Você está na página 1de 185

BIO-FUELS

Technology Status and Future Trends,


Technology Assessment and Decision Support Tools

prepared by
Sivasamy Arumugam
1
, Sergey Zinoviev
1
, Paolo Foransiero
1
, Stanislav Miertus
1
, Franziska
Mller-Langer
2
, Martin Kaltschmitt
2
, Alexander Vogel
2
, Daniela Thraen
2
, Francis
Kemausuor
1


1
ICS-UNIDO, Trieste, Italy
2
IE-Leipzig, Germany




International Centre for Science and High Technology
United Nations Industrial Development Organization
2008


i

ICS-UNIDO Publication: Prepared within the ICS-UNIDO Work program of 2007,
Pure & Applied Chemistry Area


This text can be used for non-commercial needs only. No selling, distribution, modification,
reproduction, transmission, publication, broadcasting, posting or other use of its contents can
be made without prior permission from the ICS-UNIDO



ii
Preface

One of the key issues of sustainable industrial development is the conversion from fossil feedstocks to
renewable feedstocks in various sectors such as energy production, fuel production, chemical and related
industries.

Such a conversion is driven by several factors such as fossil feedstock depletion, need for diversification
of feedstocks due to abundance of renewable resources in many countries, CO
2
neutrality of renewable
feedstocks, concerted potential development of both industry and agriculture, new openings for green
chemistry and related industries development, etc.

Obviously, there are several problems to be solved in the development of this sector. Among those, there
is a need of further development of suitable technologies for the next generation of bio-fuels, availability
of feedstocks, uncertainty of bio-feedstocks supply and their prices, risk of misconception in designing
bio-fuels strategies. Further development and assessment of various technological options and scenarios
of bio-fuels exploitation is highly needed.

The International Centre of Science and High Technology of the United Nations Industrial Development
Organization (ICS-UNIDO) has been promoting a programme on the knowledge transfer and capacity
building to developing countries in selected priority sectors including the use of renewable resources.

Due to the importance of bio-feedstocks exploitation and bio-fuels production, especially as a potential
sector for developing countries, UNIDO has recently elaborated the UNIDO Bio-fuels Strategy. In
synergy with this effort, ICS-UNIDO has been implementing a technical expertise programme on bio-fuel
production technologies, focusing on the survey and assessment of technologies, especially those for the
next generation of biofuels.

The present publication provides a survey of current technologies for the production of bio-fuels and bio-
based products, and briefly describes various feedstocks and indicates trends of further development
toward bio-refineries and integrated production of a variety of products (biofuels, chemicals, materials,
etc.) in a sustainable way.

The final part is dedicated to the assessment of various technologies and their comparison followed by a
brief description of the concept of decision support tools for technology assessment and preparation of
various bio-feedstocks exploitation scenarios.

The present publication is intended to be used as a technical guideline for initiatives on biofuels in
developing countries.


iii

Enclosed in the Annex is a preliminary survey of selected initiatives and programmes in the biofuels
sector in various parts of the world. This part is being further developed to cover other bio-based
products.

This document has been prepared within the ICS-UNIDO Work Programme 2006-2008 in the Area of
Pure and Applied Chemistry by experts and fellows of ICS-UNIDO in tight cooperation with a group of
experts from the Leipzig Institute for Energy, Germany. Valuable contributions of Sivasamy Arumugam,
Sergei Zinoviev, Paolo Fornasiero, Franziska Mueller-Langer, Martin Kaltschmitt, Alexander Vogel,
Daniela Thraen and Francis Kemausuor are appreciated.

This is a working document and may contain errors; therefore we would appreciate if readers could give
us such indications or suggestions, which will be taken into account.

Stanislav Miertus

Area Chief
Pure & Applied Chemistry, ICS-UNIDO


Trieste, July 2008




iv
Table of Contents

1.0 INTRODUCTION...............................................................................................................1
1.1 Biomass....................................................................................................................................1
1.2 Biomass to energy conversion technologies ............................................................................2
1.3 Biofuels types and generations ................................................................................................3
1.4 Bibliography ............................................................................................................................7

2.0 FIRST GENERATION OF BIOFUELS .............................................................................8
2.1 Biodiesel from vegetable oils...................................................................................................8
Direct use of vegetable oils and of their blends as fuels .............................................................8
2.2 FAME production technology ...............................................................................................11
2.3 Transesterification process technological arrangement .........................................................15
Feedstock issues ........................................................................................................................16
By-products issues .....................................................................................................................18
Effect of free fatty acid and moisture ........................................................................................19
Catalyst type and concentration................................................................................................20
Molar ratio of alcohol to oil and type of alcohol ......................................................................22
Effect of reaction time and temperature....................................................................................23
Mixing intensity .........................................................................................................................24
Effect of using organic co-solvents............................................................................................24
2.4 Product properties and quality ...............................................................................................25
Cetane number...........................................................................................................................26
Flashpoint..................................................................................................................................27
Lubricity ....................................................................................................................................27
Cold flow ...................................................................................................................................27
Fuel stability..............................................................................................................................28
Fuel energy content ...................................................................................................................28
Material compatibility...............................................................................................................29
Biodegradability ........................................................................................................................29
NOx emission.............................................................................................................................29
Other emissions .........................................................................................................................30
2.5 Safety and Precautions...........................................................................................................31
2.6 Biodiesel production via transesterification in supercritical medium....................................32
2.7 Production of biodiesel by pyrolysis and hydroprocessing of oils and fats...........................32
2.8 Market and economic aspects of biodiesel production..........................................................36
Final price .................................................................................................................................36
Production costs ........................................................................................................................36
Feedstock costs..........................................................................................................................36
Tax incentives ............................................................................................................................37
2.9 Bioethanol from sugar and Starch crops................................................................................37
Bioethanol from sugar feedstocks .............................................................................................38
Bioethanol from starch..............................................................................................................38
Bioethanol production processes ..............................................................................................39
Advantages and disadvantages of bioethanol as fuel................................................................40
2.10 Biogas by anaerobic digestion ...............................................................................................41
Biomethane by anaerobic digestion (via biogas) ......................................................................42
The process of anaerobic digestion...........................................................................................42
Digester Technology..................................................................................................................43
Types of Anaerobic Digesters....................................................................................................45
Biogas from wastes....................................................................................................................46


v
Landfill Gas ...............................................................................................................................46
Biogas upgrading to biomethane...............................................................................................48
2.11 Bibliography ..........................................................................................................................51

3.0 SECOND GENERATION OF BIOFUELS AND BIOREFINERIES..............................54
3.1 Pyrolysis and gasification of biomass....................................................................................54
3.2 Synthetic biofuels...................................................................................................................55
Bio-SNG.....................................................................................................................................58
Fischer-Tropsch fuels................................................................................................................60
Biomethanol...............................................................................................................................60
Dimethylether (DME) ................................................................................................................60
Hydro Thermal Upgrading (HTU) diesel ..................................................................................61
3.3 Cellulosic bioethanol .............................................................................................................62
Process description ...................................................................................................................63
Hydrolysis by dilute acid...........................................................................................................64
Hydrolysis by concentrated acid ...............................................................................................65
Enzymatic hydrolysis .................................................................................................................66
Bioethanol production from hydrolysed cellulosic biomass......................................................66
Hybrid and thermochemical processes for ethanol production from cellulosic biomass .........67
3.4 Bio-hydrogen .........................................................................................................................69
Thermo-chemical gasification coupled with water gas shift .....................................................70
Fast pyrolysis followed by reforming of carbohydrate fraction of bio-oil ................................71
Direct solar gasification............................................................................................................72
Biomass derived syn-gas conversion.........................................................................................72
Supercritical conversion of biomass .........................................................................................73
Microbial conversion of biomass ..............................................................................................73
Comparative analysis ................................................................................................................74
Miscellaneous novel gasification processes..............................................................................76
3.5 Biorefineries...........................................................................................................................77
The Concept...............................................................................................................................77
Biorefinery feedstocks................................................................................................................79
Biorefinery types according to feedstock and technologies ......................................................80
Current biorefineries and outlook/development trends.............................................................89
3.6 Bibliography ..........................................................................................................................92

4.0 TECHNOLOGY EVALUATION ASPECTS...................................................................97
4.1 General aspects of future biofuel production options ............................................................97
Lignocellulosic bioethanol ........................................................................................................98
Synthetic biofuels.......................................................................................................................99
Biohydrogen ............................................................................................................................100
Biogas ......................................................................................................................................100
4.2 Technology aspects..............................................................................................................101
4.3 Economic aspects.................................................................................................................101
4.4 Environmental aspects .........................................................................................................104
4.5 Bibliography ........................................................................................................................107

5.0 DECISION SUPPORT TOOLS FOR BIOFUELS AND BIOFUELS TECHNOLOGY
ASSESSMENT............................................................................................................................108
5.1 The concept of decision support tool for assessment of scenarios of biofuel production and
bio-feedstock exploitation ...........................................................................................................109
5.2 BioAs: A Decision Support Tool for Biofuel Assessment and Selection............................115
Criteria ....................................................................................................................................115


vi
Technical information .............................................................................................................116
5.3 Bibliography ........................................................................................................................118

ANNEX BIOFUELS AROUND THE WORLD.........................................................................119
A. BIODIESEL............................................................................................................................119
A.1 Europe...............................................................................................................................119
A. 2 Belgium ...........................................................................................................................121
A. 3 Czech Republic................................................................................................................122
A. 4 Estonia .............................................................................................................................123
A. 5 France ..............................................................................................................................123
A. 6 Finland.............................................................................................................................124
A. 7 Italy..................................................................................................................................125
A. 8 Germany ..........................................................................................................................126
A. 9 Slovakia ...........................................................................................................................127
A. 10 Switzerland ....................................................................................................................128
A. 11 Norway ..........................................................................................................................129
A. 12 Spain ..............................................................................................................................129
A. 13 Austria ...........................................................................................................................130
A. 14 United Kingdom............................................................................................................132
A. 15 China..............................................................................................................................133
A. 16 Singapore.......................................................................................................................134
A. 17 Thailand.........................................................................................................................134
A. 18 Malaysia.........................................................................................................................135
A. 19 India...............................................................................................................................135
A. 20 Israel ..............................................................................................................................137
A. 21 Australia.........................................................................................................................137
A. 22 Brazil .............................................................................................................................139
A. 23 United States..................................................................................................................140
A. 24 Canada ...........................................................................................................................143
A. 25 Costa Rica......................................................................................................................145
B. BIOETHANOL.......................................................................................................................146
B.1 Introduction.......................................................................................................................146
B.2 European Union ................................................................................................................149
B.3 Germany............................................................................................................................151
B.4 The Netherlands................................................................................................................152
B.5 Sweden..............................................................................................................................152
B.6 UK.....................................................................................................................................153
B.7 Italy ...................................................................................................................................153
B.8 France................................................................................................................................154
B.9 Rest of Europe ..................................................................................................................154
B.10 United States ...................................................................................................................155
B.11 Canada ............................................................................................................................160
B.12 Rest of North America....................................................................................................161
B.13 Brazil...............................................................................................................................162
B.14 Argentina ........................................................................................................................163
B.15 Colombia.........................................................................................................................164
B.16 Rest of South and Central America ................................................................................164
B.17 China...............................................................................................................................166
B.18 Thailand ..........................................................................................................................167
B.19 India ................................................................................................................................168
B.20 Philippines ......................................................................................................................169
B.21 Rest of Asia.....................................................................................................................169


vii
B.22 Australia..........................................................................................................................170
B.23 Africa ..............................................................................................................................172
Bibliography: ...........................................................................................................................175
WEB SITES: ...............................................................................................................................176



1

1.0 INTRODUCTION

1.1 Biomass
Biomass refers to material of biological origin excluding material embedded in geological
formations and transformed to fossil. Biomass can directly or indirectly be converted to biofuels
which can be of solid, liquid or gaseous forms. Major sources of biomass for energy purposes are
various types of woody and herbaceous biomass, biomass from fruits and seeds (e.g. energy
crops) as well as biomass mixtures like animal or horticultural by-products, etc. Within these
products solar energy directly or indirectly (in terms of biomass of animal by-products) by the
process of photosynthesis is stored which enables the plants to produce biomass. Biomass is
available in abundance and is cheap and its better utilization is to convert it to energy rich
products using suitable processes.

Biomass has been the most important energy source for humans since the discovery of fire, and
today it is still the main source of energy for almost half of the worlds population. The need to
increase the use of renewable energy is fundamental to make the world energy matrix more
sustainable.

Biomass is a very important renewable energy source. The total use of biomass energy is
inherently difficult to measure, especially because much of it does not involve commercial
transactions. Globally, the primary energy use of biomass in 2000 was about 52 EJ. Of this total,
roughly 45 EJ was consumed as traditional household fuel in developing countries, with some of
this converted to charcoal for urban and industrial uses. This is why biomass is only a small
percentage of primary energy in industrialized countries but is 42 % of primary energy in India
and up to 90% in the worlds poorest countries in Africa and Asia. Modern uses of biomass
comprised the remaining 7 EJ, mostly in the production of electricity and steam in industrialized
and developing countries, such as in the pulp and paper industry, but also in some production of
biofuels, as with ethanol in Brazil.

Combustion to produce thermal energy is the traditional way of using biomass, which is what
humans have been doing since they discovered fire. The positive benefits of wood combustion to
human well-being and longevity were undoubtedly enormous, but there were also costs.
Archaeologists tell that cave walls of our ancestors were coated with residues from the thick


2
smoke that would have filled the air and clogged the lungs of cave dwellers. Smoke-filled
interior spaces are still the norm for the one third of humanity that continues to rely on wood as
its primary energy source, and the particulates and noxious fumes from cooking with open fires
and inefficient and poorly ventilated stoves fuelled by wood and crop residues have substantial
health impacts. The transition in developing regions of the world from traditional technologies
using biomass to more efficient technologies using fossil fuels (propane, butane) results in a
dramatic improvement in indoor air quality and increased life expectancy.

1.2 Biomass to energy conversion technologies
Advanced technologies are now under development to convert biomass into various forms of
secondary energy including electricity, gaseous and liquid biofuels, and even hydrogen. The
purpose of biomass conversion is to provide fuels with clearly defined fuel characteristics that
meet given fuel quality standards. To ensure that these fuel quality standards are met and these
biomass based fuels can be used with a high efficiency in conversion devices (like engines,
turbines) upgrading is needed. In general, there are various options to produce alternative
transportation fuels based on biomass. Biogenous energy sources can be converted by means of
highly different supply chains into gaseous and liquid biofuels that can be used for transportation
purposes. This treatment leads to an upgrading of energy sources in terms of one or more
properties named as follows:
Energy density,
Handling,
Storage and transport,
Environmental compatibility,
Utilising of by-products and residues.

There are three principal pathways for converting biomass into biofuels (Scheme 1.1):
(i) the thermo-chemical pathway,
(ii) the physical-chemical conversion pathway,
(iii) the bio-chemical conversion pathway.
These pathways lead to production of biofuels in the form of solids (mainly charcoal), liquids
(mainly biodiesel and alcohols), or gases (mainly mixtures with methane or carbon
monoxide), which can be used for a wide range of applications, including transport and high-
temperature industrial processes.



3

Scheme 1.1: Basic pathways for the provision of final energy derived from biomass


1.3 Biofuels types and generations
Biofuel is any fuel that is derived from biomass, recently living organisms or their metabolic by-
products, such as manure from cows. It is a renewable energy source, unlike other natural
resources such as petroleum, coal and nuclear fuels. Biofuels can be grouped in generations
(thus: first, second and third generation) according to the type of technology they rely on and the
biomass feedstocks they convert into fuel.

First generation biofuels are biofuels which are produced from food crops (sugar or oil crops)
and other food based feedstock (e.g. food waste). These biofuels are on the market in
considerable amounts today and their production technologies are well established. The most
important biofuels of the first generation are bioethanol, biodiesel, and biogas. For these types of
fuels, only easily extractible parts of plants are used, such as starch-rich corn kernels, grains or
the sugar in canes and oilseeds. Remaining by-products, such as press cake from vegetable oil
production, glycerine from biodiesel production or Dried Distillers Grains with Solubles
(DDGS) from bioethanol based on starch, are typically used for fodder or chemical purposes.
Integrated concepts that are in R&D stage include the energetic use of by-products (e.g. for
process energy provision). First generation biofuels are being produced in commercial quantities
today, with almost 50 billion litres (approx. 39.5 million t) of bioethanol and 5.4 million t of


4
biodiesel produced worldwide in 2006. First generation biofuels offer the greatest short-term
potentials of biofuels today. Although they differ in properties, technical requirements,
economical aspects and potential, they can contribute to guarantee long-term mobility.

Second generation biofuels are different from the first generation types in the sense that they
can be produced from a wider range of feedstocks. The goal of second generation biofuel
processes is to extend the amount of biofuel that can be produced sustainably by using biomass
comprised of the residual non-food parts of current crops, such as stems, leaves and husks that
are left behind once the food crop has been extracted, as well as other crops that are not used for
food purposes, such as switch grass and cereals that bear little grain, and also industry waste
such as wood chips, skins and pulp from fruit pressing etc. In order to breakdown this biomass,
two main conversion pathways come into consideration:
1) hydrolysis (which can be done via chemical and bio-chemical pathways) of
lignocellulose into sugars, which can then be fermented into alcohol this technology is
best known as cellulosic bioethanol and is still in development;
2) thermo-chemical processes (use of high temperatures to pyrolyse or gasify biomass) of
lignocelluloses to a raw gas or oil. The resulting gas is then treated and conditioned into
synthesis gas (syngas), consisting mainly of carbon monoxide and hydrogen. This gas
can further be processed into different types of liquid and gaseous fuels via different fuel
syntheses. Fuels from this route are then called synthetic biofuels.
Most promising liquid synthetic biofuels, also called biomass-to-liquids (BtL), are biomethanol
and Fischer-Tropsch fuels. Gaseous synthetic biofuels are e.g. dimethylether (DME) and Bio-
SNG, which is also a form of biomethane and can be similarly used as natural gas substitute like
biogas. Alternatively, the cleaned and conditioned product gas can be converted into hydrogen.
Bio-oil obtained from biomass via pyrolysis or hydrothermal treatment can also be converted
into high quality liquid fuels by deoxygenation, e.g. Hydro Thermal Upgrading (HTU diesel).

Third generation biofuels rely on biotechnological interventions in the feedstocks themselves.
Plants are engineered in such a way that the structural building blocks of their cells (lignin,
cellulose, hemicellulose), can be managed according to a specific task they are required to
perform. For example, plant scientists are working on developing trees that grow normally, but
can be triggered to change the strength of the cell walls so that breaking them down to release
sugars is easier. In third generation biofuels, a synergy between this kind of interventions and
processing steps is then created: plants with special properties are broken down by functionally
engineered enzymes. Notably, this latter generation of biofuels is only gradually being explored.


5
Current third generation biofuels research is also placing much emphasis on algae. A viable
biofuels production scheme from algae is seen as one of the best ways to produce enough
alternative fuels to replace the use of fossil fuels. The per unit area yield of oil from algae is
estimated to be from between 5,000 to 20,000 gallons per acre, per year. Research into algae for
the mass-production of oil is mainly focused on microalgae: single-celled aquatic plants that
have many potential uses including irrigation, waste-water treatment, and the production of food,
fuels, and commodity chemicals. Microalgae have fast growth rate and high oil contents and
regarded as a very good biofuel feedstock option for the future. Current research is geared
towards finding an algal strain with a high lipid content and fast growth rate that is not too
difficult to harvest, and a cost-effective cultivation system that is best suited to that strain.

The principal biofuels of the first and second generations, including respective feedstocks and
production technologies are presented in Table 1.1.

Table 1.1: Overview of biofuels of the first and second generation and their related feedstock
and conversion processes
Generic name Chemical composition Feedstocks Technology
Vegetable oil

Straight Vegetable Oil
(SVO) triglycerides
of fatty acids
Oil crops (e.g. rape,
palm, soya, jatropha,
etc.)
Cold/hot pressing,
solvent extraction, and
purification

Biodiesel

Fatty acid methyl esters
(FAME)

SVO, waste oil, and
animal fats
Ibid. and
transesterification
Bio-ethanol

Ethanol Sugar crops (sugar
beet, sugarcane) and
starch crops (corn,
wheat)

Extraction, hydrolysis (in
the case of starch),
fermentation of sugars
1
s
t

g
e
n
e
r
a
t
i
o
n

Advanced
biodiesel

Hydrocarbons Cf. biodiesel Hydrocracking

2
n
d

g
e
n
e
r
a
t
i
o
Bio-syngas CO, H
2
, with impurities
of CO
2
, methane, etc.

Any biomass Gasification or pyrolysis


6
Bio-SNG

Synthetic (substitute)
natural gas methane

Bio-syngas Upgrading, catalytic
synthesis

FT diesel Hydrocarbons

Ibid. Ibid.
Bio-methanol Methanol

Ibid. Ibid.

Bio-DME Dimethylether

Bio-methanol Catalytic synthesis
Bio-hydrogen

Hydrogen Ibid. Water gas shift reaction
(WGSR), purification

Bio-MTBE

Methyl-tert-butyl ether Ibid. and isobutene Catalytic synthesis

Bio-oil Oxygenated
hydrocarbons and tar
Any biomass Flash pyrolysis or
hydrothermal upgrading

HTU diesel Hydrocarbons

Wet biomass Hydrothermal upgrading
and hydroprocessing

Cellulosic bio-
ethanol

Ethanol Cellulose and
hemicellulose (woody
or herbaceous crops)

Fractionation, hydrolysis,
fermentation

Bio-methane
(biogas)

Mainly methane

Wet biomass Anaerobic digestion,
conditioning/cleaning

Bio-ETBE Ethyl-tert-butyl ether Bio-ethanol or
cellulosic bioethanol,
isobutene

Catalytic synthesis
M
i
x
e
d

Bio-butanol

Butanol Cf. bio-ethanol or
cellulosic bio-ethanol
Cf. bio-ethanol or
cellulosic bio-ethanol



7
1.4 Bibliography

Boyle G. 2003, Renewable Energy, Oxford, Oxford University Press.
Dincer I. 2000, Renewable and Sustainable Energy Reviews, 4, 157.
Jaccard M. 2006, Sustainable Fossil Fuels - The Unusual Suspect in the Quest for Clean and
Enduring Energy, Cambridge University Press.
Kaltschmitt M. 2001, Energie aus Biomasse Grundlagen, Techniken und Verfahren. Springer-
Verlag, Berlin, Heidelberg.




8
2.0 FIRST GENERATION OF BIOFUELS

2.1 Biodiesel from vegetable oils
The first generation biodiesel is usually referred to as a mixture of fatty acid methyl esters
(FAME) produced from vegetable oils and animal fats via transesterification reaction. Several
production methodologies are available, which employ the use of homogeneous, heterogeneous,
or bio-catalysts. The mostly used commercial technology for biodiesel production is the
transesterification reaction of the triglyceride of the fatty acid with methanol under the basic
conditions (e.g. sodium hydroxide as the catalyst) to yield the methyl ester of the fatty acid
(biodiesel).

Biodiesel is physically similar to petroleum-based diesel fuel and can be blended with diesel fuel
in any proportion. The most common blend is a mixture consisting of 20% biodiesel and 80%
petroleum diesel, called B20. The motive for blending the fuels is to gain some of the advantages
of biodiesel while avoiding higher costs. The commonly known biodiesel blends are outlined
below.
B100 (100% biodiesel) offers the most environmental benefits. Use of B100 may however
require engine or fuel system component modification and can cause operating problems,
especially in cold weather.
B20 (20% biodiesel) offers about one fifth of the environmental benefits of B100, but can be
more broadly applied to existing engines with little or no modification.
B2 (2% biodiesel) offers little environmental or petroleum dependence benefit and could be
potentially used as an environmental marketing tool.

Feedstocks for first generation of biodiesel basically include vegetable oil obtained from oil
(energy) crops, as well as recycled oil, animal fats, etc. Different feedstocks may require
different conditions of treatment and different pretreatment technology to be adopted and the
cost and complexity of the process and the quality of the product can vary.

Direct use of vegetable oils and of their blends as fuels
Vegetable oils may be used directly as fuel in diesel engines. However, there are many problems
associated with using them directly in diesel engine (especially in direct injection engine). Some
of the problems that may be encountered include coking and trumpet formation on the injectors,
carbon deposits, oil ring sticking, thickening or gelling of the lubricating oil, and lubricating


9
problems. Other disadvantages have to do with the high viscosity of vegetable oils (about 1117
times higher than diesel fuel) and lower volatilities. These problems can be avoided by
modifying the engine.

The fuel properties of vegetable oils as listed in Table 2.1 indicate that the kinematic viscosity of
vegetable oils varies in the range of 3040 mm
2
/s at 38 C. The high viscosity of these oils is due
to their large molecular mass in the range of 600900, which is about 20 times higher than that
of diesel fuel. The flash point of vegetable oils is very high (above 200 C). The volumetric
heating values are in the range of 3940 MJ/kg, as compared to diesel fuels of about 45 MJ/kg.
The presence of chemically bound oxygen in vegetable oils lowers their heating values by about
10%. The cetane numbers (described in subsequent sections) are in the range of 3240. The
direct use of vegetable oils is generally considered to be unsatisfactory and impractical for both
direct diesel engines. The probable reasons for the problems and the potential solutions are
shown in Table 2.2.

Table 2.1: Properties of vegetable oils (Barnwal, 2005)
Vegetable oil Kinematic
viscosity at 38
o
C (mm
2
/s)
Cetane
no. (
o
C)
Heating
value
(MJ/kg)
Cloud
point (
o
C)
Pour point
(
o
C)
Flash
point (
o
C)
Density
(kg/l)
Corn 34.9 37.6 39.5 -1.1 -40.0 277 0.9095
Cottonseed 33.5 41.8 39.5 1.7 -15.0 234 0.9148
Crambe 53.6 44.6 40.5 10.0 -12.2 274 0.9048
Linseed 27.2 34.6 39.3 1.7 -15.0 241 0.9236
Peanut 39.6 41.8 39.8 12.8 -6.7 271 0.9026
Rapeseed 37.0 37.6 39.7 -3.9 -31.7 246 0.9115
Safflower 31.3 41.3 39.5 18.3 -6.7 260 0.9144
Sesame 35.5 40.2 39.3 -3.9 -9.4 260 0.9133
Soya bean 32.6 37.9 39.6 -3.9 -12.2 254 0.9138
Sunflower 33.9 37.1 39.6 7.2 -15.0 274 0.9161
Palm 39.6 42.0 - 31.0 - 267 0.9180
Babassu 30.0 38.0 - 20.0 - 150 0.9460
Diesel 3.06 50 43.8 - -16 76 0.855







10
Table 2.2: Known problems, probable cause and potential solutions for using straight vegetable
oil in diesels (Harwood, 1984)

Problem

Probable cause

Potential solution

Short-term
1. Cold weather starting High viscosity, low cetane, and low
flash point of vegetable oils

Preheat fuel prior to injection.
Chemically alter fuel to an ester.

2. Plugging and gumming of
filters, lines and injectors
Natural gums (phosphatides) in
vegetable oil. Other ash
Partially refine the oil to remove
gums. Filter oil to 4-microns.

3. Engine knocking

Very low cetane of some oils.
Improper injection timing

Adjust injection timing. Use higher
compression engines. Same as (1).

Long term
4. Coking of injector nozzles

High viscosity of vegetable oil,
incomplete combustion of fuel. Poor
combustion at part load with
vegetable oils

Heat fuel prior to injection. Switch
engine to diesel fuel when operating
at part load. Chemically alter the
vegetable oil to an ester.

5. Carbon deposits on piston and
head of engine

High viscosity of vegetable oil,
incomplete combustion of fuel. Poor
combustion at part load with
vegetable oils

Heat fuel prior to injection. Switch
engine to diesel fuel when operating
at part load. Chemically alter the
vegetable oil to an ester.
6. Excessive engine wear

High viscosity of vegetable oil,
incomplete combustion of fuel. Poor
combustion at part load with
vegetable oils. Possibly free fatty
acids in vegetable oil. Dilution of
engine lubricating oil due to blow-by
of vegetable oil

Heat fuel prior to injection. Switch
engine to diesel fuel when operating
at part load. Chemically alter the
vegetable oil to an ester. Increase
motor oil changes. Motor oil
additives to inhibit oxidation.

7. Failure of engine lubricating
oil due to polymerization

Collection of polyunsaturated
vegetable oil blow-by in crankcase
to the point where polymerization
occurs
Heat fuel prior to injection. Switch
engine to diesel fuel when operating
at part load. Chemically alter the
vegetable oil to an ester. Increase
motor oil changes. Motor oil
additives to inhibit oxidation.



11
To solve the problem of high viscosity of vegetable oils, several techniques have been
implemented. These include: blending with fossil diesel, transesterification, micro-emulsification
and thermal cracking or pyrolysis of vegetable oils. Among all these techniques the
transesterification of oils by homogeneous catalysts has seen wider application due to its easy,
cost effective process technology and also because the final products meets physico-chemical
parameters of the fuel quality

This chapter is dedicated to the issues related to the production technologies for biodiesel of the
first generation, including biodiesel produced by vegetable oil hydrocracking. Production of
other types of biodiesel, e.g. second generation synthetic biodiesel from bio-SNG (synthetic
natural gas), will be considered in the next chapters.

2.2 FAME production technology
Generally speaking, there are three basic ways for the production of methyl esters from oils and
fats:
Base catalysed transesterification of the oil (triglycerides) with methanol;
Direct acid catalyzed esterification of the free fatty acids (FFA) with methanol;
Conversion of the oil to FFA followed by esterification as described above.

The majority of the methyl esters are produced using the base catalyzed reaction because it is the
most economic for several reasons:
low temperature and pressure;
high yields and short reaction times;
direct conversion process;
simple in operation and environmentally benign.

Transesterification can be defined as the process of reacting a triglyceride (oil) with an alcohol
(e.g., methanol or ethanol) in the presence of a catalyst, such as sodium hydroxide or potassium
hydroxide, to chemically break the molecule of the oil into methyl or ethyl esters. Glycerol, also
known as glycerine, is the by-product of this reaction. The process is similar to hydrolysis,
except that alcohol is used instead of water. The general scheme of the transesterification
reaction is shown in Scheme 2.1.



12
OCOR
3
H
2
C
CH
H
2
C
OCOR
1
OCOR
2
+3CH
3
OH
OH
H
2
C
CH
H
2
C
OH
OH
+R
1
COOCH
3
+R
2
COOCH
3
+R
3
COOCH
3
catalyst
triglyceride methanol glycerol fatty acid methyl esters


Scheme 2.1: Transesterification reaction

Transesterification consists of a number of consecutive, reversible reactions. Diglycerides and
monoglycerides are the intermediates in this process. The triglyceride is converted stepwise to
diglyceride, monoglyceride and finally glycerol. The reactions are reversible, although the
equilibrium lies towards the production of fatty acid esters and glycerol. A little excess of
alcohol is used to shift the equilibrium towards the formation of esters. In the presence of excess
alcohol, the foreword reaction is pseudo-first order and the reverse reaction is found to be second
order. Transesterification is faster when catalyzed by an alkali.

The mechanism of alkali-catalyzed transesterification as described in more detail in Scheme 2.2
is formulated as three steps. The first step involves the attack of the alkoxide ion to the carbonyl
carbon of the triglyceride molecule, which results in the formation of a tetrahedral intermediate.
The reaction of this intermediate with an alcohol produces the alkoxide ion in the second step. In
the last step the rearrangement of the tetrahedral intermediate gives rise to an ester and a
diglyceride (Schwab, 1987).



13
ROH +B RO
-
+BH
+
CH
2
CH
H
2
C
R
2
OOC
R
1
OOC
OCR
3
O
+
-
OR
CH
2
CH
H
2
C
R
2
OOC
R
1
OOC
C
O
-
OR
R
3
O
CH
2
CH
H
2
C
R
2
OOC
R
1
OOC
C
O
-
OR
R
3
O
CH
2
CH
H
2
C
R
2
OOC
R
1
OOC
O
-
+ ROOCR
3
CH
2
CH
H
2
C
R
2
OOC
R
1
OOC
O
-
+ BH
+
CH
2
CH
H
2
C
R
2
OOC
R
1
OOC
O
-
+ B


Scheme 2.2: Mechanism of base catalyzed transesterification

Transesterification can also be catalyzed by Brownsted acids, preferably by sulfonic and sulfuric
acids. These catalysts give very high yields in alkyl esters but these reactions are slow, requiring
typically temperatures above 100 C and more than 3 h to complete the conversion (Schuchardt,
1998). The mechanism of acid catalyzed transesterification of vegetable oil (for a
monoglyceride) is shown in Scheme 2.3.



14
CH
2
CH
H
2
C
R
2
OOC
R
1
OOC
OCR
3
+ H
+
CH
2
CH
H
2
C
R
2
OOC
R
1
OOC
O
O
CH
2
CH
H
2
C
R
2
OOC
R
1
OOC
OCR
3
OH
+
CH
2
CH
H
2
C
R
2
OOC
R
1
OOC
OCR
3
OH
2
+
R
OH
+
C
OH
R
3
OH
+
R
CH
2
CH
H
2
C
R
2
OOC
R
1
OOC
O C
OH
R
3
OH
+
R
CH
2
CH
H
2
C
R
2
OOC
R
1
OOC
H
+
O
C
OH
R
3
O R
CH
2
CH
H
2
C
R
2
OOC
R
1
OOC
H
+
O
C
OH
R
3
O R
CH
2
CH
CH
2
R
2
OOC
R
1
OOC
ROC
OH
+
R
3
HO
+
ROC
OH
+
R
3
ROCOR
3
+H
+


Scheme 2.3: Mechanism of acid catalyzed transesterification

The protonation of carbonyl group of the ester leads to the carbocation, which after a
nucleophilic attack of the alcohol produces a tetrahedral intermediate. This intermediate
eliminates glycerol to form a new ester and to regenerate the catalyst. This mechanism can be
extended to di- and tri-glycerides.


15

The process of transesterification is affected by various factors depending upon the reaction
condition used. The effects of these factors are described below. The most relevant variables are:
the reaction temperature, the ratio of alcohol to vegetable oil, the amount and the type of catalyst,
the mixing intensity and the raw oils used.

2.3 Transesterification process technological arrangement
In scheme 2.4 the basic transesterification process flows are presented. Critical quality
parameters of the process are:
Complete reaction.
Removal of glycerol.
Removal of catalyst.
Removal of alcohol.
Absence of free fatty acids.
Low sulphur content.




Scheme 2.4: Flowchart of the esterification process

Biodiesel standards ensure that these factors in the fuel production process are satisfied. Basic
industrial tests to determine whether the products conform to the standards typically include gas
Alcohol, e.g.
methanol
Catalyst,
e.g. KOH
Oil and Fats
(Multi-Feedstock)
Pretreatment
(Optional)
Transesterification
Purification
Biodiesel
Glycerin
Fatty
Acids


16
chromatography, a test that verifies only the more important of the variables above. More
complete tests are more expensive. Fuel meeting the quality standards is very non-toxic, with a
toxicity rating of greater than 50 ml/kg.

As illustrated in the Scheme 2.3, a typical supply chain includes the production of vegetable oils
or provision of other feedstocks. The extracted and purified oils/fats undergo the conversion to
biodiesel in the production plant. The product, after purification steps is distributed to the end
user.


Scheme 2.3: Feedstock supply and biodiesel production and distribution

Feedstock issues
Any fatty acid (from vegetable oils and animal fats) can be used to prepare biodiesel, however a
sustainable choice must be made. Choosing the right biodiesel feedstock for a region is one of
the very important steps in the production of biodiesel. The right feedstock is one that adds
economic, environmental and perhaps value to the region where the fuel will be produced. Many
issues should be considered such as existing infrastructure, natural resources, local economy,
performance in vehicles, food/fuel considerations, water use, soil quality, and soil carbon
sequestration. The kind and quality of feedstock is a very important factor in a transesterification
plant as it affects corresponding material and energy flows, which are not only indicators of
technical efficiency, but also affect the economic efficiency of biodiesel production. The most
relevant feedstock parameters and their relevance for the production process and biodiesel use
are in presented in Table 2.3.



17
Table 2.3: Feedstock parameters and their relevance for biodiesel production and use (IEE,
2006)
Parameter Characterization Relevance for the biodiesel
production and use
Free fatty acids
(FFA)
significant measure of feedstock
quality, indicator for the level of
hydrolysis; FFA of native unrefined
oils and fats can be above 20.0,
refined oils/fats have FFA up to 1.0

influence degree of required
processing (e.g. catalyst demand)
and biodiesel quality (primarily cold
flow properties)
Total
contamination
proportion of unresolved impurities
(particles) in the oils/fats; mainly
affected by the oil production
procedure
relevant to glycerine quality and
unwanted glycerine caking within
the process; High total
contaminations lead to clogging of
the fuel filters and increase the
danger of damage to the injection
pump and to injection nozzles as
well as of deposits in the
combustion chamber.

Water content mainly affected by the moisture of
the seeds and refined oils/fats; with
all oils/fats the water content can
rise through storage and transport
at high temperatures water can
hydrolyze the triglycerides to
diglycerides and form free fatty
acids; relevant for disturbing the
transesterification by catalysts loss
and unwanted soap production
Cinematic
viscosity
physical-mechanical characteristic,
depending on specific melting point
influenced by the temperature, fatty
acid profile and oil-aging degrees,
whereas the kind of the oil
production procedure does not affect
viscosity

Cold flow
properties
strongly affected by temperature;
saturated fatty compounds with a
cloud point (CP) and the cold filter
plugging point (CFPP) or pour point


18
significantly higher melting points
than unsaturated fatty compounds
(PP) for fuels not suitable to
describe cold flow properties of
oils/fats, since the transition of the
liquid to the solid phase can not be
definite.

Iodine number indicator for double bindings in the
molecular structure of oils/fats. the
higher the iodine value, the more
unsaturated acids are present in the
oils/fats
high iodine number in oils/fats for
less age resisting than oils/fats with
high degree of saturation; informs
about the tendency of deposits in the
combustion chamber and at
injection nozzles.

Phosphorus
content
present in vegetable oils in the form
of phospholipids; depends on
refining grad of oils/fats (influenced
by oil production process)
decreasing oxidation stability with
rising portion of phospholipids; high
amounts of phospholipids leading to
disturbances with technical
processes (e.g. blockages of filters
and injection nozzles); avoidance of
phosphorus compounds in waste
water

Oxidation stability value, which describes the aging
condition and the shelf-life of
oils/fats
oxidation and polymerization
procedures during fuel storage,
which can lead to formation of
insoluble compounds and thus cause
e.g. filter blockage



By-products issues
An important aspect of by-products is that related with glycerol, the principal by-product of the
transesterification process. It occurs in vegetable oils at a level of approximately 10% by weight.
Crude glycerol possesses very low value because of the impurities. However, as the demand and
production of biodiesel grows, the quantity of crude glycerol generated will be considerable, and


19
its utilization will become an urgent topic. Refining the crude glycerol will depend on the
economy of production scale and/or the availability of a glycerol purification facility. It is
generally treated and refined through filtration, chemical additions and fractional vacuum
distillation to yield various commercial grades. Small to moderate scale producers who cannot
justify the high cost of purification find crude glycerol utilization or disposal to be a problem.

Larger scale biodiesel producers refine their crude glycerol and move it through markets in the
food, pharmaceutical and cosmetic industries. Moreover glycerol can be burnt as a fuel. Another
alternative is to etherify glycerol with either alcohols (e.g. methanol or ethanol) or alkenes (e.g.
isobutene) and produce branched oxygen-containing components, which could have properties
for use in fuel or solvents. An advantage of using glycerol in fuel is that, as a biocomponent, it
could be included in the renewables category (Karinen, 2006).

Effect of free fatty acid and moisture
The free fatty acid and moisture content are key parameters for determining the viability of the
vegetable oil transesterification process. To carry the base catalyzed reaction to completion; a
free fatty acid (FFA) value lower than 3% is needed. For maximum yield, the alcohol should be
free of moisture, and the Free fatty acid (FFA) content of the oil should be <0.5. Infact most
commercial plants require feedstock with FFA less than 0.1%. The higher the acidity of the oil,
the smaller the conversion efficiency. Both excess as well as insufficient amount of catalyst may
cause soap formation (Dorado, 2002).

The transesterification of beef tallow catalyzed by NaOH in the presence of free fatty acids and
water was studied by Ma et al. (1998). Without adding FFA and water, the apparent yield of beef
tallow methyl esters (BTME) was highest. When 0.6% of FFA was added, the apparent yield of
BTME reached the lowest, less than 5%, with any level of water added. The products were solid
at room temperature, similar to the original beef tallow. When 0.9% of water was added, without
addition of FFA, the apparent yield was about 17%. If the low qualities of beef tallow or
vegetable oil with high FFA are used to make biodiesel fuel, they must be refined by
saponification using NaOH solution to remove free fatty acids. Conversely, the acid catalyzed
process can also be used for esterification of these free fatty acids.

The starting materials used for base catalyzed alcoholysis should meet certain specifications. The
triglycerides should have lower acid value and all material should be substantially anhydrous.
The addition of more sodium hydroxide catalyst compensates for higher acidity, but the resulting


20
soap causes an increase in viscosity or formation of gels that interferes in the reaction as well as
with separation of glycerol (Freedman, 1984). When the reaction conditions do not meet the
above requirements, ester yields are significantly reduced. The methoxide and hydroxide of
sodium or potassium should be maintained in anhydrous state. Prolonged contact with air will
diminish the effectiveness of these catalysts through interaction with moisture and carbon
dioxide.

Most of the biodiesel produced currently is made from edible oils by using methanol and alkaline
catalyst. However, there are large amounts of low cost oils and fats that could be converted to
biodiesel. The problems with processing these low cost oils and fats are that they often contain
large amounts of free fatty acids that cannot be converted to biodiesel using alkaline catalyst.
Therefore, two-step esterification process is required for these feed stocks. Initially the FFA of
these can be converted to fatty acid methyl esters by an acid catalyzed pretreatment and in the
second step transesterification is completed by using alkaline catalyst to complete the reaction.
In one study, initial process development was performed with synthetic mixture containing 20
and 40% free fatty acid prepared by using palmitic acid. Process parameters such as molar ratio
of alcohol to oil, type of alcohol, amount of acid catalyst, reaction time and free fatty acid levels
were investigated to determine the best strategy for converting the free fatty acids to usable
esters. The work showed that the acid level of the high free fatty acids feed stocks could be
reduced to less than 1% with a two step pretreatment reaction. The reaction mixture was allowed
to settle between steps so that the water containing phase could be removed. The two-step
pretreatment reaction was demonstrated with actual feedstocks, including yellow grease with
12% free fatty acid and brown grease with 33% free fatty acids. After reducing the acid levels of
these feedstocks to less than 1%, the transesterification reaction was completed with an alkaline
catalyst to produce fuel grade biodiesel (Canakci, 2001). An investigation into the negative
influence of base catalyzed transesterification of triglycerides containing substantial amount of
free fatty acid was done by Turck et al. (2002).

Catalyst type and concentration
Catalysts used for the transesterification of triglycerides are classified as alkali, acid, enzyme or
heterogeneous catalysts. Among these, alkali catalysts like sodium hydroxide, sodium
methoxide, potassium hydroxide, potassium methoxide are more effective (Ma, 1999a). If the oil
has high free fatty acid content and more water, acid catalyzed transesterification is suitable. The
acids could be sulfuric acid, phosphoric acid, hydrochloric acid or organic sulfonic acid.



21
Methanolysis of beef tallow was studied with catalysts NaOH and NaOMe. Comparing the two
catalysts, NaOH was significantly better than NaOMe (Ma, 1998). The catalysts NaOH and
NaOMe reached their maximum activity at 0.3% and 0.5% w/w of the beef tallow, respectively.
Sodium methoxide causes formation of several by-products mainly sodium salts, which are to be
treated as waste. In addition, high quality oil is required with this catalyst (Ahn, 1995). This was
different from the previous reports (Freedman, 1984) in which ester conversion at the 6:1 molar
ratio of alcohol/oil for 1% NaOH and 0.5% NaOMe were almost the same after 60 min. Part of
the difference may be attributed to the differences in the reaction system used.

Sodium hydroxide and potassium hydroxide were used as a catalyst in the process of alkaline
methanolysis, both in concentrations from 0.4 to 2% w/w of oil. Methanolysis of soybean oil
with the 1% potassium hydroxide catalyst gave the best yields and viscosities of the esters
(Tomasevic, 2003).

Attempts have been made to use basic alkaline-earth metal compounds in the transesterification
of rapeseed oil for production of fatty acid methyl esters. The reaction proceeds if methoxide
ions are present in the reaction medium. The alkaline-earth metal hydroxides, alkoxides and
oxides catalyzed reaction proceeds slowly as the reaction mixture constitutes a three-phase
system oilmethanolcatalyst, which for diffusion reason inhibits the reaction (Gryglewicz,
1999). The catalytic activity of magnesium oxide, calcium hydroxide, calcium oxide, calcium
methoxide, barium hydroxide, and for comparison, sodium hydroxide during the
transesterification of rapeseed oil was investigated. Sodium hydroxide exhibited the highest
catalytic activity in this process. The degree to which the substrates were reacted reached 85%
after 30 min of the process and 95% after 1.5 h, which represented a close value to the
equilibrium. Barium hydroxide was slightly less active with a conversion of 75% after 30 min.
Calcium methoxide was medially active. The degree to which the substrates were reacted was
55% after 30 min., 80% after 1 h and state of reaction equilibrium (93%) was reached after 2.5 h.
The rate of reaction was slowest when catalyzed by CaO. Magnesium oxide and calcium
hydroxide showed no catalytic activity in rapeseed oil methanolysis. Acid catalyzed
transesterification was studied with waste vegetable oil. The reaction was conducted at four
different catalyst concentrations, 0.5, 1.0, 1.5 and 2.25 M HCl in the presence of 100% excess
alcohol and the result was compared with 2.25 M H
2
SO
4
and the decrease in viscosity was
observed. H
2
SO
4
has superior catalytic activity in the range of 1.52.25 M concentration
(Mohamad, 2002).



22
Although chemical transesterification using an alkaline catalysis process gives high conversion
levels of triglycerides to their corresponding methyl esters in short reaction times, the reaction
has several drawbacks: it is energy intensive, recovery of glycerol is difficult, the acidic or
alkaline catalyst has to be removed from the product, alkaline waste water require treatment, and
free fatty acid and water interfere with the reaction.

Enzymatic catalysts like lipases are able to effectively catalyze the transesterification of
triglycerides in either aqueous or non-aqueous systems, which can overcome the problems
mentioned above (Fuduka, 2001). In particular, the by-products, glycerol can be easily removed
without any complex process, and also free fatty acids contained in waste oils and fats can be
completely converted to alkyl esters. On the other hand, in general the production cost of a lipase
catalyst is significantly greater than that of an alkaline one. Table 2.4 summarizes the differences
between the various technologies used to produce biodiesel.

Table 2.4: Comparison of the different technologies to produce biodiesel
Variable Alkali catalysis Lipase catalysis Supercritical
alcohol
Acid catalysis
Reaction T(C) 60-70 30-40 239-385 55-80
FFA in raw
material
Saponified
products
Methyl esters Esters Esters
H
2
O in raw
materials
Interference with
reaction
No influence Interference with
reaction
Yields of
methyl esters
Normal Higher Good Normal
Recovery of
glycerol
Difficult Easy Difficult
Purification of
methyl esters
Repeated washing None Repeated washing
Production cost of
catalyst
Cheap Relatively
expensive
Medium Cheap


Molar ratio of alcohol to oil and type of alcohol
One of the most important variables affecting the yield of ester is the molar ratio of alcohol to
triglyceride. The stoichiometric ratio for transesterification requires three moles of alcohol and


23
one mole of triglyceride to yield three moles of fatty acid alkyl esters and one mole of glycerol.
However, transesterification is an equilibrium reaction in which a large excess of alcohol is
required to drive the reaction to the right. For maximum conversion to the ester, a molar ratio of
6:1 should be used. The molar ratio has no effect on acid, peroxide, saponification and iodine
value of methyl esters (Tomasevic, 2003). However, the high molar ratio of alcohol to vegetable
oil interferes with the separation of glycerine because there is an increase in solubility. When
glycerine remains in solution, it helps drive the equilibrium to back to the left, lowering the yield
of esters.

The base catalyzed formation of ethyl ester is difficult compared to the formation of methyl
esters. Specifically the formation of stable emulsion during ethanolysis is a problem. Methanol
and ethanol are not miscible with triglycerides at ambient temperature, and the reaction mixtures
are usually mechanically stirred to enhance mass transfer. During the course of reaction,
emulsions usually form. With methanol these emulsions quickly and easily break down to form a
lower glycerol rich layer and upper methyl ester rich layer. Instead with ethanol these emulsions
are more stable and severely complicate the separation and purification of esters (Zhou, 2003).
The emulsions are caused in part by formation of the intermediates monoglycerides and
diglycerides, which have both polar hydroxyl groups and non-polar hydrocarbon chains. These
intermediates are strongsurface active agents. In the process of alcoholysis, the catalyst, either
sodium hydroxide or potassium hydroxide is dissolved in polar alcohol phase, in which
triglycerides must transfer in order to react. The reaction is initially mass-transfer controlled and
does not conform to expected homogeneous kinetics. When the concentrations of these
intermediates reach a critical level, emulsions form. The larger non-polar group in ethanol,
relative to methanol, is assumed to be the critical factor in stabilizing the emulsions. However,
the concentrations of mono- and di-glycerides are very low, and then the emulsions become
unstable. This emphasizes the necessity for the reaction to be as complete as possible, thereby
reducing the concentrations of mono- and di-glycerides.

Effect of reaction time and temperature
The conversion rate increases with reaction time. Freedman et al. (1984) transesterified peanut,
cotton-seed, sunflower and soybean oil under the condition of methanoloil molar ratio 6:1,
0.5% sodium methoxide catalyst and 60 C. An approximate yield of 80% was observed after 1
min for soybean and sunflower oils. After 1 h, the conversion was almost the same for all four
oils (9398%). Ma et al. (1999b) studied the effect of reaction time on transesterification of beef
tallow with methanol. The reaction was very slow during the first minute due to mixing and


24
dispersion of methanol into beef tallow. From 1 to 5 min, the reaction proceeded very fast. The
production of beef tallow methyl esters reached the maximum value at about 15 min.

Transesterification can occur at different temperatures, depending on the oil used. The
transesterification of refined oil with methanol (6:1) and 1% NaOH, was studied with three
different temperatures (Ma, 1999a). After 0.1 h, ester yields were 94, 87 and 64% for 60, 45 and
32 C, respectively. After 1 h, ester formation was identical for 60 and 45 C runs and only
slightly lower for the 32 C run. Temperature clearly influenced the reaction rate and yield of
esters (Ma, 1999a).

Mixing intensity
Mixing is very important in the transesterification reaction, as oils or fats are immiscible with
sodium hydroxidemethanol solution. Once the two phases are mixed and the reaction is started,
stirring is no longer needed. Initially the effect of mixing on transesterification of beef tallow
was studied by Ma et al. (1990b). No reaction was observed without mixing and when NaOH
MeOH was added to the melted beef tallow in the reactor while stirring, stirring speed was
insignificant. Reaction time was the controlling factor in determining the yield of methyl esters.
This suggested that the stirring speeds investigated exceeded the threshold requirement of
mixing. Noureddini et al, (1998) studied the effect of mixing intensity on the transesterification
reaction of a continuous process for the conversion of vegetable oils into methyl esters of fatty
acids. They conducted several experiments at catalyst concentrations of 0.1-0.4 wt % with
varying mixing and molar ratios. They found out that higher mixing intensities and larger
catalyst concentrations both have favourable effects on the overall conversion.

Effect of using organic co-solvents
The methoxide base catalyzed methanolysis of soybean oil at 40 C (methanoloil molar ratio
6:1) shows that to form methyl esters proceeds approximately more slowly than butanolysis at 30
C. This is interpreted to be the result of a two phase reaction in which methanolysis occurs only
in the methanol phase. Low oil concentration in methanol causes the slow reaction rate; a slow
dissolving rate of the oil in methanol causes an initiation period. Intermediate mono- and di-
glycerides preferentially remain in the methanol, and react further, thus explaining the deviation
from second order kinetics. The same explanations apply for hydroxide ion catalyzed
methanolysis.



25
Using tetrahydrofuran, transesterification of soybean oil was carried out with methanol at
different concentrations of sodium hydroxide. The ester contents after 1 min for 1.1, 1.3, 1.4 and
2.0% sodium hydroxide were 82.5, 85, 87 and 96.2%, respectively. Results indicated that the
hydroxide concentration could be increased up to 1.3 wt%, resulting in 95% methyl ester after 15
min. (Boocock, 1998). Similarly for transesterification of coconut oil using THF/MeOH volume
ratio 0.87 with 1% NaOH catalyst, the conversion was 99% in 1 min.

A single-phase process for the esterification of a mixture of fatty acids and triglycerides were
investigated. The process comprised forming a single-phase solution of fatty acids and
triglyceride in an alcohol selected from methanol and ethanol, the ratio of said alcohol to
triglyceride being 15:135:1. The solution further comprised a cosolvent in an amount to form
the single phase. In a first step, an acid catalyst for the esterification of fatty acid is added. After
a period of time, the acid catalyst is neutralized and a base catalyst for the transesterification of
triglycerides is added. After a further period of time, esters are separated from the solution
(Boocock, 2001).

An improved process was investigated for methanolysis and ethanolysis of fatty acid glycerides
such as those found in naturally occurring fats and oils derived from plant and animals. The
processes comprised solubilizing oil or fat in methanol or ethanol by addition of a co-solvent in
order to form a one-phase reaction mixture, and adding an esterification catalyst. The processes
proceed quickly, usually in less than 20 min, at ambient temperatures, atmospheric pressure and
without agitation. The co-solvent increases the rate of reaction by making the oil soluble in
methanol, thus increasing contact of the reactants. The lower alkyl fatty acid monoesters
produced by the process can be used as biofuels and are suitable as diesel fuel replacements or
additives (Boocock, 1996b).

2.4 Product properties and quality
While the suitability of any material as fuel, including biodiesel, can be influenced by
contaminants arising from production or other sources, the nature of the fuel components
ultimately determines the fuel properties. Some of the properties included as specifications in
standards can be traced to the structure of the fatty esters comprising biodiesel. However, as
biodiesel consists of fatty acid esters, not only the structure of the fatty acids but also that of the
ester moiety derived from the alcohol can influence the fuel properties of biodiesel. Since the
transesterification reaction of oil or fat leads to a biodiesel fuel corresponding in its fatty acid
profiles with that of the parent oil or fat, biodiesel is a mixture of fatty esters with each ester


26
component contributing to the properties of the fuel. The properties of a biodiesel fuel that are
determined by the structure of its component fatty esters include ignition quality, heat of
combustion, cold flow, oxidative stability, viscosity and lubricity. The advantages that biodiesel
has over petroleum-based diesel are a lot. Biodiesels primary advantages lie in its effect on
emissions, cetane number, its flash point, and its lubricity.

The properties of biodiesel and diesel fuels, as given in Table 2.5, show many similarities, and
therefore, biodiesel is rated a strong candidate as an alternative to diesel.


Table 2.5: Properties of biodiesel from different oils (Barnwal, 2005)

Vegetable
Oil methyl
esters
(biodiesel)
Kinematic
viscosity
(mm
2
/s)
Cetane
no.
Lower
heating
value
(MJ/kg)
Cloud
point
(
o
C)
Pour point
(
o
C)
Flash
point
(
o
C)
Density
(kg/l)
Peanut 4.9 54 33.6 5 - 176 0.883
Soya bean 4.5 45 33.5 1 -7 178 0.885
Babassu 3.6 63 31.8 4 - 127 0.875
Palm 5.7 62 33.5 13 - 164 0.880
Sunflower 4.6 49 33.5 1 - 183 0.860
Tallow - - - 12 9 96 -
Diesel 3.06 50 43.8 - -16 76 0.855
20%
biodiesel
blend
3.2 51 43.2 - -16 128 0.859


Cetane number
The cetane number is an indication of a fuels readiness to auto ignite after it has been injected
into the diesel engine. Diesel fuel for use in on-highway engines is required to have a cetane
number of 40 or higher. Since a higher cetane number translates into higher fuel costs, most
refiners keep the cetane number of their diesel fuels between 40 and 45. Current research shows
that biodiesels higher cetane number (generally between 46 and 60 depending on the
feedstocks) shortens the ignition delay. Biodiesels lower volatility also tends to reduce the rate
at which fuel is prepared to burn during the ignition delay period (NBEP, 2007). Table 2.6 shows
the different types of biodiesel and its heat of combustion and corresponding cetane numbers.
These two factors contribute to improved combustion characteristics than occurs with fossil
diesel fuel


27

Table 2.6: Cetane number and energy content for biodiesel fuels (NBEP, 2007))
Type of Biodiesel Heat of Combustion in MJ/Kg Cetane No.
Methyl Soybean 39.8 46.2
Ethyl Soybean 40.0 48.2
Butyl Soybean 40.7 51.7
Methyl Sunflower 39.8 47.0
Methyl Peanut - 54.0
Methyl Rapeseed 40.1 -
Ethyl Rapeseed 41.4 -

Flashpoint
The flashpoint of a fuel is the temperature at which the vapours above the fuel become
flammable. Petroleum-based diesel fuels have flash points of 50 C to 80 C, so they are
considered to be intrinsically safe. Biodiesel has a flash point that is considerably higher than
fossil diesel fuel (above 160 C). This means that the fire hazard associated with transportation,
storage and utilization of biodiesel is much less than with other commonly used fuels (NBEP,
2007)

Lubricity
Lubricity can be defined as the property of a lubricant that causes a difference in friction under
conditions of boundary lubrication when all the known factors except the lubricant itself are the
same (Lee, 1995). Diesel with lower lubricity is desired for good engine performance. In the case
of diesel fuel, the fuel acts as a lubricant for the finely fitting parts in the diesel fuel injection
system. Biodiesel has much more lubricity than diesel fuel, and thus allows the engine to wear
less and last longer. Because of its solvent and lubricating properties, mechanics have reported
that engines running biodiesel look like new. Pure biodiesel and high level blends have excellent
lubricity; the addition of small amounts of biodiesel (0,25% to 2%) to diesel fuel has a dramatic
effect on the lubricity of fuel. The higher the biodiesel content in the fuel blend, the better the
fuel lubricity.

Cold flow
At low temperatures, biodiesel will gel or crystallize into a solid mass that cannot be filtered or
pumped. The engine cannot run at these temperatures. This is not a new problem for diesel


28
engine operators. Petroleum-based diesel fuel also gels but at temperatures that are lower than
for biodiesel (NBEP, 2007). The cloud point is the temperature at which crystals first start to
form in the fuel and the pour point is the lowest temperature at which the fuel will still pour from
a container. The cold filter plugging point (CFPP) is the lowest temperature at which a certain
volume of fuel can be drawn through a metal screen filter. It usually correlates well with the
lowest temperature that an engine can operate at.

Fuel stability
A fuel is considered unstable when it undergoes chemical changes that produce undesirable
consequences such as deposits, acidity or a bad smell. There are three different types of stability
commonly described in technical literature: thermal stability, oxidative stability and storage
stability. Vegetable oils are generally more susceptible to oxidative attack because they are less
saturated, that is, they contain more carbon-carbon double bonds. The short chain acids can be
volatile thus causing a foul smell and a lowering of the flashpoint. Polymerization can cause an
increase in viscosity and the formation of insoluble sediments and varnish deposits. Highly
saturated fuels, such as those made from tallow, are very resistant to oxidation and have high
Cetane numbers. However, they tend to have poor cold flow properties, often starting to
crystallize at temperatures as high as 10-15 C. Unsaturated fuels, such as those made from
soybean oil will generally stay liquid at temperatures down to 0 C (NBEP, 2007)

Fuel energy content
As can be seen in the Table 2.7, biodiesel has lower energy content (lower heating value) than
fossil diesel fuel. On a weight basis, the energy level is 13% less. But biodiesel is denser than
fossil diesel fuel, and so the energy content is only 8% less. Since diesel engines will inject equal
volumes of fuel, most diesel engine operators see a power loss of about 8%. In some cases, the
power loss may be even less than this. Biodiesels higher viscosity can decrease the amount of
fuel that leaks past the plungers in the diesel fuel injection pump.

Table 2.7. Comparison of diesel/biodiesel energy content and energy efficiency
Caloric value Fuel Density
g/cm
3

MJ/Kg MJ/dm
3

Energy efficiency
%
Diesel 0.83 42.90 35.60 38.20
Biodiesel 0.88 37.20 32.90 40.70
Variation -13 % -8 % + 7%



29

Material compatibility
Biodiesel interacts differently with materials than diesel fuel. Some metals have a catalytic effect
on the biodiesel oxidation process. Contact with these materials particularly in long-term storage
shortens material life span. Copper and copper-containing alloys such as brass and bronze are not
recommended for the storage of biodiesel. Lead, tin, and zinc are also cited as having some
incompatibility with biodiesel (Tyson, 2001). Blends of B20 or less do not seem to cause
problems within a reasonable time period. Higher level blends may however affect elastomer
materials that are used in diesel engine fuel system. While most modern diesel engines use steel
lines for the entire fuel distribution system, older engines may contain incompatible materials.
Older pumps may also contain elastomer diaphragms, seals and o-rings. These are usually made
from viton but if they are made from nitrile or natural rubbers, they will deteriorate on contact
with high levels of biodiesel (NBEP, 2007).

Biodegradability
The degradation of a compound through microbial activity in soils is called the biodegradability.
According to the standard test the result for biodiesel is that more than 95% are degraded after 21
days, fossil Diesel about 72% after 21 days (ABI, 2002). Dextrose (a test sugar used as the
positive control when testing biodegradability) degraded at the same rate. Blending biodiesel
with diesel fuel accelerates its biodegradability. For example, blends of 20% biodiesel and 80%
diesel fuel degrade twice as fast as fossil diesel alone (Levelton Engineering, 2002).


NOx emission
About 11 % of the weight of B100 is oxygen. The presence of oxygen in biodiesel improves
combustion and therefore reduces hydrocarbon, carbon monoxide, and particulate emissions as
shown in Figure 2.1; but oxygenated fuels also tend to increase nitrogen oxide emissions
(Szybist, 2005). Engine tests have confirmed the expected increases and decreases of each
exhaust component from engines without emissions controls.



30

Figure 2.1: Average emission impacts of biodiesel for heavy-duty highway engines (US EPA,
2002)

The increase in nitrogen oxide emissions from biodiesel is of enough concern that the National
Renewable Energy Laboratory (NREL) has sponsored research to find biodiesel formulations
that do not increase nitrogen oxide emissions. Adding cetane enhancers can reduce nitrogen
oxide emissions from biodiesel, and reducing the aromatic content of petroleum diesel from
31.9% to 25.8% is estimated to have the same effect. In the case of petroleum diesel, the
reduction in aromatic content can be accomplished by blending fuel. Nitrogen oxide emissions
from biodiesel blends could possibly be reduced by blending with kerosene or Fischer-Tropsch
diesel. Kerosene blended with 40 percent biodiesel has estimated emissions of nitrogen oxide no
higher than those of petroleum diesel, as does Fischer-Tropsch diesel blended with as much as
54 percent biodiesel (McCormick, 2003). Blending di-tert-butyl peroxide into B20 at 1 percent is
estimated to cost 17 cents per gallon (2002 cents), and blending 2-ethylhexyl nitrate at 0.5% is
estimated to cost 5 cents per gallon.

Oxides of nitrogen and hydrocarbons are ozone precursors. Carbon monoxide is also an ozone
precursor, but to a lesser extent than unburned hydrocarbons or nitrogen oxides. Air quality
modeling is needed to determine whether the use of biodiesel without additives to prevent
increases in nitrogen oxide emissions will increase or decrease ground-level ozone on balance
(US EPA, 2002).

Other emissions


31
When biodiesel displaces petroleum, it reduces levels of global warming gases such as CO
2
.
After the oil is extracted from soybeans, for example, it is refined into biodiesel and, when
burned, produces CO
2
and other emissions, which are returned to the atmosphere. As other
plantations of soybeans grow, they take CO
2
from the air to make the stems, roots, leaves and
seeds. Therefore this cycle does not add to the CO
2
level in the air because the next soybean crop
will reuse the same amount of CO
2
to grow.

Another important environmental factor is that biodiesel reduces tailpipe particulate matter
(PM), HC and CO emissions. These benefits occur because biodiesel contains 11% oxygen (O
2
)
by weight. The presence of O
2
allows the fuel to burn more completely, resulting in fewer
emissions from unburned fuel. This same principle also reduces air toxicity, which is associated
with the unburned or partially burned HC and PM emissions. Testing has shown that PM, HC
and CO reductions are independent of the vegetable oil used to make biodiesel.

2.5 Safety and Precautions
Biodiesel is not only non-toxic, but also almost not flammable. The fuel has to be heated up to
over 150 C before combustion can take place in an engine. The only reason it works in the
diesel engine is because diesels compress the vaporized fuel and air mixture so much so that it
combusts (without the need for spark plugs). Thus, if a biodiesel powered car, truck or boat is
involved in an accident, the resulting spill wont require specialised personnel to clean up, and
will very likely not result in a fire.

Biodiesel is a good solvent and will clean out the soot and other junk left in an engine and lines
by regular diesel fuel. This junk will eventually clog up a vehicles fuel filter. So it is
recommended that the fuel filter of a vehicle be changed after running a couple of tankfulls of
biodiesel. Also, biodiesel tends to degrade rubber which is a problem in pre-1993 vehicles that
have some of the hoses and seals made out of rubber. Such vehicle engines must be regularly
inspected for signs of swelling or degrading and the hoses and seals replaced if necessary with
synthetic lines, made out of a material called Viton. Depending on the type of oil the biodiesel
was made from, it begins to cloud up and gel (crystallize) at lower temperatures than petrodiesel.
It. For very cold temperatures, it is usually recommended that the biodiesel be blended with
fossil diesel.



32
2.6 Biodiesel production via transesterification in supercritical medium
Recently, some innovative processes for biodiesel production via transesterification of oils have
emerged. One of such new approaches proposes to use the supercritical single phase reaction.

As discussed previously, the production of biodiesel using vegetable oils is reacted with acid or
alkaline catalysts in the presence of methanol to yield methyl esters and glycerol (by products)
and the whole process follows a two phase reaction system. The rate of transesterification of
triglycerides is low and it is followed by a slow liquid-liquid separation step that involves
washing the products with a large quantity of water to purify the phases.

Production of biodiesel using supercritical methanol follows the same general reaction as the
traditional transesterification reaction. The difference is that the reaction does not require a base
or acid catalyst. Basically, methanol at supercritical conditions becomes an excellent solvent and
dissolves the feedstock so that the molecules of the reactants are in close proximity of each other
and therefore react readily without a distinct catalyst. Therefore, unlike the two-phase nature of
the biodiesel production using catalyst methods, there is no time consuming and complicated
separation of the product and catalyst. This reduces production costs, energy consumption and
water usage. Supercritical methanol forms a single phase as a result of the lower value of the
dielectric constant of methanol in the supercritical state. A co-solvent, usually a light
hydrocarbon such as propane, is also required. Without a co-solvent, the operating temperature
and pressure of the process would be about 650-750 C and 6500-9500 PSIG. Co-solvents can
decrease the critical point of methanol and allow the supercritical reaction to be carried out under
milder conditions.

2.7 Production of biodiesel by pyrolysis and hydroprocessing of oils and fats
Pyrolysis, strictly defined, is the conversion of one substance into another by means of heat with
the aid of a catalyst in the absence of oxygen or in the reductive environment, e.g. using
molecular hydrogen. It involves cleavage of chemical bonds of organic molecules to yield low
molecular weight hydrocarbons. Pyrolysis chemistry is difficult to characterize because of the
variety of reaction paths and the variety of products that may be obtained. The pyrolyzed
material can be any type of biomass, such as vegetable oils, animal fats, wood, bio-waste, etc.

Different types of pyrolysis exist, depending on the temperature, reaction time, pressure,
reagents, and catalysts involved. For example, in the case of the use of pyrolytic technologies for
biomass conversion, fast or flash pyrolysis technology is utilized for conversion of solid biomass


33
into bio-oil (pyrolysis oil) - the process also known as liquefaction. Other types of pyrolysis for
liquefaction of biomass include direct hydrothermal liquefaction, where biomass undergoes
heating and pressurizing in contact with water in the presence of alkali. Slow pyrolysis results in
predominantly gaseous products (bio-syngas). The use of pyrolysis for conversion of solid
biomass into fuels is discussed in the second generation biofuel technologies (chapter 3).

The pyrolysis of oils and fats has been investigated for more than 100 years, especially in those
areas of the world that lack deposits of petroleum. Such pyrolysis (also called cracking) can be
carried out directly (direct thermal cracking) or in the presence of catalysts. Cracking of oils in
the presence of hydrogen (used to remove oxygen from oil molecules) can be also called
hydrocracking (terms like hydroprocessing, hydrogenation, hydrogenolysis or hydrotreatment
are sometimes used). In the catalytic cracking reaction, typically, four main catalyst types are
used including transition metal catalysts, molecular sieve type catalysts, activated alumina, and
sodium carbonate. Basically, cracking represents the same principle of conversion as pyrolysis,
where it applies for conversion of higher molecular weight oils to lower molecular weight ones.
Hydrocracking involves the same reaction in the presence of hydrogen in order to yield oxygen
free fossil diesel like hydrocarbons. In this case hydrogenation and cracking reactions occur
simultaneously. The general mechanism for the thermal decomposition of a triglyceride is given
in scheme 2.4.

Hydroprocessing vegetable oils and greases by using existing hydrocracking technology will
result in a fully-deoxygenated, high-cetane (80-90), straight-chain paraffins which are superior to
minimum quality standards for diesel fuel. There are two general options available for the
hydroprocessing of the vegetable oils, i.e. co-processing in existing hydrotreaters and processing
in a separate modular unit.



34
CH
3
-(CH
2
)
5
-CH
2
-CH
2
-CH=CH-CH
2
-CH
2
-(CH
2
)
5
-CO-O-CH
2
R
CH
3
-(CH
2
)
5
-CH
2
-CH
2
-CH=CH-CH
2
-CH
2
-(CH
2
)
5
-CO-OH
CH
2
=CH-CH=CH
2
CH
3
-(CH
2
)
5
-CH
2
CH
2
-(CH
2
)
5
-CO-OH
CH
3
-(CH
2
)
3
-CH
2
CH
2
=CH
2
+
Diels Alder
-2H
2
CH
3
-(CH
2
)
3
-CH
3
H
H
CH
3
-(CH
2
)
5
-CO-OH
CH
3
-(CH
2
)
4
-CH
3
-CO
2


Scheme 2.4: General representation of the mechanism of thermal decomposition of triglycerides

Generally, the first option will greatly reduce capital cost since no new units will be required.
However, the second option will allow for optimization of the processing conditions and catalyst
for conversion of the bio-feedstock, minimize the use of expensive metallurgy for acidic bio-
feedstocks and products and allow for the removal of products such as carbon oxides and H
2
O
which interfere with the hydroprocessing catalyst.

The reactor uses a conventional commercial refinery hydrotreating catalyst and hydrogen.
Several reactions occur in the reactor: hydrocracking (breaking apart of large molecules),
hydrotreating (removal of oxygen) and hydrogenation (saturation of double bonds). The
hydrotreating options can be successfully implemented to the number of feedstocks for example,
canola oil, soya oil, yellow grease, animal tallow and tall oil (a by-product of the Kraft pulping
process).

The co-products can be easily integrated into a pre-existing refinery infrastructure. The
hydrocarbon liquid can be distilled into three basic fractions: naphtha middle distillate (CETC
SuperCetane) waxy residues. When yellow grease, animal tallow and vegetable oils are used as
feedstocks, negligible amounts of naphtha and small volumes of waxy residues are produced.
Since the naphtha fraction is so small it is usually not necessary to remove it from the
SuperCetane. The waxy residue is paraffin-rich and can be used as refinery feedstock or power
boiler fuel.


35

The middle distillate (SuperCetane) is the primary liquid product where yields of 70-80% were
achieved for yellow grease and tallow. It consists mainly of straight chain hydrocarbons in the
diesel fuel boiling range with a cetane number of about 100. The high cetane diesel fuel is a
means of reducing emissions of engine pollutants and improving fuel economy. CETC
SuperCetane also resembles conventional diesel fuel when analyzed by GC/MS, is miscible in all
proportions with diesel fuel and the sulphur content is less than 10 ppm.

As shown in scheme 2.5 the general set up uses a hydrotreating process with conventional
petroleum refinery units this same can be used for hydrotreating of different types of biomass
feedstocks to produce good quality biodiesel.



Scheme 2.5. General arrangement of the biodiesel production process via the hydroprocessing
technology (NRC, 2004)

While the processes of pyrolysis of solid biomass, coal, and cracking of bio-oil are well studied
and implemented on the large scale, the cracking technologies applied to vegetable oils and fats
for fuel production are still in the research phase. No commercial technologies or scaled up
productions have been introduced so far. There are conflicting views on the future potential of
development of the technology based on hydrocracking of vegetable oils for diesel fuel
production. While some say that the technology, once commercialized, will represent a
competitive alternative to the conventional biodiesel, the others indicate that significant
drawbacks and limitations exist. Besides higher cost of the equipment for thermal cracking and
Feed Hydrogen

Fuel gas by-
product
Water
Liquid product

Low sulphur high
cetane diesel blending
stock (Super Cetane)
Distillation column
Gas recycle
stream
Waxy paraffinic
residue
Reactor
Separator


36
pyrolysis especially for modest throughputs, there is also concern that the removal of oxygen
during the thermal processing removes any environmental benefits of using an oxygenated fuel.
On the other hand, the deoxygenated diesel fuels obtained via hydrocracking process are more
stable than FAME, which contain oxygen and double bonds and can sometimes be better suitable
for use at low temperatures. Other advantages of the biodiesel via pyrolysis and hydrocracking
processes lie in a lower amount of by-products formed and less demand for feedstock quality.

2.8 Market and economic aspects of biodiesel production
Although the use of biodiesel may provide significant environmental advantages (pollutant
emissions) when compared to the combustion of fossil fuels, the effective use of biodiesel will
depend to a large extent on its economic feasibility.

Final price
Final selling price is a critical factor in the feasibility of biodiesel. It basically depends on three
parameters: VAT (value added tax), mineral oil taxes and production price. Alternate uses for
soybean oil and even animal fats and recycled frying oils keep their price at a level where they
cannot compete directly against fossil diesel fuel. Some type of government subsidies are
necessary for the industry to develop. These subsidies, primarily in the form of tax waivers, have
been responsible for the rapid growth of biodiesel use in Europe. In various European countries,
the price of biodiesel is actually less than for petroleum diesel fuel. A mandate requiring the use
of a minimum biodiesel blend in all fuels sold for use is another type of subsidy: it removes price
as an issue since the fuel suppliers will have to pay whatever is necessary to get the biodiesel
they need for blending (NBEP, 2007)).

Production costs
In the early days, biodiesel producers were satisfied when achieving a transesterification rate
(yield) of approx. 85 - 95 % thus leaving quite a volume of potential feedstock as waste in the
glycerine phase. However, yield is the second biggest factor affecting profitability, i.e. a 10%
decline of yield reduces profitability by approximately 25 %. It is therefore crucial to transfer
any potential molecule into a fatty-acid methyl-ester; this includes all the triglycerides and Free-
Fatty-Acids (FFA). A modern and profitable process technology today is able to achieve a 100 %
yield without any expensive losses (Austrian Biofuels Institute, 2000).

Feedstock costs


37
The high price of biodiesel oil feedstock is the major obstacle to market development. As the
selling price of biodiesel must exceed the feedstock cost to cover processing, packaging,
transportation, distribution and profit, it is highly sensitive to any widening of the fossil oil-
vegetable oil price gap (Krbitz, 2002). The most promising approach to lowering the price is to
use other, less expensive feedstocks to provide a portion of the biodiesel supply. These other
feedstocks could include spoiled soybeans, beef and pork tallow, recycled restaurant frying oils
and by-products, such as soapstock, from other processes involving vegetable oils. While the
quantities of these feedstocks are not sufficient to supply a large market, they can be used as
blending agents to lower the overall costs (Levelton, 2002).

Tax incentives
Tax relief or subsidies are essential for commercial production of biodiesel due to present
processing costs to produce biodiesel, which are much higher than those for petroleum diesel.
These incentives are being used in several countries (Pramanik, 2005):
US Federal Government introduced a legislation providing for a 1 % volume reduction in
diesel fuel tax over three year period (until December 2005) for every percent of biodiesel
blended with diesel to 20 %.
Germany does not levy excise tax on biodiesel.
France offers a tax incentive of 0.35 Euro/l for VOME (Vegetable oil methyl ester) blended
with diesel.
UK has a 20 % lower tax.
Austrian tax law on Tax Reform 2000 exempts to 2 % biodiesel blend from mineral oil taxes.
Finland has adopted a tax incentive for reformulated diesel fuel.

It should be mentioned that technical, economic and environmental factors are not sufficient to
fully explain the use or non-use of a given fuel; other factors such as:
Sociological and cultural aspects
Organisational aspects
Institutional, structural and political aspects would have to be investigated, but they were
partly neglected as this would have gone beyond the scope of this work.

2.9 Bioethanol from sugar and Starch crops
Bioethanol is produced by fermenting sugars from starch and sugar biomass (e.g. cereal crops
such as corn and sugarcane). Ethanol also is produced through chemical reactions using ethylene
and steam as reactants. It can be used in pure form in specially adapted vehicles or blended with


38
gasoline provided that fuel specifications are met. For example, pure ethanol (E100) is used in
some vehicles in Brazil and a mixture of 15% ethanol and 85% gasoline by volume (E85) is also
widely used in the flexible-fuel vehicles (FFV) in Brazil, Sweden, and US. The most common
blend is 10% ethanol and 90% petrol (E10) as most standard spark-ignited engines are able to
run on E10. Vehicle engines require no modifications to run on E10 and vehicle warranties are
unaffected also. Only flexible fuel vehicles can run on E85 and E100.

Ethanol is colourless, biodegradable and has low toxicity. Ethanol has low environmental
pollution compared to other chemicals. The use of ethanol as an alternative transport fuel has
been on the ascendancy in different parts of the world for various reasons:
to reduce dependence of petroleum fuel
to reduce air pollution
mitigation of global climate change and
to create more jobs in rural areas in especially developing and least developed countries.
In general ethanol unlike gasoline contains 35% of oxygen which reduces the emissions of NO
x

and particulate matters from combustion.

Ethanol could be produced through chemical reactions or by conversion of biomass materials by
fermentation processes following three major steps: the formation of a solution of fermentable
sugars; the fermentation of these sugars to ethanol; and the separation and purification of the
ethanol, usually by distillation.

Bioethanol from sugar feedstocks
The easiest way to produce ethanol is from C
6
sugars using fermentation process. The most
common and easiest sugar to convert to ethanol is glucose (C
6
) or biomass containing higher
levels of glucose or precursors to glucose. Many microorganisms like fungi, bacteria, and yeast
can be used for fermentation of sugars but Saccharomyces cerevisiae also known as Bakers yeast
is frequently used to ferment sugar to ethanol. Sugarcane is a typical example for sugar
feedstock. Brazil is one of the principal producers of bioethanol in the world for fuel and the
principal feedstock is sugarcane. Other sugar rich biomass feedstocks are sweet sorghum, sugar
beet, and also different types of fruits. However, all these materials are in the human food chain
and these makes them usually too expensive to use for ethanol production but then waste
residues can be used as alternative feedstock for the production of bioethanol.

Bioethanol from starch


39
Starch is another feedstock which is readily available and is made up of long chains of glucose
molecules which can be fragmented into simple sugars before fermentation to produce ethanol.
Starch biomass feedstocks include tubers like sweet potato, potato, cassava and cereal grains, etc.
Starchy feedstocks undergo hydrolysis to breakdown the starch into fermentable sugars i.e.,
saccharification. The hydrolysis of starch can be carried out by mixing water with the feedstocks
to form slurry which is then heated to rupture the cell walls and different specific enzymes are
added during hydrolysis to break chemical bonds present in the starch materials.

Bioethanol production processes
Wet milling process: Cereals can be processed into ethanol by either the dry milling or the wet
milling process. In the wet milling process, cereals like corn kernel is steeped in warm water, this
helps to break down the proteins and release the starch present in the corn and helps to soften the
kernel for the milling process. The corn is then milled to produce germ, fibre and starch
products. The germ is extracted to produce corn oil and the starch fraction undergoes
centrifugation and saccharification to produce gluten wet cake. The ethanol is then extracted by
the distillation process. The wet milling process is normally used in factories producing several
hundred million gallons of ethanol every year.

Dry milling process: The dry milling process involves cleaning and breaking down the cereals
kernel into fine particles using a hammer milling process. This creates a powder with a coarse
flour type consistency. The powder contains the cereals germ, starch and fibre. In order to
produce a sugar solution the mixture is then hydrolysed or broken down into sucrose sugars
using enzymes or a dilute acid. The mixture is then cooled and yeast is added in order to ferment
the mixture into ethanol. The dry milling process is normally used in factories producing less
than 50 million gallons of ethanol every year.

Sugar fermentation process: The hydrolysis process breaks down the cellulosic part of the
biomass or corn into sugar solutions that can then be fermented into ethanol. Yeast is added to
the solution, which is then heated. The yeast contains an enzyme called invertase, which acts as a
catalyst and helps to convert the sucrose sugars into glucose and fructose (both C
6
H
12
O
6
).

The chemical reactions are shown below:



40
C
12
H
22
O
11
+ H
2
O C
6
H
12
O
6
+ C
6
H
12
O
6
Sucrose Fructose Glucose
Invertase


The fructose and glucose sugars then react with another enzyme called zymase, which is also
contained in the yeast to produce ethanol and carbon dioxide. The chemical reaction is shown
below:
C
2
H
5
OH+ 2CO
2
Fructose/Glucose Ethanol
Zymase
C
6
H
12
O
6


The fermentation process takes around three days to complete and is carried out at a temperature
of between 250 C and 300 C. Table 2.8 shows the different types of carbohydrate feedstocks
and corresponding enzymes used for the production of bioethanol.

Table 2.8: Feedstocks and enzymes used for production of bioethanol (sugar crops)
Carbohydrate
Feedstock
Main carbohydrate
to be converted
Process utilizing added
enzymes
Required enzyme
Sugar cane or
molasses
Sucrose Dextran hydrolysis Dextranase
Grains and
tubers
Starch (-1,4 linked
glucose molecules)
Starch, beta glucan and
pentosan hydrolysis
Amylase, glucoamylase,
betagulcanase, xylanase

Fractional distillation process: The ethanol, which is produced from the fermentation process,
still contains a significant quantity of water, which must be removed. This is achieved by using
the fractional distillation process. The distillation process works by boiling the water and ethanol
mixture. Since ethanol has a lower boiling point (78.3 C) compared to that of water (100 C),
the ethanol turns into the vapour state before the water and can be condensed and separated.

Advantages and disadvantages of bioethanol as fuel
Bioethanol has a number of benefits over conventional fuels. It is a renewable resource and it
reduces greenhouse gas emissions. By encouraging bioethanol use, the rural economy would
also receive a boost from growing the necessary crops. Bioethanol is also biodegradable and far
less toxic that fossil fuels. In addition, using bioethanol in older engines can help reduce the


41
amount of carbon monoxide produced by the vehicle thus improving air quality. Another
advantage of bioethanol is the ease with which it can be easily integrated into the existing road
transport fuel system. Bioethanol can be blended with conventional fuel up to 5% (E5) without
the need for engine modifications. Bioethanol is produced using familiar methods, such as
fermentation, and it can be distributed using the same petrol forecourts and transportation
systems as before.

Though ethanol based fuels has many advantages over fossil fuels, it is not compatible with
some fuel system components. It may corrode ferrous components, leave jelly-like and salt
deposits on fuel strainer screens, etc. Water content in ethanol should be less than 1.0%
otherwise, phase separation occur during blending with gasoline and moreover, ethanol will mix
with either water or gasoline but not in both. Ethanol blended gasoline can also affect electric
fuel pumps by causing internal wear and undesirable spark generation.

2.10 Biogas by anaerobic digestion
Biogas is obtained by anaerobic treatment of manure and other humid biomass materials (e.g. in
landfills), including food waste, and then upgraded to biomethane that can be feed-in into the
natural gas grid and also used in natural gas vehicles. There are conflicting views on whether or
not biogas should be refereed to as first or the second generation biofuel, because it can be
produced from a variety of biomass and not only from food crops. However, in view of the
maturity and wide use of the technology of its production, in this book biogas is considered
within the first generation, which is also an opinion shared by many experts in the field.

Biomethane can play an important role when feed-in to the natural gas grid for use as natural gas
substitute or in addition to that. Methane has a high market potential as a well known energy
carrier for the transport sector and stationary applications (heat and power) and for material
utilisation. Within the existing and well developed natural gas grid biomethane can easily be fed
in and distributed to the final consumer in industry and households. In addition to the above
mentioned advantages the combustion properties of methane are already well known and
characterised through comparably low emissions.

As shown in Scheme 2.6, biomethane can be provided either by the treatment of bio-chemically
produced biogas or by the synthesis of treated gas coming from thermo-chemical gasification. In
this chapter we will focus on the biomethane produced via the biochemical pathway.



42

Scheme 2.6: Options for the provision of biomethane

Biomethane by anaerobic digestion (via biogas)
In recent years, increasing awareness that anaerobic digesters can help control the disposal and
odour of animal waste has stimulated renewed interest in the technology. New digesters are now
being built because they effectively eliminate the environmental hazards of dairy farms and other
animal feedlots. Notably, it is often the environmental reasons rather than the digesters
electrical and thermal energy generation potential that motivate farmers to use digester
technology. This is especially true in areas where electric power costs are low. Developing rural
areas find this technology highly attractive due to the fact that it requires low cost infrastructures.

Furthermore, anaerobic digester systems can reduce fecal coliform bacteria in manure by more
than 99%, virtually eliminating a major source of water pollution. Separation of the solids during
the digester process removes about 25 % of the nutrients from manure, and the solids can be sold
out of the drainage basin where nutrient loading may be a problem. In addition, the digesters
ability to produce and capture methane from the manure reduces the amount of methane that
otherwise would enter the atmosphere. Methane is a green house gas and its release in the
atmosphere contributes to global climate change.

The process of anaerobic digestion
The process of anaerobic digestion occurs in a sequence of stages involving distinct types of
bacteria. Hydrolytic and fermentative bacteria first break down the carbohydrates, proteins and


43
fats present in biomass feedstock into fatty acids, alcohol, carbon dioxide, hydrogen, ammonia
and sulphides. This stage is called hydrolysis or, sometimes, liquefaction (not to confuse with
thermal liquefaction of biomass based on pyrolysis).

Next, acetogenic (acid-forming) bacteria further digest the products of hydrolysis into acetic
acid, hydrogen and carbon dioxide. Methanogenic (methane-forming) bacteria then convert these
products into biogas.

The combustion of digester gas can supply useful energy in the form of hot air, hot water or
steam. After filtering and drying, digester gas is suitable as fuel for an internal combustion
engine, which, combined with a generator, can produce electricity. Future applications of
digester gas may include electric power production from gas turbines or fuel cells. Digester gas
can substitute for natural gas or propane in space heaters, refrigeration equipment, cooking
stoves or other equipment. Compressed digester gas can be used as an alternative transportation
fuel.

Digester Technology
Biomass that is high in moisture content, such as animal manure and food-processing waste, is
suitable for producing biogas using anaerobic digester technology. During anaerobic digestion,
bacteria digest biomass in an oxygen-free environment (Scheme 2.7). Symbiotic groups of
bacteria perform different functions at different stages of the digestion process.




44
Scheme 2.7: Anaerobic digestion

There are four basic types of microorganisms involved. Hydrolytic bacteria break down complex
organic wastes into sugars and amino acids. Fermentative bacteria then convert those products
into organic acids. Acidogenic microorganisms convert the acids into hydrogen, carbon dioxide
and acetate. Finally, the methanogenic bacteria produce biogas from acetic acid, hydrogen and
carbon dioxide.

Controlled anaerobic digestion requires an airtight chamber, called a digester. To promote
bacterial activity, the digester must maintain a temperature of at least 18 C. Using higher
temperatures, up to 65 C, shortens processing time and reduces the required volume of the tank
by 25 to 40%. However, there are more species of anaerobic bacteria that thrive in the
temperature range of a standard design (mesophilic bacteria) than there are species that thrive at
higher temperatures (thermophilic bacteria). High-temperature digesters also are more prone to
upset because of temperature fluctuations and their successful operation requires close
monitoring and diligent maintenance.

The biogas produced in a digester (also known as digester gas) is actually a mixture of gases,
with methane and carbon dioxide making up more than 90 % of the total. Biogas typically
contains smaller amounts of hydrogen sulphide, nitrogen, hydrogen, methylmercaptans and
oxygen.

Methane is a combustible gas. The energy content of digester gas depends on the amount of
methane it contains. Methane content varies from about 55 to 80 %. Typical digester gas, with a
methane concentration of 65%, contains about 600 Btu of energy per cubic foot.

For individual farms, small-scale plug-flow or covered lagoon digesters of simple design can
produce biogas for on-site electricity and heat generation. For example, a plug-flow digester
could process 8,000 gallons of manure per day, the amount produced by a herd of 500 dairy
cows. By using digester gas to fuel an engine-generator, a digester of this size would produce
more electricity and hot water than the dairy consumes.

Larger scale digesters are suitable for manure volumes of 25,000 to 100,000 gallons per day. In
Denmark and in several other European countries, central digester facilities use manure and
other organic wastes collected from individual farms and transported to the facility.


45

Types of Anaerobic Digesters
There are three basic digester designs. All of them can trap methane and reduce fecal coliform
bacteria, but they differ in cost, climate suitability and the concentration of manure solids they
can digest.

A covered lagoon digester, as the name suggests, consists of a manure storage lagoon with a
cover. The cover traps gas produced during decomposition of the manure. This type of digester is
the least expensive of the three. Covering a manure storage lagoon is a simple form of digester
technology suitable for liquid manure with less than 3-percent solids. For this type of digester, an
impermeable floating cover of industrial fabric covers all or part of the lagoon. A concrete
footing along the edge of the lagoon holds the cover in place with an airtight seal. Methane
produced in the lagoon collects under the cover. A suction pipe extracts the gas for use. Covered
lagoon digesters require large lagoon volumes and a warm climate. Covered lagoons have low
capital cost, but these systems are not suitable for locations in cooler climates or locations where
a high water table exists.

A complete mix digester converts organic waste to biogas in a heated tank above or below
ground. A mechanical or gas mixer keeps the solids in suspension. Complete mix digesters are
expensive to construct and cost more than plug-flow digesters to operate and maintain.

Complete mix digesters are suitable for larger manure volumes having solids concentration of 3
percent to 10 percent. The reactor is a circular steel or poured concrete container. During the
digestion process, the manure slurry is continuously mixed to keep the solids in suspension.
Biogas accumulates at the top of the digester. The biogas can be used as fuel for an engine-
generator to produce electricity or as boiler fuel to produce steam. Using waste heat from the
engine or boiler to warm the slurry in the digester reduces retention time to less than 20 days.

Plug-flow digesters are suitable for ruminant animal manure that has a solids concentration of 11
percent to 13 percent. A typical design for a plug-flow system includes a manure collection
system, a mixing pit and the digester itself. In the mixing pit, the addition of water adjusts the
proportion of solids in the manure slurry to the optimal consistency. The digester is a long,
rectangular container, usually built below-grade, with an airtight, expandable cover.



46
New material added to the tank at one end pushes older material to the opposite end. Coarse
solids in ruminant manure form a viscous material as they are digested, limiting solids separation
in the digester tank. As a result, the material flows through the tank in a "plug." Average
retention time (the time a manure plug remains in the digester) is 20 to 30 days.

Anaerobic digestion of the manure slurry releases biogas as the material flows through the
digester. A flexible, impermeable cover on the digester traps the gas. Pipes beneath the cover
carry the biogas from the digester to an engine-generator set.

A plug-flow digester requires minimal maintenance. Waste heat from the engine-generator can
be used to heat the digester. Inside the digester, suspended heating pipes allow hot water to
circulate. The hot water heats the digester to keep the slurry at 25C to 40C, a temperature range
suitable for methane-producing bacteria. The hot water can come from recovered waste heat
from an engine generator fuelled with digester gas or from burning digester gas directly in a
boiler.

Biogas from wastes
Municipal sewage contains organic biomass solids, and many wastewater treatment plants use
anaerobic digestion to reduce the volume of these solids. Anaerobic digestion stabilizes sewage
sludge and destroys pathogens. Sludge digestion produces biogas containing 60 to 70% methane,
with an energy content of about 600 Btu per cubic foot.

Most wastewater treatment plants that use anaerobic digesters burn the gas for heat to maintain
digester temperatures and to heat building space. Unused gas is burned off as waste but could be
used for fuel in an engine-generator or fuel cell to produce electric power. Before use, the gas is
cleaned to remove impurities. These are principally hydrogen sulphide, halogens (fluorine,
chlorine and bromine), moisture, bacteria and solids. Biogas also contains carbon dioxide, which
cannot be removed easily.

Landfill Gas
Underground decomposition of cellulose contained in municipal and industrial solid waste
produces biogas as shown in Scheme 2.8. The digestion occurring in landfills is an uncontrolled
process of biomass decay.



47
The efficiency of the process depends on the waste composition and moisture content of the
landfill, cover material, temperature and other factors. The biogas released from landfills,
commonly called landfill gas, is typically 50% methane, 45% carbon dioxide and 5% other
gases. The energy content of landfill gas is 400 to 550 Btu per cubic foot.

Capturing landfill gas before it escapes to the atmosphere allows for conversion to useful energy.
A landfill must be at least 40 feet deep and have at least one million tons of waste in place for
landfill gas collection and power production to be technically feasible. Notably that by capturing
landfill gas will contribute to the reduction of green house gas emissions.

A landfill gas-to-energy system consists of a series of wells drilled into the landfill. A piping
system connects the wells and collects the gas. Dryers remove moisture from the gas, and filters
remove impurities. The gas typically fuels an engine-generator set or gas turbine to produce
electricity. The gas also can fuel a boiler to produce heat or steam. Further gas cleanup improves
biogas to pipeline quality, the equivalent of natural gas. Reforming the gas to hydrogen would
make possible the production of electricity using fuel cell technology.


Scheme 2.8: Landfill gas


48

Biogas upgrading to biomethane
In order to feed biogas into the gas grid, the raw biogas has to undergo two major processes to
obtain natural gas quality, namely cleaning and upgrading. For the employment in a steam
reformer the biogas has to be purified from trace components, primarily H
2
O, H
2
S and NH
3
. The
heating value, Wobbe index and other parameters, which highly depend on the CH
4
content, are
adjusted to pipeline specifications by removing the CO
2
. The upgrading process basically
consists of the separation of CH
4
and CO
2
.

A couple of technologies enable the removal of H
2
S and CO
2
combining the two steps. However,
the relevance, feasibility and sequence of the different cleaning and upgrading processes depend
on the specific gas composition and pipeline specifications. Upgrading technologies for biogas
result mainly from technological applications from the natural gas sector where partly
comparable upgrading tasks have to be solved, but at a by far larger scale. Until today, some
technologies show long time experiences but on the background of actual political discussions
and decisions not only companies change the owner - also lots of research is done in this field.
Thus, the main technologies will be characterized in the following.

Generally it has to be stated that all companies offering upgrading plants pledge to reach a
biomethane quality of more than 96% methane content and a methane loss with less than 3%.
Additionally a very high availability for all technologies is promised and confirmed in practice
with at most plants about 95 %. Practical measurements in Sweden have shown different results,
but more or less independent on the used technology. Especially it seems to be necessary to care
about the methane losses which are reported with up to 10%, because they influence ecological
and economical parameters of such plants very strongly.

Water scrubbing
One of the mainly used technologies for biogas upgrading is the water scrubbing process which
is very common in Sweden and shows a very long history of experiences. The process is based
on sorption of carbon dioxide in water at high pressures at about 10 bar and desorption of carbon
dioxide at lower pressures in another vessel to separate carbon dioxide and methane. At the same
time most trace gases as for example H
2
S are separated during this process. As water source
circulated fresh water or cleaned waste water without circulation can be used. Advantages of
waste water use are compensated by bio fouling in the scrubbers due to the relatively high
organics content.


49

Due to the long research and development period the technology has reached a very high level of
development and is proven in practice. Thus, it has to be expected, that the technology has to
compete with other technologies in economical questions but is seen as technological state of the
art.

Pressure swing adsorption
A lot of upgrading plants install pressure swing absorption technologies for biomethane
production. The technology is based on high pressure absorption of carbon dioxide at molecular
sieves or activated charcoal in at least two steps. Other trace gases have to be separated e. g. in
filter systems before carbon dioxide absorption. Absorption materials have long operation
durations but have to be exchanged from time to time. This technology shows long time
experiences and is well established and competes as solid and reliable technology.


Chemical adsorption
Chemical absorption is possible with a number of substances (e.g. Selexol, MEA, DEA and
others), which are capable to absorb carbon dioxide at ambient pressure combined with heat
demand or at raised pressure. Available information about efficiencies of the processes are very
different, but a lot of experiences exist and research is going on. The advantage of chemical
absorption is that used substances are selective for carbon dioxide (and sometimes for H
2
S too,
most times H
2
S-separation is done before upgrading) and thus no methane will be absorbed what
results in high methane concentrations in the purified gas. Absorption processes under pressure
seem to be of comparable behaviour as water scrubbing processes but ambient pressure systems
(where only pilot plants are known) promise less electrical energy demand for the continuous
process and could be of high interest for applications where only low pressure is required after
upgrading and heat is available. Summarizing it can be stated that some chemical absorption
technologies are reliable and well established and others are under development.

Membrane technologies
Due to the different molecule size of methane and carbon dioxide it is generally possible to
separate both gases through membranes. The smaller methane molecules can pass a membrane
and a gas with very high methane content and a gas with very high carbon dioxide content can be
produced. Technologies were developed for dry separation (transport through the membrane is
forced by very high pressure) and for wet separation (transport is forced by a very low


50
concentration of methane in a fluid where methane is absorbed in). Both technologies are only
known from pilot plants and efficiencies regarding methane losses and energy demand are not
known from continuously practiced processes.

Cryogenic upgrading
Additionally it is possible to use the differences in dew and condensation point of methane and
carbon dioxide. Therefore it has been shown in small scale that both gas components can be
separated by cooling down of biogas to less than -45 C at about 80 bar pressure. Applications of
this process, called cryogenic upgrading, are under research in Sweden and Germany but not
available at the market.

For safety reasons the treated biogas has to be odorised prior to being injected into public natural
gas grid. Concerning the different biogas treatment options, further technical equipment is
required, including an appropriate periphery to the gas grid, compressors adjusting required grid
pressure specifications as well as gas counter and measurement of the injected gas composition.
If biogas is injected into the nearest low-pressure gas grid (e.g. 4 to 5 bar) additional gas
compression is not required. Most of the upgrading technologies (e. g. water scrubbing and PSA)
are appropriate to provide biogas at this pressure level. In addition, a gas mixer might be
necessary to add high-calorific gas (e.g. propane or butane). This addition is an option aimed at
maximising the methane yield if biogas is upgraded to a lower gas quality as required.



51
2.11 Bibliography

Ahn E. et al. 1995, Sep. Sci. Technol., 30, 2021.
ABI (Austrian Biofuels Institute) 2000, World-wide Trends in Production and Marketing of
Biodiesel; presented at the ALTENER Seminar New Markets for Biodiesel in Modern
Common Rail Diesel Engines, University for Technology in Graz, Graz.
ABI (Austrian Biofuels Institute) 2002, Statement on the situation of the Biodiesel sector,
Vienna, 2.
Barnwal B.K. et al. 2005, Renewable and Sustainable Energy Reviews, 9, 363.
Boocock D.G.B. et al. 1996a, Biomass Bioenergy, 11, 43.
Boocock D.G.B. 1996b, Process of producing lower alkyl fatty acid esters CA 0112581.
Boocock D.G.B. et al. 1998, Journal of the American Oil Chemical Society, 75, 1167.
Boocock D.G.B. 2001, Single phase process for production of fatty acid methyl esters from
mixtures of triglycerides and fatty acids WO 0112581.
Canakci M. et al. 2001, Transactions of the American Society of Agricultural Engineers, 44,
1429.
CANMET Energy Technology Centre (CETC), NRCan, SuperCetane Technology.
Cao et al. 2005, Preparation of biodiesel from soybean oil using supercritical methanol and
co-solvent. Fuel, 84, 347-351.
Chang C.C. et al. 1947, Industrial and Engineering Chemistry, 39, 1543.
Dorado M.P. et al. 2002, Transactions of the American Society of Agricultural Engineers, 45,
525
Farrelli A. et al. 2006, Science, 311, 506.
Freedman B. et al. 1984, Journal of the American Oil Chemical Society, 61, 1638.
Fuduka H. et al. 2001, Journal of Bioscience and Bioengineering, 92, 405.
Goering et al. 1982, Journal of the American Society of Agricultural Engineers, 25, 1472.
Gryglewicz S. 1999, Bioresource Technolology, 70, 249.
HART Ethanol & Biodiesel News 2006, Vol XVIII, Issue 47, November 13.
Harwood H.J. 1984, Journal of the American Oil Chemical Society, 61, 353
Hofmann F. et al. 2005, Evaluierung der Mglichkeiten zur Einspeisung von Biogas in das
Erdgasnetz, Forschungsvorhaben im Auftrag der Fachagentur fr Nachwachsende
Rohstoffe.
Knothe G. 2005, Fuel Processing Technology, 86, 1059.
Kononova M.M. 1961, Soil Organic Matter, Its Nature, Its role in Soil Formation and in Soil
Fertility.


52
Krbitz W. 2002, 8 key trends in the production of Biodiesel world-wide, presentation, Vienna.
Levelton Engineering 2002, Assessment of Biodiesel and Ethanol diesel blends, greenhouse gas
emissions, exhaust emissions and policy issues, Ottawa.
Lee I. et al. 1995, Use of Branched-Chain Esters to Reduce the
Crystallization Temperature of Biodiesel, J. Am. Oil Chem. Soc. 72: 11551160.
Ma F. et al. 1998, Transactions of the American Society of Agricultural Engineers, 41, 1261.
Ma F. 1999a, Bioresource Technology, 70, 1.
Ma F. et al. 1999b, Bioresource Technology, 69, 289.
McCormick R.L. et al. 2003, NO
x
Solutions for Biodiesel, NREL/SR-510-31465, Golden, CO:
National Renewable Energy Laboratory.
Mohamad I.A.W. et al. 2002, Bioresource Technolology, 85, 25.
Natural Resources Canada 2004, Office of Energy Efficiency, The Addition of NRCans
SuperCetance and ROBYS Processes to GHGenius.
NBB (National Biodiesel Board) 2004, Biodiesel Production Technology Overview, Genuary.
NBEP 2007, Biodiesel Education, National Biodiesel Education Program, available at
http://www.uidaho.edu/bioenergy/
Niehaus R.A. et al. 1986, Journal of the American Society of Agricultural Engineers, 29, 683.
Noureddini Hossein et al. 1998, A Continuous Process for the Conversion of Vegetable Oils into
Methyl Esters of Fatty Acids. JAOCS, Vol. 75, No. 12
NRC 2004, The Addition of NRCans SuperCetance and ROBYS Processes to GHGenius,
Natrual Resources Canada, Office of Energy Efficiency, Ontario. Available at
http://www.ghgenius.ca/reports/NRCanSupercetaneUsedOil.pdf
Pryde E.H. 1984, Journal of the American Oil Chemical Society, 61, 1609.
Scholwin F. 2007, Biomethane from Biogas: Expectations from Established vs. new
Technologies, in Lechner (Hrsg.): Waste matters. Integrating Views, Proceedings of 2nd
BOKU Waste Conference 17-19.04.2007, Vienna, ISBN 978-3-7089-0060-5, 185-194.
Schwab A.W. et al. 1988, Journal of the American Oil Chemical Society, 65, 1781.
Schwab A.W. et al. 1987, Fuel, 66.
Sonntag N.O.V. 1979, Reactions of fats and fatty acids. Bailey's industrial oil and fat products,
vol. 1, 4th edition, ed. Swern, D., John Wiley & Sons, New York, 99.
Srivastava A. et al. 2000, Renewable and Sustainable Energy Reviews, 4, 111.
Szybist J.P. et al. 2005, Fuel Processing Technology, 86, 1109.
Thompson J.C. et al. 2006, Applied Engineering in Agriculture, 22, 261.
Tomasevic A.V. et al. 2003, Fuel Processing Technology, 81, 1.


53
Turck R. 2002, Method for producing fatty acid esters of monovalent alkyl alcohols and use
thereof. USP 0156305.
Tyson K.S. 2001, Biodiesel Handling and Use Guidelines, in: NREL - National Renewable
Energy Laboratory, Report No. TP-580-30004 (Golden CO USA).
UOP, A Honeywell Company 2006, Technology & More Newsletter, Winter.
US EPA 2002, A Comprehensive Analysis of Biodiesel Impacts on Exhaust Emissions. Draft
technical report. EPA 420-P-02-001, www.epa.gov/otaq/models/analysis/biodsl/p02001.pdf.
Wang M. et al. 1999, Effects of fuel ethanol use on fuel-cycle energy and greenhouse gas
emissions, Report, Center for Transportation Research Argonne National Laboratory.
Weisz P.B. et al. 1979, Science, 206, 57.
Zhou W. et al. 2003, Journal of the American Oil Chemical Society, 80, 367.
See also: http://en.wikipedia.org/wiki/Biodiesel
http://www.me.iastate.edu/biodiesel/
http://www.liquid-biofuels.com/FinalReport1.html
http://www.energy.gov/news/
http://ethanol.org/howethanol.html
http://www.biodiesel.org






54
3.0 SECOND GENERATION OF BIOFUELS AND BIOREFINERIES

3.1 Pyrolysis and gasification of biomass
The use of biomass as liquid fuels has attracted significant interest on pyrolysis during the past
two decades. The main advantage that pyrolysis offers over gasification is a wide range of
products that can potentially be obtained, ranging from transportation fuel to chemical feedstock.
Considerable amount of research has gone into pyrolysis in the past decade in many countries.
Any form of biomass can be used (over 100 different biomass types have been tested in labs
around the world).

Pyrolysis is a process that generally occurs under temperatures that vary from 400 C up to
650 C in total or partial absence of oxygen. Gases, liquids and solids are generated in
proportions that depend on the parameters considered, that is, the temperature and pressure of the
reactor, the residence time of the solid, liquids and gaseous phases inside the reactor, the time
and the rate of heating of the biomass particles, the reactor internal environment, and initial
conditions of the biomass.

Fast pyrolysis is one of the most recently emerging technologies used to convert biomass
feedstock into maximum bio-oil. The main characteristics of the fast pyrolysis process are: short
heating time for carbonaceous particles and vapours formed within the reactor; high heating rates
and mass-transfer coefficients; and moderate temperature from the heating source. In general,
the residence time for vapours should be lower than 1 minute. Bio-oil is formed from successive
decomposition reactions, isomerisation, cracking (split), and recombination by condensation,
polymerization, depolymerisation and fragmentation of biomass and also has high water content
in its composition.

Tar and bio-oil produced in pyrolysis is a liquid containing more than 200 components, like
acetic acid, methanol, acetic aldehyde, acetone, ethyl acetate, etc. Some components have a
value as raw material in the chemical industry. Resultants produced in biomass pyrolysis mainly
contain gases such as CO
2
, CO, CH
4
, C
2
H
4
, H
2
, etc., with a lower heat value of between 15 and
20 MJ/m
3
, therefore belonging to combustible gases of medium heat value. It is also a gas of
high quality because it doesnt contain any sulphide and nitride, and can be used directly as civil
combustible gas.



55
The gaseous components obtained from biomass pyrolysis techniques can be used to produce
fuels and chemicals using appropriate conversion technologies. To increase the yield of gaseous
products slow pyrolysis technology is used. The pyrolytic gas (bio-syngas) is generally
composed of carbon monoxide, hydrogen, water, methane, nitrogen, and other light organics and
tar (see Table 3.1). This gas can be used as fuel in turbines or boilers or be upgraded to other
biofuels.

Table 3.1: Composition of raw bio-syngas
Component Wood gas (vol%)
Nitrogen 50 - 54
Carbon monoxide 17 - 22
Carbon dioxide 9 - 15
Hydrogen 12 - 20
Methane 2 - 3
Heating value (MJ/Nm
3
) 5 - 5.9

Gasification is a thermochemical technology similar to pyrolysis which affords higher yield and
quality of bio-syngas. Gasification is carried out using high temperatures under limited supply of
oxygen or air, and, eventually, steam. During the gasification reaction, the biomass is heated by
the energy emitted via its partial combustion, whereby it undergoes a combination of drying,
pyrolysis, oxidation, and reduction processes (see scheme 3.1).

3.2 Synthetic biofuels
For an efficient production of synthetic biofuels with regard to economy of scale and biomass
transportation, costs conversion concepts that are in a medium to large-scale are required for the
production of liquid biofuels. This scale is necessary to provide sufficient amount of raw gas for
gas cleaning/conditioning and fuel synthesis as well as to produce this gas via gasification at
economically justifiable costs.

Despite the scale of a gasifier, no gasification system is a priori appropriate for biomass. Among
other criteria, chemical characteristics, physical and mechanical properties of the utilised
biomass are of importance. But, all reactors for biomass gasification are still in an R&D stage up
to now. Furthermore, previous developments on gasification were mostly not focused on syngas
production but rather on the use of product gas for heat and power generation.


56

Wet biomass
Dry biomass (C
x
H
y
O
z
) + H
2
O
Dry biomass
Producer gas (C
n
H
m
+ CO + H
2
)
+ tar (C) + H
2
O
Dry biomass
CO
2
+ H
2
O
Tar
Producer gas + O
2
CO
2
H
2
O
C
n
H
m
Dry biomass
Tar
+ CO + H
2
HEAT
drying
pyrolysis
combustion
reduction

Scheme 3.1: Schematic of chemical and physical chemical processes occurring during the
gasification of biomass and their interconnection

Depending on fuel synthesis where reactors are available specific qualities of bio-syngas at
constant compositions and large amounts have to be achieved (e.g. for the production of
100 to 1,000 m
STP
/h of FT); primarily with regard to the gas purity and the H
2
-to-CO-ratio.
Because so far no gasification system meets these requirements, appropriate gas cleaning and
conditioning system have to be applied.

During gasification, besides the main components (CH
4
, H
2
, CO and CO
2
), impurities such as
tars, coarse and fine particles, sulphur compounds, alkalis, halogen and nitrogen compounds as
well as heavy metals are also generated. Their quantities vary depending on the gasification
process.

For raw gas cleaning either low temperature wet gas cleaning or, alternatively, hot gas cleaning
can be applied. The effectiveness of wet gas cleaning (e.g. cyclone and filter, scrubbing based on
chemical or physical absorption and ZnO-bed) has been well proven for large-scale coal


57
gasification systems. Different to that, not all elements of hot gas cleaning (e.g. tar cracking,
granular beds and filters, physical adsorption or chemical absorption, ZnO-bed, physical
absorption) are of mature technology yet. Nevertheless, hot gas cleaning offers benefits for the
overall energy balance and with regard to the avoidance of contaminated sewage.

For gas conditioning available system components can be applied: hydrocarbons in the product
gas can be converted by means of an additional steam reforming step resulting in a higher H
2
/CO
ratio. To achieve the required quality for fuel synthesis the water-gas CO conversion is
conducted as final step of bio-syngas production.

Based on these and further aspects the following overall concepts for liquid biofuel production
seem to be promising (Scheme 3.2). Thereby the technological performance of FT synthesis can
be carried out in low temperature solid bed reactors or slurry reactors followed by fractioning
and hydrocracking of FT products to diesel. For the production of MeOH synthesis low pressure
synthesis is used in commercial practise by means of solid bed reactors. MeOH is also applied as
feedstock for indirect DME synthesis that is matured. In general, the complexity of an overall
system for production of synthetic biofuels strongly depends on the necessary gas cleaning and
gas conditioning. For a high efficient production the efficiency of bio-syngas production is
crucial.




58
Scheme 3.2: Concepts for liquid synthetic biofuels

With regard to that overall concept, the technologies for liquid synthetic biofuels (BtL) can be
briefly characterised as follows:

Bio-SNG
The bio-SNG (synthetic natural gas) production is characterised through the possibility that
relatively small conversion units with capacities in a range of 10 up to 100 MW
th
biomass input
can be used. Thus, the conversion of locally available lignocellulosic biomass is possible. The
production of synthetic biofuel, electricity and heat (so called tri-generation) allows high overall
efficiencies (e.g. high CO
2
mitigation potential) within the entire production process. In
comparison to the production of BtL-fuels (like Fischer-Tropsch-fuels) is the SNG production
system marked through lower technical and financial risks due to a technology that is less
complex (e.g. synthesis and fuel treatment) and basically smaller units can be used. Based on
these circumstances a rapid and easy market entrance seems to be possible.

The production of SNG can occur within a very promising concept via the steam gasification of
woody biomass with water as gasification agent, gas cleaning, subsequent methanation and up-
grading. One promising concept is shown in Scheme 3.3.


Scheme 3.3: System concept of the Bio-SNG production

Fuel
conditioning
Cooler
Cooler
ORC
Compression
Gas cleaning/
cooling
Gasification
Methanation
(two-stage)
Gas engine
Cleaning
Heat
Steam
Heat
Heat
Biomass
Electricity,
heat
SNG
Flue gas
Heat CO
2



59
For example, steam gasification using a fast internal circulating fluidised bed gasifier (FICFB)
and gas cleaning have been demonstrated successfully in a full technical scale (8 MW thermal
capacity) at the biomass combined heat and power (CHP) gasification plant in Gssing/Austria
whereas during the last years experiences in commercial use for more than 30,000 hours have
been achieved. Steam gasification leads to a producer gas with a relatively high content of
hydrogen and methane as well as a low content of nitrogen. These properties are necessary for an
efficient SNG production.

The gas cleaning of the producer gas from biomass gasification for application in gas engines
and turbines can be considered as state of the art. Additionally for the methanation process acid
components such as H
2
S, HCl and organic sulphur that could damage the catalyst have to be
removed (e.g. a rapeseed-methyl-ester (RME) scrubber). For the conversion of the producer gas
(also called bio-syngas as it is close by composition to the synthesis gas) into biomethane a
fluidized bed methanation reactor will be constructed at the demonstration plant in
Gssing/Austria. Therewith biomethane can be derived from woody biomass with an overall
efficiency of 60 to 65 % depending on the biomass assortment.

Beside the production of SNG as a biofuel for the transport sector, additionally electricity can be
generated in a gas engine as well as through the use of rejected heat in the ORC process. The
arising heat can be provided for district heating purposes as well. Currently the methanation
reactor is within a laboratory scale in operation (thermal capacity 10 kW), fed by a slipstream at
the biomass CHP gasification plant in Gssing/Austria to find out the optimal operation
conditions.

The biomass gasification for example has to be further developed, upscaled to larger scales and
improved to secure the production of a favourable bio-syngas (i.e. optimisation of the
gasification regarding the composition of the gas (primarily regarding the H
2
/CO ratio). The
optimisation and the extension of the gas cleaning and conditioning is a premise for the
production of an appropriate bio-syngas that can fulfil the necessary requirements with regard to
the gas properties. With regard to the arising problem in the field of the methanation catalyst, gas
impurities like sulphur and chlorine compounds have to be removed more intensive whereas the
gas cleaning has to be further developed. The already demonstrated methanation process has to
be scaled up on the MW range. Concerning the utilized catalyst further optimization has to be
performed to ensure a long term performance of the catalyst.



60
Within the overall SNG conversion system the single interactions between the different system
components (like gasification, gas cleaning, methanation and upgrading) have to be optimised.
Additionally, the availability and reliability of the entire facility have to be demonstrated and
improved.

Fischer-Tropsch fuels
In addition to the presented concept, there are many other different concepts to produce FT-fuels
that are characterised by the typical by-products waxes, naphtha and electricity. Basically the
technology is more complex (e.g. when compared to Bio-SNG), whereby mechanical-thermal
biomass treatment (e.g. chipping and drying of solid biofuels) is matured. The pyrolysis and
transport of pyrolysis products (e.g. slurry) is in an R&D stage. Biomass gasification is
demonstrated in small-scale. Basically, the synthesis and fuel treatment of the FT raw products
(refinery technology required) are commercially available for fossil material. The expected
overall efficiency is between 40 to 45%. Particularly in Europe, the FT-technology is on the
minds of many technology developers and mineral oil industries. The first commercial
demonstration plant with a capacity of approximately 15 kt/a FT-diesel (45 MW
th
biomass input)
was built by Choren Industries in Freiberg/Germany and its operation was programmed for the
end of 2007. However, for a broad market implementation, existing techno-economic barriers
have to be overcomed in the years to come. This is expected to happen after 2010. Until then,
several R&D demand is required, e.g. with regard to (i) further development of pyrolysis and
gasification (e.g. upscale, operation under pressure), (ii) efficient and economic gas treatment
technologies, (iii) treatment of FT-raw products (e.g. downsizing hydrocracking, use in
refineries) as well as the successful demonstration of plants availability and reliability for the use
of approved system components.

Biomethanol
Methanol synthesis and fuel treatment are not that complex as FT-fuel production. Thus, higher
overall efficiencies of up to 55% can be expected. Currently, the R&D activities are with regard
to producing an intermediate product for further fuel production (e.g. gasoline). The required
R&D demand is quite similar to FT-fuels for gasification and gas treatment. Moreover, the
adaptation of methanol synthesis units has to be carried out. This is also true for the
demonstration of biomethanol plants in pilot scale.

Dimethylether (DME)


61
DME can be used as a fuel in diesel engines, petrol engines (30% DME / 70% LPG) and gas
turbines. It works particularly well in diesel engines due to its high cetane number, which is
greater than 55 compared to diesel which is 3853. However, DME might potentially be used for
non-transport energy purposes. While the processes of gasification of solid biofuels and the gas
cleaning and conditioning to bio-syngas are similar, significant differences exists with regard to
the catalytic synthesis.

DME production synthesis can either be realised via the previous step of methanol synthesis (so
called indirect synthesis) or directly. The most common types for methanol synthesis and
catalytic dehydration of methanol (DME synthesis) are presented in the following.

A typical reactor type for methanol production is the so called adiabatic quench reactor. The
cooling is done by an internal quench (fast cooling by injection of condensate) with fresh syngas.
The reaction heat is withdrawn at the exit of the reactor, which is at the bottom. The conversion
of methanol to DME where water is chemically separated is done in a DME reactor. Usually,
fixed bed reactors are used for this purpose. First, methanol is heated to a temperature of about
250 C (e.g. by a heat exchanger) before entering the reactor, where an exothermic reaction takes
place at temperatures of 250 to 300 C and DME is formed. Afterwards, the reactor effluents are
cooled down and DME is separated from unconverted methanol and water. The conversion rate
of raw methanol to DME ranges from 86 to 88 %. Raw DME contains some water, dissolved
gases and little amounts of higher ethers, which have to be withdrawn by distillation.

Hydro Thermal Upgrading (HTU) diesel
Another basic approach to convert various types of solid biomass into fuels lies in the principle
to obtain the bio-oil as an intermediate. Bio-oil, also named bio-crude can be obtained via flash
pyrolysis or via hydrothermal treatment/ upgrading - HTU. In the later case, the thermal
pyrolytic process is carried out in an aqueous medium under elevated pressures (sub-critical
water) which improves yield and quality of the product oil, as large amount of O
2
is removed as
CO
2
during the HTU process. The HTU process is therefore advantageous for biomass
containing high water content, since drying of the biomass is not necessary.

The HTU technology uses pressures in the range of 120 180 bars, and temperature range of
300 350C. Under these conditions, the biomass is depolymerised to a hydrophobic liquid (bio-
crude), which separates from the water. Some gases such as CO, H
2
, CO
2
(90 weight %) and


62
methane is produced, along with water and organic compounds. The main product of this
reaction is a liquid (oil) consisting of various kinds of hydrocarbon. The lighter fractions of this
oil can be upgraded to diesel fuel components. HTU diesel is produced by means of a catalytic
process called hydrodeoxygenation (HDO) which is intrinsically similar to the
hydroprocessing technology discussed above for the diesel via hydrocracking of vegetable oils.
Oxygen is removed from the biocrude by treatment with hydrogen at elevated temperatures. The
upgrading of the biocrude to diesel fuel may be done at the same location as the production of
the biocrude or the biocrude may be transported to other locations, followed by local upgrading
to HTU diesel. It can be blended with fossil diesel in any proportion without the necessity of
engine or infrastructure modifications.

Presently, HTU technology is only researched in the Netherlands, where the only HTU pilot
plant is located as well. Research activities focus on the complex chemical properties of the
reactions of the HTU process, the feeding pump to achieve the required pressure, and testing of
several feedstock types. It will probably take a few more years for the process to be tested and
developed sufficiently before a commercial diesel equivalent can be produced

3.3 Cellulosic bioethanol
Since sugar and starchy materials are expensive and are in the human food chain, diverting them
to transport fuel production is not always advisable. Alternatively, lignocellulosic materials,
including waste, represent a promising option to solve the problems related to fuel-food
competition. Examples of cellulosic feedstocks are agricultural residues like stalks, leaves and
husks of food crops; forestry wastes such as sawdust and chips from timber mills, dead trees, and
tree branches; municipal solid wastes, cardboard, paper and house hold garbage products; food
processing and other industrial wastes such as liquor, by-products from paper and pulp
manufacturing industries and also energy crops, fast growing tress and grasses developed just for
this purpose. Cellulosic resources are in general very widespread and abundant. For example,
forests comprise about 80% of the worlds biomass. Being abundant and outside the human food
chain makes cellulosic materials relatively inexpensive feedstocks for ethanol production
(Badger, 2002).

Cellulose molecules consist of long chains of glucose molecules as do starch molecules, but have
a different structural configuration (Scheme 3.4). These structural characteristics plus the
encapsulation by lignin (Scheme 3.5) makes cellulosic materials more difficult to hydrolyze than
starchy materials.


63


Scheme 3.4: The Structure of Cellulose (IBWF)




Scheme3.5: Schematic structural formula of lignin (UOC)


Process description
Ethanol can be manufactured from cellulosic biomass feedstocks using acid hydrolysis,
enzymatic hydrolysis and also by thermo chemical treatment. The most commonly used process


64
is acid hydrolysis with sulphuric acid. Virtually any acid can be used; however, sulphuric acid is
most commonly used since it is usually the least expensive. There are two basic types of acid
processes: dilute acid and concentrated acid, each with variations. Scheme 3.6 illustrates the
different steps involved in the conversion of biomass to bioethanol production.



Scheme 3.6. Schematic representation of production of ethanol from cellulosic biomass

Hydrolysis by dilute acid
Dilute acid processes are conducted under high temperature and pressure, and have reaction
times in the range of seconds or minutes, which facilitates continuous processing. The first
reaction converts the cellulosic materials to sugar and the second reaction converts the sugars to
other chemicals. Unfortunately, the conditions that cause the first reaction to occur also are the
right conditions for the second to occur. Thus, once the cellulosic molecules are broken apart, the
reaction proceeds rapidly to break down the sugars into other products most notably furfural, a
chemical used in the plastics industry. Not only does sugar degradation reduce sugar yield, but
the furfural and other degradation products can be poisonous to the fermentation
microorganisms. The biggest advantage of dilute acid processes is their fast rate of reaction,
which facilitates continuous processing. Their biggest disadvantage is their low sugar yield. For
rapid continuous processes, in order to allow adequate acid penetration, feedstocks must also be
reduced in size so that the maximum particle dimension is in the range of a few millimetres.
Since 5-carbon sugars degrade more rapidly than 6-carbon sugars, one way to decrease sugar
degradation is to have a two-stage process. The first stage is conducted under mild process
conditions to recover the 5-carbon sugars while the second stage is conducted under harsher
conditions to recover the 6-carbon sugars.

Biomass
Cellulose &
hemicellulose
Ethanol Lignin and other by-products
Pretreatment Hydrolysis
F
e
r
m
e
n
t
a
t
i
o
n

&

d
i
s
t
i
l
l
a
t
i
o
n

C
5
&C
6
sugars
S
e
p
a
r
a
t
i
o
n



65
As an example, using a dilute acid process with 1% sulphuric acid in a continuous flow reactor at
a residence time of 0.22 minutes and a temperature of 237C (458F) with pure cellulose
provided a yield over 50% sugars. In this case, 0.9 t (1 ton) of dry wood would yield about 189 l
(50 gallons) of pure ethanol. The combination of acid and high temperature and pressure dictate
special reactor materials, which can make the reactor expensive. Most dilute acid processes are
limited to a sugar recovery efficiency of around 50%. The reason for this is that at least two
reactions are part of this process. Unfortunately, sugar degradation is still a problem and yields
are limited to around 272 l/t (80 gallons of ethanol/ton) of dry wood.

Hydrolysis by concentrated acid
The concentrated acid process uses relatively mild temperatures and the only pressures involved
are usually only those created by pumping materials from vessel to vessel. In the TVA
concentrated acid process, corn stover is mixed with dilute (10%) sulfuric acid, and heated to
100C for 2 to 6 hours in the first (or hemicellulose) hydrolysis reactor. The low temperatures
and pressures minimize the degradation of sugars. To recover the sugars, the hydrolyzed material
in the first reactor is soaked in water and drained several times. The solid residue from the first
stage is then dewatered and soaked in a 30% to 40% concentration of sulfuric acid for 1 to 4 hrs
as a pre-cellulose hydrolysis step. This material is then dewatered and dried with the effect that
the acid concentration in the material is increased to about 70%. After reacting in another vessel
for 1 to 4 hr at 100C, the reactor contents are filtered to remove solids and recover the sugar and
acid. The sugar/acid solution from the second stage is recycled to the first stage to provide the
acid for the first stage hydrolysis. The sugars from the second stage hydrolysis are thus
recovered in the liquid from the first stage hydrolysis. The primary advantage of the concentrated
process is the high sugar recovery efficiency, which can be on the order of over 90% of both
hemicellulose and cellulose sugars. The low temperatures and pressures employed also allow the
use of relatively low cost materials such as fibre glass tanks and piping. Unfortunately, it is a
relatively slow process and cost effective acid recovery systems have been difficult to develop.
Without acid recovery, large quantities of lime must be used to neutralize the acid in the sugar
solution. This neutralization forms large quantities of calcium sulphate, which requires disposal
and creates additional expense. Using some assumed cellulose conversion and fermentation
efficiencies, ethanol yields from glucose can be calculated for corn stover (the above-ground part
of the corn plant less the ears) as shown in Table 3.3. Similarly, ethanol yields from the xylose
can be calculated as shown in Table 3.4.



66
Thus, in this example, the total yield/t of dry stover is about 227 l (60 gallons) of ethanol. These
numbers also show how critical sugar conversion and recovery efficiencies and fermentation
efficiencies are. If one could attain 95% for both efficiencies, then the yield would be
approximately 350 l/t (103 gallons of ethanol/ton).

Enzymatic hydrolysis
Another basic method of hydrolysis is enzymatic hydrolysis. Enzymes are naturally occurring
substances that cause certain chemical reactions to occur. However, for enzymes to work, they
must obtain access to the molecules to be hydrolyzed. For enzymatic processes to be effective,
some kind of pretreatment process is thus needed to break the crystalline structure of the
lignocellulose and remove the lignin to expose the cellulose and hemicellulose molecules.
Depending on the biomass material, either physical or chemical pre-treatment methods may be
used. Physical methods may use high temperature and pressure, milling, radiation, or freezing
all of which require high-energy consumption. The chemical method uses a solvent to break
apart and dissolve the crystalline structure.

Bioethanol production from hydrolysed cellulosic biomass
After the pretreatment of cellulosic biomass using acid, alkali or enzymatic hydrolysis
carbohydrate moieties can be converted into ethanol by fermentation techniques using different
types of enzymes depending on the feedstock of sugar moieties for example, as shown in
Table 3.2. Cellulose and hemicellulose can be converted in to ethanol using cellulases and
hemicelluases respectively.

Table 3.2: Feedstocks and enzymes used for production of bioethanol (cellulosic biomass)
Carbohydrate
feedstock
Main carbohydrate to
be converted
Process utilizing
added enzymes
Required enzyme
Lignocellosic biomass
(e.g. plant residues,
bagasse etc)
Cellulose (-1,4 linked
glucose molecules) and
hemicellulose
Cellulose and
hemicellulose
hydrolysis
Cellulases,
hemicellulases

Tables 3.3 and 3.4 illustrate the production of cellulose, hemicellulose, glucose and xylose from
1000 kg of dry stover. It is interesting to note that 151 l of cellulosic bioethanol can be obtained
from 1000 kg dry stover.




67
Table 3.3: Ethanol yield from glucose
Dry stover One tonne (1000 kg)
Cellulose
Cellulose conversion and recovery efficiency
Ethanol stoichiometric yiels
Glucose fermentation efficiency
Yield from glucose
0.45
0.76
0.51
0.75
131 kg ethanol = 151 l (40 gallons)


Table 3.4: Yield of Ethanol from xylose
Dry stover One tonne (1000 kg)
Hemicellulose content
Hemicellulose conversion and recovery efficiency
Ethanol stoichiometric yield
Xylose fermentation efficiency
Yield from Xylose
0.45
0.76
0.51
0.75
131 kg ethanol = 151 l (40 gallons)

Ethanol-from-cellulose (EFC) holds great potential due to the widespread availability,
abundance, and relatively low cost of cellulosic materials. However, although several EFC
processes are technically feasible, cost-effective processes have been difficult to achieve.

Hybrid and thermochemical processes for ethanol production from cellulosic biomass
There are two ethanol production processes that currently employ thermochemical reactions in
their processes. The first system is actually a hybrid thermochemical and biological system.
Biomass materials are first thermochemically gasified and the resulting synthesis gas (a mixture
of hydrogen and carbon monoxide) is bubbled through the specially designed fermenters
containing the microorganisms capable of converting the synthesis gas into ethanol under
specific conditions.

In the second thermochemical ethanol production process the synthesis gas obtained from
biomass passes through a reactor containing the catalyst, which causes the gas to be converted
into ethanol. Numerous efforts have been made since then to develop commercially viable
thermochemical-to-ethanol processes. Ethanol yields of up to 50% have been obtained using


68
synthesis gas-to-ethanol processes. Some processes that first produce methanol and then use
catalytic shifts to produce ethanol have obtained ethanol yields in the range of 80%.


69
3.4 Bio-hydrogen
It is widely acknowledged that hydrogen is an attractive energy source to replace conventional
fossil fuels, both from the environmental and economic standpoint. It is often cited as a potential
source of unlimited clean power (Hoffman, 2002).

When hydrogen is used as a fuel it generates no pollutants, but produces water which can be
recycled to make more hydrogen. Apart from its use as a clean energy resource, hydrogen can be
used for various other purposes in chemical process industries. It is used as a reactant in
hydrogenation process to produce lower molecular weight compounds. It can also be used to
saturate compounds, crack hydrocarbons or remove sulphur and nitrogen compounds. It is a
good oxygen scavenger and can therefore be used to remove traces of oxygen to prevent
oxidative corrosion. In the manufacturing of ammonia, methanol and syngas, the use of
hydrogen is well known. The future widespread use of hydrogen is likely to be in the
transportation sector, where it will help reduce pollution. Vehicles can be powered with
hydrogen fuel cells, which are three times more efficient than a gasoline-powered engine. As of
today, all these areas of hydrogen utilization are equivalent to 3% of the energy consumption, but
it is expected to grow significantly in the years to come.

The commercially usable hydrogen currently being produced is extracted mostly from natural
gas. Nearly 90% of hydrogen is obtained by steam reformation of naphtha or natural gas.
Gasification of coal and electrolysis of water are the other industrial methods for hydrogen
production.However, these processes are highly energy intensive and not always environment-
friendly. Moreover, the fossil-fuel (mainly petroleum) reserves of the world are depleting at an
alarming rate. So, production of hydrogen by exploiting alternative sources seems imperative in
this perspective.

Biomass, as a product of photosynthesis, is the most versatile non-petroleum renewable resource
that can be utilized for sustainable production of hydrogen (Cole, 1992). Therefore, a cost-
effective energy-production process could be achieved in which agricultural wastes and various
other biomasses are recycled to produce hydrogen economically. Production of hydrogen from
renewable biomass has several advantages compared to that of fossil fuels. A number of
processes are being practised for efficient and economic conversion and utilization of biomass to
hydrogen.

Production technologies


70
There are different process routes of hydrogen-production from biomass (Milne, 2002):
1) thermochemical gasification coupled with water gas shift;
2) fast pyrolysis followed by reforming of carbohydrate fractions of bio-oil;
3) direct solar gasification;
4) miscellaneous novel gasification process;
5) biomass-derived syngas conversion;
6) supercritical conversion of biomass;
7) microbial conversion of biomass.

Among the above mentioned technologies, the processes based on thermal gasification and
pyrolysis of biomass are the most developed and can be envisaged to operate on commercial
basis in the near future.

Scheme 3.7: Pathways from biomass to hydrogen

Thermo-chemical gasification coupled with water gas shift
Gasification coupled with water gas shift is the most widely practised process route for biomass
to hydrogen (Milne, 2002). Thermal, steam and partial oxidation gasification technologies are
under development around the world. Feedstock include agricultural and forest product residues
of hard wood, soft wood and herbaceous species. Thermal gasification is essentially high-rate


71
pyrolysis carried out in the temperature range of 6001000 C in fluidized bed gasifiers. The
reaction is as follows:

Biomass + O
2
CO + H
2
+ CO
2
+ Energy

Other relevant gasifier types are bubbling fluid beds and the high-pressure high-temperature
slurry-fed entrained flow gasifier. However, all these gasifiers need to include significant gas
conditioning along with the removal of tars and inorganic impurities and the subsequent
conversion of CO to H
2
by water gas shift reaction.

CO + H
2
O CO
2
+ H
2


Fast pyrolysis followed by reforming of carbohydrate fraction of bio-oil
Pyrolysis produces a liquid product called bio-oil, which is the basis of several processes for the
development of fuel chemicals and materials. The reaction is endothermic:

Biomass + Energy Bio-oil + Char + Gas

Catalytic steam reforming of bio-oil at 750850 C over a nickel-based catalyst is a two-step
process that includes the shift reaction:

Bio-oil + H
2
O CO + H
2

CO + H
2
O CO
2
+ H
2

The overall stoichiometry gives a maximum yield of 0.172 g H
2
/g bio-oil (11.2% based on
wood).

CH
1.9
O
0.7
+ 1.26H
2
O CO
2
+ 2.21H
2


The first step in pyrolysis is to use heat to dissociate complex molecules into simple units. Next,
reactive vapours which are generated during the first step convert to hydrogen. The Waterloo
fast-pyrolysis process technology carried out at 700 C is used for the steam gasification of pine
sawdust using NiAl catalyst at a molar ratio of 1:2. It has revealed that catalytic reactivation
and high steam to biomass ratios diminish the rate of deactivation (Garca, 2002).



72
Methanol and ethanol can also be produced from biomass by a variety of technologies and used
on board reforming for transportation. Caglar and Demirbas (2001) have used pyrolysis of tea
waste to produce hydrogen, while Abedi et al. (1988) have studied hydrogen and carbon
production from peanut shells.

Direct solar gasification
The feasibility of using solar process heat for the gasification of organic solid wastes and the
production of hydrogen have been examined (Yogev, 1998; Antal, 1974). With a credit for the
wastes used, the economic projections were thought to be surprisingly favourable. Shahbazov
and Usubov (1996) have shown good hydrogen yields from agricultural wastes using a parabolic
mirror reflector. Thermal decomposition samples were studied by the method of derivative
chromatographic analysis. Rustamov et al. (1998) studied the thermo-catalytic reforming of
cellulose and wood pulp using concentrated solar energy. The possibility of obtaining hydrogen
and carbon monoxide with temperatures of 700-750 C on a Pt/Al
2
O
3
catalyst was shown in the
study.

Midilli et al. (2000) present results of the use of a palladium diaphragm to achieve solar assisted
hydrogen separations from the gases generated by pyrolysis of hazelnut shells at 500-700 C. It
was concluded that pure hydrogen gas could be efficiently separated at membrane temperatures
between 180-250 C. Walcher et al. (1996) mentioned a plan to utilize agricultural wastes in a
heliothermic gasifier.

Biomass derived syn-gas conversion
Hydrogen production from gasified biomass by sponge-iron reactor has been reported (Hacker,
1998). The sponge-iron process (or steam-iron process) offers a simple possibility to store the
energy of synthesis gas. A number of recent studies have looked into the classical steam-iron
process for upgrading synthesis gas (mainly CO and H
2
) to pure H
2
for use in fuel cells and other
energy devices. Others have worked on the purification of nitrogen containing reduction gas
from a biomass gasifier using wood and wood waste. The process involves two steps:
(1) cleaning of gas from solid biomass or coal or methane, and
(2) energy storage in sponge-iron.

Some of these studies have investigated woody biomass and commercially available sponge-
iron. The reactions are:



73
Fe
3
O
4
+ 4CO 3Fe + 4CO
2
(coal, biomass or natural gas)
3Fe + 4H
2
O Fe
3
O
4
+4H
2


Supercritical conversion of biomass
Many researchers have investigated the aqueous conversion of whole biomass to hydrogen under
low temperature but supercritical conditions. The earliest report of supercritical gasification of
wood is by Modell (1985). He studied the effect of temperature and concentration on the
gasification of glucose and maple sawdust in water, in the vicinity of its critical state (374 C, 22
MPa). No solid residue or char was produced and hydrogen gas concentration up to 18% (v/v)
was reported. The first report of extensive work on supercritical conversion of biomass-related
organics was given by Manarungson et al. (1990), where glucose at 550 C has been converted
largely into hydrogen and carbon dioxide. This was followed by a study of the uncatalysed
solvolysis of whole biomass and hemicellulose in hot compressed liquid water (Mok, 1992). The
first study showed that complete gasification of glucose can occur at 600C, 34.5 MPa and 30 s
residence time.

Following this work, a flow reactor has been used with newly discovered carbon-base catalysts
to convert water hyacinth, algae, pithed bagasse, liquid extract, glycerol, cellobiose, whole
biomass feedstock and sewage sludge to hydrogen. Wood sawdust, dry sewage sludge or other
particulate biomass can be mixed with a corn-starch gel to form a viscous paste. This paste can
be delivered to a supercritical flow reactor with a pump. Ongoing work indicates that the starch
can be reduced to 3% (w/w) and the particulate biomass increased to 10% (w/w). At the critical
pressure of water (22 MPa), the paste vapourizes without the formation of char. A packed bed of
carbon catalyst, at 650C, causes the tarry vapours to react with water to produce hydrogen,
carbon dioxide, some methane and only a trace of carbon monoxide (Antal, 1997).

Microbial conversion of biomass
Highly concentrated organic waste water is one of the most abundantly available biomass which
can be exploited for microbial conversion into hydrogen. A new and unique process has been
developed when substrates such as carbohydrates are fermented by a consortium of bacteria; they
produce hydrogen and carbon dioxide.

Municipal solid wastes and digested sewage sludge have the potential to produce large amount
of hydrogen by suppressing the production of methane by introducing low voltage electricity into
the sewage sludge. Some authors (Fascetti, 1995) have reported on the photosynthetic hydrogen


74
evolution from municipal solid wastes. Batch-wise and continuous experiments show that the
acidic aqueous stream obtained from such refuse is a good substrate for the growth of R.
sphaeroides RV. The substrate from the acidogenesis of fruit and vegetable market wastes gives
higher hydrogen evolution rates (about threefold) compared to synthetic medium. Mixed culture
of photosynthetic anaerobic bacteria provides a method of utilization of a variety of resources for
hydrogen-production (Miyake, 1990).

Hydrogen production from whey by phototropic bacteria like R. rubrum and R. capsulatus has
been discussed by Venkataraman and Vatsala (1990). Roychowdhury et al. (1988) have also
reported hydrogen generation from fermentative bacteria. Kumar and Das (2000) studied the
suitability of starchbased residues for hydrogen-production. Lactate and lactate-containing waste
water (Zurrer, 1982), cow dung slurry (Vrati, 1983), vegetable starch, sugar-cane juice and whey
(Singh, 1994), bean-product waste water (Liu, 1995), tofu waste water (Zhu, 1999) are among
other liquid biomass which are extensively used for hydrogen production.

Comparative analysis
A comparison of different process routes for hydrogen production on the basis of their relative
merits and demerits is given in Table 3.5.

In all types of gasification, biomass is thermochemically converted to a low or medium-energy
content gas. Air-blown biomass gasification results in approximately 5 MJ/m
3
and oxygen-blown
15 MJ/m
3
of gas. However, all these processes require high reaction temperature. Char (fixed
carbon) and ash are the pyrolysis by-products that are not vaporized.

Table 3.5: Merits and demerits of different processes of biomass conversion to hydrogen
Process Merits Demerits
Thermochemical
gasification
Maximum conversion can be achieved. Significant gas conditioning is
required.
Removal of tars is important.
Fast pyrolysis Produces bio-oil which is the basis of
several processes for development of
fuels, chemicals and materials.
Chances of catalyst deactivation
Solar gasification Good hydrogen yield. Requires effective collector
plates.



75
Supercritical
conversion
Can process sewage sludge, which is
difficult to gasify.
Selection of supercritical
medium.
Microbial
conversion
Waste water can also be treated
simultaneously. Also generates some
useful secondary metabolites.
Selection of suitable
microorganisms.


Some of the unburned char may be combusted to release the heat needed for the endothermic
pyrolysis reactions. For solar gasification, different collector plates (reflectors) like parabolic
mirror reflector or heliostat are required. In supercritical conversion, no solid residue or char is
produced in most of the cases. A wide variety of biomass is nowadays being used to produce
hydrogen using supercritical water. In microbial conversion of biomass, different waste materials
can be employed as substrates. These wastes are also treated simultaneously with production of
hydrogen. Some of the biomass feedstock used for the production of hydrogen is presented in
Table 3.6.

Hydrogen produced from biomass in all these processes mostly contain different gaseous
impurities like O
2
, CO, CO
2
, CH
4
and some amount of moisture. Sometimes, presence of these
gases lowers the heating value of hydrogen, in addition to posing some problems in efficient
burning of fuels. CO
2
acts as a fire extinguisher and it is sparingly soluble in water. Scrubbers
can be used to separate CO
2
. Fifty per cent (w/v) KOH solution is a good CO
2
absorbent.
Monoethnoamine can also be used as a CO
2
absorber. The presence of O
2
in the gas may cause a
fire hazard. Water solubility of O
2
is less compared to that of CO
2
. Alkaline pyrogallol solution
can be used as an absorbent of O
2
. Another important problem is the presence of moisture in the
gas mixture. It must be removed; otherwise the heating value of hydrogen will get reduced. This
can be achieved by passing the mixture either through a dryer or a chilling unit (by condensing
out vapour in the form of water). Nowadays, different membrane separation systems are being
utilized efficiently for gas purification.

Table 3.6: Some biomass feedstock used for hydrogen-production


76
Biomass feedstock Major conversion technology
Almond shell Steam gasification
Pine sawdust Steam reforming
Crumb rubber Supercritical conversion
Rice straw Pyrolysis
Microalgae Gasification
Tea waste Pyrolysis
Peanut shell Pyrolysis
Maple sawdust slurry Supercritical conversion
Starch biomass slurry Supercritical conversion
Composted municipal refuse Supercritical conversion
Kraft lignin Steam gasification
Paper and pulp waste Microbial conversion

Miscellaneous novel gasification processes
Several novel heat sources and chemistries have also been explored for hydrogen from organic
biomass. Safrany (1971) has proposed the use of a thermonuclear device to vapourize waste
organic materials in an underground, large-scale plasma process. In the 1980s, two novel
processes for hydrogen from carbonaceous materials were presented. The production of
hydrogen by the electrolysis of a mixture of coal, lime and water was tested by Thakur (1980). In
1981, an open-cycle two-step process was tested involving the reduction of In
2
O
3
by carbon
(chars) and its reoxidation by water to produce hydrogen (Otsuka, 1981).

In
2
O
3
+ C (or 2C) In
2
O +CO
2
(or 2CO), T>873K
In
2
O + 2H
2
O In
2
O
3
+ 2H
2
, T<673K

A set of biochemical reactions have been proposed to describe decomposition of water into
hydrogen and oxygen using nuclear heat and a carbon cycle (Antal, 1974). Municipal waste is
suggested as a possible source of carbon. Algae can be a by-product. Coughlin and Farooque
(1979) have showed that coals and other forms of solid carbonaceous fossil fuels could be
oxidized to oxides of carbon at the anode of an electrochemical cell and hydrogen produced at
the cathode. Gases produced are discussed as function of coal slurry concentration and electrode
potential.


77
3.5 Biorefineries

The Concept
Although a number of new bioprocesses have been commercialized it is clear that economic and
technical barriers still exist before the full potential of this bio-processing can be realized. One
area gaining considerable momentum is the biorefinery concept. A Biorefinery is a facility that
integrates biomass conversion processes and equipment to produce fuels, power, and chemicals
from biomass. In other words, a biorefinery efficiently separates biomass raw material into
individual components and converts these components into marketable products. A biorefinery
(Scheme 3.8) is similar to todays petroleum refineries, which produce multiple fuels and
products from petroleum.


Scheme 3.8: Overview of general integrated biorefinery process
Adapted from (De Jong, 2006)

A biorefinery, by producing multiple products, can take advantage of the differences in biomass
components and intermediates and maximize the value derived from the biomass feedstock. A
biorefinery might, for example, produce one or several low-volume, but high-value, chemical
products and a low-value, but high-volume liquid transportation fuel, while generating electricity
and process heat for its own use and/or for sale (NREL). The high-value products enhance
profitability, the high-volume fuel helps meet energy needs, and the power production reduces
costs and avoids greenhouse-gas emissions. The production of biofuels in the biorefinery


78
complex will service existing high volume markets, providing economy-of-scale benefits and
large volumes of by-product streams at minimal cost for upgrading to valuable chemicals. A
pertinent example of this is the glycerol by-product produced in biodiesel plants. Glycerol has
high functionality and is a potential platform chemical for conversion into a range of higher
value chemicals. The high volume product streams in a biorefinery need not necessarily be a fuel
but could also be a large volume chemical intermediate such as ethylene or lactic acid.

A key aspect of the biorefinery is the imbalance between commodity chemical needs and
transportation fuels. Using the petroleum industry as an illustrative example, presently, about 85
percent by mass of the aggregated output of petroleum refining consists of transportation fuels
and energy (gasoline, distillate fuel oil, jet fuel, residual fuel oil), with the remaining 15 percent
consisting of over half a dozen product types that add considerably to the profitability of the
overall operation while ensuring that all fractions of crude oil are used (Lynd, 2005). Products
ratios from a biorefinery will be based on the biomass feedstock option and the technological
routes through which the feedstock is processed. By analogy with crude oil, however, every
element of the plant feedstock is expected to be utilized including the low value lignin
components. The motivations for adopting the biorefinery can be economic, social or
environmental as presented in Table 3.7.

Table 3.7: Motivations for biomass refining
Economic
Motivations
Feedstock cost: May provide a means to access markets for bioproducts
whose volume is too high and/or price too low to be accessed using corn as
a raw material.
New markets: Provides potential to create new markets, such as polylactic
acid and 1,3- propanediol.
Tax incentives: Can benefit from tax incentives likely to be offered to
promote such investment.
Societal
Motivations:
Sustainable resource supply: Biomass refining has the potential to
significantly reduce both greenhouse gas emissions and the extent of non-
renewable resource depletion
Energy security: By reducing dependence on foreign oil, large-scale
biomass refining would enhance a nations energy security.
Rural economic development: By creating a large market for energy crops,
could potentially balance demand for agricultural products with current
production capacity.


79
Environmental
Motivations:
Waste management: By utilising close to 100% of the entire biomass raw
material, less waste is generated, which even reduces further, the cost of
managing waste
Source: (Adapted from Lynd, 2005)

Consistent with the interests outlined in Table 3.7, both the private and public sectors are active
today in anticipating and enabling the emergence of a new biomass refining industry.

Biorefinery feedstocks
The products from a biorefinery will be strongly influenced by the feedstock processed by the
facility. Currently most feedstocks used for fermentation processes in the ethanol industry are
based on sugar cane, corn and sugar beet. Sugars are a very abundant renewable resource and
there are many ways of transforming sugars into bioproducts. In the United States, much
attention has been paid to corn-based biorefinery, and the products have been mainly starch and
carbohydrate derivatives from it, but also smaller amounts of oil, protein, and fibre.

However biorefineries require a large and constant supply of biomass and other substrates would
have to be used if costs are going to be reduced. Such biomass could include grains such as
wheat and barley, oils, agricultural residues, waste fruits, straw, waste wood and forest
trimmings and dedicated energy crops such as switchgrass or hybrid poplar (Lasure, 2004). The
most significant impact of biorefineries on energy, greenhouse gases and climate change is likely
to come from the conversion of lignocellulosic materials to ethanol (Lynd, 2005). Due to the
huge volumes of corn grown across the globe, corn stover is and will continue to be the leading
candidate of biomass source to support a lignocellulosic biorefinery.

Vegetable oil-based products are also another excellent feedstock for biorefineries because of
their versatility. They can be used for the production of several products, such as solvents,
lubricants and hydraulic fluids, polymers, resins, printing inks, cosmetics, pharmaceuticals and
numerous other applications (Zwart, 2006). It is expected that feedstock properties will influence
design of thermochemical biorefineries in the future. Scheme 3.9 shows connections between
different types of feedstocks and different approaches of product utilization.



80

Scheme 3.9: Biorefinery: feedstocks to products evolution


Biorefinery types according to feedstock and technologies
Biorefineries can be classified based on the biomass feedstock to be processed or the
technological platform. Based on the biomass feedstock, biorefineries can be classified as:
Whole crop biorefineries
Green biorefineries
BIOMASS
Starch, lignin, hemicellulose,
cellulose, protein, oils
Glucose, xylose,
arabinose, fructose,
lactose, sucrose, starch
Biobased
syngas
Hydrogen, methanol,
mixed higher alcohols,
oxo-synthesis products,
iso- synthesis products,
Fischer-Tropsch liquids,
etc.
C3: glycerol, lactic acid, 3-hydroxy
propionate, propionic acid, malonic
acid, serine

C4: succinic acid, fumaric acid, malic
acid, aspartic acid, 3- hydroxy
butyrolactone, acetoin, threonine

C5: itaconic acid, furfural, levulinic
acid, glutamic acid, xylonic acid,
xylitol/arabitol

C6: citric/ aconitic acid, 5-
hydroxymethyl-furfural, lysine, gluconic
acid, glucaric acid, sorbital

Aromatics: gallic acid, ferulic acid &
direct polymers & gums
Polymers, fuels, lubricants, water chemicals, fertilizers,
paints, resins, pesticides, pharmaceuticals, cosmetics, etc.
Biomass components:
Platform intermediates:
Building blocks:
End products:
Feedstock:


81
Lignocellulosic biorefineries

Whole crop biorefinery. The general concept of whole crop biorefinery is to take advantage of as
much of the plant as possible and process it into products that deliver the highest possible value.
This is a step forward from plants that make use of just the grains and leave the stalks as waste
materials. The whole crop biorefinery aims to process about 100% of the biomass feedstock. As
shown in Scheme 3.10 for whole crop cereals, the feedstock is transported from the field and
milled (or pressed, in the case of oil seeds) to separate economically recoverable plant
components, which might include sugar, starch, oil, protein, and fibre. Oil is recovered from the
germ (in the case of wet milling) while the fibre remain mixed with the starch. The starch-rich
mash is then fermented to produce ethanol. The unfermented constituents of the starch-mash
from the grind operation known as distillers dried grains and solubles (DDGS) is also separated
and dried. The DDGS is gasified to CO and H
2
(syngas), which is cleaned and catalytically
converted to alcohols or hydrocarbon-based fuels (Fischer Tropsch liquids) (Brown, 2005). The
straw can also be treated to produce fuels and other chemicals, or gasified to produce similar
products as those realised from the DDGS. Depending on the feedstock and technology route,
there can be different types of biorefineries in this category.

The whole crop biorefinery concept is already being used to some extent in some industries such
as the sugar industry, starch industry and food and feed industry, where fermentation of sugars is
a key process. However, most of these plants make use of just the grains without deriving much
value from the cellulosic parts of the crop. There are opportunities to derive higher values from
some of the by-products than occurs in the present plants and to make additional use of the
whole crop biomass instead of just the grains.



82

Scheme 3.10: Whole crop biorefinery based on dry milling of cereals.
Adapted from (Kamm, 2006)

Green Biorefinery. The Green Biorefinery is a based on the integrated processing of green
biomass into multiple products. It uses fresh biomass crops, for example grass from cultivation
of permanent grass land, closure fields, nature reserves; lucerne, clover, and immature cereals
from extensive land cultivation. This biorefinery is especially suitable for places with excess
grassland. The concept is based on pressurisation of the wet biomass, resulting in a fibre-rich
press cake and a nutrient-rich press juice (Annevelink, 2006). Careful wet fractionation
technology is used as a first step (primary refinery) to isolate the contents of the green crop (or
humid organic waste goods) in their natural form (Kamm, 2006). The press cake contains
cellulose, starch, dyes, pigments, crude drugs, etc. while the green juice contains proteins,
enzymes, free amino acids and organic acids. The green juice is used for the production of such
items as ethanol, lactic acid, proteins and amino acids. The cake is used for the production of
feed pellets, other important components for the chemical industry and for conversion to syngas
and synthetic biofuels. The press cake is the largest material stream encountered in a green
biorefinery and as such the overall economic efficiency of a green biorefinery is strongly
determined by the economic efficiency of converting this fibre fraction into marketable value
added fibre products. The residues of substantial conversion are suitable for production of biogas
combined with generation of heat and electricity. Products generated in a green biorefinery
(Energytech) may be grouped as:
1. Bulk chemicals (e.g. organic acids like lactic acid, solvents like ethyl-lactate).
2. Fuels (e.g. ethanol, acetone, butanol, ester).


83
3. Food / feed (e.g. amino-acids, protein products, peptides).
4. Fibre products (e.g. fibreboards, biocomposites, insulation material).
5. Fine chemicals (e.g. flavours, chlorophyll, pigments).
6. Biogas (electricity + heat)

The advantages of the Green Biorefinery are a high biomass profit per hectare and a good link
with the agricultural production; whereas the price segment of the raw materials is still low.
Green biorefineries can process from a few tonnes of green crops per hour (farm scale process)
to more than 100 tonnes per hour (industrial scale commercial process).


Scheme 3.11: Green biorefinery system. Adapted from (Kamm, 2005)

Lignocellulosic feedstock biorefinery. Lignocellulosic feedstocks can include straw, switchgrass,
hybrid poplar, cornstover, reed, wood, paper and municipal waste and consist of three primary
chemical fractions or precursors:
hemicellulose/polyoses, sugar polymers of, predominantly, pentoses;
cellulose, a glucose polymer; and
lignin, a polymer of phenols

The lignin, hemicellulose and cellulose components are first separated and then processed. As
illustrated in Scheme 3.12, the plant material is first pretreated (some pretreatments include
dilute acid, hot water, steam explosion, and ammonia explosion) to increase the surface area of
lignocellulose, making the polysaccharides more susceptible to hydrolysis. The product streams
usually include cellulose, hexose and pentose from hydrolyzed hemicellulose, and lignin.


84

E-37
Lignocellulose
Cellulose
Lignin
Hemicellulose
Fuels, chemicals,
polymers,
materials
Sugar, raw material
Cogeneration to
produce heat and
power
Lignin raw material

Scheme 3.12: Lignocellulosic feedstock biorefinery. Adapted from (Kamm, 2005)

The cellulose can be saccharified after which the sugars are fermented to ethanol or other
fermentation products. The hemicellulose can be hydrolysed and the resulting sugars converted
into alcohol and value added chemicals via fermentation and chemical catalysis. Lignin which
is the non-carbohydrate constituent of fibre can be thermochemically converted to syngas
followed by catalytic conversion to alcohols, fuels, chemicals and other materials. The choice of
the technology applied in converting lignocellulosic feedstock determines what the final products
would be.

For any of the aforementioned biomass feedstocks and biorefinery types, one or a combination of
technologies is usually possible. To generally describe the technological approach used to treat
biomass or some of its components, a classification according to the technological pathway is
usually adopted. Such general technological pathways include biochemical, thermochemical, and
hybrid. In the former two the respective technologies (of biochemical or thermochemical nature)
constitute the primary method of converting biomass components into intermediates and/or final
products (in some cases certain intermediates can be converted to final products using a different
nature of technology). In the case of hybrid technology, the primary treatment step is
thermochemical followed by the biochemical treatment of intermediates. Listed below there are
several examples of the technological routes appertaining to these pathways:

Biochemical
Fermentation of sugars to alcohols and other products


85
Enzymatic hydrolysis of starch or cellulose to sugars (sugars can be then converted
to alcohols/chemicals via either biotechnological or chemical catalytic approaches,
and the residuals can be also treated via anaerobic digestion or gasification)
Anaerobic digestion of humid biomass to biogas (can be successively converted to
methane via chemical catalytic technology)
Thermochemical
Gasification of biomass to syngas and successive catalytic syntheses
Fast pyrolysis of biomass to bio-oil and successive extraction of chemicals or
chemical catalytic synthesis/upgrading based on bio-oil components
Hybrid thermochemical/biochemical
Production of syngas via gasification and its successive fermentation
Production of bio-oil via pyrolysis and its successive fermentation
Other chemical
Acid hydrolysis of cellulose to sugars
Transesterification of vegetable oils and esterification of fatty acids
Fine organic synthesis of drugs and speciality chemicals based on various plant
substances

One of the main technological challenges of modern biorefineries lies in the possibility to
efficiently treat the lignocellulosic feedstock. Therefore, cellulose hydrolysis to sugars
(biochemical) or its gasification to syngas (thermochemical) are usually referred to as the two
basic biorefinery approaches called sugar platform and syngas platform. However, apart
from the syngas and sugar platforms for cellulose conversion and other biochemical and hybrid
pathways for biomass treatment (e.g. biogas and pyrolysis based) a biorefinery can be, in
principle, built around other chemical technologies, like classical acid hydrolysis of cellulose,
transesterification of oils, etc.

Biochemical pathway. A classical example of this pathway is the enzymatic hydrolysis of starch
or cellulose with subsequent fermentation of sugars to produce ethanol and other products. As
shown in Scheme 3.13, fibrous feedstock must first be pre-treated (usually with strong acid) to
break down the hemi-cellulose fraction and make the remaining cellulose material more
accessible for subsequent saccharification. Cellulose enzymes are then introduced to hydrolyze
the carbohydrate material, providing a variety of sugars which are fermented to produce ethanol
and other fermentable products (the process technology is adequately addressed in section 3.3 of


86
this book). The lignin portion of the original biomass feedstock is generally unreacted
throughout this process, and is recovered to be used as a fuel or feedstock for thermochemical
conversion processes. Major technical challenges and barriers to the biochemical pathway
include dealing with the variability of biomass feedstocks, general recalcitrance of
lignocellulosic materials to biological degradation, and the need for improved effectiveness of
cellulose enzymes and fermentation organisms (Hoekman, 2008).


Scheme 3.13: Biochemical biorefinery based on lignocellulosic feedstock.
Adapted from (Brown, 2006)

Biogas technology is based on the biochemical process of anaerobic digestion of certain biomass
components (e.g. proteins, sugars, oleochemicals, etc.) and can be an important component of
certain biorefineries e.g. to process humid low value residuals coming from key product
extraction processes (green and whole crop biorefineries). The biogas can be used on site for
energy generation or be upgraded to bio-methane (see session 2.10).

In addition to the sugar platform biorefinery and energy generation via biogas, some other
biochemical processes can be used in biorefinery for chemical production. For example, proteins
contained in plants and in bio-waste can be treated by bacteria to obtain biodegradable plastics
(polyhydroxyalkanoates).

Thermochemical pathway. Thermochemical pathways generally utilize either gasification or
pyrolysis technologies as the primary processing step which yield sin-gas or bio-oil, respectively,
as the intermediates for further syntheses. The gasification technology biorefinery is shown in
Scheme 3.14. It involves high temperature thermal decomposition of the biomass, followed by
partial oxidation to produce raw synthesis gas (consisting mainly of CO and H
2
). Following
cleanup and conditioning, syngas can be reacted catalytically to produce mixed alcohols,
FischerTropsch hydrocarbons and other products depending on the catalytic technology chosen.
CO
2
Pretreatment
D
i
s
t
i
l
l
a
t
i
o
n

Saccharification
C5 & C6
Sugars
Fibrous
crop
Lignin to thermochemical conversion
Cellulose enzymes
Fermentation
Ethanol &
other
products


87
A range of syngas based technologies are well developed and applied in the petrochemical
refineries.

Pyrolysis also involves thermal decomposition of biomass, but it is done at somewhat lower
temperatures, and in the absence of oxygen, to produce a bio-oil and gases. The bio-oil contains
a range of valuable chemical components that can be extracted and used as speciality chemicals,
or be converted into other valuable products, e.g. resins, fuels, etc. For example, bio-oil can
undergo steam reforming to produce hydrogen or hydrogenation to hydrocarbon fuels. Another
possibility is to convert bio-oil to syngas via its gasification to produce better quality syngas than
that from the direct gasification of bio-mass. An example of the bio-oil based biorefinery is
given in Scheme 3.15.

Scheme 3.14: Thermochemical biorefinery using gasification and catalytic synthesis.
Adapted from (Brown, 2006)

Scheme 3.15: Thermochemical biorefinery using pyrolysis and catalytic synthesis.
Adapted from (Brown, 2006)
p
y
r
o
l
y
z
e
r

Cyclone
Bio-oil
recovery
Phase
separation
Steam
reformer
Biomass
Bio-oil vapour
Carbohydrate derived
aqueous phase
Char
H
y
d
r
o
c
r
a
c
k
e
r

Green diesel
Gas cleaning Catalytic reactor
G
a
s
i
f
i
e
r

Chemicals
Air
Syngas
CO
2
Biomass
Heat


88

Gasification and pyrolysis processes do not require enzymes or microorganisms as do
biochemical processes, they are applicable over a wide range of feedstocks, and they are
generally compatible with conventional petroleum processing technologies (Hoekman, 2008).
However, the crude producer gas and bio-oil coming from the primary gasification or pyrolysis
of biomass are low purity and cannot be fed into the known catalytic processes of classical
petroleum refinery without prior purification or upgrade. The biochemical technology
complementing the primary thermochemical treatment of biomass (hybrid pathway) can be,
therefore, a promising opportunity to valorise the raw quality thermochemical treatment
intermediates.

Hybrid thermochemical/biochemical. Syngas fermentation and bio-oil fermentation are the two
major approaches of the hybrid pathway. In the syngas fermentation process (shown in Scheme
3.16), the intermediate products (CO, CO
2
and H
2
) are fermented into metabolic products such as
alcohols, carboxylic acids and esters using autotrophic organisms. In the bio-oil fermentation
process, the anhydro-sugar is first separated from other bio-oil components and is then
hydrolysed to produce glucose. The glucose is subsequently fermented to yield ethanol as shown
in Scheme 3.17 (Brown, 2006).

Scheme 3.16: Syngas fermentation. Adapted from (Brown, 2006)



Scheme 3.17: Bio-oil fermentation. Adapted from (Brown, 2006)

p
y
r
o
l
y
z
e
r

Cyclone
Bio-oil
recovery
Phase
separation
Detoxification
Biomass
Fermenter
Ethanol
Lignin
Bio-oil vapour
Anhydrosugar &
other carbohydrate
Char
CO
2
Gasifier Gas cleaning Biomass
Bioreactor
Fuels and
chemicals
Syngas


89

Current biorefineries and outlook/development trends
The development of biorefineries has begun in some countries across Europe and the United
States. It has already proven successful in the U.S. agricultural and forest products industries,
where such facilities now produce food, feed, fibre, or chemicals, as well as heat and electricity
to run plant operations.

Multi-product corn wet mills appeared in their current form in the 1970s, prompted by the
development of commercial technology for production of high fructose corn syrup (HFCS),
which today has largely replaced sugar produced from cane in the US. Sweeteners including
HFCS, glucose, and dextrose today account for about 37 % of the output of the US corn wet
milling industry, with gluten feed and gluten meal accounting for about an additional 32 %, and
the balance consisting of starch, ethanol, carbon dioxide and corn oil (Lynd, 2005). These
processes are similar to the whole crop technology except that the DDGS in most cases is used as
feed instead of conversion into other high valued products.

According to a report by Zwart (2006), Archer Daniels Midland (ADM) has a prototype
expanded biorefinery in Decatur, US, where a large corn wet-milling plant produces industrial
enzymes, lactic and citric acids, amino acids and ethanol. Enzymes are used to convert starch to
maltodextrins and syrups, and the chemical products are used in foods, detergents and plastics.
The ethanol is used as a solvent or for transportation fuels.

Arkenol (an energy company) is marketing a biorefinery technology based on acid hydrolysis
that can produce a variety of biobased chemicals and transportation fuels. The company is
developing several biorefineries throughout the world, including a facility that will produce 90
million pounds/yr of citric acid along with 4 million gallons of fuel ethanol.

Agrologistiek BV in the Netherlands has plans to install several biorefinery pilot plants for
small-scale bioethanol production from arable crops. The idea is to realize multiple smaller
plants close to the farms. This will lower the transport costs, and create a zero emissions system.
The plant is being designed to be flexible in its use of feedstock so that several feedstock can be
processed. Some of the feedstocks will include wheat, beet, leafs, fibres, maize and grass. By-
products of the bioethanol will serve as feed for animals, biogas for energy generation and CO
2
.
The capacity of the plants will vary between 5 and 50 million litres (Annevelink, 2006).



90
In 2000 a pilot plant was built in Foxhol (Groningen) by a consortium consisting of Avebe, Plant
Research International, Nedalco, ABCTA, Agrifirm, NOM and Rabo; to study the technical and
economic feasibility of the green biorefinery concept. The aim was to make use of grass surplus
in the Netherlands. The main products were grass juice concentrate, grass protein and grass
fibres. Other products that could be manufactured based on these grass fibres are potting soil,
construction materials and filler materials for polymer extrusion products. The grass juice
concentrate can be used as feed component for pigs or to produce biofuels. Finally the grass
protein can be used as feed component for pigs and poultry (Annevelink, 2006).

A demonstration Green Biorefinery plant in Brandenburg Germany will use 24,000 tonnes per
annum fresh alfalfa, lucerne and wild grass. It will use mechanical separation (pressing) and
fermentation technology. The main products are press cake for fodder and fuels, proteins for
industry and cosmetics and heat and electricity. The primary biorefinery is combined with a
green crop drying plant.

Several other projects are being developed mainly in Europe and the United States and should be
commercialised in the next few years.

Most modern wet corn milling plants and pulp and paper mills claim to be biorefineries (and
most US ethanol plants are referred to as biorefineries) because of the variety of products at the
end of the production process. The food industry also try to add some value by supplying their
by-products to other sectors, e.g. to the feed industry. However, in these industries, the main
emphasis is still on producing their main products, and no large efforts are made yet to produce a
broad spectrum of other added-value products, like bio-chemicals or biofuels.

Over the coming decade, researchers anticipate a new category of refining processes based on
cellulosic biomass. There are some challenges, however, that have to be dealt with for
biorefining of cellulosic biomass to become fully sustainable. Major technical challenges and
barriers to the biochemical pathway include dealing with the variability of biomass feedstocks,
the difficulty of lignocellulosic materials to chemical and/or biological degradation, and the need
for improved effectiveness of cellulose enzymes and fermentation organisms. Important
technical challenges still remain with thermochemical approaches. Pre-treatment of the biomass
and physical feeding into thermal processing units are challenging. For gasification, other
challenges include minimization of tar formation, syngas cleanup, and development of effective


91
catalysts. For pyrolysis, major challenges include cleanup of the bio-oil and sufficient
stabilization of it for practical delivery and use in a petroleum refinery (Hoekman, 2008).

Future biorefineries needs intensive research in different areas such as agronomy of feedstocks,
greener catalytic process, genetically modified crops and encouraging government policies on
promotion of biofuels and co-product valorisation. Future biorefinery operations will aim to
extract high-value chemicals already present in the biomass, such as fragrances, flavoring agents,
food-related products, and high-value nutraceuticals that provide health and medical benefits
(Ragauskas, 2006). The whole biorefinery concept requires more extensive research in the
chemistry and technological applications and processes to reduce the cost of operations.




92

3.6 Bibliography

Annevelink E. et al. 2006, Official minutes of the First workshop on the possibilities of
biorefinery concepts for the industry. Held at hotel De Wageningse Berg, Wageningen,
The Netherland, 16 June 2006, Wageningen University and Research Centre (WUR) and
the Energy research Centre of the Netherlands (ECN)
Abedi J. et al. 1988, Production of primary pyrolysis oils in a vortex reactor, in Pyrolysis Oils
from Biomass: Producing, Analysing & Upgrading (Soltes E.J. and Milne T.A., Eds.),
American Chemical Society, Washington DC, 31.
Aden A. et al. 2002, Lignocellulosic Biomass to Ethanol Process Design and Economics
Utilizing Co-Current Dilute Acid Prehydrolysis and Enzymatic Hydrolysis for Corn
Stover, National Renewable Energy Lab, Golden, CO, NREL Report No. TP-510-32438.
Anon 2002, NCGA Retaliates Against Ethanol Attacks, National Corn Growers Association
News, 9, May 24, http://www.ncga.com/news/CC/volume9/ccVol9n19.html.
Antal Jr. et al. 1997, Total catalytic supercritical steam reforming of biomass, Proceedings of the
1997 US DOE Hydrogen Programme Review, Herndon, Virginia.
Antal M.J. et al. 1974, Hydrogen and food production from nuclear heat and municipal waste, in
Hydrogen Energy (Part A) (Veziroglu T. N., Ed.), Plenum, NY, 331.
Arbige M.V. 2004, Bioprocess converts glycerol to propanediol. Industrial Bioprocessing
26(11), 3.
Badger P.C. 2002, Ethanol from cellulose: A general review. In: Trends in new crops and new
uses (J. Janick and A. Whipkey, Eds), ASHS Press, Alexandria, USA.
Boerrigter H. et al. 2002, Green Diesel from Biomass via Fischer-Tropsch Synthesis, Pyrolysis
and Gasification of Biomass and Waste, Expert Meeting, Strasbourg, France, Sept 30 -
Oct 1.
Boerrigter H. et al. 2004, Thermal Bio-Refinery; High-Efficient Integrated Production of
Renewable Chemicals, (Transportation) Fuels, and Products from Biomass. Presented at
The 2nd World Conference and Technology Exhibition on Biomass for Energy,
Industry and Climate Protection in Rome, Italy, 10-14 May 2004
Brown R.C. 2006, The Role of Thermochemical Technologies in Advanced Biorefineries, North
Central Region Feedstock Workshop, Sioux Falls, USA, August 15, 2006, available at
http://ncsungrant.sdstate.org/files/feedstock/Thermochemical%20Platform%20Overview.
pdf.


93
Brown R.C. 2003, Biorenewable Resources: Engineering New Products from Agriculture, Iowa
State Press, Iowa, USA.
Brown R.C. 2006, Biomass Refineries based on Hybrid thermochemical/Biological Processing
An Overview, in Biorefineries, Biobased Industrial Processes and Products (Kamm B.,
Gruber P.R., and Kamm M., Eds.), WILEY-VCH Verlag GmbH & Co. KGaA.
Caglar A. et al. 2001, Energy Sources, 23, 739.
Cole H. et al. 1992, Energy World, 199, 15.
Coughlin R.W. et al. 1979, Nature, 279, 301.
De Jong E. et al. 2006, Biorefineries for the chemical industry-A Dutch point of view, in
Biorefineries-Industrial Processes and Products (Kamm B., Gruber P.R., Kamm M.,
Eds.), WILEY-VCH Verlag GmbH & Co. KGaA.
Energytech, Green Biorefinery-Primary Processing and Utilization of Fibres from Green
Biomass. Available from http://www.energytech.at/
Ekbom T. et al. 2003, Technical and Commercial Feasibility Study of Black Liquor Gasification
with Methanol/DME Production as Motor Fuels for Automotive Uses BLGMF.
Nykomb Synergetics AB, Chemrec, Volvo, Ecotraffic, OKQ8, STFi, Methanex, Final
Report, Altener Programme.
Ekbom T. et al. 2005, Black liquor gasification with motor fuel production BLGMF II, A
techno-economic feasibility study on catalytic Fischer-Tropsch synthesis for synthetic
diesel production in comparison with methanol and DME as transport fuels. Nykomb
Synergetics AB, STFi-Packforsk, KTH Royal Institute of Technology, Statoil, Structor
Hulthn Strth, Final Report.
Fascetti E. et al. 1995, Applied Microbiology and Biotechnology, 44, p.300.
Feng W. et al. 2004, Phase equilibria for biomass conversion processes in subcritical and
supercritical water, Chemical Engineering Journal, 98, 105-113.
Garca L. et al. 2002, Energy Fuels, 16, 1222.
Hacker V. et al. 1998, Journal of Power Sources, 71, 226.
Hacker V. et al. 2000, Journal of Power Sources, 86, 531.
Hahn-Hagerdal B. et al. 2006, Bio-ethanol--the fuel of tomorrow from the residues of today.
Trends Biotechnol, 24(12), 549-56.
Hoekman S.K. 2008, Biofuels in the U.S. Challenges and Opportunities, Renewable Energy
2008, in press.
Hoffmann P. 2002, Tomorrows energy: Hydrogen, fuel cells and the prospect for cleaner planet,
The MIT Press, Massachusetts, USA


94
IBWF, Environmental Biotechnology and Enzymes, available at
http://www.ibwf.de/env&enz_index.htm#index
Kamm B. et al. 2004, Biorefinery - systems. Chemical and Biochemical Engineering Quarterly,
18(1), 1-6.
Kamm B. et al. 2006, Biorefineries - Biobased Industrial Processes and Products (Kamm B.,
Gruber P.R., and Kamm M., Eds.), WILEY-VCH Verlag GmbH & Co. KGaA.
Kumar N. et al. 2000, Bioprocess Eng., 23, 205.
Lasure L.L. et al. 2004, Bioconversion and Biorefineries of the Future, in Applications of
biotechnology to mitigation of greenhouse warming: Proceedings of the St. Michaels II
Workshop, April 2003, Rosenberg N.J., Metting F.B., Izaurralde R.C. Eds., Battelle Press
2004.
Liu S.J. et al. 1995, Environnemental Science, 16, 42.
Lynd L.R. et al. 2005, Strategic Biorefinery Analysis: Analysis of Biorefineries, Subcontract
Report for the National Renewable Energy Laboratory, NREL/SR-510-35578, 2005.
Mabee W.E. et al. 2003, Ethanol from Lignocellulosics: Views to Implementation, Report to
IEA Bioenergy Task 39, Vancouver, Canada.
Manarungson S. et al. 1990, Advances in hydrogen energy. Hydrogen Energy Progress VIII,
International Association for Hydrogen Energy, 1, 345.
Midilli A. et al. 2000, International Journal of Hydrogen Energy, 25, 723.
Milne T.A. et al. 2002, Hydrogen from Biomass. State of the Art and Research Challenges, A
Report for the International Energy Agency Agreement on the Production and Utilization
of Hydrogen Task 16, Hydrogen from Carbon-Containing Materials
Mitchell D. 2005, Whats Online, The New York Times, July 23, Section C, Column, 1, 5,
http://www.nytimes.com.
Miyake J. et al. 1990, Hydrogen energy progress VIII, Proceedings of the 8th WHEC, Hawaii,
Pergamon Press, New York, 755.
Modell M. 1985, Gasification and liquefaction of forest products in supercritical water, in
Fundamentals of Thermochemical Biomass Conversion (Overend, R. P., Milne, T. A. and
Mudge, L. R. Eds.), Elsevier Applied Science.
Mok W.S.L. et al. 1992, Industrial & Engineering Chemistry Research, 31, 1157.
Mller-Langer F. et al. 2005, Synthetic transportation fuels based on biomass liquid versus
gaseous fuel production: an assessment of technical, economic and environmental
aspects, 14th European Biomass Conference & Exhibition Biomass for Energy, Industry
and Climate Protection, Paris, October.


95
NREL: What is a Biorefinery? National Renewable Energy Laboratory, available from
http://www.nrel.gov/biomass/biorefinery.html
Otsuka K. et al. 1981, Chemistry Letters, 23, 347.
Ragauskas A.J. et al. 2006, The Path Forward for Biofuels and Biomaterials, Science 27, 2006
Rifkin J. 2002, The hydrogen economy. Penguin Group Inc., New York, USA
Robinson J. et al. 2003, Enzyme Microb. Technol. 33, 757765.
Roychowdhury S. et al. 1998, International Journal of Hydrogen Energy, 13, 407.
Rustamov V.R. et al. 1998, International Journal of Hydrogen Energy, 23, 649.
Safrany D.R. 1971, Chemical Engineering Progress Symposium Series, 67, p.103.
Sanders J. et al. 2007, Biorefinery, the bridge between agriculture and chemistry, Presented at
the Industrial Show for Research and Development, Environmental and Process
Technology, Basel 28 September 2007
Seiffert M. 2007, Demonstration of the production and utilization of synthetic natural gas (SNG)
from solid biofuels presentation of the European project Bio-SNG, 15th European
Biomass Conference & Exhibition - From Research to Market Development, Berlin,
May.
Shahbazov S. et al. 1996, Heliothermic Unit for Gasification Agricultural Product Wastes,
Hydrogen Energy Progress XI, Proceedings of the 11th World Hydrogen Energy
Conference, 1, 951.
Singh S.P. et al. 1994, International Journal of Hydrogen Energy, 19, 37.
Spath P.L. et al. 2003, Preliminary Screening - Technical and Economic Assessment of
Synthesis Gas to Fuels and Chemicals with Emphasis on the Potential for Biomass-
Derived Syngas, National Renewable Energy Laboratory Report NREL/TP-510- 34929.
Sun Y. et al. 2002, Hydrolysis of lignocellulosic materials for ethanol production: a review.
Bioresource Technol., 83:1, 11.
Thakur L. 1980, Advances in hydrogen energy. Hydrogen Energy Progress, International
Association for Hydrogen Energy, 3, 1725.
Thuijl E.V. et al. 2003, An overview of biofuel technologies, markets and policies in Europe,
ECN-C-03-008, Amsterdam.
UOC, The Structure of Wood (1), University of Cambridge Department of Material Science and
Metallurgy, available from from
http://www.doitpoms.ac.uk/tlplib/wood/structure_wood_pt1.php
Van Gerpen J. 2005, Biodiesel Processing and Production, Fuel Processing Technology, 86,
1097-1107.


96
Van Thuijl E. et al. 2001, Biofuel production chains, Background document for modelling the
EU biofuel market using the BIOTRANS model, European commission supported
project: Clear Views on Clean Fuels, NNE5-2001-00619.
Venkataraman C. et al. 1990, Hydrogen Energy Progress VIII, 2, 781.
Vrati S. et al. 1983, Journal of Fermentation Technology, 61, 157.
Walcher G. et al. 1996, Hydrogen Project Bad Brckenau, Hydrogen Energy Progress XI,
Proceedings of the 11th World Hydrogen Energy Conference, Stuttgart, Germany, 1, 413.
Yogev A. et al. 1998, International Journal of Hydrogen Energy, 23, 239.
Zhong Ch. et al. 2002, Thermodynamic modelling of biomass conversion processes, Fluid Phase
Equilibria, 194-197, 805-815
Zhu H. et al. 1999, International Journal of Hydrogen Energy, 24, 305.
Zurrer H. et al. 1982, Studies in environmental science, 9, 31.
Zwart R.W.R. 2006, Biorefinery The worldwide status at the beginning of 2006. Wageningen
University and Research Centre (WUR) and the Energy research Centre of the
Netherlands (ECN)


97
4.0 TECHNOLOGY EVALUATION ASPECTS

Biomass is of particular interest within the ongoing discussion on sustainable mobility due to its
advantages concerning e.g. climate relevance and security of supply. Within the framework of
sustainability, the main superior targets relating to biofuels are (i) efficiency regarding system
technology and economics, (ii) environmental and climate protection regarding ecology (e.g.
greenhouse gas emissions) and (iii) energy supply security regarding biofuel potentials and
available resources. Further energy policy targets for biofuels are committed within the European
biofuels directive (2003/03/EG) up to the year 2010 and the roadmap for renewable energies
(COM (2006) 848) up to 2020. Accordingly, the share of biofuels within the transport sector has
to be increased to 5.75% (energy related) in 2010 and 10% in 2020. Besides, there are individual
country programmes, such as for example a commitment on the part of German gas suppliers to
reach a share of biomethane as natural gas substitute for transportation purposes of 10% in 2010
and 20% in 2020.

Moreover, technical targets on biofuels have to be achieved. Therefore, future biofuel production
systems should be integrated into existing technical biomass potentials. Only a reasonable mix of
promising biofuels should be implemented into the energy system under consideration of the
existing infrastructure of fuel distribution and use. These biofuels have to achieve current and
future exhaust emission standards (e.g. EURO 5/6) as well as have to be technically and
economically efficient in their production (i.e. high conversion rates and competitive costs).
To meet these future targets, there are many several promising biofuel options of the so called
2
nd
generation under discussion (i.e. bioethanol based on lingocelluloses, synthetic liquid and
gaseous biofuels, biogas and biohydrogen). These options will be evaluated in this section. The
different technologies for future biofuels options have been assessed in terms of (i) technology
aspects, (ii) economic, and (iii) environmental aspects. For that not only the production of
biofuels is considered but also their distribution and use.

4.1 General aspects of future biofuel production options
Biofuels of the future generation are currently at R&D stage, but different to the commercial
available 1
st
generation (i.e. biodiesel from vegetable oils, as well as bioethanol based on sugar
and starch) whole crops can be used for their production. Thus, they offer benefits regarding
use of land area for crop production and GHG mitigation. A simplified scheme of the whole fuel
supply chain from well to wheel (i.e. biomass production and provision, their conversion into
biofuels and the distribution and use) is shown in Scheme 4.1.


98


Scheme 4.1: Overview on biofuel production pathways

The characteristic conversion paths of the future biofuel generation are (i) the bio-chemical
conversion (i.e. using micro organism) for the production of biogas and bioethanol (EtOH) and
(ii) the thermo-chemical conversion for the production of synthetic biofuels such as Fischer-
Tropsch fuels (FT), methanol (MeOH), dimethylether (DME) and synthetic natural gas (SNG) as
well as gaseous and liquid biohydrogen (CG/LHyd). Their status of technology as well as the
R&D demand is characterised in the next sections.

Lignocellulosic bioethanol
While the production of bioethanol via fermentation of sugar and starch crops (e.g. melasse,
maize, cereals) is a matured and established technology all over the world, the use of
lignocellulosic biomass is more complex. This is due to the higher treatment expenditure for the
hydrolysis (e.g. via acids and enzymes) and saccharification of e.g. wood or straw for the
alcoholic fermentation. The ethanol treating (i.e. distillation, rectification and absolution) is
similar to the conventional ethanol plants. Currently, there are some pilot plants e.g. in Canada
(Iogen), Sweden (ETEK/SEKAB) and Spain (Abengoa). The expected plant capacity in future is
in the range of 5 to 110 MW
fuel
at a concept specific overall efficiency of approximately.
45 to 50 %. R&D demand primarily includes the development of pulping technologies for
lignocelluloses (e.g. efficient and rentable enzymes) as well as further optimisation of process
integration with ethanol plants (e.g. regarding lignin and mash residues use for energy purposes).


99

Synthetic biofuels
The production of synthetic fuels (i.e. design fuels with clearly defined properties) is
characterised by three main steps: (i) gasification of lignocellulosic biomass to a raw gas,
(ii) cleaning and conditioning of raw gas to bio-syngas, (iii) catalytic synthesis of this gas to
synthetic biofuels (i.e. FT, MeOH, DME and SNG) and (iv) final product treatment.

Regardless a long history of the development of a broad variety of system elements as well as
system layouts for the provision of liquid and/or gaseous fuels via biomass gasification, no
market breakthrough has been realised so far. One reason has to do with difficulties in
combining system elements. Additionally, some system elements are still under development.
With regard to economy of scale concepts in medium- to large-scale are required for an
efficient production of synthetic biofuels. Thus, the expected plant capacity will be in the range
of 30 to 340 MW
fuel
.

Despite the scale, for gasification system particularly, chemical characteristics, physical and
mechanical properties of the utilised biomass are of importance. But, all reactors for biomass
gasification are still in an R&D stage up to now and only available in small-scale. Depending on
fuel synthesis where reactors are available specific qualities of bio-syngas at constant
compositions have to be achieved (gas purity and the H
2
-to-CO-ratio). Because so far no
gasification system meets these requirements, appropriate gas cleaning and conditioning systems
have to be applied.

For raw gas cleaning either low temperature wet gas cleaning or, alternatively, hot gas cleaning
can be applied. The effectiveness of wet gas cleaning (e.g. cyclone and filter, scrubbing based on
chemical or physical absorption) has been proven for large-scale coal gasification systems.
Different to that, not all elements of hot gas cleaning (e.g. tar cracking, granular beds and filters,
physical adsorption or chemical absorption) are of mature technology yet. For gas conditioning
basically available and matured technologies (e.g. steam reforming, water gas shift) for
achieving bio-syngas requirements can be used; so far only limited experiences exist for the
required scale.

Based on CtL- and GtL-production fuel synthesis and conditioning, they are technically and
commercially feasible. However, there are only limited experiences regarding green syngas
and the expected plant scale. Comparing the different synthetic biofuel options, the FT raw


100
product treatment after synthesis is more extensive since refinery-similar technologies (e.g.
hydrocracker) are required.

In terms of overall efficiency, for biofuels such as SNG, high thermal efficiencies of up to 65%
can be expected. However, overall thermal efficiencies are lower (approximately 40 to 52%) for
FT, MeOH and DME based on MeOH.

Further R&D for synthetic fuel options is primarily required with regard to the development of
pyrolysis and gasification (e.g. upscale, operation under pressure) and the efficient and economic
gas treatment technologies. Moreover, the demonstration and commercial operation of the whole
chain (i.e. use of approved system components and their efficient interaction, plants availability
and reliability) is still needed. In terms of that, first results are expected from the operation of the
first demonstration plants for FT in Germany (Choren Industries), SNG in Austria and DME in
Sweden (Chemrec).

Biohydrogen
Even though biohydrogen is not a synthetic fuel, similar system components as for the
production of synthetic biofuels (i.e. gasification, gas cleaning and conditioning) can be applied.
After raw gas treatment the hydrogen rich gas is finally purified via pressure swing adsorption.
For distribution and use hydrogen need to be either compressed or liquefied. The expected plant
capacity will be in a similar range to synthetic fuels at an expected overall efficiency of about
50 to 55%. The further R&D demand is quite similar to that of synthetic biofuels.

Biogas
Biogas plants using wet and dry fermentation are a matured technology when the produced
biogas (methane-rich gas) is applied in power generation engines for CHP. To achieve natural
gas quality (e.g. for feed-in into the natural gas grid) system components for biogas treatment
(e.g. water or pressure swing adsorption) are basically available and successfully demonstrated
(e.g. in Sweden and Switzerland). A first demonstration plant in Germany is located in
Jameln/Wendland. Plant capacities, that can be expected, are in the range of up to 8 MW
fuel
at
concept and feedstock specific overall efficiencies of 43 to 86%.

Objectives of further R&D are the optimisation of process automation to increase the methane
yield in the biogas during fermentation and the upscaling of typically small-scale biogas plants.


101
Furthermore, biogas cleaning technologies need to be optimised and the feed-in of upgraded
biogas into the natural gas grid has to be demonstrated under long-run conditions.

4.2 Technology aspects
Based on the technology characteristics, the biofuel options have been compared by means of
biofuel production (i.e. stage of development, technical effort in terms of system complexity,
expected plant capacity and overall efficiency) and the biofuels suitability concerning current
fuel distribution systems as well as current use in vehicle fleets. A summary is shown in Table
4.1.

Therefore, it can be revealed that the different concepts for biofuel options of the future
generation are associated with appropriate benefits and bottlenecks along the pathway. While
e.g. FT fuels are less promising concerning the technical effort, the range of capacities (with
regard to suitable plant locations) and the overall efficiency, it seems to be very promising
regarding the implementation into the current distribution and use infrastructure.

Table 4.1: Comparison of technology aspects


4.3 Economic aspects
In addition to technical aspects, the decision on a preferable fuel is mainly driven by economic
reasons. Thus, the production costs of future generation biofuels have been analysed and
compared for exemplary overall concepts (Fig. 4.1).


102


Figure 4.1: Biofuel production costs

As the results of the economic analysis reveal, biofuel production costs show significant
differences. Based on litre fuel equivalent (FE) biomethane options (SNG and biogas) are the
most favourable. However, no cost reduction can be expected compared to biofuels of the 1
st

generation that are at a medium level of 62 ct/l
FE
.

The sensitivity analyses carried out for the determination and optimisation of influencing cost
components to the total biofuels production costs show that besides the annual full load hours
of the plant, feedstock costs and capital requirements are strongly important. It is expected that
biofuel production costs will moderately increase in future due to increasing energy prices with
expected price effects for feedstocks during broad implementation of biofuel strategies.

However, for a market implementation not only biofuel production costs but also total driving
costs relating to the well to wheel chain (WTW) are of importance for end users. Therefore, with
regard to the WTW biofuel costs involving costs of fuel distribution (i.e. via pipeline or tank)
and vehicles costs of private cars (i.e. in combustion and hybrid engines, fuel cells) the following
results (Fig. 4.2) can be indicated per vehicle km. The differences in biofuel production costs
will be lowered in terms of total driving costs as except for fuel cell application there is a
similar cost range for all biofuels, primarily dominated by vehicle use costs. The costs for biofuel
distribution play only a minor role. For a number of reasons (e.g. immature large-scale
production) the biofuels of the future generation are significantly more expensive when


103
compared to conventional oil-derived fuels such as diesel at total driving costs of approximately
0.33 /km.

Figure 4.2: Total driving costs private cars (WTW)

To identify very promising biofuel options, a relative comparison of technology and economic
aspects are shown in Fig. 4.3. Therefore, the technical feasibility has been determined including
flexibility and process requirements of the used feedstock, the technical effort, availability and
the input/output ratio biofuel production as well as the biofuel quality and the environmental
quality biofuel production (cf. Table 4.1). Based on that, very promising technologies are options
of a high technical feasibility and low biofuel production costs. This is true for bio-SNG as well
as partly for FT and biogas based on bio-waste and manure.


104

Figure 4.3: Technology versus economic aspects

4.4 Environmental aspects
The environmental effects of a product are not limited to their use or the production process.
Substantial environmental effects may also occur within the pre-chains. The most important
method to assess selected environmental effects throughout the biofuels chain is the life cycle
assessment (LCA), which can be applied to consider environmental impact categories such as the
anthropogenic green house effect indicated by the greenhouse gas emissions in the form of
CO
2
equivalents.

Biofuels options of the current and future generation have been compared relative to vehicle
kilometres according to the LCA. The results (given compared to the WTW costs) show
significant differences between the biofuel options (Fig. 4.4). Biofuels of the future generation
such as EtOH, SNG, FT and biogas promise better effects regarding GHG mitigation. However,
WTW costs are slightly higher for most of future biofuels generation compared to 1
st
generation.


105

Figure 4.4: Economic versus environmental aspects

Conclusions
To meet the future energy policy and technical targets for the implementation of biofuels into the
energy system basically many different biofuel options as well as several alternative concepts of
the future generation with different benefits and bottlenecks are under discussion. However, so
far no silver bullet could be identified.

Production concepts for biofuel option of the future generation are currently in laboratory (e.g.
biohydrogen) to pilot and demonstration stage (e.g. bioethanol, Fischer-Tropsch fuel,
dimethylether and Bio-SNG). Thus, further R&D is required with regard to (i) further
development and use of promising biofuel concepts in practice, (ii) upscaling to medium- and
large-scale plants, (iii) use of approved system components and demonstration of efficient
interaction, (iv) plant availability and reliability. For a successful market implementation existing
techno-economic barriers have to be overcome and capital risks need to be minimised.

From economic viewpoint no cost reduction can be expected from the 1
st
to the future biofuel
generation. The costs for biofuel production are dominated by feedstock costs as well as capital
related costs. Thus, concepts efficiency and plants availability are of high importance. Except for
fuel cells, costs per driven kilometre of private cars (WTW) are at a comparable level. They are
dominated by vehicle costs, while distribution costs are only less relevant. Comparing
technology and economic aspects relatively, biofuel options such as SNG as well as partly FT
and biogas based on bio-waste and manure are the most promising. Despite this, in context of the


106
entire supply chain of transportation fuels synthetic biofuels can be seen as a bridge into the
future hydrogen economy in the long term.


107
4.5 Bibliography

Mller-Langer F. 2007, Analysis and evaluation of the 2nd generation of transportation biofuels,
15th European Biomass Conference & Exhibition - From Research to Market
Development, Berlin, May.
Mller-Langer F. et al. 2007, Analyse und Evaluierung von Anlagen und Techniken zur
Produktion von Biokraftstoffen. Final report for the Energiestiftung Baden-Wrttemberg,
Leipzig.








108
5.0 DECISION SUPPORT TOOLS FOR BIOFUELS AND BIOFUELS
TECHNOLOGY ASSESSMENT

Conceptually, the Decision Support Tools (DSTs) refer to analysis, comparison, and selection of
possible options (technologies or products) according to their characteristics. Such characteristics
should be presented as the quantified data that can describe a variety of aspects of decision
makers interest. These aspects often refer to the process performance, environmental, social
economic, and risk aspects.

Technical assessment mainly refers to the process performance/efficiency, its scale, inventory of
consumables, specific physical and chemical characteristics (in the case of products comparison),
etc. The environmental impact assessment accounts for process emissions, wastes, and resource
depletion which are expressed in terms of general environmental indicators, described by the
LCA procedure. Such environmental impacts relate not only to technology operation or
particular product production, but also to the production of all material and energy consumed by
the process. The cost benefit analysis estimates unit set up and operation costs, as well as other
specific costs associated with risk, environmental and social factors. Benefits are primarily
related to the valorization of products and energy.

It is understood that the decision making process usually requires an integrated approach which
involves all the abovementioned aspects, which are usually interdependent. For example, the
thorough cost-benefit analysis can reflect various aspects, also including social to environmental.

Usually, the decision maker faces a task of taking into account a big number of factors or criteria
(technical, environmental, and other indicators) for comparison of options. Such analysis and
evaluation can be performed with the help of the Multicriteria Analysis (MCA), which is often
employed in DSTs. The MCA procedure is useful for an integrated evaluation, i.e. when more
criteria for comparison/selection should be taken into consideration or when such criteria should
be given different importance (weights). Several DSTs developed in ICS-UNIDO (DARTS,
DAWTS, SPORE, etc.) are based on the MCA.

The combined approach for decision making, represented in the DST by introduction of a variety
of different indicators and providing a possibility to manage the related information in a
comprehensible and integrated way lays the base for the adequate assessment of


109
production/feedstock exploitation scenarios and, therefore, is the key towards the development
and adoption of more sustainable practices, technologies, and overall production processes.


5.1 The concept of decision support tool for assessment of scenarios of biofuel
production and bio-feedstock exploitation

Decision support tools can offer a substantial assistance to the assessment of different biofuel
production pathways, and different bio-feedstock exploitation scenarios. The proper evaluation
of related production processes should be based on their sustainability analysis, which implies
the assessment of associated economic and environmental indicators.

The DST concept (currently under development in ICS-UNIDO) for assessment of different
scenarios of bio-feedstock exploitation is being presented. The underlying principle is presented
in Scheme 5.1.

Choice :
Technology or
sequence of
processes
Inflow
(feedstock)
Outflow 1
(material)
Inputs
(material, E)
Outputs
(env. impact)
Outflow 2
(cost, E)
Choice :
Technology or
sequence of
processes
Inflow
(feedstock)
Outflow 1
(material)
Inputs
(material, E)
Outputs
(env. impact)
Outflow 2
(cost, E)

Scheme 5.1: Concept of the DST for economic and environmental assessment of resource
exploitation scenarios

The inflow data depend mainly on the choice to be made by the user and refer to the feedstock
amount, type/quality, treatment and scale choice. The technology (treatment) block is the core
component as it reflects the choice of technology or technology sequence which helps the user to
define the scenario. Depending on this choice the system, and cost outflows (deliverables to be
valorized) are expected to change as well as material and energy inputs and the environmental
impact outputs. The inflow data can be mainly referred to as user defined (choice to be made),
while some of the data related to feedstock properties and majority of data related to technology


110
operation can be considered as default or built in (except some cases when user can define the
process linked data, e.g. if such are missing), as described in more detail in Table 5.1.


Table 5.1: Input data for the DST concept for assessment of sustainability of bio feedstock
exploitation for biofuel production scenarios
Data areas User defined Built-in
Feedstock type
amount (throughput)
(properties)
properties
moisture content
oil content
lignocelluloses content
other: FFA, P, S, I, C/H/O
Process process type
process scale
feedstock type
(energy type)
emission coefficients (process and
consumables)
efficiency
by-products/residuals amount
product/by-products properties
energy/fuel requirements (amount and type)
consumables
costs
Energy/fuel (energy type)
electricity grid
transportation
distance
fuel type
energy/fuel production, delivery, and
consumption emissions
fuel properties
heat content
fuel efficiency


Input data related to consumables (amounts and type of chemicals, fuel/energy) are mainly built
in type and are calculated based on the choice of the technology block and of the material inflow.
The same is valid for the output data or the environmental impacts produced, which are also
calculated based on the process inventory and scale. The environmental impact data refer to both
process operation/transportation and to production of all material and energy/fuel consumed by
the process/transportation. Such approach to evaluate the environmental impact is known as the


111
LCA. Adopting the basic principles and terminology of LCA, this impact can be broken down in
several categories, namely:

Output data presented environmental impact (indicators)
Smog precursors (NOx, VOC, PM)
Resource depletion (E, water, material)
Acid rain (SOx, NOx, HNO
3
, H
2
SO
4
, H
2
S, NH
3
)
Climate change (CO
2
, CH
4
, NOx, CFC)
Eutrophication (phosphates, nitrates, COD)
Human and environmental toxicity and other effects (heavy metals, POP/PTS, particulates,
residual solid waste, )

The outflow data are presented by the amount of the product to be valorized and its quality
(properties). Energy can be also considered a product of value and is therefore also taken into
consideration. The cost balance represents another important outflow which includes all
expenses related to process operation and can also include cost of feedstock and benefits derived
from the energy and products valorization. In particular the economic data refer to the
technology set-up costs (unit and installation costs), operation costs (consumables, fuel,
electricity, labour, rent, amortization, etc.), transportation costs, feedstock costs. Benefits are
based on valorisation of products, by-products, and recovered energy.

As the key point of decision support lies in allowing the user to choose options that he wishes to
compare, there is a possibility that the user selects the desired processes and therefore defines the
destination of the feedstock. In the case of bio-feedstock for bio-fuel production, and in
particular in the case of oil crops, the technological scenarios can be represented by a variety of
steps of the bio-feedstock processing train, which starts from the raw biomass (harvested food
crops) and ends with a product, such as edible oil, biodiesel, glycerol, heat, or chemicals. The
technology train therefore can include a number of processing options (processes), such as
various pre- and post-treatment steps (drying, extraction, separation, purification, etc.),
technological steps related to chemical modifications of the feedstock (oil esterification, biomass
pyrolysis and gasification, hydrolysis of cellulose, fermentation or sugars, incineration, various
glycerol processing options, etc.). The general flow diagram representing the possible pathways
of bio-feedstock exploitation in the case of oil-crop for biodiesel is presented in Scheme 5.2.



112
The overall framework of the DST is divided into four technology blocks, which are described in
more detail in Table 5.2. The first block refers to the crude crop biomass processing that yields
edible oil and bio-waste. These two can be considered as products or as intermediates in the case
where further use is made of them, e.g. biodiesel production or bio waste processing. The second
block includes the technological options for biodiesel production from oil, therefore it yields
biodiesel and glycerol as products. Another (the third) technology block refers to the processing
of other types of primary bio-based materials, such as cellulosic materials or bio-waste left after
oil production, or sugar (e.g. in the case of sugarcane). These materials can be processed via
pyrolysis, gasification, fermentation, or incineration yielding syn-gas, biogas, cellulosic ethanol
or heat. The forth block is dedicated to glycerol valorisation, which can be converted towards a
variety of chemicals (products). In principle the framework of the DST can be started with
another starting material than crop biomass, like oil or bio-waste so only the technological blocks
of interest to the user are considered as the scenario framework.

Table 5.2: Technological approaches for exploitation of different bio-feedstock/by-products
derived from oil crops
Technology module Technology options
1. Oil crop feedstock
processing
Oil extraction from seeds
- industrial pressing
- small scale (cold) pressing (depends on crop and scale)

Oil refining
- degumming
- deacidification options
- bleaching
- dehydration

chosen by user or suggested according to feedstock type and
composition/properties


113
2. Biodiesel production Type of process: transesterification, amidation, incineration
Transesterification catalyst choice
- Homo- or heterogeneous, acid or base catalysis can be used,
namely: alkali/alkaline metal alcoholates, metals/alloys,
salts, oxides, acids, zeolites, lypases, metal complexes,
polymer bound, supported catalysts, etc. also usually as
combination of different catalysts
Reactor design: continuous flow/batch, agitation type (US
assisted, mechanical)
Scale, mobility
other systems design: alcohol and catalyst recovery,
separation and clean-up steps
3. Solid residual
biomass conversion
Type: pyrolysis, gasification, incineration, fractionation
cellulose hydrolysis fermentation,
Gas cleaning and pretreatment steps
4. Glycerol based
syntheses
Microbiological and catalytic reduction
Catalytic oxidation
Selective esterification and desertification
Carbonatation
Thermal or catalytic transformation to fuel gas
Other syntheses

As it was mentioned, the inflows of starting materials, products, by-products, or intermediate
(which eventually can be starting materials or products depending on the scenario) outflows
constitute the data streams which connect different technological blocks and are characterized by
amounts and are associated with the data on quality characteristics of related materials. The
quality mainly refers to a set of chemical and physical properties and composition of the
substance which can be either defined by the user, like in the case of the starting material (seeds,
unrefined/waste oil, fat) - moisture, sulphur, phosphorous, iodine, elementary content, FFA, etc.
or, in the case of products, result from the processes involved and of type/quality of the
feedstock.


114




















Scheme 5.2: Flowchart of the DST for bio-fuels exploitation

Glycerol
Chemicals
Biodiesel
Ethanol, biogas,
syn-gas, etc.
Consum.
Impact
Technology
module 4
Crop
biomass
Consum.
Impact
Oil
Sugars or
bio-waste
Technology
module 2
Consum.
Impact
Technology
module 3
Consum.
Impact
Heat
Block 1:
Feedstock
processing
Block 2: Biodiesel production
Block 3: Sugars and bio-waste
exploitation
Block 4: Glycerol
valorization
Heat
Technology
module 1
115
5.2 BioAs: A Decision Support Tool for Biofuel Assessment and Selection
Within the ICS-UNIDO work programme there was an international collaboration project on
the development of a specific DST aimed at helping decision-makers to make a desirable
choice among different liquid and gaseous biofuels using multiple criteria. This system
provides decision support in biofuel selection using the PROMETHEE II and ELECTRE III
algorithms as its basis (Stanojevi, 2006a).

The biofuels and criteria are hierarchically organized using tree structures. A user is free to
define the tree structures by adding/deleting/ modifying biofuel groups, biofuels, criterion
groups and criteria. The initial template used in Multi-Criteria Analysis contains 6 liquid and
gaseous biofuels for Combined Heat and Power (CHP) production and 44 technological,
financial, socio-economic and environmental criteria.

The template supports the comparison of the following liquid biofuels used in CHP
production:
rape-seed vegetable oil
rape-seed methyl ester (RME)
flash pyrolysis oil and the following processes using gaseous biofuels:
slow pyrolysis (EDITTh process)
waste methanization
gasification from wood

Criteria
The technological criteria include: energy content, non-renewable energy consumed,
availability, carbon residue, sulphur content, viscosity and density. The financial criteria are
grouped into static, dynamic and risk criteria. Static criteria include standard financial and
efficiency criteria: levelized cost of biofuel, net profit plus interest to investment, long-term
debt to net worth, net cash flow to total sales, etc. Dynamic criteria are: net present value
(NPV), internal rate of return (IRR), normal and dynamic payback period, etc. Some
statistical indicators are used as risk measure: mean values for NPV and IRR, standard
deviations, intervals of variation, zero risk equivalent, etc. The socio-economic criteria
include quality of life, avoided rural depopulation, rural diversification and land
management, economic gain, incomes, economic activity, related industry support, and
116
employment generated. The environmental criteria are represented by emissions of: CO
2
,
CO, HC, NO
x
, particulates and SO
2
. The abovementioned criteria for selection of biofuels
have been defined with the help of the FIDES software (Stanojevi, 2006b; Vrane, 2002a;
Vrane, 2002b).

Technical information
BioAS is implemented using Java programming language, Java 2 Platform, Enterprise
Edition (J2EE) and three-tier architecture. In its development and implementation, Open
Source database, Integrated Development Environment (IDE) and Application Server are
used (see Figs. 5.1, 5.2).


Figure 5.1: Input screen of the BioAs software

117

Figure 5.2: Results screen of the BioAs software

118
5.3 Bibliography

Stanojevi M. et al. 2006a, Initial Results of Socio-Economic Analysis for Biofuel
Production in Europe, WSEAS Transactions on Environment and Development, Vol.
2, No. 1, ISSN 1790-5079, 38-46.
Stanojevi M. et al. 2006b, Integration Platform for Investment Projects in Goran D.
Putnik , Maria Manuela Cunha; (Eds..), Knowledge and Technology Management in
Virtual Organizations: Issues, Trends, Opportunities and Solutions, IDEA Group
Publishing, ISBN: 1-59904-165-0, 263-281;
Vrane S. et al. 2002a, Investment Decision Making, in Cornelius Leonides (Ed.) Expert
Systems Techniques and Applications, Volume 4, Chapter 31, Investment Decision
Making, Academic Press, 1107-1135;
Vrane S. et al. 2002b, Financial decision aid based on heuristics and fuzzy logic in E.
Daimiani et al. (Eds.) Knowledge based Intelligent Information Engineering systems
and allied technologies , IOS Press, ISBN 1 58603 280 1, Netherlands, 381-387.

119
ANNEX
Biofuels around the world

A. Biodiesel

This part is dedicated to a preliminary survey on the situation in the sector of biofuels
(technology development, production, use) in various counties/regions around the world.
First part (A) is focused on biodiesel situation and the second part (B) presents the situation
as pertains to bioethanol.

A.1 Europe
In Europe, consumer interest in passenger cars with diesel engines strongly increased after
the oil shortages of 1973 and 1979. The industry in Germany, France and Italy developed
energy-saving, highly efficient engines. The first research activities regarding the
development of alternative and renewable fuels were started. Commercially motivated
biodiesel-initiatives in Europe were developed as early as 1988 predominantly in Austria and
also in France, where the first industrial scale biodiesel production plants went into operation
in 1990/1991.
In 1992, reform of the Common Agricultural Policy addressed European agricultural
surpluses by idling some land used for food production through a set-aside policy. This
policy stimulated the use of set-aside land for non-food purposes. Low oil prices in the
second half of the 90s resulted in reduced interest on the part of industry and policy makers
in liquid biofuels. In 1998, the very disappointing contribution of 452.000 t coming from
biofuels reflected the situation that specific policies had been adopted in only four member
states: France contributed 58%, Germany 21%, Italy 18% and Austria 3%.
In June of the same year, as a consequence of the 1997 Kyoto Conference on Climate
Change, the EU-member states decided on a reduction of 8,1 % on the basis of 1990
emissions for 2012, a goal which can only be realised by using a considerable amount of
renewable energy sources including liquid biofuels. This urgent need for practical action to
address increasing CO
2
emissions from the transport sector resulted in a proposal for an EC
Directive promoting the use of alternative transport fuels derived from biomass and reducing
rates of excise duty on such fuels. This resulted in the enactment of DIRECTIVE
2003/30/EC OF THE EUROPEAN PARLIAMENT AND OF THE COUNCIL of 8 May
120
2003 on the promotion of the use of biofuels or other renewable fuels for transport (EU,
2003). This was followed by the EU Strategy for Biofuels in 2006.
In recent years, the production of biodiesel has made a substantial leap in the European
Union. From 1996 to 2002, the biodiesel production capacity grew fourfold to a total of 2
million tonnes. Growth has been quite phenomenal since then, with 2006 production capacity
at about 4.9 million tonnes. The number of plants producing biodiesel increased from 40 in
2003 to 94 in 2005. Industry stakeholders in Europe are given in Table A-1.

Table A-1: Biodiesel industry stakeholders in the European Union


At present, most biodiesel in the EU is processed from rapeseed and sunflower oil with a
growing percentage of recycled frying oil. The industry relied to a significant extent on
obtaining the right quantity of feedstock at a competitive price by virtue of the non-food set-
121
aside scheme, which resulted from the MacSharry Common Agricultural Policy (CAP)
Reforms introduced in 1992. Although the CAP Reforms agreed upon within the
framework of Agenda 2000 at the Berlin Summit of March 1999 provided for a reference
rate for obligatory set-aside of 10% for the period 2000/2006, the variable nature of set-aside
rates actually applied from year to year did not always offer a sustainable base for biodiesel
production.

It seems that the main possibilities for the production of conventional feedstock for liquid
biofuels production under the studied conditions will be in northern and central EU countries,
and will be reduced in Mediterranean countries because the yields are rather low in these
areas. Recycled vegetable oils and fats offer some possibilities as alternative low-cost
biodiesel feedstock whose availability is not affected by EU land use policies.

Based on the estimate from eight countries, a total of about 0,4 million t were collected in the
EU in 1999, mainly from the catering industry. The amount that could be collected is
estimated as being considerably higher, possibly from 0,7 to 1 million t. Its price is variable,
but in general about half of that of virgin oil. The countries with most practical experience on
this subject are Austria and Germany (with two new biodiesel plants of 100.000 t/a capacity
using recycling oil as the sole feedstock). Even with this alternative, the supply of feedstock
will be the limiting factor for further development of the biodiesel economy. Given the
existing production facilities, it cannot be anticipated that over 10% of diesel fuel
consumption can be replaced by biodiesel. The European Commission has acknowledged the
importance of biodiesel by including it in its Campaign for Take-Off and has defined it as a
Key Sector Action. The trend set by future projects makes it possible to situate European
production in 2003 for all types of biofuels (relatively in phase with European Union
objectives) at about 4,8 million tons. A simple projection of the present rate shows that the
goal of reaching a 7% share of consumption represented by renewable origin fuel by 2010
will not be met (11,7 million tons vs. 17 million tons) without additional efforts.

A. 2 Belgium
Interest in liquid biofuels started at the beginning of the 90s with the set-aside regulation
allowing non-food crops. Since 1991, several trials on commercial buses have been made to
promote the biodiesel utilization. A transesterification pilot plant was installed in 1994. In
1994, 9.500 ha of oilseed rape were grown on set-aside land and biodiesel was used in
122
several trials (busses in Mons (15 busses running with 20% biodiesel, 740.000 km, technical
survey, favourable results), municipal vehicles in Charleroi and Philippeville, vehicles of
AVEVE and cars in Mol (5 cars, 100% biodiesel, 300.000 km, technical survey, no major
technical problems).

In 1998 almost 5% of the biodiesel in Europe was produced in Belgium by Pantochim
(19.000 out of 390.000 t produced).

Belgium had taken biodiesel to the research, development and demonstration stage, but there
was no considerable progress to the deployment stage because of the reservations on the
overall economic and environmental performance of the fuel. After these projects, the use of
biodiesel stopped entirely in Belgium but fresh programmes began to arise in 2006 following
the approval of a tax advantage for biodiesel. There are refineries in Ertvelde (belonging to
the company Olon) and at Feluy. Biodiesel industry stakeholders in Belgium are presented
in Table A-2.

Table A-2: Biodiesel industry stakeholders in Belgium


A. 3 Czech Republic
Czech production of biodiesel was already above 60.000 m per year by the early 1990s and
is now even larger. Many of the plants are very large, including one in Olomouc which
produces almost 40.000 m per year. From the summer of 2004, Czech producers of biodiesel
for blend receive a subsidy of roughly CEK 9,50/kg. All koda diesels built since 1996 are
warranted for biodiesel use. Total production of biodiesel in 2007 was 61.000 tonnes with
2008 production capacity estimated at 203.000 tonnes. Biodiesel industry stakeholders in the
Czech Republic are presented in Table A-3.
123

Table A-3: Biodiesel industry stakeholders in Czech Republic


A. 4 Estonia
Biodiesel is available at Favora fuel stations. After 10 years of research Tomson formed
Biothompson O. Biothompson factory produces biodiesel using rapeseed and alternative
oils with a capacity soon 100000 T/y.

A. 5 France
It was in the mid-80s that Frances oil-producing and -processing industry was looking for
new markets in order to promote oil of colza that was heavily underrepresented in the food-
market at the European level.

From 1991 to 1995, a test program with the participation of all stakeholders involved
(vehicle industry - Renault, Peugeot; oil-processing industry, petrol-industry - Elf, Total;
Ministries of Agriculture and Industry; energy agencies and public transport companies) tried
to figure out the most favourable way to produce, distribute and use biodiesel. The result was
to incorporate a 5% blend with fossil diesel and commercial production started.

From the first unit pilot in the Compiegne area to the opening of the Rouen/Dico unit with
150.000 t/a of RME, the specificity of the French biodiesel activity has been to move toward
big transesterification units.
1993: demonstration plant initiation in Compigne
1994: Novaol adapting two sites for biodiesel production
1995: start-up of the Grand Couronne- site
1997: 250.000 t/a production
124
1998: downturn of biodiesel consumption due to the reduction of mandatory set- aside land
2000: 317.500 t/a authorized state quota230

Today, biodiesel is known under the trademark and common name Diester, the contraction
of Diesel and Ester. Biodiesel industry stakeholders in France are presented in Table A-4.

Table A-4: Biodiesel industry stakeholders in France




A. 6 Finland
125
Neste Oil began the production of alkyl biodiesel using the NExBTL process in summer
2007 in Finland, with a capacity of 170.000 tons/year. A contract has been signed with the
French Total, to begin production in some Total refineries in 2008.

NExBTL diesel, in contrast to rapeseed methyl ester, is a clear and colourless paraffin, and
contains no oxygen. It is used to improve the quality of petro-diesel; its quality is higher
since it has a homogenous source, namely plant-synthesized fatty acids. It doesnt require
any special engine repairs and it doesn't foul systems like ester biodiesel. It is produced by
direct hydrogenation of the plant oil (chemically, triglyceride) into alkane, water and carbon
oxides on a nickel-molybdenum catalyst. The total CO
2
produced in the entire lifecycle is
only 0,45 to 1,33 kg CO
2
/kg oil, in contrast to transesterified fuel with 1,4-2,0 kg CO
2
/kg oil,
or mineral diesel with 3,4 kg CO
2
/kg oil.

A. 7 Italy
Biodiesel production in Italy started in 1992 after a joint venture project co-financed by the
European Commission. In 1995, the Italian government provided the detaxation of 125.000t
in order to make the selling cost of biodiesel comparable to that of fossil diesel. Nevertheless,
production in 1998 was still less than 90.000 t/a. However, an increasing trend has been
recorded in the last four years.

Overall activities in Italy have been limited so far to utilise biodiesel mainly in the heating oil
segment and not in transportation as exercised in all other European countries. With the new
Directives of the European Commission in place the well established and experienced Italian
biodiesel producers will have to reconsider their marketing philosophy in order to fully
exploit the new market opportunities in the transport sector. Biodiesel industry stakeholders
in Italy are presented in Table A-5.

Table A-5: Biodiesel industry stakeholders in Italy
126


A. 8 Germany
According to the Union zur Frderung von l- und Proteinpflanzen UFOP (Union to
promote oil- and protein plants), in 2004 the sale of biodiesel through German gas stations
rose to 375,000 m, although it was only available at a limited number of outlets. In 2004,
45% of all biodiesel sales went directly to large end users, such as trucking companies.

Production capacity for biodiesel, for the most part produced from rapeseed, rose to over 2,6
million tonnes in 2006. By the end of 2007 the production capacity was expected to have
reached almost 4.5 million tonnes. Sales in Germany have doubled to 376.6 million litres
(about 99 million US gallons) from 2002 to 2004. This amount is sufficient to meet the
average yearly consumption of well over 300,000 automobiles. Diesel engines have become
increasingly popular in Germany and almost half of all newly manufactured cars are diesel
powered. This is in part due to the greater efficiency of diesel engines, the desire by
consumers to use environmentally friendlier technologies and lower taxes on diesel fuel that
make it cheaper than gasoline. With 1,900 sales points, equal to one in every ten public gas
stations, biodiesel is the first alternative fuel to be available nationwide. The industry is
expecting a surge in demand since the authorisation at the beginning of 2004, through
European Union legislation, of a maximum 5% biodiesel addition to conventional diesel fuel.
In Germany biodiesel is also sold at a lower price than fossil diesel fuel. Biodiesel industry
stakeholders are presented in Table A-6.

127
Biodiesel used to be a non-taxable fuel in Germany. In January 2007, however, the Biofuel
Quota Act went into force. As a consequence a gradual reduction of tax privileges for
biodiesel took effect and a tax of 9 cents/litre was imposed. This is supposed to be increased
gradually and will reach 45 cents/litres from 2012 onwards.

Table A-6: Biodiesel industry stakeholders in Germany


A. 9 Slovakia
128
In 1991, under the former federal system of the SFR government, an oleoprogramme was
launched, supported by initiatives for process development and quality control at the
University of Bratislava. The first small-scale plant went into operation in 1992 followed by
additional facilities for the rapeseed oil methyl ester (ROME) production developed by small
and medium entrepreneurs. In 2001, total production capacity amounted to more than
120.000 t RME. Changes in subsidy legislation paralyzed biodiesel production at the end of
2001. Table A-7 presents some of the biodiesel industry stakeholders in Slovakia.

Table A-7: Biodiesel industry stakeholders in Slovakia


A. 10 Switzerland
After initial research activities in the early 90s by various Swiss institutes (FAT, EMPA), a
small but steady production was realised. Table A-8 presents some of the stakeholders in the
Swiss biodiesel industry. BIOPETROL INDUSTRIES AG (BIOPETROL) and a group
company of Royal Vopak (Vopak) in March 2006 signed a long term co-site agreement to
construct a biodiesel facility in Rotterdam at the Vopak Terminal Botlek Noord. The plant
which covers a total surface area of 34,000 m
2
will produce 400,000 tons of biodiesel and
60,000 tons of pharma-glycerine per year.


129
Table A-8: Biodiesel industry stakeholders in Switzerland


A. 11 Norway
Biodiesel is not in common use in Norway. The three biodiesel pumps in Norway at
Lillehammer, Hadeland and Oslo are managed by the Norwegian oil-company Hydro-
Texaco. Biodiesel is also available in Bergen and supplied by Milvenn AS. Currently, all
alternative fuels are exempt from fuel tax and the carbon dioxide tax. In April 2006, the
Norwegian Pollution Control Authority (SFT) suggested that at least 2% of the sales volume
of petrol companies and other sellers of motor fuel should be biofuel and they also suggested
increasing the requirement to 4% in 2010. The Norwegian government following the
recommendations announced a new fuel standard requiring 2 percent volume share of
biofuels in 2008 and a 5 percent share in 2009. There is also a goal to reach a 7 percent
volume share (approximately 5,75 percent when measured on energy content) in 2010.

A. 12 Spain
The raw materials to produce biodiesel in Spain include traditional vegetable seed oils
(sunflower and rapeseed), alternative vegetable oils (Brassica carinata), genetically modified
vegetable oils (high oleic sunflower), and used frying oils. The traditional raw materials are
the surplus production of seed oils (sunflower and rapeseed). The use of these vegetable oils
for biodiesel production has been extensively studied.

In Spain average sunflower oil production (834 k/ha) is less favourable than in other
European countries. Therefore, it is not very profitable to use this oil in the production of
biofuels. Likewise, the yield per hectare for rapeseed oil is only a little higher than that for
sunflower oil in Spain. Conversely, the cultivation of B. carinata as an oilseed crop for
biodiesel production in the south of Spain has been of special interest, since it allows the use
130
of set-aside lands, giving higher yields per hectare than the traditional crops. The vegetable
oil obtained from B. carinata is characterized by the presence of a high concentration of
erucic acid, which is considered harmful for human consumption. Attempts to modify this
crop have resulted in the elimination of erucic acid from the oil. In this sense, there are two
sorts of B. carinata oil, low erucic and high erucic, based on its erucic acid composition. So
far, there have been a few references that describe the use of this vegetable oil as a raw
material for biodiesel production.

Spain is a major consumer of vegetable oils, mainly olive and sunflower oil, and they are not
reused many times. Consequently, the quality of the oils is not significantly affected, making
them very suitable for the production of biofuels (Energy & Fuels, 2006, 20, p. 395). It is
possible to buy biodiesel, mixed with diesel fuel, in more than 250 petrol stations around the
country. Besides, B100 is sold in two petrol stations around Pamplona, in Navarre.

A. 13 Austria
The shortage of mineral oil supply in the early 70s led to two major biofuel projects after
1979. The construction of an 80.000 t/year bioethanol plant (not realized) and the
development of the biodiesel technology were planned.

In 1982, the first production of rapeseed oil methyl esters (RME) by transesterification of
rapeseed oil and engine tests in diesel engines were carried out in Austria. One year later first
trials with recycled frying oil were conducted. In 1985, the first pilot plant world-wide for the
production of biodiesel for use in agriculture was installed in Silberberg, Styria. In 1990, the
construction of the first industrial scale biodiesel-plant was started in Aschach.

In 1992, simultaneous to the first specifications for RME in Austria (standard ONC 1190),
the first modern farmers cooperative production plant for biodiesel was put into operation in
Mureck. A short time later, a second commercial and so far the biggest plant was opened in
Bruck.

The year 1995 saw the first international conference on standardisation and assessment of
biodiesel in Vienna. In the same year 10 buses from the public fleet of Graz switched to
100% biodiesel produced from recycled frying oil.

131
Table A-9: Biodiesel industry stakeholders in Austria


In June 1999, the drafts of two regulations were signed by the Austrian Minister of the
Environment. These regulated the quality of transportation fuels (maximum of 5% biodiesel
in fossil diesel fuel) and including a regulation for a mandatory adding of 2% biodiesel in
fossil diesel fuel. Consent could not be found on the latter point, so the mandatory adding
was rejected. The blending level was limited to 3%. An amendment of the fuel regulation
132
was published in December 1999 creating the possibility of using biodiesel as a sole diesel
fuel and as a blending component to fossil diesel fuel.

In accordance with the Mineral Oil Tax Law, fuels produced from biogenic substances are
exempt from mineral oil tax. The blending of up to 2% biodiesel with diesel is also exempt
from tax. There is also a tax reduction for the blending of up to 5% biogenic fuels with
petrol. Some biodiesel industry stakeholders are presented in Table A-9.

A. 14 United Kingdom
Biodiesel is sold by a small but growing number of filling stations in B5 and B100 blend.
Some farmers have been using small plants to create their own biodiesel for farm machinery
since the 1990s. Several Co-ops and small scale production facilities have recently begun
production, typically selling fuel several pence per litre less than petrodiesel. The first large
scale plant, capable of producing 50 million litres (13 million US gallons) a year, opened in
Scotland in 2005. Biodiesel is treated like any other vehicle fuel in the UK and the
paperwork required to register as a producer is a major limiting factor to growth in the
market. Table A-10 is a list of some of the stakeholders in the UK biodiesel industry.

Table A-10: Biodiesel industry stakeholders in the United Kingdom
133


A. 15 China
In 2004, several companies started making biodiesel, and have produced more than 5,000
kilotons in a year since then. In 2006, the Bureau of Energy launched the first biodiesel buses
on Earth Day. China Biodiesel, an international renewable energy corporation focusing on
biodiesel R&D, manufacturing, marketing and investment, announced in June 2008 that it is
set to begin trial production at its new plant located in Xiamen, Fujian Province, China that
will double the companys annual production capacity to 100,000 tons, nearly four times the
2007 output. Stakeholders in the biodiesel industry in china are presented in Table A-11.

Table A-11: Biodiesel industry stakeholders in China
134


A. 16 Singapore
In January 2008, Finnish oil refiner Neste Oil said that its biodiesel plant in Singapore, its
largest, will be operational 2010 and initially cater to Europes growing biofuel requirement.
The companys investment of 550 million Euros to build the plant, which will have an annual
capacity of 800,000 tons, is in line with its strategy to become the worlds leading producer
of diesel from renewable feedstock. Meanwhile, a joint-venture announced in 2007, Van Der
Horst Biodiesel, is planning to build Singapores first biodiesel plant that uses Jatropha
curcas and not palm oil as feedstock. The plant on Jurong Island is the project of a joint
venture between the Institute of Environmental Science and Engineering, which is linked to
Nanyang Technological University, Singapore, and Van Der Horst Engineering. It will have
an annual capacity of 200,000 tons per year. Another joint venture between Wilmar Holdings
and Archer Daniels Midland Company, was expected to be operational by the end of 2006
with an initial capacity of 150,000 tons/year. Singapore is hoping that with current efforts, its
biodiesel production output should exceed one million tons per annum by 2010, and reach
three million tons per annum by 2015.

A. 17 Thailand
Thailand was the first country to launch biodiesel as a national program on July 10th 2001. It
was reported that the work was initiated by the Royal Chitralada Project, a royal -sponsored
project to help rural farmers. International co-operation among ASEAN country was also
started by the Renewable Energy Institute of Thailand and Asia-Pacific Roundtable for
Sustainable consumption and Production. The primary aims of the project in Thailand are:
135
an alternative output for excess agricultural produce
substituting diesel imports

In 2006, several biodiesel plants are operating in Thailand using the excess palm oil/palm
stearin and in some cases, waste vegetable oil as raw materials. About 15 petrol stations are
now distributing B5 in Chiangmai and Bangkok. The national biodiesel standard has been
developed based on the European standard. The target of the Government is to mandate B5
by 2011 which will require almost 4 Million litres/day of biodiesel. The raw material will
most likely come from palm oil, coconut oil, Jatropha Curcas Linn, and tallow. Several pilot
plants are now operating such as the Royal Chitralada Projects, Rajabiodiesel in Surattani,
Department of Alternative Energy Development and Efficiency, Royal Naval Dockyard, and
Tistr (www.tistr.or.th).

A. 18 Malaysia
Biodiesel called the Envo Diesel was launched by the Prime Minister Datuk Seri Abdullah
Ahmad Badawi on Tuesday 22 March 2006. Malaysia currently produces 500,000 tonnes of
biofuel annually and the government hopes to increase this number. Envo diesel blends 5%
processed palm oil (vegetable oil) with 95% petrodiesel. In contrast, EUs B5 blends 5%
methyl ester with 95% petrodiesel. Diesel engine manufacturers prefer the use of palm oil
methyl ester blends as diesel engines are designed to handle 5% methyl esters meeting the
EN14214 biodiesel standard, which palm oil cannot meet.

Malaysian Palm Oil Board (MPOB) established in 2000 is the leading government agency
entrusted to serve the countrys oil palm industry, which promotes and develops national
objectives, policies and priorities for the development of the Malaysian oil palm industry. As
the major producer of palm oil, Malaysia has provided the lead in the development of
biofuels in Asia and MPOB is promoting various biodiesel/biofuel programmes in Malaysia
as well as internationally. MPOB has played an active role in developing new technologies
which have contributed to the advancement of the Malaysian oil palm industry. A remarkable
number of more than 340 technologies including new products and services have been
launched for commercialization and adoption by industry.

A. 19 India
136
Biodiesel is now being produced locally in India for use in three-wheeler motor rickshaws.
These engines actually run on regular diesel fuel or CNG, but in the past kerosene was used
because it was far cheaper, and worked just as well. However, kerosene was dirty and wasnt
as clean-burning. Biodiesel is rapidly replacing both kerosene and diesel as a more efficient,
cheap, and clean alternative. Today plans are being chalked out to cultivate Jatropha plants
on barren land to use its oil for biodiesel production. Now it is used for Railway engines and
the plantations are recommended to plant these plants everywhere in unused areas through
government sectors. Biodiesel is being used experimentally to run state transport corporation
buses in Karnataka. University of Agriculture Sciences at Bangalore has identified many elite
lines of Jatropha Curcas and Pongamia pinnata.

The Government of India has developed an ambitious National Biodiesel Mission

to meet 20
% of the countrys diesel requirements by 2011-2012. Since the demand for edible vegetable
oil exceeds supply, the Government decided to use non-edible oil from Jatropha Curcas
oilseeds as biodiesel feedstock. The centrepiece of Indias plans for biodiesel development
and commercialization is the National Biodiesel Mission, formulated by the Planning
Commission of the Government of India. The implementation of the project consists of two
phases. In Phase I a demonstration project was to be carried out between 2003 to 2007. The
project involved the development of Jatropha oilseed nurseries, the cultivation of 400.000
hectares with Jatropha, the setting up of seed collection and Jatropha oil expression centres,
and the installation of an 80.000 Mt/year transesterification to produce biodiesel from
Jatropha oil. Phase II will consist of a self sustaining expansion of the programme leading to
the production of biodiesel to meet 20 % of the countrys diesel requirements by 2011-12.

Large scale activities have been initiated quite recently. For example, large-scale plantations
have been initiated in North-East India and Jharkhand through a Memorandum of
Understanding signed between D1 Oils and Williamson Magor. The hilly areas of the North-
East are ideal for growing this hardy, low-maintenance plant. Naturol Bioenergy Limited

(NBL), a joint venture with Energea Gmbh (Austria) and Fe Clean Energy (United States)
plans on building a 300 t/day (about 90.000 t/year) plant in Kakinada, Andhra Pradesh.
Southern Online Biotechnologies

plans on installing a 30 t/day (9,000 t/year) biodiesel plant
also in Andhra Pradesh. Feedstock will be Pogammia Pinnata/Jatropha grown on 1,000
hectares of wasteland. This cultivation will yield 6.000 tons of vegetable oil. About 9.500
137
tons of vegetable oil is required for 9.000 tons of biodiesel, and the difference will be made
up by animal fat (UNCTAD, 2006).

A. 20 Israel
Biodiesel is not yet sold on the market, it is been produced in two small-scale experiments.
The amounts produced in these experiments are up to 10,000 liters a month. The lack of
production of biodiesel in Israel is contrary with the Research and Development abilities of
the country, for Israel is a center of development for agriculture technologies. The Israel
North Recycle Group (INRG) is forecasting much progress in the biodiesel industry and has
entered into consumption agreements with municipal bodies, as part of the wider view of the
municipalities on the subject.

A. 21 Australia
The Fuel Standard (Biodiesel) Determination 2003 was signed by the Minister for the
Environment and Heritage on 18 September. The determination sets out the physical and
chemical parameters of biodiesel standard. It also sets out the associated test methods that the
Government will use to determine compliance. Biodiesel subsidies are to be phased out by
2011, after the passing of the Fuel Tax Bill 2006. Australian Farmers Fuel (SAFF) has been
retailing B100 to the public in South Australia since 2001 and now also sells B20 (marketed
as Premium Diesel) at some 52 service stations across 4 states. All of the metropolitan
trains and most of the metropolitan buses in Adelaide (capital of South Australia) operate on
a B5 blend. The South Australian Government has stated that it will soon move to B20 or
possibly higher blends. Several councils (local Governments) across Australia are using B20
(including Townsville City Council, Adelaide City Council, Sydney City Council and
Newcastle City Council). In February of 2005 the first retail outlet for biodiesel opened in the
Sydney suburb of Marrickville. It offers B20 and B50 blends to the general public, and caters
to qualified fleets wishing to utilize B100. 2006 saw the first rollout of biodiesel by a service
station network. Gull, a Western Australian based company, introduced B20 biodiesel to
several Gull service stations on April 3 which has since expanded to a total of 21 sites of
purchase. In addition, pure biodiesel (B100) along with other blends can be purchased in
bulk. Gull is also involved with the Western Australian Government to provide B5 biodiesel
for use in Transperth buses. Eventually the fleet will be provided with B10 or B20 blends.
Currently seven percent of Transperths bus fleet is running biodiesel. More recent news is
the launch of reeFUEL biodiesel in Sustainable Townsville, North Queensland. reeFUEL
138
sells only B100 and as of September, 2006, was selling 50,000 litres per week into a
community of about 160,000. This is believed to be the highest penetration of biodiesel per
capita in Australia (www.reefuel.com). Biodiesel industry stakeholders are shown in Table A-
12. The status of various biodiesel plants as at 2007 is shown in Table A-13.

Table A-12: Biodiesel industry stakeholders in Australia


Table A-13: Status of biodiesel plants in Australia
Biodiesel
plant
Location Owner capacity
(ML*)

feedstock Status
ABG Narangba
Biodiesel Plant

S-E Qld

ABG Limited

160

Tallow; used cooking
oil;
vegetable oil
Not in
production

Eco Tech
Biodiesel Plant
S-E Qld

Gull Group

30 Tallow; used cooking
oil
In production

BP Bulwer
Island Refinery
Brisbane, Qld

BP Australia
Ltd
Tallow (testing
hydrogenation
Not in
production
139
process)
BIA Biodiesel
Plant

Maitland,
NSW

Biodiesel
Industries
Australia
20 Used cooking oil;
vegetable oil
In production

BPL Biodiesel
Plant

Wodonga, Vic

Biodiesel
Producers
Limited
60 Tallow; used cooking
oil

In production
Laverton
Biomax Plant
Melbourne,
Vic
Smorgon Fuels
Pty Ltd
30 Tallow; used cooking
oil; vegetable oil

Moama
Biodiesel Plant
NSW/Vic
border
Future Fuels

30 Tallow; used cooking
oil
In production

ARF Largs
Plant

Adelaide, SA

Australian
Renewable
Fuels Limited

45 Tallow

Not in
production

ARF Picton
Plant

S-WA

Australian
Renewable
Fuels Limited
45 Tallow

Not in
production

Natural Fuel
Biodiesel Plant
Darwin, NT

Natural Fuel
Limited
138 Palm Oil, soya &
jatropa oil
Not in
production
ML Million Litres (as of 1 January 2008)


A. 22 Brazil
Brazil opened a commercial biodiesel plant in March 2005. Speaking at the opening of plant,
the Brazilian president said the National Biodiesel Program, besides producing fuel, will
produce a lot of social inclusion, most of all in the poorest regions. The plant, located in
Cssia, in the state of Minas Gerais, has a capacity to produce 12 million litres (3.17 million
gallons) of biodiesel per year and will be the first of many. Feedstocks can be a variety of
sunflower seeds, soybeans, or castor beans. The finished product will be currently a blend of
gas oil with 2% biodiesel and, after 2011, 5% biodiesel, both usable in unmodified diesel
engines. As of 2005, there were 3 refineries and 7 that are planned to open. These three
factories were capable of producing 45.6 million litres of biodiesel per year. In Brazil, castor
bean is the best option to make biodiesel, because it is easier to plant and costs less than
soybean, sunflower or other seeds.

Petrobras (the Brazilian national petroleum Company) launched an innovative system,
making biodiesel (called H-Bio) from the petroleum refinery. Petrobras will construct a new
193 million gallon a year plant for biodiesel production, and will make biodiesel directly
from seeds rather than oils. The new technology is being tested at a pilot plant in Guamare,
Rio Grande do Norte state. Petrobras is already constructing three smaller plants that will
140
each have a capacity of 15 million gallons a year. Total project investment for the four plants
will be $300 million.

A. 23 United States
Biodiesel has found diverse application in several states in the US. From military fleets to
long-haul truckers, farm equipment to heating oil supplies, and of course, in diesel vehicles
gas tanks, biodiesel is cropping up all over the U.S. as the clean fuel made primarily from
soybeans. From 2004 to 2006, biodiesel consumption in the U.S. grew from 25 million
gallons per year to 225 million gallons. Biodiesel is commercially available in most oilseed-
producing states in the United States. As of 2005, it is somewhat more expensive than fossil
diesel, though it is still commonly produced in relatively small quantities (in comparison to
petroleum products and ethanol). Many farmers who raise oilseeds use a biodiesel blend in
tractors and equipment as a matter of policy, to foster production of biodiesel and raise public
awareness. It is sometimes easier to find biodiesel in rural areas than in cities. Similarly,
some agribusinesses and others with ties to oilseed farming use biodiesel for public relations
reasons. As of 2003 some tax credits are available in the U.S. for using biodiesel. In 2004
almost 30 million US gallons (110,000 m) of commercially produced biodiesel were sold in
the U.S., up from less than 0.1 million US gallons (380 m) in 1998. Projections for 2005
were 75 million gallons produced from 45 factories and 150 million gallons (570 million
liters). Due to increasing pollution control requirements and tax relief, the U.S. market is
expected to grow to 1 or 2 billion US gallons (4,000,000 to 8,000,000 m) by 2010. The price
of biodiesel in the United States has come down from an average $3.50 per US gallon
($0.92/l) in 1997 to $1.85 per US gallon ($0.49/l) in 2002. This appears economically viable
with current petrodiesel prices, which as of 09/19/05 varied from $2.648 to $3.06.

Soybeans are not a very efficient crop solely for the production of biodiesel, but their
common use in the United States for food products has led to soybean biodiesel becoming
the primary source for biodiesel in that country. Soybean producers have lobbied to increase
awareness of soybean biodiesel, expanding the market for their product. A pilot project in
Unalaska/Dutch Harbor, Alaska is producing fish oil biodiesel from the local fish processing
industry in conjunction with the University of Alaska Fairbanks. It is rarely economic to ship
the fish oil elsewhere and Alaskan communities are heavily dependent on diesel power
generation. The local factories project 3.5 million tonnes of fish oil annually. In March 2002,
141
the Minnesota State Legislature passed a bill which mandated that all diesel sold in the state
must contain at least 2% biodiesel. The requirement took effect on June 30, 2005.

In March 2006, Washington State became the second state to pass a 2% biodiesel mandate,
with a start-date set for December 1, 2008. In 2005, U.S. entertainer Willie Nelson was
selling B20 biodiesel in four states under the name BioWillie. By late 2005 it was available
at 13 gas stations and truck stops (mainly in Texas). Most purchasers were truck drivers. It
was also used to fuel the buses and trucks for Mr. Nelson's tours as well as his personal
automobiles. In February of 2006, a team of high school students showed a sports-economy
car, fueled by soybean bio-diesel. The car can go from zero to 60 in four seconds and gets
more than 50 miles to the gallon. The car, called the K-1 Attack, combines a Volkswagen-
built diesel engine for the rear wheels, and an AC Propulsion electric drive for the front
wheels for short bursts of speed.

Table A-14. Biodiesel industry stakeholders in the United States

142
143


A. 24 Canada
According to the Canadian Ministry for Agriculture and Rural Development, Biodiesel
development in Canada is in early stages of development and commercialization compared to
Europe and the United States. Commercial companies include Rothsay/Laurenco (Quebec),
Milligan Bio-Tech (Saskatchewan), Agri-Green biodiesel (British Columbia) and Ocean
Nutrition (Nova Scotia). Current biodiesel production and utilization has targeted the
transportation sector (fleet and public transit demonstration). Considering the forecasted
consumption of diesel fuel in 2010, a 5% renewable content in diesel fuel nationally would
require the production of 1.5 billion litres of biodiesel (Canola Council of Canada). This
would require about on-fifth of Canadas projected canola production in 2010.

Rothsay of Ville Ste Catherine, Quebec produces 35.000 m of biodiesel per year. The
Province of Nova Scotia uses biodiesel in some public buildings for heating as well as (in
more isolated cases) for public transportation. Halifax Regional Municipality has converted
its bus fleet to biodiesel, with a future demand of 7.500 m of B20 (20% biodiesel fuel
mixture) to B50reducing biodiesel content in low temperatures to avoid gelation issues
and 3.000 m split between B20 and B100 for building heat. The municipality forecasts a
greenhouse gas reduction of over 9.000 tonnes CO
2
equivalents (4.250 tonnes from fleet use
and 5.000 tonnes from building heating) if fully implemented. Private sector uptake is slower
but not unheard of possibly due to a lack of price differential with petroleum fuel and a lack
of federal and provincial tax rebating. Ocean Nutrition Canada produces 6 million gallons
(23.000 m) of fatty acid ethyl esters annually as a byproduct of its Omega-3 fatty acid
processing. This surplus is used by Wilson Fuels to produce blended biodiesel for use as
transportation and heating fuel. Wilson Fuels have also opened a biodiesel station in
144
Moncton, New Brunswick. In Ontario, Biox Corporation of Oakville is building a biodiesel
processing plant in the Hamilton harbour industrial lands, due for completion in the first half
of 2006. There are also a few retail filling stations selling biodiesel to motorists in Toronto
and Unionville. Manitoba has seen a rush of building in biodiesel plants in 2005 and 2006,
starting in June 2005 with Bifrost Bio-Diesel in Arborg, Manitoba. In addition, biodiesel is
made by individuals and farmers for personal use. BioFuel Canada Ltd has small scale
affordable plants for farmers and off-road users. Along Canada's western-most coast, in
British Columbia the cooperative association proves a successful structure for micro-
economy-of-scale biodiesel production reaching the end-user. Vancouver Biodiesel Co-op,
WISE Energy and Island Biodiesel Co-op are notable examples.

Both James Richardson International (JRI) and Louis Dreyfus two companies that deal in
agricultural produce in Canada recently announced their intentions to build canola crushing
facilities in Yorkton, Saskatchewan. The $100 million JRI facility will have the capacity to
crush 840,000 tonnes of canola annually while the Louis Drayfus operation ($90 million) will
process 850,000 tonnes of oilseed a year. Both companies project to start crushing in late
2008. Canada will take advantage of a variety of feedstocks including soybean, canola, waste
greases and animal fats. The Canadian rendering industry produces about 400,000 tonnes of
animal and yellow grease annually for the oleochemical and animal feed industry. In
addition, Canada is the worlds largest exporter of canola with ideal growing conditions in
Western Canada. Some biodiesel plants in Canada are presented in Table A-15.

Table A-15: Biodiesel Plants in Canada
Plant Name City Province Feedstock Capacity
Milligan
BioTech
Foam Lake SK Canola Oil 1.000.000 l
BIOX Hamilton ON Tallow 66.000.000 l
Rothsay Montreal QC Animal Fats/Yellow Grease 30.000.000 l
Canadian
Bioenergy *
Sturgeon County AB Canola 225.000.000 l
*Plant currently under construction.


145
A. 25 Costa Rica
Costa Rica is a large producer of crude palm oil and this has spurred interest in biodiesel.
Currently several small biodiesel production projects are starting in the country. There are
also biodiesel reactor manufacturers in Costa Rica which provide equipment to the Central
American and Caribbean region.

146
B. Bioethanol

B.1 Introduction
The use of ethanol as fuel (bioethanol) has far surpassed the use of biodiesel as global
production of fuel ethanol soared above 13 billion US gallons in 2007 (RFA, 2008). Global
leaders in the industry, US and Brazil, continue to expand their production capacities.
Between January 2007 and January 2008, 29 new ethanol biorefineries were commissioned
in the US alone, increasing production capacity from 5.493 million US gallons to 7.888
million. There were huge increases in production capacity in Brazil as well. Together, the US
and Brazil produced about 88% of global fuel ethanol productions in 2007 as shown in
Figure B-1. With current expansion programmes in both Brazil and US, it is expected that the
two countries will lead fuel ethanol productions for a long time yet.

Source: Based on fi gures from RFA, 2008
Brazil
38.3%
Canada
1.6%
Rest of World
2.4%
China
3.7%
EU
4.4%
USA
49.6%


Figure B-1. Top five fuel ethanol producers worldwide, 2007 (% of global production)

Brazil is the largest exporter of fuel ethanol, exporting mostly to the EU and the US. On the
other hand and interestingly the US is the largest importer, importing mainly from the
South American Region, notwithstanding its huge production figures. Feedstock for the
production of fuel ethanol has mainly been sugarcane as in the case of Brazil, and corn as in
147
the case of the US. Other feedstocks include wheat, cassava sugar beet, grain sorghum, sweet
potatoes, etc. Aside the US and Brazil, other leading producing countries/regions include
China, India, Canada and the EU. Several countries and regional bodies have targets for the
gradual incorporation of ethanol fuel blends into their transportation fuel market. As can be
seen in Table B-1, some countries, such as Brazil, are more ambitious in the targets they have
because of long experiences in the production and use of ethanol fuel. Other countries are
being very cautious because of the potential for fuel ethanol feedstocks to compete with food
production: infact several organisations have reportedly linked the current global food crisis
to the production of biofuels in general, but more to bioethanol because of the use of corn as
feedstock in several countries, including the US. South Africa, in its recently unveiled
Biofuels Industry Strategy has pledged to stay away from the use of corn as feedstock for
fuel ethanol. Annual fuel ethanol productions by the major producers are presented in Table
B-2. It is evident from Table B-2 that whilst the US and Brazil have had enormous growth in
production capacity in the past few years (in the case of the US, about 100% growth in
production between 2004 and 2007), production in the other countries have remained fairly
stable.

Table B-1. Global ethanol blending requirements

Country/Region Blending Requirements
Brazil All gasoline must contain between 20 and 25% anhydrous ethanol.
Currently, the mandate is 23%.

Canada By 2010, 5% of all motor vehicle fuel must be ethanol or biodiesel.

France Set target rates for incorporation of biofuels into fossil fuels (by energy
content). Calls for 5.75% in 2008, increasing to 10% in 2010.

Germany Mandates 8% biofuels in motor fuels by 2015, 3.6% coming from ethanol.

Lithuania Gasoline must contain 7-15% ETBE. The ETBE must be 47% ethanol.

Poland Mandatory National Biofuel Goal Indicators calling for biofuels to
represent a set percentage of total transportation fuel use. 2008s standard
is 3.45%, on an energy content basis.

Argentina Requires the use of 5% ethanol blends by 2010.

148
Thailand Gasoline in Bangkok must be blended with 10% ethanol.

India Requires 5% ethanol in all gasoline.

China Five Chinese provinces require 10% ethanol blends Heilongjian, Jilin,
Liaoning, Anhui, and Henan.

The Philippines Requires 5% ethanol blends in gasoline beginning in 2008. The
requirement expands to 10% in 2010.

Bolivia Expanding ethanol blends to 25% over the next five years. Current blend
levels are at 10%.

Colombia Requires 10% ethanol blends in cities with populations over 500,000.

Venezuela Phasing in 10% ethanol blending requirement.

South Africa Biofuels to account for 2% of total fuel production by 2013 but will
exclude staple maize as a feedstock.
Source: RFA, 2008

In this section an overview of various country/regional programmes on bioethanol are
presented. Specifically, the section addresses issues that include: current production
capacities; plans for further expansion and addition of new plants; production and
consumption targets in place for bioethanol use and projects outlined to help meet the targets.
Because countries in the EU work closely together on biofuel programmes and due to the fact
that they have regional biofuel targets in place quite apart from the various country targets,
the section on Europe will begin with an introduction of the general EU regional programme
after which specific country programmes will be presented.

Table B-2. World Ethanol/Fuel Ethanol Production (millions US gallons)

Annual Fuel Ethanol
Production by
Country/Region (2007)
Annual Ethanol Production (All Grades)
by Country/Region (2004-2006)
Country 2007 2006 2005 2004
US 6498 4855 4264 3535
Brazil 5019 4491 4227 3989
EU 570
149
China 486 1017 1004 964
Canada 211 153 61 61
Thailand 79
Colombia 79
India 52 502 449 462
Central America 39
Australia 26 39 33 33
Turkey 15
Pakistan 9 24 24 26
Peru 8
Paraguay 5
France 251 240 219
Russia 171 198 198
South Africa 102 103 110
UK 74 92 106
Germany 202 114 71
Ukraine 71 65 66
Source: RFA, 2008


B.2 European Union
Since 1993, there has been a gradual growth in the production and use of bioethanol in the
EU. Production went up 9 times from 60 million litres in 1993 to 526 million litres in 2004
(Vierhout, 2005). Available data indicates that 2007 fuel ethanol production for the EU stood
at 570,3 million US gallons (RFA, 2008). Consumption is mainly in Germany, Sweden,
France and Spain, but also in Poland, UK, The Netherlands, Hungary, etc. About 90% of the
bioethanol consumed in the EU in 2006 was produced in the region with Germany producing
70% of its consumption, Spain 60% and Sweden 50%. In Sweden there are over 800 E85
filling stations and in France over 130 E85 service stations with 550 more under
construction.

In order to promote the market breakthrough for bioethanol in the EU, the Bioethanol for
Sustainable Transport (BEST), supported by the European Commission was launched as a
150
demonstration project. The duration of the project is from January 2006 to December 2009.
The BEST project has inspired the sale of Flexible Fuel Vehicles (FFVs)
1
in most EU
countries.

Table B-3. Overview of BEST demonstration activities
Site Flexi Fuel
Cars
E95
Buses
Low Blend
Petrol
Low Blend
Diesel
E85 Fuel
Pumps
E95 Fuel
Stations
Stockholm 4.250 127 25 5
Biofuel Region 2.500 15 E-diesel 55 3
Rotterdam 2.955 3 E10 E-diesel 12 1
Somerset 300 E10 5
Basque Country 200 E10 4
Nanyang 100 4 30 1
Madrid 25 5 1 1
La Spezia 10 3 E-diesel 2 1
Sao Paulo 3 1
Total 10.532 160 135 13

Within the programme, it is expected that over 10.000 FFVs, 160 bioethanol buses and 150
fuelling stations (E85 and E95) will be demonstrated in 10 strategically chosen European
cities and regions. Table B-3 gives an overview of the number of vehicles and fuel stations
that are expected to be implemented in these target regions. Table B-4 gives an overview of
some of the bioethanol industry stakeholders in Europe

Table B-4. Bioethanol industry stakeholders in Europe

Name Web Address Logo Description

1
Flexible fuel vehicles (FFVs) are capable of operating on gasoline, E85, or a mixture of both. Unlike natural
gas and propane driven biofuel vehicles, flexible fuel vehicles contain one fueling system, which is made up of
ethanol compatible components and is set to accommodate the higher oxygen content of E85. Other than their
fuelling capability and ethanol compatible components, FFVs are similar to their conventional gasoline
counterparts. Their power, acceleration, payload, and cruise speed are comparable whether running on ethanol
or gasoline.


151
Sdzucker
Bioethanol GmbH
http://www.bio-pro.de/en/
index.html

Produces 260.000 cubic metres

of bioethanol
annually as well as 260.000 tons of high-value
protein feed (ProtiGrain), yielding 30
million kWh of electricity.

Bio-Ethanol
Rotterdam (BER)
http://www.ber-rotterdam.com

Established in March 2006. Expected to
produce bioethanol without making use of
fossil fuels and to capture all CO
2
from the
entire production process.

Lyondell Chemie
Nederland BV
wiebe.schipper@lyondell.com

Supplier of petrol components to well-known
oil companies, as well as to independent fuel
suppliers, in many European countries.

Futura Petroleum
Limited Harvest
Biofuels BV
www.futura-petroleum.com

Harvest Biofuels will use around 375,000 tons
of grain per year to produce 110.000 tons of
bioethanol. Operations to begin in 2009.

Royal Nedalco B.V.
www.nedalco.nl

Subsidiary company of the Dutch sugar
producer Cosun

SAAB
http://www.saabbiopower.co.uk/

Global vehicle manufacturer.

British Sugar
http://www.britishsugar.co.uk/

British Sugar is the leading supplier of sugar
to the UK market.

Green Spirit
http://www.greenspiritfuels.com

Launched in June 2005 by grain trader Wessex
Grain to produce and market bioethanol


B.3 Germany
In February 2006, E85, a fuel consisting of 85 percent ethanol and 15 percent petrol, was
launched on the German market by Sdzucker Bioethanol GmbH. Cargill, an international
agricultural company in the UK planned to commence the construction of a bioethanol plant
in Germany in autumn 2007. The plant which is expected to be completed in spring 2009,
will lead to the production of up to one million hectolitres (100.000 m
3
) of bioethanol per
year. The German government has mandated an 8% biofuels in energy content by 2015, and
it is expected that bioethanol will contribute 3,6%. Germany has exempted bioethanol based
152
on hemicelluloses from any tax until 2012. The city of Brandenburg in Germany is one of the
participants of the BEST demonstration project and has already started selling E85 from
some fuel stations and has as well acquired a fleet of FFVs. The BEST project in Germany is
being supported by a bioethanol producer, Nordbrandenburger BioEnergie; vehicle
manufacturing giant, Ford and the ministries of Economies, Rural Development,
Environment and Consumer Protection. In the meantime, the German government has
cancelled plans to increase the compulsory ratio of bioethanol in regular petrol to 10% in
2009 from 5%, due to vehicle incompatibility.

B.4 The Netherlands
The Netherlands in an effort to meet the EU requirements of biofuels in transport fuels has
began the gradual phasing out of regular gasoline and is making efforts to make blended
fuels available to all consumers. Most of the planned initiatives are still at the construction
stages and include an annual production of at least 110.000 tons of bioethanol per year from
350,000 tons of wheat by BER BV; Bio-ETBE production capacity of at least 600.000 tons
per year by Lyondell Chemie Nederland BV, among others. Producer Abengoa Bioenergy
Netherlands has begun building a bioethanol facility in Rotterdam Europoort, which is
expected to be operational before the close of 2009. The facility is expected to process 1,25
million tonnes of cereals, and has the capacity to produce 480 million litres of bioethanol a
year. The first E85 fuel station was opened in the city of Rotterdam in June 2006 as part of
the BEST programme. Currently a few gas stations in the country sell E85 to the public. It is
expected that by the end of December 2009, over 4.000 cars and buses will be running on
E85 and E95 in Rotterdam alone. There are currently no incentives (such as tax exemption)
to the sale of bioethanol in The Netherlands making the cost to the consumer quite high and
militating against its usage. It is estimated that about 12 million dry tons of ligno-cellulosic
biomass could be realised each year for the production of approximately 2,5 million tons of
bioethanol annually in The Netherlands.

B.5 Sweden
Sweden is leading Europe in encouraging the use of bioethanol as an eco-friendly renewable
fuel source. E85 already accounts for 2,5% of fuel for Swedish road transport, by far the
highest proportion in any European market. From 1994 to 2004 the number of filling stations
for bioethanol in Sweden grew from 1 to 100. There has since been a phenomenal growth
following an act of parliament that requires all Swedish gas filling stations to offer at least
153
one alternative fuel. This has led to the situation where every fifth car in Stockholm now
drives at least partially on alternative fuels, mostly ethanol. The number of bioethanol
stations in Europe is highest in Sweden, with over 800 stations. Meanwhile new bioethanol
pumps are opening at an average rate of between five and ten a week. Stockholm is expected
to introduce a fleet of Swedish-made electric hybrid buses in its public transport system on a
trial basis in 2008. These buses will use ethanol-powered internal-combustion engines and
electric motors the vehicles diesel engines will use ethanol. The Swedish bioethanol
programme has enjoyed tremendous government support: government legislation requires
that bioethanol E85 sell for 25% less at the pump than petrol and 50% of government agency
fleet cars must be eco-friendly vehicles. Sweden is currently Europes third largest producer
of bioethanol, but still imports large amounts from Brazil and Spain.

B.6 UK
The first bioethanol pump in the UK was opened by Morrisons, a supermarket chain in 2006.
British Sugar, a leading agro-processing company in the UK began production of
bioethanol in September 2007 making it the first company to produce commercial bioethanol
in the UK. The company produces up to 55.000 tonnes (70 million litres) of bioethanol every
year and uses around 110.000 tonnes of sugar. This is equivalent to 650.000 tonnes of sugar
beet. Beet supplied to British Sugar for bioethanol production is grown on existing farm land.
Another plant to be constructed by Green Spirit at Henstridge in Somerset is expected to start
production before the close of 2008. Annual production is expected to be 105.000 tonnes.

B.7 Italy
Italian chemical group Mossi & Ghisolfi (M&G) plans to build a 200.000 tonne bioethanol
plant in the northern region of Piedmont, Italy. The investment of around 100 million will
build the largest bioethanol plant in Italy by 2009 and a further 120 million will be pumped
into research to convert the plant to process cellulose feedstock. The new plant, which will
help Italy meet its bioethanol target of 1 million tonnes by 2010, would initially use 600.000
tonnes of corn as feedstock. The plant aims to cover 60% of its feedstock needs with local
supplies. M&G started a research project in 2007 aimed at converting the future bioethanol
plant from corn to fibre sorghum or common cane, which use less fertiliser and does not
require irrigation. The group aims to launch a demonstration plant by 2012, which will have
a 20.000 tonne annual output (Biofuel News, 2008).

154
B.8 France
Until 2004 it was not allowed to blend the bioethanol directly in conventional fuels.
Bioethanol was converted into ETBE before it was blended in gasoline, to a maximum of
15%. Since 2004 however, direct blending of ethanol is allowed, but so far it occurs only on
a very limited scale (van Walwijk, 2005). The French Government has meanwhile set up an
ambitious plan designed to quickly implement a bioethanol E85 programme across France,
encompassing a number of critical industry stakeholders. A charter to this effect was signed
at the end of 2006. As part of the plan, the Government has decided that at least 500
bioethanol E85 pumps should be installed in Frances biggest cities and at its motorway
service stations in 2007, with the major objective to ease Frances dependency on fossil fuels.

Other objectives set out in the charter include:
The French government to define the normative and technical framework necessary
for launching E85 in France.
Vehicle manufacturers to offer a diversified and high performance range of flex-fuel
vehicles.
Fuel retailers to put in place an E85 bioethanol retail distribution network throughout
France.
Producers to develop a competitive and sustainable bioethanol production system.
All of the partners to put in place the necessary economic and tax conditions to enable
E85 bioethanol to be competitive with fossil fuels and, in particular, for the French
government to assure the fiscal conditions necessary for it to be competitive.

B.9 Rest of Europe
Bulgaria: Spanish biofuels producer GreenFuel has announced it will invest $71,8 million in
a biodiesel and bioethanol plant in Bulgaria. Construction of the plant, which will have an
annual capacity of 110.000 tonnes of biodiesel and 60 million tonnes of bioethanol, was
expected to begin in early 2008 in Pleven near the Danube river. Raw material will be
sourced from local farmers.

Hungary: A group of Hungarian, Swiss, Austrian and German investors are planning to build
a 84 million bioethanol plant in Oroshza, southeast Hungary. The plant is expected to
begin operation in the third quarter of 2009 and will have an annual capacity of 100.000 m
3

of ethanol, 18MW of electricity and 60 million m
3
of biogas from approximately 280,000
155
tonnes of grain, mainly corn. Magyar Bioenergetikai (Mabio), a Hungarian bioenergy
company, also announced plans to launch a 190 million investment to build two bioethanol
plants in 2007. The two plants were to be built in Csabacsd and Szabadszlls, and were
expected to cost 97 million and 93 million respectively.

Russia: The Russian government will support the construction of 30 new ethanol plants to
produce biofuels, a programme that begins in 2008 and also includes the upgrading of
existing plants. Russia will aim to ultimately produce 2 million tonnes of ethanol a year and
will be relying on such feedstock as timber waste and saw dust. The production will initially
be for export.

Austria: An Austrian Bioethanol Directive, which came into effect on 1 October 2007, entails
tax breaks for high bioethanol concentrations (85%) in gasoline mixtures. AGRANA (a sugar
and starch producing company), took advantage of this directive and set about building the
first industrial-scale bioethanol production plant in Austria. The plant is located in
Pischelsdorf and has a capacity of up to 240,000 m
3
a year. Corn, wheat and sugar beet will
serve as the raw material of the plant.

Czech Republic: The Czech Republic has announced an $18 million second generation
biofuel project in Lovosice, Bohemia. The plant, which is expected to begin operation in
2012, will be funded by the government and developed by state-oil company Cepro.

B.10 United States
As of 2006, the United States (US) produced and consumed more fuel ethanol fuel than any
other country in the world. The production of fuel ethanol soared to 6,5 billion gallons in
2007, a 32% increase from the 4,9 billion gallons of ethanol (all grades) produced in 2006.
This production volume is reported to have displaced the need for 228 million barrels of oil
which was more oil than the US imported from Iraq and nearly half that from Venezuela in
2007. The displacement of 228 million barrels of oil in 2007 saved Americans $16.5 billion,
an average of $45 million a day. As at January 2008, there were 139 ethanol bio-refineries in
production and 61 plants under construction. The Renewable Fuels Association (RFA, 2008)
expects that once all of the new construction currently underway is complete, the US ethanol
industry will be able to supply more than 13 billion gallons of ethanol, representing nearly
10% of the nations gasoline demand. Currently, the major feedstock for US fuel ethanol
156
production is corn. In May 2008, the first cellulosic ethanol plant was commissioned by
Verenium Corporation in Jennings, Louisiana. The new facility is expected to produce 1,4
million gallons of next-generation ethanol per year from agricultural waste left over from
sugarcane processing.

Ethanol is currently blended into more than 50% of the gasoline sold in the US, the majority
as E10. The use of E85 has begun gradually and some 1.400 stations in the US offer E85. As
of January 2008, three states Missouri, Minnesota and Hawaii required ethanol to be
blended with gasoline motor fuel. In some states, such as California, Minnesota, Missouri
and Texas, ethanol is blended in every gallon of gasoline sold. Many cities are also required
to use an ethanol blend due to non-attainment of federal air quality goals. There are more
than 6 million FFVs in the US capable of utilising E85. The industry is enjoying some good
measure of infrastructural development to accommodate greater volumes of ethanol.
Development in the areas of transport, storage and higher blends are well on course. In 2005,
the US government mandated the use of 7,5 billion gallons of biofuels per year by 2012.
President George W. Bush in his latest State of the Union address, called on the country to
produce 35 billion gallons of renewable fuel a year by 2017. Beside the huge production
figures, the US remains the main importer of fuel ethanol, accounting for 31% of global
imports in 2005. Imports rose from 46 million gallons in 2002 to 654 million in 2006. Table
B-5 presents some of the bio-refineries in the US ethanol industry. Figure B-2 is a map of
existing bio-refineries and those under construction in the ethanol industry. The use of corn
as the predominant feedstock for fuel ethanol production in the US as opposed to sugar cane
in Brazil, has raised concerns of food shortages, especially in view of the recent global food
crisis. The price of corn rose by 23 percent in 2006 and by some 60 percent over the past two
years, largely because of the US fuel ethanol programme. Because it is the worlds largest
corn exporter, ethanol expansion programmes in the US has contributed to a decline in grain
stocks to a low level and has put upward pressure on world cereal prices (World Bank, 2007).

Table B-5. Bio-refineries and production capacity in US

Company Location Feedstock
Current
Capacity
(mgy
1
)
Under
Construction/
Expansions
(mgy
1
)
157
Abengoa Bioenergy Corp. York, NE Corn/milo 55
Aberdeen Energy* Mina, SD Corn 100
Absolute Energy, LLC* St. Ansgar, IA Corn 100
ACE Ethanol, LLC Stanley, WI Corn 41
Adkins Energy, LLC* Lena, IL Corn 40
Advanced Bioenergy Fairmont, NE Corn 100
AGP* Hastings, NE Corn 52
Agri-Energy, LLC* Luverne, MN Corn 21
Al-Corn Clean Fuel* Claremont, MN Corn 35 15
Denison, IA Corn 48 Amaizing Energy, LLC*
Atlantic, IA Corn 110
Archer Daniels Midland Decatur, IL Corn 1,070 550
Arkalon Energy, LLC Liberal, KS Corn 110
Aventine Renewable Energy, LLC Pekin, IL Corn 207 226
Badger State Ethanol, LLC* Monroe, WI Corn 48
Big River Resources, LLC* West Burlington, IA Corn 52
BioFuel International Clearfield, PA Corn 110
Blue Flint Ethanol Underwood, ND Corn 50
Bonanza Energy, LLC Garden City, KS Corn/milo 55
Bushmills Ethanol, Inc.* Atwater, MN Corn 40
Calgren Pixley, CA Corn 55
Cardinal Ethanol Harrisville, IN Corn 100
Blair, NE Corn 85 Cargill, Inc.
Eddyville, IA Corn 35
Central Indiana Ethanol, LLC Marion, IN Corn 40
Central MN Ethanol Coop* Little Falls, MN Corn 21.5
Chief Ethanol Hastings, NE Corn 62
Chippewa Valley Ethanol Co.* Benson, MN Corn 45
East Kansas Agri-Energy, LLC* Garnett, KS Corn 35
Elkhorn Valley Ethanol, LLC Norfolk, NE Corn 40
ESE Alcohol Inc. Leoti, KS Seed corn 1.5
Ethanol Grain Processors, LLC Obion, TN Corn 100
First United Ethanol, LLC (FUEL) Mitchell Co., GA Corn 100
Front Range Energy, LLC Windsor, CO Corn 40
Gateway Ethanol Pratt, KS Corn 55
Glacial Lakes Energy, LLC* Watertown, SD Corn 100
Lakota, IA Corn 95 Global Ethanol/Midwest Grain
Processors Riga, MI Corn 57
158
Golden Grain Energy, LLC* Mason City, IA Corn 110 50
Golden Triangle Energy, LLC* Craig, MO Corn 20
Grand River Distribution Cambria, WI Corn 40
Grain Processing Corp. Muscatine, IA Corn 20
Granite Falls Energy, LLC* Granite Falls, MN Corn 52
Greater Ohio Ethanol, LLC Lima, OH Corn 54
Shenandoah, IA Corn 50 Green Plains Renewable Energy
Superior, IA Corn 50
Iowa Falls, IA Corn 105
Fairbank, IA Corn 115
Menlo, IA Corn 100
Hawkeye Renewables, LLC
Shell Rock, IA Corn 110
Heartland Corn Products* Winthrop, MN Corn 100
Aberdeen, SD Corn 9 Heartland Grain Fuels, LP*
Huron, SD Corn 12 18
Heron Lake BioEnergy, LLC Heron Lake, MN Corn 50
Holt County Ethanol O'Neill, NE Corn 100
Homeland Energy New Hampton, IA Corn 100
Husker Ag, LLC* Plainview, NE Corn 26.5
Scotland, SD Corn
Prairie Horizon Agri-Energy, LLC Phillipsburg, KS Corn 40
Quad-County Corn Processors* Galva, IA Corn 27
Range Fuels Soperton, GA Wood waste 20
Red Trail Energy, LLC Richardton, ND Corn 50
Redfield Energy, LLC * Redfield, SD Corn 50
Reeve Agri-Energy Garden City, KS Corn/milo 12
Renew Energy Jefferson Junction, WI Corn 130
Siouxland Energy & Livestock
Coop*
Sioux Center, IA
Corn 60
Siouxland Ethanol, LLC Jackson, NE Corn 50
Southwest Iowa Renewable
Energy, LLC *
Council Bluffs, IA
Corn 110
Sterling Ethanol, LLC Sterling, CO Corn 42
Loudon, TN Corn 67 38 Tate & Lyle
Ft. Dodge, IA Corn 105
The Andersons Albion Ethanol
LLC
Albion, MI
Corn 55
The Andersons Clymers Ethanol, Clymers, IN Corn 110
159
LLC
The Andersons Marathon Ethanol,
LLC
Greenville, OH
Corn 110
Tharaldson Ethanol Casselton, ND Corn 110
Trenton Agri Products, LLC Trenton, NE Corn 40
United Ethanol Milton, WI Corn 52
United WI Grain Producers, LLC* Friesland, WI Corn 49
Utica Energy, LLC Oshkosh, WI Corn 48
Albert City, IA Corn
Western New York Energy, LLC Shelby, NY Corn 50
Western Plains Energy, LLC* Campus, KS Corn 45
Western Wisconsin Renewable
Energy, LLC*
Boyceville, WI
Corn 40
White Energy Hereford, TX Corn/Milo 100
Russell, KS Milo/wheat
starch
48
Plainview, TX Corn 100
Yuma Ethanol Yuma, CO Corn 40
* locally-owned

1
mgy Million Gallons per Year
Updated: April 2, 2008
Source: RFA, 2008


Figure B-2. US Ethanol bio-refinery locations
160
Source: RFA, 2008

B.11 Canada
Before the end of 2002, there were five fuel ethanol producing plants in Canada: one each in
Alberta, Saskatchewan and Manitoba and two in Ontario, with a total production capacity of
about 175 million litres of fuel ethanol per year. Several other plants were in the planning
stage, with the potential to significantly increase Canadas annual production. There are
currently over 1000 retail outlets in Canada selling ethanol-blended gasoline, with ethanol
sales totalling about 240 million litres per year. 2007 production of fuel ethanol stood at
211,3 million gallons. The Canadian government has mandated that by 2010, 5% of all motor
vehicle fuel must be ethanol or biodiesel. The governments of Ontario, Manitoba and
Saskatchewan have mandated the use of between 5 and 10% ethanol in gasoline over the next
few years. In order to achieve this, the ethanol portion of blended gasoline receives an
exemption from the federal excise tax of 10 cents per litre on gasoline. The governments of
British Columbia and Quebec have committed to exempt the ethanol portion of low-level
ethanol blends from their road taxes when an ethanol plant is built in those respective
provinces. British Columbia currently offers a road tax exemption on the ethanol portion of
E-85. The Government of Canada has announced a $4 million funding under the federal
ecoAgriculture Biofuels Capital (ecoABC) initiative for IGPC Ethanol, a subsidiary of the
Integrated Grain Processors Co-operative, in Aylmer, Ontario for a 150 million litre ethanol
plant. The plant, which is expected to be completed in November 2008, has also received
equity investment from farmers totalling close to $15,5 million. Some of the ethanol plants in
Canada are presented in Table B-6.

Table B-6. Ethanol Plants in Canada

Plant Name City Province Feedstock Capacity (l)
Permolex Red Deer AB Wheat 40.000.000
Husky Energy Lloydminster SK Wheat 130.000.000
Terra Grain Fuels* Belle Paine SK Wheat 150.000.000
Poundmaker Lanigan SK Wheat 12.000.000
NorAmera Bioenergy Weyburn SK Wheat 25.000.000
North West Bio-Energy* Unity SK Wheat 25.000.000
161
Husky Energy Minnedosa MB Wheat 10.000.000
Iogen Ottawa ON Wheat Straw 2.000.000
IGPC* Aylmer ON Corn 150.000.000
Greenfield Ethanol* Hensall ON Corn 200.000.000
Greenfield Ethanol Tiverton ON Corn 26.000.000
Greenfield Ethanol Chatham ON Corn 150.000.000
Greenfield Ethanol* Johnstown ON Corn 200.000.000
Greenfield Ethanol Varennes QC Corn 120.000.000
Collingwood Ethanol* Collingwood ON Corn 50.000.000
Suncor Energy St. Clair ON Corn 200.000.000
* Plant currently under construction
Source: Canadian Renewable Fuels Association, 2008 (www.greenfuels.org/)

B.12 Rest of North America
Jamaica: Ethanol production in Jamaica dates back to 1985, when two producing plants were
opened for export to the USA. Currently, production capacity is over 160 million gallons of
ethanol per year. The first plant commissioned has an annual capacity of 10 million gallons.
The second plant was installed in 1986 and has a capacity of 42 million gallons per year. The
Petroleum Corporation of Jamaica is planning to begin the sales of E10 ethanol blend from
its supply station in place of unleaded gasoline in the second half of 2008. Global Energy
Ventures (GEV), an oil and gas inspection company, is expected to construct a 60 million
gallon ethanol dehydration plant in Port Esquivel, St. Catherine. The plant will process some
5 million gallons of hydrous ethanol each month. The company will in the near future also
construct four 3,5 million gallon storage tanks on a 370.000 square metre plot of land.
Petrojam, a state-owned oil refinery, will expand production of ethanol by constructing a new
60 million gallon plant. The company already has 40 million gallon ethanol dehydration plant
in operation. Jamaica Ethanol Processing operates a 60 million gallon ethanol plant in East
Kingston. Jamaica exported 75.193.188 gallons of ethanol to the US in 2007.

Dominican Republic: The government of the Dominican Republic has authorised E7,5 blends
for sale across the country. In order to achieve this target, the government has offered an
incentive package for utilities for renewable energy production. Brazils Infinity Bio-Energy
and Dominican Bioethanol Boca Chica are investing up to $200 million towards the
establishment of an ethanol plant to produce ethanol from sugarcane, an abundant crop in the
162
country. Another biofuels producer Bio E Group has plans to construct two ethanol plants
that will produce 35 million gallons a year of ethanol.

El Salvador: Southridge Ethanol Inc., a US ethanol company, has plans to construct a 5
million gallon a year ethanol plant in El Salvador. The company will import ethanol from
Brazil, dehydrate it, and re-export the fuel to the US, supplementing this programme with
production at the plant from more than 4,500 acres of sugarcane cultivation.

Mexico: In 2007, the Mexican president vetoed the Mexican biofuels bill because it placed
too much emphasis on corn and sugarcane. The President instead called for a bill that placed
more emphasis on algae and cellulosic biofuels. Because of this, the countrys Agriculture
Ministry has stated that the country will focus on beets, yucca root and sorghum as
feedstocks for biofuels in 2008.

Cuba: Cuba is said to have the potential to produce between 2 billion and 3.2 billion gallons
a year of sugarcane ethanol.

B.13 Brazil
Talk of fuel ethanol and straight away Brazil comes to mind. Up until 2006, Brazil was the
largest producer of the commodity in the world. The country is a pioneer when it comes to
the use of fuel ethanol as road transport fuel, indeed the production of sugar and subsequently
ethanol has been the backbone of Brazils economy for centuries. Even though the
production of ethanol had existed in Brazil for a long time, its serious use as transportation
fuel received national attention during the oil crisis of the 1970s. The government,
encouraged by skyrocketing oil prices launched a programme to replace gasoline with
bioethanol. The comprehensive programme included state support for ethanol plant
construction, tax incentives for bioethanol-powered cars, and a massive expansion of the
ethanol fuel pump network. As a result, in the early 1980s when the rest of the world suffered
under the highest oil prices ever, almost all cars sold in Brazil ran on bioethanol. Brazil
requires that all gasoline must contain between 20 and 25% anhydrous ethanol and provides
preferential tax treatment to producers of bioethanol currently, the mandate is 23%. Total
fuel ethanol production in Brazil was over 5 billion US gallons in 2007 out of which about
200 million gallons was transported to the US alone. In all, Brazil exported over 3 billion
litres of ethanol in 2007, but this figure is expected to reach 3,91 billion litres in 2008-09.
163
Efforts are already underway to achieve this impressive target. Brenco (a Brazilian
Renewable Energy Company) has begun construction on its sugar and ethanol mill in Alto
Taquari. The facility will process 3 million tonnes of sugarcane and produce 275 million
litres of ethanol by 2009. Many other ethanol production plants are in the planning stage.
There are plans to invest $1 billion to build a 1.100km, 4 million litre a year ethanol pipeline
extending from Alto Taquari in Mato Grosso State to Santos, the countrys largest port, in
Sao Paulo state on the countrys south Atlantic seaboard. But Brazils success in the sector is
also influenced by strong internal patronage. In 2006, the number of bioethanol-powered cars
in Brazil hit the 2 million mark, and flex-fuel cars accounted for more than three-quarters of
the nations new car sales. Many of the worlds largest car producers, including General
Motors, Ford, Peugeot, Volkswagen, Fiat and Renault now have a presence in the Brazilian
flex-fuel car market. The Brazilian government plans to raise ethanol production by 40%
between 2005 and 2010 and has offered incentives, set technical standards, and invested in
supporting technologies and market promotion. Meanwhile, sugar mills in Brazils major
producing region will turn 58% of 2008s sugarcane crop yield into ethanol as against 56% in
2007 and 51% in 2005, as rising oil prices generate demand for alternative fuels. The
southern Brazilian state of Paran the countrys second largest ethanol producer and the
fourth largest sugar producer will construct 10 ethanol and sugar plants due for operation in
2010. Planted area for sugarcane is set to increase by 16,5% in the 2007/2008 crop season. In
the 2005/2006 crop season, the state of Paran ranked as the second largest Brazilian ethanol
producer with production exceeding 1 million m
3
, against 9.951 million m
3
in the state of So
Paulo.

B.14 Argentina
Argentina plans to use its enviable position as the worlds second largest exporter of corn and
as one of the major producers of sugarcane to implement a vigorous bioethanol programme.
In April 2006, the Argentine Congress approved a law aimed at promoting the use and
production of biofuels in the country. The highlights of the law include a 5% mandatory use
of biodiesel and bioethanol in all diesel oil and gasoline consumption from the beginning of
2010 and provision of strong fiscal incentives through tax exemptions for 15 years. The law
emphasises that the programme would sponsor small and medium enterprises and that
projects would be owned by local agricultural producers and be located in regional and rural
economies. The projected gasoline consumption for Argentina for 2010 is 3,5 million tons,
which means that the country should be producing about 175.000 tons of ethanol fuel by
164
2010 to be in the position to meet the projected demand for ethanol blends. Such production
demands about 540.000 tons of corn, which is just 2% of current production and 77.000
hectares of agricultural land for cultivation, also just 2% of total planted area. Total grain
production in Argentina in 2006/2007 was estimated at 95 million tons and fuel ethanol
production for 2007 was 5,2 million US gallons. On the whole, there are 23 refineries in the
country (IICA, 2007) with an installed daily capacity of 1.5 million litres of ethanol and with
an annual potential production of more than 400 million litres.

B.15 Colombia
Fuel ethanol production in Columbia at the end of 2007 was 74.9 million US gallons.
Colombian law currently mandates a 10% ethanol blend in gasoline and a 5% blend of bio-
diesel in fossil diesel by the year 2008. The existing ethanol plants, owned by sugar mills,
supplies about 70% of the amount needed to meet the governments 10% target. There are
plans to produce ethanol from beetroot and cassava to meet the remaining 30% of the target
but uncertainties in feedstock supply and investment costs are demanding some sort of
caution. The Ministry of Energy is planning some projects which are expected to be
completed and operational by early 2009. The Colombian Government tax reform in 2003
established an exemption for bio-diesel and ethanol from paying VAT, the Global Tax
impuesto global and the additional local tax sobretasa. A comprehensive programme was
launched in 2005 (Table B-7) which sought to substantially raise the production capacity of
ethanol in Colombia for local consumption and export.

Table B-7. Estimated production and consumption of ethanol in Colombia

2006 2007 2008 2009 2010
Hectares of sugarcane 43.000 66.000 108.000 146.000 193.000
Litres per day 983.000 1.670.500 2.677.200 3.548.400 4.574.000
Number of plants 6 9 15 21 28
Consumption (l/d) 1.370.000 1.370.000 1.430.000 1.510.000 2.390.000
Exports (Litres) 1.247.200 2.038.400 2.184.000
Source: Inter-American Institute for Cooperation on Agriculture, 2007

B.16 Rest of South and Central America
165
Bolivia: In July 2005, the Bolivian Government passed a law which mandates a 10-20%
ethanol blend with gasoline by 2010 using sugarcane as feedstock. The then government
planned incentives and was prepared to lead project towards realising the target. But the
countrys new leader has openly expressed, on several occasions, his lack of support for the
biofuels industry due to its possible competition with the food industry for feedstock.
Presently, there are 15 ethanol plants under construction.

Peru: Peru has established an optional ethanol blend of 7,8%. Current production and sales of
biofuels in general is in six regions and it is expected that all the regions could be producing
and marketing biofuels by 2010. Even though there are no specific financial incentives for
biofuel producers, sugarcane, the principal feedstock, enjoys a tax exemption as any other
agricultural product. Investments in biofuels were expected to reach $250 million by the end
of 2007. Grupo Romero plans to invest $40 million in the production of bioethanol, and the
Casa Grande group has 15.000 hectares of sugarcane available for the production of
bioethanol. Maple Ethanol has 10 thousand hectares for the growing of sugarcane, and plans
to invest $32 million in the production of ethanol, projecting an investment of $100 million
for the same product.

Cuba and Venezuela have agreed to co-operate on the construction of 11 ethanol plants. The
plants, which are expected to use sugarcane as a feedstock, will not only provide transport
fuel but also use the bagasse from the sugarcane to provide electricity. Venezuela has already
begun construction on 17 domestic ethanol plants.

Uruguay: Gulf Ethanol Corporation, an American company that develops ethanol pre-
processing and production technologies, began negotiations in 2007 to produce feedstocks
and build ethanol production plants in Uruguay. Uruguay is currently working on a national
strategy to begin using ethanol as a major energy source and also for export.

Paraguay: Paraguay adopted a legal framework in 1999 to encourage the blending of fossil
fuels with ethanol. That blend is currently 18%. Steps are also being taken to increase tax
incentives. The Paraguay government in March 2007 announced plans to export at least $50
million worth of biofuels within four years as part of a wider energy project that aims to
attract foreign investment. As part of the plans, the government is developing a national plan
with a goal of producing about 300 million litres of biofuels before the end of 2011.
166

Costa Rica and Trinidad & Tobago and are all major exporters of ethanol to the US.
Together with El Salvador and Jamaica, they exported a total of 230,5 million gallons of fuel
ethanol to the US in 2007. Brazil and Ecuador have signed agreements for the two countries
to jointly produce biofuels and explore for oil in Ecuador.

Costa Rica launched a pilot project in February 2006 to distribute ethanol in the Pacifico
Central region and Guanacaste. Sixty-four gas stations are offering a blend of ethanol with
gasoline, between 5% and 8% as part of the pilot study. There are plans to begin distribution
of ethanol blended gasoline in 2008 and 2009, and the demand for ethanol is expected to
increase from 88 million litres in 2006 to 153 million litres in 2018 (IICA, 2007).

B.17 China
China, the worlds third largest producer of ethanol is on the road to expanding its bioethanol
production and is requiring a 10% ethanol blend in five of its provinces. An ethanol-based
fuel pilot programme is currently ongoing in five cities in its central and north-eastern
regions in a move to create a new market for its surplus grain and reduce consumption of
petroleum. The cities include Zhengzhou, Luoyang and Nanyang in central Chinas Henan
province, and Harbin and Zhaodong in Heilongjiang province, northeast China. The
government has been putting more emphasis on ethanol fuel development rather than
biodiesel for the simple reason that China lacks feedstock resources for biodiesel production.
Chinas Renewable Energy Law which came into effect on January 1, 2006 is paving the way
for further national funding for renewable energy initiatives. Companies producing
bioethanol and other renewable energy enjoy interest support for loans and favourable tax
policies, and several regions in China are obliged to sell ethanol fuel. Fuel ethanol production
in 2007 was close to 0,5 billion gallons, most of which was produced by four major players,
presented in Table B-8. There are four government-sponsored ethanol fuel plants with total
annual capacity of over 1 million tonnes. China plans to increase production capacity to
around 3,2 billion gallons over the next ten years. More than 80% of Chinas ethanol is made
from grains including corn, wheat and rice, and also from cassava. About 10% is made from
sugar, 6% from paper pulp waste residue and the rest from ethylene by synthetic processing.
There is a strong push for the use of non-food crops as bioethanol feedstock. Hong Kong-
based J.I.C. Technology is planning to invest a total of $207 million in three renewable
energy projects across China, a pilot cellulosic ethanol plant and two wind farms, in an
167
answer to the call to move away from the use of food crops as bioethanol feedstock.
Meanwhile, the Guangxi Zhuang Autonomous Region has become the 10th Chinese region
to replace petrol with bioethanol. Petrol stations in all cities of the region began selling
bioethanol fuel in April 2008. Guangxi is the first Chinese locality to commercially produce
ethanol fuel with cassava instead of grain. The region is the largest production base for
cassava in China, accounting for more than 60% of the countrys annual production.

Table B-8. Leading producers of bioethanol in China (As of July 2006)

Producer Location Background Capacity
Heilongjiang
Huarun Ethanol

Harbin City,
Helongjiang
Province

Founded in 1996. Fuel
ethanol product line was
initiated in 2001
71 million gallons/year

Jilinn Fuel
Ethanol
Jinlin City,
Jinlin Province
Total investment
2.89 billion RMB
97 million gallons/year

Henan Tianguan
Group

Nanyang City,
Henan Province

State-owned enterprise

97 million gallons/year
Expansion to 162
million gallons possible
Anhui BBCA
Group

Suzhou City,
Anhui Province

Listed on Shenzhen stock
exchange
194 million gallons/year

Source: Chervenak, 2006

B.18 Thailand
The use of ethanol as alternative fuel in Thailand became popular in the year 2001, with
strong interest from one of the countrys MPs and a group of researchers. Meanwhile a
couple of pilot projects had been running since the late 1980s and 90s. The National Ethanol
Committee was set up as an agency comprising government and private sector personnel to
promote the use of ethanol fuel. Thailand became the first country in Asia to announce a
National Policy for both bioethanol and biodiesel in 2000 and 2001 respectively. The Thai
government has targeted a 10% ethanol in gasoline by the year 2012. As the second largest
global sugar exporter, Thailand has set a tentative ethanol production target of 1 billion litres
168
by 2010, to be used in alternative fuel blends. There are currently over 4.000 fuel stations
that sell E10 in Thailand and these are expanding. Sales increased from 5 million litres per
month in 2004 to about 50 million litres in 2005. The main feedstock for bioethanol in
Thailand are sugarcane and cassava with annual productivity of sugarcane over 75 million
tons in 2004/2005. There were seven ethanol plants as of 2006 with a combined capacity of
955.000 litres per day. Overall capacity was expected to shoot up to 2,17 million litres per
day by the end of 2007 once eight new ethanol producers began operating. Meanwhile, total
fuel ethanol production at the end of 2007 was 79,2 million US gallons.

B.19 India
In order to boost the agricultural sector and reduce environmental pollution, the Government
of India have been examining for quite some time, the supply of ethanol blended gasoline in
the country. The Government of India through its Ministry of Petroleum and Natural Gas
introduced pilot plants and trial of 5% ethanol fuel addition to gasoline. There were three
pilot projects two in Maharashtra and one in Uttar Pradesh during April and June 2001 and
these pilot projects were tasked to supply 5% blended gasoline to the retail outlets under their
respective supply areas. Following the success of the pilot programme and further R&D work
in the field, the Ministry of Petroleum & Natural Gas announced its decision to cover all the
states in the country except north-eastern states, for 5% ethanol blending from November 1,
2006. Fuel ethanol production in 2007 amounted to 52,8 million US gallons. Some of the
ethanol plants in India are listed in Table B-9.

Table B-9. Ethanol Plants in India

Plant Name Capacity in LPD*
Shetimal Sahakari Prakriya Sanstha Ltd., Herwad Fuel Ethanol
Plant
30.000
Patil Alco & Allied Industries Pvt. Ltd., Kolhapur 30.000
Precious Alco & Petro India Pvt. Ltd., Wai. 30.000
Wallams (I) Agro Products & Power Ltd., Islampur 30.000
Vamshi Exports, Uttar Pradesh 60.000
XL Telecom Ltd 150.000
Khandoba Distilleries 150.000
169
Arvind Mills
AShri Kedarnath Agro and Sugar Products Ltd. 150.000
Om Sai Industries 45.000
Astral Poly Technik Ltd. 60.000
Anantha Energy Ltd. 100.000
* LPD Litres per day

B.20 Philippines
In the Philippines, the Department of Energy (DOE) is implementing an Alternative Fuels
Programme to reduce dependence on imported oil and to provide cheaper and more
environment-friendly alternatives to fossil fuels. There are four sub-programmes under the
alternative fuels programme and one of them is the development of bioethanol. The President
signed Parliament ratification into law in 2007 which mandates a minimum 5% ethanol blend
by volume in all gasoline fuels, being distributed and sold in the country by 2009. This is
expected to rise to 10% by 2011. Following the signing of this bill, several projects are being
planned to produce enough alternative fuels to meet the demand. Two companies, FE Clean
Energy (US) and bioenergy developer and investor Bronzeaok (UK) are building what could
be the countrys first ethanol plant using sugarcane as feedstock. Philippines-based Eastern
Petroleum and Chinese Guanxi Estates are planning to invest $30 million in an ethanol plant,
capable of producing 200.000 litres of ethanol from cassava, in the southern Sarangani
province of the Philippines. The plant is expected to begin operations in 2010 and will be
using cassava as feedstock. Meanwhile the Philippine government in 2008 also signed a $30
million project agreement to expand the capacity of the countrys first bioethanol plant from
145.000 litres a day to 200.000 litres. Bronzeoak has also formed a joint venture with
Zabaleta to construct two sugarcane ethanol plants in a $147.2 million investment that will
result in 30 million gallon a year plants at Southern Bukidnon and Pampanga. Negros Green
Energy Resources is expected to begin a 37 million gallon a year sweet sorghum ethanol
plant in Negros Occidental in the Philippines. The ultimate plant capacity is estimated at 75
million gallons a year.

B.21 Rest of Asia
Indonesia: The National Biofuel Development Committee in Indonesia is expected to
propose a 1% biofuels mandate in the 2008 legislative session. The proposed mandate looks
170
to increase biofuels consumption from 7 million gallons a year to 158 million gallons a year.
By 2010, Indonesia expects to substitute 10% of fossil fuel usage with biofuels using cassava
and molasses as feedstock.

Japan: The Japanese petroleum company Nippon Oil is leading the way to boost production
of ethanol to blend into petrol in response to appeals by the government to increase its drive
to cut carbon emissions. Nippon Oil and its counterparts in the refinery industry plan to help
raise bioethanol consumption to as much as 500.000 kilolitres in oil equivalent. The increase
is more than twice the target for the year ending March 2011, of 210.000 kilolitres.

Kazakhstan will enforce laws in 2008 to regulate its biofuel industry and plans to construct
two plants in the next 2 years. Kazakhstan could produce up to 1 billion litres of bioethanol.
The country harvested a record 20,1 million tonnes of wheat in 2007.

South Korea plans to invest $21.8 million by 2010 on biofuels. The government will begin a
project to acquire the technologies for the production of biobutanol and other synthetic crude
oil from biomass, coal and natural gas. The project will involve 29 private companies,
research institutes and universities. Funding will be shared between the government and a
number of private companies. If the first-stage project is successful, the country will be able
to build test-bed facilities for the production of biobutanol, natural gas hydrate and other
clean energy by the end of 2010.

B.22 Australia
Fuel ethanol development in Australia has without doubt enjoyed tremendous government
support since its inception in 2001 when the Federal Government announced an ambitious
target of producing 350 million litres of biofuel each year by 2010, with bioethanol
accounting for over 80% of this target. Fuel ethanol produced in 2007 amounted to 26,4
million US gallons. Fuel ethanol produced in Australia enjoys a fuel tax credit against excise
of 38,143 cents per litre up until 1 July 2011, when effective fuel tax will then begin to be
applied incrementally from 2,5 cents per litre to an end cap of 12,5 cents per litre. In July
2003, a Biofuels Capital Grants Programme was introduced, which provided capital grants
totalling $38 million for new and expanded projects producing biofuels from renewable
resources. Currently there are three commercial bioethanol producers: Manildra located in
Nowra, New South Wales; CSR located in Sarina, Queensland and Rocky Point Distillery,
171
located south of Brisbane in Queensland. Manildra produces bioethanol from waste starch
whilst CSR and Rocky Point Distillery produce bioethanol from low-grade molasses. The
three plants together have a production capacity of about 160 million litres per annum. A
number of other prospective producers have projects at various stages of development. Over
200 filling stations in Queensland sell E10 to the public, with E85 expected to come on board
some time soon. The Queensland Government is providing $7,5 million in funding over three
years to encourage more service stations to change their pumps and sell bioethanol. In
partnership with the Queensland Government, the Cane growers organisation launched a
regional billboard campaign in March 2007 to promote the renewable fuels industry. The
Premier of New South Wales in 2006 announced his intention to have a mandatory inclusion
of 10% bioethanol in all petrol sold in the state by 2011, which will provide bioethanol
producers with sufficient market security and certainty. Ethanol plants in Australia existing
ones and those in preparation are presented in Table B-10.

Table B-10. Ethanol Plants in Australia

Ethanol Plant

Location Owner Feedstock Status
(at 31.12.08)
Capacity
(million litres)
Austcane
Ethanol Plant
Queensland Austcane Ltd Sugar Plant in planning stage 60
Sarina Distillery Sarina,
Queensland
CSR Ethanol Molasses In full production 32
Pinkenba Biofuel
Project
Queensland Primary Energy Sorghum Plant in planning stage 160
Rocky Point
Distillery
Brisbane,
Queensland
Heck Group Sugar, Grain Plant expansion in
planning stage
35
Dalby Bio-
Refinery
Queensland Dalby Bio-
Refinery Pty Ltd
Sorghum Commissioned set for
October 2008
80
Manildra
Ethanol Plant
Nowra, New
South Wales
Manildra Group

Starch (by-
product from
flour milling)
Plant expansion in
planning stage
120
East
Rockingham
Bioethanol
Project
West Perth,
Western
Australia
Grainol Ltd

Wheat Plant in planning stage 190
Kwinana Biofuel
Project
Western
Australia
Primary Energy Wheat and
barley
Plant in planning stage 160

Source: Modified from Biofuels Association of Australia, 2008
172

B.23 Africa
Africa produced 160 million gallons of ethanol (all grades) in 2006 out of a total global
production of 13,5 billion gallons. This represents a mere 1% of global productions. Existing
small-scale ethanol plants in Africa can be found mostly in Southern Africa and active
participants include South Africa, Malawi, Swaziland, Mauritius, Kenya and Zimbabwe.
Some large-scale projects have been lined up in several countries by international companies
hoping to take advantage of Africas rich agricultural resources. Currently most of the
ethanol produced in Africa is used for other purposes apart from transportation fuel. A close
look at global fuel ethanol production figures for 2007 (Table B-1) indicates that Africa
produces very little as compared to the rest of the world. South Africa is the largest ethanol
producer in Africa and produces close to 65% (2006 figures) of the total productions as
shown in Table B11.

Table B-11. Africa Ethanol Production (all grades) by Country (Millions of US gallons)

Country 2004 2005 2006
South Africa 110 103 102
Zimbabwe 6 5 7
Kenya 3 4 5
Swaziland 3 3 5
Mauritius 6 3 2
Egypt 8
Nigeria 8
Malawi 4
Other 20

Other commercial ethanol producing countries are Egypt, Zimbabwe and Nigeria. Ethanol
programmes that produce a blend of ethanol and gasoline for use in existing fleets of motor
vehicles have been implemented in Malawi, Zimbabwe and Kenya since the early 1980s.
Fresh programmes are springing up in several countries including South Africa, Ethiopia,
Nigeria, Sudan, Ghana, among others. Several African countries have draft biofuels strategy
documents in place, with targets for biofuels blends, but these are yet to be passed into law.
173

Ethanol Africa, a South African bioethanol company has began the construction of the
countrys first large-scale bioethanol plant in Bothaville with plans for eight more as part of a
R7 billion investment in the inland maize farming region. The plant when completed should
be capable of producing 158 million litres of bioethanol annually from 375.000 tonnes of
corn. The plant will also have the potential to produce 108.000 tonnes of animal feed. There
are concerns of feedstock acquisition however in light of recent proclamations by the South
African government to as much as possible exclude corn from its bioethanol programmes.
South Africa recently approved the final draft of its biofuels industrial framework strategy,
but has excluded maize from the production of biofuels.

Ethanol production in Zimbabwe started in 1980 with an annual production capacity of 40
million litres a year. The country began using E15 as transportation fuel but this was later
changed to E12. Recent economic depression in Zimbabwe has seriously hampered the
ethanol programme but there are plans to re-establish vigorous fuel ethanol projects in the
country.

Kenya has for sometime been using the E20 blend without any significant effect on engine
performance. An ethanol plant was constructed in the 1980s in Kenya that used surplus
molasses as feedstock. E10 became a popular fuel in those years when the plant was
operational, but was discontinued due to uncompetitive pricing of gasoil in those days,
making the ethanol programme unprofitable. The plant was revived in 2001 and has since
been doing quite well. As of March 2005, the plant was producing 30.000 litres per day of
ethanol (all grades). This figure is projected to increase to 250.000 litres per day in the near
future.

There are two ethanol plants in Malawi with combined production capacity of 18 million
litres per year. One of the plants uses molasses obtained from a nearby sugar mill as
feedstock for the production of ethanol. Malawi currently uses unleaded fuel with 10%
ethanol blend at its refineries.

In Ethiopia, of the three existing sugar factories, one produces about 8 million litres of
ethanol annually. The government is planning to expand the other two factories while a third
one is under construction. Upon completion of the expansion and construction programmes,
174
the four factories together are expected to produce about 128 million litres of ethanol
annually. The Ethiopian government has come to an agreement with the oil importing
companies to produce fuels with 5% biofuel blends for the countrys fuel retail stations.

A sugar producing company in Sudan, Kenana, has plans to import Brazilian technology for
the production of ethanol. The company intends to begin productions soon and could
potentially produce up to 70 million litres annually by 2014.

Uganda plans to produce enough ethanol for blends of up to 15% in the countrys gasoil. In
view of this, the Sugar Corporation of Uganda has requested for additional land from the
government to increase its sugar production from the current 50.000 tonnes to 100.000
tonnes per annum.

Nigerias first ethanol refinery is expected to be constructed by Global Biofuels. At full
capacity the refinery will produce 1,5 million litres of ethanol a day using sorghum as
feedstock. Plantations feeding the refinery will span seven Nigerian states, with further
expansions being discussed. The refinery has been endorsed by The Nigerian National
Petroleum Corporation (NNPC). In 2006, Nigeria awarded two oil concessions to INC
Resources after it committed $4 billion to an ethanol project in the northern state of Jigawa.

Sweden-based ethanol producer SEKAB Group is planning and preparing for the production
of ethanol in Tanzania and Mozambique. The first factories are expected to be operational by
2011 and will be followed by many new projects in the coming 30 years. SEKAB is aiming
to support and lead efforts for biofuels in these countries through the long-term development
of over 400.000 hectares of feedstock for bioenergy production.









175
Annex Bibliography:
Bhandhumanyong P. 2007. Development of ethanol as transportation fuel in Thailand. The
National Metal and Materials Technology Centre, National Science and Technology
Development Centre, Bangkok, Thailand
Biofuel News, available from http://www.biofuels-news.com/news/italy_plansplant.html
Chervenak M. Industrial biotechnology in China. Feature Commentary.
http://www.liebertonline.com/doi/pdf/10.1089/ind.2006.2.174?cookieSet=1
Darkwah L. et al. 2007. Background Paper on Biofuels Industry Development in Africa at
the AU/Brazil/UNIDO High Level Seminar on Biofuels in Africa 30th July 1st
August, 2007. Addis Ababa, Ethiopia
IICA 2007, Agroenergy and Biofuels Atlas of the Americas. Inter-American Institute for
Cooperation on Agriculture, San Jose, Costa Rica
Janssen R. et al. 2007. Bioethanol for Europe The EU Project BEST (Bioethanol for
Sustainable Transport). 15th European Biomass Conference and Exhibition 7 11
May 2007, Berlin, Germany; p.2284-2287.
Karekezi S. 2008, Biofuels in East Africa. Background Paper for the International
Conference on Scaling-up Renewables in Africa. 16-18 April 2008, Dakar, Senegal.
Karekezi S. Et al. 2008, Scaling Up Bioenergy in Africa. Thematic Background Paper for
International Conference on Renewable Energy in Africa Jointly organised by the
Government of Senegal, the African Union, German Federal Ministry of Economic
Cooperation and Development (BMZ) and the United Nations Industrial
Development Organisation (UNIDO) 16-18 April 2008, Dakar, Senegal.
RFA 2007, Ethanol Industry Outlook 2007: Building New Horizons. Renewable Fuels
Association, Washington DC.
RFA 2008, Ethanol Industry Outlook 2008: Changing the Climate. Renewable Fuels
Association, Washington DC.
UNCTAD 2006, An Assessment of the Biofuels Industry in India.
UNCTAD/DITC/TED/2006/6, United Nations Conference on Trade and
Development.
Urbanchuk J. M. et al. 2005, Economics of a Queensland Ethanol Industry. Prepared for the
Queensland Department of State Development and Innovation, Queensland,
Australia.
Van Walwijk M. 2005, Biofuels in France 1990 2005. PREMIA report, available at
http://www.premia-eu.org/public_files/D4c_France-biofuels_MvW-Sept2005.pdf.
176
Vierhout R. 2005, Challenges for the EU Bio-ethanol Fuel industry. F.O.Lichts World
Ethanol 2005, Amsterdam, 2-4 November 2005.
World Bank 2007, World Development Report 2008: Agriculture for Development. The
International Bank for Reconstruction and Development / The World Bank.



Web sites:
journeytoforever.org
www.veggieavenger.com
www.biodieselnow.com
http://www.best-europe.org/
http://www.biofuels-news.com/
http://www.frost.com/
http://www.ethanolmarketplace.com/

Europe
www.ebb-eu.org
http://www.uepa.be/home.php
www.koal2.cop.fi/leonardo/
http://ec.europa.eu/energy/res/sectors/bioenergy_en.htm

Austria
www.biodiesel.at
http://www.biofuelsassociation.com.au/

France
www.villediester.asso.fr

Germany
www.ufop.de

Italia
www.assobiodiesel.it
177

UK
www.biofuels.fsnet.co.uk/biobiz.htm
www.biodiesel.co.uk
http://www.saabbiopower.co.uk/
www.britishbioethanol.co.uk/

Canada
www.greenfuels.org

USA
www.eere.energy.gov/cleancities/afdc/
www.biodiesel.org
www.nrel.gov
www.veggieoilcoop.org
www.buolderbiodiesel.com
www.grease-works.com
www.gobiodiesel.com
www.biodieselamerica.org

Australia
www.biodiesel.org.au

Malaysia
www.mpob.gov.my

India
http://www.ethanolindia.net/ethanol_govt.html

Philippines
http://www.doe.gov.ph/AF/Biofuels.htm

Africa
http://www.biopact.com/

Você também pode gostar