalytic
deoxygenation of stearic acid for production of biodiesel. Ind Eng Chem Res 45:5708–5715
Snåre M, Kubičková I, Mäki-Arvela P, Eränen K, Wärnå J, Murzin DY (2007) Production of
diesel fuel from renewable feeds: kinetics of ethyl stearate decarboxylation. Chem Eng J
134:29–34
4 Biodiesel Purification and Upgrading Technologies 99

Snåre M, Mäki-Arvela P, Simakova I, Myllyoja J, Murzin DY (2009) Overview of catalytic


methods for production of next generation biodiesel from natural oils and fats. Russian J Phys
Chem B 3:1035–1043
Sotelo-Boyas R, Liu Y, Minowa T (2010) Renewable diesel production from the hydrotreating of
rapeseed oil with Pt/Zeolite and NiMo/Al2O3 catalysts. Ind Eng Chem Res 50:2791–2799
Sousa L, Zotin J, da Silva VT (2012) Hydrotreatment of sunflower oil using supported
molybdenum carbide. Appl Catal A: General 449:105–111
Srifa A, Faungnawakij K, Itthibenchapong V, Assabumrungrat S (2015) Roles of monometallic
catalysts in hydrodeoxygenation of palm oil to green diesel. Chem Eng J 278:249–258
Srinivasan S (2009) The food v. fuel debate: A nuanced view of incentive structures. Renew
Energy 34:950–954
Stanislaus A, Marafi A, Rana MS (2010) Recent advances in the science and technology of ultra
low sulfur diesel (ULSD) production. Catal Today 153:1–68
Stellwagen DR, Bitter JH (2015) Structure–performance relations of molybdenum-and tungsten
carbide catalysts for deoxygenation. Green Chem 17:582–593
Stöcker M (2008) Biofuels and biomass-to-liquid fuels in the biorefinery: catalytic conversion of
lignocellulosic biomass using porous materials. Angew Chem Int Ed 47:9200–9211
Stojković IJ, Stamenković OS, Povrenović DS, Veljković VB (2014) Purification technologies for
crude biodiesel obtained by alkali-catalyzed transesterification. Renew Sustain Energy Rev
32:1–15
Tremblay A, Cao P, Dubé MA (2008) Biodiesel production using ultralow catalyst concentrations.
Energy Fuels 22:2748–2755
Treybal RE (1980) Mass transfer operations, 3rd edn. McGraw-Hill, New York, United States
Twaiq FA, Zabidi NA, Bhatia S (1999) Catalytic conversion of palm oil to hydrocarbons:
performance of various zeolite catalysts. Ind Eng Chem Res 38:3230–3237
UOP H G D (2017) UOP, Honeywell Green Diesel. https://www.uop.com/processing-solutions/
renewables/green-diesel/#biodiesel. Accessed 02 June 2017
van der Graaf S, Schroën C, Boom R (2005) Preparation of double emulsions by membrane
emulsification—a review. J Membr Sci 251:7–15
Van Gerpen J, Shanks B, Pruszko R, Clements D, Knothe G (2004) Biodiesel production
technology: August 2002–January 2004, NREL/SR-510-36244. National Renewable Energy
Laboratory
Veljković VB, Banković-Ilić IB, Stamenković OS (2015) Purification of crude biodiesel obtained
by heterogeneously-catalyzed transesterification. Renew Sustain Energy Rev 49:500–516
Venkatesan S (2013) Adsorption. In: Ramaswamy S, Huang H-J, Ramarao BV (eds) Separation
and purification technologies in biorefineries. Wiley, Chichester, United Kingdom
Vinh TQ, Loan NTT, Yang X-Y, Su B-L (2011) Preparation of bio-fuels by catalytic cracking
reaction of vegetable oil sludge. Fuel 90:1069–1075
Wall J, van Gerpen J, Thompson J (2011) Soap and glycerin removal from biodiesel using
waterless processes. Trans ASABE 54:535–541
Wang H, Yan S, Salley SO, Ng KS (2012) Hydrocarbon fuels production from hydrocracking of
soybean oil using transition metal carbides and nitrides supported on ZSM-5. Ind Eng Chem
Res 51:10066–10073
Wang H, Yan S, Salley SO, Ng KS (2013) Support effects on hydrotreating of soybean oil over
NiMo carbide catalyst. Fuel 111:81–87
Wang J, Ge X, Wang Z, Jin Y (2001) Experimental studies on the catalytic distillation for
hydrolysis of methyl acetate. Chem Eng Technol 24:155–159
Wang Y, Nie J, Zhao M, Ma S, Kuang L, Han X, Tang S (2010) Production of biodiesel from
waste cooking oil via a two-step catalyzed process and molecular distillation. Energy Fuels
24:2104–2108
Wang Y, Wang X, Liu Y, Ou S, Tan Y, Tang S (2009) Refining of biodiesel by ceramic membrane
separation. Fuel Process Technol 90:422–427
Wei C-Y, Huang T-C, Yu Z-R, Wang B-J, Chen H-H (2014) Fractionation for biodiesel
purification using supercritical carbon dioxide. Energies 7:824–833
100 H. Bateni et al.

Wijffels RH, Barbosa MJ, Eppink MH (2010) Microalgae for the production of bulk chemicals and
biofuels. Biofuels, Bioprod Biorefin 4:287–295
Yang C, Li R, Cui C, Liu S, Qiu Q, Ding Y, Wu Y, Zhang B (2016) Catalytic hydroprocessing of
microalgae-derived biofuels: a review. Green Chem 18:3684–3699
Yang RT (2003) Adsorbents: fundamentals and applications. Wiley, New Jersey, United States
Yang Y, Chen J, Shi H (2013a) Deoxygenation of methyl laurate as a model compound to
hydrocarbons on Ni2P/SiO2, Ni2P/MCM-41, and Ni2P/SBA-15 catalysts with different
dispersions. Energy Fuels 27:3400–3409
Yang Y, Ochoa-Hernández C, de la Peña O’shea VCA, Coronado JM, Serrano DP (2012) Ni2P/
SBA-15 as a hydrodeoxygenation catalyst with enhanced selectivity for the conversion of
methyl oleate into n-octadecane. ACS Catal 2:592–598
Yang Y, Ochoa-Hernández C, Pizarro P, Víctor A, Coronado JM, Serrano DP (2015) Influence of
the Ni/P ratio and metal loading on the performance of NixPy/SBA-15 catalysts for the
hydrodeoxygenation of methyl oleate. Fuel 144:60–70
Yang Y, Wang Q, Zhang X, Wang L, Li G (2013b) Hydrotreating of C18 fatty acids to
hydrocarbons on sulphided NiW/SiO2–Al2O3. Fuel Process Technol 116:165–174
Yilmaz G, Jongboom R, Van Soest J, Feil H (1999) Effect of glycerol on the morphology of
starch–sunflower oil composites. Carbohydr Polym 38:33–39
Yori J, D’Ippolito S, Pieck C, Vera C (2007) Deglycerolization of biodiesel streams by adsorption
over silica beds. Energy Fuels 21:347–353
Zarchin R, Rabaev M, Vidruk-Nehemya R, Landau MV, Herskowitz M (2015) Hydroprocessing
of soybean oil on nickel-phosphide supported catalysts. Fuel 139:684–691
Zhao C, Brück T, Lercher JA (2013) Catalytic deoxygenation of microalgae oil to green
hydrocarbons. Green Chem 15:1720–1739
Zhao H, Baker GA (2013) Ionic liquids and deep eutectic solvents for biodiesel synthesis: a
review. J Chem Technol Biotechnol 88:3–12
Zhu X, Lobban LL, Mallinson RG, Resasco DE (2011) Bifunctional transalkylation and
hydrodeoxygenation of anisole over a Pt/HBeta catalyst. J Catal 281:21–29
Chapter 5
Applications of Biodiesel By-products

Hajar Rastegari, Hossein Jazini, Hassan S. Ghaziaskar


and Mohammad Yalpani

Abstract This chapter reviews the applications of biodiesel by-products. The


biodiesel production process is predominantly carried out through transesterifica-
tion of triglycerides with methanol. Besides the desirable methyl esters, this process
provides other products including oil cake/meal, crude glycerol, methanol, and
biodiesel wash-water. Oil cake/meal are solid residues obtained after oil extraction
from the seeds. The by-product of the transesterification step is crude glycerol. To
remove the impurities, crude methyl esters will be washed out which result in the
biodiesel wash-water that is another potential by-product. Many applications are
known for the aforementioned by-products and also the unreacted methanol. This
chapter starts with a brief introduction on biodiesel process. It will continue with
reviewing applications of biodiesel process’s by-products and unreacted methanol.

5.1 Introduction

Biodiesel is identified as one of the alternative resources of energy which is


composed of methyl/ethyl esters of long chain fatty acids such as lauric, palmitic,
stearic, oleic, etc. The main advantages of biodiesel in either pure or blended forms,
are as follows: being renewable and biodegradable, producing less smoke and
particulates, higher cetane number, as well as lower carbon monoxide and hydro-
carbon emissions (Ma and Hanna 1999; Encinar et al. 2007; Antolin et al. 2002).
Moreover, biodiesel is known for having no sulfur; therefore, reducing sulfur
dioxide emissions. However the commercialization of this fuel demands for further
reductions in the production costs (Vicente et al. 2004).

H. Rastegari  H. Jazini  H. S. Ghaziaskar (&)


Department of Chemistry, Isfahan University of Technology, Isfahan 84156-83111, Iran
e-mail: ghazi@cc.iut.ac.ir
M. Yalpani
R&D Department, Farzin Chemicals Sepahan Co., Montazerie Industrial Complex,
Villashahr, Isfahan 85131-14461, Iran

© Springer Nature Switzerland AG 2019 101


M. Tabatabaei and M. Aghbashlo (eds.), Biodiesel,
Biofuel and Biorefinery Technologies 8,
https://doi.org/10.1007/978-3-030-00985-4_5
102 H. Rastegari et al.

Transesterification of renewable feedstocks, such as vegetable oils, animal fats,


waste cooking oil, and algal oil, with methanol/ethanol, produce crude biodiesel
and crude glycerol (Ma and Hanna 1999; Di Serio et al. 2008; Huber et al. 2006).
The conventional biodiesel production process is shown in Fig. 5.1. Besides desired
alkyl esters, by-products are also generated throughout the production cycle.
Pressing and extraction of oil from oil seed plants lead to the generation of oil cake/
meal as one of these by-products. The other by-product named crude glycerol, is
produced during the transesterification reaction (Kolesarova et al. 2011). Moreover,
to remove the impurities such as soaps, short chain fatty acids, excess methanol,
catalyst residues, etc., crude biodiesel is generally washed generating biodiesel
wash-water (wastewater) as another potential by-product (Kolesarova et al. 2011).
Therefore, the most important by-products from biodiesel production cycle include
pressed oil cake/meal, crude glycerol, and biodiesel wash-water.
As an alternative to diesel fuel, biodiesel must be technically feasible, eco-
nomically competitive, environmentally acceptable, and readily available (El
Diwani et al. 2009). However, the process cost in many parts of the world is still
higher than that of the petroleum-derived diesel and it is necessary to explore ways
to reduce the costs. One convenient way is to convert the by-products to higher
added value products (Kolesarova et al. 2011). This chapter reviews some options
in order to lower the costs of the biodiesel process through the aforementioned
approach.

Fig. 5.1 Process flow diagram of conventional biodiesel production process


5 Applications of Biodiesel By-products 103

5.2 Oil Cake/Meal Applications

The solids remaining after oil extraction from oil seeds are mainly composed of
proteins, fibers, and minerals. Depending on the method of oil extraction, two basic
types of solid by-products are generated. The solid residue left after simple
mechanical press, is called oil cake and if this cake is further extracted by a solvent,
another type, known as oil meal is produced (Kolesarova et al. 2011). In fact, the
main difference between oil cake and oil meal is their oil contents (Kolesarova et al.
2011). More specifically, the oil content is reduced from 12–20 wt% to about 8 wt%
in oil cake through pressing. While by using hexane-based extraction, oil content
will be further reduced to about 1–3 wt% in oil meal (Kolesarova et al. 2011).
Oil cake/meal can be classified into two categories including, edible and
nonedible. Edible ones have a high nutritional value, especially containing protein
in the range of 15–50 wt%. The variety, growing conditions, and extraction
methods are the effective parameters on the composition of these by-products
(Ramachandran et al. 2007). High amounts of protein make them suitable animal
feeds while the nitrogen, phosphorous, and potassium contents of nonedible types
such as castor, karanja, and neem make them suitable organic nitrogenous fertil-
izers. Some of these oil cake/meal have been found to increase the nitrogen uptake
of plants, as they retard the nitrification of soil. They also protect the plants from
soil nematodes, insects, and parasites; thereby offer great resistance to infection
(Ramachandran et al. 2007). Nevertheless, alternative applications for oil cake/meal
should be sought considering the overproduction of oil cake/meal and the increased
emphasis placed on cost reduction throughout biodiesel production process
(Ramachandran et al. 2007). Different applications of oil cake/meal are presented in
Fig. 5.2.

Fig. 5.2 Different applications of oil cake/meal


104 H. Rastegari et al.

5.2.1 Animal Feed

Oil cake/meal is rich in protein and so they can be used predominantly for feed
applications in poultry, ruminants, fish, and swine industries (Ramachandran et al.
2007; Alshelmani et al. 2016; Zhou et al. 2016; Hamdi et al. 2016). The inclusion
of fermented palm kernel cake in poultry diets was investigated where 15 wt% of
the fermented cake could replace up to 30 wt% of yellow corn to reduce the costs of
feed consumption in the poultry industry (Alshelmani et al. 2016). Addition of palm
kernel cake to lamb supplements favorably affected the fatty acids profile of the
muscles in ruminant animals (Freitas et al. 2017). This was ascribed to the fact that
meat from ruminant animals typically contains mostly saturated fatty acids in the
range of 40–60 wt%, which are associated with coronary heart disease in humans.
Thus, increasing the content of polyunsaturated fatty acids could lead to meat
products perceived as healthier (Freitas et al. 2017). Addition of palm kernel cake to
lamb supplements up to the level of 30 wt% was found useful as the produced meat
could be considered healthy for human consumption without posing the risk of
increasing saturated fat levels in diet and the consequent health consequences
(Freitas et al. 2017). Olive cakes inclusion in dairy diets reportedly increased
mono-unsaturated fatty acids content and lowered saturated fatty acid content in the
produced milk, however, high levels of Cu in olive leaves could restrict the use of
this by-product in practical feeding (Molina-Alcaide and Yáñez-Ruiz 2008).

5.2.2 Fertilizers/Pesticides

Some oil cake/meal like brassicaceae, sunflower, and Ethiopian mustard, are con-
sidered to be suitable organic nitrogenous fertilizers because of their nitrogen,
phosphorous, and potassium contents. This application may represent an opportu-
nity to reduce the environmental impact related to chemical fertilizer production
and to improve the fertility of soil through organic carbon supply and soil biological
activity strengthening (Mazzoncini et al. 2015). Moreover, some oil cake/meal such
as those of Brassicaceae oilseeds contain a series of biologically active compounds,
mainly isothiocyanates, that have been shown effective in controlling weeds, insect
pests, nematodes, and soil-borne pathogens (Mazzoncini et al. 2015). Overall,
repeated application of oil meals could increase the organic nitrogen of the soil and
reduce nitrate leaching to the same level as the unfertilized system, in comparison
with ammonium nitrate supply (Mazzoncini et al. 2015).
5 Applications of Biodiesel By-products 105

5.2.3 Biotechnological Applications

Oil cake/meal can be utilized as a potential raw material in bioprocesses as an excellent


substratum for the growth of microorganism supplying them with essential nutrients
(Ramachandran et al. 2007). They have been widely used for the production of amino
acid, enzymes, mushrooms, ethanol, organic acids, single-cell protein, antibiotics,
bio-pesticides, vitamins, and other bio-chemicals (Ramachandran et al. 2007).
High protein content and availability are the main factors for selection of oil
cake/meal as the starting material to produce different amino acids. Essentially the
process consists of isolating the protein and enzymatically digesting the isolate. The
final product contains almost the essential and non-essential amino acids depending
on the used oil cake/meal (Ramachandran et al. 2007; Murti 1965; Sarker et al.
2015). Oil cakes can be used as substrate for enzyme production using fungal
species in solid state fermentation. They are ideally suited nutrient supports ren-
dering both carbon and nitrogen sources. Lipase (Oliveira et al. 2017), a-amylase
(Ramachandran et al. 2004), phytase (Sabu et al. 2002), protease (Mahanta et al.
2008), and glutaminase (Kashyap et al. 2002), tannase (Sabu et al. 2005) are some
of the enzymes produced using oil cakes as nutrient source.
The effect of supplementing spent rice straw substrate with extra organic
nitrogen (in the form of oil cakes) was studied for production of mushroom
Pleurotus sajor-caju. Supplemented substrate proved to be better in enhancing the
mushroom yields (up to 12 times vs. those grown on un-supplemented spent straw).
These mushrooms showed increased protein, fat, and decreased carbohydrate
contents (Sabu et al. 2005; Shashirekha et al. 2002; Bano et al. 1993).
Palm kernel cake has been shown to be a promising feedstock for production of
bioethanol due to its high carbohydrate content, facile disruption and susceptibility
to enzymatic hydrolysis (Raita et al. 2016). Ethanol yields of greater than 90 wt% of
the theoretical maximum have been achieved by a process involving mild steam
explosion pretreatment, enzymatic hydrolysis by the core mannanase and cellulase,
and fermentation of the oligosaccharide-rich hydrolysate by Geobacillus ther-
moglucosidasius TM242 (Raita et al. 2016). Such a conversion process provides the
basis for an efficient and economic production of ethanol from palm kernel cake as
part of an integrated biorefinery. Moreover, as a low-cost lignocellulosic feedstock,
palm kernel cake could be considered as a potential source of fermentable sugars for
further application in biobutanol production (Shukor et al. 2016). Organic acids
production could also be accomplished through fermentation processes of oil cake/
meal (Moresi and Parente 2014; Jain et al. 1989). Oil cake/meal has also been
reportedly used as carbon source in production of antibiotics and antimicrobials
(Farzana et al. 2005; Vidyarthi et al. 2002).
106 H. Rastegari et al.

5.2.4 Mosquito Repellent

Disposal of nonedible oil cakes of jatropha and karanja is a major concern because
they contain some toxic components. Due to their toxic nature, these cakes can
neither be used as animal feed nor can be used as fertilizer. While these waste oil
cakes can be used to develop formulations as green pesticides for mosquito control
(Pant et al. 2016). The efficacy of these cakes against mosquitoes is due to the
presence of active compounds such as karinjin in karanja and phorbol esters in
jatropha (Pant et al. 2016).

5.2.5 Other Applications

Some oil cakes including linseed, mustard, and neem have been used for growing
an endo-parasite of nematodes (Gupta et al. 2005). Sesame and mustard oil cakes
have also been used as substitute for animal protein hydrolysate, used in the
treatment of protein malnutrition, because of their high crude protein content
(Ramachandran et al. 2007). Considering their high oil content, oil cakes have high
energetic values and therefore, they could be suitable substrates for combustion or
pyrolysis, however because of the large quantity of ash and high emissions of
nitrogen oxides, advanced purification technology would be required (Yorgun et al.
2001). Since oil cakes and meals contain a high portion of digestible substances,
some of them have also been used as substrates for the production of biogas (Barik
and Murugan 2015; Kanchanasuta and Pisutpaisal 2016; Jabłoński et al. 2016).
Biogas from de-oiled cakes can provide energy for heating, cooking, lighting, and
engine applications. In addition, the digested slurry of the cakes containing nitrogen
and phosphorus can be directly used in the agricultural sector as fertilizer
(Kolesarova et al. 2011; Barik and Murugan 2015).

5.3 Glycerol Applications

Glycerol is a triol present in the form of its esters (triglycerides) in oil feedstocks. It
is generated at about 10 wt% as the main by-product during the biodiesel pro-
duction process (Johnson and Taconi 2007). The continued increase in the pro-
duction of biodiesel, produces excessive amounts of glycerol which has encouraged
research to develop new applications for glycerol. In its pure form, glycerol has
diverse applications in different fields, especially in the pharmaceuticals, food, and
cosmetics industries (Johnson and Taconi 2007). However, the glycerol-rich stream
generated during biodiesel production, named crude glycerol has a purity in the
range of 60–80 wt% (Gholami et al. 2014). This crude glycerol contains many
impurities such as methanol, organic and inorganic salts, water, vegetable colors,
5 Applications of Biodiesel By-products 107

fatty acid methyl esters, free fatty acids, mono and di-glycerides, soap, etc. (Xiao
et al. 2013). The purification of crude glycerol is an expensive process, requiring
expensive processing equipment (Xiao et al. 2013). Hence, it is economically viable
only for the large-scale biodiesel producers to refine the crude glycerol, while many
small-scale producers lack the resources to refine the crude glycerol (Pagliaro and
Rossi 2010). On the other hand, since the crude glycerol contains methanol, it
cannot be safely discarded into the environment. Therefore, proper disposal options
are essentially limited to anaerobic digestion or transportation to larger biodiesel
plants where necessary refining could be performed (Pagliaro and Rossi 2010). To
overcome this challenge, much research has focused on the utilization of crude
glycerol as a building block in numerous reaction pathways to synthesize
value-added chemicals some of which are reviewed in this section.

5.3.1 Synthesis of Various Chemicals

Different derivatives of glycerol have been synthesized by utilizing glycerol and


different organic reactants (Len and Luque 2014; Bagheri et al. 2015). Processes for
glycerol valorization can be broken down in two classes: (1) oxidation or reduction
of glycerol into other three carbon compounds, (2) reaction of glycerol with other
molecules to form new species.
Figure 5.3 presents the block diagram of different reactions of glycerol. One of
the reactions in which glycerol can be converted into valuable products is glycerol
oxidation. Glycerol oxidation is a complex pathway of reactions yielding a range of
possible products, namely dihydroxyacetone, hydroxypyruvic acid, glyceric acid,
tartaric acid, oxalic acid, mesoxalic acid, glyoxylic acid, glyceraldehyde, and gly-
colic acid (Bagheri et al. 2015). Dihydroxyacetone is used as a tanning agent in the
cosmetics industry and as an active substance in sunless tanning lotions. Beside
these main utilizations, it also has other applications for instance as a monomer in
polymeric biomaterials. Hydroxypyruvic acid is a flavor component in cheese and
has been previously used for synchronization of fruit maturation. Dihydroxyacetone
and hydroxypyruvic acid are possible starting materials for D, L-serine synthesis
(Zheng et al. 2008).
The oxidation of the primary hydroxyl groups of glycerol produces glycer-
aldehyde, leading to further oxidation producing carboxylic acids like glyceric acid,
tartronic acid, and mesoxalic acid. Glyceric acid can be used for treatment of skin
disorders, and as an anionic monomer of packaging material for exothermic and
volatile agents. In its ester form, glyceric acid can act together with a quaternary
ammonium salt as an effective and biodegradable fabric softener. Tartronic acid can
be used as a potentiating agent or adjuvant to increase the blood absorption of a
tetracycline antibiotic. Moreover, it can be used to scavenge dissolved oxygen in
alkaline water. Mesoxalic acid has a potential for use as a complexing agent and as
a precursor in organics synthesis and has been recently found to possess anti-HIV
activity (Zheng et al. 2008; Villa et al. 2015). It should be noted that these
108 H. Rastegari et al.

Fig. 5.3 Block diagram for


different reactions through
which biodiesel glycerol can
be converted into valuable
products

derivatives have a limited market due to the costly nature of the conventional
production processes involved (Bagheri et al. 2015). On the contrary, for glycerol
oxidation, inexpensive oxidizing agents such as air, oxygen, hydrogen peroxide, or
bleach can be used (Zheng et al. 2008), and therefore, by using these oxidizing
agents and the inexpensive feedstock, i.e., glycerol, the process costs could be
reduced substantially. This would in turn facilitate further utilization of these
valuable derivatives at industrial scale (Villa et al. 2015).
Glycerol hydrogenolysis is another reaction employed for glycerol valorization.
The removal of the side –OH group of glycerol forms 1,2-propandiol, while dis-
sociation of the middle –OH group generates 1,3-propandiol. Competing reactions
in glycerol hydrogenolysis include breaking one C–C bond to form ethylene glycol.
Over-hydrogenolysis of C–C and C–O bonds in glycerol or in its derived products
will further generate monobasic alcohols and alkanes such as propanol, ethanol,
methanol, and methane (Wang et al. 2015). Another attractive methodology for
glycerol valorization is aqueous phase reforming in the absence of any external
hydrogen addition leading to the production of propanediols, important commodity
5 Applications of Biodiesel By-products 109

chemicals (Roy et al. 2010). For instance, 1,3-Propanediol is an important diol in


the production of polyesters, polycarbonates, and polyurethanes. Moreover, it is
also used in the manufacturing of carpet and textile fibers exhibiting unique
properties in terms of chemical resistance, light stability, elastic recovery, and dye
ability (Wang et al. 2015). 1,2-propanediol is often used as a less toxic alternative
used in nutrition products, solvent for colorings and flavors, wetting agent in
tobaccos, additive in cosmetics, as well as a constituent of break or hydraulic fluids,
lubricants, and anti-freezing agents. It is also extensively used as feedstock in the
preparation of polyester resins, in fiber manufacturing, manufacturing and in
pharmaceutical industry. Furthermore, it can be used as a starting material for
solvents, emulsifiers, and plasticizers (Villa et al. 2015).
Two important chemicals can be produced directly by dehydration of glycerol,
namely acrolein and 3-hydroxypropionaldehyde. However, dehydration of glycerol
is not easy, since the C=C bond is thermodynamically more favorable than the C=O
bond (Bagheri et al. 2015). Acrolein is an important bulk chemical used to produce
acrylic acid, glutaraldehyde, methionine, acrylic acid esters, super absorber poly-
mers, and detergents. Acrolein can be used as an herbicide to control the growth of
aquatic plants without further modifications (Zheng et al. 2008). Most indirect
application of acrolein concern the synthesis of acrylic acid, which is used, for
example, as a starting material for synthesizing superabsorbent polymers
(Katryniok et al. 2010). The second largest indirect application of acrolein is in the
synthesis of DL-methionine, an essential amino acid, which cannot be synthesized
by living organisms (Katryniok et al. 2010). 3-Hydroxypropionaldehyde has
applications in healthcare, food industry, and in the chemical industry (Zheng et al.
2008). Due to its broad antimicrobial activity, it is used as a probiotic strain in
human and animal diets. Moreover, since it is capable of inhibiting the growth of
several human pathogens, it is regarded attractive for use as a food preservation and
conservation agent. 3-Hydroxypropionaldehyde is also a useful chemical in the
fixation of biological tissues (Vollenweider and Lacroix 2004).
Among the various chemical routes aimed at glycerol valorization, etherification
is capable of producing oxygenated additives for diesel fuel. One of the important
additives is produced through etherification of glycerol with isobutylene. The
mixtures of mono-, di-, and tri-alkyl glycerols are suitable for use as oxygenates in
diesel fuel (Bagheri et al. 2015). The etherification of glycerol generate polyglyc-
erols; low polymerization level and oxygenated compounds (Zheng et al. 2008).
Polyglycerols and polyglycerol esters are gaining prominence in products such as
surfactants, lubricants, cosmetics, and food additives (Zheng et al. 2008). Glucosyl
glycerol is another ether of interest found in Japanese traditional fermented foods
such as sake, miso, and mirin and its use in foods, beverages, and cosmetics is
expected to increase. It can also suppress the action of intestinal disaccharides in
rats and may be useful in decreasing caloric intake (Zheng et al. 2008).
In recent years, esterification of glycerol has also been a particularly active area
of research in which mono-, di- and tri-esters are formed. Glycerol monoesters are
useful as nonionic surfactants and emulsifiers. They are widely used in bakery
products, margarines, dairy products, as well as sauces. In the cosmetic industry,
110 H. Rastegari et al.

they are added as texturing agents for improving the consistency of creams and
lotions. In addition, because of their excellent lubricant and plasticizing properties,
they are used in textile processing and formulation of oils for various types of
machinery (Zheng et al. 2008). Glycerol diesters have been utilized as a cocoa
butter blooming agent and as an intermediate in the synthesis of structural lipids
(Zheng et al. 2008). The catalytic esterification of glycerol with acetic acid produces
glycerol acetates, namely monoacetin, diacetin and triacetin (Bagheri et al. 2015;
Rezayat and Ghaziaskar 2009; Rastegari and Ghaziaskar 2015; Rastegari et al.
2015; Shafiei et al. 2017). Monoacetin is used in the manufacturing of dynamite,
tanning leather, and cryogenics, while it is also regarded as raw material for pro-
duction of biodegradable polyester, food additives, explosive, and smokeless
powder (Bagheri et al. 2015). Diacetin is used as solvent for various dyes, plasti-
cizer, as well as in the preparation of softening agents and printing ink (Bagheri
et al. 2015). Triacetin is used as antiknock additives for gasoline, as cold and
viscosity improver of biodiesel, while it is also used in the production of photo-
graphic films, and in the perfumery industry (Bagheri et al. 2015). Recently, the
transesterification of glycerol with methyl- and ethyl- acetate was introduced as an
alternative pathway for the synthesis of glycerol acetates (Khayoon and Hameed
2013).
Another important series of glycerol derivatives, are glycerol carbonates and
dicarbonate. A typical method for obtaining carbonate derivatives of glycerol is its
transesterification with ethylene carbonate or di-alkyl carbonate. Glycerol carbon-
ate, a stable and colorless liquid, is a useful solvent for plastics and resins, such as
cellulose acetate, nylon, nitrocellulose, and polyacrylonitrile. It reacts readily with
phenols, alcohols, and carboxylic acids when heated, forming the glycerol ethers or
esters of these materials, for example, polyesters, polycarbonates, polyurethanes,
and polyamides (Zheng et al. 2008).
Glycerol can be treated with nitrating agents to form a solution containing
dinitroglycerol. The solution is treated with a cyclizing agent to convert dinitro-
glycerol into glycidyl nitrate, which is then polymerized into poly-(glycidyl nitrate).
Poly-(glycidyl nitrate) has been recognized as an energetic polymer potentially
suitable for use in propellants, explosives, gas generators, and pyrotechnics (Zheng
et al. 2008). Another promising derivatives of glycerol are cyclic acetals and ketals
produced in the reaction of glycerol with aldehydes and ketones, respectively
(Nanda et al. 2014a, b, c, 2016; Shirani et al. 2014; Gorji and Ghaziaskar 2016).
Various acetals and ketals can be used as ignition accelerators and antiknock
additives. A second application of glycerol acetals is the use as scent or flavor, e.g.,
the acetalization of phenyl acetaldehyde or vanillin with glycerol leading to hya-
cinth or vanilla flavors. Glycerol acetals can also be used as the basis for surfac-
tants. The properties of these acetals and ketals can be changed through
transacetalization with long chained molecules to produce very polar as well as very
nonpolar acetals and ketals.
Studies of glycerol halogenation have focused on the production of
1,3-dichloropropanol, an intermediate in epichlorohydrin synthesis. Epichlorohydrin
is a colorless liquid with a pungent, garlic-like odor, and is moderately soluble in
5 Applications of Biodiesel By-products 111

water, but miscible in most polar organic solvents (Anitha et al. 2016). It is an
important raw material for the production of materials such as epoxide resins,
synthetic elastomers, and sizing agents for the papermaking industry (Zheng et al.
2008). The preparation of epichlorohydrin form glycerol shows significant advan-
tages compared with the present process based on propene: First of all the feedstock
is renewable and the water consumption is lower. In addition to that, the chlorinated
residues are reduced and the chlorination agent is hydrochloric acid instead of the
more expensive chlorine used in the conventional process (Anitha et al. 2016).
Pyrolysis of glycerol has also been studied in steam and supercritical water with
acrolein, formaldehyde, and acetaldehyde as the major products at lower temper-
atures. These products appear to result from dehydration and fragmentation of
glycerol. At higher temperatures, other products such as carbon dioxide, molecular
hydrogen, ethylene, and methane are formed, indicative of more complex chemistry
(Zheng et al. 2008). Gasification of glycerol for the production of hydrogen and
syngas for energy purposes has also been explored (May et al. 2010). In addition to
the above reactions, glycerol has also been used in the aqueous phase reforming
process. This process is carried out at quite low temperatures and moderate pres-
sures, producing different chemicals (gases and liquids). The gas phase has a high
H2 content while the liquid phase is a complex mixture of different organic com-
pounds such as alcohols, ketones, acids, esters, paraffin, aldehydes, and other
oxygenated compounds in water (Remón et al. 2016).

5.3.2 Production of Hydrogen Gas

Hydrogen is a clean energy source whose demand has increased in recent years.
The combustion products of hydrogen when burned completely with air include
water, oxygen, and nitrogen (Avasthi et al. 2013). In comparison with gasoline, the
emission of nitrogen oxides is far less. Hydrogen can be generated in many different
ways, such as glycerol conversion. Theoretically, 1 mol of glycerol can produce up
to 4 mol of hydrogen. Available methods to convert glycerol into hydrogen are
steam reforming, partial oxidation gasification, auto thermal reforming, aqueous
phase reforming, and supercritical water reforming processes (Avasthi et al. 2013).
In the steam reforming process, glycerol reacts with water vapor in the presence
of a catalyst such as platinum, palladium, and nickel, producing mainly hydrogen,
carbon dioxide, and carbon monoxide. This process is the most preferred process
because hydrogen is simultaneously removed from water, which increases the yield
of the reaction (Schwengber et al. 2016). However, the high consumption of energy
for vaporization of the reaction mixture reduces the energy efficiency of the process.
Moreover, a large amount of water is necessary to facilitate the gasification of
carbon, preventing its deposition in the form of coke on the catalyst (Schwengber
et al. 2016).
Partial oxidation reforming of the glycerol occurs under atmospheric pressure,
with quantities of oxygen below the optimal stoichiometric value for complete
112 H. Rastegari et al.

combustion (Schwengber et al. 2016). This process is an exothermic process whose


efficiency depends directly on controlling the amount of oxygen that enters the
mixture. Because it involves rapid consumption of oxygen and high temperatures,
the reaction causes various parallel reactions (Schwengber et al. 2016). The stoi-
chiometric coefficients in auto thermal reforming differ from those of partial oxi-
dation reforming due to the inclusion of water vapor in the combustion process,
which could increase the production of hydrogen. The main expected advantages
are high efficiency and compactness of the hydrogen generator system (Schwengber
et al. 2016).
Aqueous phase reforming is usually performed at high pressures and moderate
temperatures in a continuous stream, transforming glycerol into aqueous phase
without pre-vaporization. In this process, first, glycerol is decomposed, after which
it can be converted into carbon dioxide and hydrogen by the water–gas shift
reaction. Hydrogen can be consumed by intermediate compounds such as carbon
monoxide and hydroxides, and by dehydration reactions (Schwengber et al. 2016).
The main advantage of this process is that it takes place in the liquid phase, unlike
other available technologies in which the processes occur in the gas phase, plus the
fact that most biomass-based liquids do not vaporize easily (Schwengber et al.
2016).
Supercritical water is an emerging and promising medium to obtain hydrogen by
reforming of glycerol. It stands out for operating with a high-pressure and low-
temperature system, in which the water is at its critical point (374 °C, 218 atm),
which provides interesting properties such as low viscosity, high diffusivity, and
low dielectric constant (Schwengber et al. 2016). This process is advantageous
because the reactivity of water increases at the supercritical point, and the presence
of catalyst favors the reaction. Another advantage is that at temperatures above
600 °C and pressures higher than its critical point, water becomes a strong oxidant.
As a result, carbon is preferentially oxidized to CO2 and low concentrations of CO
are formed. The hydrogen atoms from water and glycerol form H2. The product is
composed of H2, CO2, CH4, and CO (Schwengber et al. 2016). These reforming
processes reveal that obtaining H2 from glycerol is favorable, resulting in high
yields and conversions. Among these methods, steam reforming is the most widely
applied method in the chemical industry. Overall, the hydrogen produced from raw
glycerol is a good alternative to efficiently valorize the excess of glycerol generated
by biodiesel industries. It should be noted that the generated hydrogen can also be
used in fuel cells to produce energy in the form of heat and electricity (Schwengber
et al. 2016).

5.3.3 Fuel Additives

Branched oxygen-containing molecules can be used as fuel additives in order to


improve fuel properties (Anitha et al. 2016). Transformation of glycerol into
oxygenated compounds through etherification, esterification, transesterification, and
5 Applications of Biodiesel By-products 113

acetalization/ketalization is a potential route for glycerol valorization (Mota 2012;


Climent et al. 2014). The glycerol-based ethers mixture is used as flash point
reducer for diesel, with a blending ratio of 10% (v/v). This mixture affects the
combustion process and has been found effective in reducing particulate matter
emissions (Beatrice et al. 2014). Melero et al. (2010) investigated the effects of
oxygenated compounds produced from glycerol on the quality parameters of soy-
bean biodiesel. The glycerol-derived oxygenated compounds studied were: a
mixture of tert-butylated glycerol (5 wt% mono-tert-butyl glycerol, 55 wt%
di-tert-butyl glycerol, 38 wt% tri-tert-butyl glycerol), a glycerol ketal
(2,2-dimethyl-4-hydroxymethyl-1,3-dioxolane 98 wt%), triacetin, and a mixture of
glycerol esters (6 wt% monoacetin, 45 wt% diacetin, 47 wt% triacetin). The mixture
of tert-butylated glycerol obtained from the etherification of glycerol with iso-
butylene reportedly led to the best performance. However, it should be noted that
unfortunately, isobutylene availability in refineries is usually very limited (Melero
et al. 2010).
Esterification of glycerol with acetic acid results in the production of mono-
acetin, diacetin, and triacetin. These derivatives are useful intermediates/products as
biofuel additives. In particular, triacetin is used as fuel additive for gasoline to
increase its octane number and to improve certain properties of biodiesel.
Monoacetin is an intermediate used to synthesize a new ketal derived from glycerol
as feedstock, namely, 2,2-dimethyl-1,3-dioxolan-4-yl methyl acetate (Gorji and
Ghaziaskar 2016). Addition of this ketal to biodiesel could not only improve the
viscosity of the fuel but also could assist in meeting the requirements established by
the European and American Standards (EN 14214 and ASTM D6751, respectively)
for other important parameters, such as flash point and oxidation stability. It should
be mentioned that this ketal does not improve cold properties as much as triacetin
does (Garcia et al. 2008).
Acetalization of glycerol with benzaldehyde produces branched five and six
membered acetals, 2-phenyl-1,3-dioxolane-4-yl and 2-pheny-1,3-dioxan-5-ol,
respectively. These acetals are versatile additives for diesel fuels. The addition of
glycerol/acetone ketal, i.e., solketal, to gasoline, in the range of 1-5% (v/v),
reportedly led to a significant decrease in gum formation, indicating a potential
antioxidant property of this glycerol derivative, while it also improved the octane
number of the gasoline (Mota et al. 2010).

5.3.4 Feedstock for Biological and Non-biological Processes

Except the above products, glycerol is well known as a carbon source for biological
processes such as anaerobic digestion and fermentation. In anaerobic digestion,
biogas is generated while acetate, propionate, and butyrate are common interme-
diates (Chen et al. 2016). In this process, a significant methane yield can be
observed (Lopez et al. 2009). Crude glycerol has the potential to be used as a
co-substrate for anaerobic digestion of sewage sludge (Silvestre et al. 2015).
114 H. Rastegari et al.

Glycerol can be fermented into many types of products, such as 1,3-propanediol,


acetate, butyrate, butanol, ethanol, lactate, hydrogen, and other organic acids
(Dabrock et al. 2012; Rossi et al. 2012; Oh et al. 2011).
In addition to that, since glycerol is non-toxic, nonflammable and non-volatile,
and has high energy density (6.26 kWh L−1), it could be used for power generation
in non-biological processes such as fuel cells. In these fuel cells it is possible to
control the activity and selectivity of the catalysts toward a given product. The
beneficial features of direct oxidation of glycerol in fuel cell applications are the
generation of power and the production of oxygenated products, which are costly
and challenging to produce by catalytic and biological mechanisms (Gallezot 1997;
Ilie et al. 2011).

5.3.5 Wastewater Treatment

Nitrate is one of the most abundant ground and surface-water contaminants which is
hazardous to human health. Photocatalytic reduction processes are promising routs
for nitrate contaminated water remediation. In these processes, solar energy is
exploited to convert nitrate into non-toxic final by-products such as nitrogen gas
instead of undesirable by-products requiring additional treatments (Lucchetti et al.
2017). Photocatalytic nitrate reduction is normally carried out in the presence of
organic molecules acting as hole scavengers. Formic acid, oxalic acid, methanol, or
ethanol have been generally used as hole scavengers (Lucchetti et al. 2017). The
highest reaction rates obtained are generally attributed to very polar organic hole
scavengers. In line with that, glycerol, the major by-product of the biodiesel
manufacturing process, was also investigated as a hole scavenger (Lucchetti et al.
2017). Moreover, the potential of glycerol as a carbon source in the removal of
nitrates, nitrite, and phosphorus were examined by several studies (Akunna et al.
1993; Bernat et al. 2015; Costa et al. 2018; Yang et al. 2018). The results obtained
showed that the denitrification efficiency was nearly 100% in the presence of
glycerol as a carbon source (Bodík et al. 2009; Guerrero et al. 2011; Cyplik et al.
2013).

5.4 Methanol Applications

One of the key roadblocks to the development of a biodiesel production process is


the reversibility of the transesterification reaction when full conversion is desired.
Hence, most of the commercial processes are conducted using excess methanol. To
decrease the overall methanol consumption in the process, methanol recycling
would be necessary. Therefore, excess methanol is generally separated from the
polar phase (crude glycerol) through distillation. It can then be reused in the next
cycles of the biodiesel process (Cao et al. 2008).
5 Applications of Biodiesel By-products 115

5.5 Biodiesel Wash-Water Applications

After the transesterification reaction, crude biodiesel is separated from the glycerol
phase. The biodiesel phase contains some impurities including acylglycerols, free
fatty acids, soap, residual alcohol, catalyst, free glycerol, and salts (Daud et al.
2015). These impurities can reduce the quality of biodiesel and negatively affect
engine performance (Atadashi et al. 2011) by increasing injector fouling and the
emissions of aldehydes and acrolein and by decreasing engine durability (Berrios
and Skelton 2008). Hence, biodiesel must be purified before being used as diesel
fuel in order to meet the requirements of the international quality standards such as
EN 14214. Washing process is commonly used for purification to enhance biodiesel
quality. As industrial level, this process is commonly performed via two techniques:
wet washing or dry washing (Berrios and Skelton 2008). In the wet washing
process, distilled/deionized hot water (50–60 °C), mineral acid (pH of 2–5)
(Atadashi et al. 2011), organic solvents such as petroleum ether (Siler-Marinkovic
and Tomasevic 1998), and solution of citric acid in methanol 50% (Tomasevic and
Siler-Marinkovic 2003) can be used to remove impurities. This methodology has a
good purification performance, however, the introduction of additional water to the
process offers many disadvantages including increased cost and production time as
well as the generation of biodiesel wash-water (wastewater) (Berrios and Skelton
2008). In the dry washing process, contaminants are removed from crude biodiesel
by adsorption or passing crude biodiesel through a bed of ion exchange resin
(Atadashi et al. 2011). Different adsorbent materials are used for treating crude
biodiesel such as magnesium silicate or Magnesol (Low et al. 2011), calcium
magnesium silicate, and other inexpensive bio-sorbents. BD10 Dry and Purolite
PD206 are two types of ion exchange resins used for biodiesel dry purification
(Berrios and Skelton 2008). The advantage of this treatment method is that the
impurities are absorbed to form a solid waste product instead of a liquid, and
therefore, no wastewater is produced (Leung et al. 2010; Atadashi et al. 2011;
Berrios and Skelton 2008). However, at the present time, these adsorbents are not
recycled because the cost of recycling is about the same as that of buying new ones.
Therefore, the spent material are disposed of to landfill or used for other applica-
tions (Veljković et al. 2014). Magnesium silicate is not toxic and the contaminants
absorbed from biodiesel have nutritional values, therefore, it can be used as com-
post and animal feed additive. Although dry washing process offers the advantage
of being waterless, it is reported that the products obtained through this process
hardly meet the requirements of the EN 14214 standards (Leung et al. 2010). On
that basis, water washing is still the most frequently used process for biodiesel
purification in which successive water washing steps are employed (Daud et al.
2015).
Depending on the quantity of the impurities contained in biodiesel, washing
steps should be repeated 2–5 times. Under the conventional process (alkali cat-
alyzed transesterification) for every 100 L biodiesel produced, about 20 L of raw
wastewater is discharged, assuming no prior acid pretreatment to remove free fatty
116 H. Rastegari et al.

acids which otherwise will increase this value (Kolesarova et al. 2011). The raw
biodiesel wash-water is a viscous liquid with an opaque white color and contains
high amounts of oil and grease and low contents of nitrogen and phosphorous
(Jaruwat et al. 2010; Ngamlerdpokin et al. 2011; Chavalparit and Ongwandee
2009). This wastewater has high levels of COD in the range of 18–800 g L−1 due to
the presence of different impurities including glycerol, soap, methanol, FFAs,
catalyst and a portion of methyl esters as well as trace mono-, di- and triglycerides
(Ngamlerdpokin et al. 2011; Chavalparit and Ongwandee 2009; Srirangsan et al.
2009). Biological treatment of this wastewater is difficult because both high levels
of COD and low nitrogen concentrations could inhibit the growth of most
microorganisms (Srirangsan et al. 2009; Kolesarova et al. 2011). Since biodiesel
manufacturing process results in large amounts of such highly polluting wastewater,
its treatment should be carried out effectively. On the other hand, the resultant
treated water could be reused in the biodiesel production process and improve the
economic viability and water footprint of the whole process. Alternatively, this
wastewater in its raw form (untreated) could be used as a feedstock for the pro-
duction some value-added products.

5.5.1 Biodiesel Wash-Water Reuse

Several treatment processes have been developed to treat biodiesel wash-water such
as physicochemical treatments, electrochemical treatments, coupled chemical and
electrochemical treatments, advanced oxidation technologies, biological treatments,
and integrated treatment processes (Veljković et al. 2014). The physicochemical
treatment techniques are applied for the removal of oil and greases, organic and
inorganic components, difficult to decompose non polar organic substances, high
salt concentration, etc. This process involves adsorption, acidification (pH adjust-
ment), and coagulation/flocculation processes or their combination, and are even-
tually followed by a physical treatment such as sedimentation, filtration, or
floatation (Veljković et al. 2014). Adsorption is a versatile, easily operated physical
treatment process in which various type of adsorbents, including peat, bentonite
clay, activated carbon, agricultural waste, and chitosan are used (Zhang et al. 2010).
In this process the adsorbed pollutants from wastewater creates a film of pollutant
molecules on the surface of the solid particles of adsorbent, suspended in a stirred
tank or packed in a column. In most cases, adsorption is not used alone for biodiesel
wastewater treatment.
Acidification is usually used as a pretreatment stage prior to other treatment
procedures used on biodiesel wastewater. In the acidification technique, pH
adjustment is performed by adding an acid to destabilize and destroy oil emulsions
in wastewater through changing the chemical structure of the functional groups.
Moreover, acidification breaks up soaps into a salt and a free fatty acid. In this way,
oily impurities are separated and residual oil and free fatty acids are recovered prior
to the coagulation process. Different chemicals are used to adjust the pH before the
5 Applications of Biodiesel By-products 117

coagulation process, such as acids or calcium oxide (Veljković et al. 2014).


Coagulation and flocculation occur in successive steps. Coagulation is in fact a
chemical process that involves neutralization of charges whereas flocculation is a
physical process. Coagulation is performed by adding a coagulant to wastewater to
destabilize colloidal suspensions in a reasonable time. At this step, a rapid mixing is
needed to disperse the coagulants uniformly in the aqueous solution. Then, the
small destabilized particles are flocculated into larger, settleable aggregates, under
the conditions of gentle mixing. The most widely used coagulants are inorganic
salts of iron and aluminum (e.g., ferry chloride or aluminum sulfate) while a
pre-polymerized inorganic compound (e.g., poly-aluminum chloride) is used to
clump fine particles together into larger aggregates (Daud et al. 2015). Coagulation
and flocculation can be used as a preliminary or intermediary step between other
treatment methods. pH adjustment can enhance the efficiency of this process.
Hence, coagulation/flocculation is frequently preceded by pH adjustment.
A common sequence of operations is pH adjustment, coagulation/flocculation, and
sedimentation or floatation (Veljković et al. 2014).
Another physical operation to remove dispersed oil drops or solids is dissolved
air floatation. The removal is achieved by dissolving compressed air in wastewater
and then releasing the air at atmospheric pressure. The released air form tiny
bubbles, which adhere to oil drops and solids, could cause the suspended matters to
float to the surface of the water where they are subsequently removed by a skimmer
(Veljković et al. 2014). This method is generally employed for treating wastewater
before biological treatment and is considered effective for removing solid particles
of low density from suspensions and clarifying low turbidity values. Moreover, it is
successful for separating oily materials from aqueous emulsions after chemical
pretreatment (Veljković et al. 2014).
Electrochemical techniques may offer an attractive alternative as they are more
environmentally friendly processes performed in the absence or reduced amounts of
chemicals and indeed electrons are the only reactants added to the process. Other
advantages of these techniques are simple equipment, easy operation, short treat-
ment time, fast sedimentation, and smaller sludge generation (Veljković et al.
2014). Electrochemical treatment is performed as electrocoagulation and
hydrothermal electrolysis (Veljković et al. 2014). Electrocoagulation is used to
remove contaminants that are generally more difficult to remove by filtration or
chemical treatment systems such as emulsified oil and suspended solids. This
technology removes impurities from wastewater by introducing highly charged
polymeric metal hydroxide species. Anode of the electrolytic cell is usually alu-
minum in which the oxidation occurs while water molecules are reduced in the
cathode. Metal and hydroxyl ions produced in the anode and cathode react in the
aqueous medium and form amorphous metallic hydroxide species with large sur-
faces. These species neutralize the electrostatic charges on suspended solids and oil
droplets to facilitate the agglomeration or coagulation. Then precipitates are
removed either by sedimentation or floatation by using hydrogen produced at the
cathode (Veljković et al. 2014). The electrocoagulation treatment is often combined
with a chemical treatment to improve its efficiency. The combined treatment can be
118 H. Rastegari et al.

conducted in two ways. In the one-step process, the electrochemical treatment is


carried out in the presence of a coagulant and/or an oxidant. While in the two step
process, first acidification of the biodiesel wastewater is carried out by adding a
mineral acid to recover crude biodiesel, and then the electrochemical treatment is
carried out as the second step. Usually, a biological post-treatment of the effluent
wastewater is needed because the process cannot reduce the levels of contaminants
such as grease and oil, COD and BOD below the permitted limits (Veljković et al.
2014).
In the hydrothermal electrolysis, wastewater is converted into valuable products
using subcritical water as the reaction medium (Veljković et al. 2014). When
electrolysis is performed, the water vapor molecules around the anode are ionized
or activated and then bombard each other, generating ions, free hydroxyl radicals,
and sometimes hydrogen atoms. In addition, in the liquid phase reaction zone,
several liquid water molecules are broken and reformed into gaseous products
(hydrogen, hydrogen peroxide and oxygen). Therefore, several oxidants are gen-
erated including OH radicals which are among the strongest oxidants and can
oxidize many stable organic compounds (Veljković et al. 2014). The main products
generated in this process are glyceraldehyde, glycolaldehyde, and formic, lactic,
acetic and glycolic acids (Veljković et al. 2014).
Advanced oxidation processes are a set of physicochemical treatment procedures
to remove impurities from wastewater by oxidation through reactions with highly
reactive oxidizing species, mainly hydroxyl radicals, able to degrade organic
compounds (Veljković et al. 2014). The subsets of these processes are ozonation
and ozone related processes (O3/H2O2, UV/O3), heterogeneous (TiO2/UV) and
homogeneous (photo-Fenton process) photo-catalysis, electro-oxidation, etc.
(Veljković et al. 2014). The mechanism of hydroxyl radicals production depends on
the type of advanced oxidation process used. In the ozonation process, ozone is
generated onsite, by imposing a high voltage alternating current across a dielectric
discharge gap through which the air or oxygen passes. Once dissolved into water,
ozone undergoes a reaction with hydroxyl ions leading to the formation of
hydrogen peroxide in charged form. Then, a second ozone molecule reacts with this
species to produce the ozonide radical. Finally, the ozonide radical produces
hydroxyl radical upon protonation (Veljković et al. 2014). In the photocatalytic
oxidation with TiO2, the irradiation of a photocatalytic surface leads to the pro-
duction of an exited electron and an electron gap. The highly reactive electron gap
will react with the adsorbed water molecule onto the catalyst surface and hence,
hydroxyl radicals will be produced and redox processes take place at the surface of
titanium dioxide (Veljković et al. 2014). In the photo-Fenton process, Fe2+ is
oxidized to Fe3+ by decomposing hydrogen peroxide in an acidic aqueous solution
to form hydroxyl radicals. Then, the Fe3+ ions are reduced back to Fe2+ ions by UV
light, producing an additional hydroxyl radical. These hydroxyl radicals oxidize a
wide range of organic and inorganic compounds. This process is one of the most
efficient advanced oxidation processes when a high reduction of COD is required
(Veljković et al. 2014).
5 Applications of Biodiesel By-products 119

Microbial degradation is another method for biodiesel wash-water treatment


because of the high COD level of this wastewater. Although the biodegradable
organic compounds contained in wastewater serve as nutrients for microorganisms
but as mentioned earlier, the biological treatment of biodiesel wash-water is difficult
because its composition is not suitable for microbial growth (Veljković et al. 2014).
More specifically, this wastewater has a high pH value and is poor in nutrients
needed for microbial growth, except for carbon source (residual oil, methanol, and
glycerol), and hence, minimum amounts of important nutrients should be supple-
mented (Veljković et al. 2014). Nevertheless, it should also be noted that high
methanol concentrations are toxic to methanogens, while too high residual oil
contents can inhibit the microbial growth. Moreover, free fatty acids could also
inhibit the anaerobic digestion process while high salinity of the biodiesel
wastewater can negatively influence some microorganisms, such as methanogens.
Therefore, a biological treatment of biodiesel wastewater should be undertaken
under optimum operating conditions, including biodegradation of the residual oil
(Veljković et al. 2014). If performed successfully, anaerobic digestion of biodiesel
wash-water could not only lead to the generation of treated water for reuse in the
biodiesel production process but also could generate biogas which could be sub-
sequently used for heat and power generation.
Another treatment technique investigated for biodiesel wash-water is microbial
fuel cells (Sukkasem et al. 2011). Microbial fuel cells are performed under
anaerobic conditions and are aimed at harvesting energy. This treatment offers high
COD removal efficiencies but it is costly due to the expensive materials used such
as platinum or gold metal catalysts, proton exchange membranes, mediators, and
graphite electrodes (Sukkasem et al. 2011).

5.5.2 Biodiesel Wash-Water as Feedstock for Production


of Value-Added Products

Attempts have also been made to utilize biodiesel wash-water as a source of carbon
to produce value-added products such as biogas, biohydrogen, and bioethanol, or as
a carbon source for bioreactors treating acid mine drainage (Jaruwat et al. 2010;
Phukingngam et al. 2011). Biomethane has also been produced by the anaerobic
co-digestion of a mixture of glycerol and the biodiesel wash-water. In fact, biodiesel
wash-water was used as a replacement for the clean water needed to dilute biodiesel
glycerol and to meet some nutrient requirements (Siles et al. 2010).
In a study, concurrent polyhydroxy alkanoates (PHA) production and COD
removal in shaken flasks was studied utilizing biodiesel wash-waters and a PHA
producing microbial seed with and without residual ethanol (Coats et al. 2007). In a
different attempt, the biodiesel wash-water having a high COD (300–500 g/L) and
methanol (6–10% by vol.) obtained from a commercial plant based on
base-catalyzed transesterification was used as a carbon substrate for denitrification
120 H. Rastegari et al.

of the nitrogen-rich sludge liquor from an aerobic sludge digestion in a sequential


batch reactor. The concentration of nitrogen (NH4) in the effluent was substantially
reduced throughout the process to less than 10 mg/L (Malá and Malý 2010).
The feasibility of using biodiesel wash-water to which some nutrients were
added for growing alga Chlorella protothecoides in photo-bioreactors was reported
(Lamers 2009). The alga showed an excellent ability to remediate biodiesel
wastewater by converting glycerol, methanol, and catalyst residues into the biomass
or other organics. After algae harvesting, the residual water contained small
amounts of unharvested biomass and unused substrates, which were successfully
removed by the diatomaceous earth filtration. The advantage of this process was
that the algal lipids were viable for biodiesel production while the remediated water
could be reused (Lamers 2009). Having a different approach for valorization of
biodiesel wash-water, an up flow bio-filter circuit system has also been developed
for electricity generation from this wastewater, without any chemical pretreatments
or nutrient addition (Sukkasem et al. 2011; Sukkasem and Laehlah 2013). Finally, it
should be mentioned that if biodiesel wash-water is methanol-free, it can also be
used for farm irrigation without any purifications (Sukkasem et al. 2011).

References

Akunna JC, Bizeau C, Moletta R (1993) Nitrate and nitrite reductions with anaerobic sludge using
various carbon sources: Glucose, glycerol, acetic acid, lactic acid and methanol. Water Res
27:1303–1312
Alshelmani MI, Loh TC, Foo HL, Sazili AQ, Lau WH (2016) Effect of feeding different levels of
palm kernel cake fermented by Paenibacillus polymyxa ATCC 842 on nutrient digestibility,
intestinal morphology, and gut microflora in broiler chickens. Animal Feed Sci Technol
216:216–224
Anitha M, Kamarudin SK, Kofli NT (2016) The potential of glycerol as a value-added commodity.
Chem Eng J 295:119–130
Antolin G, Tinaut FV, Briceno Y, Castano V, Perez C, Ramirez AI (2002) Optimization of
biodiesel production by sunflower oil transesterification. Bioresource Technol 83:111–114
Atadashi IM, Aroua MK, Abdul Aziz AR, Sulaiman NMN (2011) Refining technologies for the
purification of crude biodiesel. Appl Energy 88:4239–4251
Avasthi KS, Reddy RN, Patel S (2013) Challenges in the production of hydrogen from glycerol a
biodiesel byproduct via steam reforming process. Procedia Eng 51:423–429
Bagheri S, Julkapli NM, Yehye WA (2015) Catalytic conversion of biodiesel derived raw glycerol
to value added products. Renew Sust Energy Rev 41:113–127
Bano Z, Shashirekha MN, Rajarathnam S (1993) Improvement of the bioconversion and
biotransformation efficiencies of the oyster mushroom (Pleurotus sajor-caju) by supplemen-
tation of its rice straw substrate with oil seed cakes. Enzyme Microb Technol 15:985–989
Barik D, Murugan S (2015) Assessment of sustainable biogas production from de-oiled seed cake
of karanja-an organic industrial waste from biodiesel industries. Fuel 148:25–31
Beatrice C, Di Blasio G, Guido C, Cannilla C, Bonura G, Frusteri F (2014) Mixture of glycerol
ethers as diesel bio-derivable oxy-fuel: Impact on combustion and emissions of an automotive
engine combustion system. Appl Energy 132:236–247
Bernat K, Kulikowska D, Uchniewski K (2015) Glycerine as a carbon source in nitrite removal
and sludge production. Chem Eng J 267:324–331
5 Applications of Biodiesel By-products 121

Berrios M, Skelton RL (2008) Comparison of purification methods for biodiesel. Chem Eng J
144:459–465
Bodík I, Blštáková A, Sedlácek S, Hutnan M (2009) Biodiesel waste as source of organic carbon
for municipal WWTP denitrification. Bioresource Technol 100:452–2456
Cao P, Dube MA, Tremblay AY (2008) Methanol recycling in the production of biodiesel in a
membrane reactor. Fuel 87:825–833
Chavalparit O, Ongwandee M (2009) Optimizing electrocoagulation process for the treatment of
biodiesel wastewater using response surface methodology. J Environ Sci 21:1491–1496
Chen Y, Wang T, Shen N, Zhang F, Zeng RJ (2016) High-purity propionate production from
glycerol in mixed culture fermentation. Bioresource Technol 219:659–667
Climent MJ, Corma A, Iborra S (2014) Conversion of biomass platform molecules into fuel
additives and liquid hydrocarbon fuels. Green Chem 16:516–547
Coats ER, Loge FJ, Smith WA, Thompson DN, Wolcott MP (2007) Functional stability of a mixed
microbial consortium producing PHA from waste carbon sources. Appl Biochem Biotechnol
136–140:909–925
Costa DD, Gomes AA, Fernandes M, da Costa Bortoluzzi RL, Magalhães MDLB, Skoronski E
(2018) Using natural biomass microorganisms for drinking water denitrification. J Environ
Manage 217:520–530
Cyplik P, Juzwa W, Marecik R, Powierska-Czarny J, Piotrowska-Cyplik A, Czarny J,
Drozdzynska A, Chrzanowski L (2013) Denitrification of industrial wastewater: influence of
glycerol addition on metabolic activity and community shifts in a microbial consortium.
Chemosphere 93:2823–2831
Dabrock B, Bahl H, Gottschalk BY, Lee WJ, Chien LJ, Chang JS (2012) High yield biobutanol
production by solvent-producing bacterial microflora. Bioresource Technol 113:58–64
Daud NM, Abdullah SRS, Hasan HA, Yaakob Z (2015) Production of biodiesel and its wastewater
treatment technologies: a review. Process Saf Environ Protect 94:487–508
Di Serio MG, Lanza B, Mucciarella MR, Russi F, Iannucci E, Marfisi P, Madeo A (2008) Effects
of olive mill wastewater spreading on the physico-chemical and microbiological characteristics
of soil. Int Biodeterior Biodegrad 62:403–407
El Diwani G, Attia NK, Hawash SI (2009) Development and evaluation of biodiesel fuel and
by-products from jatropha oil. Int J Environ Sci Technol 6:219–224
Encinar JM, Gonzalez JF, Rodriguez-Reinars A (2007) Ethanolysis of used frying oil: biodiesel
preparation and characterization. Fuel Process Technol 88:513–522
Farzana K, Shah SN, Butt FB, Awan SB (2005) Biosynthesis of bacitracin in solid state
fermentation by Bacillus licheniformis using defatted oil seed cakes as substrate. Pak J Pharm
Sci 18:55–57
Freitas TB, Felix TL, Pedreira MS, Silva RR, Silva FF, Silva HGO, Moreira BS (2017) Effects of
increasing palm kernel cake inclusion in supplements fed to grazing lambs on growth
performance, carcass characteristics, and fatty acid profile. Anim Feed Sci Technol 226:71–80
Gallezot P (1997) Selective oxidation with air on metal catalysts. Catal Today 37:40–418
Garcia E, Laca M, Perez E, Garrido A, Peinado J (2008) New class of acetal derived from glycerin
as a biodiesel fuel component. Energy Fuel 22:4274–4280
Gholami Z, Abdullah AZ, Lee KT (2014) Dealing with the surplus of glycerol production from
biodiesel industry through catalytic upgrading to polyglycerols and other value added products.
Renew Sust Energy Rev 39:327–341
Gorji YM, Ghaziaskar HS (2016) Optimization of solketalacetin synthesis as a green fuel additive
from ketalization of monoacetin with acetone. Ind Eng Chem Res 55:6904–6910
Guerrero J, Tayà C, Guisasola A, Baeza JA (2011) Glycerol as a sole carbon source for enhanced
biological phosphorus removal. Water Res 46:2983–2991
Gupta RC, Vaish SS, Singh RK, Singh NK, Singh KP (2005) Oil cakes as media for growing
Catenaria anguillulae Sorokin, a facultative endoparasite of nematodes. World J Microbiol
Biotechnol 21:1181–1185
122 H. Rastegari et al.

Hamdi H, Majdoub-Mathlouthi L, Picard B, Listrat A, Durand D, Znaïdi IA, Kraiem K (2016)


Carcass traits, contractile muscle properties and meat quality of grazing and feedlot Barbarine
lamb receiving or not olive cake. Small Rumin Res 145:85–93
Huber GW, Iborra S, Corma A (2006) Synthesis of transportation fuels from biomass: chemistry,
catalysts, and engineering. Chem Rev 106:4044–4098
Ilie A, Simoes M, Baranton S, Coutanceau C, Martemianov S (2011) Influence of operational
parameters and of catalytic materials on electrical performance of direct glycerol solid alkaline
membrane fuel cells. J Power Sources 196:4965–4971
Jabłoński SJ, Kułażyński M, Sikora I, Łukaszewicz M (2016) The influence of different
pretreatment methods on biogas production from Jatropha curcas oil cake. J Environ Manage
Available from: https://doi.org/10.1016/j.jenvman.2016.06.001
Jain MK, Datta R, Zeikus JG (1989) High value organic acids fermentation-emerging processes
and products. Bioprocess Eng the first generation 36:6–398
Jaruwat P, Kongjao S, Hunsom M (2010) Management of biodiesel wastewater by the combined
processes of chemical recovery and electrochemical treatment. Energy Convers Manag
51:531–537
Johnson DT, Taconi KA (2007) The glycerin glut: Options for the value added conversion of
crude glycerol resulting from biodiesel production. Environ Prog 26:338–348
Kanchanasuta S, Pisutpaisal N (2016) Waste utilization of palm oil decanter cake on biogas
fermentation. Int J Hydrogen Energy 41:15661–15666
Kashyap P, Sabu A, Pandey A, Szakacs G, Soccol CR (2002) Extracellular L-glutaminase
production by Zygosaccharomyces rouxii under solid state fermentation. Process Biochem
38:307–312
Katryniok B, Paul S, Bellière-Baca V, Rey P, Dumeignil F (2010) Glycerol dehydration to acrolein
in the context of new uses of glycerol. Green Chem 12:2079–2098
Khayoon MS, Hameed BH (2013) Yttrium-grafted mesostructured SBA-3 catalyst for the
transesterification of glycerol with methyl acetate to synthesize fuel oxygenates. Appl Catal A:
Gen 460:61–69
Kolesarova N, Hutnan M, Bodik I, Spalkova V (2011) Utilization of biodiesel by-products for
biogas production. J Biomed Biotechnol ID No. 126798
Lamers A (2009) Algae oil from small scale low input water remediation site as feedstock for
biodiesel conversion. Guelph Eng J 2:24–38
Len C, Luque R (2014) Continuous flow transformations of glycerol to valuable products: an
overview. Sust Chem Process 2:1–10
Leung DYC, Wu X, Leung MKH (2010) A review on biodiesel production using catalyzed
transesterification. Appl Energy 87:1083–1095
Lopez JAS, Santos MAM, Perez AFC, Martin AM (2009) Anaerobic digestion of glycerol derived
from biodiesel manufacturing. Bioresource Technol 100:5609–5615
Low SC, Gan GK, Cheong KT (2011) Separation of methyl ester from water in a wet
neutralization process. J Sustain Energy Environ 2:15–19
Lucchetti R, Onotri L, Clarizia L, Di Natale F, Di Somma I, Andreozzi R, Marotta R (2017)
Removal of nitrate and simultaneous hydrogen generation through photocatalytic reforming of
glycerol over in situ prepared zero-valent nano copper/P25. Appl Catal B: Environ 202:539–549
Ma F, Hanna MA (1999) Biodiesel production: a review. Bioresource Technol 70:1–15
Mahanta N, Gupta A, Khare SK (2008) Production of protease and lipase by solvent tolerant
Pseudomonas aeruginosa PseA in solid state fermentation using Jatropha curcas seed cake as
substrate. Bioresource Technol 99:1729–1735
Malá J, Malý J (2010) Waste water from biodiesel production as a carbon source for denitrification
of sludge liquor in SBR. Chem Biochem Eng 24:211–217
May A, Salvadó J, Torras C, Montané D (2010) Catalytic gasification of glycerol in supercritical
water. Chem Eng J 160:751–759
Mazzoncini M, Antichi D, Tavarini S, Silvestri N, Lazzeri L, D’Avino L (2015) Effect of defatted
oilseed meals applied as organic fertilizers on vegetable crop production and environmental
impact. Ind Crop Product 75:54–64
5 Applications of Biodiesel By-products 123

Melero JA, Vicente G, Morales G, Paniagua M, Bustamante J (2010) Oxygenated compounds


derived from glycerol for biodiesel formulation: influence on EN 14214 quality parameters.
Fuel 89:2011–2018
Molina-Alcaide E, Yáñez-Ruiz DR (2008) Potential use of olive by-products in ruminant feeding:
a review. Anim Feed Sci Technol 147:247–264
Moresi M, Parente E (2014) Fermentation (industrial), production of some organic acids (citric,
gluconic, lactic and propionic). In: Reference module in food science, from encyclopedia of
food microbiology, 2nd edn, pp 804–815
Mota CJ, da Silva CX, Rosenbach N Jr, Costa J, da Silva F (2010) Glycerin derivatives as fuel
additives: the addition of glycerol/acetone ketal (solketal) in gasolines. Energy Fuel 24:2733–
2736
Mota CJA (2012) Valorization of glycerol by-product of biodiesel production. In: Advances in
biodiesel production. Woodhead Publishing, Cambridge
Murti CR (1965) Oil cake meal for preparation of protein hydrolysate. Biotechnol Bioeng 7:285–
293
Nanda MR, Yuan Z, Qin W, Ghaziaskar HS, Poirier MA, Xu CC (2014a) A new continuous-flow
process for catalytic conversion of glycerol to oxygenated fuel additive: catalyst screening.
Appl Energy 123:75–81
Nanda MR, Yuan Z, Qin W, Ghaziaskar HS, Poirier MA, Xu CC (2014b) Catalytic conversion of
glycerol to oxygenated fuel additive in a continuous flow reactor: process optimization. Fuel
128:113–119
Nanda MR, Yuan Z, Qin W, Ghaziaskar HS, Poirier MA, Xu CC (2014c) Thermodynamic and
kinetic studies of a catalytic process to convert glycerol into solketal as an oxygenated fuel
additive. Fuel 117:470–477
Nanda MR, Zhang Y, Yuan Z, Qin W, Ghaziaskar HS, Xu CC (2016) Catalytic conversion of
glycerol for sustainable production of solketal as a fuel additive: a review. Renew Sust Energy
Rev 56:1022–1031
Ngamlerdpokin K, Kumjadpai S, Chatanon P, Tungmanee U, Chuenchuanchom S, Jaruwat P,
Lertsathitphongs P, Hunsom M (2011) Remediation of biodiesel wastewater by chemical- and
electrocoagulation: a comparative study. J Environ Manag 92:2454–2460
Oh BR, Seo JW, Heo SY, Hong WK, Luo LH, Joe MH, Park DH, Kim CH (2011) Efficient
production of ethanol from crude glycerol by a Klebsiella pneumoniae mutant strain.
Bioresource Technol 102:3918–3922
Oliveira F, Souza CE, Peclat VR, Salgado JM, Ribeiro BD, Coelho MA, Belo I (2017)
Optimization of lipase production by Aspergillus ibericus from oil cakes and its application in
esterification reactions. Food Bioprod Process 102:268–277
Pagliaro M, Rossi M (2010) The Future of glycerol. The Royal Society of Chemistry Publishing,
Cambridge
Pant M, Sharma S, Dubey S, Naik SN, Patanjali PK (2016) Utilization of biodiesel by-products for
mosquito control. J Biosci Bioeng 121:299–302
Phukingngam D, Chavalparit O, Somchai D, Ongwandee M (2011) Anaerobic baffled reactor
treatment of biodiesel-processing waste water with high strength of methanol and glycerol:
reactor performance and biogas production. Chem Pap 65:644–651
Raita M, Ibenegbu C, Champreda V, Leak DJ (2016) Production of ethanol by thermophilic
oligosaccharide utilising Geobacillus thermoglucosidasius TM242 using palm kernel cake as a
renewable feedstock. Biomass Bioenerg 95:45–54
Ramachandran S, Patel AK, Nampoothiri KM, Francis F, Nagy V, Szakacs G, Pandey A (2004)
Coconut oil cake a potential raw material for the production of a-amylase. Bioresource Technol
93:169–174
Ramachandran S, Singh SK, Larroche C, Soccol CR, Pandey A (2007) Oil cakes and their
biotechnological applications a review. Bioresource Technol 98:2000–2009
Rastegari H, Ghaziaskar HS (2015) From glycerol as the by-product of biodiesel production to
value-added monoacetin by continuous and selective esterification in acetic acid. J Ind Eng
Chem 21:856–861
124 H. Rastegari et al.

Rastegari H, Ghaziaskar HS, Yalpani M (2015) Valorization of biodiesel derived glycerol to


acetins by continuous esterification in acetic acid: focusing on high selectivity to diacetin and
triacetin with no byproducts. Ind Eng Chem Res 54:3279–3284
Remón J, Ruiz J, Oliva M, García L, Arauzo J (2016) Effect of biodiesel-derived impurities (acetic
acid, methanol and potassium hydroxide) on the aqueous phase reforming of glycerol. Chem
Eng J 299:431–448
Rezayat M, Ghaziaskar HS (2009) Continuous synthesis of glycerol acetates in supercritical
carbon dioxide using Amberlyst 15. Green Chem 11:710–715
Rossi DM, da Costa JB, de Souza EA, Peralba MCR, Ayub MAZ (2012) Bioconversion of
residual glycerol from biodiesel synthesis into 1,3-propanediol and ethanol by isolated bacteria
from environmental consortia. Renew Energy 39:223–227
Roy D, Subramaniam B, Chaudhari RV (2010) Aqueous phase hydrogenolysis of glycerol to
1,2-propanediol without external hydrogen addition. Catal Today 156:31–37
Sabu A, Sarita S, Pandey A, Bogar B, Szakacs G, Soccol CR (2002) Solid state fermentation for
production of phytase by Rhizopus oligosporus. Appl Biochem Biotechnol 102:251–260
Sabu A, Pandey A, Daud MJ, Szakacs G (2005) Tamarind seed powder and palm kernel cake: two
novel agro residues for the production of tannase under solid state fermentation by Aspergillus
Niger ATCC 16620. Bioresource Technol 96:1223–1228
Sarker AK, Saha D, Begum H, Zaman A, Rahman MM (2015) Comparison of cake compositions,
pepsin digestibility and amino acids concentration of proteins isolated from black mustard and
yellow mustard cakes. AMB Express 5:22–27
Schwengber CA, Alves HJ, Schaffner RA, da Silva FA, Sequinel R, Bach VR, Ferracin RJ (2016)
Overview of glycerol reforming for hydrogen production. Renew Sust Energy Rev 58:259–266
Shafiei A, Rastegari H, Ghaziaskar HS, Yalpani M (2017) Glycerol transesterification with ethyl
acetate to synthesize acetins using ethyl acetate as reactant and entrainer. Biofuel Res J 4:565–
570
Shashirekha MN, Rajarathnam S, Bano Z (2002) Enhancement of bioconversion efficiency and
chemistry of the mushroom, Pleurotus sajor-caju (Berk and Br.) Sacc. produced on spent rice
straw substrate, supplemented with oil seed cakes. Food Chem 76:27–31
Shirani M, Ghaziaskar HS, Xu CC (2014) Optimization of glycerol ketalization to produce solketal
as biodiesel additive in a continuous reactor with subcritical acetone using Purolite PD206 as
catalyst. Fuel Process Technol 124:206–211
Shukor H, Abdeshahian P, Al-Shorgani NKN, Hamid AA, Rahman NA, Kalil MS (2016)
Saccharification of polysaccharide content of palm kernel cake using enzymatic catalysis for
production of biobutanol in acetone butanol ethanol fermentation. Bioresource Technol
202:206–213
Siler-Marinkovic S, Tomasevic A (1998) Transesterification of sunflower oil in situ. Fuel
77:1389–1391
Siles JA, Martín MA, Chica AF, Martín A (2010) Anaerobic co-digestion of glycerol and waste
water derived from biodiesel manufacturing. Bioresour Technol 101:6315–6321
Silvestre G, Fernández B, Bonmatí A (2015) Addition of crude glycerine as strategy to balance the
C/N ratio on sewage sludge thermophilic and mesophilic anaerobic co-digestion. Bioresource
Technol 193:377–385
Srirangsan A, Ongwandee M, Chavalparit O (2009) Treatment of biodiesel wastewater by
electrocoagulation process. Environ Asia 2:15–19
Sukkasem C, Laehlah S, Hniman A, Sompong O, Boonsawang P (2011) Up flow bio-filter circuit
(UBFC): Biocatalyst microbial fuel cell (MFC) configuration and application to biodiesel
wastewater treatment. Bioresour Technol 102:10363–10370
Sukkasem C, Laehlah S (2013) Development of a UBFC biocatalyst fuel cell to generate power
and treat industrial waste waters. Bioresour Technol 146:749–753
Tomasevic AV, Siler-Marinkovic SS (2003) Methanolysis of used frying oil. Fuel Process Technol
81:1–6
Veljković VB, Stamenković OS, Tasić MB (2014) The wastewater treatment in the biodiesel
production with alkali-catalyzed transesterification. Renew Sust Energy Rev 32:40–60
5 Applications of Biodiesel By-products 125

Vicente G, Martinez M, Aracil J (2004) Integrated biodiesel production: a comparison of different


homogenous catalysts systems. Bioresource Technol 92:297–305
Vidyarthi AS, Tyagi RD, Valero JR, Surampalli RY (2002) Studies on the production of B.
thuringiensis based biopesticides using wastewater sludge as a raw material. Water Res
36:4850–4860
Villa A, Dimitratos N, Chan-Thaw CE, Hammond C, Prati L, Hutchings GJ (2015) Glycerol
oxidation using gold-containing catalysts. Account Chem Res 48:1403–1412
Vollenweider S, Lacroix C (2004) 3-Hydroxypropionaldehyde: applications and perspectives of
biotechnological production. Appl Microb Biotechnol 64:16–27
Wang Y, Zhou J, Guo X (2015) Catalytic hydrogenolysis of glycerol to propanediols: a review.
RSC Adv 5:74611–74628
Xiao Y, Xiao G, Varma A (2013) A universal procedure for crude glycerol purification from
different feedstocks in biodiesel production: experimental and simulation study. Ind Eng Chem
Res 52:14291–14296
Yang G, Wang D, Yang Q, Zhao J, Liu Y, Wang Q, Zeng G, Li ZX, Li H (2018) Effect of acetate
to glycerol ratio on enhanced biological phosphorus removal. Chemosphere 196:78–86
Yorgun S, Sensoz S, Kockar OM (2001) Flash pyrolysis of sunflower oil cake for production of
liquid fuels. J Anal Appl Pyrolysis 60:1–12
Zhang MH, Zhao QL, Bai X, Ye ZF (2010) Adsorption of organic pollutants from coking
wastewater by activated coke. Physico Chem Eng Asp 62:140–146
Zheng Y, Chen X, Shen Y (2008) Commodity chemicals derived from glycerol, an important
biorefinery feedstock. Chem Rev 108:5253–5277
Zhou X, Beltranena E, Zijlstra RT (2016) Effects of feeding canola press-cake on diet nutrient
digestibility and growth performance of weaned pigs. Anim Feed Scie Technol 211:208–215
Chapter 6
Economic Risk Analysis and Critical
Comparison of Biodiesel Production
Systems

Seyed Soheil Mansouri, Carina L. Gargalo, Isuru A. Udugama,


Pedram Ramin, Mauricio Sales-Cruz, Gürkan Sin and
Krist V. Gernaey

Abstract In this chapter, the importance of risk assessment in biodiesel-based


economy is first discussed. The importance of risk analysis to identify the most
promising production schemes is also discussed from an economic point of view.
Next, a systematic framework for economic risk assessment of biodiesel production
processes and its associated by-products is presented. The application of the
framework is highlighted through the production of 1,2-propanediol and
1,3-propanediol as value-added products from glycerol, which are critically
assessed in terms of its techno-economic performance through the estimation of
economic indicators, net present value (NPV), and minimum selling price (MSP).
The Monte Carlo method with Latin Hypercube Sampling (LHS) is used to prop-

S. S. Mansouri  C. L. Gargalo  I. A. Udugama  P. Ramin  G. Sin 


KristV. Gernaey (&)
Department of Chemical and Biochemical Engineering, Process and Systems Engineering
Center (PROSYS), Technical University of Denmark, Building 229,
2800 Kongens Lyngby, Denmark
e-mail: kvg@kt.dtu.dk
S. S. Mansouri
e-mail: seso@kt.dtu.dk
C. L. Gargalo
e-mail: carlour@kt.dtu.dk
I. A. Udugama
e-mail: isud@kt.dtu.dk
P. Ramin
e-mail: pear@kt.dtu.dk
G. Sin
e-mail: gsi@kt.dtu.dk
M. Sales-Cruz
Departamento de Procesos y Tecnología, Universidad Autónoma Metropolitana –
Cuajimalpa, Colonia Santa Fe Cuajimalpa, Delegación Cuajimalpa de Morelos,
05300 Mexico City, Mexico
e-mail: asales@correo.cua.uam.mx

© Springer Nature Switzerland AG 2019 127


M. Tabatabaei and M. Aghbashlo (eds.), Biodiesel,
Biofuel and Biorefinery Technologies 8,
https://doi.org/10.1007/978-3-030-00985-4_6
128 S. S. Mansouri et al.

agate the market price and technical uncertainties to the economic indicator cal-
culations and to quantify the respective economic risk. In order to decrease the
economic risk, the integrated production of the product as a module added to the
biodiesel plant was tested as an alternative scenario. Using the integrated concept of
utilizing the waste glycerol stream in biodiesel plants contributes to the diversifi-
cation of the product portfolio for vegetable oil-based biorefineries, and in turn
improves cost-competitiveness and robustness against market price fluctuations.
The developed generic framework can be applied to other biodiesel by-products to
assess the potentials of obtaining value-added products from them. Finally, future
perspectives and other approaches toward economic production of biodiesel with
lower risks are highlighted. The framework proposed in this work is to provide
some detailed perspectives to facilitate the economic risk analysis of biodiesel
production for any given technology.

6.1 Introduction

Over the past few decades, there has been a growing concern about the depletion of
fossil fuels and the need for renewable sources of energy. Furthermore, issues with
climate change and its societal impacts have become major drivers to seek solutions
to overcome with them. Therefore, it has become a priority for many policy and
decision-makers. However, there remains a challenge for chemical and biochemical
industries to search for more sustainable products and processes. To this end, one
alternative approach to overcome these challenges is moving toward bio-based
economy and bio-based production processes/products. This approach may lead to
sustainable long-term development such that bio-based chemicals and fuels (e.g.,
biodiesel) may play a key role in order to replace the large dependence on fossil
sources.
Sustainability has various attributes at the three bottom lines being societal,
economic, and environmental (ecological) aspects. Thus, it is a multidimensional
assessment metric by nature. Therefore, the decision-making becomes a
multi-criteria and multi-objective process when it comes to biorefinery concepts. As
a result, developing these concepts toward implementation is complex and chal-
lenging. The complexity does not only come from the multi-criteria evaluation
techniques required to develop the biorefinery concepts, it is also associated with a
large set of data required as input parameters. These data are usually obtained from
various sources with different degrees of uncertainty associated with them
(Mansouri et al. 2013; Gargalo and Sin 2015). It is worthy to clarify that from
ecological economics standpoint, natural capital cannot be substituted by
human-made capital, and the economic aspects presented here should be considered
as an additional incentive for feedstock suppliers and producers to promote
fossil-free energy production. It is no doubt that production of biodiesel production
would increase the capacity of fuel production, contribute to the diversity of
resources and increase energy security in short and long term.
6 Economic Risk Analysis and Critical Comparison … 129

An array of assumptions and simplifications are made to manage the complexity


associated with biorefinery design problems at the early stage of process devel-
opment. This is because usually, at early stages of process development, actual
information are either not available or they are reported partially. Taking into
account the alternative processing technologies, feedstocks, and products, a large
set of potentially feasible processing routes become available. Therefore, evaluating
and ranking these alternatives using economic and sustainability metrics is required
to identify the most sustainable process alternative. Hence, uncertainty associated
with techno-economic parameters/criteria during screening the biorefinery process
systems has to be appropriately handled (Turton 2009).
Many decision-makers are usually baffled when they have to make decisions
under uncertain conditions. This concern is not only associated with returning
economic value but also with the extent of risk that each decision carries. Quantified
based on the uncertainty for which the probability distribution is known (or pro-
jected), risk is equal to the sum of probabilities of outcome(s) (likelihood of
occurrence) times the projected loss as a consequence of the outcome(s) (Gargalo
et al. 2016). Thus, risk-based decision-making is advantageous by providing
organized and easy-to-use information regarding the potential unwanted outcomes
and their economic loss. In this way, managers can make more informed and
realistic choices regarding project feasibility. Recently, a comprehensive review has
been published by Živković et al. (2017) where they provide an overview of the
technological, economic, environmental, social, toxicological, and human health
risk considerations of biodiesel production and use.
There exist several examples of methodologies to compare biorefinery process
systems on predefined criteria/indicators. For example, Azapagic et al. (2006)
proposed a methodology toward identification of more sustainable process systems
which was dominated by life cycle school of thought. Their methodology guides
the process designer through different design stages to enable integration of tech-
nical, economic, environmental, and social criteria. They demonstrated that appli-
cation of this methodology brings the understanding on how sustainability
considerations could be integrated into process design from project initiation
through preliminary to detailed design. Sacramento-Rivero (2012) has proposed a
methodology applicable to biorefineries where a sustainability scale is used based
on an absolute reference and normalized for sustainability indicators applicable to
biorefineries. The resulting framework consists of 5 indicator categories, and a total
of 14 metrics; this tool can be used to evaluate the sustainability of biorefinery
designs, and also for performance evaluations. Furthermore, authors in (You et al.
2012; Gong and You 2015; Yue et al. 2013) present a literature review on pro-
gramming techniques explored to identify the optimal alternative through single- or
multi-objective optimization.
In this chapter, an overview of the importance of economics in industrial bio-
diesel production is first provided. This is followed by describing a generic
techno-economic framework to perform risk analysis on biorefinery system. In
order to demonstrate the applicability of the developed framework, a case study
130 S. S. Mansouri et al.

related to biodiesel production is presented. This is then followed by drawing a set


of conclusions based on the discussions provided throughout the chapter.

6.2 Importance of Economics in Industrial Biodiesel


Production

The promise of producing a renewable alternative to the use of fossil fuels has
allowed biodiesel to attract significant funding from both industry and government
institutes. Biodiesel represents both an environmentally and societally responsible
alternative to the more establish crude oil-based production of diesel and other
related production. As described in the previous sections, the technology used in
biodiesel production is also well established. However, with all these technical
capabilities as well as societal and environmental pressure, industrial-scale bio-
diesel production has not taken off and only a fraction of the world’s diesel pro-
duction is derived from renewable routes. The main reason for this lack of
adaptation is the lack of lucrative economic drivers as well as the economic
uncertainties that are related with many aspects of biodiesel productions especially
the market variability in both product and feedstock. Economic assessment could be
beneficial to discover and promote strategies to increase the competitiveness of
biodiesel productions and also increase their commercial availability. This analysis
can serve as a driving force to promote development of new processes technologies
and can be also used to predict the production cost of a plant.
To analyze the economic potential of the biodiesel production, different aspects
should be considered. First, the cost analysis of the feedstock offered by the bio-
mass suppliers and the willingness of the buyers for a profitable bioenergy pro-
duction should be taken into account. The feedstock cost is highly dependent on the
crops used, e.g., sugar beet or sunflower, and chosen harvesting technologies for
where the crop is grown. This cost is also influenced by the input materials used,
e.g., fertilizers or herbicide. It is possible to reduce feedstock price using cheaper
but effective raw materials, e.g., vegetable oil. Second, the conversion cost
including biodiesel conversion processes should be considered. Different produc-
tion technologies would result in different production costs. This cost is also
impacted by the production scale of the plant. Third, the potential of competing
markets created by the price effects after increased production of biofuel should be
considered. This often happens due to limited availability of resources for agri-
culture, forestry, and green energy production. Hence, biofuel production should be
established effectively and also be supported to stay competitive in the energy
market. Moreover, the availability of resources and production of commodities in
market should be assessed with relation to the location of production and location of
demand. For instance, the effect of production of biofuel on balance of trades in has
been assessed in details for United States (National Research Council 2011).
6 Economic Risk Analysis and Critical Comparison … 131

One way to make biofuel competitive and attractive from economic standpoint is
not only focussed on the primary biofuel products, e.g., bioethanol and biodiesel, but
also the production of coproducts. The coproduct production could be enhancement
of already existing coproducts or introducing new coproducts. This could be certain
livestock and poultry feeds for animal feed industry (Makkar 2012) or production of
glycerin. There are different measures defined which can be used to perform a cost–
benefit analysis. Agricultural yield (Ton/ha), defined as annular production of crops
per area, is a good reference value which lumps regional factors such as conditions of
the soil, climate conditions, production efficiency, and production development. The
yield values can be easily accessible from Eurostat (2017) and have been used to
perform yield-sensitive cost assessments for the member states in Europe (Toro
Chacón 2004). To determine the degree of competitiveness, the biodiesel production
costs should be compared with fossil fuel counterparts. This comparison is recom-
mended to include sufficient period of time and exclude taxes. Short-term price
fluctuations may influence a realistic assessment. Nevertheless, it might be chal-
lenging to make an assessment for the favor of biodiesel production due to lower
energy content of biodiesel compared to fossil fuels. However, inclusion of the value
of subproducts such as glycerin may offset some of the cost.
Furthermore, the economic aspects should also take into account the nontech-
nical limitations that are imposed by local government on the agricultural com-
modities for biofuel production. Lack of subsidies or tax incentives would
negatively impact the biodiesel production. Compared with other manufacturing
industry, less systematic approaches are followed for biodiesel production industry
to identify and address possible process hazards at different stages of the produc-
tion. Lack of comprehensive guidelines and also reported incidents may discourage
investors to provide funding for biodiesel production. Hazard identification process
is especially less known for small- and medium-sized scale biodiesel productions.
The hazard could be related to materials (e.g., Ethanol, methanol, and sulphuric
acid), operation and handling (e.g., storage of toxic materials), or design and
installation (e.g., requirements for electrical installations). Evaluations of such
factors that could potentially discourage the investors to support biodiesel pro-
duction, especially for smaller production scales, may need to be accounted for.

6.3 Systematic Framework for Economic Risk Assessment

The economic risk of developing a biodiesel project can be challenging. Especially


considering the volatile market environment for many of the biodiesel products that
are correlated to the price of crude oil, these biodiesel-derived products need to
compete with their petrochemical-derived counterparts. The economic complexity is
further compounded by the fact that many biodiesel feedstocks can be used for other
purposes, food being the main and are subject to changes. The technology used in
biodiesel production can also bring about economic risk both in the development and
operations in particular when the technology used is a novel technology that has not
132 S. S. Mansouri et al.

been extensively field tested. As such, from an economic perspective, it is necessary


to systematically assess the techno-economics of a biodiesel project where each
stage of the development process receives sufficient information in the most efficient
manner. Figure 6.1 presents a techno-economic framework that can be used to
systematically analyze and direct industrial biodiesel projects.

Background/
Case Study
Data collection

Information
Market value,
Availability 1. Economic potential Economic outlook
Tools
Spreadsheet based Calculation

Available? No
Stop
Revert to previous Step

Yes
TRL Identification/
Experimental Evidence
Tools 2. Technology
TRL assessment of project
Spreadsheet based Readiness Level (TRL)
Calculation
Yes

Acceptable
? No
Stop
Revert to previous step
Yes
Information
Costs of Equipment, Recovery and Risk
of TRL
3. Detailed economic TRL assessment of Resource/
Tools
Econoimic model
analysis technology pairs
Monte carlo
LOPA

Stop go No
decision
Stop
Revert to previous step
Yes

Industrial implementation

Fig. 6.1 A techno-economic framework for systematically analyzing and directing industrial
biodiesel projects
6 Economic Risk Analysis and Critical Comparison … 133

6.3.1 Step 1: Economic Potential Screening

In order to carry out a detailed techno-economic work, it is necessary to identify all


potential products that can be generated from a given feedstock quickly and effi-
ciently. This would be a process-oriented approach, where the process technology/
pathway is fixed and the products assessed are numerous. In comparison, a
product-oriented approach product is known. A product-oriented approach is rec-
ommended when the aim is to produce a specific product (or a set of products) and
the decision-maker wants to evaluate several paths for its production. This is the
case of a retrofit problem, where one wants to change or adapt the existing plant so
as to have more production routes and/or to have different sources of feedstock
being converted in the plant, whereas a process-oriented approach should be used
when the user is aiming to evaluate a number of paths to synthesize a set of
products from a certain (already selected) raw material and, therefore, the question
is the selection of product portfolio. Then, it would result in a completely new plant.
Based on this, the system boundaries are defined.
Once an approach has been decided on, the options generated must be evaluated
quickly and efficiently. In this instance, an economic potential analysis would
provide an efficient ranking method. To carry out the economic potential analysis,
the current market value of the products and the feedstock (including utility
chemicals and energy) together with a first-pass estimate desired production and
feed rates can be used for calculating annual economic potential as follows:

annual cash inflow ðnet revenueÞ ¼


total annual sales  anual production cost
   
ton $
total annual sales ¼ production ratei  market pricei
year ton
i ¼ 1; . . .; I

annual production cost ¼ variable operating costs

This simple yet efficient method of analysis has two distinct advantages:
1. The ranking and identification of the best feedstock and product combinations,
and
2. Understand if a product development through a biodiesel pathway is econom-
ically feasible.
The stop/go criteria in this step will depend on the investors but typically a
project that has a negative economic potential would be overlooked since the
process is essentially “destroying” the value of the feedstocks.
134 S. S. Mansouri et al.

6.3.2 Step 2: Technology Readiness Level (TRL) Assessment

The objective of Technology Readiness Level (TRL) assessment is to establish the


design space from which the best potential alternative will be identified. In this
respect, all the data required for the subsequent steps are gathered into a multidi-
mensional matrix (database) of process technologies, storing all the relevant data
required to perform the subsequent steps, and representing the alternatives within
the boundaries. The database includes (i) possible processing networks; (ii) the
parameters needed to solve the generic block model equations (e.g., stoichiometry,
conversion, mixing, product separation, waste separation) (Quaglia et al. 2015;
Cheali et al. 2014); and (iii) techno-economic data (such as prices of raw materials
and products and capital and operating costs). The required data can be obtained by
literature review as well as from real plant data and/or process simulation. It is
important to note that, when considering bioprocesses, it is common that real data
and literature data are unavailable or incomplete, and therefore data validation and/
or consolidation together with process simulation is often required (Gargalo et al.
2016). Lastly, the generic block model equations (see Quaglia et al. 2015; Cheali
et al. 2014 for more details) are solved using a mathematical solver, where the
processes are simulated in order to obtain the mass and energy balances, fulfilling
fundamental data requirements for the next steps.
The concept of TRL is an important parameter in developing a process tech-
nology from a concept to an implementable solution (Parasuraman 2000;
Technology Readiness Level 2012). The TRL concept allows the users to gauge the
level of development of technology with TRL-1 representing a conceptual idea to
TRL9 representing an implemented and in-operation design. The TRL framework
originally proposed by NASA has been adapted by many other disciplines
including in assessing newly developed processes in the area of chemical engi-
neering (Parasuraman 2000). Thereby, TRL evaluation allows us to identify and
differentiate between more established technologies and novel technologies. The
objective of this step is to assess the technological readiness of a given production
process and to understand if the process is sufficiently developed for industrial-scale
implementation. In general, higher TRL processes are preferred over lower TRL
processes as the high TRL represents a process that requires minimal development
costs and can be readily implemented. A process with lower TRL values will
require further development which needs to be added as into the initial capital
expenditure.

6.3.3 Step 3: Detailed Economic Analysis

In the previous two steps, an investor can identify if a project is worth further
investigating from both a technical and an economic point of view. The purpose of
step 3 is to accurately evaluate a project so a final stop/go decision to implement a
6 Economic Risk Analysis and Critical Comparison … 135

project can be taken. In this first part of a detailed economic analysis, a deter-
ministic approach can be taken. A deterministic analysis is relatively less complex
method and will take into account all the economic variables such as project
development expenses, capital expenditure, operational expenditure, borrowing
costs, etc. As such, this analysis would provide a good economic outlook of a
proposed project.
However, this analysis does not inherently take into account number of factors:
1. Economic uncertainties bought on market fluctuations,
2. Economic uncertainties bought on by variation in project development/CAPEX,
and
3. Economic uncertainties related to operational and development failure.

6.3.3.1 Deterministic Techno-Economic Analysis

To make an investment decision, the forecasted profit from an investment must be


assessed relative to some quantitative profitability measurements in regard to the
investment needed to generate that profit. As discussed in earlier work (Gargalo
et al. 2016), for the project evaluation, the economic model used is the discounted
cash flow rate of return (DCFROR) which considers the time value of money. This
model uses future cash flow predictions and discounts them to get a present value
estimate, which is then used to analyze its potential for investment. The DCFROR
takes into consideration that money to be received or paid at some time in the future
is viewed as having less value than an equal amount received or paid today.
Because the goal of investing is to increase the shareholder’s wealth, an investment
is worth undertaking if it creates value, and the investment creates value if it creates
more value than it costs, considering the time value of money (discounted cash in-
and outflows). Thus, a process to be economically viable/acceptable has to present a
net present value (NPV) higher than zero (at which point it breaks even, value
created is equal to costs). This means that it is expected to generate more revenue
than could be obtained by gaining the discount rate, representing the time value of
money which is the amount that could be obtained by investing in other alternatives
(Peters et al. 2003; Towler and Sinnott 2013). The discount rate is usually set by the
investors, and it reflects the minimum rate of return that the company has decided to
accept for a new investment, which is in fact a way for the company to adjust to the
built-in risk of a project. Therefore, by setting a target discount rate, this model
enables the calculation of the NPV as presented in Eqs. 6.1–6.8, and the minimum
selling price. A brief description of the assumptions for the economic model is
presented in Table 6.1.
NPV is therefore used to assess the economic viability of the projects within the
design space, which is then used to comprehensively compare the process alter-
natives. This method is especially recommended when uncertainties are present and
thus risk is a challenge (Short et al. 1995).
136 S. S. Mansouri et al.

Table 6.1 Average prices and respective standard deviations


Products and raw materials Mean ($/kg) References
Crude glycerol (60% w/w) 0.42 ICIS (n.d.)
1,2-Propanediol (1,2-PDO) 1.662 Techno-orbichem (n.d.)
1,3-Propanediol (1,3-PDO) 2.02 Davies (2014), Molel et al. (2015)

In order to generate the base case solutions and to identify the first ranking of
solutions, the DCFROR is initialized. To this end, the problem is formulated and
solved by maximizing NPV for each of the processing networks using the nominal
input parameters, such as prices of product(s) and raw material(s), capital, and
operating costs (data previously collected).

X
T
Ct
NPV ¼  C0
t¼1 ð1 þ rÞt
t ¼ 1; . . .; T ðinteger time periodsÞ ð6:1Þ
C t ¼ net cash inflow during the lifetime of the plant
C 0 ¼ total initial investment costs

X
T
NPV ¼ Annual present value  ðAnnual Fixed Investment CostÞ ð6:2Þ
t¼2

X
T  
1
Annual present value ðnet present revenueÞ ¼ annual cash inflow 
t¼1 ð1 þ r Þt
 
1
¼ discount factor
ð1 þ r Þt
r ¼ discount rate ¼
ð6:3Þ

annual cash inflowðnet revenueÞ ¼


¼ total annual sales  total anual production cost  income tax  loan payment
ð6:4Þ
   
ton $
total annual sales ¼ production ratei  market pricei
year ton ð6:5Þ
i ¼ 1; . . .; I

total annual production cost ¼ variable operating costs þ fixed operating costs
ð6:6Þ
6 Economic Risk Analysis and Critical Comparison … 137

 
$
variable operating costs
year
     
ton $ ton
¼ raw materialsl  market pricel þ utilities k
year ton year
       
$ ton $ $
 market pricek þ waste disposal  price þ repairs
ton year ton ton
   
$ $
þ operating supplies þ royalities
year year
k ¼ 1; . . .; K ðtotal number of utilitiesÞ
l ¼ 1; . . .; L ðtotal number of raw materialsÞ
ð6:7Þ

fixed operating costs ¼ labor ðoperators  salariesÞ þ maintenance


ð6:8Þ
þ property insurance and tax

6.3.3.2 Stochastic Analysis

While NPV analysis can be used to quantify the market and technology uncer-
tainties within a design space, stochastic analysis techniques are better equipped to
characterize the uncertainty and incorporate it in the economic metrics for project
evaluation. To this end, a comprehensive economic assessment needs to be per-
formed to characterize the uncertainties and incorporate such uncertainties in the
economic metrics for project evaluation.
As such, this requires the selection of an uncertainty analysis technique. One
method that can be used is the Monte Carlo technique which has two main steps. First,
the identified sources of uncertainty are characterized using appropriate statistical
distribution functions. The distribution function—portraying the behavior of the
uncertainty sources—can be obtained from (i) historical data if available (e.g., feed-
stock and product prices) and, when there is no data available, using (ii) expert review/
literature survey assuming uniform distribution (Helton and Davis 2003). Expert
reviews refer to information on typical ranges of variation that can be used as first
estimates and can be obtained from published sources and engineering textbooks such
as (Peters et al. 2003) or (Towler and Sinnott 2013). Therefore, we are able to evaluate
the project’s performance under uncertainty where worst-case scenarios are also
depicted. However, the framework is systematic and flexible, and the input data,
namely, uncertainty distributions, can be added or modified according to the updated
information the user may have at their disposal. As an example, historical data on the
raw materials and products market prices are collected and fitted through appropriate
distribution functions after analyzing their behavior over the years.
Once this step is completed, a sampling method such as Latin Hypercube
Sampling (LHS) is used to sample from the parameter space (Sin et al. 2009)
defined in the previous step, where the user needs to, a priori, specify the total
138 S. S. Mansouri et al.

sample number, N. After the sampling, one obtains a sampling matrix with N rows
and p columns, where N is the total number of samples and p refers to the number
of uncertain parameters for which the sampling is performed. After LHS, the
economic equations, describing the calculation of NPV, are evaluated/solved using
parameter values from the sampling matrix (usually called Monte Carlo simula-
tions). This procedure leads to N solutions for each sample set which result in
distributions of the model output values (i.e., NPV and Sc values) as a consequence
of the propagation of the parameter uncertainties from the input. By incorporating a
probabilistic interpretation, the distribution of the outputs will be analyzed by
means of empirical cumulative distribution functions (ECDF) and used as input
data in the next step.
Alternatively, risk analysis techniques such as Layer of Protection Analysis
(LOPA) (Willey 2014) or other uncertainty analysis techniques can also be used in
this analysis process. However, in comparison with the deterministic analysis, the
stochastic analysis using any method will require accurate detailed market infor-
mation and will cost both time and money to develop. These extra costs and
requirements are compensated by the better qualified NPV information that can be
derived.

6.3.3.3 Final Stop/Go Decision

The primary objective of this overall framework is to systematically assess the


techno-economic feasibility of the overall project. As such, this final stop/go
decision will dictate if the project will progress to the industrial implementation
phase. To this end, this stop/go decision will utilize available information that has
been generated in all previous steps. In general, a project with a positive NPV at a
given discount rate will be having a go decision. With the inclusion of the
uncertainty information, the stop/go decision can become more complicated as the
variability in the NPV and the corresponding uncertainty needs to be factored in. In
this instance, the risk is estimated as the probability of occurrence of certain event
times the consequence of that same event to happen. In this work, the economic risk
(Riskecon ) is quantified by the probability of failing to achieve the targeted NPV
(“being lower or equal to”) times the magnitude of the consequence of that hap-
pening (“loss of profit”). Therefore, the economic risk is given by the probability of
the project being non-profitable (NPV < 0) times the loss of profit in the event of
that happening (consequence). The respective mathematical description is presented
in Eq. 6.9.
X X
Riskecon ¼ Pi  M i ¼ PðNPV  0Þ  ðNPVi Þ ð6:9Þ
i i

where i is the occurrence of the undesirable event, Pi is the probability of that event
to occur, and Mi is the magnitude of the consequence (loss of profit in MM$).
6 Economic Risk Analysis and Critical Comparison … 139

6.4 Case Study

The production of value-added chemicals from glycerol is expected to act as a


potential driving force to increase global sustainability of the biodiesel industry
(Yang et al. 2012; Almeida et al. 2012). To this end, several potential pathways for
the conversion of glycerol into value-added products have been proposed. These are
1,2-propanediol (1,2-PDO) and 1,3-propanediol (1,3-PDO) (Almeida et al. 2012;
De Jong 2011). This is due to the fact that both products have several known
applications, but also due to their high impact on the market as potential replace-
ments (if not all in part) of their petro-based counterparts. 1,2-PDO (propylene
glycol) has a broad range of uses, from industrial applications for the production of
unsaturated polyester resins, coolants, and antifreeze, to paints and coatings.
Although smaller, there is also a market for higher grade product for the health and
personal care applications. Similarly, the 1,3-PDO commercialization also targets
industrial applications such as textiles, coatings, and engineering plastics, and its
large expected growth is as a building block for the production of biopolymers.

6.4.1 Step 1: Economic Potential

Upgrading regular glycerol into the two value-added products of 1,2-PDO and
1,3-PDO should in principle have a strong economic potential. However, it is
important to validate this hypothesis prior to proceeding to the future steps. For the
purpose of this case study, we will base our calculation upon a medium-sized
biodiesel plant in Europe (Green Business 2010). Being glycerol produced in
stoichiometric quantities to biodiesel (10 kg biodiesel to 1 kg of glycerol), it pro-
vides approximately 10 kton glycerol/year, corresponding to 10% (w/w) of the total
annual biodiesel production (100 kton/year). The total production rates of 1,2-PDO
and 1,3-PDO are 7740 ton/year and 5500 ton/year, respectively. This information
together with information available in Table 6.2 can be used to carry out a simple
economic potential calculation.
In this instance, the economic potential of the two projects can be calculated as

Economic Potential 1; 2-PDO ¼ Revenue generated  cost of raw materials


¼ 1:662 $=kg  7; 740 ton=year  0:42  10; 000
¼ USD 8:6 Million
Economic potential 1; 3-PDO ¼ 2:02  5; 500 ton=year  0:42  10; 000
¼ USD 6:9 Million

Based on the above simple economic potential calculations, both projects should
proceed to the next step.
140 S. S. Mansouri et al.

Table 6.2 Summary of the assumptions used for the discounted cash flow rate of return (Peters
et al. 2003; Humbird et al. 2011; Waldron 2014)
Parameter Assumption
Plant life (years) 20
Discount rate (mar) 10%
Depreciation period (Years) 10 (MACRS system)
Equity 40%
Loan interest 5%
Loan term (Years) 10
Construction period (Years) 2
% Spent in Year −1 60%
% Spent in Year 0 40%
Start-up time (Years) 0.50
Product production/Feedstock use (% of Normal) 50%
Variable costs (% of Normal) 75%
Fixed cost (% of Normal) 100%
Income tax rate 35%
Cost year for analysis 2014

6.4.2 Step 2: Technology Readiness Level (TRL) Assessment

In a typical industrial case study, this step needs to be performed to identify the
technologies that are available to carry out the transformation process of glycerol
into 1,2-PDO and 1,3-PDO, respectively. However, in this case study, this has
already been established by authors in (Živković et al. 2017). The 1,2-PDO pro-
duction process is given by the sequential processes of dehydrogenation–hydro-
genation of glycerol via hydroxyacetone, whereas 1,3-PDO is obtained through
glycerol fermentation by an engineered strain of K. pneumoniae followed by
reactive extraction with isobutyraldehyde.
The technology used to transform glycerol into 1,2-PDO is standard chemical
engineering unit operations that use established technology and process designs. In
comparison, the 1,3-PDO uses a fermentation pathway that is slightly less estab-
lished at full-scale implementations. As such, 1,2-PDO should have a higher TRL
than 1,3-PDO. However, for the purpose of this case study, both the process
pathways are considered as established technologies with pilot and full-scale
industrial units in operation to simplify the analysis. Hence, both the proposals will
fulfill the go criteria required to proceed to the next step.
6 Economic Risk Analysis and Critical Comparison … 141

6.4.3 Step 3: Detailed Economic Analysis

This step represents key stop/go criteria in the decision-making process as it will
combine information that has been generated in the past sections to make an
informed decision about the economics of the project as well as the economic
uncertainties attached to it. To this end, this part consists of a compressive analysis
that takes into account the following factors
1. Economic uncertainties bought on market fluctuations,
2. Economic uncertainties bought on by failure of technology,
3. Economic uncertainties bought on by variation in project development/Capital
expenditures (CAPEX), and
4. Economic uncertainties bought on by operational failure.
However, in this instance, both points 2 and 4 do not need to be taken into
consideration due to the simplification made above and assuming no major oper-
ational failure.

6.4.3.1 Economic Model Development

Based on these simplifications and the available information on estimated mass and
energy balances, the input data, and by following the generic block model equations
as presented in (Quaglia et al. 2015; Cheali et al. 2014), Eq. 6.10 can be derived.
The corresponding key assumptions are presented in Table 6.2.
" !
X
T
Ct XT
1   X X X
NPV ¼ t  C0 ffi t xp  P p  xrm  Prm þ xutl  Putl þ xwm  Pwm þ f ðFCIÞ
t¼1 ð1 þ rÞ t¼1 ð1 þ rÞ k l m
!#
XT
 MACRSðFCIÞ  C0
t¼2

ð6:10Þ
 
where xp ; Pp ; xrmi ; Prmi ; xutl ; Putl ; xwm ; Pwm ; FCI correspond, respectively, to sales
volume, product price, raw material(s) inflow, price of raw material(s), utilities
needed, price of utilities, waste outflow(s), waste(s) treatment price(s), and fixed
capital investment.
Furthermore, the techno-economic data required for the Step 3 is also collected
such as spot market prices of products and raw material(s) (Table 6.2). In this table,
MACRS denotes modified accelerated cost recovery system which is the current tax
depreciation system in the United States.
142 S. S. Mansouri et al.

6.4.3.2 Deterministic Economic Assessment

To obtain the first ranking of solutions, the problem is solved by maximizing the
NPV of the alternatives within the design space. To this end, all the input data
required for the calculation of Eq. 6.10 are defined beforehand and have already
been collected or calculated (e.g., fixed and variable operating costs). In this study,
the fixed and variable operating costs were calculated based on the factorial
methodology. The capital investment was calculated based on dividing the pro-
cessing routes/pathways into three sections as shown in Fig. 6.1. The purchased
equipment costs were then estimated based on the product-specific references as
reported in (Gargalo et al. 2016) and using the power law equation for the pre-
diction of the purchased capital investment (Peters et al. 2003), where the expo-
nential is 0.7 and 0.75 for chemical and biochemicals processes (Stanbury 1995),
respectively. A summary of the data required for the application of the economic
model is presented in Table 6.3 along with the obtained deterministic NPV and
MSP values.

6.4.3.3 Stochastic Economic Assessment—Monte Carlo Technique

In this step, the uncertainty related to key model inputs is described and charac-
terized so as to understand its impact on the techno-economic indicators such as
NPV and MSP. Therefore, the Monte Carlo technique is used in order to propagate
the uncertainty from the inputs to the model outputs.

Characterization of Sources of Uncertainty

The sources of uncertainty identified (Ɵj,…,M) are, as shown in Eq. 6.10, the crude
glycerol price (Prm ) and product’s price (Pp ), and the fixed capital investment (FCI).
These have been identified in the previous study as key input parameters which
convey high uncertainty to the techno-economic evaluation (Gargalo et al. 2016). In
this step, they are characterized using suitable statistical distribution functions. To
this end, as presented in Table 6.4, the stochastic assessment is based upon the

Table 6.3 Purchased capital cost to be used for the factorial methodology remaining needed
parameters. NPV and MSP are estimated at the conditions reported in Table 6.2
Purchased Fixed Total product Utilities Sales NPV MSP
capital cost, capital cost without (MM$/ (MM @10 @10
E′ (MM$) investment, depreciation, y) $/y) (MM ($/
[References] FCI (MM$) TPC ($/kg) $) kg)
1,3-PDO 5.347 (Molel 22.278 2.045 1.971 10.6 −28.1 2.71
et al. 2015)
1,2-PDO 4.713 16.280 1.421 2.848 12.2 −6.0 1.79
(Cabaniss
et al. 2014)
6 Economic Risk Analysis and Critical Comparison … 143

Table 6.4 Identified sources of uncertainty for economic analysis


Glycerol historical prices Lognormal distribution (0.048 std.) (Gargalo et al. 2016)
(2007–2014)
1,2-PDO historical prices Lognormal distribution (0.28 std.) (Gargalo et al. 2016)
(2007–2014)
1,3-PDO historical prices Lognormal distribution (0.35 std.) (Gargalo et al. 2016)
(2007–2014)
Fixed capital investment Uniform distribution (−20 to +50%) (Christensen and Dysert
2005; Cheali et al. 2015)

Table 6.5 Summary of 1,2-PDO 1,3-PDO


results for the calculation of
economic risk for 1,2-PDO Frequency of selection 327 173
and 1,3-PDO Pr (NPV  0) @IRR10% 0.768 0.908
Pr (NPV  0) @IRR24% 0.95 0.965
Risk @IRR10% (MM$) 17.8 40.7
Risk @IRR24% (MM$) 25.4 45.4

2.2

1.8

1.6

1.4
price ($/kg)

1.2

0.8

0.6

0.4

0.2
0 2 4 6 8 10 12 14
time (2007 - 2014)

Fig. 6.2 Input data uncertainty for the products and feedstock: prices of crude glycerol (gray line),
1,3-PDO (blue line) and 1,2-PDO (violet line), in Europe, from 2007 to 2014

following: historical price data for the crude glycerol, 1,2-PDO and 1,3-PDO prices
(as presented in Fig. 6.2); and variability of the fixed capital investment over a
typical range of variations.
144 S. S. Mansouri et al.

LHS and Monte Carlo Sampling Methodology

Through LHS sampling, the uniform coverage of the uncertain space is achieved,
where a rank correlation method is used so as to reflect the correlation between the
uncertain parameters in the generated future scenarios/samples (Towler and Sinnott
2013; Short et al. 1995). Thus, 500 future scenarios were created from the input
uncertainty domain corresponding to 500 realizations of uncertainty. The output of
this step was therefore an N  M matrix, where N represents the number of samples
and M represents the uncertain parameters, as presented in Eq. 6.11. Furthermore,
as presented in Eq. 6.12, to identify the optimal alternative under stochastic con-
ditions, Monte Carlo simulations were performed, where the problem was formu-
lated and solved deterministically for each one of the 500 samples, generating 500
sets of possible economic indicators (NPV and MSP). The frequencies of selection
of 1,2-PDO and 1,3-PDO were 65% and 35%, respectively (also presented in
Table 6.5).
2 3
Generation of N samples with LHS from the ssi;j  ssi;M
6 .. 7 ð6:11Þ
uncertainty domain of Ɵj,…,M SNM ¼ 4 ... ..
. . 5
ssN;j  ssN;M
Monte Carlo simulations: Deterministic maxf ðx; yÞ ¼ maxNPV
optimization problem solved for each scenario ssi,j 2 3
yi;k    yi;K
! Mapping and analysis of solutions 6 .
YNK ¼ 6 .. .. 7
7
4 .. . . 5 ð6:12Þ
ssN;k  yN;K

Stop/Go Decision

Aiming to assist the decision-maker to further invest or reject a given project, the
framework guides the user through a techno-economic feasibility analysis under
uncertainties. It therefore enables the decision-maker to take informed decisions on
whether and where to invest, based on economic risk, acknowledging that the
economic parameters are not static but ever-changing. Thus, considering a system
under uncertainty, if one detects and analyses the risk of the objective function
reaching a value lower or equal to a given target/level/threshold, this can be
expressed as the probability of failure to achieve the target times the consequence of
failing to achieve that same target. A summary of the economic risk results is
presented in Table 6.5. Figure 6.3 also graphically presents the results for both
products under analysis.
In order to further investigate the yearly effect of the discount rate on the
economic model, an additional scenario analysis was performed by testing
the model at a discount rate of 24%. This reflects a more cautious attitude of the
6 Economic Risk Analysis and Critical Comparison … 145

Fig. 6.3 Cumulative distribution function for the 1,2-PDO and 1,3-PDO production from
glycerol. The highlighted area represents the risk of the project being non-profitable. Blue
represents NPV obtained for a discount rate of 10%. Red represents NPV obtained for a discount
rate of 24%

investor toward the potential investment, leading to a higher economic risk, as can
be observed in Fig. 6.3 and in Table 6.5.
The valorization of crude glycerol through the production of 1,2-PDO is
potentially a better investment option since it has lower risk of being non-viable/
non-profitable than 1,3-PDO, having an impending profit loss of 18 MM$ (at 10%
discount rate), as presented in Table 6.5. However, the stop/go decision will depend
on the type of investor that is financing the project. For example, venture capitalists
and government agencies might take on the project for multitude of economic,
societal, or environmental reasons. Even with these considerations, the 1,2-PDO
project will likely get the go decision. The technology that is used in the 1,2-PDO
process is an established process that has pilot- and full-scale plants. Hence, the
implementation of the process does not require separate development steps.
A detailed design based on required plant capacity can be developed and built.

6.5 Conclusions

In this chapter, a generic techno-economic methodology for risk analysis of


biorefinery processes was proposed. The application of the methodology was
demonstrated in a biodiesel-related case study. Nowadays, being glycerol a surplus
product, the robustness of two glycerol-based biorefinery concepts was tested
herein for the valorization of glycerol. The uncertainties in the input parameters
were taken on board and propagated to the techno-economic assessment by using
the Monte Carlo technique enhanced with Latin hypercube sampling. The
146 S. S. Mansouri et al.

corresponding economic risk was quantified as the probability of failing to achieve


a positive NPV. Thus, it was demonstrated that 1,2-PDO is the potential best
alternative to add value to the glycerol surplus, since it is the design that presented
the lowest economic risk. Hover, both alternatives present high risk of failure and
impending loss of 18 and 25 MM$ for 1,2-PDO and 1,3-PDO, respectively.
Moreover, this concept potentially contributes to the diversification of the product
portfolio for vegetable oil-based biorefineries contributing to more economically
sustainable solutions.

References

Almeida JRM, Fávaro LCL, Quirino BF (2012) Biodiesel biorefinery: opportunities and
challenges for microbial production of fuels and chemicals from glycerol waste. Biotechnol
Biofuels 5:48. https://doi.org/10.1186/1754-6834-5-48
Azapagic A, Millington A, Collett A (2006) A methodology for integrating sustainability
considerations into process design. Chem Eng Res Des 84:439–452. https://doi.org/10.1205/
cherd05007
Cabaniss S, Park D, Silvinsky M, Wagoner J, Chen D, You F (2014) Design G2
Cheali P, Gernaey K, Sin G Uncertainties in early-stage capital cost estimation of process design—
a case study on biorefinery design. Front Energy Res 3 (2015). https://doi.org/10.3389/fenrg.
2015.00003
Cheali P, Quaglia A, Gernaey KV, Sin G (2014) Effect of market price uncertainties on the design
of optimal biorefinery systems—a systematic approach. Ind Eng Chem Res 53:6021–6032
Christensen P, Dysert LR (2005) Cost estimate classification system—as applied in engineering,
procurement, and construction for the process industries
Davies P (2014) Chemical business focus—a monthly roundup and analysis of the key factors
shaping world chemicals markets. Bio-materials and inter mediates including biobased
chemicals, bio-polymers and their petrochemical equivalents
De Jong E (2011) Bio-based chemicals value added products from biorefineries. A report prepared
for IEA bioenergy-task, 36
Eurostat (2017) http://ec.europa.eu/eurostat
Gargalo CL, Sin G (2015) Sustainable process design under uncertainty analysis: targeting
environmental indicators. Comput Aided Chem Eng 37:2579–2584. https://doi.org/10.1016/
B978-0-444-63576-1.50124-2
Gargalo CL, Carvalho A, Gernaey KV, Sin G (2016a) A framework for techno-economic and
environmental sustainability analysis by risk assessment for conceptual process evaluation.
Biochem Eng J 116:146–156. https://doi.org/10.1016/j.bej.2016.06.007
Gargalo CL, Cheali P, Posada JA, Carvalho A, Gernaey KV, Sin G (2016b) Assessing the
environmental sustainability of early stage design for bioprocesses under uncertainties: an
analysis of glycerol bioconversion. J Clean Prod 139:1245–1260. https://doi.org/10.1016/j.
jclepro.2016.08.156
Gargalo CL, Cheali P, Posada JA, Gernaey KV, Sin G (2016c) Economic risk assessment of early
stage designs for glycerol valorization in biorefinery concepts. Ind Eng Chem Res 55:6801–
6814. https://doi.org/10.1021/acs.iecr.5b04593
Gong J, You F (2015) Sustainable design and synthesis of energy systems. Curr Opin Chem Eng
10:77–86. https://doi.org/10.1016/j.coche.2015.09.001
Green Business (2010) Factbox: biodiesel plants in the EU
6 Economic Risk Analysis and Critical Comparison … 147

Helton JC, Davis FJ (2003) Latin hypercube sampling and the propagation of uncertainty in
analyses of complex systems. Reliab Eng Syst Saf 81:23–69. https://doi.org/10.1016/S0951-
8320(03)00058-9
Humbird D, Davis R, Tao L, Kinchin C, Hsu D, Aden A, Schoen P, Lukas J, Olthof B, Worley M,
Sexton D, Dudgeon D (2011) Process design and economics for biochemical conversion of
lignocellulosic biomass to ethanol process design and economics for biochemical conversion of
lignocellulosic biomass to ethanol, golden, Colorado 80401
ICIS (n.d.) ICIS dashboard price history—crude glycerol, 25 Aug 2010–25 Aug 2015
Makkar HPS (2012) Biofuel co-products as livestock feed: opportunities and challenges. Food and
Agriculture Organization of the United Nations, Rome
Mansouri SS, Ismail MI, Babi DK, Simasatitkul L, Huusom JK, Gani R (2013) Systematic
sustainable process design and analysis of biodiesel processes. Processes 1:167–202. https://
doi.org/10.3390/pr1020167
Molel E, Phillips H, Smith A (2015) 1,3-Propanediol from crude glycerol, Pennsylvania
National Research Council (2011) Renewable fuel standard: potential economic and environmen-
tal effects of U.S. biofuel policy. The National Academies Press, Washington, D.C.
Parasuraman A (2000) Technology readiness index (Tri): a multiple-item scale to measure
readiness to embrace new technologies. J Serv Res 2:307–320. https://doi.org/10.1177/
109467050024001
Peters MS, Timmerhaus KD, West RE (2003) Plant design and economics for chemical engineers.
McGraw-Hill
Quaglia A, Gargalo C, Sin G, Gani R (2015) Systematic network synthesis and design: problem
formulation, superstructure generation, data management and solution. Comput Chem Eng
72:68
Sacramento-Rivero JC (2012) A methodology for evaluating the sustainability of biorefineries:
framework and indicators. Biofuels Bioprod Biorefin 6:32–44. https://doi.org/10.1002/bbb.335
Short W, Packey DJ, Holt T (1995) A manual for the economic evaluation of energy efficiency and
renewable energy technologies. National Renewable Energy Laboratory, US Department of
Energy, Colorado. DE-AC36-83CH10093
Sin G, Gernaey KV, Neumann MB, van Loosdrecht MCM, Gujer W (2009) Uncertainty analysis
in WWTP model applications: a critical discussion using an example from design. Water Res
43:2894–2906. https://doi.org/10.1016/j.watres.2009.03.048
Stanbury PF (1995) Principles of fermentation technology. Pergamon
Technology Readiness Level, NASA (2012). https://www.nasa.gov/directorates/heo/scan/
engineering/technology/txt_accordion1.html
Toro Chacón FA (2004) Techno-economic assessment of biofuel production in the European
Union, Karlsruhe
Towler G, Sinnott R (2013) Economic evaluation of projects. In: Chemical engineering design,
2nd edn. Elsevier, pp 389–429. http://dx.doi.org/10.1016/B978-0-08-096659-5.00009-2
Towler G, Sinnott R (2013) Chemical engineering design: principles, practice and economics of
plant and process design, 2nd edn. Elsevier
Turton R (2009) Analysis, synthesis and design of chemical processes. Prentice Hall
Waldron KW (2014) Advances in biorefineries: biomass and waste supply chain exploitation.
Woodhead Publishing
Willey RJ (2014) Layer of protection analysis. Procedia Eng 84:12–22. https://doi.org/10.1016/j.
proeng.2014.10.405
Techno-orbichem (n.d.) Techno-orbichem: monopropylene glycol
Yang F, Hanna M, Sun R (2012) Value-added uses for crude glycerol—a byproduct of biodiesel
production. Biotechnol Biofuels 5:13. https://doi.org/10.1186/1754-6834-5-13
You F, Tao L, Graziano DJ, Snyder SW (2012) Optimal design of sustainable cellulosic biofuel
supply chains: Multiobjective optimization coupled with life cycle assessment and input-output
analysis. AIChE J 58:1157–1180. https://doi.org/10.1002/aic.12637
148 S. S. Mansouri et al.

Yue D, Kim MA, You F (2013) Design of sustainable product systems and supply chains with life
cycle optimization based on functional unit: general modeling framework. Mixed-integer
nonlinear programming algorithms and case study on hydrocarbon biofuels. ACS Sustain
Chem Eng 1:1003–1014. https://doi.org/10.1021/sc400080x
Živković SB, Veljković MV, Banković-Ilić IB, Krstić IM, Konstantinović SS, Ilić SB, Avramović
JM, Stamenković OS, Veljković VB (2017) Technological, technical, economic, environmen-
tal, social, human health risk, toxicological and policy considerations of biodiesel production
and use. Renew Sustain Energy Rev 79:222–247. https://doi.org/10.1016/j.rser.2017.05.048
Chapter 7
Techno-economical Aspects of Biodiesel
Plants

Syed Taqvi, Mohamed Elsholkami and Ali Elkamel

Abstract The main aim of this chapter is to provide an overview of the technical
and economical characteristics of biodiesel production plants. A literature review of
various techno-economic feasibility studies of biodiesel production is conducted.
Moreover, outcomes of these evaluations are reported to present potential com-
merciality and near-term technical viability of biodiesel plants from different
regions. In this chapter, various technological possibilities and economic aspects
involved in the production of biodiesel are outlined. Significant effort is made to
ensure that common assumptions are used as the basis for the comparison among
conversion technologies. In addition, this analysis constitutes of the input feedstock
needed, environmental considerations, energy balances, and detailed economic
assessment including the cost component structures for biodiesel. Conclusions
made in this study are merely indicative of the expected performance of biodiesel
production plants based on the current state of public knowledge. These findings
can aid in conducting significant comparisons between different technologies for
producing biodiesel and exploring scenarios with optimal biodiesel production
configurations.

7.1 Overview

Countries worldwide strive to achieve the objective of national independence for


energy and reduction in greenhouse gas (GHG) emissions, which have stimulated
the development of biofuel-based energy production technologies. Among the
commercially producible renewable biofuels, biodiesel is generally considered to be
one of the simplest to produce. Biodiesel is characterized by having comparable
performance and efficiency to petroleum-based diesel fuel, while having consid-
erably lower emissions associated with its combustion. The global demand for

S. Taqvi  M. Elsholkami  A. Elkamel (&)


Department of Chemical Engineering, University of Waterloo, 200 University Avenue West,
Waterloo, ON N2L 3G1, Canada
e-mail: aelkamel@uwaterloo.ca

© Springer Nature Switzerland AG 2019 149


M. Tabatabaei and M. Aghbashlo (eds.), Biodiesel,
Biofuel and Biorefinery Technologies 8,
https://doi.org/10.1007/978-3-030-00985-4_7
150 S. Taqvi et al.

biodiesel was forecasted to increase by 14% from 2016 to 2020, which amounts to
an increase of approximately 4.7 billion liters (Glystra et al. 2020).
In addition to limitations associated with the availability of sufficient feedstock
to achieve required production levels, the considerably higher production cost of
biodiesel is a barrier to its replacement for conventional diesel. There are various
factors that contribute to the technical and economic feasibility of the production of
biodiesel, which play a significant role in determining its cost competitiveness in
comparison with the use of petroleum-based diesel. These factors include the type
of feedstock used, catalyst employed, retrofits implemented to achieve certain
production objectives, the final application considered for the produced biodiesel,
and the geographic region in which the production plants exist or are proposed to
set in (Huang et al. 2016; Lee and Ofori-Boateng 2013). In order to identify the
various socioeconomic impacts of biodiesel production and to assess the economic
sustainability of production processes, there are several studies in the literature that
focused on analyzing the various technical and economic aspects in the production
of biodiesel.
Extensive techno-economic assessments for biodiesel production processes that
mainly focus on the technical and economic aspects of the production stages are
available in the literature. For example, Huang et al. (2016) analyzed the technical
and economic feasibility of biodiesel and ethanol coproduction from sugarcane.
Other authors, such as Albuquerque et al. (2016), Apostolakou et al. (2009), and
Marchetti et al. (2008), assessed the feasibility of using alternative feedstocks.
However, majority of these studies excluded the economics of the feedstock pro-
duction stages depending only on ex-factory price data, even though capital and
operational costs associated with these stages account to approximately more than
60% of total costs (Marchetti and Errazu 2008; Enguidanos et al. 2002).
The aim of this chapter is to highlight the various factors that affect the technical
and economic sustainability of biodiesel production processes, as well as the impact
of these processes on the communities they serve. Each of these factors may prove
to be vital than the other in different scenarios. This will be presented in the
following subsections based on a review of various reports in the literature that
focus on the evaluation of various technical and economic aspects of biodiesel
production processes.

7.2 Technical Aspects

At the initial stage, biodiesel plants are assessed for their technical feasibility. This
involves determining whether production of biodiesel is viable under a set of
conditions in a specified environment. In addition, the control of specific variables
within the process that aid in producing the maximum yield is also presented.
7 Techno-economical Aspects of Biodiesel Plants 151

7.2.1 Raw Materials

Among the various techniques employed, transesterification is the most common


method of converting biomass into biodiesel. A number of raw materials are
consumed in the production of biodiesel through this process. These mainly involve
the feedstock and the alcohol. The feedstock refers to the natural resources (i.e.,
biomass) from which the oil is extracted. The alcohol is used to transform this oil
into biodiesel, producing glycerol (glycerine) as a by-product. From a technical
perspective, the ability of a certain feedstock to convert into biodiesel, the con-
version efficiency, and the availability of that particular feedstock affect the process.
Vegetable oils and animal fats have been known to be the primitive sources of
biodiesel. Nevertheless, feasibility studies on a lot of other biomass have been
carried out recently, such as (micro)algae, jatropha, palm oil, soybean, brown
grease (industrial, commercial, and municipal sewage traps’ grease), yellow grease
(waste cooking oil), restaurants and rendered animal fats, industrial hemp
(Cannabis sativa Linn), macauba, oilseeds, and stillingia. However, they vary in
their conversion efficiencies in order to produce biodiesel.
The availability of the feedstock in order to ensure a steady flow of raw materials
is another factor affecting technical feasibility. A particular feedstock may be found
at a specific time of the year or it may not be readily available in a predicted
quantity. For example, the supply for vegetable oil may be less in winter due to the
inability of producing fresh crops at relatively larger scale. In the latter scenario,
predicting the amount of waste cooking oil collected from restaurants or grease
removed from grease traps installed in equipment may be difficult (Gao et al. 2012).
If the availability of feedstock is hindered, the production of biodiesel will be
limited by the quantity collected. In contrast, algae is a biomass that is highly
productive and grows far more yield than any other biomass used for the production
of biodiesel (Gao et al. 2012; Mata et al. 2010). It requires less land area to cultivate
(i.e., 5.87–13.69 L/m2) and has high photosynthetic efficiency as compared to other
crops (Mata et al. 2010; Nagarajan et al. 2013; Wu et al. 2012). In addition, they
can be easily cultivated in almost any environments and are not perceived for
human consumption (Mata et al. 2010).

7.2.2 Temperature and Pressure

The conditions at which a particular process operates at can influence the yield and
quality of the product and/or by-product(s). In biodiesel production, the reaction
conditions have been found to strongly affect the rate of reaction (Barnwal and
Sharma 2005). Barnwal and Sharma conducted a study on investigating the
potential of biodiesel production from vegetable oil in India. They found that
the pretreatment (i.e., refining, pre-esterification, and removal of fatty acids) of the
vegetable oil was required if the operating conditions were mild (i.e., 60–70 °C and
152 S. Taqvi et al.

1 atm) (Barnwal and Sharma 2005). However, when operating at high conditions
(i.e., 240 °C and 9000 kPa), no pretreatment was necessary. Nevertheless, maxi-
mum yield was observed at a temperature range of 60–80 °C (Barnwal and Sharma
2005). In another study conducted on a biodiesel plant in Greece, a conversion of
98% was seen when operating at 65 °C and 20 psi (Skarlis et al. 2012).
In some cases, the operating conditions are set at a higher level, as dictated by
the process requirements. A study, evaluating the effect of homogeneous and
heterogeneous catalysts, showed that the latter type of catalysts operated at high
temperatures (Kiss et al. 2010). A study conducted by Marchetti et al. depicted that
heterogeneous catalysts operated at a temperature of 150 °C while homogeneous
catalysts operated at 60 °C (Marchetti et al. 2008).

7.2.3 Catalyst

Catalysts, in biodiesel production, are classified according to two characteristics:


(i) pH (acidic or basic) and (ii) phase (homogeneous or heterogeneous). Each of the
two properties affects the technical viability differently. According to a study,
producing biodiesel from vegetable oil in India, it is found that alkali metal
alkoxides performed much better than acidic catalysts. A concentration of 0.5–1%
of alkali catalysts resulted in a 94–99% conversion of vegetable oil into esters
(Barnwal and Sharma 2005). In another work, acidic catalysts were seen to operate
at relatively higher temperatures than alkali catalysts (Zhang et al. 2003).
A study, evaluating homogeneous and heterogeneous catalysts, found that a
relatively higher yield was obtained using heterogeneous catalysts, in biodiesel
production, than homogeneous catalysts. The yield relative to the vegetable oil used
was about 100.3% and 99.5% for heterogeneous and homogeneous catalysts,
respectively (Kiss et al. 2010). As mentioned earlier, heterogeneous catalysts
operate at a higher operating temperature (Marchetti et al. 2008). Moreover, higher
purity in glycerine is yielded (Kiss et al. 2010).

7.2.4 Properties of Biodiesel

Certain biomass, like vegetable oil, can be utilized directly into diesel equipment, in
its raw form. However, this reduces the problem-free operating lifetime of the
engine (Barnwal and Sharma 2005). First, the vegetable oil is a lot more viscous
than biodiesel which leads to injection problems. Second, it has relatively larger
particles which makes it difficult to process within the engine. Last, the inefficient
mixing of vegetable oil with air leads to incomplete combustion (Barnwal and
Sharma 2005). Therefore, the need for processing of vegetable oil into biodiesel
rises. There are certain standards, such as those set by American Society for Testing
and Materials (ASTM), defining the quality of the biodiesel that is quite compatible
7 Techno-economical Aspects of Biodiesel Plants 153

with diesel engines. These standards focus on physical and chemical properties
such as kinematic viscosity, cetane number, heating value, cloud point, and density
(Barnwal and Sharma 2005). A particular source of biomass may possess the
potential to be used as biodiesel. However, due to its poor quality (i.e., not up to the
standard), it may harm the equipment and may not be widely acceptable. Therefore,
these properties affect the technical feasibility of a certain process.

7.2.5 Reactor

The type of reactor used to convert a particular biomass into biodiesel may also
affect its viability. A study, focusing on using microbial oil as a source to biodiesel,
showed that continuous flow reactor had a relatively lower yield (Koutinas et al.
2014). Another review study compared cases between the use of batch reactors and
continuous reactors (Mata et al. 2010). Additionally, the reactor design, size, and
shape could also affect its performance. In extension, it affects the overall process
outcome (i.e., yield and quality) (Ofori-Boateng et al. 2012).

7.2.6 Environment

Characteristics of the environment tend to affect the growth of a particular biomass.


Since harvesting feedstock is a vital process in biodiesel production, the efficiency
with which the process occurs affects the entire biodiesel production. For
microalgae feedstock, the effects of an open versus closed culture system on
feedstock cultivation were studied (Mata et al. 2010). The amount and quality of
light, made available to the samples, was stated to be the most important factor. In
addition, the levels of oxygen, carbon dioxide, pH, salinity, and several other
factors were considered to influence the feedstock growth (Mata et al. 2010;
Nagarajan et al. 2013; Mulugetta 2009).

7.3 Economic Factors

Once the process has been proven to be technically viable, the economics of the
project is studied to determine whether there are sufficient resources to implement
the plan. Several types of costs are involved that need to be met by the outcome of
the process, during its lifetime. If a certain process does not generate the target
revenue and/or cover its expenses within the project lifetime, it is considered
infeasible.
154 S. Taqvi et al.

7.3.1 Raw Materials

Raw materials were seen to have a significant influence on the technical feasibility
of a biodiesel plant. When studying the economic practicality of a biodiesel plant,
the cost associated with the raw materials is a major factor. This cost comprises
purchasing/cultivating/harvesting the raw materials and/or collecting them from
various locations within the region. If biodiesel was to be produced via transes-
terification, this would involve the cost of the feedstock, alcohol, and catalyst used
in the reaction.
In almost all biodiesel plant cost–benefit analyses, the cost of raw materials is the
single most expensive item on the expense list. Even in the case where waste
products, such as yellow and brown grease, are utilized, acquiring them is chal-
lenging. In a study, conducted on the feasibility of biodiesel production from waste
cooking oil, 76–80% of the operating cost was the cost of acquiring the raw
material (i.e., waste cooking oil) (Marchetti et al. 2008). In scenarios where virgin
raw materials (i.e., vegetable oil) are used, the cost of raw materials can account for
up to 90% of total production cost of the biodiesel plant (Apostolakou and Kookos
2009). In all, Zhang et al. found virgin oil to be about 2–3 times more expensive
than waste cooking oil (Zhang et al. 2003).

7.3.2 Capital Cost

Capital cost mainly refers to the cost of equipment installed in order to produce
biodiesel from the feedstock. In some studies, the cost of the land, used for feed-
stock cultivation and/or processing, is also included in the capital cost (Nagarajan
et al. 2013). Capital cost accounts for about 69–92% of the total investment cost of
a biodiesel plant (Skarlis et al. 2012).
Equipment cost can vary significantly based on the type and configuration of
process, number of units employed, geometry of vessels, and material of con-
struction. The most expensive equipment is the one used for transesterification
(Fortenbery 2005). Nevertheless, the cost associated with equipment can be low-
ered with appropriate selection of technology. For example, coupling of the oil
extraction process with transesterification, using ultrasound technology, can reduce
the capital cost significantly (Nagarajan et al. 2013). In another case, equipment
cost of a process involving heterogeneous catalysts was 40% cheaper than the
process which involved homogeneous catalysts (Marchetti et al. 2008). Moreover,
less number of units are required if the process is simple. In the case where virgin
vegetable oil is used, the production involves less complex operations. The cost of
raw materials is observed to be high but with a lower capital cost (Zhang et al.
2003).
Zhang et al. studied a process where an acidic catalyst was used along with an
additional hexane extraction process, in the production of biodiesel from waste
7 Techno-economical Aspects of Biodiesel Plants 155

vegetable oil (Zhang et al. 2003). A larger column was required to handle the
increase in load, due to hexane/water addition. Furthermore, two stainless steel
columns were needed, as opposed to carbon steel, due to the acidic nature of the
catalyst that may promote corrosion (Zhang et al. 2003). On the contrary, the lowest
capital cost was observed when an alkali catalyst was used in the production of
biodiesel from virgin vegetable oil (Zhang et al. 2003).

7.3.3 Operating Costs

Operating cost, also referred to as manufacturing cost, is the cost incurred during
the operation of the plant. These include costs of, but not limited to, raw materials,
catalysts, utilities, and labor. In contrast, as seen in this section, raw materials have
been stated separately as they account for a major proportion of the costs. In
addition, the cost of catalyst may be included within the cost of raw materials.
However, the different types of catalysts and their effect on the operating costs will
be discussed here.

7.3.3.1 Catalysts

Heterogeneous catalysts are cheaper than homogeneous catalysts (Kiss et al. 2010).
They have lower maintenance costs and are more profitable since they have a higher
yield. However, they have higher energy requirement as they operate at higher
temperatures (Kiss et al. 2010). On the other hand, alkaline catalysts are preferred
over acidic catalysts as they are less corrosive to equipment (Barnwal and Sharma
2005). Also, they require roughly eight times less alcohol to react with, as com-
pared to acidic catalysts (Zhang et al. 2003). Yet, twice as much amount of alkaline
catalyst is needed for production in comparison with acidic catalysts (Marchetti
et al. 2008). Additionally, the cost of biodiesel from waste oil using acidic and
alkaline catalysts were $0.8/L and $0.9/L, respectively, as found in a case study
(Karmee et al. 2015).

7.3.3.2 Utilities

The second most important type of operating cost is the cost associated with util-
ities. Several processes involved in biodiesel production are energy intensive, and
significant energy loss is attributed to them. With efficient reactor/heat transfer
design, cost of utilities can be reduced significantly (Ofori-Boateng et al. 2012).
According to an exergy analysis, the oil extraction process experiences the highest
exergy loss while transesterification suffers the least (Ofori-Boateng et al. 2012).
Nevertheless, using a heat-integrated design for equipment can aid in reducing
energy loss by up to 45% (Ofori-Boateng et al. 2012).
156 S. Taqvi et al.

The type of feedstock used can, also, help to reduce energy expended during the
cultivation process. For example, algae, used for biodiesel production, is known to
be energy independent as they can grow in a wide spectrum of climatic conditions
(Gao et al. 2012). Moreover, not much attention is needed to promote their growth
(Mata et al. 2010). In the case of catalysts, as stated in the previous section,
heterogeneous catalysts have a higher energy requirement than homogeneous cat-
alysts. Yet, they have a higher yield than homogeneous catalysts. A 1% increase
from 99 to 100% could lead to a 10% decrease in operating cost per metric ton
(Kiss et al. 2010). According to a study on a 100,000 mt biodiesel plant, this
decrease resulted in US$ 1 million worth of savings, annually (Kiss et al. 2010).

7.3.3.3 Labor

Labor charges (i.e., wages, salaries, etc.) are a significant part of the operating cost,
associated with a biodiesel plant. However, none of the above factors depicts any
significant effects on labor costs (Kiss et al. 2010). Nevertheless, as more
automation is introduced into the industry, with technological advancement over
time, there may be an expected decrease in labor costs. In contrast, if a particular
biodiesel plant is overstaffed, decreasing the labor cost will certainly reduce the
operating costs, thus making the project more economically favorable.

7.3.4 Credit Value

There are many items that may not have a significant effect on the production of
biodiesel. Also, they may not necessarily fall under the main scope of the analysis.
Nonetheless, they can be utilized in a manner to support in making the biodiesel
plant more economic. These may include credit values pertaining to CO2 emissions,
meal revenue, and by-products.

7.3.4.1 Carbon Dioxide (CO2)

Biofuel, in general, is known to be renewable and has a relatively much less


environmental impact than fossil fuels (Gao et al. 2012). A feasibility study, on
palm oil biodiesel transition from petro-diesel in Mexico, showed emissions
reduction of about 148 million tons of carbon dioxide (Lozada 2010). The emission
surplus can be sold to other industries and/or countries that have exceeded their
carbon emissions (Gao et al. 2012). Thus, accounting it for a credit and making it
more economic to produce biodiesel.
7 Techno-economical Aspects of Biodiesel Plants 157

7.3.4.2 Meal

After oil has been extracted from the biomass, it can be used as feed (i.e., meal
cake) for the livestock and/or as fertilizers (Mulugetta 2009). Several studies have
conducted sensitivity analyses, observing the impact of meal price on the economic
viability of biodiesel production (Mulugetta 2009; Bender 1999; Singhabhandhu
and Tezuka 2010; Kenkel and Holcomb 2008). In a study, conducted on a biodiesel
plant in Africa, it was found that palm oil was a better feedstock than jatropha as it
could be used for both, meal cake for animals and fertilizers (Mulugetta 2009).
Consequently, adding revenue generated from both products to make biodiesel
production more economically sustainable.

7.3.4.3 By-product

The main by-product of the transesterification of biomass is glycerin. Similar to


meal price, many studies have conducted sensitivity analyses on the market price of
glycerin (Kiss et al. 2010; Singhabhandhu and Tezuka 2010; Vlysidis et al. 2011;
Kumaran et al. 2011). Unlike crushed feedstock, glycerin has numerous applica-
tions and a wide market across the globe (Singhabhandhu and Tezuka 2010).
Glycerin is used in the production of personal/oral care products, medicine, food
and beverages, tobacco, polyether polyols, alkyd resins, and several other products
(Singhabhandhu and Tezuka 2010). Such a wide demand leads to the generation of
much larger revenue. Therefore, the amount of yield of glycerin and its purity plays
an important role in its selling price. With the help of sensitivity analysis, other
control variables may be optimized to operate in the most economic scenario.

7.4 Conclusions

The depletion of oil reserves and increase in GHG emissions with its associated
effects on global warming encouraged the search for alternatives to conventional
fossil fuel use in order to preserve energy security and achieve global emission
reduction targets. Biofuel, particularly biodiesel, is a renewable fuel, and one of the
feasible alternatives. In addition to biodiesel being nontoxic and biodegradable, it
has few better properties than petroleum-based diesel, such as improved lubrication
and burning properties. Even though the overall energy balance of biodiesel is
dependent on the type of feedstock utilized, most studies showed a positive overall
energy balance for biodiesel production. In addition, biodiesel has considerably
improved emission parameters in comparison to conventional diesel fuel.
The production of agricultural products for energy production purposes has
increased worldwide. Based on the analyses conducted from the review of studies
in the literature, and by considering the governmental policies (i.e., energy, agri-
cultural, etc.) associated with the assumptions incorporated in them, the following
158 S. Taqvi et al.

conclusions were made. Generally, most studies stated that it is very difficult to
sustain biodiesel production without government intervention through subsidies and
incentives that are required in order to facilitate the production and marketing of
biodiesel effectively. Biodiesel production is considered to be technically feasible;
however, its total cost is approximately 2–3.5 times higher than conventional
petroleum-based diesel. Moreover, projects are considered economically feasible if
the price of the biodiesel is kept high, taxes are increased on petroleum-based
diesel, and/or subsidies are provided to the biodiesel industry to make it more
competitive to petroleum-based diesel. Depending on the type of feedstock used,
biodiesel can be cost competitive with conventional petroleum-based diesel even
without government policy subsidies. However, this can be achieved through
economies of scale from large production capacities, which requires a large capital
investment. This is because the cost of raw materials is a critical factor in deter-
mining the cost of biodiesel manufacturing. Biodiesel is not cost competitive in
comparison with conventional diesel if the economic assumptions incorporated in
the price estimation mechanism do not include positive externalities and a monetary
quantification of environmental and socioeconomic impacts. Therefore, until
improvements and cost reductions in feedstock preparation and processing stages
are achieved, the cost competitiveness of biodiesel production will depend on
external monetary incentives. Though biodiesel is generally found to have a pos-
itive economic and environmental impact on societies, there are several socioeco-
nomic aspects that need to be addressed in order to improve the effective
penetration and development of biodiesel production technologies in the energy
production industry.

References

Albuquerque A, Danielski L, Stragevitch L (2016) Techno-economic assessment of an alternative


process for biodiesel production from feedstock containing high levels of free fatty acids.
Energy Fuels 30(11):9409–9418
Apostolakou A, Kookos I, Marazioti C, Angelopoulos K (2009) Techno-economic analysis of a
biodiesel production process from vegetable oils. Fuel Process Technol 90:1023–1031
Barnwal K, Sharma MP (2005) Prospects of biodiesel production from vegetable oils in India.
Renew Sustain Energy Rev 9:363–378
Bender M (1999) Economic feasibility review for community-scale farmer cooperatives for
biodiesel. Biores Technol 70:81–87
Enguidanos M, Soria A, Kavalov B, Jensen P (2002) Techno-economic analysis of bio-diesel
production in the EU: a short summary for decision-makers. European Commission
Fortenbery TR (2005) Biodiesel feasibility study: an evaluation of biodiesel feasibility in
Wisconsin. University of Wisconsin, Madison
Gao Y, Gregor C, Liang Y, Tang D, Tweed C (2012) Algae biodiesel—a feasibility report. Chem
Cent J 6(S1):1–16
Glystra C, Barlett S, Fox J (2016) World biodiesel production/consumption to rise 14% by 2020:
OECD/FAO. S&P Global Platts, 5 July 2016
Huang H, Long S, Singh V (2016) Techno-economic analysis of biodiesel and ethanol
co-production from lipid-producing sugarcane. Biofuels Bioprod Biorefin 10(3):299–315
7 Techno-economical Aspects of Biodiesel Plants 159

Karmee SK, Patria RD, Lin CSK (2015) Techno-economic evaluation of biodiesel production
from waste cooking oil—a case study of Hong Kong. Int J Mol Sci 16:4362–4371
Kenkel P, Holcomb R (2008) Feasibility of on-farm or small scale oilseed processing and biodiesel
production. In: Integration of agricultural and energy systems, Atlanta
Kiss FE, Jovanovic M, Boskovic GC (2010) Economic and ecological aspects of biodiesel
production over homogeneous and heterogeneous catalysts. Fuel Process Technol 91:1316–
1320
Koutinas AA, Chatzifragkou A, Kopsahelis N, Papanikolaou S, Kookos IK (2014) Design and
techno-economic evaluation of microbial oil production as a renewable resource for biodiesel
and oleochemical production. Fuel 116:566–577
Kumaran P, Mazlini N, Hussein I, Nazrain M, Khairul M (2011) Technical feasibility studies for
Langkawi WCO (waste cooking oil) derived-biodiesel. Energy 36:1386–1393
Lee K, Ofori-Boateng C (2013) Economic sustainability assessment of biofuels production from
oil palm biomass. Springer, Singapore
Lozada I, Islas J, Grande I (2010) Environmental and economic feasibility of palm oil biodiesel in
the Mexican transportation sector. Renew Sustain Energy Rev 14:486–492
Marchetti JM, Errazu AF (2008) Technoeconomic study of supercritical biodiesel production
plant. Energy Convers Manag 49(8):2160–2164
Marchetti J, Miguel V, Errazu A (2008) Techno-economic study of different alternatives for
biodiesel production. Fuel Process Technol 89:740–748
Mata TM, Marins AA, Caetano NS (2010) Microalgae for biodiesel production and other
applications: a review. Renew Sustain Energy Rev 14:217–232
Mulugetta Y (2009) Evaluating the economics of biodiesel in Africa. Renew Sustain Energy Rev
13:1592–1598
Nagarajan S, Chou SK, Cao S, Wu C, Zhou Z (2013) An updated comprehensive techno-economic
analysis of algae biodiesel. Biores Technol 145:150–156
Ofori-Boateng B, Keat TL, JitKang L (2012) Feasibility study of microalgal and jatropha biodiesel
production plants: exergy analysis approach. Appl Therm Eng 36:141–151
Singhabhandhu A, Tezuka T (2010) A perspective on incorporation of glycerin purification
process in biodiesel plants using waste cooking oil as feedstock. Energy 35:2493–2504
Skarlis S, Kondili E, Kaldellis J (2012) Small-scale biodiesel production economics: a case study
focus on Crete Island. J Clean Prod 20:20–26
Vlysidis A, Binns M, Webb C, Theodoropoulos C (2011) A techno-economic analysis of biodiesel
biorefineries: assessment of integrated designs for the co-production of fuels and chemicals.
Energy 36:4671–4683
Wu LF, Chen PC, Huang AP, Lee CM (2012) The feasibility of biodiesel production by
microalgae using industrial wastewater. Biores Technol 113:14–18
Zhang Y, Dube M, McLean D, Kates M (2003a) Biodiesel production from waste cooking oil: 1.
Process design and technological assessment. Biores Technol 89:1–6
Zhang Y, Dube M, McLean D, Kates M (2003b) Biodiesel production from waste cooking oil: 2.
Economic assessment and sensitivity analysis. Biores Technol 90:229–240
Chapter 8
Biodiesel Production and Consumption:
Life Cycle Assessment (LCA) Approach

Mohammad Ali Rajaeifar, Meisam Tabatabaei, Mortaza Aghbashlo,


Saeed Sadeghzadeh Hemayati and Reinout Heijungs

Abstract Like all energy carriers including renewable energies, the production to
combustion cycle of biodiesel should also be assessed from the sustainability point
of view. Life cycle assessment (LCA) is a promising approach capable of assisting
decision makers to find the environmental consequences of the existing or future
biodiesel production plans. For instance, for different feedstocks, production
technologies, downstream processes implemented, etc., an LCA of biodiesel pro-
duction cycles could result in different recommendations ranging from agricultural
practices to production and combustion stages. Despite the fact that an ISO standard
is available for conducting LCA studies, there are still many challenging issues
faced when performing LCA studies concerning biodiesel production and con-
sumption. These challenges include the functional unit, the choice of system
boundaries, the impact categories to be assessed, the treatment of land use change,

M. A. Rajaeifar (&)  M. Tabatabaei (&)


Biofuel Research Team (BRTeam), Karaj, Iran
e-mail: mohamad_rajaei@ut.ac.ir
M. Tabatabaei
e-mail: meisam_tab@yahoo.com; meisam_tabatabaei@abrii.ac.ir
M. Tabatabaei
Microbial Biotechnology Department, Agricultural Biotechnology Research Institute of Iran
(ABRII), Agricultural Research, Education and Extension Organization (AREEO), Karaj, Iran
M. Aghbashlo
Faculty of Agricultural Engineering and Technology, Department of Agricultural Machinery
Engineering, University College of Agriculture and Natural Resources, University of Tehran,
P.O. Box 4111, Tehran, Iran
S. S. Hemayati
Sugar Beet Seed Institute, Agricultural Research Education and Extension Organization
(AREEO), P.O. Box 31585-4114, Karaj, Iran
R. Heijungs
Department of Econometrics and Operations Research, Vrije Universiteit Amsterdam,
Amsterdam, The Netherlands
R. Heijungs
Institute of Environmental Sciences, Leiden University, Leiden, The Netherlands

© Springer Nature Switzerland AG 2019 161


M. Tabatabaei and M. Aghbashlo (eds.), Biodiesel,
Biofuel and Biorefinery Technologies 8,
https://doi.org/10.1007/978-3-030-00985-4_8
162 M. A. Rajaeifar et al.

and biogenic carbon. The present chapter provides a systematic overview of the
above-mentioned topics with the aim of shedding light on various aspects of LCA
of biodiesel production and consumption cycle.

8.1 Introduction

The modern world is heavily dependent on fossil fuels for satisfying its primary
needs, particularly, in the industrial and transportation sectors (Rajaeifar et al.
2017a, b). In fact, more than 80% of the current world’s energy consumption is
fossil-based and projections indicate a continuation of this trend till at least the year
2040 (Ashokkumar et al. 2017). Although concerns with respect to fossil fuel
depletion have been considered over time, the major challenge regarding the huge
consumption rate of fossil fuels is the environmental consequences caused by their
combustion. More specifically, air pollution and the subsequent risks for human
health and the environment on one hand and anthropogenic GHG emissions and
their subsequent global warming impacts on the other hand are among the most
grave challenges faced on a worldwide scale (Hosenuzzaman et al. 2015; Nicoletti
et al. 2015; Aghbashlo et al. 2017).
Alternative energy carriers such as biofuels have been widely considered as
replacement for fossil sources in order to address the above-mentioned challenges.
Biofuels offer numerous advantages including non-toxicity, biodegradability, better
emission profiles, renewability, domestic production in many countries, capability
to be used as transportation fuels, stimulating the agricultural sector and improving
its economic balance, creation of new job opportunities, and providing energy
security (Demirbas 2009; Wiloso and Heijungs 2013; Rajaeifar et al. 2016).
Nevertheless, there are controversial sides to biofuel production and consumption
as well which have been the subject of debates among the global scientific com-
munity. These controversies include (1) competition with agricultural food/feed/
fiber products and their impacts on the food/feed/fiber price and (2) direct and
indirect land use change impacts (Wiloso and Heijungs 2013) which could sig-
nificantly affect their GHG reduction benefits (Malça and Freire 2011). Moreover,
the relatively high cost of biofuels production has necessitated government supports
for their promotion (e.g., through subsidies, price guaranteed, lower taxes or tax
exemptions) (Rajaeifar et al. 2013).
Among different commercial biofuels, biodiesel is a promising alternative for
petroleum diesel and has recently attracted a huge deal of attention in the trans-
portation fleets around the world (Demirbas 2009; Jiaqiang et al. 2016). Biodiesel,
also known as mono-alkyl esters of different long chain fatty acids, is derived from
a variety of renewable lipid sources (Ghobadian et al. 2009). Possible feedstocks
used for biodiesel production are generally classified into three different groups, i.e.,
(1) first-generation feedstock (mainly edible oils), (2) second-generation feedstock
(mainly nonedible or waste oils), and (3) third-generation feedstock (mainly related
to algal biomass but to a certain extent linked to utilization of CO2 as feedstock
8 Biodiesel Production and Consumption … 163

(Lee and Lavoie 2013)). First-generation biodiesels are readily available and widely
used due to the fact that they can be produced from a wide range of feedstocks and
through well-developed production technologies. Nevertheless, their production
and development has triggered a debate on controversial competition with agri-
cultural food/feed/fiber products while it has also led to direct and indirect land use
change impacts.
The second-generation biodiesel fuels have been able to rectify the problems
associated with their first-generation counterparts, but they may also create an
indirect competition between the biodiesel industry and the industries in which
waste feedstocks are currently used. Moreover, nonedible and waste-oriented oil
feedstocks generally require several extra energy-intensive processes during feed-
stock preparation, which could also potentially increases indirect land use change
impacts (Singh et al. 2011). The third-generation biodiesels are assumed to be free
of such problems. However, several studies have shown that industrial-scale algal
cultivation also requires a considerable deal of nitrogen and phosphorous supple-
mentation used in form of fertilizers. This may seriously endanger the potential
advantages of the third-generation feedstocks since the upstream activities of fer-
tilizer production impose heavy burdens on the environment. For example, in
comparison to rapeseed biodiesel, biodiesel from microalgae needs 55–111 times
more nitrogen fertilizer—i.e., 8–16 tons/ha/year (Demirbas 2011). Such consider-
ations suggest that even when biodiesel would be environmentally superior during
combustion; it may have downsides during production. As such, a life cycle per-
spective is needed. In addition to that, given the free-fall of the prices of petroleum
products in response to the recent developments, e.g., emerging of the shale oil
extraction technology, the economic viability of algal biodiesel for short-term and
medium-term applications is also questionable.
Among the advantages of biodiesel is its environmentally friendly emission
profile compared with petroleum diesel, i.e., decreased emissions of CO, unburned
hydrocarbons (UHC), and particulate matter (PM), as well as decreased smoke
opacity (Kumara et al. 2009; Lee et al. 2011). Moreover, biodiesel contains no
sulfur and aromatic compounds in its chemical structure leading to a cleaner
combustion compared with its diesel counterpart. Nevertheless, it has been reported
that biodiesel generally increases tailpipe emissions of CO2 and NOx (Sheehan et al.
1998; Mohammadi et al. 2012).
In spite of all the mentioned benefits associated with biofuels utilization as an
alternative for fossil fuels, sustainability assessments criteria should still be taken
into account during decision-making and policy-making processes. Based on the
definition presented by the World Commission on Environment and Development,
the term ‘‘sustainable development’’ is defined as ‘‘development that meets the
needs of the present without compromising the ability of future generations to meet
their own needs’’(UNCED 1992). Accordingly, three important dimensions of
sustainability, namely social, environmental and economic form the backbone of
sustainability standards which must be considered in sustainability assessment of
any products or services as much as possible (Elkington 1997). Similarly, the
general principles of sustainable biofuel production and consumption could be
164 M. A. Rajaeifar et al.

easily defined but establishing a sound framework in order to efficiently charac-


terize these impacts is quite challenging due to the complicated interactions among
these three different dimensions (Singh et al. 2013).
Currently, environmental assessments—consisting of their very own frameworks
—are generally accompanied by the other types of sustainability assessments, e.g.,
social and economic assessments, or they are solely used for environmental sus-
tainability assessment purposes. Life cycle assessment (LCA) is one of the
assessment methods widely used for inspecting the environmental impacts of a
product/system (Guinée and Heijungs 2017) and is also the most widely used
technique for assessing the environmental balance in biofuel production and con-
sumption chains.
LCA is generally defined as a tool or approach that helps to assess the envi-
ronmental impacts of a product/service throughout its life cycle (Guinée 2002; Lin
et al. 2013). More specifically, it is capable of attributing the possible consequent
threats to the human health, natural ecosystems, and resources through different
damage assessment mechanisms. From the methodological point of view, LCA
deals with such questions by using a system approach, i.e., considering the product
(in this case biodiesel) as “a product system” or in better words, as “a function
system”. In fact, this approach considers the entire life cycle of a product/service,
from extraction of natural resources to the final waste management of the disposed
product, or so-called from “cradle to grave” (Guinée and Heijungs 2017). Of
course, a legitimate question may arise about the necessity of employing LCA in
some cases where the best scenario could be easily found by intuition. The answer
is that even when dealing with the simplest problems, the reality could be much
more complex and a systems approach is required to map the whole life cycle and
all potentially relevant environmental impacts (Guinée and Heijungs 2017). For
instance, it may be advocated that using electric vehicles is simply way better than
driving gasoline-driven vehicles from the environmental perspective because they
are “zero emission” vehicles. However, an LCA study showed that the results could
be heavily dependent on the source of electricity and/or consumer’s behavior
(Hawkins et al. 2013).
ISO standards–14040-46—proposed a standardized method for conducting LCA
studies, in which many criteria have been defined and guidelines have been pro-
posed ranging from basic issues, i.e., goal and scope definition, inventory analysis,
life cycle impact assessment, and interpretation to emerging crucial and compli-
cated problems, i.e., eco-efficiency assessment (ISO14045 2012) and water foot-
print (ISO14046 2014). In fact, the ISO series 14040 has been the most successful
attempt in harmonizing LCA studies to date. Nevertheless, there are still many
challenging issues faced when performing LCA studies in practice concerning
bioenergy feedstocks. This is due to the fact that such systems directly or indirectly
involve an agricultural stage which brings some complex and challenging issues in
estimating the real environmental impacts. Moreover, indirect inclusion of agri-
cultural stage implies more agricultural cultivation in other parts of the world, and
thus increases the uncertainty in the environmental impacts calculations. There are
also many other complicated factors coupled with an increased agricultural
8 Biodiesel Production and Consumption … 165

cultivation including economic, market, land occupation, and agricultural man-


agement issues which may add to the uncertainty level of the cycle. In addition to
that defining a functional unit, choosing of system boundaries, selecting impact
categories, and the treatment of land use change as well as biogenic carbon are the
most prominent technical and practical issues which should be tackled in order to
increase quality and usefulness of LCA results.
The present chapter provides a systematic overview of the above-mentioned
topics with the aim of shedding light on various aspects of LCA of biodiesel
production and consumption cycles. Section 8.2 discusses the general stages in life
cycle of different biodiesels including the main area in data collection and scenario
design in LCA studies. In Sect. 8.3, some issues and challenges faced in conducting
LCA of biodiesel production/consumption systems are comprehensively elaborated.
Finally, some recommendations to perform more accurate LCA studies on biodiesel
production/consumption systems are included in Sects. 8.2 and 8.3.

8.2 Biodiesel Life Cycle Stages: A Brief Description

Before conducting an LCA study on a given biofuel, it is very important to


understand and determine every stage of the life cycle. This could help to perform a
comprehensive LCA in form of “well-to-wheel” in case of biofuels. More specif-
ically, this would avoid overlooking a stage/substage in the life cycle and could also
help with detailed inventory data collection at the time of performing the project or
collecting data from databases. Neglecting a stage/substage in the life cycle causes
increased uncertainty, increased time and costs related to recalculating the neglected
stage/substage in the life cycle while also making the comparison of the results
incorrect or impossible. Overlooking one or some of the stages/substages involved
in the life cycle is an error observed in some studies on LCA of biodiesel (Rajaeifar
et al. 2017b). For example, there are studies in which the scope of the study did not
clearly define the inclusion of the combustion stage, while other studies failed to
define transportation of diesel and biodiesel from the production source to the point
of use, transportation of goods (input materials) to the agricultural farms, etc.
Therefore, all stages involved in the life cycle of a given biodiesel must be
determined before conducting an LCA study and those stages must be clearly
mentioned through the scope of the study and illustrated in the proposed system
boundary.
Stages involved in the life cycle of biodiesel may be defined based on the
feedstock used for biodiesel production, i.e., from the first- to third-generations.
Figures 8.1, 8.2, and 8.3 briefly show the possible stages involved in the life cycle
of these three biodiesel generations. It should be mentioned that different tech-
nologies may include more or fewer substages, but the general scheme of the stages
involved in different biodiesel generations is similar to the ones presented in these
figures. Based on Fig. 8.1, LCA of biodiesel production/consumption using
first-generation feedstock generally encompasses the following main stages:
166 M. A. Rajaeifar et al.

Fig. 8.1 Simplified flow diagram of the well-to-wheel processes involved in the first-generation
biodiesels’ life cycle

Fig. 8.2 Simplified flow diagram of the well-to-wheel processes involved in the
second-generation biodiesel life cycle (a waste oils and b animal fat)

Fig. 8.3 Simplified flow diagram of the well-to-wheel processes involved in the third-generation
biodiesel life cycle
8 Biodiesel Production and Consumption … 167

agricultural cultivation, transportation, oil milling (oil extraction), as well as bio-


diesel production and combustion. The agricultural cultivation stage could further
be divided into the following substages: upstream activities for the production of
agricultural inputs (e.g., seed, fertilizers, pesticides, fuels, etc.), production of
capital goods (e.g., agricultural buildings and machinery), and fieldwork operations
(e.g., land preparation, planting of seed/seedlings, fertilizing, tillage, harvesting,
etc.). It should be mentioned that there is a difference between annual and perennial
crops in the agricultural cultivation substages. In other words, perennial crops
generally need pre-nursery, nursery, and immature plantation (or two of these)
substages before annual plantation activities. These substages may take place
during several years and must be included in the assessment. In this regard, the
consumption of agricultural inputs, emissions originated from upstream activities
for the production of these inputs as well as emissions originated from the
above-mentioned pre-cultivation substages should also be considered. More
specifically, all the agricultural activities (from pre-nursery activities to the agri-
cultural actives in each cultivation season) should be taken into account throughout
the lifetime of a crop (e.g., 25–28 years for palm oil trees and 25–30 years for olive
trees) and the inventory for a cultivation year should be obtained as the average of
these years. This approach has been well employed by many researchers (Schmidt
2007; Choo et al. 2011; Van Zutphen and Wijbrans 2012; Rajaeifar et al. 2016)
while there are also a number of studies in which these substages were left out.
Since the second-generation biodiesel feedstocks are generally considered to be
waste, useless or low price fat/oils, analyses do not include the agricultural culti-
vation stage, and thus no environmental burdens are carried from their first life
(Fig. 8.2a, b). It should be noted that in the case of nonedible oil feedstock
specifically cultivated for biodiesel production, their agricultural cultivation stage
should also be included. As for animal fats (Fig. 8.2b), the upstream activities
related to animal husbandry and slaughterhouse are generally excluded since fat is
usually traded at far lower prices in comparison with meat (i.e., an increase in
demands for animal fats would not serve as a motivation for meat producers to
increase their meat production) and moreover, a proportion of animal fats is gen-
erally subjected to disposal in many parts of the world. This approach has been used
by many reports published previously (Dufour and Iribarren 2012; Jørgensen et al.
2012; Escobar et al. 2014; Rajaeifar et al. 2017b). For the third-generation bio-
diesels (Fig. 8.3), the agricultural cultivation stage includes the cultivation of algae
in pounds. In such systems, a pre-cultivation stage, i.e., the cultivation of an algal
strain in photo-bioreactors/indoor ponds to be used as inoculum (seed culture) for
the open ponds, should also be considered (Sander and Murthy 2010). It should also
be noted that dewatering is a different stage, which must be inventoried separately.
In the transportation stage, all the relevant transportation activities are included
and the inventory data generally include the consumption of materials by vehicles
(from fuels to engine oils and filters), production of capital goods, and construction
of the infrastructure. However, the calculation of the last two items is difficult and
could bring about uncertainties since vehicles or roads used could, in general, have
other applications rather than being solely used for biodiesel transportation. Based
168 M. A. Rajaeifar et al.

on Fig. 8.1, transportation of the agricultural outputs to the oil mill plant, trans-
portation of the extracted oil to the biodiesel production plant, and transportation of
biodiesel to the point of use are the main substages which are generally included in
the assessment. It is also worth quoting that transportation of agricultural inputs to
the farms as well as transportation of input materials to the oil mill and biodiesel
production plants are generally included in their related stages rather than in the
transportation stage (Escobar et al. 2014; Rajaeifar et al. 2014). The transportation
substages for the third-generation feedstocks are the same as those of the
first-generation ones (Fig. 8.3). However, for the second-generation biodiesels
(Fig. 8.2a, b), there is no agricultural output transportation and instead, collection
and transportation of feedstock from the point of generation to biodiesel production
plants (or in the case of animal fats, the transportation of feedstock to rendering
plants and then to biodiesel production plants) are considered. Nevertheless, if a
consequential approach is employed, agricultural stage and its relevant substages
might also be potentially included for the second-generation biodiesels as well.
Overall, a more detailed inventory data for transportation in a life cycle could be
helpful in further optimizing the transportation distances based on the final results
and the environmental hotspots found in the life cycle.
In the oil mill stage, the agricultural output is converted into oil and meal. There
are many oil extraction methods such as cold pressing, pressing and extraction by
organic solvents, microwave or ultrasound-assisted methods (Moreno et al. 2003;
Shah et al. 2005; Rajaeifar et al. 2013) with their own pros and cons. It should be
mentioned that most modern oil extraction technologies are based on lowering the
volume of wastewater produced in the oil milling process, while more efficient
methods for treating the generated wastewater are also in development (Hodaifa
et al. 2013; Lim et al. 2014; Liew et al. 2015; Yu et al. 2017). From the LCA point
of view, the oil mill stage could further be divided into the following substages:
upstream activities for the production of input materials needed for oil milling (e.g.,
chemicals, fuels, electricity, etc.), production of capital goods, and oil mill plant
operations (e.g., oil extraction, wastewater treatment, meal drying, etc.). This is also
applicable to the third-generation feedstocks (Fig. 8.3). As mentioned earlier,
transportation of the input materials to the oil mill plant is generally included in this
stage as well. For the second-generation feedstocks, there is generally no oil mill
stage included unless in the case of animal fats where further rendering is required
(Fig. 8.2b).
The oil extracted in oil mill stage is transported for further conversion into
biodiesel in the biodiesel production stage. The type of the conversion technology
used for a dedicated biofuel may have a significant impact on its life cycle emis-
sions, at a lower magnitude in comparison with feedstock production (i.e., agri-
cultural cultivation) stage though (Wiloso and Heijungs 2013; Altamirano et al.
2016). Among the different methods used for biodiesel production, transesterifi-
cation has been considered by far as the best method (Baskar and Aiswarya 2016)
and is the most prominent technology used at commercial scale as well (Stojković
et al. 2014). More specifically, transesterification is the reaction of triglyceride
molecules present in fat or oils with an alcohol resulting in the formation of
8 Biodiesel Production and Consumption … 169

mono-alkyl esters (biodiesel) and glycerol (Ma and Hanna 1999). The conventional
transesterification reaction is mainly highlighted by heating and stirring the reaction
mixture (to stimulate a quick contact between reagents), consumption of a high
amount of energy for heating and stirring, relatively high temperature (i.e., slightly
below methanol boiling point of 65 °C), and using homogeneous/heterogeneous
acid or base catalysts (Sáez-Bastante et al. 2015; Rajaeifar et al. 2017a). However,
different attempts have been made with an aim of introducing new techniques in
order to further enhance conventional biodiesel production from different aspects of
energy consumption, time, biodiesel conversion efficiency, wastewater generation,
and production costs. Some of these techniques include nanocatalytic technology,
ultrasound-assisted, microwave-assisted, in situ transesterification, supercritical
(catalytic or non-catalytic), subcritical, and membrane-assisted techniques
(Georgogianni et al. 2008; Motasemi and Ani 2012; Sáez-Bastante et al. 2015;
Rajaeifar et al. 2017a; Tran et al. 2017). The biodiesel production stage can further
be divided into substages, such as upstream activities for the production of input
materials needed for biodiesel production (e.g., chemicals, electricity, etc.), pro-
duction of capital goods and biodiesel production plant operations (e.g., biodiesel
production, biodiesel refining, wastewater treatment, etc.). This is also applicable to
the second- and third-generation feedstocks. Similar to the oil mill stage, trans-
portation of the input materials to the biodiesel production plant is generally
included in this stage as well.
The combustion stage is the final stage in a ‘well-to-wheel’ life cycle of bio-
diesel in which tailpipe emissions from stationary or mobile engines running on
biodiesel are measured. The required inventory data on tailpipe emissions can be
collected through laboratory chassis dynamometer tests (steady-state operation also
known as bench-scale examination) or real-world tests. Laboratory chassis
dynamometer tests are commonly performed based on standard driving cycles at
considerably less costs and experimental burdens while real-world tests need rig-
orous operational considerations and impose higher costs for monitoring emissions
when the vehicle is in motion. The combustion stage only considers tailpipe
emissions, and thus has no further substages. It is also worth quoting that in case of
using biodiesel–diesel blends or additives in biodiesel, the upstream activities
related to diesel or additive production and transportation must also be inventoried
separately (Xue et al. 2012; Rajaeifar et al. 2017b). Moreover, there is no difference
between the different generation feedstock in preparing the inventory data for this
stage.
170 M. A. Rajaeifar et al.

8.3 LCA and Biodiesel Production/Consumption Systems:


Some Issues and Challenges

At the first glance, applying LCA in biodiesel production/consumption systems


seems like the other products or services in which LCA could be practically
applied. In another word, the choices within the four main phases of LCA (i.e., goal
and scope definition, inventory analysis, impact assessment, and interpretation)
seem clear and apparently easy to be made, similar to LCAs of other products.
However, a more in-depth look would reveal that in practice, conducting an LCA of
biodiesel production/consumption systems is more difficult and complicated due to
data variability (mainly in the agriculture, biodiesel production, and combustion
stages) as well as the additional challenges and uncertainties in the currently used
methodological approaches. More specifically, such systems mainly involve an
agricultural stage, or they would imply more agricultural cultivation in other parts
of the world, thus introducing challenges and complexities in the calculation
methods as well as increasing the level of uncertainty in the environmental impacts
calculations. The other reason is that biodiesel development is coupled with many
other complicated factors originated from changes in demand and supply chains in
the market (locally or globally) or agricultural land occupation. Moreover, there are
also technical and practical issues in the methodology of LCA; some of which are
still the subjects of ongoing discussions among academics (i.e., functional unit, the
choice of system boundaries, the impact categories to be assessed, the treatment of
land use change, and biogenic carbon). These issues and challenges have also
caused a wide range of outcomes even for apparently similar biofuel life cycles. The
present section provides a systematic overview of the above-mentioned topics and
challenges.

8.3.1 Goal and Scope Definition

The goal and scope definition is a very important initial step in every LCA study.
This is due to the fact that goal and scope definition is the starting point of a
research work which could directly affect many choices used throughout the course
of the study. Choosing attributional or consequential approach alongside choosing
the system boundaries, functional unit and dealing with multifunctional processes,
as well as the types of required inventory data are among the methodological
choices which follow the goal and scope definition (Wiloso and Heijungs 2013).
The goal of an LCA study determines the context of the study, its intended
application, and targeted audience while the scope definition outlines the type of
methodology to be used in the subsequent modeling (Baumann and Tillman 2004;
Wolf et al. 2010).
It should also be noted that a well-defined scope and boundaries are essential for
guaranteeing a well-defined goal (Curran 2017). Therefore, the following six
8 Biodiesel Production and Consumption … 171

aspects are recommended to be addressed and documented during the goal defi-
nition process (Wolf et al. 2010):
• The intended application(s)
• The reasons for carrying out the study
• Limitations regarding the method, assumptions, and impact categories used
• The intended audience
• Whether the results are to be used in comparative assertions and planned to be
disclosed to the public
• The commissioner of the study and other influential actors.
Scope definition embraces a set of major choices which must be clearly
explained through the course of each LCA study, i.e., studied system or process and
its function, system boundaries, functional unit, modeling approach (consequential
or attributional), as well as the reference system or flow to be used. Moreover,
scope definition should lead to the determination of the following issues: the type of
required inventory data and the data quality requirements, life cycle inventory
(LCI) modeling framework, treatment of multifunctional processes and products,
impact categories to be covered, selection of life cycle impact assessment (LCIA)
method and if included—data normalization and weighting factors, and treatment of
uncertainties (Wolf et al. 2010; Heijungs and Wiloso 2014). Moreover, neglecting,
removing or merging stages/substages must be clearly explained in the scope
definition as it could be misleading when interpreting the results. It is worth quoting
that unlike what is suggested by the ISO standard on goal and scope definition, no
concrete details on system boundaries, impact categories, and treatment of uncer-
tainty are allowed to be implemented at this stage (Heijungs and Wiloso 2014). In
better words, such details should be collected and analyzed in the inventory phase
and the impact assessment phase of the study, respectively, and not in the goal and
scope definition.

8.3.1.1 System Boundaries

As one of the most important aspects of the goal and scope definition, system
boundaries influence data collection, background data choices, and foreground
modeling aspects (Baitz 2017). System boundaries must be well designed in a way
that correctly present an overall perspective of the life cycle stages involved, main
relevant (unit) processes/flows, main input(s)/output(s), excluded activities (stages/
substages), as well as included and excluded emissions from different flows (e.g.,
tailpipe emissions, wastewater emissions, emission to air/water/soil). Based on the
ISO definitions, system boundaries simultaneously separate the analyzed system
from the rest of the technosphere as well as the ecosphere (Wolf et al. 2010). Other
dimensions beside technical aspects need to be specified clearly as well, i.e.,
geographic (spatial) and time (temporal) boundaries (Curran 2017). It is also worth
quoting that excluding any stages/substages/processes/emissions during an LCA is
172 M. A. Rajaeifar et al.

only permitted if they are estimated to have no significant impacts on the overall
conclusions of the study. Examples are construction of capital goods in some cases,
human labor, some internal transportation of materials within production facilities,
manufacture and transport of packaging materials which are not associated with the
final product, maintenance and operation of support equipment, identical stage/
substage/process/emissions in comparative LCA studies, or in cases where cutoff
may be the only solution (i.e., when the system is theoretically infinitely large). The
brief description presented in the previous section could help in arranging conve-
nient system boundaries in LCA of different biodiesel feedstocks and prevent
arbitrary system boundaries definition. A word of warning should also be added
regarding the importance of the inclusion of a reference system when defining and
describing the system boundaries of a study. This generally applies when the goal
of an LCA study encompasses a comparison between the main system under
investigation and the other systems of comparison value.

8.3.1.2 Functional Unit

The product or process being studied through LCA is described and quantified
through a functional unit (FU) specified in relation to the nature of a system,
geographical, and time boundaries. In other words, the functional unit is a quan-
tified description of the performance of a product system (Weidema et al. 2004). An
appropriate functional unit is the one that positively reflects the reality of the
problem. This, in fact, could be achieved when the FU is driven by the main
questions or goals of the LCA study. Choosing a proper FU is very important in
LCA studies since different choices of functional units from the same system may
lead to different results when compared to each other (Wiloso and Heijungs 2013).
For instance, comparing two types of paint on a per liter basis may yield a different
preference compared to comparing the same paints on a per square meter basis.
For LCA of biodiesel systems, the most common FU used in the studies are
generally classified into the following four groups (Cherubini and Strømman 2011):
Input-oriented FUs: these types of FUs describe the performance of a system based
on input biomass (either in mass or energy unit) and are appropriate to show the
best uses for a given biomass feedstock. Examples of such FU applications are 1 kg
corn produced, 1 kg or barrel of waste cooking oil collected and transported, and,
etc.
Output-oriented FUs: calculating and evaluating the performance of a system
based on the unit of output delivered is performed through these types of FUs. It is
regularly reported that output-oriented FUs are the most common type in LCA of
bioenergy and seems the best option for these systems (Cherubini and Strømman
2011) unless the system delivers multi-outputs and need an allocation procedure
like what generally happens in biorefineries. Such FUs generally show the per-
formance of a given biodiesel based on the calorific value of biodiesel (MJ, kWh),
8 Biodiesel Production and Consumption … 173

mass or volume of biodiesel produced (kg or L), or driving distance of a vehicle (in
km) fueled by a given biodiesel blend.
Agricultural land use oriented FUs: here the evaluation of a biodiesel system is
performed based on the hectare of land area required to produce the biodiesel
feedstocks. This type of FU is convenient for the first-generation feedstocks.
Although the application of this type of FU was rarely reported in the literature, it
could lead to driving helpful results at policy level since the biomass could bring a
competition in land occupation with food/feed/fiber products. Moreover, this type
of FU directly shows the efficiency of agricultural management in a dedicated
occupied agricultural land. Cherubini and Strømman (2011) remarked that relative
land use efficiency (i.e. the use of scarce land resources as efficiently as possible) is
found using this type of FU. An example of using agricultural land use as an FU
could be found in Lim and Lee (2011) study in which 1-year use of one-hectare
palm oil plantation was considered as a FU to produce both biodiesel and
bioethanol.
Time-oriented FUs: these types of FUs refer to a period of activity performed by a
system, e.g., yearly, monthly or based on a season activity.
Input- and time-oriented FUs as well as the ones based on the unit of agricultural
land use do not facilitate a proper comparison between biofuels and their fossil
counterparts. Probably, for this reason, output-oriented FUs are the most common
type used in the LCA of bioenergy systems (Cherubini and Strømman 2011). This
also applies to LCA of biodiesels, for which results based on the volume of bio-
diesel produced (or combusted) (in liters) seems more perceptible for the public as
they see and understand what they finally pay for. This is also the case for policy
makers as they are generally offered reports in which consumption of fossil fuels as
well as projected substitutions by alternative fuels is presented on a volume basis
(i.e., in liters). Nevertheless, it has been reported that when the best use of a given
biomass feedstock as bioenergy (heat, electricity, biofuel) is the main question of
the study, functional units in the form of one MJ or kWh are more appropriate
(Wiloso and Heijungs 2013). Based on literature studies, there are two different
perceptions of FU as a unit of energy: (1) the calorific value of biodiesel in forms of
MJ and (2) MJ or kWh of useful energy. It is worth quoting that only the second
perspective lead to find the best use of a given biomass feedstock as bioenergy since
they consider the efficiency of different systems, e.g., efficiency of diesel engines in
power plants alongside conversion and transmission losses, while the first per-
spective only help to find the best technologies, treatment methods or feedstock for
biodiesel production.
For LCA studies in which the combustion stage is included or comparing bio-
diesel and petroleum diesel for transportation is the main question, FUs in the form
of distance traveled (in km or miles) by vehicles is more appropriate. However,
when measuring tailpipe emissions, the working conditions must be completely
similar in order to be able to compare systems based on this type of FU. Moreover,
all the experimental details regarding the vehicle traveling must also be reported,
e.g., the type of vehicle (s) used along with their model and age, vehicle speed,
174 M. A. Rajaeifar et al.

passenger load (number), and route specifications including the length and grade of
the road used in the experiments. Overall, in order to enhance understanding of the
system under study and avoid misleading conclusions, using several functional
units could be more helpful. This important issue is commonly neglected by the
methodological standards for bioenergy systems (Cherubini and Strømman 2011).

8.3.1.3 Attributional and Consequential LCA

In developing LCA methodologies, the distinction between attributional LCA


(ALCA) and consequential LCA (CLCA) should be taken into consideration. The
specification of the type of LCA used should be firstly shown in goal and scope
definition step. This could further influence methodological and data choices for the
LCI and LCIA as subsequent steps. The goal of an ALCA study is to assess the
environmental burdens attributed to a product/service assuming the current situation
(of technology, market, economy, and supply chains) or so-called ‘a status quo
situation’(Wiloso and Heijungs 2013). This approach describes the environmentally
relevant physical flows to and from a life cycle and its subsystems (Ekvall and
Weidema 2004). A complete set of procedures and recommendation for a clear goal
and scope definition when ALCA is applied were introduced by Martin (2017). It
should be highlighted that the static nature of ALCA would not permit this type of
LCA to be the central core in decision-making processes for policy makers espe-
cially in case of biofuels such as biodiesel and bioethanol. This is due to the fact
that this approach does not have the capability of showing the possible changes in
environmental impacts regarding possible choices, especially indirect land use
change (iLUC) impacts originated from biofuel development. However, ALCA is
useful in highlighting the environmental hotspots of the current production systems
and determining the differences between feedstocks, production processes, and
efficiencies with respect to the overall environmental burdens (Wiloso and Heijungs
2013).
CLCA, on the other hand, expands the system boundaries of an attributional
approach so as to embrace possible external consequences in response to possible
decisions and changes, and consequently estimates their effects on environmental
flows (resource use and emissions) of a given product/service (Finnveden et al.
2009). Therefore, CLCA is in principle more effective and attractive for strategic
planning when biodiesel and bioethanol development is the main concern. This
approach mainly relies on additional economic data like marginal production costs,
elasticity of supply and demand (Ibenholt 2002). The CLCA methodology differs
from ALCA not only in goal and scope definition, but also in system boundaries,
FU, LCI, and treatment of multifunctional processes (Thomassen et al. 2008). More
importantly, CLCA encompasses the indirect effects especially ones related to the
land use changes and it employs marginal data while ALCA does not include the
indirect effects and it uses average data. This specification has a twofold structure,
one which shows CLCA as a more comprehensive method with its advantages. The
other aspect shows the complexity of modeling the indirect effects and thus
8 Biodiesel Production and Consumption … 175

increased uncertainty introduced by this method which may cause to remove the
advantages came from this approach. More specifically, the results made by CLCA
are strongly sensitive to the assumptions employed by the modeling. Therefore, all
the assumptions should be kept tracked rigorously and should be clearly identified
in the final assessment report (Prox and Curran 2017). Figure 8.4 shows two dif-
ferent approaches employed for assessing the environmental impacts of using waste
cooking oil (WCO) and poultry fat (PF) based biodiesel blends in urban buses in
Iran (Rajaeifar et al. 2017b). Based on the figure, when considering ALCA of WCO
and PF biodiesel blends, feedstock collection, transportation, biodiesel production
and combustion stages are considered inside the system boundaries of the study. In
such situation, all the relevant (background and foreground) data need to be col-
lected as average data considering the current technologies while changes in
demands and supply chains in the market as a result of biodiesel development are
not considered.
By contrast, considering a CLCA of WCO and PF biodiesel blends (Fig. 8.4),
the indirect effects of using these fuels are included in addition to the mentioned
stages. More specifically, when using WCO and PF for biodiesel production, their
demand in the market is bound to increase and consequently (in the most likely
situation) their previous users should find alternative oils or so-called ‘marginal
oils’. Accordingly, removed WCO is compensated by palm oil while removed PF is
substituted by a mixture of palm and soybean oils (Jørgensen et al. 2012; Rajaeifar

Fig. 8.4 Attributional and consequential approaches in assessing the environmental impacts of
using waste cooking oil (WCO)- and poultry fat (PF)-based biodiesel blends in urban buses in Iran.
With Permission from Rajaeifar et al. (2017b)
176 M. A. Rajaeifar et al.

et al. 2017b). In this regard, Malaysia was identified as the marginal producer of
palm oil in the global market while Argentina was identified as the marginal
supplier of soybean oil to Iran. Therefore, the compensation of removed WCO oil
and rendered PF (for biodiesel production) from the market implies more agricul-
tural cultivation in these countries. Consequently, this leads to an increase in the
production of some coproducts (i.e., palm kernel meal in Malaysia and soybean
meal in Argentina). Increased production of such coproducts has also an indirect
effect, i.e., decreased production of their marginal products. Since Brazil was
identified as the marginal producer of soybean meal in the global market, the
increased agricultural cultivation and consequent increase in the production of palm
kernel meal and soybean meal leads to a decrease in the production of soybean meal
(as well as its soybean oil) in Brazil. In this LCA approach, the indirect land use
change impacts are also included and the marginal (background and foreground)
data are collected considering the futuristic technologies.
Overall, while the main challenges in performing ALCA for biofuel systems
(including biodiesels) focus on allocation procedures, the main challenges when
using CLCA approach are generally attributed to quantifying the indirect effects of
developing biofuels on the other cycles, i.e., food/feed/fiber (Wiloso and Heijungs
2013). This mainly includes quantifying/modeling the iLUC impacts as one of the
major challenges faced by regulators when making specific choices among various
biofuel alternatives s (Plevin et al. 2015). This issue has been neglected by many
research studies on LCA of biofuels while taking into account such impacts could
undermine the benefits attributed to the substitution of biofuels with their fossil
counterparts (Searchinger et al. 2008; Zamagni et al. 2012; Ben Aoun and Gabrielle
2016). Likewise, unresolved debates in direct/indirect impacts from agricultural
cultivation as well as allocation of coproducts have caused problems leading no
clear distinction between ALCA and CLCA in most regionalized policy guidelines
(Brander et al. 2008; van Dam et al. 2010). It seems that improving the global
economic interaction models as well as developing the methods for more accurately
calculating land use change emissions are the master keys for these challenges.

8.3.2 Inventory Analysis

LCI is the phase of collecting and quantifying inputs and outputs (as a flow model
including all the emissions) inside the defined system boundary of a product
throughout its life cycle. Accordingly, the quality of collected data as well as
methods for quantifying emissions are the main concerns in the inventory analysis
(Heijungs and Wiloso 2014). Therefore, data sources, quality, and their collection
procedure as well as methodology applied to calculate emissions must be clearly
and unambiguously presented in this stage from the starting point of a given life
cycle to its end. Below, some typical LCI-related problems faced in biodiesel
studies are discussed.
8 Biodiesel Production and Consumption … 177

8.3.2.1 Agricultural Field Emissions

One of the most important steps in the LCI of biodiesel systems is the estimation of
agricultural field emissions. The stage may be involved as the main stage like what
is usually happened in first-generation feedstocks or may be indirectly involved in
such agricultural activities happened as an indirect effect of some second-generation
biodiesels. In the case of third-generation feedstock, agricultural cultivation stage
mainly includes the cultivation of algae in pounds which is also an important but
less complex issue.
The agricultural stage has been identified as the stage which introduces a lot of
complexity and uncertainty in LCA of biofuel systems (Wiloso and Heijungs 2013).
The complexity is the result of interactions between soil–water–air and chemical/
organic nutrients. The uncertainty encompasses a wider concept related to indirect
land use change impacts while also in the simplest form (without considering any
indirect effect), the variability in soil structure and climate as well as in agricultural
practice management scenarios could lead to substantial variations in LCA results.
In order to calculate field crop emissions, the nutrient balance must be identified
by focusing on the most important nutrient inputs to and outputs from a farm. The
main cycles which generally must be considered are carbon and nitrogen as well as
the other elements such as phosphorous and potassium. Moreover, there are some
undesirable inputs such as heavy metals which must be inventoried as well. The
main activities responsible for the field crop emissions (or so-called ‘on site
emissions’) are applying fertilizers/pesticides and land transformation. It is also
worth quoting that soil and climate conditions affect the level of emissions as well.
Overall, the following emissions are generally considered (Harris et al. 2015;
Nemecek et al. 2016; Khoshnevisan et al. 2017):
1. Carbon dioxide emissions (to air) due to urea application (if any urea fertilizer
is used).
2. Carbon dioxide emissions (to air) due to changes in carbon pools as a result of
land transformation, occupation and restoration activities.
3. N2O and CH4 emissions (to air) as a result of land use change.
4. NH3 and NOx emissions (to air) as well as direct and indirect N2O emissions (to
air) due to the application of N-based fertilizers.
5. Nitrate leaching (to groundwater) due to the application of N-based fertilizers.
6. Nitrate run-off (to surface water) due to the application of N-based fertilizers.
7. Phosphorus leaching (to groundwater) due to the application of P-based
fertilizers.
8. Phosphate run-off (to surface water) due to the application of P-based
fertilizers.
9. Heavy metal emissions (to agricultural soil) and groundwater due to the
application of N, P, and K-based fertilizers.
10. Tailpipe emissions (to air) from diesel/gasoline combustion during farm
operations.
178 M. A. Rajaeifar et al.

11. Emissions of pesticides (to agricultural soil).


There are three approaches for determining the field crop emissions, i.e.,
(1) measuring actual emission rates from the studied system, (2) applying emission
values derived from literature in a case-by-case procedure, and (3) estimating
potential emission rates through structured estimation methods (Brentrup et al.
2000). Using the first approach in order to determine field crop emissions is money-
and time-consuming, and not applicable for many types of emissions. Moreover,
field measurements generally show great variations and generally are representative
of specific conditions at the time of measurement which is not necessarily appro-
priate for LCA purposes. A new update of emission values is often required when
using emission values from the literature (the second approach) while the quality of
the values is questioned (Brentrup et al. 2000).
The third approach, i.e., using structured estimation methods in order to estimate
emission rates, is the most employed approach in LCA studies in which field crop
emissions are considered. The term ‘structured’ refers to considering different
impacts of soil condition, climate and agricultural practices alongside the dosage of
a given nutrient as input. These types of methods can easily be employed with less
effort, cost and uncertainty compared to direct measurements or values derived from
the literature. Moreover, since the estimation methods simplify the complex con-
ditions using structured frameworks, the quality of the estimated emission rates can
be updated and improved in time. There are many types of models and guidelines
for estimating the field crop emissions, some of them only model limited number of
emissions, while others model most of emissions. Some examples are IPCC (2006),
Nemecek et al. (2016), Birkved and Hauschild (2006), etc.
While methods for estimating field crop emissions are being further developed
consistently, becoming more accurate on a daily basis, many challenges still remain
in calculating agricultural field emissions mainly due to the (1) data gaps in
long-term soil quality dynamics, (2) lack of LCIs on many active ingredients of
pesticides, (3) lack of understanding of the pesticides’ emissions mechanism when
used in field operations, (4) inclusion of N2O and CH4 emissions from land use
change activities, (5) modeling the emissions of NO3 based on different
pedo-climatic conditions and different management options as well as (6) adopting
data related to the periods before and after agriculture in the assessments.

8.3.2.2 Land Use and Land Use Change

The issue of land use and land use change has become a very critical and important
issue since it has been identified as a great contributor to global GHG emissions
(Watson et al. 2000). The term land use is generally used for land transformation
(period before agricultural activities), land occupation (period during agricultural
activities), and land restoration (period after agricultural activities). Although there
is no exact place for land use and land use change in the LCA framework, most of
the studies and guidelines consider land transformation and land restoration
8 Biodiesel Production and Consumption … 179

activities as land use changes (Wiloso and Heijungs 2013), while land occupation is
generally considered as land use activities. The main issue is not about the use of
these terms instead of each other, but rather quantifying the impact of land use
change using a sound and accurate framework is the current challenge in LCA
studies. More specifically, despite the existing consensus on the inclusion of land
use change impacts when assessing the environmental impacts of biofuel devel-
opment, the resulting indicators suffer from a considerable deal of heterogeneity,
significant inaccuracies, as well as high uncertainties (Cherubini and Strømman
2011; Fritz et al. 2013).
In the case of first-generation biodiesels (as well as the other first-generation
biofuels), the increase of biodiesel usage will lead to an increase in the demand for
the feedstock used for biodiesel production, with a subsequent feedstock shortage,
and thus increased market prices of the feedstocks used. This situation would
motivate those active in the agriculture industry, e.g., farmers for increasing their
outputs through different mechanisms, i.e., (1) intensifying crop management
systems to improve yields, (2) transforming uncultivated lands into agricultural
land, and (3) substituting food/feed/fiber crops by energy crops (Ben Aoun and
Gabrielle 2016). The second and third approaches could trigger land use change
mechanisms either in direct or indirect ways. It is also worth quoting that the
development of second-generation biodiesels could also result in such land use
change impacts when the feedstock used for biodiesel production, e.g., PF and
WCO, would also be used as raw materials in other industries. In such cases, the
removed WCO or PF must be compensated by producing and importing oil crops
elsewhere, presumably mainly fulfilled by the second and third mechanisms.
Direct Land Use Change (dLUC)
In biodiesel production/consumption systems, direct land use change (dLUC for
short) takes place when new agricultural land is assigned to the production of
feedstock for biodiesel production and the feedstock displaces a prior land use (e.g.,
conversion of degraded tropical rainforest or alang-alang grass lands to oil palm
plantation). This substitution in land use may include the situation in which feed-
stock for biodiesel displaces other crops (used for food/feed/fiber purposes) in a
cropping system. These situations as dLUC may cause changes in the carbon stock
of the assigned land and result in global warming impact. The carbon stock gen-
erally exists in five different pools, i.e., above ground vegetation, below ground
vegetation, dead wood, litter, and (most importantly) soil (Cherubini and Strømman
2011).
Depending on which type of land is transformed for the production of biodiesel
feedstock, the net carbon emissions due to dLUC can be positive or negative. More
specifically, when land with a high carbon stock (e.g., natural forest areas, pasture,
and peatlands) is converted to agricultural land, a loss of carbon stocks may affect
the whole carbon balance, strongly undermining the environmental performance of
the given biodiesel compared with its fossil counterpart (Reijnders and Huijbregts
2008; Panichelli et al. 2009). Moreover, these types of transformation would lead to
a decrease in biodiversity (Salaa et al. 2009). On the contrary, if biodiesel
180 M. A. Rajaeifar et al.

feedstocks are grown on set-aside or degraded lands or when perennial crops (e.g.,
oil palm) replace annual row crops, dLUC can contribute to increases in the carbon
stock, and thus has a positive effect on the GHG balance of the biodiesel system
(Styles and Jones 2007; Wu et al. 2008).
It is also worth quoting that the changes in carbon stocks are highly dependent
on the previous and current agricultural practices, post-harvest activities, climate,
and soil characteristics (Cherubini and Strømman 2011), and thus they require
site-specific quantification. Nevertheless, methods have been developed to gener-
alize and estimate changes in carbon pools by means of literature references, default
values, or software tools capable of modeling soil carbon dynamics.
Indirect Land Use Change (iLUC)
Indirect land use change (iLUC for short) occurs when developments in biodiesel
production cause changes in land use elsewhere. This situation generally happens
when feedstock production for biodiesel takes place on lands currently used for
food/feed/fiber crops and the demand for the former land use (i.e., food, feed, fiber)
remains. Therefore, the removed agricultural production will relocate to other
places in order to maintain the balance in the global market and prevent competi-
tions among biodiesel and food/feed/fiber production domains (Gnansounou et al.
2008). Clearly, market mechanisms are the core of agricultural cultivation dis-
placement. The existing challenges in quantifying iLUC impacts is mainly due to
the complexity and speculative nature of the mechanisms involved in land use
change. Such shortcomings could result in (1) lack of consensus on using one
method for estimating iLUC impacts, (2) variability in iLUCs results, and
(3) uncertainty of the LCA conclusions.
Although increased pressure on land worldwide, and thus the creation of iLUC
impacts has been mainly associated with the development of first-generation
feedstock (Ben Aoun and Gabrielle 2016), the development of second-generation
feedstock may also bring about such effects when the feedstock is already used by
other sectors rather than biofuels sector. For example, when the development of
WCO and PF biodiesel is considered (Fig. 8.4), demands for WCO and rendered
PF in the market are bound to increase and consequently (under the most possible
circumstances) their previous users will inevitably try to find alternative sources of
oil. Accordingly, the removed WCO could be compensated for instance, by palm
oil while a mixture of palm and soybean oils could substitute removed PF. This
situation necessitates more agricultural production elsewhere which causes iLUC
impacts (Rajaeifar et al. 2017b). If the cultivation of feedstock for biodiesel pro-
duction takes place on fallow, marginal, or degraded lands where no conventional
crops are grown, no iLUC happens and the GHG balance can turn out to be more
favorable (Cherubini and Strømman 2011).
There are three main approaches in dealing with iLUC effects including (1) use
of historical data and statistical analysis, (2) using experts’ opinion, and (3) ap-
plying economic equilibrium models. The first approach employs historical data
from different sources and statistically analyzes them in order to identify possible
relationships between the rate of feedstock production (for biofuels including
8 Biodiesel Production and Consumption … 181

biodiesel) in a given country and land use change (Ben Aoun and Gabrielle 2016).
There are number of studies which employed this approach for estimating land use
change impacts (Kim and Dale 2011; Overmars et al. 2011). In the second
approach, the experts are supposed to have an understanding of the underlying
market mechanisms in order to chase the possible iLUCs location and quality and
predict their magnitude. This generally happens through estimating cause-effect
relations in the market as well as through the simplification of market mechanisms
(Bauen et al. 2010; Ben Aoun et al. 2013). Examples of employing such approaches
are Dalgaard et al. (2008), Schmidt (2010), Reinhard and Zah (2011), Escobar et al.
(2014) and Rajaeifar et al. (2017b).
Compared with the first and second approaches, the third approach could be
more accurate and effective in estimating iLUC impacts since it is capable of
modeling economic processes in the market. In fact, when using historical data or
expert-based opinion approaches, market mechanisms are simplified and some
other activities that can lead to land use change may not be considered. Therefore,
the prediction of iLUC might not be accurate enough (Ben Aoun et al. 2013). The
economic equilibrium models are based on the theory of perfect markets in econ-
omy. Accordingly, the response of supply and demand to price changes creates an
equilibrium in which demand equals supply. This, in fact, forms the basis of the
estimation for the iLUC impacts. The economic equilibrium models generally
include two types of equilibrium models, i.e., partial and general equilibrium
models which have been comprehensively explained by Ben Aoun and Gabrielle
(2016).
Despite the existence of various methods for estimating iLUC impacts, there is a
global interest in using economic models for estimating land use change impacts.
This is ascribed to the fact that these models could generally determine the con-
sequences of additional demands created by the development of biodiesel
production/consumption systems on global land use. Nevertheless, economic
models still need to be substantially improved in order to be more accurate and
reliable. This could be achieved by making the estimation of markets response to
price changes and producers/consumers preferences more realistic and less uncer-
tain. Improving the quality of different databases concerning carbon stocks changes,
fertilizer use, and gaseous emissions of nitrogen could also enhance the accuracy of
the models used and reduce their uncertainty and variations, generally occurring
when using different methods for estimating land use change impacts.
Apart from land use change attributed to biodiesel production, water footprint
should also be taken into consideration. This is further highlighted given the
anticipated water scarcity challenge and the resultant conflicts among countries in
the near future. Therefore, a word of warning should also be added regarding the
importance of the water availability and water footprint assessments in future LCA
frameworks and guidelines. In another word, even if the challenges concerning the
quantifying iLUC impacts will be resolved; the water availability as well as water
footprint of biodiesel production/consumption systems will still be a limiting factor
for future development of this alternative fuel. In this regard, water resources and
availability estimations as well as water footprint assessment must also be included
182 M. A. Rajaeifar et al.

in decision-making for anticipating the possible impacts attributed to the indirect


land use effect of biodiesel development.

8.3.2.3 Conversion Processes and Combustion Emissions

Biodiesel production/consumption systems include conversion processes through


which the raw feedstock for biodiesel production is converted into biodiesel.
Inventory data at this stage mainly include the upstream activities for the production
of energy and materials used in the conversion process, emissions from the pro-
duction of capital goods as well as emissions arisen from the conversion reaction
itself. Data gaps and uncertainties in the conversion process are mainly attributed to
developing technological routes, particularly, on an industrial scale. In this regard,
the real impacts of the current production technology or the considered marginal
technology could be under- or overestimated and therefore, a sensitivity analysis
would be necessary.
As mentioned before, the combustion stage is the final stage in a ‘well-to-wheel’
life cycle of biodiesel in which tailpipe emissions from vehicles using biodiesel are
measured. More specifically, the required inventory data on tailpipe emissions can
be collected through laboratory chassis dynamometer tests or real-world tests. The
most important issue when collecting inventory data for this stage is to ensure that
the data would be representative of the real-world conditions. In line with that, the
driving cycles chosen for laboratory chassis dynamometer must be selected or
designed carefully according to the real-world conditions of the area/routes/fleet at
which biodiesel is intended to be used. In case of real-world tests, it is recom-
mended to perform a set of real-driving tests considering statistical criteria (e.g.,
using statistical designs), ambient factors (e.g., temperature, humidity, and pas-
senger counts), and vehicle factors (engine model year, engine type, mileage
traveled before and during the experiments, engine oil type and viscosity).
Moreover, it is suggested to select various routes and various traffic situations in
different weekdays for performing real-world driving tests. These would increase
the validity of and options for extrapolating the results, especially when assessing
tailpipe emissions in a transportation fleet is considered.

8.3.2.4 Biogenic Carbon

Like other biofuel systems, biodiesel production/consumption systems can absorb


and sequester CO2 from the atmosphere. More specifically, agricultural crops are
capable of fixing atmospheric CO2 during their growth period, while the absorbed C
is released when the crops/resultant products are subjected to the combustion
process. Nevertheless, this is not a sufficient reason for assuming biodiesel
production/consumption systems as carbon neutral ones, i.e., assuming all emis-
sions arisen from biodiesel combustion as biogenic. In fact, such systems require a
significant amount of fossil inputs whose consumption (or production) increase the
8 Biodiesel Production and Consumption … 183

atmospheric CO2 level. These fossil inputs are generally consumed during planting,
fertilization, harvesting, oil extraction, transportation, as well as biodiesel produc-
tion and combustion. For example, large amounts of N-based fertilizers are con-
sumed annually for agricultural fertilization and the production of these fertilizers
impose a significant deal of environmental burden, i.e., GHG emissions released to
the atmosphere.
As for biodiesel combustion, it is generally assumed by many studies that all the
CO2 emissions arisen from biodiesel combustion have a biogenic nature. However,
from the chemical point of view, biodiesel is made through a reaction between an
alcohol and triglycerides, and thus the C atoms of the alcohol used also contribute
to the resultant methyl esters. Therefore, if the alcohol such as methanol (the
common alcohol in biodiesel production) used in the biodiesel production stage is
of fossil origin, not all the CO2 emissions from the combustion of biodiesel could be
regarded as biogenic. Therefore, it is important to distinguish the emissions asso-
ciated with the biogenic and non-biogenic carbon moieties; this is not a facile job
though. In better words, it is suggested to calculate their partitioning among all
major carbon-based tailpipe emissions such as CO2, CO, PM and non-methane
hydrocarbons (NMHC) (Sheehan et al. 1998). It is important to take into account all
carbon-based tailpipe emissions due to the fact that under real-world conditions,
engines do not completely combust all the carbon in the fuel, and thus the whole
carbon contained in a given fuel is not combusted as CO2. It is also worth men-
tioning that the carbon emission components other than CO2, could not be con-
sidered as biogenic since they will not be absorbed throughout the cycle, i.e., over
the plant’s growth period.
Another challenging issue in biogenic carbon cycles occurs when the agricultural
production stage results in coproducts (e.g., palm oil and palm kernel oil). Under
such circumstances, a part of the absorbed CO2 is allocated to each of the
coproducts (Wiloso and Heijungs 2013). Using an appropriate allocation method is
the center of current debate among academics and more research is needed in order
to reach a global consensus. The final challenge regarding the biogenic nature of
biodiesel is attributed to a time difference between CO2 fixation and release.
Development of dynamic LCA methods could help to account for such situations.
Although there is no consensus regarding how to treat biogenic carbon, the most
important issue is to avoid double counting of CO2 emissions. To achieve that,
(1) the inclusion or exclusion of carbon sequestration, (2) the inclusion or exclusion
of biogenic carbon, and (3) calculating the share of biomass-oriented carbon in final
CO2 released by biodiesel combustion must be explicitly stated in the inventory
analysis of a study.

8.3.2.5 Multifunctional Unit Processes

The issue of multifunctional unit processes occurs when a unit process yields more
than one functional flow. Multifunctional unit processes can take place under three
main conditions, i.e., coproduction, combined waste processing, and recycling
184 M. A. Rajaeifar et al.

(Wiloso and Heijungs 2013). Under such conditions, it is generally preferred to


determine the environmental burdens of a given product among the others. An
example in biodiesel production/consumption systems is biodiesel production from
palm oil in which it is important to determine the environmental burdens attributed
to palm oil and distinguishing it from the environmental burdens related to palm
kernel oil. The ISO standard (ISO14044 2006) describes multifunctionality options
rather not in the goal and scope definition, but in the later stage of the Inventory
analysis. This avoids predefining or dictating a preferred option and instead could
help with considering the relevancy and adequacy of the options quantitatively in
the inventory phase (Baitz 2017).
In order to deal with multifunctional unit processes, there are three approaches
available as suggested by existing guidelines but in different order of priority, i.e.,
subdivision, system expansion (including substitution), and allocation (or parti-
tioning). Based on the ISO standard, it is preferred to avoid burden allocation
mainly using the first two methods. Nevertheless, using subdivision requires precise
separation between the main processes considering the related mass and energy
balances. Furthermore, using system expansion or product displacement could be
more difficult and uncertain, as it requires comprehensive investigation of the whole
market of all the output products, including so-called “avoided” processes.
A major challenge in the third option, i.e., allocation, is determining the criteria
needed for attributing emissions, waste, and upstream inputs to various coproducts.
The criteria often used include ratios of mass, energy, and economic value. There
are also studies which tried to avoid allocation through choosing an appropriate FU,
e.g., input-oriented FUs (Cherubini and Strømman 2011) as well as the two options
elaborated earlier. A detailed discussion of the possible allocation methods, with
their advantages and disadvantages, can be found in the literature (Ekvall and
Finnveden 2001; Curran 2007; Heijungs and Guinée 2007).

8.3.3 Impact Assessment

LCIA is the third phase in conducting a life cycle study. In LCIA, the inventory
data are reflected in the impact categories (Wolf et al. 2010). This phase consists of
a number of activities including the selection of impact category, classification,
characterization, normalization, grouping, weighting, and data quality analysis.
Among these activities, the first three are mandatory, while the rest are optional
(ISO14044 2006). The most important fact about impact assessment is to properly
choose a set of relevant impact categories to measure the potential environmental
burdens of a given biodiesel production/consumption system in all environmental
dimensions, i.e., human health, ecosystem quality, and resources. In this regard, the
set of chosen impact categories must be comprehensive as much as possible.
LCA studies on biodiesel production/consumption systems generally fall into
three types of impact assessment, i.e., energy input–output analysis, global
warming, and other life cycle impact categories (Cherubini and Strømman 2011).
8 Biodiesel Production and Consumption … 185

Energy input–output analysis or energy analysis is aimed at quantifying the effi-


ciency of a given renewable energy production (i.e., biodiesel) and determines the
possible nonrenewable energy savings through biodiesel production. The most
important index in these types of studies is fossil energy ratio (FER) which shows
the actual benefit obtained from a biodiesel production/consumption system by
taking into account the amount of fossil resources consumed in the life cycle of
biodiesel production/consumption. Nevertheless, it should be mentioned that the
energy input–output analysis mostly shows the technical feasibility of the biodiesel
production/consumption systems rather than being an impact assessment method in
principle (Wiloso and Heijungs 2013). However, its respective results could be later
analyzed and interpreted from the environmental point of view. The transportation
distances and methods as well as the type of allocation method used to allocate
energy flow between coproducts (if required) are the most important issues which
could significantly affect the results of energy analysis studies and, therefore, these
must be performed accurately, including a sensitivity analysis.
Global warming is one of the most common indices used in LCA of biodiesels
by which a list of GHG emissions arisen from all the processes involved in the life
cycle are collected and then translated into CO2 equivalents. Although global
warming is included in most of LCA studies performed on biodiesel production/
consumption, the issue of biogenic carbon as well as land use change still remains
as the main challenge, mostly neglected in many research studies as elaborated
previously. These issues could significantly affect the global warming contribution
of a given biodiesel production/consumption system and turn the results from
favorable to unfavorable, or the other way around. Other life cycle impact cate-
gories include a variety of impact categories, e.g., eutrophication, acidification,
aquatic ecotoxicity, terrestrial ecotoxicity, carcinogens, noncarcinogens, respiratory
organics/inorganics, etc., which have been used in a number of LCA studies of
biodiesel (Panichelli et al. 2009; Xue et al. 2012; Rajaeifar et al. 2016; Rajaeifar
et al. 2017a; Sousa et al. 2017).
Overall, a common weakness of most studies conducted on biodiesel LCA is
attributed to the lack of sufficient coverage of impact categories. In fact, failure to
address key impact categories may bring about incomplete or unreliable informa-
tion, creating biased decisions. Therefore, it has been suggested to choose a set of
impact categories in agreement with the goal and scope of the study, while con-
sidering a default minimum in order to reduce the risk of biased decisions (Wiloso
and Heijungs 2013). Moreover, due to the fact that water scarcity challenge is
expected to introduce more serious troubles in the near future, water footprint must
also be included as an important impact category in future LCA studies on biodiesel
production. The most important issue regarding the current impact categories in
LCA studies is the fact that all the current impact categories consider the potential
impact or maximum possible impact only. In this regard, future attempts could also
be devoted to including the exposure effects of different emissions in LCA studies
since a system with higher level of pollution but in low population (humans or
biodiversity) areas may perform environmentally better compared with a system
with lower level of environmental emissions but at more populated areas.
186 M. A. Rajaeifar et al.

8.3.3.1 Regionalized Impact Assessment

Regionalization in LCA studies is generally performed for inventories and impact


assessment methods. Collection of activity data as well as region-specific back-
ground data is required in order to accurately show all the relevant processes
involved and to enable the application of geographically explicit impact assessment
models (Morais et al. 2016). In view of the impact assessment methods used, there
are some environmental impacts which could have different effects on the human
health/ecosystem quality depending on the characteristics of the receiving envi-
ronment (i.e., the location of the activity). The variations in the resulting impacts of
such impact categories generally relate to the characteristics of both the emitting
source and the receiving environment (Finnveden et al. 2009).
When the system boundary of a study includes an agricultural stage, the use of
regionalized impact assessment methods could be of importance since the impacts
analyzed are clearly site-dependent. Therefore, impact assessment methods must
meet some criteria which clearly reflect the regional conditions of the agricultural
system included in a biodiesel production/consumption system under investigation.
Regionalization is not of concern for global impact categories such as global
warming or stratospheric ozone depletion since these impacts are independent of
where the emissions occur. Instead, there are some impact categories which are
often regional or even local in nature and thus, they need to be set on regional scale.
Although regionalization in LCA studies is rarely performed in practice (Mutel and
Hellweg 2009), recent developments in regionalization have focused on enhancing
characterization methods for regional impact categories in order to be able to apply
them in a more consistent way in different regions all over the world (Hauschild and
Huijbregts 2015) and to further improve the accuracy of LCIA methods. Among the
attempts aimed at the regionalization of impacts is the implementation of impact
assessment relating to water use (Boulay et al. 2011; Verones et al. 2013) and land
use (Elshout et al. 2014), as well as acidification and terrestrial eutrophication
(Seppälä et al. 2006). Besides, attempts have also been made in developing
site-dependent characterization for LCIA which are appropriate for processes in
Europe, the USA, and some other certain countries (Finnveden et al. 2009).
The main aim in regionalizing an impact category must be devoted to addressing
how the impacts will manifest themselves on local or regional scales. To achieve
this goal, accurate assessments must be performed in order to provide a process
under investigation with spatial variability so that it can be applied on all geo-
graphical scales. Moreover, in order to obtain more representative impact estimates
in relation to location-dependent impacts, the inherent differences associated with
variability in soil types and complex interactions with local climates must be
considered as well (Wiloso and Heijungs 2013).
8 Biodiesel Production and Consumption … 187

8.4 Conclusions

Biodiesel has attracted a great deal of attention throughout the world as a promising
substitute for petroleum diesel. From an LCA point of view, biodiesel production/
consumption systems face similar challenges as other biofuel systems. This is due
to the fact that such systems directly or indirectly involve an agricultural stage
which brings about several complex and challenging issues in estimating the real
environmental impacts. In addition to that, there is a necessity of reaching a con-
sensus on methods used for selecting a functional unit, setting system boundaries,
selecting impact categories, allocating multifunctional processes as well as quan-
tifying land use changes and biogenic carbon. Most importantly, future policy
decisions should be focused on the best use of fertile lands for climate change
mitigation as well as the best use of water for producing goods in the agriculture
sector. Therefore, the quantification of the land use change impacts as well as water
footprint originated from the development of feedstock for biodiesel are very
important issues for the development of more efficient environmentally friendly
biodiesel production/consumption systems in the future. As a consequence, estab-
lishing a sound and scientific framework in order to efficiently characterize land use
change impacts and water use assessments in future LCA frameworks is of prime
importance for future LCA studies.

References

Aghbashlo M, Tabatabaei M, Mohammadi P, Khoshnevisan B, Rajaeifar MA, Pakzad M (2017)


Neat diesel beats waste-oriented biodiesel from the exergoeconomic and exergoenvironmental
point of views. Energy Convers Manag 148:1–15
Altamirano CAA, Yokoyama L, de Medeiros JL, Araújo OdQF (2016) Ethylic or methylic route to
soybean biodiesel? Tracking environmental answers through life cycle assessment. Appl
Energy 184:1246–1263
Ashokkumar V, Salim MR, Salam Z, Sivakumar P, Chong CT, Elumalai S, Suresh V, Ani FN
(2017) Production of liquid biofuels (biodiesel and bioethanol) from brown marine macroalgae
Padina tetrastromatica. Energy Convers Manag 135:351–361
Baitz M (2017) Attributional life cycle assessment. In: Curran M (ed) Goal and scope definition in
life cycle assessment. Springer, Dordrecht. The complete world of life cycle assessment.
Springer, Dordrecht
Baskar G, Aiswarya R (2016) Trends in catalytic production of biodiesel from various feedstocks.
Renew Sustain Energy Rev 57:496–504
Bauen A, Chudziak C, Vad K, Watson P (2010) A causal descriptive approach to modelling the
GHG emissions associated with the indirect land use impacts of biofuels. E4Tech, London
Baumann H, Tillman A-M (2004) The Hitch Hiker’s guide to LCA. An orientation in life cycle
assessment methodology and application. External Organization
Ben Aoun W, Gabrielle B (2016) Life cycle assessment and land-use changes: effectiveness and
limitations. In: Gnansounou E, Pandey A (eds) Life-cycle assessment of biorefineries. Elsevier
Ben Aoun W, Gabrielle B, Gagnepain B (2013) The importance of land use change in the
environmental balance of biofuels. In: Oilseeds and fats, crops and lipids, vol 1
188 M. A. Rajaeifar et al.

Birkved M, Hauschild MZ (2006) PestLCI—a model for estimating field emissions of pesticides in
agricultural LCA. Ecol Model 198:433–451
Boulay A-M, Bulle C, Bayart J-B, Deschênes L, Margni M (2011) Regional characterization of
freshwater use in LCA: modeling direct impacts on human health. Environ Sci Technol
45:8948–8957
Brander M, Tipper R, Hutchison C, Davis G (2008) Technical paper: consequential and
attributional approaches to LCA: a guide to policy makers with specific reference to
greenhouse gas LCA of biofuels. Econometrica Press
Brentrup F, Küsters J, Lammel J, Kuhlmann H (2000) Methods to estimate on-field nitrogen
emissions from crop production as an input to LCA studies in the agricultural sector. Int J Life
Cycle Assess 5:349–357
Cherubini F, Strømman AH (2011) Life cycle assessment of bioenergy systems: state of the art and
future challenges. Biores Technol 102:437–451
Choo YM, Muhamad H, Hashim Z, Subramaniam V, Puah CW, Tan Y (2011) Determination of
GHG contributions by subsystems in the oil palm supply chain using the LCA approach. Int J
Life Cycle Assess 16:669–681
Curran MA (2007) Co-product and input allocation approaches for creating life cycle inventory
data: a literature review. Int J Life Cycle Assess 12:65–78
Curran MA (2017) Overview of goal and scope definition in life cycle assessment. Goal and scope
definition in life cycle assessment. In: LCA compendium—The complete world of life cycle
assessment. Springer, pp 1–62
Dalgaard R, Schmidt J, Halberg N, Christensen P, Thrane M, Pengue WA (2008) LCA of soybean
meal. Int J Life Cycle Assess 13:240
Demirbas A (2009) Biodiesel from waste cooking oil via base-catalytic and supercritical methanol
transesterification. Energy Convers Manag 50:923–927
Demirbas MF (2011) Biofuels from algae for sustainable development. Appl Energy 88:3473–
3480
Dufour J, Iribarren D (2012) Life cycle assessment of biodiesel production from free fatty acid-rich
wastes. Renew Energy 38:155–162
Ekvall T, Finnveden G (2001) Allocation in ISO 14041—a critical review. J Clean Prod 9:197–
208
Ekvall T, Weidema BP (2004) System boundaries and input data in consequential life cycle
inventory analysis. Int J Life Cycle Assess 9:161–171
Elkington J (1997) Cannibals with forks. In: The triple bottom line of 21st century, vol 73
Elshout PM, van Zelm R, Karuppiah R, Laurenzi IJ, Huijbregts MA (2014) A spatially explicit
data-driven approach to assess the effect of agricultural land occupation on species groups. Int J
Life Cycle Assess 19:758–769
Escobar N, Ribal J, Clemente G, Sanjuán N (2014) Consequential LCA of two alternative systems
for biodiesel consumption in Spain, considering uncertainty. J Clean Prod 79:61–73
Finnveden G, Hauschild MZ, Ekvall T, Guinée J, Heijungs R, Hellweg S, Koehler A,
Pennington D, Suh S (2009) Recent developments in life cycle assessment. J Environ Manage
91:1–21
Fritz S, See L, Valin H (2013) Current issues and uncertainties in estimating global land
availability for biofuel production. Taylor & Francis
Georgogianni K, Kontominas M, Pomonis P, Avlonitis D, Gergis V (2008) Conventional and
in situ transesterification of sunflower seed oil for the production of biodiesel. Fuel Process
Technol 89:503–509
Ghobadian B, Rahimi H, Nikbakht A, Najafi G, Yusaf T (2009) Diesel engine performance and
exhaust emission analysis using waste cooking biodiesel fuel with an artificial neural network.
Renew Energy 34:976–982
Gnansounou E, Panichelli L, Dauriat A, Villegas JD (2008) Accounting for indirect land-use
changes in GHG balances of biofuels: review of current approaches
Guinée J, Heijungs R (2017) Introduction to life cycle assessment. In: Sustainable supply chains.
Springer, pp 15–41
8 Biodiesel Production and Consumption … 189

Guinée JB (2002) Handbook on life cycle assessment operational guide to the ISO standards. Int J
Life Cycle Assess 7:311
Harris Z, Spake R, Taylor G (2015) Land use change to bioenergy: a meta-analysis of soil carbon
and GHG emissions. Biomass Bioenergy 82:27–39
Hauschild MZ, Huijbregts MAJ (2015) Introducing life cycle impact assessment. In: Life cycle
impact assessment. Springer, pp 1–16
Hawkins TR, Singh B, Majeau-Bettez G, Strømman AH (2013) Comparative environmental life
cycle assessment of conventional and electric vehicles. J Ind Ecol 17:53–64
Heijungs R, Guinée JB (2007) Allocation and ‘what-if’scenarios in life cycle assessment of waste
management systems. Waste Manag 27:997–1005
Heijungs R, Wiloso EI (2014) Life cycle assessment of bioenergy systems. In: Wang L
(ed) Sustainable bioenergy production, pp 99–114
Hodaifa G, Ochando-Pulido J, Rodriguez-Vives S, Martinez-Ferez A (2013) Optimization of
continuous reactor at pilot scale for olive-oil mill wastewater treatment by Fenton-like process.
Chem Eng J 220:117–124
Hosenuzzaman M, Rahim N, Selvaraj J, Hasanuzzaman M, Malek A, Nahar A (2015) Global
prospects, progress, policies, and environmental impact of solar photovoltaic power generation.
Renew Sustain Energy Rev 41:284–297
Ibenholt K (2002) Materials flow analysis and economic modelling. In: Handbook of industrial
ecology. Edward Elgar, Cheltenham, pp 177–184
IPCC (2006) Guidelines for national greenhouse gas inventories, Agriculture, forestry and other
land use, vol. 4. Intergovernmental Panel on Climate Change
ISO14044 (2006) Environmental management. Life cycle assessment requirements and guidelines.
ISO 14044 International Standard International Organization for Standardization
ISO14045 (2012) Environmental management—eco-efficiency assessment of product systems—
principles, requirements and guidelines. ISO 14045 International Standard International
Organization for Standardization
ISO14046 (2014). Environmental management—water footprint—principles, requirements and
guidelines. ISO 14046 International Standard International Organization for Standardization
Jiaqiang E, Liu T, Yang W, Li J, Gong J, Deng Y (2016) Effects of fatty acid methyl esters
proportion on combustion and emission characteristics of a biodiesel fueled diesel engine.
Energy Convers Manag 117:410–419
Jørgensen A, Bikker P, Herrmann IT (2012) Assessing the greenhouse gas emissions from poultry
fat biodiesel. J Clean Prod 24:85–91
Khoshnevisan B, Rafiee S, Tabatabaei M, Ghanavati H, Mohtasebi SS, Rahimi V, Shafiei M,
Angelidaki I, Karimi K (2017) Life cycle assessment of castor-based biorefinery: a well to
wheel LCA. Int J Life Cycle Assess 1–18
Kim S, Dale BE (2011) Indirect land use change for biofuels: testing predictions and improving
analytical methodologies. Biomass Bioenergy 35:3235–3240
Kumara AS, Maheswarb D, Reddyc KVK (2009) Comparison of diesel engine performance and
emissions from neat and transesterified cotton seed oil. Jordan J Mech Ind Eng 3
Lee RA, Lavoie J-M (2013) From first-to third-generation biofuels: challenges of producing a
commodity from a biomass of increasing complexity. Anim Front 3:6–11
Lee W-J, Liu Y-C, Mwangi FK, Chen W-H, Lin S-L, Fukushima Y, Liao C-N, Wang L-C (2011)
Assessment of energy performance and air pollutant emissions in a diesel engine generator
fueled with water-containing ethanol–biodiesel–diesel blend of fuels. Energy 36:5591–5599
Liew WL, Kassim MA, Muda K, Loh SK, Affam AC (2015) Conventional methods and emerging
wastewater polishing technologies for palm oil mill effluent treatment: a review. J Environ
Manag 149:222–235
Lim S, Lee KT (2011) Parallel production of biodiesel and bioethanol in palm-oil-based
biorefineries: life cycle assessment on the energy and greenhouse gases emissions. Biofuels
Bioprod Biorefin 5:132–150
190 M. A. Rajaeifar et al.

Lim SL, Wu TY, Clarke C (2014) Treatment and biotransformation of highly polluted
agro-industrial wastewater from a palm oil mill into vermicompost using earthworms. J Agric
Food Chem 62:691–698
Lin J, Babbitt CW, Trabold TA (2013) Life cycle assessment integrated with thermodynamic
analysis of bio-fuel options for solid oxide fuel cells. Biores Technol 128:495–504
Ma F, Hanna MA (1999) Biodiesel production: a review. Biores Technol 70:1–15
Malça J, Freire F (2011) Life-cycle studies of biodiesel in Europe: a review addressing the
variability of results and modeling issues. Renew Sustain Energy Rev 15:338–351
Martin B (2017) Attributional life cycle assessment. In: Curran MA (ed) Goal and scope definition
in life cycle assessment. LCA compendium—The complete world of life cycle assessment.
Springer, pp 123–143
Mohammadi P, Nikbakht AM, Tabatabaei M, Farhadi K, Mohebbi A (2012) Experimental
investigation of performance and emission characteristics of DI diesel engine fueled with
polymer waste dissolved in biodiesel-blended diesel fuel. Energy 46:596–605
Morais TG, Teixeira RF, Domingos T (2016) Regionalization of agri-food life cycle assessment: a
review of studies in Portugal and recommendations for the future. Int J Life Cycle Assess
21:875–884
Moreno AO, Dorantes L, Galíndez J, Guzmán RI (2003) Effect of different extraction methods on
fatty acids, volatile compounds, and physical and chemical properties of avocado (Persea
americana Mill.) oil. J Agric Food Chem 51:2216–2221
Motasemi F, Ani FN (2012) A review on microwave-assisted production of biodiesel. Renew
Sustain Energy Rev 16:4719–4733
Mutel CL, Hellweg S (2009) Regionalized life cycle assessment: computational methodology and
application to inventory databases. Environ Sci Technol 43:5797–5803
Nemecek T, Schnetzer J, Reinhard J (2016) Updated and harmonised greenhouse gas emissions for
crop inventories. Int J Life Cycle Assess 21:1361–1378
Nicoletti G, Arcuri N, Nicoletti G, Bruno R (2015) A technical and environmental comparison
between hydrogen and some fossil fuels. Energy Convers Manag 89:205–213
Overmars KP, Stehfest E, Ros JP, Prins AG (2011) Indirect land use change emissions related to
EU biofuel consumption: an analysis based on historical data. Environ Sci Policy 14:248–257
Panichelli L, Dauriat A, Gnansounou E (2009) Life cycle assessment of soybean-based biodiesel
in Argentina for export. Int J Life Cycle Assess 14:144–159
Plevin RJ, Beckman J, Golub AA, Witcover J, O’Hare M (2015) Carbon accounting and economic
model uncertainty of emissions from biofuels-induced land use change. Environ Sci Technol
49:2656–2664
Prox M, Curran MA (2017) Consequential life cycle assessment. In: Curran MA (ed) Goal and
scope definition in life cycle assessment. LCA compendium—The complete world of life cycle
assessment. Springer, pp 145–160
Rajaeifar MA, Abdi R, Tabatabaei M (2017a) Expanded polystyrene waste application for
improving biodiesel environmental performance parameters from life cycle assessment point of
view. Renew Sustain Energy Rev 74:278–298
Rajaeifar MA, Akram A, Ghobadian B, Rafiee S, Heidari MD (2014) Energy-economic life cycle
assessment (LCA) and greenhouse gas emissions analysis of olive oil production in Iran.
Energy 66:139–149
Rajaeifar MA, Akram A, Ghobadian B, Rafiee S, Heijungs R, Tabatabaei M (2016) Environmental
impact assessment of olive pomace oil biodiesel production and consumption: a comparative
lifecycle assessment. Energy 106:87–102
Rajaeifar MA, Ghobadian B, Davoud Heidari M, Fayyazi E (2013) Energy consumption and
greenhouse gas emissions of biodiesel production from rapeseed in Iran. J Renew Sustain
Energy 5:063134
Rajaeifar MA, Tabatabaei M, Abdi R, Latifi AM, Saberi F, Askari M, Zenouzi A, Ghorbani M
(2017b) Attributional and consequential environmental assessment of using waste cooking
oil-and poultry fat-based biodiesel blends in urban buses: a real-world operation condition
study. Biofuel Res J 4:638–653
8 Biodiesel Production and Consumption … 191

Reijnders L, Huijbregts M (2008) Biogenic greenhouse gas emissions linked to the life cycles of
biodiesel derived from European rapeseed and Brazilian soybeans. J Clean Prod 16:1943–1948
Reinhard J, Zah R (2011) Consequential life cycle assessment of the environmental impacts of an
increased rapemethylester (RME) production in Switzerland. Biomass Bioenergy 35:2361–
2373
Sáez-Bastante J, Ortega-Román C, Pinzi S, Lara-Raya F, Leiva-Candia D, Dorado M (2015)
Ultrasound-assisted biodiesel production from Camelina sativa oil. Biores Technol 185:116–
124
Salaa O, Saxa D, Lesliea H (2009) Biodiversity consequences of increased biofuel production. In:
Howarth RW, Bringezu S (eds) Environmental consequences and interactions with changing
land use. Proceedings of the scientific committee on problems of the environment (scope)
international biofuels project rapid assessment, Gummersbach Germany, Cornell University,
Ithaca NY, USA
Sander K, Murthy GS (2010) Life cycle analysis of algae biodiesel. Int J Life Cycle Assess
15:704–714
Schmidt J (2007) Life cycle assessment of rapeseed oil and palm oil. Thesis, Part 3: Life cycle
inventory of rapeseed oil and palm oil. Department of Development and Planning, Aalborg
University, Aalborg
Schmidt JH (2010) Comparative life cycle assessment of rapeseed oil and palm oil. Int J Life Cycle
Assess 15:183–197
Searchinger T, Heimlich R, Houghton RA, Dong F, Elobeid A, Fabiosa J, Tokgoz S, Hayes D, Yu
T-H (2008) Use of US croplands for biofuels increases greenhouse gases through emissions
from land-use change. Science 319:1238–1240
Seppälä J, Posch M, Johansson M, Hettelingh J-P (2006) Country-dependent characterisation
factors for acidification and terrestrial eutrophication based on accumulated exceedance as an
impact category indicator (14 pp). Int J Life Cycle Assess 11:403–416
Shah S, Sharma A, Gupta M (2005) Extraction of oil from Jatropha curcas L. seed kernels by
combination of ultrasonication and aqueous enzymatic oil extraction. Biores Technol 96:121–
123
Sheehan J, Camobreco V, Duffield J, Graboski M, Shapouri H (1998) Life cycle inventory of
biodiesel and petroleum diesel for use in an urban bus. Final report. National Renewable
Energy Lab, Golden, CO (US)
Singh A, Olsen SI, Nigam PS (2011) A viable technology to generate third-generation biofuel.
J Chem Technol Biotechnol 86:1349–1353
Singh A, Olsen SI, Pant D (2013) Importance of life cycle assessment of renewable energy
sources. In: Life cycle assessment of renewable energy sources. Springer, pp1–11
Sousa VM, Luz SM, Caldeira-Pires A, Machado FS, Silveira CM (2017) Life cycle assessment of
biodiesel production from beef tallow in Brazil. Int J Life Cycle Assess 22:1837–1850
Stojković IJ, Stamenković OS, Povrenović DS, Veljković VB (2014) Purification technologies for
crude biodiesel obtained by alkali-catalyzed transesterification. Renew Sustain Energy Rev
32:1–15
Styles D, Jones MB (2007) Energy crops in Ireland: quantifying the potential life-cycle greenhouse
gas reductions of energy-crop electricity. Biomass Bioenergy 31:759–772
Thomassen MA, Dalgaard R, Heijungs R, De Boer I (2008) Attributional and consequential LCA
of milk production. Int J Life Cycle Assess 13:339–349
Tran D-T, Chang J-S, Lee D-J (2017) Recent insights into continuous-flow biodiesel production
via catalytic and non-catalytic transesterification processes. Appl Energy 185:376–409
UNCED (1992) Report of the united nation conference on environment and development
van Dam J, Junginger M, Faaij AP (2010) From the global efforts on certification of bioenergy
towards an integrated approach based on sustainable land use planning. Renew Sustain Energy
Rev 14:2445–2472
Van Zutphen J, Wijbrans R (2012) LCA GHG emissions in production and combustion of
Malaysian palm oil biodiesel. J Oil Palm Environ Health (JOPEH) 2
192 M. A. Rajaeifar et al.

Verones F, Saner D, Pfister S, Baisero D, Rondinini C, Hellweg S (2013) Effects of consumptive


water use on biodiversity in wetlands of international importance. Environ Sci Technol
47:12248–12257
Watson R, Noble I, Bolin B, Ravindranath NH, Verardo D, Andrasko K, Apps M, Brown S,
Farquhar G, Goldberg D (2000) Summary for policymakers: land use, land-use change, and
forestry
Weidema B, Wenzel H, Petersen C, Hansen K (2004) The product, functional unit and reference
flows in LCA. Environ News 70:1–46
Wiloso EI, Heijungs R (2013) Key issues in conducting life cycle assessment of bio-based
renewable energy sources. In: Life cycle assessment of renewable energy sources. Springer,
pp 13–36
Wolf M-A, Chomkhamsri K, Brandao M, Pant R, Ardente F, Pennington DW, Manfredi S, de
Camillis C, Goralczyk M (2010). ILCD handbook-general guide for life cycle
assessment-detailed guidance
Wu M, Wu Y, Wang M (2008) Mobility chains analysis of technologies for passenger cars and
light duty vehicles fueled with biofuels: application of the Greet model to project the role of
biomass in America’s energy future (RBAEF) project. Argonne National Laboratory (ANL)
Xue X, Collinge WO, Shrake SO, Bilec MM, Landis AE (2012) Regional life cycle assessment of
soybean derived biodiesel for transportation fleets. Energy Policy 48:295–303
Yu N, Xing D, Li W, Yang Y, Li Z, Li Y, Ren N (2017) Electricity and methane production from
soybean edible oil refinery wastewater using microbial electrochemical systems. Int J
Hydrogen Energy 42:96–102
Zamagni A, Guinée J, Heijungs R, Masoni P, Raggi A (2012) Lights and shadows in consequential
LCA. Int J Life Cycle Assess 17:904–918
Chapter 9
Exergy-Based Sustainability Analysis
of Biodiesel Production and Combustion
Processes

Mortaza Aghbashlo, Meisam Tabatabaei, Mohammad Ali Rajaeifar


and Marc A. Rosen

Abstract Sustainability has become a relevant issue for the biodiesel industry. As a
consequence, increasingly advanced engineering methods and metrics are being
applied to make decisions on biodiesel production and combustion systems in order
to achieve the most thermodynamically, economically and environmentally sound
synthesis pathways and conditions. Among the various approaches developed,
exergy-based methods exhibit significant promise for the quantitative and qualitative
evaluation of energy conversion and biofuel production processes. Exergy-based
analyses provide valuable insights into the performance, costs and environmental
impacts of biodiesel production and combustion systems. In this chapter, after briefly
describing the exergy concept and its theoretical background, an overview is pro-
vided of the most important researches relating to the application of this approach and
its extensions for analyzing biodiesel production and combustion systems. In general,
quantifying exergy destruction rate and exergy efficiency is the greatest focus of
researchers globally when applying exergy method in this domain. However,
applications of extended exergy-based methods like exergoeconomic and exer-
goenvironmental analyses, as comprehensive decision-making paradigms for eval-
uating, optimizing, and retrofitting biodiesel production and combustion processes,
are limited. Future research is needed into finding the most efficient, cost-effective,

M. Aghbashlo (&)
Faculty of Agricultural Engineering and Technology, Department of Agricultural Machinery
Engineering, University College of Agriculture and Natural Resources, University of Tehran,
P.O. Box 4111 Tehran, Iran
e-mail: maghbashlo@ut.ac.ir
M. Tabatabaei  M. A. Rajaeifar
Biofuel Research Team (BRTeam), Karaj, Iran
M. Tabatabaei
Microbial Biotechnology Department, Agricultural Biotechnology Research Institute of Iran
(ABRII), Agricultural Research, Education and Extension Organization (AREEO), Karaj, Iran
M. A. Rosen
Faculty of Engineering and Applied Science, University of Ontario Institute of Technology,
2000 Simcoe Street North, Oshawa, ON L1H 7K4, Canada

© Springer Nature Switzerland AG 2019 193


M. Tabatabaei and M. Aghbashlo (eds.), Biodiesel,
Biofuel and Biorefinery Technologies 8,
https://doi.org/10.1007/978-3-030-00985-4_9
194 M. Aghbashlo et al.

and environmental-friendly routes for biodiesel synthesis and its subsequent uti-
lization, using exergoeconomic and exergoenvironmental approaches together with
advanced knowledge- and evolutionary-based optimization techniques.
Nomenclature
A Ash percentage (%)
C Carbon percentage (%)
Cp Specific heat capacity (kJ/kg K)
:
E Energy flow rate (kW)
ex Specific exergy (kJ/kg)
:
Ex Exergy flow rate (kW)
G Gibbs free energy (kJ/mol)
h Specific enthalpy (kJ/kg)
H Hydrogen percentage (%)
:
IP Exergetic improvement potential rate (kW)
:
m Mass flow rate (kg/s)
n Mole number (–)
:
n Molar flow rate (mol/s)
N Nitrogen percentage (%)
O Oxygen percentage (%)
P Pressure (kPa)
qLHV Lower heating value (kJ/kg)
:
Q Heat rate (kW)
R Gas constant (kJ/kg K)

R Universal gas constant (kJ/mol K)
s Specific entropy (kJ/kg K)
S Sulfur percentage (%)
SI Exergetic sustainability index (–)
T Temperature (°C or K)
x Mole fraction (–)
:
W Work flow rate (kW)

Greek Symbols
u Chemical exergy factor (–)
η Exergy efficiency (%)
e Standard chemical exergy (kJ/mol)

Subscript
0 Dead state
ch Chemical
dest Destruction
9 Exergy-Based Sustainability Analysis of Biodiesel Production … 195

f Fuel
in Input
i, j, k Numerator
ki Kinetics
l Loss
out Output
ph Physical
po Potential
P Product
R Reactant
s Source
w Work

9.1 Introduction

Although biodiesel is a renewable and clean form of bioenergy, the need for large
amounts of energy, water, and materials during its feedstock production, synthesis,
and purification is an important challenge of the biodiesel industry (Aghbashlo et al.
2018b). In recent years, both energy and environmental concerns have spurred
research into the development of resource-efficient and environmentally benign
biodiesel production technologies. It is also important that produced biodiesel be
converted into useful work in an environmentally benign and cost-effective manner.
Therefore, intensive research efforts have been devoted worldwide to determine the
most productive and sustainable biodiesel combustion conditions. It can be seen in
the literature that the productivity and sustainability of biodiesel production and
combustion processes have been investigated using various tools, such as emission
analysis, emergy analysis, life cycle assessment, energy analysis, and exergy-based
methods (Aghbashlo et al. 2018a). Among these, exergy-based analyses have
become increasingly popular. By integrating the concepts of both the first and
second laws of thermodynamics, the exergy approach is a comprehensive thermo-
dynamic tool. It has been acknowledged by many to be a powerful engineering tool
for the analysis, design, and optimization of many energy conversion processes.
Exergy methods have been used in various engineering disciplines to address both
energy and environmental concerns, due in part to its ability to quantify and locate
the sources of thermodynamic inefficiencies (Aghbashlo et al. 2018c).
In simple terms, exergy is the maximum capacity of a given form of energy or
material to produce useful work when it comes to complete equilibrium with a
reference environment (Aghbashlo et al. 2018d, 2018f). Exergy can also be
regarded as a measure of energy quality (Aghbashlo et al. 2018g). Due to this
ability, exergy-based methods have gained growing international acceptance
over the last few decades as effective methods for evaluating the efficiency, pro-
ductivity, and sustainability of various biofuel production and utilization
196 M. Aghbashlo et al.

Fig. 9.1 Representation of


the association of
sustainability and
environmental aspects of an
energy conversion system
with its exergy efficiency
(Dincer and Rosen 2005).
With permission from
Elsevier. Copyright © 2018

systems (Aghbashlo and Rosen, 2018; Aghbashlo et al. 2018e). In the real world,
all energy and material transformations involve entropy generation due to irre-
versibilities, and the dissipation of a portion of useful energy. Accordingly, the
exergy concept is closely related to sustainability and sustainable development, as
portrayed by Dincer and Rosen (2005) in Fig. 9.1. There, the environmental impact
is seen to approach zero and the sustainability simultaneously rises as the exergy
efficiency approaches 100% and vice versa.
In addition to standard exergy analysis, several researchers have enhanced the
quality of conclusions derived from exergy analysis by integrating thermodynamics
principles with economic and environmental constraints. Exergoeconomic and
exergoenvironmental analyses have been shown to be pioneering examples of
enhanced exergy-based methods for micro/macro-level analyses of energy con-
version systems from exergy/economic and exergy/environmental perspectives.
These approaches have been further extended by Tsatsaronis and Morosuk (2008a,
b) to determine endogenous/exogenous and avoidable/unavoidable parts of exergy
destruction as well as its corresponding costs and environmental impacts.
The main goal of this chapter is to describe the concept of exergy and its
theoretical formulations, as well as its use in analyzing biodiesel production and
combustion systems. Further, an overview is presented of the most important
applications of exergy-based approaches to biodiesel production and combustion
systems. Opportunities and advantages of exergy-based analyses over other avail-
able techniques as well as their limitations and disadvantages in analyzing biodiesel
production and combustion processes are also discussed.

9.2 Exergy Analysis

9.2.1 Exergy Concept

Energy is subject to a conservation law for processes, whether reversible and


irreversible, while exergy is destroyed because of irreversibilities (except during
reversible processes). The quantity of exergy destruction during an energy con-
version process is proportional to the amount of entropy generated. Generally,
9 Exergy-Based Sustainability Analysis of Biodiesel Production … 197

the exergy concept integrates the concepts of conservation of energy and non-
conservation of entropy.
Figure 9.2 illustratively compares energy, entropy, and exergy concepts
(Aghbashlo et al. 2013). The quantity of energy at the inlet and outlet sections
remains constant under steady-state conditions according to the first law of ther-
modynamics. However, the quantity of entropy at the outlet is larger than at the
inlet according to the second law of thermodynamics. The quantity of exergy at the
outlet is smaller than at the inlet due to the fact that exergy is destroyed during the
process due to irreversibilities, which lead to entropy generation. Overall, energy
analysis cannot provide insights on the irreversibility aspects of energy conversion
processes, while exergy analysis can, allowing it to provide meaningful and useful
guidelines for process and design improvement.

9.2.2 Mass, Energy, and Exergy Balances for Steady-State


Biodiesel Production and Combustion Processes

The mass balance for steady-state biodiesel production and combustion processes
can be written in the rate form as
X X
m_ in ¼ m_ out ð9:1Þ

Fig. 9.2 Comparison among energy, entropy and exergy concepts for a typical steady-state heat
transfer process through a wall (Adopted from Shukuya and Hammache 2002). The right-hand
side of the system is warmer than the left-hand side, and the circles on either side of the figure
indicate the vibrations of particles
198 M. Aghbashlo et al.

The energy balance can be expressed as follows:


X X
E_ in ¼ E_ out ð9:2Þ

The exergy balance can be written as below:


X X X
_ in 
Ex _ out ¼
Ex _ dest
Ex ð9:3Þ
X X X
m_ in exin  m_ out exout ¼ _ dest
Ex ð9:4Þ

The total exergy rate of all streams involved in the biodiesel production and
combustion processes can be computed as the summation of its physical, chemical,
kinetic, and potential exergy rates as follows:

_ ¼ Ex
Ex _ ph þ Ex
_ ch þ Ex
_ po þ Ex
_ ki ð9:5Þ

The kinetic and potential exergies of all streams can be neglected due to their
negligible effects on the total exergy. The thermodynamic properties (e.g., specific
enthalpy and specific entropy) of most pure streams can be easily obtained from the
literature and thermodynamics textbooks and, accordingly, their physical exergy
rates can be determined as follows:

_ ph ¼ m_ ðh  h0  T0 ðs  s0 ÞÞ
Ex ð9:6Þ

The physical exergy rates of mixed liquid and gas streams, respectively, can be
evaluated as follows:
   
_ T
Ex ¼ m_ Cp T  T0  T0 ln
ph
ð9:7Þ
T0
     
_ ph T P
Ex ¼ m_ Cp T  T0  T0 ln þ RT0 ln ð9:8Þ
T0 P0

The chemical exergy rate of streams can be computed according as follows:


!
X  X
_ ¼ n_
Ex ch
xi ei þ R T0 xi lnðxi Þ ð9:9Þ
i i

Standard chemical exergies of many organic and inorganic substances involved


in biomass and combustion processes can be obtained from the literature.
Moreover, several theoretical and semi-theoretical mathematical models have been
published for estimating the specific/standard chemical exergies of organic com-
ponents. For example, Song et al. (2012) developed the following expression for the
specific chemical exergies of organic components:
9 Exergy-Based Sustainability Analysis of Biodiesel Production … 199

exch ¼ 363:439C þ 1075:633H  86:308O þ 4:14N þ 190:798S  21:1A ð9:10Þ

The specific chemical exergy of fuel containing biodiesel (in a combustion


process) can be expressed as

exch ¼ uqLHV ð9:11Þ

The chemical exergy factor of the fuel (u) can be obtained as follows:
        
H O S H
u ¼ 1:0401 þ 0:1728 þ 0:0432 þ 0:2169  1  2:0268
C C C C
ð9:12Þ

The following basic equation can be used for determining the standard chemical
exergies of some inorganic substances, which cannot be found in the literature:
X X
e ¼ DG þ nj e j  nk e k ð9:13Þ
P R

The exergy rate of shaft work produced/consumed is equal to the energy rate of
the generated/utilized work. That is,

_ w ¼ W_
Ex ð9:14Þ

The heat loss rate to the environment from a control mass/volume can be found
using an energy balance. The heat loss rate can be converted to the corresponding
exergy rate as follows:
 
_ l ¼ Q_ l 1  T0
Ex ð9:15Þ
Ts

The exergetic efficiency of a diesel engine running on a biodiesel-containing fuel


can be written as follows:

W_
g¼  100 ð9:16Þ
_ ch
Ex f

The universal exergy efficiency of the biodiesel production system can be


obtained as follows:
 
_ out
Ex _ dest
Ex
g¼  100 ¼ 1  100 ð9:17Þ
Ex_ in Ex_ in
200 M. Aghbashlo et al.

Furthermore, the functional exergetic efficiency can be defined for a biodiesel


production system as the ratio of the product(s) exergy to the total exergy supplied
to the process. Note that this index can be a more realistic metric and more con-
textually relevant compared with the universal definition for assessing the perfor-
mance of biodiesel production processes. In addition, the normalized exergy
destruction can be defined for the process as the ratio of exergy destroyed to the
product(s) exergy. Using these indicators, various biodiesel production routes can
be meaningfully compared regardless of their feedstocks, capacities, arrangements,
operating conditions and locations.
Some additional exergetic performance parameters can be computed for a bio-
diesel production system. For instance, the exergetic improvement potential rate of
the process can be determined as follows:
 
_ ¼ ð1  gÞ Ex
IP _ in  Ex
_ out ð9:18Þ

and the exergetic sustainability index as follows:

100
g ¼ 100  ð9:19Þ
SI

9.2.3 Methodology for Exergy Analysis of Biodiesel


Production and Utilization Systems

For an exergy analysis of a biodiesel production or combustion process, a system


boundary (control mass/volume) first needs to be identified. Then, a diagram
showing mass and energy flows prepared, and all inlet and outlet mass and energy
flows labeled/numbered. Figures 9.3 and 9.4 show typical examples of diagrams of
biodiesel production and combustion systems, respectively.
Using the developed schematic diagram, mass, energy, and exergy balance
equations can be written for all subunits and the overall system. Chemical and
physical exergies of all mass and energy flows can be determined according to the
equations presented earlier. Finally, exergetic performance parameters of all sub-
units and the overall system can be determined. The results can be presented
numerically or graphically. For example, Figs. 9.5 and 9.6 present typical Sankey
diagrams for biodiesel production and combustion systems, respectively.
9 Exergy-Based Sustainability Analysis of Biodiesel Production … 201

Fig. 9.3 Schematic of a biodiesel production plant (Blanco-Marigorta et al. 2013). With
permission from Elsevier. Copyright © 2018

Fig. 9.4 Inputs and outputs


for exergy analysis of diesel
engine (Aghbashlo et al.
2015). With permission from
Elsevier. Copyright © 2018

9.2.4 Application of Exergy Analysis for Biodiesel


Production Processes

Although biodiesel is a viable alternative to petrodiesel, this environmentally


compatible fuel only can compete if its production routes are shown to be ther-
modynamically, economically, and environmentally sound. Traditionally, energy
analysis is used for addressing these issues. However, energetic indicators cannot
provide sufficient insights regarding the productivity and sustainability of biodiesel
production routes. This is because of the weaknesses of energetic indicators in
quantifying irreversibilities of energy conversion processes. During the couple past
202 M. Aghbashlo et al.

Fig. 9.5 Exergy flow diagram of biodiesel production process from canola oil (Talens et al.
2007). With permission from Elsevier. Copyright © 2018

decades, increased worldwide efforts have been undertaken to apply exergy-based


approaches for understanding and decision-making regarding the productivity and
sustainability of biodiesel production pathways, largely in response to both energy
and environmental concerns. Table 9.1 summarizes many of the most significant
applications of exergy-based approaches for analyzing and optimizing biodiesel
synthesis processes from various feedstocks.
Hovelius and Hansson (1999) determined the overall energy and exergy uti-
lization for rapeseed methyl ester production under Swedish conditions. Nitrogen
fertilizers and diesel fuels exhibited the highest exergetic contributions to the
production of rapeseed methyl ester. Talens et al. (2007) applied exergy analysis to
a biodiesel production process from waste cooking oil and found that the overall
exergy loss of the process was 492 MJ per ton of biodiesel produced. The exergetic
9 Exergy-Based Sustainability Analysis of Biodiesel Production … 203

Fig. 9.6 Exergy flow diagram of yellow grease methyl ester combustion process (Tat 2011). With
permission from Elsevier. Copyright © 2018

renewability index of biodiesel production from vegetable oils was improved by


10.5% by replacing the required fossil fuel input with solar energy (Hou and Zheng
2009). According to the extended exergy accounting analysis carried out by Peiró
et al. (2010), biodiesel production from waste cooking oil is a resource-efficient
pathway compared with the biodiesel obtained from rapeseed oil. More specifically,
the production of 1 ton of biodiesel from waste cooking oil used 25.15 GJ less
resources than from rapeseed methyl ester. That is, the extended exergy content of
the methyl ester produced from waste cooking oil is about 0.67 of that from
rapeseed oil biodiesel.
Sorguven and Özilgen (2010) reported that the amount of exergy lost in an algal
biodiesel production chain is much lower than the exergetic content of the biodiesel
evolved. More specifically, the exergy of the produced biodiesel is almost 213 times
of what is consumed in the production chain. Demir et al. (2011) found that almost
25% of the exergy supplied to a biodiesel synthesis plant is lost because of internal
thermodynamic irreversibilities. de Mora et al. (2012) presented a new framework
called “exergy return on investment” on the basis of exergy cost theory for eval-
uating the sustainability of various biodiesel production routes, from crop culti-
vation through to production. In simple terms, the proposed method seeks the
amount of biodiesel required to produce 1 kg of biodiesel. According to the results
reported based on the developed concept, all the investigated pathways for biodiesel
production from rapeseed, sunflower, and palm oil are sustainable. Ofori-Boateng
et al. (2012) demonstrated that biodiesel production from microalgae and jatropha
is not thermodynamically feasible since the majority of the exergy supplied to their
chains is lost due to thermodynamic inefficiencies. However, biodiesel production
from jatropha was shown to be a resource-efficient route compared with biodiesel
obtained using microalgae. Blanco-Marigorta et al. (2013) found that about 95% of
the exergy supplied to a biodiesel plant producing methyl esters from jatropha is
Table 9.1 Selected important applications of exergy-based approaches for analyzing and optimizing biodiesel synthesis from various feedstocks
204

Feedstock Biodiesel production Method of Data type Remarks References


method analysis
Rapeseed Esterification of hot Overall exergy Real data for Overall energy and exergy use for rapeseed methyl Hovelius and
water pressed oil utilization Swedish ester production was found to be 17,900 GJ/m3 and Hansson (1999)
conditions 16,700 GJ/m3 biodiesel, respectively
Waste Conventional Exergy flow Published and Exergetic efficiency of the process was computed as Talens et al.
cooking oil transesterification analysis industrial data 98.6% (2007)
Vegetable Conventional Classical exergy Simulation Exergetic renewability indexes of biodiesel production Hou and Zheng
oil transesterification analysis data obtained using fossil fuel and solar energy were found to be (2009)
using Aspen 89.4% and 99.99%, respectively
Plus
Waste Conventional Extended exergy Real data for Extended exergy contents of 1 ton of biodiesel Peiró et al. (2010)
cooking oil transesterification accounting Spanish obtained from waste cooking oil and rapeseed oil were
and conditions found to be 51.9 GJ and 77.1 GJ, respectively
rapeseed oil
Algae Conventional Cumulative Simulated Overall exergy loss in biodiesel production from algae Sorguven and
transesterification exergy data was determined as 0.19 GJ/ton of biodiesel, while the Özilgen (2010)
consumption exergy of the produced biodiesel was 40 found to be
GJ/ton
Canola Conventional Classical exergy Real data Overall mean energy and exergy efficiencies of the Demir et al.
transesterification analysis process were found to be 96.5% and 75.6%, (2011)
respectively
Rapeseed, Conventional Exergy cost Real data from Exergy return on investment was found to be 4.37, de Mora et al.
sunflower transesterification theory European 4.78 and 5.31 for rapeseed, sunflower, and palm oil (2012)
and palm countries
oil
(continued)
M. Aghbashlo et al.
Table 9.1 (continued)
Feedstock Biodiesel production Method of Data type Remarks References
method analysis
Microalgae Conventional Overall exergy Simulated 64% of the overall exergy supplied to biodiesel Ofori-Boateng
and transesterification utilization data using production chain from microalgae was destroyed, et al. (2012)
jatropha Aspen Plus while this value was 44% for jatropha methyl ester
production chains
Jatropha Conventional Classical exergy Simulated Overall exergy efficiency of the process was found to Blanco-Marigorta
curcas transesterification analysis data using be 63% et al. (2013)
Aspen Plus
African oil Convectional Exergetic Experimental Exergy-based Renewability Performance Indicator was Velásquez et al.
palm transesterification indicator based on and field data found to be 1.2 for biodiesel production from African (2013)
classical exergy oil palm
analysis
Canola oil Conventional Overall exergy Real data for Overall input exergy of methyl and ethyl ester Antonova et al.
transesterification utilization Belarus production processes was found to be 31.0 GJ/ha yr (2015)
conditions and 33.3 GJ/ha yr, respectively
Cooking Acid-catalyzed Exergy-based Simulated Energy requirements for methanol recovery, biodiesel Fu et al. (2015)
waste transesterification self-heat data purification and glycerol purification steps of the
recuperation optimized process declined by 83.5%, 88.4% and
58.8%, respectively, compared with the base process
9 Exergy-Based Sustainability Analysis of Biodiesel Production …

Pure Mechanical Classical exergy Simulated Optimized point using energy and exergy approaches Amelio et al.
triolein transesterification analysis data using reduced the energy consumption by 45.5% and 46%, (2016)
Aspen Plus respectively
Waste Low power, high Classical exergy Experimental Functional exergetic efficiency of the process ranged Aghbashlo et al.
cooking oil frequency analysis data from 9 to 91% (2017)
ultrasound-assisted
transesterification
205
206 M. Aghbashlo et al.

destroyed in the transesterification reactor, making it the most significant part of the
plant in terms of exergy loss.
Velásquez et al. (2013) compared first- and second-generation ethanol produc-
tion processes with oil palm biodiesel production from renewability and sustain-
ability viewpoints using an exergy-based “renewability performance indicator”.
Interestingly, both ethanol production pathways were found to be unsustainable,
while biodiesel production from oil palm was shown to be sustainable due to its
lower exergy destruction. Antonova et al. (2015) calculated the overall exergy
utilization for assessing the sustainability of biodiesel production from canola under
Belarus conditions. They reported that the canola cultivation stage makes the
highest contribution to the overall exergy utilization in biodiesel production due to
the application of fertilizers, chemicals, and fossil fuels with high standard exergy
values. Fu et al. (2015) reduced the energy requirement of biodiesel production by
71% using exergy-based self-heat recuperation technology that recovers both
sensible and latent waste heat. Amelio et al. (2016) compared energy and exergy
approaches for optimizing biodiesel production process from pure triolein. The
optimized point obtained using exergy was different than that achieved using
energy analysis. Aghbashlo et al. (2017) selected a methanol to oil molar ratio of
6:1, an ultrasonic irradiation time of 10 min, and a temperature of 60 °C as the best
conditions, from an exergetic viewpoint, for synthesizing biodiesel from waste
cooking oil using a low power, high frequency ultrasound-assisted reactor.
According to the findings of the above-mentioned surveys, exergy-based
approaches can be useful sustainability assessment tools with a wide variety of
applications for various biodiesel production pathways. These methods have been
applied to date only to a few categories of biodiesel production routes, but their
potential applications are much broader. Also, some authors have attempted to
demonstrate the advantages of exergy-based approaches over energy-based indi-
cators, generally confirming the superiority of exergy methods as demonstrated by
various researchers. It is noted that the exergetic contents of externalities like
capital, labor, and environmental impact spent to synthesize a given amount of
biodiesel have usually been neglected by investigators in measuring the life cycle
exergy utilization of biodiesel production. This issue can be addressed through the
application of Extended Exergy Accounting, such as done by Peiró et al. (2010).
The majority of studies carried out in this domain used simulated data for exergy
analysis, while accurate determinations of the exergetic contents of externalities for
such assessments are complex and not straightforward.
Overall, little research has been reported on the application of exergy-based
methods for analyzing, optimizing and retrofitting industrial biodiesel production
plants, likely due in part to commercial barriers to the publication of scientific/
technological details. Future studies should include exergoeconomic and exer-
goenvironmental analyses for micro/macro-level analyses of commercial biodiesel
production plants from exergy/economic and exergy/environmental perspectives,
respectively. In addition, the exergetic performance of biodiesel production plants
can be significantly enhanced by applying the methods described herein, such as
the use of the self-heat recuperation technology describe by Fu et al. (2015).
9 Exergy-Based Sustainability Analysis of Biodiesel Production … 207

Pinch technology is another promising approach to reduce heating and cooling


loads of biodiesel production plants by recovering waste heat and integrating
processes. Renewable energy resources can be incorporated in biodiesel production
plants for providing the majority of heat and electricity required by the process, as
examined by Hou and Zheng (2009). Valorization of glycerin as the main
by-product of biodiesel production plants into added-value fuels and chemicals
using the biorefinery concept can also be applied, as it can provide an efficient
approach to enhance economic revenues. However, exergy-based approaches like
exergoeconomic and exergoenvironmental analyses need to be applied to confirm
the cost-effectiveness and environmental benefits of such modifications and
additions.

9.2.5 Application of Exergy Analysis for Combustion


of Biodiesel and Its Blends

The brake thermal efficiency (bte) of diesel engines is based on the first law of
thermodynamics, expressing the fact that energy can neither be created nor be
destroyed. Hence, this performance indicator cannot provide information on the
irreversibility aspects of combustion processes. Moreover, the quality of various
energy and material flows cannot be appropriately recognized using the first law of
thermodynamics. For these reasons, the analysis of brake thermal efficiency is not
sufficient for identifying the most eco-friendly fuel blends and the best engine
operational conditions. It is fortunate that valuable insights on the local and overall
irreversibilities of combustion processes can be attained using exergy analysis. The
main objectives of applying exergy analysis to combustion processes are normally
to determine advantageous or optimal combustion conditions and fuels blends, to
diminish the environmental impacts of engines, and to quantify the origins of
thermodynamic inefficiencies. It is well documented that exergetic parameters,
particularly exergy efficiency, are much more suitable than brake thermal efficiency
to assess the performance of various fuel blends. The exergy efficiency of a diesel
engine can be defined as the ratio of shaft work generated by the engine to the
exergetic content of fuel blend introduced to the combustion chamber. In other
words, this value indicates the capability of a given diesel engine to convert
chemical exergy of fuel to shaft work exergy. Table 9.2 lists many of the most
significant applications of exergy analysis for evaluating the performance of diesel
engines fueled by biodiesel and its blends.
Gokalp et al. (2008) found that the exergy efficiency of standard diesel fuel was
improved by the addition of soybean oil methyl ester, in that enhanced combustion
was achieved by the homogeneous mixture in the chamber. Benjumea et al. (2009)
found that elevating the altitude lowered the exergy efficiency of a diesel engine
working with palm oil biodiesel and neat diesel. This could be related to the
advance in injection and combustion timings at higher altitudes. A similar argument
Table 9.2 Selected significant applications of exergy analysis for evaluating the performance of diesel engines fueled with biodiesel and its blends
208

Fuel blend(s) Basal Engine characteristics Exergy Remarks References


fuel efficiency
(%)
B100 Diesel Four-cylinder, four-stroke TZDK Basak " Exergy efficiency of engine increased with Gokalp et al.
50%D + 50%B engine; 40 kW; displacement = 3.14 L " biodiesel addition into neat diesel (2008)
80%D + 20%B #
95%D + 5%B #
B100 Diesel Four-cylinder ISUZU engine; 59 kW; # Exergy efficiency of engine running on palm oil Benjumea
displacement = 2499 cm3 biodiesel was lower than on neat diesel at altitudes et al. (2009)
of both 500 and 2400 m
B100 (soybean Diesel Four-cylinder, John Deere 4045T " Soybean methyl ester had slightly higher exergy Caliskan and
methyl ester) engine; 66.5 kW " efficiency than high-oleic soybean oil methyl ester Hepbasli
B100 (high-oleic and neat diesel (2011)
soybean oil methyl
ester)
B100 Diesel Four-cylinder, water-cooled Steyr " Exergy efficiency of neat biodiesel operation was Sekmen and
engine; 46 kW higher compared with diesel fuel Yilbaşi
(2011)
B100 (yellow B100 Four-cylinder, turbocharged, John Deere # Exergy efficiency of engine did not increase using Tat (2011)
grease methyl (SME) 4045 T engine; displacement = 4.5 L # cetane improver (2-ethylhexyl nitrate, EHN)
ester) #
B (soybean methyl
ester) + 0.75%wt%
EHN
B (soybean methyl
ester) + 1.5%wt%
EHN
(continued)
M. Aghbashlo et al.
Table 9.2 (continued)
Fuel blend(s) Basal Engine characteristics Exergy Remarks References
fuel efficiency
(%)
D + Biogas Diesel Single cylinder, Kirloskar engine; # Exergy efficiencies of two fuel blends were much Sahoo et al.
B + Biogas 5.2 kW; displacement = 661 cc # lower than neat diesel at various engine loads (2012)
B10 Diesel One-cylinder, Kirloskar engine; 3.5 kW # Adding koroch seed oil methyl ester to mineral Gogoi
B20 # diesel reduced exergy efficiency of engine (2013)
B30 #
B40 #
B100 (palm Diesel One-cylinder, Kirloskar engine; 5.2 kW; " Two biodiesels had better exergetic performance Jena and
biodiesel) displacement = 661 cc " compared with neat diesel Misra (2014)
B100 (karanja
biodiesel)
B20 Diesel Three-cylinder Perkins engine; 34 kW; nc Biodiesel addition to diesel did not negatively López et al.
B50 displacement = 2500 cm3 nc affect exergy efficiency of engine (2014)
B80 nc
B100 nc
B5 Diesel Four-cylinder, turbocharged OM314LA # Maximum exergy efficiency was obtained for B5 Aghbashlo
B5 + 25 EPS g/L EUll; 81 kW # blend containing 50 g EPS/L biodiesel et al. (2015)
biodiesel "
B5 + 50 EPS g/L #
9 Exergy-Based Sustainability Analysis of Biodiesel Production …

biodiesel
B5 + 75 EPS g/L
biodiesel
B5 Diesel Four-cylinder, turbocharged, MT 4.244 # B10 exhibited a slightly higher exergy efficiency Meisami and
B10 engine; displacement = 3.99 L " than neat diesel Ajam (2015)
B15 nc
B20 #
B30 #
(continued)
209
Table 9.2 (continued)
210

Fuel blend(s) Basal Engine characteristics Exergy Remarks References


fuel efficiency
(%)
B5 + 30 ppm B5 Four-cylinder, turbocharged IDEM nc Exergetic performance parameters of B5 and B20 Aghbashlo
MWCNTs-CeO2 engine; 81 kW nc blends were not negatively affected by adding et al. (2016)
B5 + 60 ppm nc MWCNTs-CeO2
MWCNTs-CeO2
B5 + 90 ppm
MWCNTs-CeO2
B20 + 30 ppm B20 nc
MWCNTs-CeO2 nc
B20 + 60 ppm nc
MWCNTs-CeO2
B20 + 90 ppm
MWCNTs-CeO2
B100 Diesel Four-cylinder, Mitsubishi Canter; # Peanut biodiesel exhibited slightly lower exergy Hürdoğan
displacement = 3907 cc efficiency than mineral diesel (2016)
B5 Diesel One-cylinder engine; # B20 was found to be the optimal fuel blend Nemati et al.
B10 displacement = 662 cm3 # according to quasi-dimensional multi-zone (2016)
B20 " simulation approach
B100 "
B10 (apricot) Diesel One-cylinder Kirloskar engine; " Fuel blend containing 20% v/v Argemone Singh et al.
B10 (argemone) 3.75 kW; displacement = 661.45 " Mexicana biodiesel was found to be best in terms (2016)
B10 (karanja) " of biodiesel exergetic performance parameters
B10 (nahar) "
B10 (neem) "
B20 (apricot) "
B20 (argemone) "
B20 (karanja) "
B20 (nahar) "
M. Aghbashlo et al.

B20 (neem) "


(continued)
Table 9.2 (continued)
Fuel blend(s) Basal Engine characteristics Exergy Remarks References
fuel efficiency
(%)
B100 (methyl Diesel One-cylinder engine # Exergy efficiency of sunflower ethyl ester was Yamik
ester) " higher than that of methyl ester (2015)
B75 (methyl ester) nc
B50 (methyl ester) nc
B25 (methyl ester) #
B100 (ethyl ester) "
B75 (ethyl ester) "
B50 (ethyl ester) nc
B25 (ethyl ester)
B20 Diesel Four-cylinder, turbocharged, Mitsubishi " Maximum exergy efficiency was found for neat Caliskan and
B50 Fuso diesel engine; 96 kW; " biodiesel from waste cooking oil Mori (2017)
B100 displacement = 3 L "
The symbol #/" represents the trend of variations reported in each study and nc shows no change compared with basal fuel
9 Exergy-Based Sustainability Analysis of Biodiesel Production …
211
212 M. Aghbashlo et al.

was also presented for the lower exergy efficiency of biodiesel compared with
petrodiesel. Caliskan and Hepbasli (2011) reported that the exergy efficiency of a
diesel engine when using both soybean methyl ester and high-oleic soybean oil
methyl ester was higher than when using neat diesel. The exergetic performance of
a diesel engine was slightly higher when operating on soybean methyl ester than on
petroleum diesel fuel (Sekmen and Yilbaşi 2011). Tat (2011) concluded that exergy
efficiency of a diesel engine can be increased by lowering the cetane number and
prolonging the ignition delay and that a higher level of premixed combustion might
lead to better exergetic performance of the engine. Sahoo et al. (2012) found that
supplementing biogas to the engine manifold decreased the exergy efficiency of an
engine working with jatropha biodiesel and neat diesel compared with neat diesel.
This can be attributed to the fact that feeding CO2-rich biogas into the combustion
chamber can dilute the biogas-air mixture, which in turn decreases the combustion
temperature and diminishes engine performance.
Jena and Misra (2014) argued that the presence of oxygen in palm and karanja
biodiesels can lead to better exergetic performance compared with neat diesel. López
et al. (2014) concluded that olive–pomace oil biodiesel and its blends with neat
diesel fuel can replace the use of diesel fuel without any unfavorable changes in the
exergetic performance of the DI diesel engine. Meisami and Ajam (2015) proposed
B5 as an affordable blend based on an economic analysis carried out by considering
emissions, fuel cost, fuel consumption, and engine power loss. Gogoi (2013) found
that adding koroch seed oil methyl ester to mineral diesel reduced the exergy effi-
ciency of a diesel engine. Aghbashlo et al. (2015) improved the exergy efficiency of
a diesel running on a B5 blend by adding 50 g of expanded polystyrene (EPS) per
liter of biodiesel. Aghbashlo et al. (2016) reported that cerium oxide immobilized on
amide-functionalized multiwall carbon nanotubes (MWCNTs-CeO2) did not sig-
nificantly affect the exergetic performance of a diesel engine working with B5 and
B20 blends while mitigating the harmful emissions substantially.
Singh et al. (2016) reported that fuel blends containing 10 and 20% v/v apricot,
Argemone, karanja, nahar, and neem biodiesels exhibited better exergy efficiencies
compared with mineral diesel. Peanut biodiesel exhibited a slightly lower exergy
efficiency than mineral diesel (Hürdoğan 2016). Nemati et al. (2016) simulated and
analyzed exergetically a diesel engine fueled with biodiesel derived from waste
cooking oil at full load operation, using a quasi-dimensional, multi-zone model.
The best combustion performance was observed for B20. Yamik et al. (2015)
compared the exergetic performance of sunflower methyl and ethyl esters and their
blends with mineral diesel and concluded that 75 and 50% ethyl ester/diesel fuel
blends were the most in terms of exergy efficiency. Caliskan and Mori (2017)
compared the effects of two after treatment systems, i.e., diesel oxidation catalyst
and diesel particulate filter, on the exergetic performance of an engine using various
diesel/biodiesel blends. Also, an exergoeconomic analysis was carried out based on
the exergy–cost–energy–mass (EXCEM) method. Both treatment approaches,
particularly their combination, were shown to be thermodynamically and
environmentally viable. Enhancing the quantity of biodiesel in fuel blends
improved the exergetic and exergoeconomic performance parameters of the engine.
9 Exergy-Based Sustainability Analysis of Biodiesel Production … 213

Overall, B100 synthesized from waste cooking oil was found to be an exergetically,
exergoeconomically and environmentally sound fuel compared with B20, B50, and
mineral diesel.
Generally, the effect of using biodiesel as a diesel substitute or additive on
engine exergy efficiency is somewhat controversial because of the fact that there is
no general consensus among the researchers working on this subject. Some of the
previously published reports imply that substituting biodiesel for petroleum diesel
could significantly reduce the exergy efficiency of diesel engines (Benjumea et al.
2009; Gogoi 2013; Hürdoğan 2016). This can be ascribed to the lower heating
values of methyl/ethyl esters compared with neat diesel and its blends. Yet,
numerous other investigators state that using biodiesel in neat or blended forms can
improve the exergy efficiency of diesel engines (Gokalp et al. 2008; Caliskan and
Hepbasli 2009; Sekmen and Yilbaşi 2011; Jena and Misra 2014; Singh et al. 2016;
Caliskan and Mori 2017). In general, it can be concluded that improvements in the
exergy efficiencies of biodiesel-fueled engines can be attributed, on one hand, to the
structural oxygen content of biodiesel and, on the other hand, to the higher lubricity
of biodiesel. These effects can improve combustion during the diffusion combustion
period and reduce frictional losses (Sekmen and Yilbaşi 2011). In contrast with the
two groups of studies discussed above (i.e., reports that substituting biodiesel for
petroleum diesel could significantly reduce or increase the exergy efficiency of
diesel engines), López et al. (2014) report that there is no statistically significant
difference between the exergy efficiency of a diesel engine fueled with olive–
pomace oil biodiesel and neat diesel. In this case, it seems that the structural oxygen
of biodiesel compensates for its lower heating value, offsetting the adverse effect of
heating value on the exergy efficiency of the diesel engine. Note that engine
characteristics, engine operating conditions, experimental conditions, biodiesel fatty
acid ester profile, and experimental uncertainties all can affect the exergy efficiency
of biodiesel-fueled diesel engines.
Even though the most commonly used biodiesel blend on the market is B5 (5%
biodiesel, 95% petroleum diesel) as approved by ASTM, some authors use bio-
diesel mixes up to 80% or even 100% in their studies. In general, parameters like
cetane number, viscosity, oxygen content, and calorific value of diesel/biodiesel
blends can potentially affect the exergy efficiency of engines working with these
fuels. However, there are often synergistic and antagonistic interactions among
these factors. Such interactions can, therefore, positively or negatively shift the
trend of exergy efficiency, depending on engine characteristics and operating
conditions as well as biodiesel fatty acid ester profile. This explains why some
researchers observe an increase in the exergy efficiency of engines running on
diesel/biodiesel blends compared with neat diesel fuel, while others observe
otherwise for different quantities and qualities of the added biodiesel. It is con-
cluded that a trade-off among these factors needs to be considered in order to
determine the optimum diesel/biodiesel blends for maximizing exergy efficiency.
Generally, the literature survey reported here indicates that exergy analysis has
been applied for some diesel/biodiesel blends. Future studies should be directed
toward the use of exergy analysis for diesel engines working with such diesel/
214 M. Aghbashlo et al.

biodiesel additives such as metallic and polymeric based additives as well as


antioxidants. Although, experimental studies of combustion processes are often
laborious, expensive, and time-consuming, and they involve measurement errors,
they are important. But, simulations of diesel engines working with various diesel/
biodiesel blends, and analyses using continuity, mass, momentum, energy, entropy,
and exergy balance equations are often more advantageous over experimental
studies due to the savings in cost and time like. The study carried out by Nemati
et al. (2016) is one example. Preparing fuel blends and studying their combustion
also is expensive and time-consuming, and simulations can contain the designing/
redesigning phase-out of the loop using models already developed in the design
step. Finally, the application of exergoeconomic and exergoenvironmental analyses
along with advanced optimization techniques for decision-making regarding the
operational conditions of diesel engines and fuel compositions need to be taken into
account in future studies. Such comprehensive optimization processes could yield
in-depth insights on the relationships among thermodynamic, economic and the
environment performance.

9.3 Conclusions

Exergy analysis has been used for analyzing and optimizing various biodiesel
production and combustion systems because of its capability to reveal thermody-
namic inefficiencies, as well as irreversibilities. Interest in applying exergy analysis
to biodiesel production and combustion processes is growing because of both
energy and environmental concerns. Additionally, exergy methods have the
potential to enhance the integration of thermodynamic performance with economic
and environmental constraints, providing more comprehensive approaches such as
exergoeconomic and exergoenvironmental analyses. These can be used to make
decisions on the productivity and sustainability aspects of biodiesel production and
combustion processes. For these reasons, it is almost certain that the exergy
approach and its extensions will serve as powerful tools for designing and devel-
oping resource-efficient, cost-effective, and eco-friendly biodiesel production and
combustion processes. Exergy-based analyses of some biodiesel production plants
have been reported in the literature, while exergoeconomic and exergoenviron-
mental analyses have rarely been used for biodiesel production and combustion
systems. Future work should be directed to the application of advanced
exergy-based methods for analyzing available and new biodiesel production and
combustion systems and to exploring strategies to enhance their thermodynamic,
economic, and environmental performance, simultaneously. The overall natural
resource consumption of various biodiesel production routes can be best-assessed
using farm-to-tank or farm-to-shaft strategies. This can be performed by combining
exergy and life cycle assessment concepts into a single framework called exergetic
life cycle assessment. However, in such studies, non-energetic/immaterial exter-
nalities should be converted to exergetic terms or resource equivalents using the
9 Exergy-Based Sustainability Analysis of Biodiesel Production … 215

extended exergy accounting concept. It is also suggested that renewable energy


resources be used in order to provide all or a portion of energy required for bio-
diesel production and purification processes. Valorization of the glycerin produced
in the transesterification process into added-value chemicals and fuels can improve
the overall exergoeconomic and exergoenvironmental performances of biodiesel
production plants.

Acknowledgements The authors gratefully acknowledge financial support from the Iran National
Science Foundation (Grant no. 96005466).

References

Aghbashlo M, Mobli H, Rafiee S, Madadlou A (2013) A review on exergy analysis of drying


processes and systems. Renew Sustain Energy Rev 22:1–22
Aghbashlo M, Tabatabaei M, Mohammadi P, Pourvosoughi N, Nikbakht AM, Goli SAH (2015)
Improving exergetic and sustainability parameters of a DI diesel engine using polymer waste
dissolved in biodiesel as a novel diesel additive. Energy Convers Manag 105:328–337
Aghbashlo M, Tabatabaei M, Mohammadi P, Mirzajanzadeh M, Ardjmand M, Rashidi A (2016)
Effect of an emission-reducing soluble hybrid nanocatalyst in diesel/biodiesel blends on
exergetic performance of a DI diesel engine. Renew Energy 93:353–368
Aghbashlo M, Tabatabaei M, Hosseinpour S, Khounani Z, Hosseini SS (2017) Exergy-based
sustainability analysis of a low power, high frequency piezo-based ultrasound reactor for rapid
biodiesel production. Energy Convers Manag 148:759–769
Aghbashlo M, & Rosen MA (2018) Exergoeconoenvironmental analysis as a new concept for
developing thermodynamically, economically, and environmentally sound energy conversion
systems. J Cleaner Prod 187:190–204
Aghbashlo M, Mandegari M, Tabatabaei M, Farzad S, Soufiyan MM, Görgens JF (2018a) Exergy
analysis of a lignocellulosic-based biorefinery annexed to a sugarcane mill for simultaneous
lactic acid and electricity production. Energy 149:623–638
Aghbashlo M, Tabatabaei M, Hosseinpour S (2018b) On the exergoeconomic and exergoenvi-
ronmental evaluation and optimization of biodiesel synthesis from waste cooking oil
(WCO) using a low power, high frequency ultrasonic reactor. Energy Convers Manage
164:385–398
Aghbashlo M, Tabatabaei M, Jazini H, Ghaziaskar HS (2018c) Exergoeconomic and exergoenvi-
ronmental co-optimization of continuous fuel additives (acetins) synthesis from glycerol
esterification with acetic acid using Amberlyst 36 catalyst. Energy Convers Manage 165:183–194
Aghbashlo M, Tabatabaei M, Khalife E, Shojaei TR, Dadak A (2018d) Exergoeconomic analysis
of a DI diesel engine fueled with diesel/biodiesel (B5) emulsions containing aqueous nano
cerium oxide. Energy 149:967–978
Aghbashlo M, Tabatabaei M, Rastegari H, Ghaziaskar HS (2018e) Exergy-based sustainability
analysis of acetins synthesis through continuous esterification of glycerol in acetic acid using
Amberlyst® 36 as catalyst. J Cleaner Prod 183:1265–1275
Aghbashlo M, Tabatabaei M, Rastegari H, Ghaziaskar HS, Shojaei TR (2018f) On the exergetic
optimization of solketalacetin synthesis as a green fuel additive through ketalization of
glycerol-derived monoacetin with acetone. Renew Energy 126:242–253
Aghbashlo M, Tabatabaei M, Rastegari H, Ghaziaskar HS, Valijanian E (2018g) Exergy-based
optimization of a continuous reactor applied to produce value-added chemicals from glycerol
through esterification with acetic acid. Energy 150:351–362
216 M. Aghbashlo et al.

Amelio A, Van de Voorde T, Creemers C, Degrève J, Darvishmanesh S, Luis P, Van der


Bruggen B (2016) Comparison between exergy and energy analysis for biodiesel production.
Energy 98:135–145
Antonova ZA, Krouk VS, Pilyuk YE, Maksimuk YV, Karpushenkava LS, Krivova MG (2015)
Exergy analysis of canola-based biodiesel production in Belarus. Fuel Process Technol
138:397–403
Benjumea P, Agudelo J, Agudelo A (2009) Effect of altitude and palm oil biodiesel fuelling on the
performance and combustion characteristics of a HSDI diesel engine. Fuel 88(4):725–731
Blanco-Marigorta AM, Suárez-Medina J, Vera-Castellano A (2013) Exergetic analysis of a
biodiesel production process from Jatropha curcas. Appl Energy 101:218–225
Caliskan H, Hepbasli A (2011) Exergetic cost analysis and sustainability assessment of an internal
combustion engine. Int J Exergy 8(3):310–324
Caliskan H, Mori K (2017) Environmental, enviroeconomic and enhanced thermodynamic
analyses of a diesel engine with diesel oxidation catalyst (DOC) and diesel particulate filter
(DPF) after treatment systems. Energy 128:128–144
de Mora EF, Torres C, Valero A (2012) Assessment of biodiesel energy sustainability using the
exergy return on investment concept. Energy 45(1):474–480
Demir B, Utlu Z, Kocar G, Hepbasli A (2011) Exergy assessment of biodiesel production process:
application. J Energy Inst 84(4):236–245
Dincer I, Rosen MA (2005) Thermodynamic aspects of renewables and sustainable development.
Renew Sustain Energy Rev 9(2):169–189
Fu Q, Song C, Kansha Y, Liu Y, Ishizuka M, Tsutsumi A (2015) Energy saving in a biodiesel
production process based on self-heat recuperation technology. Chem Eng J 278:556–562
Gogoi TK (2013) Exergy analysis of a diesel engine operated with koroch seed oil methyl ester
and its diesel fuel blends. Int J Exergy 12(2):183–204
Gokalp B, Soyhan HS, Saraç Hİ, Bostan D, Şengün Y (2008) Biodiesel addition to standard diesel
fuels and marine fuels used in a diesel engine: effects on emission characteristics and first-and
second-law efficiencies. Energy Fuels 23(4):1849–1857
Hou Z, Zheng D (2009) Solar utility and renewability evaluation for biodiesel production process.
Appl Therm Eng 29(14):3169–3174
Hovelius K, Hansson PA (1999) Energy-and exergy analysis of rape seed oil methyl ester
(RME) production under Swedish conditions. Biomass Bioenergy 17(4):279–290
Hürdoğan E (2016) Thermodynamic analysis of a diesel engine fueled with diesel and peanut
biodiesel. Environ Prog Sustain Energy 35(3):891–897
Jena J, Misra RD (2014) Effect of fuel oxygen on the energetic and exergetic efficiency of a
compression ignition engine fuelled separately with palm and karanja biodiesels. Energy
68:411–419
López I, Quintana CE, Ruiz JJ, Cruz-Peragón F, Dorado MP (2014) Effect of the use of olive–
pomace oil biodiesel/diesel fuel blends in a compression ignition engine: preliminary exergy
analysis. Energy Convers Manag 85:227–233
Meisami F, Ajam H (2015) Energy, exergy and economic analysis of a diesel engine fueled with
castor oil biodiesel. Int J Engine Res 16(5):691–702
Nemati P, Jafarmadar S, Taghavifar H (2016) Exergy analysis of biodiesel combustion in a direct
injection compression ignition (CI) engine using quasi-dimensional multi-zone model. Energy
115:528–538
Ofori-Boateng C, Keat TL, JitKang L (2012) Feasibility study of microalgal and jatropha biodiesel
production plants: exergy analysis approach. Appl Therm Eng 36:141–151
Peiró LT, Méndez GV, Sciubba E, i Durany XG (2010) Extended exergy accounting applied to
biodiesel production. Energy 35(7):2861–2869
Sahoo BB, Saha UK, Sahoo N (2012) Diagnosing the effects of pilot fuel quality on exergy terms
in a biogas run dual fuel diesel engine. Int J Exergy 10(1):77–93
Sekmen P, Yilbaşi Z (2011) Application of energy and exergy analyses to a CI engine using
biodiesel fuel. Math Comput Appl 16(4):797–808
9 Exergy-Based Sustainability Analysis of Biodiesel Production … 217

Shukuya M, Hammache A (2002) Introduction to the concept of exergy. Low exergy systems for
heating and cooling of buildings, IEA ANNEX37 Finland, pp 41–44
Singh N, Kumar D, Sarma AK, Jha MK (2016) Exergy conceptual-based study for comparative
thermodynamic performance of ci engine fueled with petroleum diesel and biodiesel blends.
J Energy Eng 04016030
Song G, Xiao J, Zhao H, Shen L (2012) A unified correlation for estimating specific chemical
exergy of solid and liquid fuels. Energy 40(1):164–173
Sorguven E, Özilgen M (2010) Thermodynamic assessment of algal biodiesel utilization.
Renewable Energy 35(9):1956–1966
Talens L, Villalba G, Gabarrell X (2007) Exergy analysis applied to biodiesel production. Resour
Conserv Recycl 51(2):397–407
Tat ME (2011) Cetane number effect on the energetic and exergetic efficiency of a diesel engine
fuelled with biodiesel. Fuel Process Technol 92(7):1311–1321
Tsatsaronis G, Morosuk T (2008a) A general exergy-based method for combining a cost analysis
with an environmental impact analysis. Part II—application to a cogeneration system. In:
Proceedings of the ASME IMECE, file IMECE2008-67219, Boston, MA
Tsatsaronis G, Morosuk T (2008b) A general exergy-based method for combining a cost analysis
with an environmental impact analysis. Part I–theoretical development. In: Proceedings of the
ASME international mechanical engineering congress and exposition, vol 31
Velásquez HI, De Oliveira S, Benjumea P, Pellegrini LF (2013) Exergo-environmental evaluation
of liquid biofuel production processes. Energy 54:97–103
Yamık H, Özel G, Açikkalp E, İçingür Y (2015) Thermodynamic analysis of diesel engine with
sunflower biofuel. Proc Inst Civil Eng-Energy 168(3):178–187
Chapter 10
“Omics Technologies” and Biodiesel
Production

Reza Sharafi and Gholamreza Salehi Jouzani

Abstract Biodiesel is being considered as a renewable fuel candidate to com-


pletely or partially replace fossil diesel. The most important challenge in devel-
opment of different generations of biodiesel is input cost, low oil yield in the
sources, and lack of efficient technologies for biodiesel production. Recent devel-
opments in next-generation sequencing technologies (NGS) and new “omics”
methodologies have provided excellent opportunities for high-throughput func-
tional genomic surveys in different organisms. In this context, different “Omics”
technologies have been widely used to enhance the oil yield in oil-producing plants
and microorganisms. This chapter reviews the existing studies revolving around
new “Omics” technologies used to enhance the oil and biodiesel production in the
promising plant for biodiesel production, Jatropha, as a sample.

10.1 Omics Technologies: An Introduction

The revolutionary advances in the molecular biology technologies have signifi-


cantly improved life sciences during the last decades. New “omics” technologies
have become available for different families of cellular molecules, including genes,
RNAs, proteins, and metabolites (Raupach et al. 2016). These new valuable tools
have provided wonderful opportunities for high-throughput phylogenic studies
among different individuals, communities and species, structural and functional
genomics studies, discovery of biochemical pathways and metabolomes, and
genetic and metabolic engineering (Debnath et al. 2010). This progress has been
realized via the development of next-generation sequencing technologies, such as
Illumina/Solexa, ABI/SOLiD, 454/Roche, and Helicos. Currently, these NGS
technologies are used in a variety of contexts, such as de novo sequencing or

R. Sharafi  G. S. Jouzani (&)


Department of Microbial Biotechnology, Agricultural Biotechnology Research Institute of
Iran (ABRII), Agricultural Research, Education and Extension Organization (AREEO),
Fahmideh Blvd, P.O. Box: 31535-1897, Karaj, Iran
e-mail: gsalehi@abrii.ac.ir

© Springer Nature Switzerland AG 2019 219


M. Tabatabaei and M. Aghbashlo (eds.), Biodiesel,
Biofuel and Biorefinery Technologies 8,
https://doi.org/10.1007/978-3-030-00985-4_10
220 R. Sharafi and G. S. Jouzani

re-sequencing of whole genomes of different organisms, identification of microbial


communities in different ecosystems (metagenomics studies), discovery of new
genes, transcription factors and other regulatory elements (promoters, enhancers, or
suppressors), microRNAs, and metabolic pathways (Morozova and Marra 2008;
Raupach et al. 2016).
The appendix “Omics” has been applied to various previously used biological
terms to create names for fields of endeavor like genome, transcriptome, proteome,
metabolomes, phenome, metagenome, etc. Accordingly, “Omics” technologies
include genomics, transcriptomics (gene expression profiling), proteomics, meta-
bolomics, metagenomics, meta-transcriptomics, meta-proteomics, and meta-
metabolomics which are known as high-throughput technologies. These sciences
are able to generate huge amounts of data about microorganisms, genes, regulatory
elements, RNAs, proteins, metabolites, and pathways in a short period of time and
at substantially lower costs than the traditionally used technologies (Debnath et al.
2010; Raupach et al. 2016; Schneider and Orchard 2011; Kircher and Kelso 2010;
Mardis 2013; van Dijk et al. 2014). In the following, we have tried to present a brief
definition for some important “Omics” methodologies (Fig. 10.1).

Fig. 10.1 Different “Omics” methodologies and their brief definitions


10 “Omics Technologies” and Biodiesel Production 221

10.1.1 Genomics

Genomics is the science of studying genomes (complete set of DNA within a single
cell) of different organisms. Commonly, genomics is focused on structure, function,
evolution, and mapping of genomes, and therefore, there are many subbranches,
such as structural genomics, functional genomics, evolutionary genomics, etc.

10.1.2 Transcriptomics

Commonly, transcriptome includes the set of all RNA molecules in one cell or a
population of cells. It is sometimes used to refer to all RNAs, or just mRNA,
depending on a particular experiment. Thus, transcriptomics is a large-scale study
of RNAs or study of all RNAs expressed by a cell or an organism.

10.1.3 Proteomics

Proteome includes the set of total proteins in one cell or a population of cells;
therefore, proteomics refers to the study of all proteins of an organism.

10.1.4 Metabolomics

The metabolome refers to all produced metabolites in a biological cell, tissue,


organ, or organism, as end products of cellular processes. Hence, metabolomics is
the systematic study of all or a part of metabolites (biochemical fingerprints).

10.1.5 Metagenomics

Metagenomics is the study of genetic materials of all or a group of microbial


communities recovered directly from environmental samples. The broad field may
also be referred to as environmental genomics, ecogenomics, or community
genomics. In this branch of science, studies are performed without isolation and
purification of microbes. As a result, the genetic materials (DNA of all microbiome)
are extracted directly from environmental samples.
222 R. Sharafi and G. S. Jouzani

10.1.6 Meta-transcriptomics

Meta-transcriptomics is the study of transcriptome of all microbial communities


recovered directly from environmental samples. While metagenomics reveals which
microbes are present in an environmental sample and what genomic potential they
might offer, meta-transcriptomics reveals about their activity at the transcription
level, e.g., the genes that are highly expressed in a specific microbial environment.

10.1.7 Meta-proteomics

Meta-proteomics refers to the study of all proteins recovered directly from envi-
ronmental samples at a specific time (Zampieri et al. 2016).

10.1.8 Meta-metabolomics

Meta-metabolomics is the study of all metabolites recovered directly from envi-


ronmental samples at a specific time.

10.2 Omics and Biodiesel Production: An Overview

Given the increasing energy demands and decreasing global fossil fuel reserves as a
result of world’s population growth, industry development on one hand and the
environmental challenges faced such as climate change and global warming on the
other hand, developing renewable platforms for fuel production is of significant
importance. These renewable platforms include solar, water, wind, nuclear energy,
and bioenergy. Biomass-based energy production has attracted a great deal of
attention because of the huge biomass production capacity globally (e.g., agricul-
tural and municipal wastes/residues) and the possibility to mitigate a considerable
fraction of the harmful environmental impacts of fossil fuels (Balaman and Selim
2015; Li-Beisson and Peltier 2013).
Liquid biofuels, such as bioethanol and biodiesel, have shown great potentials,
especially in the transportation systems. The first-generation biofuels include the
application of edible crops, such as rapeseed, sugarcane, sugar beet, and maize as
well as vegetable oils and animal fats. However, these biofuels have had significant
unfavorable impacts on the global food market and food security, especially in
countries with vulnerable economies, and therefore, their wide application has
generated a lot of controversies. In addition, increasing the current global area under
cultivation (14 million hectares) is impractical because of the abovementioned
10 “Omics Technologies” and Biodiesel Production 223

problems. The second-generation biofuels include fuel production from the biomass
of inedible plants, agricultural and forest harvesting residues, etc. The
third-generation biofuels are based on the application of photosynthetic microor-
ganisms such as microalgae. The problem with these two last generations is the lack
of efficient technologies for economic biofuel production (Tabatabaei et al. 2011;
Li-Beisson and Peltier 2013; Najafi et al. 2009; FAO Reports 2007, 2008; IEA
Report 2004).
Biodiesel is an oil-based biofuel containing a mixture of fatty acids alkyl esters
mainly produced by transesterification, and is well known as a renewable, nontoxic,
and biodegradable fuel. The physicochemical properties of biodiesel are very
similar to those of petroleum diesel; hence, it could be used as an alternative to
diesel in conventional engines without the need for any modifications. Higher
cetane number, higher flash point and lubricity, absence of sulfur, and lower aro-
matic content compared with the petroleum diesel are other useful attributes of
biodiesel (Demirbas 2009; Fjerbaek et al. 2009; Taher and Al‐Zuhair 2016).
Commonly, the first generation of biodiesel is generated from fatty acids
triglycerides found in oil crops, such as those from soybean, rapeseed, canola,
sunflower, and palm crops (Taher and Al-Zuhair 2016). Performing breeding
programs to improve oil yield and quality for biodiesel production is of importance
to reduce the input cost and also to increase the biodiesel production efficiency.
However, as mentioned earlier, application of such crops as source for biodiesel
production has negative impacts on food security, especially in the developing
countries. Therefore, concentration has to be placed on the development of
nonedible oilseeds as main bioenergy crops. The most important inedible oil plants
with potential use for biodiesel production include jatropha (physic nut: with 30–
35% extractable oil in seeds), castor beans, neem, mahua, cheura, camelina, and
pongamia. However, current oil yields from these plants are insufficient to meet the
bioenergy demands, and therefore, genetic improvements to enhance oil yield are of
importance (del Pilar Rodriguez et al. 2016). In fact, recent developments in NGS
technologies have encouraged their application in breeding programs in order to
enhance oil yield and quality (Fig. 10.2). In this context, different “Omics” tech-
nologies have been widely used to enhance the oil yield in oil crops (first-generation
biodiesel), inedible plants (second-generation biodiesel), and also in oil-producing
microorganisms (third-generation biodiesel).
Different genomic resources such as molecular markers, linkage maps, expressed
sequences tags (ESTs), and genome sequences are powerful tools to speed up the
genetic improvements in terms of oil yield and quality through marker-aided
selection (MAS) or genome sequencing (Yue et al. 2013). For instance, “Omics”
technologies could be used to identify genes involved in oil production, to discover
oil production pathways, as well as to improve oil quality and quantity using MAS
or genetic engineering programs (Fig. 10.2).
Molecular markers are known as a group of the most important “Omics” tech-
nologies used in plant improvements during the last decades. Molecular
marker-based techniques are helpful for the breeding programs of plants, especially
for MAS. The advantages of molecular markers compared with the phenotypic
224 R. Sharafi and G. S. Jouzani

Marker aided
selection for
more oil
production

Removing the Identification of


genes encoding genes involving
the toxins in oil
production
Omics
applications in
biodiesel
production

Cauterization of
Improving the
genetic
quality of oil for
diversity in oil
biodiesel
crops and
production
microorganisms
Genetic
engineering to
enhance oil
production

Fig. 10.2 Some applications of new “Omics” technologies to improve the quantity and quality of
oil production in oil crops and microorganisms for biodiesel production

markers are their high numbers and stability, detectability in all plant tissues
regardless of age, and their independency to the environmental conditions. During
the last three decades, a wide number of molecular markers, such as random
amplified polymorphic DNA (RAPD), sequence characterized amplified regions
(SCAR), restriction fragment length polymorphism (RFLP), simple sequence
repeats (SSR), inter-simple sequence repeat (ISSR), amplified fragment length
polymorphism (AFLP), and single-nucleotide polymorphism (SNP), have been
developed and used for the evaluation of genetic diversity, phylogenic studies, and
also for MAS programs of different oil crops and microorganisms (Agarwal et al.
2010; Ceasar and Ignacimuthu 2011). MAS and transgenic approaches have been
considered as the most promising solutions, which can be used to improve oil yield
and quality, and also to enhance the tolerance to biotic and abiotic stresses in
bioenergy crops. For instance, it has been proposed that genetic engineering of
identified candidate genes involved in oil biosynthesis, such as glycerol-3-
phosphate acyltransferase (GPAT), lysophosphatidic acid acyltransferase (LPAT),
10 “Omics Technologies” and Biodiesel Production 225

and diacylglycerol acyltransferase (DGAT), would enhance lipid content in seeds of


bioenergy crops. Moreover, through metabolic and genetic engineering, it would be
possible to overexpress the genes encoding proteins and enzymes involved in oil
production (Grover et al. 2013).
The following sections review the existing studies revolving around new
“Omics” technologies used to enhance oil and biodiesel production in Jatropha as a
promising bioenergy crop for biodiesel production.

10.2.1 Omics Studies in Physic Nut (Jatropha curcas)

Among biofuel plants, physic nut (J. curcas) has been known as a plant with high
potentials for biodiesel production, because of its high seed oil content, easy
propagation, rapid growth, short gestation period, and adaptation to a wide range of
agro-climatic conditions (Kumar and Sharma 2008; Wu et al. 2015). However,
J. curcas has not been domesticated for producing biodiesel. During the last dec-
ades, conventional breeding methods, based on phenotypic selection strategies,
have been used to increase the productivity of J. curcas, and some new cultivars
with improved oil yields have been obtained. It should be quoted that conventional
breeding has shown significant impacts on improving different traits, making
possible the release of cultivars possessing traits of interest, such as high yield, early
maturation, resistance to pests and diseases, and tolerance to drought and cold.
However, this procedure is slow, especially for perennial crops such as Jatropha
(Laviola et al. 2017). Therefore, different molecular-based methods, such as MAS
and genomic selection (GS) using a diverse range of molecular tools, such as DNA
markers, linkage maps, as well as genome, proteome, and transcriptome sequenc-
ing, have been developed and some are being used in genetic improvement for
sustainable production of biodiesel (Yue et al. 2013).

10.2.1.1 Conventional Molecular Markers

Some molecular markers have been successfully applied in Jatropha with a view to
characterize and improve the plant for enhanced oil production. During the last two
decades, many research groups have used different molecular markers to assess
genetic diversity and phylogenic relationship between a diverse number of Jatropha
species and genotypes, and also to identify the center of origin of this plant. The
molecular markers used include nrDNA-ITS (Pamidimarri and Reddy 2014), SNP
(Montes et al. 2014), AFLP and MS-AFLP (Mastan et al. 2016; Pamidimarri et al.
2008a, b; Tatikonda et al. 2009; Pioto et al. 2015), target region amplification
polymorphism (TRAP) (Franco et al. 2014), SSR (Raposo et al. 2014; Montes et al.
2014), ISSR (Basha and Sujatha 2007), RAPD (Verma et al. 2016; Pamidimarri and
Reddy 2014; Kole et al. 2015; Ram et al. 2008; Pamidimarri et al. 2008a, b) and
ISSR (Basha and Sujatha 2009; Kumar et al. 2008), and single primer amplification
226 R. Sharafi and G. S. Jouzani

reaction (SPAR) (Ranade et al. 2008). Ceasar and Ignacimuthu (2011) reviewed the
molecular markers used for studying genetic diversity among Jatropha species and
genotypes.
Ram et al. (2008) used RAPD markers to assess the genetic diversity of 12
species of Jatropha. In another study, a phylogenetic relationship among seven
species of Jatropha, namely, J. curcas, J. glandulifera, J. gossypifolia,
J. integerrima, J. multifida, J. podagrica, and J. tanjorensis, was established by
RAPD and AFLP techniques (Pamidimarri et al. 2008a). They could identify the
species with maximum relatedness (J. curcas and J. integerrima) and propose that
performing interspecies hybrid crosses between these two species would be suc-
cessful. In addition, RAPD and AFLP markers have been used to characterize toxic
and nontoxic varieties of J. curcas. Both markers showed genetic similarity
between nontoxic and toxic varieties about 0.92 (Pamidimarri et al. 2008b). In
another study, the genetic diversity of 42 accessions of J. curcas grown in different
regions of India and Mexico has been studied using RAPD and ISSR markers. The
genetic variation among the Indian germplasms was less than that of Mexican
genotypes, and some specific ISSR and two SCAR markers were identified for
Indian germplasm, which can be used as identification keys for them (Basha and
Sujatha 2007). An AFLP analysis was performed to assess the diversity in 48 elite
germplasm collections of J. curcas. The results confirmed that the accessions
coming from Andhra Pradesh have high diversity as they are placed in different
phylogenetic relations, whereas the accessions coming from Chhattisgarh, India
showed occurrence of high number of unique/rare fragments (Tatikonda et al.
2009). Recently, Pamidimarri and Reddy (2014) used RAPD, AFLP, and
nrDNA-ITS sequences to evaluate the intraspecific genetic diversity and to identify
and confirm the probable center of origin of J. curcas L. They showed that the
overall genetic diversity of J. curcas is narrow; however, the maximum and min-
imum genetic diversity was observed for collection of Mexico and India, respec-
tively. The high diversity of Mexican collection confirmed that this country may be
the center of origin and diversity of Jatropha. So, evaluation of genetic diversity of
Jatropha species is helpful for selection of parents for breeding programs and also
for collection, conservation, and characterization of Jatropha genetic resources in
future (Ceasar and Ignacimuthu 2011).

10.2.1.2 QTL Mapping and Linkage Map

During the process of genome-wide identification, construction of linkage map is


considered as the first step in order to find out associations between markers and
traits. To construct a linkage map and high-resolution mapping, it is necessary to
prepare one or several segregating populations, containing 4000–5000 individuals
with segregating DNA markers (Doerge 2002; Yue et al. 2013). Thus, having
populations with high genetic diversities is of importance. However, due to the very
low genetic diversity of J. curcas populations, preparation of high-resolution
linkage maps is very difficult. Due to this problem, recent works have focused on
10 “Omics Technologies” and Biodiesel Production 227

generation of new families using interspecies crosses (Wang et al. 2011a, b; Yue
et al. 2013). Using interspecific crosses between J. curcas and J. integerrima,
quantitative traits loci (QTL) mapping programs were performed and as a result,
some QTLs linked to seed weight, fatty acid composition, and vegetative growth
characteristics were characterized (Liu et al. 2011; Sun et al. 2012). In another
work, this kind of interspecies cross (J. curcas  J. integerrima) and backcross
were used to construct linkage maps. To achieve this, 506 codominant DNA
markers (SSRs and SNPs) were identified in the segregating population, and 11
linkage groups were constructed, which could be used for QTL mapping of agro-
nomic traits, MAS, as well as for cloning genes responsible for phenotypic varia-
tions (Wang et al. 2012). King et al. (2015) performed a QTL analysis for a number
of agronomic traits (i.e., plant height, stem diameter, number of branches, total
seeds per plant, 100-seed weight, seed oil content, and fatty acid composition) in
two mapping populations of J. curcas. Based on the data obtained from physical
mapping derived from genome sequencing of Jatropha, and using the candidate
genes approach, they could identify and localize a large number of genes (such as
phosphatidate phosphatase genes contributing to the biosynthesis of storage lipids)
onto the genetic map. Recently, Clarke (2016) identified and mapped QTLs for seed
oil content, seed oil composition (oil quality), and oil yield in J. curcas by
designing SNP markers using SSR markers derived from genome sequencing data.
The identified QTLs belonged to seed oil content, seed oil composition (palmitate,
stearate, oleate, and linoleate contents), seed weight, number of branches, and seed
yield. Therefore, based on these reports, it could be concluded that the application
of QTL strategies could help breeders in the future with improving the traits of
interest, especially those related to oil yield and quality.

10.2.1.3 Genome Sequencing and Genome-Wide Selection (GWS)

During the last decade, the genomes of some bioenergy crops with potentials for
biodiesel production have been sequenced. Examples include Glycine max (soy-
bean) (Schmutz et al. 2010), Brassica rapa (Wang et al. 2011a, b), Phoenix
dactylifera (date palm) (Al-Dous et al. 2011), Ricinus communis L. (castor bean)
(Chan et al. 2010), J. curcas (Sato et al. 2011; Wu et al. 2015), Hevea brasiliensis,
Elaeis guineensis (oil palm) (Singh et al. 2013; Pootakham et al. 2015), Lindera
glauca (Lin et al. 2017), Camelina sativa (Mudalkar et al. 2014), Pongamia pinnata
(Jiang et al. 2017), Helianthus annuus (sunflower) (Staton et al. 2012), and
Arabidopsis (Cao et al. 2011). Such huge deal of data could open up new ways to
discover genes and pathways involved in oil production, and also genes affecting
the quality of oil in these plants.
Whole-genome sequencing and the development of genetic maps of J. curcas
are important components in molecular breeding and genetic improvement of this
plant. The genome of J. curcas has been sequenced by several research institutes
and companies (Yue et al. 2013). Sato et al. (2011) were the first to publish some
information about the genome sequence of Jatropha. Then, Hirakawa et al. (2012)
228 R. Sharafi and G. S. Jouzani

reported a new assembly of Jatropha genome (with a total length of 297.7 Mb


consisting of 39,277 contigs) by integrating de novo assembly of a total of 537
million paired-end reads from the Illumina sequencing platform into the previous
genome assembly (Hirakawa et al. 2012; Yue et al. 2013). Later, Wu et al. (2015)
reported the draft genome of J. curcas L with a total length of 320.5 Mbp con-
taining 27 172 putative protein-coding genes. They established a linkage map
containing 1208 markers while also identified about 5268 gene families (such as
those encoding ribosome-inactivating proteins and oil biosynthesis pathway
enzymes), out of which 13887 gene families had been previously identified in the
castor bean genome (Wu et al. 2015). These reports confirmed that the genomics
tools could make a significant contribution to the acceleration of genetic
improvements in Jatropha for increasing oil yield and quality in the near future.
Genome-wide selection (GWS) or genomic selection (GS) is a new form of
MAS breeding which uses high‐throughput genotyping and whole‐genome
molecular markers to improve the quantitative and complex traits (polygenic traits)
in large segregating populations, and has been applied to multiple crops, including
wheat, maize, and barley. GWS has attracted a great deal of attention in plant and
animal breeding programs due to its high efficiency in performance prediction of
offspring in the segregating populations by associating marker information with
phenotypic information. The GWS strategy is different from the traditional MAS, as
it does not use QTL mapping and a test of significant markers, and all markers are
randomly modeled and included in both the training and validation populations of
the GWS model (de Azevedo Peixoto et al. 2017; Meuwissen et al. 2001; Lorenz
et al. 2011, 2012; Burgueño et al. 2012; Heslot et al. 2012).
Recently, de Azevedo Peixoto et al. (2017) used modeling and GWS in Jatropha
to estimate the genetic parameters for grain yield (GY) and the weight of 100 seeds
(W100S) using restricted maximum likelihood (REML). Their results demonstrated
the applicability of genome-wide prediction to identify useful genetic sources of
GY and W100S for Jatropha breeding.

10.2.1.4 Other New “Omics”-Based Molecular Markers

Recently, thanks to the “Omics”-based studies performed on Jatropha, a large


number of EST sequences of this plant have been made available in the public
domain. This opportunity could open up new ways for the identification of can-
didate genes for several agronomic characteristics, including oil content in seeds
(Grover et al. 2014). De novo assembly of transcript sequences produced by NGS
technologies offers an efficient and rapid methodology to identify novel expressed
genes (Laosatit et al. 2016). Transcriptome studies are of importance as well and
could lead to a better understanding of the function and biological process of
genomes.
Several projects have been performed to sequence expressed sequence tags
(ESTs) of J. curcas in order to detect SNP marker for this plant (Yue et al. 2013;
Moniruzzaman et al. 2016). By sequencing and evaluation of more than 30,000
10 “Omics Technologies” and Biodiesel Production 229

ESTs from developing and germinating seeds, Costa et al. (2010a, b) could find
some SNPs to construct a linkage map of Jatropha. They recognized many genes
involved in lipid synthesis, the degradation pathway, and seed toxicity. King et al.
(2015) sequenced the transcriptome of developing seeds of Jatropha using 454
high-throughput sequencing and generated 46 Mb of raw sequence data, including
95,692 sequences using a single sequencing run. Natarajan and Parani (2011)
developed cDNAs from developing seeds, flowers, roots, embryos, and mature
leaves of J. curcas using 454 pyrosequencings. In another work, Silva-Junior et al.
(2011) identified a total of 18,225 SNPs in 11.9 giga bases of J. curcas by
sequencing pooled samples. Another research group detected 2482 informative
SNPs by sequencing 148 global collections of J. curcas lines and found that a
narrow level of genetic diversity existed among the indigenous genotypes as
compared with the exotic genotypes of J. curcas (Gupta et al. 2012). In a different
work, 9844 unique sequences (1070 contigs and 3595 singletons) were detected by
sequencing ESTs derived from cDNA libraries of Jatropha (Chen et al. 2011).
Some genes related to the fatty acid biosynthesis pathway have also been recog-
nized by construction of full-length enriched cDNA library from developing seeds
(Yadav et al. 2011). Gu et al. (2012) sequenced over 50,000 EST clones by con-
structing several cDNA libraries from different tissues using Sanger sequencing.
Grover et al. (2014) analyzed about 43,000 ESTs of Jatropha for microsatellites
and fatty acid metabolism related sequences. They showed that about 8% of the
unigenes revealed the presence of microsatellites, and a total of 931 sequences
participated in the pathways related to fatty acid or lipid metabolism. Recently,
Laosatit et al. (2016) sequenced the transcriptomes of apical meristem tissues of
two J. curcas accessions and four Jatropha-related species. The results obtained
indicated that 1,683 unique proteins were orthologous across all four Jatropha spp.
They also identified 269 EST-SSR markers, out of which 20 candidate SNPs were
identified in four coding sequences including one gene relating to biosynthesis
pathways of phorbol esters. Laviola et al. (2018) identified more than 4007 DArTs
(dominant) and 747 polymorphic SNPs (codominant) markers in the genotyping
process of different Jatropha genotypes. These results opened new possibilities for
the operational use of markers for selection for quantitative traits, and mainly for the
reduction of the time necessary to select superior Jatropha genotypes for breeding
programs, consequently accelerating the domestication of the species (Laviola et al.
2017). Achieving an efficient reference transcriptome (all transcripts, coding and
noncoding, large and small RNA) is of importance for transcriptomics studies.
Therefore, some works have been focused on searching for a complete reference
transcriptome of J. curcas (Moniruzzaman et al. 2016).
Identifying Candidate Genes for Important Traits
All “Omics” technologies, such as genomics, transcriptomics, and proteomics, can
be used to identify the candidate genes for target traits. For instance, using tran-
scriptome analysis, it is possible to find and characterize ESTs and their source
genes. It is also possible to develop microarray systems to study expression profiles
of different genes in different tissues at different developing stages. During the last
230 R. Sharafi and G. S. Jouzani

decade, researchers could identify and characterize genes involved in lipid synthesis
pathways, tolerance to biotic and abiotic stresses, as well as other important traits
using Omics technologies, such as cDNA libraries and RNA-seq analysis (Tang
et al. 2007; Liu et al. 2017; Tian et al. 2016).
Previously, the JcERF gene, which is an ERF subfamily member, was isolated
from Jatropha, and functionally characterized as a putative AP2/EREBP domain
containing transcription factor (Tang et al. 2007). Fatty acyl carrier thioesterase A
(FATA) plays an important role in the process of fatty acid synthesis metabolism.
Liu et al. (2017) studied the expression pattern of JcFATA gene using
semi-quantitative RT-PCR and qRT-PCR. Their results showed that JcFATA gene
could express in multiple tissues of J. curcas, and is especially highly expressed in
roots, flowers, and leaves. Expression of the gene under the control of the JcFATA
gene promoter in Arabidopsis thaliana confirmed the high expression of the protein
in roots, flowers, and leaves. It is well known that the seed size has an important
role in oil yield of Jatropha, and therefore, it has been taken into account as the
main goal of Jatropha breeding to enhance its oil yield. However, the genetic
regulation of this trait has not been fully understood in Jatropha. Recently, Tian
et al. (2016) cloned CYP78A98 gene from Jatropha. They showed that the gene
was highly expressed in different tissues, such as seed, apex, stem, and leaf. The
protein was accumulated in endoplasmic reticulum (ER), and its N-terminus
hydrophobic domain was essential for the correct protein localization in ER.
Subsequently, they overexpressed the gene in tobacco, which resulted in an increase
in seed size, weight, as well as protein and fatty acid contents (Tian et al. 2016).
Biodiesel quality is highly dependent on the fatty acid composition of seed oil.
For instance, 16- to 18-carbon fatty acids ratio has a direct impact on biodiesel
quality and its combustion performance. Commonly, the length of fatty acid chains
is determined by the plastidial fatty acid synthase enzymes (three b-ketoacyl-acyl
carrier protein synthases (KASI, II and III)). Yu et al. (2015) could functionally
characterize a putative JcKASIII gene from J. curcas. They confirmed that this gene
plays an important role in the regulation of fatty acid composition, as its expression
in Arabidopsis resulted in a significant increase of palmitic acid and increased the
16- to 18-carbon fatty acids ratio. In another work, Yang et al. (2016) performed a
comparative transcriptomic study with seeds of Vernicia and Jatropha at the initial
and fast stages of oil accumulation to characterize gene regulation during the oil
accumulation process. Transcriptome analysis revealed a number of species-specific
fatty acid desaturases (FAD2, FADX, FAH12, etc.), oleocins, and PDAT proteins
involved in or regulating fatty acid compositions in seeds of Vernicia and Jatropha.
The b-ketoacyl-acyl carrier protein synthase I (KASI) is involved in de novo
fatty acid biosynthesis in many organisms. Xiong et al. (2017) isolated two putative
KASI genes from J. curcas. Both JcKASI genes were expressed in multiple tissues,
most strongly in filling stage seeds of J. curcas. Additionally, the proteins were
both localized to the plastids. Expressing JcKASI-1 in the Arabidopsis kasI mutant
rescued the mutant’s phenotype and restored the fatty acid composition and oil
content in seeds to wild type, but expressing JcKASI-2 in the Arabidopsis kasI
mutant resulted in only a partial rescue. This implies that JcKASI-1 and JcKASI-2
10 “Omics Technologies” and Biodiesel Production 231

exhibit partial functional redundancy and that the KASI genes play a universal role
in regulating fatty acid biosynthesis, growth, and development in plants. In another
work, to identify novel genes expressed during stress in J. curcas, cDNA library
constructed from salt-stressed roots was screened and 32 full-length genes that
could confer abiotic stress tolerance were obtained (Eswaran et al. 2010). Other
genes, including JcDof1 DNA-binding and transcriptional activation activities,
flowering-time-related Dof transcription factor gene (Yang et al. 2010), accA,
accB1, accC, and accD genes encoding the subunits of heteromeric ACCase (Gu
et al. 2011), and stearoyl-acyl carrier protein desaturase (SAD) (Tong et al. 2006),
have also been isolated and characterized in J. curcas. In another work by Gu et al.
(2012), 68 fatty acids and lipid biosynthetic genes were identified using cDNA
library of developing endosperm.
Curcin, a type 1 ribosome-inactivating protein, is the major toxic protein found
in Jatropha seeds. The cDNA of the gene curcin2 was obtained from J. curcas
leaves, which were affected by drought, temperature stress, and fungal infection
(Tong et al. 2007). Recently, three curcin genes have also been isolated and
characterized from J. curcas MD44. These results provided useful information and
research materials for further functional study of curcin proteins and genetic
engineering of J. curcas (Wu et al. 2017). As Jatropha is known as a drought
tolerant plant, gene discovery in this plant will be very useful for providing a source
of genetic information for achieving abiotic stress tolerance improvements in other
crops. Recently, a newly developed microarray system was used to identify genes
involved in Jatropha drought tolerance and about 1300 genes contributing to
dehydration and/or recovery were characterized (Cartagena et al. 2015).
Genetic Engineering to Enhance Oil Production
Jatropha can be multiplied by vegetative propagation as well as through seeds.
Thus, different breeding strategies, including clone breeding, line breeding, hybrid
breeding, and population breeding, have been used for its domestication and
breeding to enhance oil yield and other economic traits. In addition, some sup-
porting techniques such as recurrent backcrossing, marker assisted and genomic
selection, mutation breeding, and genetic engineering can be used in the framework
of the four abovementioned breeding strategies (Montes and Melchinger 2016).
Efficient application of genetic engineering for Jatropha improvement requires elite
genotypes with optimized tissue and organ culture and also transformation proto-
cols. Therefore, currently the majority of the research works in this field are focused
on the optimization of tissue culture and genetic engineering protocols and also on
the identification of the genes associated with important traits of Jatropha (Jaganath
et al. 2014; Montes and Melchinger 2016; Xiong et al. 2017; Yang et al. 2016).
During the last decade, different strategies and protocols based on
Agrobacterium-mediated and microprojectile bombardment transformation have
been developed for this plant. For instance, Agrobacterium-mediated cotyledon disc
transformation (Li et al. 2008; Pan et al. 2010), Agrobacterium-mediated leaf
explants transformation (Kumar et al. 2010), Agrobacterium-mediated transfor-
mation using leaf and hypocotyl segments (Misra et al. 2012), Agrobacterium-
232 R. Sharafi and G. S. Jouzani

mediated transformation using vacuum infiltration and vegetative in vivo cleft


grafting method without the aid of tissue culture (Jaganath et al. 2014; Nanasato
et al. 2015), and microprojectile bombardment using embryo axes (Purkayastha
et al. 2010; Joshi et al. 2011) have been optimized for this purpose.
A proper oil for biodiesel production should have favorable oxidative stability
and cold flow properties; so, it is necessary that the produced oil contain more
monounsaturated fatty acids (oleic and palmitoleic acids), less polyunsaturated
acids, and less saturated acids. New insights into the genes involved in Jatropha oil
biosynthesis and metabolism as well as transcriptional control are beginning to
unfold biotechnological enhancement of oil content and quality by direct genetic
engineering (Ye et al. 2013; Moniruzzaman et al. 2016). Three key enzymes, i.e.,
diacylglycerol acyl transferase, lysophosphatidic acid acyltransferase, and
glycerol-3-phosphate acyltransferase, play important roles in lipid biosynthesis
pathway. Some previous research works have shown that the overexpression of
these enzymes could increase the seed oil content in Brassica and Arabidopsis
(Sharma et al. 2008; Maisonneuve et al. 2010; Jain et al. 2000). Hence, it is possible
to increase the oil content and enhance the seed oil composition in Jatropha.
The EST analysis in Jatropha has shown that different enzymes, such as carboxyl
transferase of the ACCase b subunit, the biotin carboxyl carrier protein of ACCase,
acyl ACP thioesterase A, lysophosphatidic acid acyl transferase, acyl-CoA binding
protein, malonyl-CoA:ACP transacylase, glycerol-3 phosphate acyl transferase,
beta-keto acyl ACP synthase II, 3-keto acyl ACP reductase, beta-keto acyl ACP
synthase I, acyl carrier protein x-3-fatty acid desaturase, x-6-fatty acid desaturase,
and long-chain acyl-CoA synthetase, are involved in lipid biosynthesis pathways in
Jatropha (Natarajan et al. 2010; Moniruzzaman et al. 2016). Accordingly, these
genes of the beta-oxidation pathway can be good sources for oil yield and quality
engineering. In addition, identification, cloning, and transformation of the genes
involved in Jatropha tolerance to biotic (diseases and pests) and also abiotic
stresses (e.g., drought, salinity, and cold) are of importance in breeding programs of
this plant.
It should be noted that the application of genetic engineering for Jatropha
breeding is currently at starting phases. Nevertheless, some successful works have
been reported during the last decade. It is well known that the enzyme FAD2 is a
key enzyme contributing to the linoleic acid biosynthesis in plants. Qu et al. (2012)
using genome-wide analysis could identify three putative fatty acid desaturase
genes in Jatropha (JcFAD2s). They down-regulated the expression of one of these
genes (JcFAD2-1) using RNAi technology and as a result, the oleic acid contents of
the achieved transgenic plants were significantly increased (>78%) against that of
the parent. This down-regulation also reduced the polyunsaturated fatty acids
contents (<3%). It is well known that SDP1 is a specific lipase for the first step of
TAG catabolism in Arabidopsis, and its deficiency causes more oil accumulation in
plants. Kim et al. (2014) generated JcSDP1 deficiency in transgenic Jatropha using
RNAi technology, which resulted in 13–30% higher total seed storage lipid and 7%
compensatory decrease in protein content. Misra et al. (2012) isolated the JcDGAT1
gene encoding a key enzyme contributing to fatty acids biosynthesis pathway from
10 “Omics Technologies” and Biodiesel Production 233

J. curcas seeds. The transgenic Arabidopsis plants showed a gradual increase in


total oil content from early seed development to maturation (by up to 30–41%)
without any phenotypic changes.
It should be quoted that despite the promising application of Jatropha as biofuel
crop, its seeds contain phorbol esters and curcins, which are known as
skin-irritating and cancer-promoting toxins and reduce nutritional and commercial
value of the seeds and oil obtained from this plant (Gu et al. 2015). Recently, some
works successfully used RNAi technology to reduce curcins in Jatropha, and could
reduce the toxin content by up to 98% (Patade et al. 2014; Gu et al. 2015). Li et al.
(2016) used a specific RNAi technology to down-regulate a phorbol ester gene
(JcCASA: casbene synthase gene), and could effectively reduce the gene expression,
which caused a final reduction of phorbol esters level in seeds by about 85%.
Other transgenic Jatropha plants have also been developed recently for other
agronomic traits, such as salinity tolerance (Jha et al. 2013; Yang et al. 2014),
drought tolerance (Tsuchimoto et al. 2012), freezing tolerance (Wang et al. 2015),
pest resistance using Bacillus thuringiensis cry genes (Gu et al. 2014), fungal
disease resistance (Franco et al. 2016), and geminivirus resistance (Ye et al. 2014).
Finally, it is important to note that as Jatropha is a nonedible plant, therefore, the
majority of the biosafety issues concerning GM crops and human health could be
avoided. In better words, the application of genetic engineering for molecular
breeding of this plant would be more applicable compared with edible biofuel
crops.

10.3 Conclusions

Recent revolutionary developments in NGS and “Omics” technologies have pro-


vided excellent opportunities to accelerate and promote breeding programs for
different organisms. In this context, different “Omics” technologies have been
widely used to enhance oil yield, oil quality, and tolerance to biotic and abiotic
stresses in bioenergy crops, such as Jatropha. Oil-rich seeds of Jatropha have a
high potential to contribute to sustainable production of biodiesel and reduction of
atmospheric carbon dioxide. However, in spite of the tremendous progress in
Jatropha domestication and breeding using conventional breeding strategies, the
yield of oil produced by the seeds of this plant is not suitable and economic at this
stage yet, and therefore, lots of crop improvement efforts would be required in the
future. It could be anticipated that the application of MAS, GWS, and genetic
engineering would significantly assist researchers in overcoming the challenges
faced in biodiesel production from Jatropha.
234 R. Sharafi and G. S. Jouzani

References

Agarwal M, Shrivastava N, Padh H (2010) Advances in molecular marker techniques and their
applications in plant sciences. Plant Cell Rep 27:617–631
Al-Dous EK, George B, Al-Mahmoud ME, Al-Jaber MY, Wang H, Salameh YM, Al-Azwani EK,
Chaluvadi S, Pontaroli AC, DeBarry J, Arondel V (2011) De novo genome sequencing and
comparative genomics of date palm (Phoenix dactylifera). Nat Biotechnol 29(6):521–527
Balaman ŞY, Selim H (2015) Biomass to energy supply chain network design: An overview of
models, solution approaches and applications. In: Handbook of bioenergy. Springer
International Publishing, pp 1–35
Basha SD, Sujatha M (2007) Inter and intra population variability of Jatropha curcas (L.)
characterized by RAPD and ISSR markers and development of population specific SCAR
markers. Euphytica 156:375–386
Basha SD, Sujatha M (2009) Genetic analysis of Jatropha species and interspecific hybrids of
Jatropha curcas using nuclear and organelle specific markers. Euphytica 168:197–214
Burgueño J, de los Campos G, Weigel K, Crossa J (2012) Genomic prediction of breeding values
when modeling genotype  environment interaction using pedigree and dense molecular
markers. Crop Sci 52(2):707–719
Cartagena JA, Seki M, Tanaka M, Yamauchi T, Sato S, Hirakawa H, Tsuge T (2015) Gene
expression profiles in Jatropha under drought stress and during recovery. Plant Mol Biol Rep
33(4):1075–1087
Cao J, Schneeberger K, Ossowski S, Günther T, Bender S, Fitz J, Koenig D, Lanz C, Stegle O,
Lippert C, Wang X (2011) Whole-genome sequencing of multiple Arabidopsis thaliana
populations. Nat Genet 43(10):956–963
Chan AP, Crabtree J, Zhao Q, Lorenzi H, Orvis J, Puiu D, Melake-Berhan A, Jones KM,
Redman J, Chen G, Cahoon EB (2010) Draft genome sequence of the oilseed species Ricinus
communis. Nat Biotechnol 28(9):951–956
Chen MS, Wang GJ, Wang RL, Wang J, Song SQ, Xu ZF (2011) Analysis of expressed sequence
tags from biodiesel plant Jatropha curcas embryos at different developmental stages. Plant Sci
181(6):696–700
Ceasar SA, Ignacimuthu S (2011) Applications of biotechnology and biochemical engineering for the
improvement of Jatropha and biodiesel: a review. Renew Sustain Energy Rev 15(9):5176–5185
Clarke J (2016) Quantitative trait locus mapping of oil yield and oil quality related traits in the
biofuel crop Jatropha curcas. Doctoral dissertation, University of York
Costa GGL, Cardoso KC, Del Bem LEV, Lima AC, Cunha MAS, de Campos-Leite L et al (2010)
Transcriptome analysis of the oil-rich seed of the bioenergy crop Jatropha curcas L. BMC
Genom 11
Costa GG, Cardoso KC, Del Bem LE, Lima AC, Cunha MA, de Campos-Leite L, Vicentini R,
Papes F, Moreira RC, Yunes JA, Campos FA (2010b) Transcriptome analysis of the oil-rich
seed of the bioenergy crop Jatropha curcas L. BMC Genom 11(1):462
de Azevedo Peixoto L, Laviola BG, Alves AA, Rosado TB, Bhering LL (2017) Breeding Jatropha
curcas by genomic selection: a pilot assessment of the accuracy of predictive models.
PLoS ONE 12(3):e0173368
Debnath M, Pandey M, Malik CP (2010) Analysing the potentiality of Jatropha using “Omics
Technology”. J Pl Sci Res 26(1):29–51
del Pilar Rodriguez M, Brzezinski R, Faucheux N, Heitz M (2016) Enzymatic transesterification of
lipids from microalgae into biodiesel: a review
Demirbas A (2009) Progress and recent trends in biodiesel fuels. Energy Convers Manag 50:14–34
Doerge RW (2002) Mapping and analysis of quantitative trait loci in experimental populations.
Nat Rev Genet 3(1):43–52
Eswaran N, Parameswaran S, Sathram B, Anantharaman B, Kumar GRK, Tangirala SJ (2010)
Yeast functional screen to identify genetic determinants capable of conferring abiotic stress
tolerance in Jatropha curcas. BMC Biotechnol 10:23
10 “Omics Technologies” and Biodiesel Production 235

FAO (2007) Sustainable bioenergy: a framework for decision makers. United Nations Energy
FAO (2008) The state of food and agriculture 2008. Biofuels: prospects, risks and opportunities.
http://www.fao.org/publications/sofa-2008/en
Fjerbaek L, Christensen KV, Norddahl B (2009) A review of the current state of biodiesel
production using enzymatic transesterification. Biotechnol Bioeng 102(5):1298–1315
Franco MC, Gomes KA, de Carvalho Filho MM, Harakava R, Carels N, Siqueira WJ, Latado RR,
de Argollo Marques D (2016) Agrobacterium-mediated transformation of Jatropha curcas leaf
explants with a fungal chitinase gene. Afr J Biotech 15(37):2006–2016
Franco MC, Marques DDA, Siqueira WJ, Latado RR (2014) Micropropagation of Jatropha curcas
superior genotypes and evaluation of clonal fidelity by target region amplification polymor-
phism (TRAP) molecular marker and flow cytometry. Afr J Biotechnol 13(38)
Grover A, Kumari M, Singh S, Rathode SS, Gupta SM, Pandey P, Gilotra S, Kumar D, Arif M,
Ahmed Z (2014) Analysis of Jatropha curcas transcriptome for oil enhancement and genic
markers. Physiol Mol Biol Plants 20(1):139–142
Grover A, Patade VY, Kumari M, Gupta SM, Arif M, Ahmed Z (2013) Omics approaches in
biofuel production for a green environment. In: Press CRC (ed) OMICS: applications in
biomedical, agricultural, and environmental sciences. Taylor and Francis Group, Boca Raton,
pp 623–636
Gu K, Yi C, Tian D, Sangha JS, Hong Y, Yin Z (2012) Expression of fatty acid and lipid
biosynthetic genes in developing endosperm of Jatropha curcas. Biotechnol Biofuels 5:47
Gu KY, Chiam H, Tian DS, Yin ZC (2011) Molecular cloning and expression of heteromeric
ACCase subunit genes from Jatropha curcas. Plant Sci 180(4):642–649
Gu K, Mao H, Yin Z (2014) Production of marker-free transgenic Jatropha curcas expressing
hybrid Bacillus thuringiensis d-endotoxin Cry1Ab/1Ac for resistance to larvae of tortrix moth
(Archips micaceanus). Biotechnol Biofuels 7(1):68
Gu K, Tian D, Mao H, Wu L, Yin Z (2015) Development of marker-free transgenic Jatropha
curcas producing curcin-deficient seeds through endosperm-specific RNAi-mediated gene
silencing. BMC Plant Biol 15(1):242
Gupta P, Idris A, Mantri S, Asif MH, Yadav HK, Roy JK et al (2012) Discovery and use of single
nucleotide polymorphic (SNP) markers in Jatropha curcas L. Mol Breed 30(3):1325–1335
Heslot N, Yang H-P, Sorrells ME, Jannink J-L (2012) Genomic selection in plant breeding: a
comparison of models. Crop Sci 52(1):146–160
Hirakawa H, Tsuchimoto S, Sakai H, Nakayama S, Fujishiro T, Kishida Y et al (2012) Upgraded
genomic information of Jatropha curcas L. Plant Biotechnol 29(2):123–130
IEA (2004) Biofuels for transport, International Energy Agency. http://www.iea.org/textbase/
nppdf/free/2004/biofuels2004.pdf. Accessed 17 July 2010
Jaganath B, Subramanyam K, Mayavan S, Karthik S, Elayaraja D, Udayakumar R,
Manickavasagam M, Ganapathi A (2014) An efficient in planta transformation of Jatropha
curcas (L.) and multiplication of transformed plants through in vivo grafting. Protoplasma 251
(3):591–601
Jain R, Coffey M, Lai K, Kumar A, MacKenzie S (2000) Enhancement of seed oil content by
expression of glycerol-3-phosphate acyltransferase genes. Biochem Soc Trans 28:959–960
Jha B, Mishra A, Jha A, Joshi M (2013) Developing transgenic Jatropha using the SbNHX1 gene
from an extreme halophyte for cultivation in saline wasteland. PLoS ONE 8(8):e71136
Jiang Q, Yen SH, Stiller J, Edwards D, Scott PT, Gresshoff PM (2017) Genetic, biochemical, and
morphological diversity of the legume biofuel tree Pongamia pinnata. Plant Genet Genom
Biotechnol 1(3):54–67. ISSN 2332-2012
Joshi M, Mishra A, Jha B (2011) Efficient genetic transformation of Jatropha curcas L. by
microprojectile bombardment using embryo axes. Ind Crops Prod 33(1):67–77
Kim MJ, Yang SW, Mao HZ, Veena SP, Yin JL, Chua NH (2014) Gene silencing of
Sugar-dependent 1 (JcSDP1), encoding a patatin-domain triacylglycerol lipase, enhances seed
oil accumulation in Jatropha curcas. Biotechnol Biofuels 7(1):36
236 R. Sharafi and G. S. Jouzani

King AJ, Montes LR, Clarke JG, Itzep J, Perez CA, Jongschaap RE, Visser RG, van Loo EN,
Graham IA (2015) Identification of QTL markers contributing to plant growth, oil yield and
fatty acid composition in the oilseed crop Jatropha curcas L. Biotechnol Biofuels 8(1):160
Kircher M, Kelso J (2010) High-throughput DNA sequencing—concepts and limitations.
BioEssays 32:524–536
Kole PR, Bhat KV, Chaudhury R, Malik SK (2015) Genetic variation among Jatropha curcas L.
using dominant molecular marker collected from different agro-climatic regions of India.
Indian J Genet 75(2):267–270
Kumar RS, Parthiban KT, Rao MG (2008) Molecular characterization of Jatropha genetic
resources through inter-simple sequence repeat [ISSR] markers. Mol Biol Rep 36:1951–1956
Kumar N, Anand KV, Pamidimarri DS, Sarkar T, Reddy MP, Radhakrishnan T, Kaul T,
Reddy MK, Sopori SK (2010) Stable genetic transformation of Jatropha curcas via
Agrobacterium tumefaciens-mediated gene transfer using leaf explants. Ind Crops Prod 32
(1):41–47
Laosatit K, Tanya P, Somta P, Ruang-Areerate P, Sonthirod C, Tangphatsornruang S,
Juntawong P, Srinives P (2016) De novo transcriptome analysis of apical meristem of
Jatropha. Plant Mol Biol Rep 34(4):786–793
Laviola BG, Rodrigues EV, Teodoro PE, de Azevedo Peixoto L, Bhering LL (2017) Biometric and
biotechnology strategies in Jatropha genetic breeding for biodiesel production. Renew Sustain
Energy Rev 76:894–904
Laviola BG, Alves AA, Rosado TB, Bhering LL, Formighieri EF, de Azevedo Peixoto L (2018)
Establishment of new strategies to quantify and increase the variability in the Brazilian
Jatropha genotypes. Ind Crop Prod 117:216–223
Li C, Ng A, Xie L, Mao H, Qiu C, Srinivasan R, Yin Z, Hong Y (2016) Engineering low phorbol
ester Jatropha curcas seed by intercepting casbene biosynthesis. Plant Cell Rep 35(1):103–114
Li M, Li H, Jiang H, Pan X, Wu G (2008) Establishment of an Agrobacteriuim-mediated
cotyledon disc transformation method for Jatropha curcas. Plant Cell Tissue Organ Cult 92
(2):173–181
Li-Beisson Y, Peltier G (2013) Third-generation biofuels: current and future research on
microalgal lipid biotechnology. OCL 20(6):D606
Lin Z, An J, Wang J, Niu J, Ma C, Wang L, Yuan G, Shi L, Liu L, Zhang J, Zhang Z (2017)
Integrated analysis of 454 and Illumina transcriptomic sequencing characterizes carbon flux
and energy source for fatty acid synthesis in developing Lindera glauca fruits for woody
biodiesel. Biotechnol Biofuels 10(1):134
Liu P, Wang C, Li L, Sun F, Liu P, Yue G (2011) Mapping QTLs for oil traits and eQTLs for
oleosin genes in Jatropha. BMC Plant Biol 11(1):132
Liu Y, Yang Y, Yin X, Li L, Zhu H, Lu J, Shi Y (2017) Expression of JcFATA gene in Jatropha
curcas and its promoter cloning and analysis. J Agric Biotechnol 25(2):214–221
Lorenz AJ, Chao S, Asoro FG, Heffner EL, Hayashi T, Iwata H et al (2011) Genomic selection in
plant breeding: knowledge and prospects. Adv Agron 110:77
Lorenz AJ, Smith KP, Jannink J-L (2012) Potential and optimization of genomic selection for
Fusarium head blight resistance in six-row barley. Crop Sci 52(4):1609–1621
Maisonneuve S, Bessoule J-J, Lessire R, Delseny M, Roscoe TJ (2010) Expression of rapeseed
microsomal lysophosphatidic acid acyltransferase isozymes enhances seed oil content in
Arabidopsis. Plant Physiol 152:670–684
Mardis ER (2013) Next-generation sequencing platforms. Annu Rev Anal Chem 6:287–303
Mastan SG, Rathore MS, Ghosh A (2016) Molecular characterization of genetic and epigenetic
divergence in selected Jatropha curcas L. germplasm using AFLP and MS-AFLP markers.
Plant Gene 8:42–49
Meuwissen THE, Hayes BJ, Goddard ME (2001) Prediction of total genetic value using
genome-wide dense marker maps. Genetics 157(4):1819–1829. pmid:11290733
Misra P, Toppo DD, Mishra MK, Saema S, Singh G (2012) Agrobacterium tumefaciens-mediated
transformation protocol of Jatropha curcas L. using leaf and hypocotyl segments. J Plant
Biochem Biotechnol 21(1):128–133
10 “Omics Technologies” and Biodiesel Production 237

Moniruzzaman M, Yaakob Z, Khatun R (2016) Biotechnology for Jatropha improvement: a


worthy exploration. Renew Sustain Energy Rev 54:1262–1277
Montes JM, Melchinger AE (2016) Domestication and Breeding of Jatropha curcas L. Trends
Plant Sci 21(12):1045–1057
Montes JM, Technow F, Martin M, Becker K (2014) Genetic diversity in Jatropha curcas L.
assessed with SSR and SNP markers. Diversity 6(3):551–566
Morozova O, Marra MA (2008) Applications of next-generation sequencing technologies in
functional genomics. Genomics 92(5):255–264
Mudalkar S, Golla R, Ghatty S, Reddy AR (2014) De novo transcriptome analysis of an imminent
biofuel crop, Camelina sativa L. using Illumina GAIIX sequencing platform and identification
of SSR markers. Plant Mol Biol 84(1–2):159–171
Najafi G, Ghobadian B, Tavakoli T, Yusaf T (2009) Potential of bioethanol production from
agricultural wastes in Iran. Sustain Energy Rev, Renew. https://doi.org/10.1016/j.rser.2008.
08.010
Nanasato Y, Kido M, Kato A, Ueda T, Suharsono S, Widyastuti U, Tsujimoto H, Akashi K (2015)
Efficient genetic transformation of Jatropha curcas L. by means of vacuum infiltration
combined with filter-paper wicks. In Vitro Cell Dev Biol-Plant 51(4):399–406
Natarajan P, Kanagasabapathy D, Gunadayalan G, Panchalingam J, Sugantham PA, Singh KK,
Madasamy P (2010) Gene discovery from Jatropha curcas by sequencing of ESTs from
normalized and full-length enriched cDNA library from developing seeds. BMC Genom 11
(1):606
Natarajan P, Parani M (2011) De novo assembly and transcriptome analysis of five major tissues of
Jatropha curcas L. using GS FLX titanium platform of 454 pyrosequencing. BMC Genomics
12(1):191
Pamidimarri DVNS, Pandya N, Reddy MP, Radhakrishnan T (2008a) Comparative study of
interspecific genetic divergence and phylogenic analysis of genus Jatropha by RAPD and AFLP
Genetic divergence and phylogenic analysis of genus Jatropha. Mol Biol Rep 36:901–907
Pamidimarri DVNS, Singh S, Mastan SG, Patel J, Reddy MP (2008b) Molecular characterization
and identification of markers for toxic and non-toxic varieties of Jatropha curcas L. using
RAPD, AFLP and SSR markers. Mol Biol Rep 36:1357–1364
Pamidimarri DS, Reddy MP (2014) Phylogeography and molecular diversity analysis of Jatropha
curcas L. and the dispersal route revealed by RAPD, AFLP and nrDNA-ITS analysis. Mol Biol
Rep 41(5):3225–3234
Pan J, Fu Q, Xu ZF (2010) Agrobacterium tumefaciens-mediated transformation of biofuel plant
Jatropha curcas using kanamycin selection. Afr J Biotech 9(39):6477–6481
Patade VY, Khatri D, Kumar K, Grover A, Kumari M, Gupta SM, Kumar D, Nasim M (2014)
RNAi mediated curcin precursor gene silencing in Jatropha (Jatropha curcas L.). Mol Biol
Rep 41(7):4305–4312
Pioto F, Costa RS, França SC, Gavioli EA, Bertoni BW, Zingaretti SM (2015) Genetic diversity by
AFLP analysis within Jatropha curcas L. populations in the State of São Paulo, Brazil.
Biomass Bioenergy 80:316–320
Pootakham W, Jomchai N, Ruang-areerate P, Shearman JR, Sonthirod C, Sangsrakru D,
Tragoonrung S, Tangphatsornruang S (2015) Genome-wide SNP discovery and identification
of QTL associated with agronomic traits in oil palm using genotyping-by-sequencing (GBS).
Genomics 105(5):288–295
Purkayastha J, Sugla T, Paul A, Solleti SK, Mazumdar P, Basu A, Mohommad A, Ahmed Z,
Sahoo L (2010) Efficient in vitro plant regeneration from shoot apices and gene transfer by
particle bombardment in Jatropha curcas. Biol Plant 54(1):13–20
Qu J, Mao HZ, Chen W, Gao SQ, Bai YN, Sun YW, Geng YF, Ye J (2012) Development of
marker-free transgenic Jatropha plants with increased levels of seed oleic acid. Biotechnol
Biofuels 5(1):10
Ram SG, Parthiban KT, Kumar RS, Thiruvengadam V, Paramathma M (2008) Genetic diversity
among Jatropha species as revealed by RAPD markers. Genet Resources Crop Evol 55:803–
809
238 R. Sharafi and G. S. Jouzani

Ranade SA, Srivastava AP, Rana TS, Srivastava J, Tuli R (2008) Easy assessment of diversity in
Jatropha curcas L. plants using two single-primer amplification reaction (SPAR) methods.
Biomass Bioenergy 32:533–540
Raposo RS, Souza IGB, Veloso MEC, Kobayashi AK, Laviola BG, Diniz FM (2014)
Development of novel simple sequence repeat markers from a genomic sequence survey
database and their application for diversity assessment in Jatropha curcas germplasm from
Guatemala. Genet Mol Res 13(3):6099–6106
Raupach MJ, Amann R, Wheeler QD, Roos C (2016) The application of “-omics” technologies for
the classification and identification of animals. Org Divers Evol 16(1):1–12
Sato S, Hirakawa H, Isobe S, Fukai E, Watanabe A, Kato M et al (2011) Sequence analysis of the
genome of an oil-bearing tree, Jatropha curcas L. DNA Res 18(1):65–76
Schmutz J, Cannon SB, Schlueter J, Ma J, Mitros T, Nelson W, Hyten DL, Song Q, Thelen JJ,
Cheng J, Xu D (2010) Genome sequence of the palaeopolyploid soybean. Nature 463
(7278):178–183
Sharma N, Anderson M, Kumar A, Zhang Y, Giblin EM, Abrams SR et al (2008) Transgenic
increases in seed oil content are associated with the differential expression of novel
Brassica-specific transcripts. BMC Genom 9:619
Silva-Junior O, Rosado T, Laviola B, Pappas M, Pappas G, Grattapaglia D (2011) Genome-wide
SNP discovery from a pooled sample of accessions of the biofuel plant Jatropha curcas based
on whole-transcriptome Illumina resequencing. BMC Proc 5:P57
Singh R, Ong-Abdullah M, Low ETL, Manaf MAA, Rosli R, Nookiah R, Ooi LCL, Ooi SE,
Chan KL, Halim MA, Azizi N (2013) Oil palm genome sequence reveals divergence of
interfertile species in Old and New worlds. Nature 500(7462):335–339
Staton SE, Bakken BH, Blackman BK, Chapman MA, Kane NC, Tang S, Ungerer MC, Knapp SJ,
Rieseberg LH, Burke JM (2012) The sunflower (Helianthus annuus L.) genome reflects a
recent history of biased accumulation of transposable elements. Plant J 72(1):142–153
Schneider MV, Orchard S (2011) Omics technologies, data and bioinformatic principles. Methods
Mol Biol 719:3–30
Sun F, Liu P, Ye J, Lo L, Cao S, Li L, Yue G, Wang C (2012) An approach for Jatropha
improvement using pleiotropic QTLs regulating plant growth and seed yield. Biotechnol
Biofuels 5(1):42
Tabatabaei M, Tohidfar M, Salehi Jouzani G, Safarnejad M, Pazouki M (2011) Biodiesel
production from genetically engineered microalgae: future of bioenergy in Iran. Renew Sustain
Energy Rev 15(4):1918–1927
Taher H, Al‐Zuhair S (2016) The use of alternative solvents in enzymatic biodiesel production: a
review. Biofuels Bioprod Biorefin
Tang MJ, Sun JW, Liu Y, Chen F, Shen SH (2007) Isolation and functional characterization of the
JcERF gene, a putative AP2/EREBP domain containing transcription factor, in the woody oil
plant Jatropha curcas. Plant Mol Biol 63(3):419–428
Tatikonda L, Wani SP, Kannan S, Naresh B, Hoisington DA, Devi P (2009) AFLP-based
molecular characterization of an elite germplasm collection of Jatropha curcas L. A biofuel
plant. Plant Sci 176:505–513
Tian Y, Zhang M, Hu X, Wang L, Dai J, Xu Y, Chen F (2016) Over-expression of CYP78A98, a
cytochrome P450 gene from Jatropha curcas L., increases seed size of transgenic tobacco.
Electron J Biotechnol 19:15–22
Tong L, Shu-Ming P, Wu-Yuan D, Dan-Wei M, Ying X, Meng X, Fang C (2006) Characterization
of a new stearoyl-acyl, carrier protein desaturase gene from Jatropha curcas. Biotechnol Lett
28:657–662
Tong L, Wei MD, Ying X, Yuan DW, Meng X, Wei QR, Fang C (2007) Cloning and
characterization of a stearoyl-ACP desaturase gene from Jatropha curcas. J Shanghai Univ
(English Edition) 11(2):182–188
Tsuchimoto S, Cartagena J, Khemkladngoen N, Singkaravanit S, Kohinata T, Wada N, Sakai H,
Morishita Y, Suzuki H, Shibata D, Fukui K (2012) Development of transgenic plants in
Jatropha with drought tolerance. Plant Biotechnol 29(2):137–143
10 “Omics Technologies” and Biodiesel Production 239

van Dijk EL, Auger H, Jaszczyszyn Y, Thermes C (2014) Ten years of next-generation sequencing
technologies. Trends Genet 30:418–426
Verma KC, Singh US, Verma SK, Gaur AK (2016) Molecular profiling of Jatropha curcas L.
collected from different geographical locations of India. Int J Ambient Energy 37(1):20–23
Wang CM, Liu P, Yi CX, Gu KY, Sun F, Li L et al (2011a) A first generation microsatellite- and
SNP-based linkage map of Jatropha. PLoS ONE 6(8):e3632
Wang L, Gao J, Qin X, Shi X, Luo L, Zhang G, Yu H, Li C, Hu M, Liu Q, Xu Y (2015) JcCBF2
gene from Jatropha curcas improves freezing tolerance of Arabidopsis thaliana during the
early stage of stress. Mol Biol Rep 42(5):937–945
Wang X, Wang H, Wang J, Sun R, Wu J, Liu S, Bai Y, Mun JH, Bancroft I, Cheng F, Huang S
(2011b) The genome of the mesopolyploid crop species Brassica rapa. Nat Genet 43
(10):1035–1039
Wang CM, Liu P, Sun F, Li L, Liu P, Ye J, Yue GH (2012) Isolation and identification of miRNAs
in Jatropha curcas. Int J Biological Sci 8(3):418
Wu L, Goh ML, Tian D, Gu K, Hong Y, Yin Z (2017) Isolation and characterization of curcin
genes with distinct expression patterns in leaves and seeds of Jatropha curcas L. Plant Gene
9:34–44
Wu P, Zhou C, Cheng S, Wu Z, Lu W, Han J, Chen Y, Chen Y, Ni P, Wang Y, Xu X (2015)
Integrated genome sequence and linkage map of physic nut (Jatropha curcas L.), a biodiesel
plant. Plant J 81(5):810–821
Xiong W, Wei Q, Wu P, Zhang S, Li J, Chen Y, Li M, Jiang H, Wu G (2017) Molecular cloning
and characterization of two b-ketoacyl-acyl carrier protein synthase I genes from Jatropha
curcas L. J Plant Physiol 214:152–160
Yadav HK, Ranjan A, Asif MH, Mantri S, Sawant SV Tuli R (2011) EST-derived SSR markers in
Jatropha curcas L. development, characterization, polymorphism, and transferability across
the species/genera. Tree Genet & Genomes 7(1):207–219
Yang J, Yang MF, Wang D, Chen F, Shen SH (2010) JcDof1, a Dof transcription factor gene, is
associated with the light-mediated circadian clock in Jatropha curcas. Physiologia Plantarum
139(3):324–334
Yang D, Zhang H, Peng K, Chen L, He H, Huang X, Qin J, He G, Zhang D (2016) Differential
gene regulation of lipid synthesis in the developing seeds of two biodiesel tree species,
Jatropha and Vernicia. Int J Agric Biol 18(6)
Yang H, Yu C, Yan J, Wang X, Chen F, Zhao Y, Wei W (2014) Overexpression of the Jatropha
curcas JcERF1 gene coding an AP2/ERF-Type transcription factor increases tolerance to salt
in transgenic tobacco. Biochemistry (Moscow) 79(11):1226–1236
Ye J, Hong Y, Qu J, Wang C (2013) Improvement of Jatropha curcas oil by genetic transformation.
In: Jatropha, challenges for a new energy crop. Springer, New York, pp 547–562
Ye J, Qu J, Mao HZ, Ma ZG, Rahman NE, Bai C, Chen W, Jiang SY, Ramachandran S, Chua NH
(2014) Engineering geminivirus resistance in Jatropha curcas. Biotechnol Biofuels 7(1):149
Yu N, Xiao WF, Zhu J, Chen XY, Peng CC (2015) The Jatropha curcas KASIII gene alters fatty
acid composition of seeds in Arabidopsis thaliana. Biol Plant 59(4):773–782
Yue GH, Sun F, Liu P (2013) Status of molecular breeding for improving Jatropha curcas and
biodiesel. Renew Sustain Energy Rev 26:332–343
Zampieri E, Chiapello M, Daghino S, Bonfante P, Mello A (2016) Soil metaproteomics reveals an
inter-kingdom stress response to the presence of black truffles. Sci Rep 6

You might also like