Download as pdf or txt
Download as pdf or txt
You are on page 1of 54

Biodiesel Feedstocks Technologies

Economics and Barriers Assessment of


Environmental Impact in Producing and
Using Chains Armen B. Avagyan
Visit to download the full and correct content document:
https://textbookfull.com/product/biodiesel-feedstocks-technologies-economics-and-ba
rriers-assessment-of-environmental-impact-in-producing-and-using-chains-armen-b-a
vagyan/
More products digital (pdf, epub, mobi) instant
download maybe you interests ...

Environmental Impact Assessment Theory and Practice


Anji Reddy Mareddy (Auth.)

https://textbookfull.com/product/environmental-impact-assessment-
theory-and-practice-anji-reddy-mareddy-auth/

Introduction To Environmental Impact Assessment 5th


Riki Therivel

https://textbookfull.com/product/introduction-to-environmental-
impact-assessment-5th-riki-therivel/

Chemistry for Environmental Engineering Armen S.


Casparian

https://textbookfull.com/product/chemistry-for-environmental-
engineering-armen-s-casparian/

Environmental Law and Economics Theory and Practice 1


;b Edition Michael G. Faure

https://textbookfull.com/product/environmental-law-and-economics-
theory-and-practice-1-b-edition-michael-g-faure/
Environmental Sustainability in Asian Logistics and
Supply Chains Xiaohong Liu

https://textbookfull.com/product/environmental-sustainability-in-
asian-logistics-and-supply-chains-xiaohong-liu/

Biodiesel Production with Green Technologies 1st


Edition Aminul Islam

https://textbookfull.com/product/biodiesel-production-with-green-
technologies-1st-edition-aminul-islam/

Handbook of Distributed Generation Electric Power


Technologies Economics and Environmental Impacts 1st
Edition Ramesh Bansal (Eds.)

https://textbookfull.com/product/handbook-of-distributed-
generation-electric-power-technologies-economics-and-
environmental-impacts-1st-edition-ramesh-bansal-eds/

Microcosm Manual for Environmental Impact Risk


Assessment From Chemicals to Whole Effluent Toxicity
WET Yuhei Inamori

https://textbookfull.com/product/microcosm-manual-for-
environmental-impact-risk-assessment-from-chemicals-to-whole-
effluent-toxicity-wet-yuhei-inamori/

Foundations of Environmental Economics Wolfgang


Buchholz

https://textbookfull.com/product/foundations-of-environmental-
economics-wolfgang-buchholz/
Armen B. Avagyan · Bhaskar Singh

Biodiesel:
Feedstocks,
Technologies,
Economics and
Barriers
Assessment of Environmental Impact in
Producing and Using Chains
Biodiesel: Feedstocks, Technologies, Economics
and Barriers
Armen B. Avagyan • Bhaskar Singh

Biodiesel: Feedstocks,
Technologies, Economics
and Barriers
Assessment of Environmental Impact
in Producing and Using Chains
Armen B. Avagyan Bhaskar Singh
President and Sole Founder Department of Environmental Sciences
R&I Center of Photosynthesizing Central University of Jharkhand
Organism Ranchi, Jharkhand, India
Yerevan, Armenia

ISBN 978-981-13-5745-9 ISBN 978-981-13-5746-6 (eBook)


https://doi.org/10.1007/978-981-13-5746-6

© Springer Nature Singapore Pte Ltd. 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 Singapore Pte Ltd.
The registered company address is: 152 Beach Road, #21-01/04 Gateway East, Singapore 189721,
Singapore
Overview

This book is devoted to the biodiesel industry, its limitations, and the evaluation of
the full manufacturing chain to allow making conclusions for further improvements.
In 2016, total energy investment was more than $1.7 trillion, shares of oil and gas
were $649 billion, and those for renewables (transport and heat) were $19 billion.
G20 nations responsible for more than three-quarters of global GHG emissions
contributed an average of $71.8 billion/year of public finance for fossil fuel projects
and only $18.7 billion/year for renewable energy. The adverse effect of fossil fuel
subsidies is to divert the use of public funds from other necessary expenses such as
social spending, health, clean energy, and energy access for the poor. Termination of
these subsidies can reduce GHG emissions by 21% and deaths caused by fossil fuel
air pollution by 55%. However, it will be a difficult decision as the world oil demand
will increase to 16.5–19.1 billion liters per day by 2040 from approximately
14.3 billion liters per day in 2014, and the continued extraction and combustion of
fossil fuels will create further severe environmental challenges.
Air pollution policy is closely connected with climate change, public health,
energy, transport, trade, and agriculture. Overall, the Earth has been pushed to the
brink and the damage is becoming increasingly obvious. The transport sector
remains a foremost source of air pollutants, which has stimulated the production of
biofuels to become one of the most rapidly rising markets in the current bioeconomy.
There, policies fill an imperative role, and more than 50 countries have applied
biofuel blending targets as well as other measures such as tax incentives. Introduc-
tion into the market generally depends on mandatory biofuel blending of fuel
resulting from low oil prices and subsidy contributions. Biofuel composition allows
decreasing emissions of sulfur oxides or nitrous oxides and CO from vehicles
compared with diesel and petroleum fuels.
Oil derived from oilseed plantations/crops is the most commonly used feedstock
for the production of biodiesel. Simultaneously, the UK Royal Academy of Engi-
neering and 178 Netherlands scientists determined that some biofuels, such as diesel
produced from food crops, have led to more emissions than those produced by fossil
fuels. Therefore, reevaluation of the full cycle of biodiesel production for optimal

v
vi Overview

solutions is necessary. Our analysis shows that production of fertilizers causes


greenhouse gas (GHG) emissions of 0.9–1.2 kgCO2e/l biodiesel. It was also indi-
cated that the use of fertilizers produces additional emissions that exceed emissions
from their production by 2- to 5.5 fold. This book confirms that the production and
use of fertilizers for cultivation of biodiesel feedstock crops generate much greater
GHG emissions than are mitigated by the use of biodiesel. To address this dilemma,
future biofuel development requires producers to shift to the use of feedstock yields
originated from organic agriculture approaches, including the use of microalgae.
Advanced biofuels (second-generation, or 2G) produced from nonfood crops,
woody or grassy materials, straw, animal fat, forest residues, sawmill by-products,
waste cooking oil, etc. and third-generation (3G), from algae, are considered suitable
replacements for first-generation (1G) biofuels because their feedstocks can be
grown in marginal lands that are usually not suitable for crop cultivation and do
not directly compete with food production or land use. The advantages of algae as
feedstock include the highest efficiency in converting solar energy, absorption of
CO2 (exhaust gases can be used for aeration, heating, and as nutrients) and pollut-
ants, etc., and therefore are recognized as a better feedstock for future fuels.
Microalgae cultivation is not limited by water resources. On the other hand, our
analysis proves that photoautotrophic growth of microalgae has no potential for
mitigation of GHG emissions and can be applied only for other purposes. Efforts to
increase biofuel volume from plants and other sources must be directed at transfor-
mation of wastes into low-carbon transportation fuels and chemicals that are already
addressing the challenges related to global sustainable development. At the same
time, cultivation of microalgae in photobioreactors can have only limited application
as there are well-known difficulties from high energy consumption in cleaning the
internal and external walls of the reactors for the use of alternative nutrition sources
as such as wastewaters and the relatively low installation.
Another barrier for the commercialization of algae for biofuel is related to high
capital investment, the operation costs of biomass production (fertilizers, energy,
freshwater) and technological challenges. To the question: why have algae benefits
not been realized in biofuel production? the response is found in a combination of
economics, technology, and political issues, and our analyses provide that the main
barrier to realization of the algal biofuel potential is ineffective international and
governmental policies which create difficulties in coupling the goals of economic
development and environmental activity. Therefore, it is a methodical mistake to
compare the cost of microalgal biofuel production with the subsidized price of oil
fuel and 1G biofuel as the governments have paid part of these fuel prices from
public funds. Algae biofuel benefits include the prospect of a high flow of invest-
ments without producing subsidies, if business models and an environmentally
driven approach can be achieved through payments for mitigation of waste and air
pollution. A transformative model can be created wherein all elements generate
direct economic, societal, or environmental benefits aimed at the well-being of
people and of nature. Each company whose activity directly or indirectly decreases
pollution must receive payment for its Life Conserve product. The book shows that
Overview vii

the world must create new activities that include legislation, regulations, and guid-
ance addressed to promote corporations and other companies in mitigating environ-
mental challenges through a new economic model and instruments. Only the
development of the Global Life Conserve Industry can provide solutions to problems
associated with sustainable development.
Macroalgae can be converted into bio-oil and its lipids then be separated for
biodiesel production. However, the high lipid content of some microalgae compared
to macroalgae has centered attention on the production of biodiesel from microalgae.
Also, it is doubtful that sufficient seaweed can be harvested to provide significant
quantities of transport fuel because of the technological and economic barriers.
Research Areas, Societies and Potential Audiences of the Book This book
includes analysis of feedstocks, technologies, economics, and barriers of the biodie-
sel industry, as well as biodiesel crop production impacts on greenhouse emission
mitigation, with a problem-solving approach. Its profile is multidisciplinary within
the areas of environmental policy, sustainable development and climate change,
renewables, economics of the production chain, biotechnology, agriculture, bio-
materials and market analysis, biodiesel feedstock production, technologies of
biomass conversion, greenhouse gas emission mitigation, and life cycle assessment.
Contents

1 Introduction. Links to International Policy and Markets . . . . . . . . . 1


References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2 Biodiesel from Plant Oil and Waste Cooking Oil . . . . . . . . . . . . . . . . 15
2.1 Properties of Biodiesel and Production . . . . . . . . . . . . . . . . . . . . 16
2.2 Forest, Marginal Land, and Wasteland for the Supply
of Feedstock for Biodiesel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.3 EROEI of Feedstocks Used in the Synthesis of Biodiesel . . . . . . . 29
2.4 Catalysts: Efficiency, Cost Aspects, and Leaching . . . . . . . . . . . . 33
2.5 Refining Methods of Crude Biodiesel . . . . . . . . . . . . . . . . . . . . . 40
2.5.1 Water Washing of Crude Biodiesel . . . . . . . . . . . . . . . . . 43
2.5.2 Dry Washing of Crude Biodiesel . . . . . . . . . . . . . . . . . . . 44
2.6 Transportation and Storage of Biodiesel . . . . . . . . . . . . . . . . . . . 46
2.6.1 Transportation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
2.6.2 Storage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
2.7 Life Cycle Assessment of Biodiesel . . . . . . . . . . . . . . . . . . . . . . 51
2.8 Deterrents and Proponents of the Commercialization
of Biodiesel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
2.9 Barriers in the Production of Biodiesel . . . . . . . . . . . . . . . . . . . . 63
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
3 Biodiesel from Algae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
3.1 History and Classification of Algae . . . . . . . . . . . . . . . . . . . . . . 79
3.1.1 Algae to Biofuel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
3.2 Technologies of Algae Cultivation and Economics . . . . . . . . . . . 81
3.2.1 Microalgae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
3.2.2 Macroalgae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
3.3 Algal Biomass Conversion to Biodiesel . . . . . . . . . . . . . . . . . . . 95
3.3.1 Microalgae Biomass Conversion to Biodiesel . . . . . . . . . . 95
3.3.2 Macroalgae Biomass Conversion to Biodiesel . . . . . . . . . 101
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102

ix
x Contents

4 Barriers in the Biofuel-Producing Chain and Revision of


Environmental Impacts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
About the Authors

Armen B. Avagyan graduated with degrees from the


Biological Department of Yerevan State University and
the Economics Department of Armenian State Univer-
sity before pursuing postgraduate studies at Moscow
State University. He subsequently worked as a senior
researcher at the Agriculture Institute and Technological
Institute of Amino Acids; director of Yerevan Vitamin
Plant, Institute of Biotechnology; and deputy director of
Nairit Chloroprene Rubber Plant. He founded the
Research and Industry Center of Photosynthesizing
Organisms, Feed Additives and Physiologically Active
Compounds and was an advising expert to EU Horizon
2020. He is a member of the American Chemical Soci-
ety, Society of Chemical Industry (USA), and serves on
the editorial boards of several international journals.
Dr. Avagyan has received the “International Presi-
dents Award for Iconic Achievement” and “Top
100 Professionals” (IBC, England), “The Albert Ein-
stein Award for Excellence” (ABI, USA), the 2018
“Albert Nelson Marquis Lifetime Achievement
Award” (USA), and other awards.
He has published more than 100 works, including the
following monographs: Theory of Global Sustainable
Development Based on Including of Microalgae in Bio
and Industrial Cycles; New Design and Building of
Biological System and Algae to Energy and Sustainable
Development; Technologies, Resources, Economics and
System Analyses; and New Design of Global Environ-
mental Policy and Live Conserve Industry.

xi
xii About the Authors

Bhaskar Singh graduated from Ranchi University


(BSc, environment and water management), completed
his MSc in environmental sciences at Jiwaji University,
and received his MPhil (environmental science and
engineering) from Pondicherry University, India, with
a Gold Medal awarded in 2006. He obtained his PhD
from the prestigious Indian Institute of Technology
(BHU), Varanasi, UP, India, in 2010, before conducting
postdoctoral research at Durban University of Technol-
ogy, South Africa. He is currently serving as an assistant
professor in the Department of Environmental Sciences,
Central University of Jharkhand, Ranchi, India. He also
serves as a reviewer for several international journals
from Elsevier, Springer, and the ACS. He is a recipient
of the prestigious CSIR senior research fellowship and is
a CSIR research associate.
Dr. Singh’s current research interest lies in the syn-
thesis of biodiesel from plants and used cooking oil and
in the development of novel heterogeneous catalysts. He
has co-edited two books with Springer: Algae and Envi-
ronmental Sustainability and Phytoremediation Poten-
tial of Bioenergy Plants. His publications include
40 research and review papers in peer-reviewed and
high-impact international journals as well as 15 book
chapters.
Chapter 1
Introduction. Links to International Policy
and Markets

Abstract In 2016, the total energy investment was more than $1.7 trillion, shares of
oil and gas were $649 billion, and renewables (transport and heat) only $19 billion.
G20 nations responsible for more than three-quarters of global greenhouse gas
(GHG) emissions contributed an average of $71.8 billion/year of public finance
for fossil fuel projects and only $18.7 billion/year for renewable energy. The adverse
effects of fossil fuel subsidies consist of the diversion of public funds from other
necessary expenses such as social spending, health, clean energy, and energy access
for the poor. Termination of these subsidies can reduce GHG emissions by 21% and
reduce deaths caused by fossil fuel air pollution to 55%. However, this course will be
a challenging scenario as the world oil demand will increase from approximately
14.3 billion liters per day in 2014 to 16.5 to 19.1 billion liters per day by 2040,
whereas the continued extraction and combustion of fossil fuels will create severe
environmental challenges.
Air pollution policy is closely connected with climate change, public health,
energy, transport, trade, and agriculture. Overall, the Earth has been pushed to the
brink and the damage is becoming increasingly obvious. This realization stimulates
the production of biofuels, which has become one of the most rapidly rising markets
in the current bioeconomy. Biofuel composition results in less sulfur oxides, nitrous
oxides, and carbon monoxide emissions during its combustion compared with fossil
fuels. Thus, this chapter focuses on the transport sector, which remains the foremost
source of air pollutants. More than 50 countries have applied a biofuel blending
target as well as other measures such as tax incentives. In 2015, support for biofuels
achieved only $26 billion. For 2021, projected biofuel volume can exceed $41 billion
at a compound annual growth rate of 3.8%.
Among advanced biofuels, algal biomass provides a low risk of causing indirect
land use change; as it does not compete directly for agricultural land for the food and
feed markets, this is recognized as the better future feedstock. The global algae
biofuel market volume was expected to be $5.96 billion in 2018 and will reach
$10.73 billion by 2025.
Simultaneously, the UK Royal Academy of Engineering and 178 Netherlands
scientists determined that some biofuels, such as diesel produced from food crops,
have led to more emissions than those produced by the fossil fuels. This statement
requires reevaluating the full cycle of biodiesel production to find optimal solutions.

© Springer Nature Singapore Pte Ltd. 2019 1


A. B. Avagyan, B. Singh, Biodiesel: Feedstocks, Technologies, Economics and
Barriers, https://doi.org/10.1007/978-981-13-5746-6_1
2 1 Introduction. Links to International Policy and Markets

Keywords Biodiesel · Biomass conversion economics · Feedstock · Investment ·


Environmental policy · Climate change · Pollution

Abbreviations

Btu British thermal unit (equivalent 1055 joules)


CAGR annual growth rate
CO2e carbon dioxide equivalent for a gas
EROI energy return on investment
FAME fatty acid methyl ester
GHG greenhouse gases
HC unburned hydrocarbon
HVO hydro-treated vegetable oil
MBOE million barrels of oil equivalent
MJ megajoule
Mtoe million tonnes of oil equivalent (energy defined as the amount of energy
released by burning 1 ton of crude oil)
NOx oxides of nitrogen
PM particulate matter
RFS Renewable Fuel Standard

Our world environment is one of the main challenges for economic and social
progress. Climate change impacts include rising temperatures, changing precipita-
tion patterns, and intensifying of extreme weather events (Avagyan 2012–2013,
2018; WEF 2018; WEO 2016). Air pollution produces climate change, which is
closely connected with energy, transport, trade, and agriculture. On the whole the
Earth has been pushed to the brink of ecosystem collapse, and the damage is
becoming increasingly clear (Avagyan 2012–2013, 2018, WEF 2018). Currently,
trends in ecosystems are calling out to us that our life support systems are unstable,
out of balance, and the challenges of the twenty-first century mandate extensive
transformations of our current environmental and energy systems (Avagyan 2018;
FAO-IPCC 2017).
The Sun is the basic source of Earth’s energy. Following the origin of life on
Earth, solar energy and chemical elements were captured and stored in living matter
(Avagyan 2018). Over millions of years, in geochemical processes oil derived from
ancient fossilized organic materials such as algae and zooplankton, coal from dead
plant matter, and natural gas formed when layers of decomposing plant and animal
matter were exposed to high heat and pressure beneath the surface of the Earth.
Oil is the world’s leading fuel (one third of global energy consumption). Increase
in oil consumption averages 254 million liters (l) per day, or 1.6% more than the
10-year average (BP 2017). The Dated Brent oil average price was $0.28 per liter in
1 Introduction. Links to International Policy and Markets 3

2016 ($0.33 per liter in 2015), which was the lowest level since 2004 (BP 2017). The
rising oil demand has promoted low prices of consumer-led fuels such as gasoline
(BP 2017). In 2017, oil prices increased and then stabilized because of the OPEC/
non-OPEC agreement (BP 2017). The price of Brent crude reached $0.42 per liter in
December 2017 (Singgih 2018). For the first time since 2009, diesel demand fell in
2016, which was also connected with the industrial slowdown (BP 2017).
In 2016, global natural gas consumption was increased by 1.5% (63 billion cubic
meters, or slower than the 10-year average of 2.3%), nuclear power generation
increased by 1.3% (9.3 million tonnes of oil equivalent, Mtoe), and hydroelectric
power generation increased by 2.8% (27.1 Mtoe) (BP 2017). However, world coal
production decreased by 6.2% (231 Mtoe) because of the largest downturn; China
and US production fell by 7.9% (140 Mtoe) and 19% (85 Mtoe), respectively)
(BP 2017).
In 2016, total energy investment was more than $1.7 trillion, that for shares of oil
and gas were $649 billion, and for renewables (transport and heat) only $19 billion
(IEA 2017a). The G20 nations responsible for more than three fourths of global
greenhouse gas (GHG) emissions contributed an average of $71.8 billion/year of
public finance for fossil fuel projects and only $18.7 billion/year for renewable
energy (Doukas et al. 2017). The unfavorable consequences of fossil fuel subsidies
are the diversion of public funds from other necessary expenditures such as social
services, health, clean energy, and energy access for the poor (OCI 2017). Elimi-
nating fossil fuel subsidies will be a key role in the development of a green economy
(OECD 2017), possibly reducing GHG emissions by 21% and deaths caused by
fossil fuel air pollution to 55% (Coady et al. 2017). The United Nations (UN), the
Organisation for Economic Co-operation and Development (OECD), the global
leaders of G20 and G8 (now G7) countries, etc. have announced their intent to
move away from environmentally harmful fossil fuel subsidies but find the steps
difficult (Avagyan 2018; OCI 2017; WEC 2016). Overall, at the present time OECD
countries continue to maintain subsidies for potentially environmentally harmful
fossil fuels, and the share directed to environment and energy objectives has
remained unchanged (OECD 2017). Only a few countries provide high support for
clean energy finance as compared with fossil fuel finance (France, Mexico, and
Australia) (OCI 2017). However, termination of fossil fuel subsidies will be a
difficult decision because, according to expectations, world oil demand will increase
from approximately 14.3 million liters per day in 2014 to 16.5–19.1 million liters per
day by 2040 (Avagyan 2018; Coady et al. 2017; IEA 2017a). However, the
continued extraction and combustion of fossil fuels are creating serious environ-
mental damage such as global warming and ocean acidification.
According to the COP22, leaders of the 48 nations that constitute the Climate
Vulnerable Forum jointly declared efforts toward achieving 100% renewable energy
in their respective nations (REN21 2017). However, the world investments are not
moving enough fast to develop advanced biofuels, which is a problem for the further
growth of the renewables branch (IBRD-WB-IEA 2017). The good news is that in
2017 the World Bank announced a suspension of financing upstream oil and gas
investments after 2019 (WEF 2018).
4 1 Introduction. Links to International Policy and Markets

The transport sector remains the foremost source of air pollutants (Atabani et al.
2012; Avagyan 2008, 2010, 2011, 2012–2013, 2017, 2018; Bugarski et al. 2011;
Rahman 2015; U.S. Department of Energy 2014; WEO 2016). The more environ-
mentally friendly way to decrease GHG emissions associated with energy produc-
tion is to develop energy from carbon-neutral or reduced carbon emission sources
(Sayre 2010). At the popular level, if we use fossil fuel, we increase GHG emissions
from previously conserved bioenergy. If biofuels feedstocks are obtained from raw
plant materials, we begin a new conservation of bioenergy, reducing the СO2
atmospheric content, and only in the combustion chamber do we return СO2 to the
atmosphere. Further, biofuel composition allows decreasing sulfur oxides, nitrous
oxides, and carbon monoxide upon combustion as compared with diesel and petro-
leum fuels (Fernández et al. 2012; Liu et al. 2013). Therefore, governments are
promoting the development of biofuels manufacturing as an answer to the challenges
of climate change, energy security, and health (Avagyan 2008, 2010, 2012–2013,
2018; GVR 2017; New 2017). Governmental policies that have an imperative role in
the increase of biofuel markets were started in the second half of the 2000s with
shifting the market structure by increasing biofuel production (OECD-FAO 2017).
The overall demand for biofuels was promoted by mandatory blending of fossil fuels
with biofuels (OECD-FAO 2017). More than 50 countries have applied a biofuel
blending target, combined with other measures such as tax incentives (European
Parliament 2015).
The biofuels market is one of the fastest rising markets in the current bioeconomy
(Avagyan 2018). In 2006, biofuel manufacturing was 49 billion liters (3% of the
gasoline and diesel fuel market), whereas in 2015–2016 the biofuels market reached
135–155 billion liters (Kotrba 2017; REN21 2017). For 2021, the projected biofuel
volume can exceed $41 billion at a compound annual growth rate (CAGR) of 3.8%
(Kotrba 2017). By 2030, the produced volume could be as much as 30% of the world
global transport fuel mix (New 2017). According to the evaluation in energy terms,
the existing biofuel production share of bioethanol is 72% or 99 billion liters,
biodiesel is 23%, and hydro-treated vegetable oil (HVO) is 4% (REN21 2017).
This growth has been aided by an objective aimed to decrease GHG emissions in the
transport sector with minimal changes to vehicle stocks and fuel distribution infra-
structure, to promote progress in energy security and the sustainable development of
the agricultural sector and the rural economy, while reducing oil price volatility
(European Parliament 2015; OECD-FAO 2017; WEO 2016). In 2015, biofuels
support recorded only $26 billion (WEO 2016).
The US and Brazil are the major biofuel producers (70% volume), followed by
Germany, Argentina, China, and Indonesia (EPA 2017; REN21 2017).
The US consumption of renewable biofuels was 1.144 quadrillion British thermal
units (Btu) for ethanol and 0.261 quadrillion Btu for biodiesel in 2016 (EIA 2017a)
or 0.7 MBOE/day (6% of US road transport energy use), which was supported by the
Renewable Fuel Standard (RFS) (WEO 2016). Biofuel volume will increase as much
as 1 MBOE/day by 2025 and 1.4 MBOE/day by 2040 (share of ethanol up to 80%
and biodiesel to 20%) (WEO 2016). In 2017, the RFS 2 required the blending of
73 billion liters of renewable fuels, including 16.2 billion liters of advanced biofuels
1 Introduction. Links to International Policy and Markets 5

and 1.2 billion liters of cellulosic biofuels (REN21 2017). According to RFS
2, 136.3 billion liters of biofuels was planned for blending into gasoline and diesel
fuels by 2022 (later changed to 83.3 billion liters) (WEO 2016).
In 2015, US biodiesel cut GHG emissions by 18 million tons carbon dioxide
equivalent (CO2e) or the equivalent GHG emissions of 3.8 million cars (NBB 2016).
The most common US biodiesel blends are sold as B5 or B20 (5% and 20%
biodiesel, respectively, blended with petroleum) and B100 (100% biodiesel, which
requires some engine modifications to avoid maintenance and performance prob-
lems) (Avagyan 2012–2013, 2018; EPA 2015; NBB 2017; U.S. Department of
Energy 2016; EIA 2017b). Biodiesel B100 fuel can reduce tailpipe emissions as
much as 47% by particulate matter (PM), 67% for unburned hydrocarbon (HC),
sulfates up to 100%, and CO by 48%; this fuel improves the lubricity of some
engines, but increases NOx emissions as much as 10% (EPA 2015). Other studies
have estimated that biodiesel combustion leads to reducing GHG emissions by at
least 50% and often as much as 86% compared to petroleum diesel fuel, with a
similar energy return on investment (EROI) without major modifications to engines,
and with more energy in comparison with ethanol (Adnan 2015; Araújo et al. 2017;
Avagyan 2012–2013, 2018; BDI 2017; Liu et al. 2013; NBB 2016; U.S. Department
of Energy 2016).
Brazil is the second largest producer of biodiesel worldwide (blended up to 7% of
diesel in 2014 and projected to increase to 10% in 2019) (Avagyan 2012–2013,
2018; Lima et al. 2017) and has the highest share (21%) of bioethanol used in road
transport (GII 2017). Its biodiesel production accounted for 4 billion cubic meters
(m3), mostly from soybean oil (77.7%) followed by beef tallow (18.8%), cotton oil
(2%), and other greasy compounds (1.5%) (Lima et al. 2017). Only Brazil has
non-mandated biofuel demand because policies in the main states favor hydrous
ethanol (OECD-FAO 2017).
The European Union (EU) energy policy is established in the Energy and Climate
Change Package (CCP) and the Fuel Quality Directive (European Parliament 2015;
Flach et al. 2017). This package includes the “20/20/20” obligatory requirements to
2020 with 20% share of renewable energy. The Renewable Energy Directive (RED)
is part of the CCP. The EU Renewable Energy Directive set sustainability require-
ments for a liquid biofuels target of 10% for renewable energy in transport by 2020
(globally, 176 countries have policy targets for renewables) (REN21 2017), and
biofuels use will increase to 0.5 MBOE/day in 2025 and to 0.7 MBOE/day in 2040
(15% of road transport energy demand) (EC 2017; GII 2017). The implementation of
the EU RED is under the responsibility of the Member States (MS) (Flach et al.
2017). The MS countries independently developed the policy but it is controlled by
the EC. The EC established guidelines of 2014 for MS aimed to renew their systems
for reducing distortions in energy markets because these resulted in high energy
prices. The EU GHG mitigation of biomass fuels must be a minimum of 80% after
2020 and 85% after 2025. However, the EU transport sector shows the slowest
growth of renewable (0.5% per year from 2005 to 2014, the renewable energy share
being 5.9–6.0% by 2014–2015 compared to the planned target of 10% for 2020)
(EC 2017). This slow progress could be attributed to difficulties including regulatory
6 1 Introduction. Links to International Policy and Markets

uncertainty and a late uptake of advanced biofuels (EC 2017; UFOP 2017). Simul-
taneously, Europe is the largest consumer of biodiesel (54% of world demand)
(North America, Latin America, and Asia account for about 17%, 14%, and 12%,
respectively) (MR & C 2017). Currently, European biodiesel production has fallen
by 5% or to 10.7 billion liters (REN21 2017) and did not reach the expected level
(was 10.9 Mtoe instead of 14.4 Mtoe) (EC 2017).
The main European centers of biofuels production are in Germany, France,
Sweden, and Spain (GII 2017). Germany is the largest European biodiesel producer
(3.0 billion liters), followed by France (1.5 billion liters) (REN21 2017). Currently,
EU domestic ethanol prices are dependent on the import of significant volumes of
duty-free ethanol from foreign markets (Flach et al. 2017). The European biodiesel
standard, EN 14214, is a standard published by the European Committee for
Standardization [also used is the EN 590 specification for biodiesel blends up to
7% fatty acid methyl ester (FAME)] (Transport Policy 2017), and the specification
for biodiesel in Germany is DIN V 51606. The price of EU biodiesel ranges between
€0.71 and €0.81 per liter (EC 2010; ENMC 2018).
The Indonesia government plans to meet its local biodiesel consumption target
from 2.14 million m3 up to 3.5 million m3 in 2018 (Singgih 2018). China biofuel
policies have focused mainly on ethanol production: an E10 mandate is set in
4 provinces and 27 cities, but production has been constrained and, historically, no
blending was allowed to take place outside these areas (REN21 2017; van Dyk et al.
2016). The production of grain-based biofuels (cassava, sweet potato, sweet sor-
ghum) has increased from 2430 Ttoe in 2001 to 4000 Ttoe in 2015 (Xie et al. 2017).
Among the second-generation or advanced biofuels, China has only one demonstra-
tion or commercial-scale facility based on cellulosic feedstocks (van Dyk et al.
2016). DuPont, Beta Renewables, and Novozymes have announced plans for the
construction of commercial-scale cellulosic ethanol facilities in China but the current
low price of oil has limited their activities (van Dyk et al. 2016).
In Japan in 2014, more than 30 companies and organizations from the aerospace,
fuel, engineering, finance, and research interests created a group named Initiatives
for Next Generation Aviation Fuels to create a plan for using nationally found
aviation biofuels by 2020 (Green Air 2015). The National Standard of Canada
biodiesel is CAN/CGSB-3.524, and the Indian specification for biodiesel comes
under IS 15607.
First-generation (1G) biofuels constitute the largest amount of biofuels world-
wide, produced mainly from wheat, corn, sugarcane, soybean, rapeseed, and other
food crops (Araújo et al. 2017; Avagyan 2008; Avagyan 2012–2013, 2018; Xie et al.
2017; WEO 2016). The share of vegetable oil in biofuel production increased from
less than 1% in 2000 to the current 12%–14% (OECD-FAO 2017). In the US,
soybean oil constitutes about 50% of the feedstock for biodiesel, and 25% of US
soybean oil fabrication is used for biodiesel production (U.S. Department of Energy
2016; WEO 2016). Other feedstocks include yellow grease, canola oil, corn oil,
white grease, tallow, other recycled oils, poultry fat, other vegetable oils, palm oil,
and miscellaneous other sources (WEO 2016). EU biodiesel is mainly produced
from rapeseed oil (amounting to 50% of the total feedstocks) (EC 2017).
1 Introduction. Links to International Policy and Markets 7

The main question of the sustainability of 1G biofuels from food crops is their
impact on the food supply with large-scale use (competition for land and water;
physical availability and access) and trade of biomass associated with increased food
and feed prices (debate of food versus fuel) (Araújo et al. 2017; Avagyan 2008,
2010, 2011, 2012–2013, 2018; European Parliament 2015; Lane 2012; Rocca et al.
2015; Troustle et al. 2012; WEO 2016; WWDR 2012, 2017). Hence, the growth of
1G biofuel might not produce the complete range of expected green-economy
benefits (WWDR 2012). The possibility of grain-based biofuels to replace fossil
fuels in transport is extremely limited (Avagyan 2008, 2010, 2011, 2012–2013,
2018). The U.S. Energy Independence and Security Act (EISA) has capped corn
grain contributions and requires that other crops (perennial grasses, trees, algae, etc.)
be increasingly used to supply ethanol. EISA requires 20% or more GHG reduction
for any renewable fuel production facility constructed after 2007: 50% reduction for
advanced biofuels, 50% reduction for biomass-based diesel, and 60% reduction for
cellulosic biofuels (EIA 2017a). The content of the new EU RED II of 2016
envisaged decreasing the maximum contribution of food crop-based biofuels by
limiting their share as well as by established sustainability criteria for biomass of
forest and palm oil and increased support for advanced biofuels (share of 1.5% in
2021 to 6.8% by 2030) (Flach et al. 2017). According to the RED II the EU will
produce 27% renewable sources of its consumed energy near 2030 and a cap on
conventional crop biofuels will decrease from 7% (2021) to 3.8% (decrease of these
biofuels volume is also established in other decisions) (EU 2015; European Parlia-
ment 2015, 2017; Rocca et al. 2015; UFOP 2017).
Advanced biofuels (second-generation, 2G) produced from non-food crops,
woody or grassy materials, straw, animal fat, forest residues, sawmill by-products,
waste cooking oil, etc. and third-generation (3G), from algae, are considered to be
suitable replacements for first-generation biofuels because their feedstocks can be
grown in marginal lands that are usually not suitable for crop cultivation and do not
directly compete with food production or land use (Avagyan 2018). However, the
production capacity of all advanced biofuels plants was estimated to be only
5.4 billion liters in 2013 (WER 2016). With sustained support and advances in
technology, biofuels consumption will increase to 4.2 MBOE/day in 2040 or
account for 8% of total energy use (ethanol 65%, biodiesel 30%, aviation biofuel
5%) (WEO 2016).
The US advanced biofuel biodiesel fuel market (50% life cycle carbon emission
reduction) rose from about 94.6 million liters in the early 2000s to more than
10.6 billion liters in 2016 (the US on-road diesel market is about 132.5 billion to
154.8 billion liters). The objective was to manufacture about 10% by 2022 (NBB
2017). However, a large amount of biomass is required for feedstock production:
more than half a million tons by dry weight is required for a biorefinery production
of 150,000 ton/year of lignocellulosic ethanol (WEO 2016). Simultaneously, the US
specifically excluded biomass from federal forests (Bracmort et al. 2011). Cellulosic
biofuel has some potential but significant complications that must be overcome to
reach competitiveness (Avagyan 2012–2013).
8 1 Introduction. Links to International Policy and Markets

The EU list of advanced biofuels includes algae, biomass fraction of industrial


waste and mixed municipal waste, bio-waste, straw, animal manure and sewage
sludge, palm oil mill effluent and empty palm fruit bunches, tall oil and tall oil pitch,
crude glycerin, bagasse, grape marc and wine lees, nutshell, husks, cobs cleaned of
corn kernels, biomass fraction of waste and residues from forestry and forest-based
industries, used cooking oil, certain animal fats, and molasses as a by-product from
refining sugarcane or sugar beets (Directive of the European Parliament and of the
Council amending Directive 98/70/EC relating to the quality of petrol and diesel
fuels and amending Directive 2009/28/EC on the promotion of the use of energy
from renewable sources, Brussels, 17.10.2012 COM (2012) 595 final) (European
Parliament 2015). The present EU production volume of cellulosic ethanol, about
60 million liters, is planned to increase to about 200 million liters in 2021 (Flach
et al. 2017). Feedstock potential for advanced renewable fuels is very large, but
production facilities at the commercial scale are still limited (EC 2017). However,
advanced biofuels production at large scale is yet to be developed in the EU
(European Parliament 2015; UFOP 2017). Only Italy established a mandate for the
use of advanced biofuels (gasoline and diesel must contain at least 1.2% of advanced
biofuel, with the aim to increase to 2% by 2022) and Denmark (0.9% by 2020). Since
2017, the GHG emissions saving by the use of biofuels should be at least 50% and,
by 2018, 60% for biofuels produced in new installations of the EU.
The development of biofuels markets, especially for advanced biofuels technol-
ogies, is dependent on crude oil prices and policy decisions, which cause uncer-
tainties about future biofuel developments (Avagyan 2018; IEA 2017a; OECD-FAO
2017; WEO 2013; 2016). Since August 2014, the substantial decline in the cost of oil
has dipped below $0.25/l, putting a dent in the biofuel market (Avagyan 2018; Casey
2017; IEA 2017a). The 27% transportation fuel replacement by biofuel will allow
mitigation of emissions at 2.1 GtCO2e/year by 2050 (IEA 2017a). For this purpose, it
is necessary to create a stable long-term policy framework to motivate investors,
increase end-user confidence, and allow the rapid expansion of this industry (IEA
2017a).
Algae have remained the most effective tool for the primary accumulation of
bioenergy (Avagyan 2008, 2010, 2012–2013, 2018; Sayre 2010). Algal biomass
also offers a low risk of indirect land use change and does not compete directly for
agricultural land for the food and feed markets, thus being recognized as a better
future fuel feedstock (Avagyan 2008, 2010, 2012–2013, 2018; European Parliament
2015; IEA 2017b; Walsh et al. 2016). Microalgae have higher oil yields, as much as
3- to 17 fold compared with oil-producing terrestrial plants (Avagyan 2008, 2010,
2012–2013, 2017, 2018; Freyberg 2012; IEA 2017b; Patel et al. 2016; Piloto-
Rodríguez et al. 2017; Santanu 2017). Algae can produce more than 30 times
more energy per unit area in comparison with 1G biofuel, with oil yield at
100 times more per acre than soybeans or other terrestrial oil-producing crops
(Avagyan 2012–2013, 2018; Future of Working 2016). It is apparent that microalgae
biodiesel is also very competitive by water footprint compared to other conventional
feedstocks and especially less for the microalga Chlorella vulgaris (Avagyan 2018;
Yang et al. 2011). The global algae biofuel market volume is expected to be
References 9

$5.96 billion in 2018 and will reach $10.73 billion by 2025 (GVR 2017). The total
global demand for liquid fuels can be achieved through microalgae cultivation on an
area of 1.92 million km2, about 21% of US land area (Greene et al. 2016).
Microalgae biodiesel also has technical advantages compared to lignocellulosic
biodiesel, with a higher caloric value (30 and 29 MJ/kg for Chlorella
(Auxenochlorella) protothecoides and Microcystis aeruginosa, respectively) and
lower viscosity and density than plant-based biodiesel (Costa and de Morais
2011). For microalgae with a lipid content of about 40%, biodiesel yields may
reach 40–50 ton/ha/year, which exceeds the most promising yields from land-
based crops (Schlagermann et al. 2012).
Simultaneously, the UK Royal Academy of Engineering estimated that some
biofuels, such as diesel produced from food crops, have led to more emissions than
those produced by the fossil fuels (Carrington 2017). On December 4, 2017,
178 Netherlands scientists sent an open letter to the Dutch cabinet with the request
that food crop-based biofuels be deleted from the European energy policy because of
these increased GHG emissions (Lane 2017). Therefore, it is necessary to reevaluate
the full cycle of biodiesel production to find optimal solutions for the fuels dilemma.

References

Adnan H (2015) Market analysis. Universiti Teknologi Malaysia. http://www.utm.my/trans-


marinebiodiesel2015/market-analysis/
Araújo K, Mahajan D, Kerr R, da Silva M (2017) Global biofuels at the crossroads: an overview of
technical, policy, and investment complexities in the sustainability of biofuel development.
Agriculture 7:32. https://doi.org/10.3390/agriculture7040032
Atabani AE, Silitonga AS, Badruddin IA, Mahlia TMI, Masjuki HH, Mekhilef S (2012) A
comprehensive review on biodiesel as an alternative energy resource and its characteristics.
Renew Sustain Energy Rev 16:2070–2093. https://doi.org/10.1016/j.rser.2012.01.003
Avagyan AB (2008) A contribution to global sustainable development: inclusion of microalgae and
their biomass in production and bio cycles. Clean Technol Environ Policy 10:313–317. https://
doi.org/10.1007/s10098-008-0180-5
Avagyan AB (2010) New design & build biological system addressed to global environment
management and sustainable development through including microalgae and their biomass in
production and bio cycles. J Environ Prot 1:183–200. https://doi.org/10.4236/jep.2010.12023
Avagyan AB (2011) Water global recourse management through the use of microalgae addressed to
new design & build biological system and sustainable development. Clean Technol Environ
Policy 13:431–445. https://doi.org/10.1007/s10098-010-0321-5
Avagyan AB (2012–2013) Theory of global sustainable development based on including of
microalgae in bio and industrial cycles. New design and building of biological system. Amazon.
ISBN-13: 978-1484000335, ISBN-10: 1484000331, ASIN: B00A7BIV9O
Avagyan AB (2017) Environmental building policy by the use of microalgae and decreasing of
risks for Canadian oil sand sector development. Environ Sci Pollut Res 24:20241–20253.
https://doi.org/10.1007/s11356-017-9864-x
Avagyan AB (2018) Algae to energy and sustainable development. technologies, resources,
economics and system analyses. New design of global environmental policy and live conserve
industry. Amazon, ISBN-13: 978-1718722552, ISBN-10: 1718722559, ASIN: B07DFQBFFD,
209 pp
10 1 Introduction. Links to International Policy and Markets

BDI (2017) Production statistics. http://biodiesel.org/production/production-statistics


BP (2017) Statistical review of world energy June 2017. https://www.bp.com/content/dam/bp/en/
corporate/pdf/energy-economics/statistical-review-2017/bp-statistical-review-of-world-energy-
2017-full-report.pdf
Bracmort K, Schnepf R, Stubbs M, Brent D, Yacobucci BD (2011) Cellulosic biofuels: analysis of
policy issues for Congress. https://fas.org/sgp/crs/misc/RL34738.pdf
Bugarski AD, Janisko SJ, Cauda EG, Noll JD, Mischler SE (2011) Diesel aerosols and gases in
underground mines: guide to exposure assessment and control. Report of Investigations 9687.
DHHS (NIOSH) Publication No. 2012–101. https://www.cdc.gov/niosh/mining/userfiles/
works/pdfs/2012-101.pdf
Carrington D (2017) Biofuels needed but some more polluting than fossil fuels, report warns. The
Guardian. https://www.theguardian.com/environment/2017/jul/14/biofuels-need-to-be-
improved-for-battle-against-climate-change
Casey T (2017) $8 million for algae biofuel from U.S. Department of Energy. TriplePundit. http://
www.triplepundit.com/2017/07/8-million-algae-biofuel-u-s-department-energy/
Coady D, Parry I, Sears L, Shang B (2017) How large are global fossil fuel subsidies? World Dev
91:11–27. https://doi.org/10.1016/j.worlddev.2016.10.004
Costa JAV, de Morais MG (2011) The role of biochemical engineering in the production of biofuels
from microalgae. Bioresour Technol 102:2–9. https://doi.org/10.1016/j.biortech.2010.06.014
Doukas A, De Angelis K, Ghio N, Trout K, Bast E, Bossong K (2017) Talk is cheap: how G20
governments are financing climate disaster. Oil Change International, U.S. Friends of the Earth,
Sierra Club, WWF European Policy Office. http://priceofoil.org/content/uploads/2017/07/talk_
is_cheap_G20_report_July2017.pdf
EC (2010) Annex 5: subsidy level indicators for the case studies. http://ec.europa.eu/environment/
enveco/taxation/pdf/Annex%205%20-%20Calculations%20from%20the%20case%20studies.
pdf
EC (2017) Revision for phase 4 (2021–2030). https://ec.europa.eu/clima/policies/ets/revision_en,
https://ec.europa.eu/clima/policies/ets/allowances_en, https://ec.europa.eu/clima/news/eu-emis
sions-trading-system-landmark-agreement-between-parliament-and-council-delivers-eus_en
EIA (U.S. Energy Information Administration) (2017a) U.S. renewables consumption,
U.S. Department of Energy. https://www.eia.gov/renewable/
EIA (2017b) Monthly biodiesel production report. U.S. Department of Energy. https://www.eia.
gov/biofuels/biodiesel/production/
ENMC (2018) Biodiesel prices in Portugal and in Europe. http://www.enmc.pt/en-GB/activities/
biofuels/indicators/biodiesel-prices-in-portugal-and-in-europe/, http://www.enmc.pt/static-img/
2017-02/2017-02-01152931_f7664ca7-3a1a-4b25-9f46-2056eef44c33$$72f445d4-8e31-416a-
bd01-d7b980134d0f$$69d43ccf-ad8b-4c6a-a3e9-109a63af93da$$File$$pt$$1.pdf
EPA (2015) Green remediation best management practices: clean fuel & emission technologies for
site cleanup. https://www.epa.gov/sites/production/files/2015-04/documents/clean-fuel-emis-
gr-fact-sheet.pdf
EPA (2017) Global greenhouse gas emissions data. https://www.epa.gov/ghgemissions/global-
greenhouse-gas-emissions-data
EU (2015) Directive (EU) 2015/1513 of the European Parliament and of the Council of 9 September
2015 amending Directive 98/70/EC relating to the quality of petrol and diesel fuels and
amending Directive 2009/28/EC on the promotion of the use of energy from renewable sources.
http://eur-lex.europa.eu/legal-content/EN/TXT/?uri=CELEX%3A32015L1513
European Parliament (2015) Briefing: EU biofuels policy. Dealing with indirect land use change.
http://www.europarl.europa.eu/RegData/etudes/BRIE/2015/545726/EPRS_BRI(2015)
545726_REV1_EN.pdf
European Parliament (2017) MEPs call for clampdown on imports of unsustainable palm oil and
use in biofuel. http://www.europarl.europa.eu/news/en/press-room/20170329IPR69057/meps-
call-for-clampdown-on-imports-of-unsustainable-palm-oil-and-use-in-biofuel
References 11

FAO-IPCC (2017) Expert meeting on climate change, land use and food security. Rome, Italy,
January: 23–25. http://www.ipcc.ch/pdf/supporting-material/EM_FAO_IPCC_report.pdf
Fernández AFG, González-López CV, Fernández SJM, Molina GE (2012) Conversion of CO2 into
biomass by microalgae: how realistic a contribution may it be to significant CO2 removal? Appl
Microbiol Biotechnol 96(3):577–586. https://doi.org/10.1007/s00253-012-4362-z
Flach B, Lieberz S, Rossetti A (2017) EU biofuels annual 2017. https://gain.fas.usda.gov/Recent%
20GAIN%20Publications/Biofuels%20Annual_The%20Hague_EU-28_6-19-2017.pdf
Freyberg T (2012) Micro organisms: maximum biogas. Waste Manage World May–June: 44–46.
http://www.waste-management-world.com/index/display/article-display/7705280701/articles/
waste-management-world/volume-13/issue-3/features/micro-organism-maximum-biogas.html
Future of Working (2016) 7 advantages and disadvantages of algae biofuel. https://futureofworking.
com/7-advantages-and-disadvantages-of-algae-biofuel/
GII (Global Information, Inc.) (2017) Biofuels market forecast 2017–2024. https://www.
giiresearch.com/report/ink451147-global-biofuels-market-forecast.html
Green Air (2015) Japanese initiative plots map to commercialisation of aviation biofuels in time for
2020 Tokyo Olympics. http://www.greenaironline.com/news.php?viewStory=2105
Greene CH, Huntley ME, Archibald I, Gerber LN, Sills DL, Granados J, Tester JW, Beal CM,
Walsh MJ, Bidigare RR, Brown SL, Cochlan WP, Johnson ZI, Lei XG, Machesky SC, Redalje
DG, Richardson RE, Kiron V, Corless V (2016) Marine microalgae: climate, energy, and food
security from the sea. Oceanography 29:10–15. https://doi.org/10.5670/oceanog.2016.91
GVR (Grand View Research) (2017) Algae biofuel market estimates & trend analysis by applica-
tion (Transportation, Others), by region (North America, Europe, Asia Pacific, Rest of World),
by country, and segment forecasts, 2018–2025. http://www.marketwatch.com/story/algae-bio
fuel-market-worth-1073-billion-by-2025-growth-rate-88-grand-view-research-inc-2017-03-20-
5203127
IBRD-WB-IEA (International Bank for Reconstruction and Development, World Bank, Interna-
tional Energy Agency) (2017) Sustainable energy for all global tracking framework progress
toward sustainable energy. https://doi.org/10.1596/978-1-4648-1084-8. http://gtf.esmap.org/
data/files/download-documents/eegp17-01_gtf_full_report_for_web_0516.pdf
IEA (2017a) World energy investment 2017. https://www.iea.org/publications/wei2017/
IEA (2017b) State of technology review – algae bioenergy. In: IEA bioenergy inter-task strategic
project, Task 39. http://www.ieabioenergy.com/wp-content/uploads/2017/02/IEA-Bioenergy-
Algae-report-update-Final-template-20170131.pdf
Kotrba R (2017) Global biodiesel market to surpass $41 billion by 2021. https://www.
grandviewresearch.com/press-release/global-biodiesel-market
Lane J (2012) RFS waiver could raise feed costs for livestock, dairy producers: new study. Biofuels
Digest. http://www.biofuelsdigest.com/bDigest/2012/10/11/rfs-waiver-could-raise-feed-costs-
for-livestock-dairy-producers-new-study/
Lane J (2017) Letter sent to Dutch cabinet by 178 Dutch scientists urges end to food-based biofuels.
Biofuels Digest. http://www.biofuelsdigest.com/bdigest/2017/12/04/letter-sent-to-dutch-cabi
net-by-178-dutch-scientists-urges-end-to-food-based-biofuels/, http://brandstofbrief.nl/
Lima MA, Linhares FG, Mothe GA, Castro MPP, Sthel MS (2017) Study of gaseous emissions
derived from the combustion of diesel/beef tallow biodiesel blends. Sustain Environ 2:210–222.
https://doi.org/10.22158/se.v2n2p210
Liu X, Saydah B, Eranki P, Colos LM, Mitchel BG, Rhodes J, Clarens AF (2013) Pilot-scale data
provide enhanced estimates of the life cycle energy and emissions profile of algae biofuels
produced via hydrothermal liquefaction. Bioresour Technol 148:163–171. https://doi.org/10.
1016/j.biortech.2013.08.112
MR & C (Merchant Research & Consulting, Ltd) (2017) Biodiesel: 2017 world market outlook and
forecast up to 2021. https://mcgroup.co.uk/researches/biodiesel
NBB (2017) What is biodiesel? http://biodiesel.org/what-is-biodiesel/biodiesel-basics
12 1 Introduction. Links to International Policy and Markets

NBB (U.S. National Biodiesel Board) (2016) Scientists agree biodiesel a key to global carbon
reduction. http://biodiesel.org/news/news-display/2016/05/17/scientists-agree-biodiesel-a-key-
to-global-carbon-reduction
New P (2017) World market for biofuels. In: 7th international congress and expo on biofuels &
bioenergy date and venue, Toronto, Canada. https://biofuels-bioenergy.conferenceseries.com/
market-analysis-pdfs/market%20analysis%202%20nd%20conf161105011838.pdf
OCI (Oil Change International) (2017) Fossil fuel subsidies overview. http://priceofoil.org/fossil-
fuel-subsidies/
OECD (2017) Green growth indicators 2017. http://www.oecd.org/environment/indicators-model
ling-outlooks/Highlights_Green_Growth_Indicators_2017.pdf
OECD-FAO (2017) Agricultural Outlook 2017–2026. https://doi.org/10.1787/agr_outlook-2017-
en
Patel B, Guo M, Izadpanah A, Shah N, Hellgardt K (2016) A review on hydrothermal pre-treatment
technologies and environmental profiles of algal biomass processing. Bioresour Technol
199:288–299. https://doi.org/10.1016/j.biortech.2015.09.064
Piloto-Rodríguez R, Sánchez-Borroto Y, Melo-Espinosa EA, Verhelst S (2017) Assessment of
diesel engine performance when fueled with biodiesel from algae and microalgae: an overview.
Renew Sustain Energy Rev 69:833–842. https://doi.org/10.1016/j.rser.2016.11.015
Rahman M (2015) Influences of biodiesel chemical compositions and physical properties on engine
exhaust particle emissions. Queensland University of Technology. https://eprints.qut.edu.au/
82754/8/Md_Rahman_Thesis.pdf
REN21 (Renewable Energy Policy Network for the 21st century) (2017) Renewables 2017 global
status report (Paris: REN21 Secretariat). ISBN 978-3-9818107-6-9. http://www.ren21.net/wp-
content/uploads/2017/06/17-8399_GSR_2017_Full_Report_0621_Opt.pdf
Rocca S, Agostini A, Giuntoli J, Marelli L (2015) Biofuels from algae: technology options, energy
balance and GHG emissions. EU JRC. https://doi.org/10.2790/125847
Santanu B (2017) Evaluating sustainable economic development. Clean Technol Environ Policy
19:1815–1816. https://doi.org/10.1007/s10098-017-1400-7
Sayre R (2010) Microalgae: the potential for carbon capture. Bioscience 60:722–727. https://doi.
org/10.1525/bio.2010.60.9.9
Schlagermann P, Göttlicher G, Dillschneider R, Rosello-Sastre R, Posten C (2012) Composition of
algal oil and its potential as biofuel. J Combust 2012:14. https://doi.org/10.1155/2012/285185
Singgih VP (2018) More companies to enjoy biodiesel subsidy. The Jakarta Post. http://www.
pressreader.com/indonesia/the-jakarta-post/20180102/281526521435571
Transport Policy (2017) EU: fuels: biofuel specifications. http://www.transportpolicy.net/standard/
eu-fuels-biofuel-specifications/
Troustle R, Marti D, Rosen S, Westcott P (2012) Why have food commodity prices risen again?
USDA. http://www.ers.usda.gov/media/126752/wrs1103.pdf.http://www.national-economists.
org/gov/trostle11.pdf
U.S. Department of Energy (2014) FY 2008/2009 progress report for fuels technologies. Energy
efficiency and renewable energy office of vehicle technologies. https://www.energy.gov/sites/
prod/files/2014/03/f8/2008-2009_fuels_technologies.pdf
U.S. Department of Energy (2016) 2016 billion-ton report: advancing domestic resources for a
thriving bioeconomy, volume 1: economic availability of feedstocks. In: Langholtz MH, Stokes
BJ, Eaton LM (leads), ORNL/TM-2016/160. Oak Ridge National Laboratory, Oak Ridge,
TN. 448 pp. doi:https://doi.org/10.2172/1271651
UFOP (Union for the promotion of oil and protein plants) (2017) Biodiesel 2016/2017 report on
progress and future prospects – excerpt from the UFOP annual report. https://www.ufop.de/
files/5115/1309/0426/UFOP-Biodiesel_2016-2017_EN.pdf
van Dyk JS, Li L, Leal DB, Hu J, Zhang X, Tan T, Saddler J (2016) The potential of biofuels in
China. IEA. http://task39.sites.olt.ubc.ca/files/2013/05/The-Potential-of-biofuels-in-China-
IEA-Bioenergy-Task-39-September-2016.pdf
References 13

Walsh MJ, Van Doren LG, Sills DL, Archibald I, Beal Colin M, Lei XG, Huntley ME, Johnson Z,
Greene CH (2016) Algal food and fuel coproduction can mitigate greenhouse gas emissions
while improving land and water-use efficiency. Environ Res Lett 11:114006. http://iopscience.
iop.org/article/10.1088/1748-9326/11/11/114006/meta
WEC (World Energy Council) (2016) World energy perspectives. Non-tariff measures: next steps
for catalysing the low carbon economy. https://www.worldenergy.org/wp-content/uploads/
2016/08/Full-report__Non-tariff-measures_next-steps-for-catalysing-the-low-carbon-economy.
pdf
WEF (World Economic Forum) (2018) The global risks report 2018. http://www3.weforum.org/
docs/WEF_GRR18_Report.pdf
WEO (2016) Part B: Special focus on renewables. IEA. https://www.iea.org/media/publications/
weo/WEO2016SpecialFocusonRenewableEnergy.pdf
WEO (World Energy Outlook) (2013) Water for a sustainable world. IEA. https://www.iea.org/
publications/freepublications/publication/WEO2013.pdf
WER (World Energy Resources) (2016) World energy council. https://www.worldenergy.org/wp-
content/uploads/2017/03/WEResources_Bioenergy_2016.pdf
WWDR (2012) Managing water report under uncertainty and risk. UNESKO. http://www.unesco.
org/new/en/natural-sciences/environment/water/wwap/wwdr/wwdr4-2012/
WWDR (2017) Wastewater: the untapped resource. United Nations World Water Assessment
Programme, Paris. http://www.unwater.org/publications/world-water-development-report-
2017/
Xie X, Zhang T, Wang L Huang Z (2017) Regional water footprints of potential biofuel production
in China. Biotechnol Biofuels 10:95. https://doi.org/10.1186/s13068-017-0778-0
Yang J, Xu M, Zhang X, Hu Q, Sommerfeld M, Chen Y (2011) Life-cycle analysis on biodiesel
production from microalgae: water footprint and nutrients balance. Bioresour Technol
102:159–165. https://doi.org/10.1016/j.biortech.2010.07.017
Chapter 2
Biodiesel from Plant Oil and Waste
Cooking Oil

Abstract Vegetable oil derived from oilseed plantations or crops is the most
commonly used feedstock for the production of biodiesel. These oils, primarily
including those obtained from rapeseed, sunflower, and palm, are transesterified
with methanol in the presence of an alkaline catalyst to reduce their viscosity so that
their fuel properties are comparable to those of diesel fuel. Because the techno-
economic viability and overall sustainability of advanced biofuels produced from
algae or lignocellulosic biomass are yet to be proved, the oilseed plants are likely to
dominate the scene in the near future as well. Utilization of edible oils grown on
agricultural land has led to the infamous “food versus fuel” dilemma. Developing
nations such as India that rely heavily on imported edible vegetable oil cannot afford
to divert their agrarian land/produce toward biodiesel production. National policy,
accordingly, has restricted the development of biodiesel feedstock plantations
(nonedible oilseed-bearing plants) to wastelands or marginal lands. However,
large-scale alteration of the ecologically diverse landscape should be avoided at
any cost as that could negate the positive attributes of biodiesel production. The
energy return on energy investment (and, more recently, exergy) has become a vital
sustainability indicator for alternative sources of energy. A wide range of energy
return values for biodiesel production has been reported in the literature, but there is
a general sense of agreement of its sustainability for nonedible plant oil-based
biodiesel production. The high production cost of biodiesel is a significant imped-
iment for its successful commercialization. The high cost of production is mainly
attributed to the cost of feedstock (70–80%), and as a result, there is a growing need
for diversion of recycled vegetable oil toward biodiesel production. This concern
mainly holds true for China where almost the entire biodiesel production is derived
from waste cooking oil. A few techno-economic studies have also highlighted the
impact of the choice of catalyst on the overall capital investment and manufacturing
cost. The catalyst affects the process in terms of the degree of conversion of the feed
to biodiesel and the downstream purification requirements. Efficient heterogeneous
catalysts appear to be economically and environmentally more appealing than their
homogeneous counterparts. Biodiesel is safer than diesel and offers easy handling
and transport options. Depending on the fatty acid profile of the feedstock used, the
long-term storage of biodiesel can be problematic. However, the stability of biodie-
sel can sometimes be improved using blending, winterization, and antioxidant

© Springer Nature Singapore Pte Ltd. 2019 15


A. B. Avagyan, B. Singh, Biodiesel: Feedstocks, Technologies, Economics and
Barriers, https://doi.org/10.1007/978-981-13-5746-6_2
16 2 Biodiesel from Plant Oil and Waste Cooking Oil

additives. Despite the general belief in biodiesel as a sustainable fuel, life cycle
assessment studies are needed to account for the direct and indirect impacts that
often are feedstock- and location specific. Although there are certain deterrents to the
mass-scale production of biodiesel, there are numerous opportunities to address the
associated concerns.

Keywords Fossil fuels · Climate change · Renewable energy · Biodiesel · Vegetable


oil · Biofuel policy

2.1 Properties of Biodiesel and Production

Energy is needed in all sectors of society. Energy can be derived from either
renewable or nonrenewable sources. The transport sector is dependent on
nonrenewable fuels because the percentage of renewables has been quite low, except
for liquid biofuels. Supported by policy measures, production of liquid biofuels is on
the rise, and the industry is witnessing unprecedented growth worldwide (Tyner
2008). Among the renewable fuels available, liquid fuels have an important role in
providing energy security. Existing fleets of vehicles are overwhelmingly powered
by internal combustion engines that are predominantly fueled by liquid petroleum
products. The transport sector remains one of the largest consumers of liquid
petroleum fuels (Maibach et al. 2008).
All the major producers (and exporters) of biodiesel derive their biodiesel from
edible agricultural produce grown on arable land, except China, which is the leading
producer of waste cooking oil-derived biodiesel with an annual production of more than
1 billion liters. The choice of feedstock for biodiesel production depends on its
availability in a particular region and the economic value of the feedstock. Apart
from plants and animals, the lipid also can be obtained from other organisms, including
molds, yeast, fungi, and algae that can accumulate up to more than 70% (w/w)
intracellular lipids. The oleaginous microorganisms can be utilized for large-scale
lipid fermentation (Zhao et al. 2018). The lignocellulosic biomass can be converted
to biodiesel in three steps: (1) hydrolysis of lignocellulosic biomass to carbohydrate,
(2) fermentation of a carbohydrate with microorganisms to produce oil, and (3) conver-
sion of microbial oil to biodiesel through transesterification (Yousuf 2012). However,
the techno-economic viability of biodiesel derived from lignocellulosic biomass is yet
to be ascertained (Klein-Marcuschamer et al. 2010). Similarly, the cost-competitiveness
and environmental dimensions of large-scale biodiesel production from microalgae are
still a matter of extensive research (Nagarajan et al. 2013; Beal et al. 2015).
One of the most attractive attributes of biodiesel is its miscibility in diesel, and
lower biodiesel blends (up to 20%) have properties comparable to diesel (Ma and
Hanna 1999). This property allows the use of lower biodiesel blends in existing
compression ignition internal combustion engines without modification. Although
denser, the energy density of biodiesel is lower than diesel (by approximately 12%
per gallon), and hence more is needed in the combustion chamber to generate
comparable heat (Tyson and McCormick 2006). Straight vegetable oil (SVO) has
2.1 Properties of Biodiesel and Production 17

a much higher viscosity than diesel, and its use in a diesel engine is not
recommended. Formation of carbon deposits, poor pumping, and atomization are
some of the significant troubles commonly encountered when SVOs are combusted
in a diesel engine (Misra and Murthy 2010). Biodiesel is produced when an
oleaginous feedstock is subjected to a chemical reaction, called transesterification,
in the presence of small-chain alcohol to break the long chain of triacylglycerol into
three small chains of monoalkyl esters (biodiesel) (Meher et al. 2006). In this
process, the viscosity of the parent material is reduced by a factor of 10 to 15. In
addition to biodiesel, glycerol is also obtained, which after refining could be
valorized in numerous applications (Ullah et al. 2016).
Biodiesel can be made from a wide variety of materials including vegetable oils,
animal fats, and any material that contains triacylglycerols or fatty acids. Feedstocks
have different types and amounts of fatty acids. The fatty acids have different
properties in terms of their melting point, caloric value, cetane number, vulnerability
to oxidation, etc., and thus the fuel properties of biodiesel developed from different
feedstocks will vary (Knothe and Razon 2017). The viscosity of biodiesel must be
low enough that it does not pose a problem during combustion in engines. Viscous
biodiesel may cause an alteration in the injection spray characteristics of the fuel.
Hence, the fuel pump might be damaged by viscous biodiesel.
Apart from being renewable, the major advantages associated with biodiesel are a
substantial reduction in emission of pollutants that contribute to global warming and
are also deleterious to human and plant health. When compared to diesel, biodiesel
has a better lubricant property, higher cetane number, better biodegradability,
negligible emission of sulfur and aromatics, higher flash point, and substantially
less emission of carbon monoxide, unburned hydrocarbons, and soot compounds on
combustion (Demirbas 2009). Further, energy security, employment opportunities,
and wasteland reclamation potential add to the advantages of biodiesel production.
However, comparatively high production cost, good compatibility only for lower
blends, the requirement of land for oilseed plantation and associated land use and
land cover changes, modest energy and greenhouse gas balance, poor cold flow
properties, and often a higher emission of nitrogen oxides are major challenges
facing the biodiesel industry (Bozbas 2008).
At the present time, approximately 34% of energy is derived from liquid fuel,
including all types of biofuels (Amelio et al. 2016). The production of fatty acid
alkyl ester is not new and dates back to the 1980s, when researchers in South Africa
produced biodiesel from edible oil (sunflower seeds) by transesterification.
Biodiesel is a renewable fuel that could provide energy security and energy
independence to countries across the globe. However, the properties of biodiesel
and the methods adopted in its production are of importance in its sustainability and
acceptability among users. The diesel engine was originally designed to run on a
variety of fuels including vegetable oil (Knothe 2010). After the discovery of diesel
(fuel), the design of the engine was altered to run on diesel. As a result, using straight
vegetable oil in unmodified diesel engines may cause problems. The challenges arise
particularly in terms of the higher viscosity of vegetable oils. Finite availability,
uneven spatial distribution, environmental pollution, and climate change are among
the major drivers for the exploration and development of alternative forms of energy.
18 2 Biodiesel from Plant Oil and Waste Cooking Oil

Fatty
Acid FAME

Fatty Glycerol Catalyst Glycerol


Acid 3 Methanol FAME
Heat+Stirring
Fatty
Acid FAME

Biodiesel
Triacylglycerol Fatty Acid Methyl Ester
Major Constituent of (FAME)
Vegetable oil/animal fat

Fig. 2.1 Scheme of catalytic transesterification

These challenges have again diverted global attention toward vegetable oil as a
renewable substitute for diesel. The viscosity of vegetable oils is approximately 10 to
15 times higher than diesel and their direct use in diesel engines is not recommended.
High viscosity leads to poor pumping and atomization of the fuel and greater buildup
of carbon deposits in the combustion chamber. Several attempts have been made to
lower the viscosity of vegetable oil such as blending it with diesel, microemulsion,
pyrolysis of vegetable oil, hydro-processing, and transesterification (Bezergianni
et al. 2010).
Among these approaches, the most extensively studied method is the
transesterification of vegetable oil, which transforms each triester (triacylglycerol,
TAG) molecule in vegetable oil to three molecules of monoalkyl esters of long-chain
fatty acids, and a molecule of glycerol is released as a by-product. These fatty acid
alkyl esters are popularly known as biodiesel. Because vegetable oil is composed of
different fatty acids, ranging in carbon chain length from 14 to 24, biodiesel derived
from the transesterification of vegetable oil is invariably a mixture of fatty acid
esters. The general scheme of transesterification is shown in Fig. 2.1.
Transesterification, as the name suggests, involves the transfer of ester linkages
between the fatty acids and glycerol in the TAG to that between fatty acids and
monohydric alcohol. The parent molecule is a triester whereas the biodiesel mole-
cule has only one ester linkage (Ma and Hanna 1999). After transesterification, the
molecular weight of vegetable oil is reduced from its typical initial range of
860–900 g mol1 to 270–290 g mol1 whereas the corresponding decrease in
viscosity is approximately 10- to 12 fold. Transesterification of the oleaginous
feedstock is performed in the presence of an alkoxy group donor alcohol (usually
methanol or ethanol), and a catalyst is used to accelerate the rate of the reaction.
However, a noncatalytic approach popularly known as supercritical
transesterification is increasingly gaining popularity (Meher et al. 2006). The super-
critical conditions involve high temperature and high pressure to maintain methanol
in its supercritical state, which facilitates the conversion of the feed to biodiesel. As
no catalyst is involved in supercritical transesterification, the downstream purifica-
tion operations are simplified. Despite the advantages offered by the supercritical
2.1 Properties of Biodiesel and Production 19

process, the catalyzed processes remain the most commonly adopted strategy
(Marchetti and Errazu 2008). The preference mainly lies in the lesser amount of
energy required for the catalyzed process. The catalyzed transesterification process
can use either homogeneous or heterogeneous catalysts. Homogeneous catalysis has
been performed using alkali metal hydroxides/methoxide (alkaline catalysts) and
mineral acids (acidic catalysts). There are inherent advantages and disadvantages in
both types of homogeneous catalysts. Alkaline catalysts, albeit highly efficient, are
highly sensitive to free fatty acids (FFAs) and moisture content in the feedstocks
(Leung et al. 2010). Researchers have advised against the direct use of alkaline
catalysts for feedstocks rich in FFAs and moisture as the presence of excess FFAs
leads to saponification whereas moisture promotes hydrolysis of the feedstock
(Ma and Hanna 1999), so prior removal of excess moisture and FFA management
is necessary. Two-step transesterification for FFA-rich feedstocks is recommended.
In the first step the FFAs are allowed to react with methanol in the presence of an
acidic catalyst (H2SO4), converting the FFAs to fatty acid methyl ester (FAME).
Excess mineral acid is removed by glycerol wash and the water formed by esterifi-
cation is separated out. The esterified oil is then amenable to alkaline
transesterification. Acidic catalysts, on the other hand, are insensitive to FFAs and
moisture in the feedstock but require higher reaction time, higher temperature, and a
higher methanol-to-oil ratio. Further, utilization of acidic catalysts necessitates the
use of corrosion-resistant storage tanks and reactors and incorporation of additional
safety measures (Leung et al. 2010).
Because the homogeneous catalyst remains in the same phase as the oil, its
recovery and reuse are a daunting task. Moreover, the catalyst fractions need to be
separated from crude biodiesel. These challenges have spurred worldwide interest in
the exploration and development of heterogeneous catalysts (Lam et al. 2010). A
heterogeneous catalyst offers reusability, easy recovery, and less complicated down-
stream purification operations (Helwani et al. 2009a). Numerous types of heteroge-
neous catalysts have been tested with varying degrees of popularity in terms of
efficiency, selectivity, reusability, and catalyst lixiviation (Helwani et al. 2009b).
The conversion of any given feedstock for a given catalyst is primarily dependent
on the oil-to-alcohol molar ratio, reaction time, and reaction temperature. Hence, for
any combination of catalyst and feedstock, these variables require optimization to
determine the optimal combination of factor levels and also to ensure efficient use of
resources (Leung and Guo 2006). Optimization experiments have been performed
using the traditional approach of varying one factor at a time while keeping the
others constant. However, to ensure efficient use of the collected data, advanced
statistical operations based on response surface methodology are gaining popularity
(Bakkiyaraj et al. 2016). Matlab, Minitab, and Design Expert are common software
platforms for the designing and analysis of experiments.
Stoichiometrically, a 1:3 molar ratio of oil to alcohol is sufficient for the complete
conversion of the TAG to biodiesel. As transesterification is a reversible reaction, an
excess of alcohol is invariably involved to promote the forward reaction. The
resulting excess methanol must be recovered to ensure compliance with biodiesel
standards. Apart from three molecules of biodiesel, each molecule of TAG upon
Another random document with
no related content on Scribd:
Coppieters, 146

Cornelia, 211, 213, 218

Cornelis, 150, 210, 213, 228, 249

Cornelis Vriens, 178

de Corte, 141

Cortebosc, 120

Cortvriend, 141

de Coster, 134, 140

Cottun, 110, 111, 113

Courtebourne, 118

Cousebourne, 118

de Coussemaker, 143

de Craecker, 141

Cromatic, 120

Crudeniers van Oostburch, 77, 78

Cuern, Cuerne, Kuren, 74

Cunibert, 162

Cuonbercht, 162

Cuonrath, 162

Cupers van Damme, 76

Curen, Kuren, 74

de Curte, 137, 141


Curtebrona, Curtebrune, 118

Curtrycke, Kortrijk, 76

Cusebrona, 118

de Cuupere, 140

van Cuyck, 137, 138

de Cuyper, 138, 140

Cyrick, 132

Daenel van Bruheze, 180

Dagbert, 134

Daknam, 73

Dalem, 126

Dalen, 87

Dalhem, 126

Dalle, 117

Damme, 76

Dammekens, 146

Dams, 146

Dansers van Everghem, 81

Daryncbarners van den Vryen, 76, 79

De Backer, 140

— Backere, 138, 140

— Baecker, 134
— Beer, 143

— Beyer, 134

— Bie, 143

— Bisschop, 140

— Blauwe, 141

— Bleecker, 139

— Bleye, 143

— Boeve, 141

Deborah, 251

De Borchgrave, 143 [300]

De Brabander, 143

— Brauwere, 143

— Broere, 143

— Bruycker, 141

— Bruyne, 141

— Buck, 143

— Caesemaecker, 140

— Caluwe, 141

— Ceuninck, 138, 140

— Clerk, De Clercq, 140

— Coen, 142

— Coninck, 140
— Conynck, 134

— Cooman, 140

— Corte, 141

— Coster, 134, 140

— Coussemaker, 134

— Craecker, 141

— Curte, 137, 141

— Cuupere, 140

— Cuyper, 138, 140

— Dapper, 142

— Decker, 140

Dedda, 231

Dedde, 212

Dedmer, Dethmar, 232

Deetsje (Deetje), 251

De Dryver, 139

— Gaye, 143

— Gheele, 141

— Ghellinck, 146

— Gheselle, 141

— Graeve, 140

— Grave, 134
— Grendel, 134

— Gryse, 141

— Groote, 134, 141

— Gruyter, 140

— Haas, 137

— Haene, 134

— Jaegher, 140

— Jong, 137

— Jonghe, 138, 141

— Ketelaere, 139

— Keyser, 140

— Kock, 140

— Koninck, 140

— Krijger, 141

— Laeter, 134

— Langhe, 141

— Lantsheere, 140

Delden, 59

De Leener, 141

— Leeuw, 143

— Lepelaere, 143

— Lepeleire, 137, 143


Delft, 6, 66

Delfzijl, 55

De Maesschalk, 140

— Man, 134, 141

— Mangelaere, 139

— Meersman, 140

— Meester, 141

— Meulemeester, 140

— Meulenaere, 137, 140

— Meuleneire, 140

— Meunynck, 134

— Meyer, 140

— Moerloose, 141

— Mol, 134

— Moor, 143

— Mulder, 140

— Munter, 140

— Muynck, 137, 140

— Muyter, 141

— Naeyer, 139

Dendermonde, 73, 76

Den Dooven, 141


Den Duyts, 143

De Neve, 141

Den Haeze, 143

Denterghem, 71, 77

De Pachtere, 140

— Paepe, 140

— Pannemaecker, 139

— Pauw, 143

— Poorter, 134, 141

— Potter, 139

— Praeter, 141

— Prince, 140

— Proost, 140

— Puydt, 143

— Raeve, 143

— Ram, 180

— Ridder, 140

Derk, Dirk, Durk, 232, 249, 251

Derkje, Dirkje, Durkje, 232, 250

Dermonde, Dendermonde, 76

De Rode, 134

— Roo, 141
— Rouck, 143

De Rudder, 140

— Russcher, 141

— Ruysscher, 141

— Ruyter, 140

— Ruywe, 134

— Rycke, 142

— Saegher, 139

— Scheemaecker, 139

— Scheirder, 140

— Schepper, 139

— Schodt, 134

— Schoenmaker, 140

— Schrijver, 140

— Schuyter, 141

— Seeldraeyer, 140

— Sloover, 141

— Smedt, De Smet, 139

— Smidt, 134

— Smyttere, 134

— Spiegelaere, 140

Desschel, 71, 72
De Staercke, 142

— Staute, 142

Desteldonk, 73

De Stuynder, 134

— Surgeloose, 141

— Swarte, 134

— Taeye, 141

— Tavernier, 140

— Temmaeker, 134

Detje, 231

De Turck, 143

Deunynck, 146

De Valck, 143

Deventer, 59, 87

De Vilder, 134

— Vincke, 143

— Vis, 143

— Visscher, 140

— Vliegher, 141

— Vogelaere, De Voghelaere, 140

— Vos, 134, 143

— Vreese, 143
— Vriendt, 141

— Vriese, 143

— Vulder, 141

— Vylder, 139

— Waegemaeker, 134

— Wagenaere, 140

— Waegeneire, 140

— Waele, 143 [301]

De Wandeleer, 141

— Weert, 140

— Wevere, 139

Dew Gowerts, 264

De Witte, 141

— Wolf, 143

Deyaert, 144

Deynse, 77, 81

De Zitter, 134

— Zutter, 140

— Zwaef, 143

D’Haese, 138, 143

D’Hane, 143

D’Hollander, 143
D’Hondt, 143

D’Hoog, 141

D’Hooghe, 137

D’Huyvetter, 137, 139

Dick, 150, 210

Dickedeuters van Dunum, 57

Didde, 250

Diddeken, 177

Didderic Janssoen van den Broeke, 174

Diederik, 217, 232, 249

Diederika, 217, 232

Dienders van Sint-Anna ter Mude, 77

Diependal, 117

Dierick, 146

Diest, 7, 71, 73

Dietwara, 231

Dieuwke, Dieuwertje, 231

Diewertje, 132

Diksmude, 6, 74, 76

Dirc Dieric Gueldensoens soen, 176

Dirck Cransmaker, 268

Dirck Glaesmaecker, 267


Dirck Stadbode, 268

Dirck Steenwerper, 272

Dirck die Weder, 180, 195

Dirc Udemans Swolfssoen, 177, 194

Dirc Wautgers, 178

Dirk, Durk, 217, 226, 232, 249, 251


Dirkje, Durkje, 217, 232

Dirre, 250

Diryck Swerts, 182

Ditmar, Dietmar, 232

Diuva, 231

Dixmude, 76

Djoere, Djurre, 212

Doaike (Dooike), 215

Doaite (Dooite), 214

Doaitse (Dooitse), 215, 258

Doaitsen (Dooitsen), 274

Donye (Dooye), 212

Dobbelaere, 141

Doctor Dionisius Dodo, Doctor Dodo, 266

Doda, 231

Dodden van Niedorp, 63

Dode, 274

Dodo, 229, 230

Doed, 274

Doede, 212, 227, 229, 274


Doederus, 227

Doed Juyssma, 260

Doeke, 214, 215, 221, 229

Doekele, 214, 221, 222, 229

Doeke Martena, 260

Doekje, Doekeltje, 231

Doesburg, 60

Doetje, 231

Doetlage, 121

Dohem, 126

Doitse, 221

Dokkum, 5, 14, 21, 86

Dolislaeger, 140

Dolphyn, 143

Dominicus, 207

Doofpotten van Holwierde, 56

Dordrecht, 66

Douay, 76

Dou Sixz, Douwe Sikkes, 275

Doutsje (Doutje) Douwtsen, 231, 275

Douvres, 110, 111

Douwe, 132, 133, 203, 207, 212, 275


Douwes, 259

Douwtsen, 203, 215, 221, 231, 275

Dowe, 273

Dowe Bottez, 258

Dowe Decker, 268

Dowe Ryoerdsz, 258

Dowe Scomaker, 267

Doythye, 274

Doythye Feyckez, 258

Draadsma, 237, 238, 245

Draaiers van Aalst, 74

Draaisma, 245

Drachten, 36, 237

Dragstra, 236

Drapeniers van Comene, 77

Driehornstic, 120

Driel, 60, 61

Dries, 148

Driessens, 146, 147

Drijvers van Zoersel, 73

Drinkers van Sint-Winoks-Bergen, 6, 72, 76

Driuwpôllen van Woudsend, 37


Dronrijp, 36

Drooghaerts van Werveke, 77, 79, 81

Dubbeltjessnijders van Os, 67

Duco, 215, 229

Duerinck, 146

Duinkerke, 6, 50, 95

Duiveldragers van Hoorn, 63

Duivelshoopen van Aartswoud, 63

Dukke, 217, 232, 249, 250

Dúmkefretters van Sneek, 7, 8, 14, 15, 20

Dunkercke, 76, 82

Dunum, 57

Durinckx, 146

Duva, 231

Duye, 217, 232, 249, 250, 251

Duyff (Duif, Duifje), Duveke, 272, 275

Duyff Jelles, 275

Duyff wedue, 275

Duyvis, 259

Duyfken, 132

Dweilstikken van Jorwerd, 36

Dworp, 73
Dye, 249, 250

Dykema, Dyksma, 240

Dykstra, 237

Dykwoartels van Langweer, 37 [302]

Dyre Cnoopsoen van Zoemeren, 174

Eabe, 212, 221, 229

Eabele, 214, 221, 222, 229

Eade, 212, 229

Eadsger, 200

Eadske, 231

Eage, 212

Eagele, 214

Eale, 212

Ealke, 215

Ealse, 214

Ealtse, 215

Ealtsje, 215

Eartepotten van Baayum, 36

Eartepûlen van Spannum, 36

Easge, 258

Easge Sikkes, 273

Eastbourne, 117
Easterboarn, 117

Eatse, 215

Eauwe (Eeuwe), 212

Ebbele, 214

Ebbing, 216

Ebe, 176, 212

Ebke, 215

Eckington, 105

Edam, 63, 88

Ede, Edo, 211, 212, 229, 230

Ede Kremer, 269

Ede Wagenaer, 269

Edel, Edele, 101, 219

Edelmar, 212

Edgar, Edger, 201, 233

Edington, 105

Edleff, 201

Edou, Edouw, 202, 277

Edse, 214

Edser, 200

Edsert, 200

Edsard, Edzard, 200, 227


Eduard, 211

Edwer, Eedwer, 201

Edzard, Edsard, 200, 227

Ee, 121

Eeckman, 134

Eecloo, 77

van Eeke, 134

Eekje, 250

Eelck, 262

Eelck Onsta, 260

Eelco, Eelke, 101

Eelcoma, 101

Eelke, Eelco, 101, 205, 215, 218, 220, 226, 230, 232, 249

Eelkema, 101

Eelkje, 101, 217, 218, 220, 226, 232, 250

Eelse, 101, 214

Eelsma, 101

Eelsma-state, 101

Eelswert, 101

Eeltjes, 101

Eeltse, 101, 215

Eeltsje (Eeltje), 101, 214, 215, 218, 220, 226, 230, 232
Eendepullen van Oostzaan, 63

Eente, Eento, 214

Eernewoude, 7, 35

Eestrum, 36

Egbert, Eibert, 200, 212, 232

Egbert Kuper, 269

Egbrecht, Ekbrecht, 232

Egga Jellazoen, 273

Egge, Eggo, 132, 212, 229

Egilbald, 201

Egmond aan Zee, 63, 64

Ehlingen, 101

Eidse, 214

Eilard, Eilert, Eilhart, Agilhart, 200, 232

Eilbrand, 201

Eilof, Agilolf, 200

Eise, Eiso, 132, 214, 229, 239

Eisinga, 239

Eisselsdorf, 166

Eite, 214

Eitse, 215

Ek, 60

You might also like