You are on page 1of 122

Biomass in the

energy industry
An introduction

Biomass in the energy industry


An introduction

Supported by BP, as part of the multi-partner


Energy Sustainability Challenge, which explores the
implications for the energy industry of competing
demands for water, land and minerals.

The handbook also provides key data about crops


species and biomass types that are already in
production or are being researched for biomass.
The data includes plant characteristics, suitable
growth conditions, required inputs and agricultural
practices, co-products and alternative markets, as
well as yield and energy productivity indicators.
The handbook offers a valuable guide for policy
makers, businesses and academics on the
characteristics of major biomass crops and the
issues related to sustainable and responsible use
of biomass for energy.
Biomass in the energy industry An introduction
shows:
n

What role biomass plays in the global


energy context.
n What fundamental knowledge is required to
understand bioenergy systems.
n How biomass is converted to energy and what
technological developments are under way.
n Why it is vital to view use of biomass for energy
from socio-economic, environmental and
political perspectives.
n What is the potential for bioenergy and how this
potential can be realized.
n Where can biomass feedstocks be grown and
what are the key characteristics of biomass
crops already in production or being researched
for biomass.
Published by BP p.l.c.
2014 BP p.l.c.
ISBN 978-0-9928387-1-3

9 780992 838713

Biomass in the energy industry An introduction

Biomass in the energy industry An introduction


is a study that provides contextual knowledge
required for assessing the potentials and issues
of using biomass for energy. The book is based
on literature research and review by colleagues
at the Energy Biosciences Institute
(www.energybiosciencesinstitute.org) and
is part of the Energy Sustainability Challenge
(www.bp.com/energysustainabilitychallenge)
series of handbooks. This book addresses the need
for having a holistic view of the benefits and risks
associated with bioenergy by studying the subject
from agricultural, energy, environmental,
technological, socio-economic and political
perspectives. The book emphasizes that realizing
the potential of biomass energy as a major player
in carbon emissions reduction needs careful
consideration of environmental aspects and
competing demands of food, water, energy and
other resources. Clear and consistent supportive
policies are also required to facilitate significant
financial investments for developing biomass
conversion technologies and improving
performance of biomass crops.

Plant
Planttypes
types
BP Biomass Handbook
Table 3.1 (20 December 2013)
Draft produced by ON Communication

Perennial
Plant characteristics icons in chapter 6Perennial

Herbaceous
Herbaceous

Feedstock

Conversion

Hydrolysis and fermentation

Lignocellulosic biomass

Herbaceous

Chemical process
Pre-process

Oil crops

Fuel for heat


and/or power

Anaerobic digestion

(Wood, straw,
energy crop, etc.)

Liquid fuels,
transport fuels
Biodiesel

Hydrogenation
Bioethanol
Transesterification
Other catalysis

Thermochemical process

(Rape, soy, palm, etc.)

Gasification
Pyrolysis

Other liquids

Woody

C 4C 4

Annual

Perennial
C3C3
C C4
Grain
or seed 4Herbaceous
Woody

Annua

Perenn

Grain or seed

Photosynthetic pathway
Photosynthetic pathway
Propagation
method
Propagation
method
Propagation
method
C3

C4

C3

C4

Plant
Plant
Plant
Plant
types
types
types
types
Plant
Plant
types
types

C3

C4

C3

C4

Annual
Annual
Annual
Annual

Annual
Annual
Seed
Stem
Rhizome
Seed
Stem
Rhizome Micropropagation
Micropropagation
Perennial
Perennial
Perennial
Perennial
cutting
or
cuttings
cutting
orroot
root
cuttings
Woody
Woody
Woody
Woody
Grain
Grain
Grain
orGrain
seed
oror
seed
seed
or seed
Herbaceous
Herbaceous
Herbaceous
Herbaceous
Propagation method
Propagation method
Perennial
Perennial
Woody
Woody
Grain
Grain
or or
seed
seed
Herbaceous
Herbaceous
Current
dominant
energy
use
Current
dominant
energy
use
Photosynthetic
Photosynthetic
Photosynthetic
Photosynthetic
pathway
pathway
pathway
pathway

Photosynthetic pathway

Photosynthetic
Photosynthetic
pathway
pathway

Gaseous fuel
Syngas

C 3 CC33 C 3

Schematic diagram of bioenergy production pathways. Feedstocks on the


left of the diagram are converted via a range of processes to solid, liquid or
gaseous fuels on the right. No attempt is made to show relative scales of
each process

C 4 CC44 C 4

DDC 4C 4Micropropagation
Seed
Stem EE C 3C 3 Rhizome
Rhizome
Micropropagat
CAM Stem
C3orCC3root
C4 CC4 4 C4
3 C3cuttings
cutting
cutting
or root cuttings
Bioethanol
Bioethanol Biodiesel
Biodiesel Heat
Heatand
andpower
power Biogas
Biogas
C3 C3
C4 C 4

Seed

Biogas

Combustion

 Table 3.1
Bioenergy production routes

C 3C 3

Energy

Biochemical process

Woody
Woody Grain
Grainororseed
seed

Plant types pathway


Photosynthetic
Photosynthetic
pathway
Plant types

Plant types

Sugar and starch crops

Annual
Annual

Current dominant energy use


Current dominant energy use
Other
Other
Propagation
Propagation
Propagation
Propagation
method
method
method
Plant life cyclemethod
E
Bioethanol

Propagation
Propagation
method
method
E
D
Car
Weight
Barrel
Car
Weight
Barrel
Biodiesel HeatAnnual
and power Perennial
Bioethanol
Biogas
Biodiesel Heat and power Biogas
D

Other

Car

Seed
Seed
Seed
Seed

Stem
Stem
Stem
Stem

Rhizome
Rhizome
Rhizome
Rhizome
Micropropagation
Micropropagation
Micropropagation
Micropropagation

Seed
Seed

Stem
Stem

Rhizome
Rhizome Micropropagation
Micropropagation

cutting
cutting
cutting
cutting
or root
ororroot
root
or
cuttings
root
cuttings
cuttings
cuttings
Other

cutting
cutting
oruse
or
root
root
cuttings
cuttings
Current
Current
Current
Current
dominant
dominant
dominant
dominant
energy
energy
energy
energy
use
use
use
Primary energy use
Current
Current
dominant
dominant
energy
energy
use
use
Weight

E EE E

Barrel

Car

Weight

D DD D

Bioethanol
Biodiesel
Biodiesel
Heat
Heat
Heat
and
Heat
and
and
power
and
power
power
power
Biogas
Biogas
Biogas
Bioethanol
Bioethanol
Biogas
E EBioethanol
DBiodiesel
DBiodiesel
Power
Powerusage
usage
Bioethanol
Bioethanol
Biodiesel Heat
Heat
and
and
power
power Biogas
Biogas
Other
Other
Other
OtherBiodiesel

Icons shown in grey indicate pre-commercial stages of adoption.

Other
Other
CarCar
CarCar Weight
Weight
Weight
Weight Barrel
Barrel
Barrel
Barrel
Power usage

CarCar

usage
Weight
Weight PowerBarrel
Barrel

Power
Power
Power
Power
usage
usage
usage
usage
Power
Power
usage
usage

Barrel

Biomass in the
energy industry
An introduction

First published 2014

Acknowledgements

We make no representation, express or implied, with regard to


the accuracy of the information contained in this handbook and
cannot accept any legal responsibility for any errors or omissions
that may have been made.

The insights and technical information presented in this


document were shaped by the research of many academic
scientists associated with the Energy Sustainability Challenge
(www.bp.com/energysustainabilitychallenge) and the Energy
Biosciences Institute (www.energybiosciencesinstitute.org).
In particular, Prof. Steven Long, Prof. Chris Somerville, Dr Heather
Youngs and Dr Caroline Taylor of the Energy Biosciences Institute
provided useful data, insights and perspectives throughout the
drafting.

Copyright 2014 BP p.l.c.


All rights reserved. No part of this handbook may be reproduced,
stored in a retrieval system, transmitted or utilized in any form or
by any means, electronic, mechanical, photocopying, recording or
otherwise, without written permission from Cameron Rennie, BP
International Ltd.
Printing: Pureprint Group Limited, UK, ISO 14001, FSC certified
and CarbonNeutral.
Paper: This handbook is printed on FSC-certified Cocoon Silk.
This paper has been independently certified according to the rules
of the Forest Stewardship Council (FSC) and the inks used are all
vegetable-oil based.
This handbook was written and edited based on literature
review by Dr Sarah Davis, Ohio University; John Pierce, BP
Chief Bioscientist; and John Simmons, ON Communication;
analysis and research by Dr William Hay, Researcher for Global
Change Solutions and Reza Haghpanah at SPENTA ; and project
management by Sharon Rynders, BP and Morag Ashfield,
ON Communication.
Designed, illustrated and produced by ON Communication,
www.oncommunication.com
For more information
BP contact: Sharon Rynders
www.bp.com/energysustainabilitychallenge
Published by BP p.l.c., London, United Kingdom
ISBN 978-0-9928387-1-3
Reference citation
Davis, S.C., Hay, W. & Pierce, J. (2014),
Biomass in the energy industry: an introduction.
The Biomass handbook is part of a series that reflects the work
of the BP-sponsored Energy Sustainability Challenge.
The other titles are:
Water in the energy industry An introduction
Materials critical to the energy industry An introduction
(2nd edition)
These books can be downloaded at:
www.bp.com/energysustainabilitychallenge

In addition, we would like to thank the following people for their


guidance in the writing and structuring of the handbook and for
their technical review: Dr Gran Berndes, Chalmers University;
Prof. Dr Marcos Buckeridge, University of Sao Paulo;
Dr Steven L Fales, Iowa State University; Dr Angela Karp,
Rothamsted Research; Prof. Dr Iris Lewandowski, Universitat
Hohenheim; Dr William Parton, Colorado State University;
Dr Jeremy Woods, Imperial College.
We thank Matthew Trainer, Data and GIS Specialist in the
Voinovich School at Ohio University, for the creation of detailed
biome maps.
We are also grateful for the analytical insights and valued
contributions from many colleagues within BP.
In acknowledging our gratitude to these individuals and
institutions, we do not imply that they either endorse or agree
with any statements or views expressed in this handbook.

Contents

Contents and About this book 3

Foreword by John Pierce
BP Chief Bioscientist 4

Foreword by Stephen P Long FRS
Gutgsell Endowed Professor of Plant Biology at the
University of Illinois, and Chief and Founding Editor,
Global Change Biology 5

About this book


The intent of this book is to provide an introduction to the
potentials and issues associated with utilizing biological
materials (biomass) for energy. Detailed information
is provided on various biological materials, including
currently important crops and those thought to have
future potential. Contextual information associated with
agriculture, energy and environmental considerations is
also provided.

Units of area and Units of energy 67


Setting the context 8

Global energy use


Biomass and bioenergy
Overview of agroecosystems
Water use in agriculture
Agricultural production of energy crops

Global ecosystems and land classifications


Land types
Plant functional features
Metrics of biomass productivity
Energy issues and greenhouse gas accounting

Important concepts 22

3 Bioenergy potential 34

Current bioenergy production


Global potential bioenergy production
How might this global potential be realized?
Developments in biomass conversion technologies

Economics, the environment and politics 48


The socio-economic drivers and impacts of bioenergy
Environmental sustainability
The politics of biomass


5 Where can biomass feedstocks be grown? 58
Growing regions (biomes)
Regional characteristics: comparison table

6


Biomass feedstock crops 70


Introduction to the selected crops
Biomass crops: comparison table
Biomass feedstock crops: complete list of references


Glossary 115

Foreword by John Pierce


BP Chief Bioscientist

Energy is at the foundation of all economic activity.


It heats us, it cools us and it lights our lives. It drives
our transport, communication and computer systems,
and provides the heat and mechanical work required to
transform materials into a dazzling array of useful forms.
Energy utilization is strongly correlated with economic
well-being, and we have become adept at deriving energy
from a wide range of sources. Energy is abundantly
available and we use it abundantly. The majority of
our energy derives from fossil fuels, and their use results
in increasing concentrations of carbon dioxide in the
atmosphere with worrying consequences for our climate.
While we are likely to find fossil resources to fulfil our
energy needs for many years to come, the pressing need
to understand the effects of our energy use on our finite
atmosphere, land and water resources has resulted in a
burgeoning effort to find alternative renewable forms of
energy with lower environmental impacts. However, any
activity as large in scale as energy production requires very
careful assessment and understanding of likely impacts.
Approaches that are renewable in one dimension may be
less so in another.
Sunlight is earths primary source of energy.
Indeed, our fossil sources of energy derive from plant
photosynthesis that took place long ago. The difficulty
with all contemporary solar energy conversion methods
is the dispersed nature of sunlight and the need to
concentrate it into energy vectors that are more readily
useable.Geological processes did this for us in making
fossil fuels in rich, concentrated deposits, but renewable
energies dependent on the sun require us to gather
the energy produced over large areas. As a result, the
use of such renewable energies requires significant
new approaches for collection and distribution, and a
sophisticated approach to land utilization.

The use of renewable biological materials as energy


sources is an area of increasing focus. Plants cover the
earth profusely and, using energy provided by the sun,
convert carbon dioxide and water into useful organic
compounds on a truly massive scale. To take advantage
of this fecundity to effectively and sustainably provide a
significant source of our energy needs without degrading
other aspects of our environment will require diligent
work to understand the scale of effort involved, the nature
of the plants themselves, and how to conduct large-scale
agriculture and forestry in the most environmentally
responsible manner.
The impact of biomass and land availability on energy
production is one of many questions being addressed
in BPs Energy Sustainability Challenge programme.
Researchers from a number of leading universities are
collaborating in this programme to establish trusted data
on the land, water, materials and ecosystems footprints
of different energy pathways. BP is pleased to support
this contribution concerning the role of biomass in
energy production. We would like to thank all involved,
and especially our colleagues at the Energy Biosciences
Institute and the reviewers, who have together helped to
ensure that this book is factual and well-founded.
In detailing attributes of crops and biological resources
currently in use for energy production, as well as emergent
energy crops and issues associated with large-scale energy
production from agriculture, we hope this book will
provide an accessible overview and contribute towards a
more sound understanding of the use of biomass in energy.

Foreword by Stephen P Long FRS


Gutgsell Endowed Professor of Plant Biology at the University of Illinois,
and Chief and Founding Editor, Global Change Biology

The carbon dioxide (CO2) concentration of our atmosphere


has been monitored at the Mauna Loa Observatory, high
over the central Pacific, since 1959, when it was 316
parts per million (ppm) of air. Since then CO2 has risen
at an ever-accelerating rate and on 13 May 2013 reached
400ppm a 27% increase in just over half a century, and
more than 50% higher than the global pre-industrial level.
Because of the differences in isotopic composition of
carbon in the biosphere and that in fossil fuels, it has been
shown that most of this increase is due to our use of fossil
fuels. We have sufficient known fossil-fuel reserves to
continue to increase CO2 concentration in the atmosphere
to two and three times current levels. The physical laws
of thermodynamics and radiative exchange tell us that
if we increase the concentrations of long-wave radiation
trapping gases, such as CO2, the earth will warm. While
the details of climate change are uncertain, that climate is
changing, and will continue to change substantially, is a
fact. If this is not addressed, then food supply, biodiversity
and our most vital ecosystem services are threatened.
At first sight bioenergy derived from plants would
seem an ideal solution. Plants use energy from the sun to
assimilate CO2 and trap chemical energy in the form of
plant biomass. When the biomass or fuels derived from the
biomass are combusted, the same CO2 is returned to the
atmosphere. Thus energy is obtained for heat and work,
with no net effect on the CO2 level in the atmosphere. In
practice, some energy input is required to grow the crops,
transport biomass and produce fuels. But with the possible
exception of early corn ethanol operations, these almost
always produce considerably more energy than they
consume. Brazil has shown, through learning by doing,
a rapid pace of improvement of its sugarcane ethanol
system. By 2010 this change resulted in the sale of
more ethanol for its automobiles than gasoline, and the
production of a large proportion of dry-season electricity
via combustion of the sugarcane bagasse.

Despite the seeming value of bioenergy, progress has been


dogged by opposition, some well-intentioned and some
grossly over-exaggerated, based on single issues that fail
to recognize the wider feedstock options. Headlines such
as Food versus fuel, Water versus biofuel, Dirtier
than coal and The next kudzu have contributed
to the emergence of policies inhibitory to progress,
particularly toward more sustainable biofuels from
perennial feedstocks. Indeed, we now have laws aimed at
lowering invasive risk that apply to a crop if it is grown for
bioenergy, but exempt if used for food, despite the fact that
biologically, invasive risk will be the same, whatever the
end use.
Bioenergy could be a major part of reduction of net
carbon dioxide emissions, especially given the huge
potential to improve agricultural productivity and
sustainability. Realizing this key goal requires policies
based on a holistic view of risks and benefits, as well as
recognition of the different bioenergy feedstock options.
Until now there has been no such holistic overview so,
for the first time, this book provides one, outlining in
one place a contemporary and forward-looking view of
the issues. In addition, each current and proposed major
feedstock is objectively analysed for key properties,
including yield, agronomy, pests and diseases, handling
logistics, environmental benefits and invasiveness.
Particularly important is its illustration of the
opportunities presented by a wide range of perennials that
could provide substantial environmental benefits, restore
ecosystem services to degraded land and use land unsuited
to major food crops.

BP Biomass Handbook
Figure X.XX (10 December 2013)
Draft produced by ON Communication

1010 1011 1012 1013 1014 1015 1016 1017 1018 1019 1020

1,000,000

510,000,000km2
150,000,000km2
5,500,000km2

Surface of earth

Global land area

Amazon rainforest

1,800,000km2 Global maize harvest


700,000km2
110,000km

Texas

Cuba

10,000km2 Lebanon
1,000km2 Hong Kong
100km2
44km

3.4km

Paris

Vatican City

Central Park, New York

1.4km2 Hyde Park, London


100Ha 1 square kilometre (1km2)
44Ha
10,000m
7,000m

Tiananmen Square, Beijing

1 hectare (ha)
Professional football playing area

4,047m2 1 acre
420m2 Basketball court
10m2 Standard parking space
4.2m2 Table tennis table
1m2 Average bath towel
Note: all values approximated to two significant figures apart from unit conversions.

 Figure X.XX
Area comparisons

6 | 1Introduction
6

billion
billion

109

million
billion

108

million

107

thousand
billion

1,000

106

billion

100

105

etc.

10

104

10,000,000

103

100,000

102

10,000

10

thousand

square metres
m2

Units of area

BP
BP Biomass
Biomass Handbook
Handbook
Figure
Figure X.XX
X.XX (10
(10 December
December 2013)
2013)
Draft
Draft produced
produced by
by ON
ON Communication
Communication

105

106

107

108

109

1010 1011 1012 1013 1014 1015 1016 1017 1018 1019 1020

10

102

kilo
k

104

105

mega
M

107

108

giga
G

1010 1011 1012 1013 1014 1015 1016 1017 1018 1019 1020

10
10

100
100

10,000
10,000

100,000
100,000

10,000,000
10,000,000

etc.
etc.

thousand
thousand
billion
billion

109
billion
billion

million
million

106

1,000,000
1,000,000

1,000
1,000

thousand
thousand

103

exa
E

104

billion
billion
billion
billion

103

peta
P

102

million
million
billion
billion

10

tera
T

11

square metres
joules
m2

Units of energy

2
510,000,000km
Surface of earth
World energy consumption
in 2010

150,000,000km
Global
area
Global annual
oilland
production
2

2
5,500,000km
Amazon
rainforest
Solar energy
received on
earth every
minute

5 x 1020J
1.9 x 1020J

6 x 1018J

1,800,000km Global maize harvest


3.2 x 1016J
Average power plant annual
output
2

700,000km2 Texas
Oil passing through the Strait of Hormuz each hour 4.6 x 1015J
110,000km2 Cuba
1 x 1012J Typical road tanker full of gasoline
10,000km2 Lebanon
5 x 1011J Energy content of one hectare of miscanthus
1,000km2 Hong Kong
3 x 1010J One tonne of bioethanol
100km2 Paris
5.7 x 109J Energy in one barrel of oil
44km2 Vatican City
3.6 x 109J One megawatt hour (MWh)
3.4km2 Central Park, New York
1.6 x 107J Energy content of 1kg of maize
1.4km2 Hyde Park, London
1 x 107J Recommended human daily calorific intake
100Ha 1 square kilometre (1km2)
3.6 x 106J One kilowatt hour (kWh)
44Ha Tiananmen Square, Beijing
10,000m2

6
2.2 x 10
J Dietary energy in 100g dark chocolate
1 hectare
(ha)

Running
a largearea
television for one hour
1 x 106J football
playing
7,000m2 Professional
x 105J Dietary energy in one large apple
acre
4,047m2 14.2
2 x 103J One kilocalorie or dietary Calorie
4.18
Basketball court
420m

One British
Standard
space thermal unit (btu) = 1,055J
x 103Jparking
10m21.055
2
Table
tennis table
4.2m
One calorie
4.18J
2
bathgram
towel(nearly one litre) of air through one degree Celsius
1m
Heating one
1J Average

Note: all values approximated to two significant figures apart from unit conversions.

 Figure X.XX
Energy comparisons
 Figure X.XX
Area comparisons

1Introduction | 7
7

World population is forecast to reach

8.3 billion by 2030, and societies


are becoming more affluent.
Global energy use is rising with
population growth and increased
consumer demand, and there
are concerns about the resulting
carbon dioxide emissions to the
atmosphere. Renewable energy
offers a mechanism to reduce carbon
emissions, and its production is
expected to grow faster than overall
energy growth through to 2035.
Biomass the solid matter in

biological organisms can be


converted into biofuels, heat and
power, and biogas. Global use of
bioenergy is expected to more than
double by 2035, with heat and power
being the largest consumers. Liquid
transport fuels currently account for
less than 5% of current bioenergy,
though production is rising fast.
Most biofuels are currently derived
from crops that are also food for
humans and animals, but non-food
plants are also being investigated for
their potential as biomass crops.

Societies throughout the world have

converted significant areas of forests,


savannah and shrubland into crop
and pasture lands, while advances in
technology and plant breeding have
led to prodigious yield increases in
commercial food and fibre crops.
Similar improvements are expected
for dedicated energy crops; these
could potentially be grown on land
currently less suitable for food
production.
While economic and practical

realities will limit the rate and


extent of change possible, there is
substantial land and technology
available for improving the overall
output of both food and biomass.

1 Setting the context


This chapter provides an introduction to the use of biomass
for energy in the context of global energy use, the evolution
of agriculture and projections for future uses of biomass.

Global energy use


Population and income growth are the key drivers of the
The worlds energy use is complex and changing. Total
growing demand for energy. These factors and the
energy use rises with population and economic activity,
development of energy-hungry technologies from the time
and technological and commercial innovations affect the
of the Industrial Revolution to the present day are reflected
type of energy used. As a result,
the amount
and type of
BP Biomass
Handbook
in the energy demand curves shown in Figure 1.2. By 2030
energy use varies throughout
the1.1
world
depending
on both
Figure
(10 February
2014)
the world population is projected to reach 8.3 billion,
technology and available Draft
resources.
produced by ON Communication
which means an extra 1.3 billion people will need energy;
Figure 1.1 shows how sources of energy differ across
and world income in 2030 is expected to be roughly double
regions and by level of economic development. Biomass
the 2011 level in real terms[1].
currently provides a very small portion of energy use
in developed countries, whereas biomass is a primary
energy source for heating and cooking in many developing
countries. The reliance on biomass in Africa relative to all
other regions is clearly shown, as is the relatively small
contribution from renewable sources worldwide.
Figure 1.1
World energy use in percentage
terms by region and by source
in 2010. Each square in the regional
stripe represents 1%. The major
fossil and renewable fuels are
shown in different colours. Circles
show tonnes of CO2 emissions per
capita in 2009[2, 3].

tCO2/person

Per cent energy use by source


World

4.7

OECD
Americas

17.1

OECD
Europe

7.2

OECD
Asia
Oceania

9.0

Eastern
Europe/
Eurasia

5.1

Non-OECD
Asia

4.6

Africa

0.9

Latin
America

2.8

Middle
East

 Figure 1.1

5.9

Biomass

Hydro

Other renewables

Nuclear

Gas

Oil

Coal

1 Setting the context | 9

Figure
1.2Handbook
BP
Biomass
Global
use
(EJ)2013)
of the six most important energy sources since 1850. Historically, biomass use is mainly
Figure 1.2 in
(20exajoules
Decmber
the traditional
useby
of ON
fuelwood;
the renewable curve includes all modern renewable sources except biomass.
Draft
produced
Communication
Major technology advances are shown and also significant changes in energy source: coal replacing biomass in the
Industrial Revolution; the increase in oil with the rise of the internal combustion engine; and gas for heating and
power generation[4].

500

Primary energy (EJ)

400
Microchip

Commercial
aviation
300
Nuclear
energy
Television

200
Vacuum
tube
100

Steam
engine

Electric
motor

Gasoline
engine

0
1850

1950

1900
Biomass

Coal

Oil

Gas

2000

Nuclear

Other renewables

BP Biomass Handbook

EJ

Billion toe

1.3
Figure 1.3
(11Figure
June 2014)
Despite increasing energy efficiency, energy consumption
Draft produced by ON Communication
is on the rise globally as shown in Figure 1.3. World
The increase in energy demand in billion tonnes of oil equivalent
(toe), excluding biomass used for heat and cooking from 1990 to
primary energy consumption is projected to grow by
the present day, and projections until 2035. The effect of the global
1.5% per year from 2012 to 2035, adding 41% to global
economic crisis from 2008 can be seen clearly[5].
consumption by 2035. The fastest-growing fuels are
renewables (including biofuels), with growth averaging

Figure
6.4%
per 1.2
year from 2012 to 2035. Nuclear (2.6% per year)
18
and hydro (2.0% per year) are both projected to grow
700
faster than total energy consumption.
Among fossil fuels, natural gas use has grown the
15
600
fastest (1.9% per year), followed by coal (1.1% per year)
[5]
and oil (0.8% per year) . The lower relative growth rates
of fossil fuels, however, apply to a very large base of use.
12
500
On an absolute energy basis, for example, coal use grew
the most in the period 2000 10 and the additional use
400
of coal constituted almost 50% of the total increase in
9
energy use.
300
6
200
3

100

0
1990

10 | 1 Setting the context

2005

2020

Oil

Gas

Coal

Nuclear

Hydro

Other renewables
(includes biofuels)

Note: 1toe equals approximately 42 gigajoules.

Figure 1.3

2035

Biomass and bioenergy


Biomass refers to the matter in all biological organisms,
but in the context of energy it is most commonly used to
mean the solid material that can be harvested or collected
from biological organisms primarily from plants. This
meaning is used throughout this handbook (definitions for
terms used in this book that are unfamiliar or that have
various meanings in common usage may be found in the
Glossary at the end of this handbook). Major components
of biomass include sugars, starches and oils produced from
plants. These are extracted in great abundance today for
energy production. The term biomass also includes the
heterogeneous material found in even greater abundance in
materials such as wood, plant stems and husks. All these
materials can be converted into an energy form useful for
heat, power and transport fuel.
Bioenergy is a general term referring to energy derived
from any renewable biological material from plant matter,
animals or organic wastes derived from plant and animal
matter. In this handbook we focus primarily on the
conversion of materials from plants (biomass) into
bioenergy.
Bioenergy is produced from biomass in a number of ways,
including:

Biofuels: liquid fuels mainly used for transport, produced


by a variety of thermochemical and biochemical
processes. These fuels can come from a wide range of
plant and animal materials. The predominant forms
today are bioethanol (derived from fermentation of
sugars) and biodiesel (from esterification of plant and
animal oils) with increasing amounts of the oils being
treated with hydrogen to create hydro-treated vegetable
oils (HVO) suitable for diesel use. A variety of new fuel

molecules are also at the research and early commercial


demonstration phases. For instance, liquid biofuels
derived from lignocellulosic biomass that has undergone
thermochemical or biochemical processes are just
beginning to appear in commercial quantities.

Heat and power: this includes the traditional form of


bioenergy in which plant materials (such as wood or
grasses) are collected and burned for heat. This heat can
be used to generate electrical power as well. Early, largescale adoption has often involved a mixture with coal for
the generation of power as shown in Figure 1.4.
Biogas: a combustible gas produced by the anaerobic
digestion of biological material. Biogas consists of a
number of different compounds and hydrocarbons,
the main ones being methane and carbon dioxide. It is
produced from a wide range of materials, particularly
wastes (especially those with relatively high water
content), and from landfill. It is used primarily for
electricity and heat generation. Biogas from anaerobic
digestion should not be confused with syngas, which
can be derived from both fossil and renewable sources of
carbon (biomass). Syngas has a very different chemical
composition (being composed of carbon monoxide and
hydrogen) and is both made and used in a very different
manner. Biogas itself can be converted to syngas for use
in the production of fuels, but this is not a significant
current practice.

Production routes of biomass to various fuels are


discussed in greater detail in Chapter 3 with a summary
provided in Table 3.5 and a schematic outline on the inside
front cover of this book.

Figure 1.4
Woody biomass being blended with coal at
a Colorado electricity generating plant to
provide a mixed feedstock boiler fuel[6].

1 Setting the context | 11

Biomass can also be used on a large scale in the production


of industrial chemicals. Current processes primarily focus
on converting starch and sugar into the desired chemical
products, using microorganisms modified to produce
the chemical of interest. In many ways these processes
resemble those used to produce biofuels such as ethanol.
This is an area of much research and commercial interest,
and new processes and facilities continue to appear. Given
the volumes of materials needed for the chemical sector,
the overall use and demand for biomass for chemicals is
much lower than that for energy.
Globally and traditionally, the largest use of bioenergy is
for so-called direct use. This traditional use of bioenergy
is mainly for heating and cooking, using biomass sources
such as wood, charcoal, crop residues and animal dung.
Much of it is used in small domestic stoves and open
fires, and statistical data are therefore limited. Even
in OECD countries, two-thirds of total bioenergy use
is for heating, much of it sourced through forestry
management. Figure 1.5 is a chart originally published by
the International Energy Agency (IEA) depicting the use
of bioenergy by sector in 2010 along with the potential
use in 2035. The future estimates are based on the IEAs
New Policies Scenario, which takes into account broad
policy commitments and plans to address energy-related
challenges, even if the specific measures to implement
these commitments are yet to be defined.
In this assessment, the total amount of traditional
biomass consumed is expected to decline slightly over
time, as access to modern fuels increases around the
world. Excluding the traditional use of biomass, global
primary use of bioenergy is expected to more than double
from 22 exajoules (EJ) in 2010 to nearly 50EJ by 2035,
growing at an average rate of 3.3% per year. Provision of
heat and power are projected to be the largest consumers of
non-traditional bioenergy, potentially growing from nearly
17EJ in 2010 to more than 37EJ by 2035. Together, these
two sectors account for about two-thirds of the additional
consumption of bioenergy in the IEA scenario.

A little more than 10% of current non-traditional


bioenergy is in the form of liquid fuels for transport
(i.e. biofuels). Brazil and the US are the largest producers
of bioethanol, and Germany is the largest producer of
biodiesel. The use of biomass for electricity generation
(such as bagasse in Brazil and woodchip- and pellet-fuelled
power generators in the UK) accounts for just over 20% of
current non-traditional bioenergy.
Bioelectricity continues to grow in both OECD and
non-OECD nations. In 2011 more than 35 countries
had bioelectricity capacities exceeding 100 megawatts
(MW). Total generation has increased by more than
170 terawatt-hours (TWh) (0.6EJ) from 2000, reflecting an
8% annual growth rate over the past decade[7]. With more
than 100 countries enacting renewable electricity targets,
bioelectricity is expected to grow. The IEA estimates that
electricity generated from biomass could grow to 530TWh
(1.9EJ) in 2017 and possibly to more than 1,470TWh (5.3EJ)
in 2035, depending on the cost and availability of biomass.
While much work has been done to map the potential
for global biogas production, there is little reliable
data about current biogas production levels in many
countries. While the contribution (in energy terms) is
relatively small, biogas was used to produce roughly 3%
of electricity use in Germany[8], provided heating and
cooking fuel to nearly 40 million Chinese households[9],
and made up 64% of the gas use for transportation in
Sweden in 2010[10]. There is increasing production and
local use of biogas from landfills, and growing interest in
utilizing anaerobic digestion of biomass for biogas and
production of electricity.
Heat and power production are, and are expected to
continue to be, the largest uses of biomass, enabled by
well-known and widely practised combustion technology.
However, biofuels for transport are also expected to more
than double by 2035, and significant research is under way
to provide more cost-effective conversion technologies
to enable more penetration into the transport sector with
fewer environmental impacts than are currently associated
with liquid fuels.

Figure
1.5
BP Biomass
Handbook
Use of
bioenergy
sector in 2014)
2010 and 2035 (projected by the IEA for conditions where new policies are implemented).
Figure
3 (10by
February
Use is
estimated
to riseby
from
in 2010 to 79EJ in 2035. The proportion used for heat by traditional methods (heating
Draft
produced
ON53EJ
Communication
and cooking) is projected to fall considerably; the proportion used for heat via modern methods of production remains
almost unchanged; while proportions used for power and transport by modern methods make significant increases[2].

2010
Other
5.7%

2035
Heat
22.4%

Other
5.5%

Power
8.5%
Traditional
58.8%

Transport
4.6%

Traditional
36.5%

Heat
24.6%

Power
22.5%
Transport
10.9%

Total 53EJ

 Figure 2.6a
Use of bioenergy by sector in 2010 and 20351

12 | 1 Setting the context

Total (projected) 79EJ

Although biofuels currently supply only a small fraction


of liquid transport fuels, production has been rapidly
rising. Crops used to produce both food and fuel dominate
today, though crops grown specifically for energy use
are expected to increase in the future. While future
projections of biofuels production are notoriously difficult
to quantify, and the types of crops used and fuels made are
subject to numerous technical, economic and political
considerations, recent analyses suggest that biofuels could
constitute approximately 6% of liquid transport fuels
by 2035[2], with around three-quarters of the production
BP
BPBiomass
BiomassHandbook
Handbook
continuing
to come
from North and South America[1].
Figure
Figure1.6
1.6(20
(20December
December2013)
2013)
Figures
1.6
and
1.7
provide
example perspectives on the
BP
Biomass
Handbook
Draft
Draftproduced
producedby
byON
ONCommunication
Communication
Figureand
1.6 future
(20 December
current
mix of 2013)
feedstocks for the supply of

biofuels to 2020. This projection, from the UN FAO,


shows negligible growth of ethanol based on grains
such as corn, the most significant source today, with
growing volumes from sugarcane and, later in the
projection, cellulosic biomass crops. Biodiesel is
produced, and is expected to be produced, at much
lower volumes (note different scales used in the graphs).
Vegetable oil is projected to remain the most widely
used source of biodiesel by volume. Waste oils, fats and
tallows, as well as new crops with high oil yields and an
ability to grow in diverse habitats (as exemplified by
Jatropha in Figure 1.7), are anticipated to become a
richer part of the mix.

Draft produced by ON Communication

Figure 1.6
Current and future mix of volumes of bioethanol supplied for fuel use projected annually to 2020[11].
160
160
Coarse
Coarsegrains*
grains*

160
140
140

Other**
Other**
Coarse
grains*
Sugarcane
Sugarcane
Other**

140
120
120

Sugar
Sugarbeet
beet
Sugarcane
Cellulosic
Cellulosic
Sugar beet

Billion
Billion
litreslitres
Billion litres

120
100
100
100
8080

Wheat
Wheat
Cellulosic

80
60
60

Wheat

60
40
40
40
20
20
20
00

20082008- 2011
2011 2012
2012 2013
2013 2014
2014 2015
2015 2016
2016 2017
2017 2018
2018 2019
2019 2020
2020
2010
2010
2008- 2011 2012 2013 2014 2015 2016 2017 2018 2019 2020
2010 grains
* *Cereal
Cereal
grainsother
otherthan
thanwheat
wheatand
andrice.
rice.
****Production
Productionfrom
fromother
othersources
sourcesisiscomposed
composedofofresidues
residuesofofallallkinds,
kinds,ininparticular
particular
* wood,
Cereal
grains
other
than
wheat
and
rice.and
wood,as
aswell
well
asasfrom
from
roots
roots
and
and
tubers
tubers
andsugar
sugarproduction
productionby-products.
by-products.
** Production from other sources is composed of residues of all kinds, in particular
wood, as well as from roots and tubers and sugar production by-products.

Figure 1.7
Current and future mix of volumes of biodiesel supply projected annually to 2020[11].
4040
40
35
35

Vegetable
Vegetableoil
oil

Billion
Billion
litreslitres
Billion litres

35
30
30

Biomass-based
Biomass-based
Vegetable
oil
Waste
Wasteand
andby-products
by-products
Biomass-based

30
25
25

Jatropha
Jatropha
Waste
and by-products

25
20
20

Jatropha

20
15
15
15
10
10
10
55
050
0

2014 2015
2015 2016
2016 2017
2017 2018
2018 2019
2019 2020
2020
20082008- 2011
2011 2012
2012 2013
2013 2014
2010
2010
2008- 2011 2012 2013 2014 2015 2016 2017 2018 2019 2020
2010

1 Setting the context | 13

The bioenergy production chain consists of four main


stages as shown in Figure 1.8. Biomass is grown, collected
and often treated or densified into transportable forms
such as bales, chips, billets or pellets to allow economical
movement to the conversion facility. Following conversion
to an alternate energy carrier such as electricity, heat,
steam, liquid fuel or gas, the bioenergy can be distributed
for end use in homes, vehicles and industry.
BP Biomass
Handbook
Biomass
cultivation can encompass a wide variety of
Figure 1.8
(18 November
2013)
practices
used in conventional
agriculture and forestry.
Draft produced by ON Communication
Where residues or wastes are the biomass source for
energy, the production chain may begin with collection

and treatment, since primary cultivation is for other


purposes including food, feed or timber production.
There are many ways to convert raw biomaterials into
fuels. Active development worldwide is improving the
various process efficiencies and widening their utility for
different feedstocks (see Chapter 3 for more details on
conversion processes).
As the use of bioenergy grows, there is increasing
interest in developing new types of crops that produce
a large amount of biomass per hectare and are grown
specifically to supply energy. These plants are known as
energy crops, examples of which are detailed in Chapter 6.

Figure 1.8
The four main stages of bioenergy production.
Biomass
cultivation

Collection,
densification,
transport and storage

Conversion to
energy carrier

Distribution
and end use

What makes an energy crop?


The first criterion for any energy crop is that it should
 Figure 1.8be productive in terms of biomass yield per hectare to
minimize the land area required. Second, the physical
and chemical characteristics of the crop must be
suitable for the conversion technology that will convert
it into biofuel, biogas or power. For energy crops to
make significant impacts on energy use, they must be
grown on a large scale, so questions of sustainability
(economic, environmental and social) must also be
considered. The following characteristics are key traits
that facilitate environmental sustainability:
Characteristics of an ideal bioenergy crop [1214]

High-energy yield per unit growing area.


Low-input, low-cost processing requirements.
Low greenhouse gas (GHG) emissions and energy
requirements.
Easy to establish.
Tolerant to extreme and/or variable environments.
High efficiency of nutrient use.
High efficiency of water use.
Provide additional ecosystem services
and/or co-products.
Suitable for a range of conversion processes into
various forms of bioenergy.
Productive on soils and topographies less suited to
food crops.
Low- or zero-invasive potential.
Unrelated to native or major weed species to avoid
spread of genes and potential disruption of native
ecosystems.

Inevitably, no single crop will meet all of these


requirements, and the importance of a particular
requirement depends on location. Chapter 6 of this
handbook lays out the extent to which alternative
energy crops address these requirements in different
locations.
As a general rule, perennial crops meet many of
these requirements: they do not require annual tillage
and planting, and they often recycle nutrients and add
carbon to the soil. As explained in Chapter 3, most
14 | 1 Setting the context

advanced biomass crops (both current and potential) are


perennials, either trees or perennial grasses. Complex
choices between maximizing yield and maximizing
sustainability, however, have to be made. For example
perennial grasses such as switchgrass and miscanthus
can be highly productive in temperate environments.
In the autumn these grasses transfer their nutrients to
the root system so they can be retained over winter,
even if the plant is harvested. Production in temperate
environments is limited to the warm period of the
year, and is therefore less productive than in moist
subtropical and tropical environments where production
is possible throughout the year. On the other hand, in
environments with no dormant season, material must
be harvested green. This gives higher yields but there
are costs associated with drying fresh grass, and in this
scenario there is less recycling of nutrients to the root
system.
Geographical, topographical and social factors can
also affect the choice of bioenergy cropping systems.
As a crop, oil palm, for example, is not inherently less
sustainable than other plantation crops. But many
areas suitable for oil palm are naturally forested, so the
conversion of carbon-rich peat-swamp forests in parts
of South-East Asia can have environmental impacts
that outweigh any benefit from producing renewable
biofuel. Conversely, the introduction of perennial grass
feedstocks on intensively managed land can create net
carbon and GHG benefits through restoration of soil
carbon and interception and reuse of nutrients.
Indeed, the use of perennials as bioenergy sources can
broaden opportunities for sustainable agricultural
production: some grasses can stabilize eroding slopes,
others tolerate saline soils, while succulent plants
thrive in semi-arid areas.
Trade-offs between different crops are inherent in all
forms of agriculture. As large new areas of bioenergy
crops are contemplated, it is essential to minimize
environmental impacts. Land resources are not infinite,
so production of bioenergy crops must be carefully
balanced with other uses for land, such as food, animal
feed and material production.

Overview of agroecosystems
Figure 1.9

After the emergence of agriculture about


10,00012,000 years ago, agricultural productivity
increased slowly until the advent of mechanization in
the 18th century, when horse-drawn drills, reapers and
threshing machines allowed more land to be cultivated
and also brought greater yields. The invention of the
steam engine and steam plough increased productivity
further and allowed previously unproductive land to be
tilled. These gains were enhanced when tractors driven by
internal combustion engines came to the market in the
early 1900s. A modern tractor and plough can till many
times more land in a day than a person with a horse, and
combine harvesters reap in an hour what teams of people
could only achieve in days.
Other agricultural practices have also improved
dramatically. Crop rotation began in the 18th century. In
the 19th century, more sophisticated fertilizer treatments
emerged, along with increasing numbers of research
organizations that advised farmers on best practice and
introduced new strains of plants. The 20th century saw
genetics and chemistry playing an increasing part in
agriculture. The introduction of hybrid maize and dwarf
wheat allowed dramatically improved yields; the Haber
Bosch process for fixing nitrogen allowed ammonia-based
fertilizers to become cheaply available; and organic
chemists developed a wide range of effective pesticides
and herbicides. In the late 20th century and through
to today, genetic marker-assisted plant breeding and
genetic engineering (also known as genetic modification
or genetic manipulation) have become hallmarks of the
next revolution in agriculture in many parts of the world,
spearheaded by large research organizations.
The application of all these technologies has driven
prodigious improvements in the yields of many crops.
Since 1961, for example, average yields of sugarcane and
corn have increased by 41% and 166% respectively, as
shown in Figure 1.9. The most favoured growing locations
have seen even more dramatic yield improvements. These
improvements have occurred steadily over decades while,
at the same time, different crops have been developed to

Increasing yields of sugarcane and maize grain over 50 years


compared with yields in 1961. Moisture contents of 70% for
sugarcane and 15% for maize grain were assumed[15].
180%

1961 yields
Maize grain: 1.65 dry tonne/ha
Sugarcane: 15.08 dry tonne/ha

160%
140%

2011 yields
Maize grain: 4.38 dry tonne/ha
Sugarcane: 21.33 dry tonne/ha

120%
100%
80%
60%
40%
20%
0%

2011

2007

2003

1999

1995

1991

1987

1983

1979

1975

1971

1967

1963

-20%

Year
World average maize grain yield increase
World average sugarcane yield increase

respond to the changing demands for food, feed, fuel and


materials. Future energy crops, such as those discussed in
this handbook, would be expected to benefit from the same
types of investment.
In addition to the changing nature of agriculture, the
amount of land devoted to it has expanded significantly.
Across the globe, cropland and pasture has expanded at the
expense of primary forest, savannah and shrublands, as
shown in Figure 1.10. Growth has been driven primarily by
a rise in population and by the changing dietary intake of
more affluent consumers.

BP Biomass
BP Biomass
Handbook
Handbook
Figure Figure
1.9 (18 1.9
November
(18 November
2013) 2013)
Figure 1.10
Draft produced
Draft produced
by ON by
Communication
ON Communication
Change in land use during various periods since 1765[16]. The figures on top of each category show the land-use change
between 1765 and 2005, in million km3.
50

23.7
40

30

30

+8.3

+0.4

5.4

6.8

1.7

+0.5

+9.4

+19.3

1765

1765 1900

1900 2000

2000 2005

Pasture

Pasture

Cropland

Cropland

Urban

Urban

Others

Others

Shrubland

Savanna

0
Shrubland

Savanna

10

Grassland

10

Secondary
forest
Grassland

20

Primary
forest
Secondary
forest

20

Million km2

40

Primary
forest

Million km2

50

2005

1 Setting the context | 15


 Figure
 Figure
1.9
1.9

BP Biomass Handbook

Million tonnes

1.10Figure
(10 December
Increasing demand for meat proteins, as shown in Figure Figure
1.11 2013)
Draft produced by ON Communication
1.11, driven by population growth, economic growth
The growth in demand for meat proteins for developed and
and changing dietary habits, is directing more and
developing countries[17, 18].
more resources into meat production. From an energy
perspective, livestock production is quite inefficient.
Demand for meat proteins (million metric tons)
Intensive beef production, for example, commonly
250
utilizes grains for feed, and can require 620kg grain/kg
[17]
213
beef produced . While there is continuous development
200
in methods to improve the efficiency of meat production,
it is estimated that 70% of all agricultural land is used
in pastoral, mixed-system and intensive livestock
150
production. Food, feed and energy uses will all compete
111
for available land.
114
100
Despite the overall increase in land area devoted to
98
77
agriculture, there are areas where farming has been
50
abandoned across large regions. Some of this abandoned
32
agricultural land has become reforested and is now valued
for recreation, biodiversity and important carbon stocks
0
(growing forests remove substantial amounts of carbon
1997
1974
2020 (projected)
dioxide from the atmosphere). Many re-established
Developing countries
Developed countries
forests, such as large areas of the eastern US, are actively
managed for wood resources. Residual wastes from timber
extraction and saw milling have increasingly been used for
out of agricultural production due to changing land
energy in the wood-products industry and can potentially
ownership or altered economic incentives. Developing
provide bioenergy feedstock to other sectors (see Forest
recently abandoned land for bioenergy production would
biomass box below).
have less environmental impact than developing land on
Abandoned agricultural land that has not returned
which indigenous ecosystems have regenerated. It would
to forest or native ecosystems has, in many places,
 Figure 1.10
also help maintain food production capacity that might be
been developed for urban and residential use. Recently
required in the future.
abandoned land, however, may also be shifting in and

Forest biomass
Woody biomass is used for 80% of traditional primary
energy use, totalling nearly 32EJ per year and supplying
nearly 2 billion people with heat and cooking fuel. In
developed nations, wood typically supplies less than 5%
of primary energy. In the US, for example, wood is used
to supply 1% of the electricity supply and 2% of primary
energy, mainly to industrial users. Finland and Sweden
are exceptions with nearly 19% of primary energy
generated as heat and power from woody biomass[19,20].
The total potential for woody biomass could be
as high as 110EJ per year (EJ/yr), according to the
Intergovernmental Panel on Climate Change; however,
the sustainable and acceptable limits of forest biomass
use are still under debate and the use of forest biomass
for energy is controversial. Historic depletion of forest
resources in many parts of the world has instilled
caution in communities considering re-expanding
use of wood biomass for energy. In parts of the US,
Canada, the EU and China, forest biomass is actually
accumulating. Growing stock in the EU has increased
nearly half a per cent per year for the past 23 years and
US forest biomass has increased by 10% in a 10-year
period. In the US and Canada, less than 1% of available
forest biomass is currently harvested for all uses. The
increase in tree stand density, increased dead woody
biomass, and increasing climate stress have been
implicated in more frequent and more severe forest

16 | 1 Setting the context

fires. Whether increased forest management will result


in a sustainable and acceptable supply of biomass for
bioenergy is not yet clear.
The situation for tropical forests is still worrisome.
Forests in South America and Africa are still
experiencing net losses, although deforestation has
slowed in many regions including the Brazilian Amazon.
Although an increase in tropical forest plantations for
fruit, oil seed and timber production may eventually be
a source of residual biomass for energy in some regions,
such biomass may not be considered acceptable by
some stakeholders. See Chapter 4 in this handbook for a
discussion of sustainability criteria in policy.
Forest ecosystems can be very productive and offer
some advantages to herbaceous energy crops. Trees store
carbon, both above ground and in the soil, over a very
long time period and so can be left as standing stocks.
Herbaceous crops must be harvested before or soon after
senescence or they will degrade and release their carbon.
Trees can be grown on steeply sloping land and tolerate
a wide range of soils and hydrologies. Finally, forests
can provide a more diverse set of ecosystem services,
but because forests are more complex and require longer
rotations to accumulate biomass, a careful and detailed
understanding of each forest system is required to assess
long-term sustainability.

Water use in agriculture


All biomass requires water to grow. The intensity of water
use is determined by the volume of water withdrawn from
local freshwater sources and subsequently consumed in
their growth. Worldwide, about 80% of cropland is rainfed
(not irrigated[21]) and provides about 60% of global crop
production [22]. The remaining 20% of cropland, about
250 Mha[23, 24], is irrigated during at least part of the
growing season and yields about 40% of all production.
Freshwater withdrawals for agricultural irrigation
constitute some 2,700km3 of water (or about 70% of
world withdrawals) as shown in Figure 1.12, a Sankey
diagram that illustrates the fate of water withdrawn for
human use[25]. While many factors impact crop yield,
the dominant factor in determining where irrigation is
needed is the amount of rainfall. This varies dramatically:
from desert regions where precipitation is rare to regions
with more than a thousand millimetres of precipitation
per year. As with food crops, the amount of water used
for irrigation of biomass crops will be highly dependent
on local conditions and the type of plant used.Common
practice to date for the production of liquid biofuels has
been to grow crops where little or no irrigation is needed.
As a result, the intensity of water use for growing these
crops is much lower than for agriculture on average, with
estimates of about 0.5% of world freshwater withdrawals
for 2010 biofuels production[26].
The water cycle (in which plants participate) involves
the set of processes by which water circulates between
the earths oceans, atmosphere and land. It involves
precipitation as rain and snow, drainage in streams and
rivers, and return to the atmosphere by evaporation

and transpiration. This water cycle provides essential


ecosystem services. Regardless of whether vegetation is
native or non-native, evaporation and transpiration affect
the water flows into local streams and rivers. Vegetation
thus plays a part in flood control.
Plants obtain the water they need from the soil via their
roots. Soil water comes from precipitation, groundwater
and from irrigation. Water is lost from the soil by
evaporation, drainage and uptake by plants, with different
types of plant cover withdrawing water at different rates.
Plants take up far more water than they ultimately use in
photosynthesis or store within their structure.
The remainder is released into the atmosphere (in a
process called transpiration) to be recycled as rain. Plant
canopies intercept some rainfall before it reaches the
soil; this water is then lost through direct evaporation
from the leaves. This loss can be particularly important
in densely planted agricultural crops with complete
canopy closure. Because the periods of highest rainfall
in the year may be out of phase with crop demand for
water, storage of water in the soil plays an important
role in supporting crops. The capacity of the soil to
store water that is accessible to plants depends on the
soil texture. Sands absorb water rapidly, but can store
little and drain quickly. Clay soils will not absorb
water rapidly, making run-off and erosion more likely.
Once they do absorb water, clay particles bind water
molecules, so some of the water is not available to the
crop. Soil organic matter is critical in increasing the
water-holding capacity of soils.

Figure 1.12
Global Sankey diagram for annual fresh water withdrawn for human use [25]. From left to right, the diagram illustrates the continental
distribution of withdrawals, the sectors (agriculture, industry, domestic) in which the water is used, the services provided by the water, and
finally the return of the water to the hydrological cycle. Share of agriculture in total withdrawals is shown in yellow. In the final (right-hand)
segments, changes in water quality during its use are indicated in different colours. The red segment indicates where energy is used in
treating wastewater. The vertical width of each bar in the diagram is proportional to the volume of fresh water involved, measured in cubic
kilometres (km3), and numerical amounts are provided with labels, also in km3.

Regions

Total
applied water

Services

Post-use
treatment

Hydropower

Destination

Outflow to salt sinks

Africa
Atmosphere

Asia

North America
Industry
~ 775

Oceania
Europe
Recycled water

Domestic
~ 380

Energy
~ 470

Domestic
and industry
~ 790

Organic and inorganic compounds;


nitrogen and phosphorous.
High temperature
(effect on ecosystems).

Agriculture
Food
~ 2700
~ 2700
Biofuels

Latin America

Pollutant parameters
of return flow

Return flows into


surface water
and percolation
into groundwater
Untreated
wastewater
Sewage
treatment

Industrial effluent (silt and rock particles


and sufactants, heavy metals, high
biological oxygen demand, blood
contaminants).
Domestic effluent (excreta, urine
and faecal sludge) together with
grey water (kitchen and bathing
wastewater).

Recycled water
(direct to supply)

1 Setting the context | 17

Evaporation and transpiration provide another essential


ecosystem service by cooling continental surfaces that
could otherwise be much warmer at the height of summer.
At the same time these processes provide water to the
air, which in turn falls as rain elsewhere. Indeed a major
concern of Amazon deforestation is that it could cause

increased regional droughts because less water will be


evaporated. Other types of ecosystem services provided by
plants include reduced loss of nitrate and other elements
due to changes in peak flow drainage, and reduced soil
erosion.

Agricultural production of energy crops


A primary driver for promoting energy crops is the desire to
lower the amount of GHG associated with energy
production and to find cheaper and more sustainable ways
to produce biomass. It can be hard, however, to quantify
the issues: economic data are difficult to pin down, with
yields and prices for crops varying dramatically over time
and regionally. There is also an imbalance in the amount of
data available: we have plenty of information about
traditional crops used in our food systems, whereas largescale production of energy crops is still in its infancy. As a
result, much of the information on energy crops is based on
extrapolation from small datasets and emerging research
findings. Chapter 3 of this handbook provides more details
on the energy potential from biomass and issues associated
with its large-scale production.
Although food production is the main purpose of
agriculture, our farming systems also produce a wide range
of non-food goods and services. As the worlds population
grows and demands more resources from a finite amount of
land, there is a need to prioritize these items. Goods and
services arising from agriculture include food crops and

livestock, energy, organic materials (such as wood and


cotton), specialist materials (such as bioplastics and other
large-volume chemicals), carbon sinks, biodiversity and
other ecosystem services. Ultimately, land will be used in
a way that gives the highest economic return, and this will
differ by location and by prevailing policy and regulation.
Our need for food remains paramount. To feed our
growing population, we will need higher yields of existing
crops, as well as new crops that can be grown on land
currently less suitable for agriculture. Figure 1.13 provides
a view of the distribution of land areas around the world
with differing suitabilities for agriculture. Productivity of
crops grown in different regions can vary manyfold with
crop genetics and agricultural practice significantly
impacting output. Major land areas, for example in Africa
and the Ukraine, are currently producing substantially
suboptimal yields.
One way to achieve significant increases in biomass
production without compromising food resources is by
converting marginal and abandoned land to bioenergy
production. This could be done by replacing poorly

Figure 1.13
Suitability of land with appropriate levels of inputs for pasture and rainfed crops[27].

Agricultural suitability across the globe

n Closed forest

n
n Irrigated area
n Land prime or well suited for agriculture
Inland water bodies

18 | 1 Setting the context

n Land suited for agriculture


n Land unsuited to poorly suited for agriculture
n Protected area
n Urban area

performing crops with better alternatives as they become


available and developing crops that can be grown on saline,
water-logged or arid lands that cannot (economically
speaking) support food production. Investments in
bioenergy systems might even rehabilitate such land so
that it could be used for future food production.
Many agricultural systems already yield more than one
type of product simultaneously. For instance, when cereal
crops are used to produce bioethanol, between a quarter
and a third of the total weight of the grain is available
after processing as a high-protein feed (known as distillers
grains). Cotton produces cotton oil and cottonseed meal
as well as fibre (cellulosic cotton lint). Excess straw can

be used for heat and power production or as feedstock


for lignocellulosic biofuels. Similarly, the use of wood
harvested from rubber and palm-oil plantations at the
end of a plantations economic life does not displace
the production of rubber or palm oil, but facilitates the
replanting of such crops.
All of these approaches and more (increasing
productivity and yield, diversifying crop patterns,
bringing marginal land into cultivation and developing
multi-use crops) will be needed to achieve significant
expansion of agricultural output while minimizing
environmental damage.

Chapter references
[1] BP (2013), BP energy outlook 2030. BP, London, UK.
Available from: http://www.bp.com/content/dam/bp/
pdf/statistical-review/BP_World_Energy_Outlook_
booklet_2013.pdf [accessed July 2013].
[2] IEA (2012), World Energy Outlook 2012. International
Energy Agency (IEA), Paris.
[3] The World Bank, Data Indicators Databank CO2
Emissions. Available from: http://data.worldbank.org/
indicator/EN.ATM.CO2E.PCcountries/1W?display
=graph [accessed February 2014].
[4] Adapted from Nakicenovic, N. (2009), Supportive
policies for developing countries: a paradigm shift.
Background paper prepared for World Economic and
Social Survey 2009.
[5] BP (2014), BP energy outlook 2035. BP,
London, UK. Available from: http://www.bp.com/
content/dam/bp/pdf/Energy-economics/EnergyOutlook/Energy_Outlook_2035_booklet.pdf
[accessed February 2014].
[6] Kryzanowski, T. (2009), Big energy win with biomass,
enrG Magazine. Available from:
http://www.altenerg.com/back_issues/index.phpcontent_id=231.htm [accessed July 2013].
[7] International Energy Agency (2013), Tracking clean
energy progress 2013: IEA input to the clean energy
ministerial. OECD/IEA, Paris. Available from:
http://www.iea.org/publications/TCEP_web.pdf
[accessed February 2014].
[8] German Biogas Association (2013), Entwicklung
des jhrlichen Zubaus von neuen Biogasanlagen in
Deutschland (Stand 11/2013). Available from:
http://www.biogas.org/edcom/webfvb.nsf/id/
DE_Branchenzahlen/$file/14-07-01_Biogas%20
Branchenzahlen_2013-Prognose_2014.pdf [accessed
February 2014].
[9] Global Methane Initiative, Country profile: China.
Available from: https://www.globalmethane.org/
documents/ag_cap_china.pdf [accessed February
2014].
[10] Swedish Energy Agency (2011), Biogas in
Sweden factsheet. Available from: http://www.
energimyndigheten.se/Global/Internationellt/
Exportfr%C3%A4mjande%20o%20Bilateralt/Biogas_
Sweden_Faktablad_HR.pdf [accessed February 2014].

[11] OECDFAO (2011), OECDFAO agricultural outlook


20112020. Available from: http://www.oecd.org/site/
oecd-faoagriculturaloutlook/48178823.pdf
[accessed February 2014].
[12] Dale,V. H., Kline, K. L., Wright, L. L., Perlack, R. D.,
Downing, M. & Graham, R. L. (2011), Interactions
among bioenergy feedstock choices, landscape
dynamics, and land use, Ecological Applications, vol.
21, pp. 10391054.
[13] Davis, S. C., Boddey, R. M., Alves, B. J., Cowie, A. L.,
George, B. H., Ogle, S. M., Smith, P., van Noordwijk,
M. & van Wijk, M. T. (2013), Management swing
potential for bioenergy crops, GCB Bioenergy, vol. 5,
pp. 623638.
[14] US Department of Energy (2006), Breaking
the biological barriers to cellulosic ethanol:
a joint research agenda. Report from the
December 2005 workshop, DOE/SC-0095.
Department of Energy Office of Science. Available
from: http://genomicscience.energy.gov/
biofuels/2005workshop/2005low_feedstocks.pdf
[accessed February 2014].
[15] Food and Agricultural Organization of the United
Nations. FAOSTAT database. Available from:
http://faostat3.fao.org/home/index.html#HOME
[accessed February 2014].
[16] Adapted from Meiyappan, P. & Jain, A.K. (2012),
Three distinct global estimates of historical land-cover
change and land-use conversions for over 200 years,
Frontiers of Earth Science, vol. 6, no. 2, pp. 122139.
[17] Giovannucci, D., Scherr, S., Nierenberg, D.,
Hebebrand, C., Shapiro, J., Milder, J. & Wheeler,
K. (2012), Food and agriculture: the future of
sustainability. United Nations Department of
Economic and Social Affairs, Division for Sustainable
Development, New York. Available from:
http://www.un.org/esa/dsd/dsd_sd21st/21_pdf/
agriculture_and_food_the_future_of_sustainability_
web.pdf [accessed October 2013].
[18] Kanaly, R., Manzanero, L., Foley, G., Panneerselvam,
S. & Macer, D. (2010), Energy flow, environment
and ethical implications for meat production.
UNESCO, Bangkok. Available from: http://unesdoc.
unesco.org/images/0018/001897/189774e.pdf
[accessed October 2013].

1 Setting the context | 19

[19] Pelkonen, P., Hakkila, P., Karjalainen, T. &


Schlamadinger, B. (2000), Woody biomass as an
energy source challenges in Europe, European Forest
Institute, pp. 7378.
[20] White, E. M. (2010), Woody biomass for bioenergy
and biofuels in the United States a briefing paper,
General Technical Report PNW-GTR-825. US
Department of Agriculture, Forest Service, Pacific
Northwest Research Station, Portland, US.
[21] International Institute for Applied Systems Analysis
(IIASA)/Food and Agriculture Organization of the
United Nations (FAO) (2010), Global agro-ecological
zones (GAEZ v3.0). IIASA, Laxenburg and FAO,
Rome. Available from: http://www.fao.org/nr/gaez/
en/# [accessed February 2014].
[22] WorldBank.org, Water resource management.
Available from: http://water.worldbank.org/topics/
agricultural-water-management [accessed July 2013].
[23] Siebert, S. et al. (2010), Groundwater use for irrigation
a global inventory, Hydrology and Earth System
Sciences, vol. 14, no. 10, pp. 186380.

20 | 1 Setting the context

[24] Portmann, F.T., Siebert, S. & Dll, P. (2010),


MIRCA2000Global monthly irrigated and
rainfed crop areas around the year 2000: a new highresolution data set for agricultural and hydrological
modeling, Global Biogeochemical Cycles, vol. 24,
no. 1.
[25] Curmi, E., Richards, K., Fenner, R., Allwood,
J.A., Kopec, G. & Bajzelj, B. (2013), An integrated
representation of the services provided by global
water resources, Journal of Environmental
Management, vol. 129, pp. 456462.
[26] Williams, E.D. & Simmons, J.E. (2013), Water in the
energy industry: an introduction. BP, London, UK.
[27] van Velthuizen, H. et al. (2006), Mapping biophysical
factors that influence agricultural production and
rural vulnerability, Environmental and Natural
Resources, series 11. FAO & IIASA, Rome.

Land is often classified according to


potential use for agriculture. Broader
climatic and ecosystem classifications
are also used. Both systems are useful
in considering which crops to grow and
where to grow them.

Agriculture involves a series of


operations to produce crops, and modern
techniques are highly mechanized.
Understanding the energy needs and
GHG emissions of these processes is
important for understanding whether
growing biomass crops will be economic
and energy efficient at a particular site.

Different plants employ different

methods for utilizing carbon dioxide and


water, allowing for plant growth under
a wide range of conditions. The way in
which each type of plant grows and
develops affects its productivity and
its potential yield.
Various raw materials derived from

plants including sugars, starch,


cellulose, lignin and oils are used
in energy production. The physical
properties and chemical complexity
of these materials affect the ease and
energy efficiency of processing them
into biofuels.

22

2 Important concepts
Understanding bioenergy production systems requires background
knowledge in many diverse fields. This chapter sets out some of the
fundamental concepts. There is also a comprehensive glossary at the
end of this handbook.

Global ecosystems and land classifications


and sustainability of production will depend on these
land characteristics, together with management
and crop choice.
Land areas can be categorized by grouping ecosystems
that have similar climatic, biotic and abiotic conditions.
Each such group constitutes a biome, and different biomes
give rise to different types of use biomes are described
more fully in Chapter 5. Figure 2.1 summarizes the global
areas of different biomes, the types of feedstock or existing
vegetation, and their carbon productivity and end uses. In
short, the diagram connects biomes to the various services
they provide [1].

Land is commonly classified according to its actual


or potential use for agriculture, and this is sometimes
described as land capability classes. In the context of this
handbook, however, where we review wide-ranging
biomass crops (both potential and realized), we look at land
more holistically because land not normally prioritized for
agricultural use can sometimes support carefully selected
biomass crops. Traditional land classification takes into
account soil type, previous land use, fertility, water
availability, potential for erosion and accessibility for
agriculture. These remain important indicators of land
that might support bioenergy production. Productivity
Figure 2.1

Sankey diagram showing global biomes, vegetation, carbon productivity and end uses[2]. There are a number of nomenclatures currently
used for biomes and those listed on the left represent a slightly modified classification by Ramankutty and Foley[1]. The left-hand column,
slice 1, shows areas of potential natural vegetation, while slice 2 shows actual land use after appropriation for human use. Slice 3 shows
land productivity in Petagrams or Pg (billion tonnes) of carbon per year and also the amount of carbon stored in these lands in Pg. The
final three slices show how this carbon moves through harvesting and processing to final services. Also shown are losses to net primary
productivity (NPP0) attributed to lower productivity on conversion from natural vegetation. Food, fibre and fuel account for only a small
amount of final use. The majority of fixed carbon is available for use by other species, is ultimately respired (heterotrophic respiration) and
contributes to a variety of services, collectively called ecosystem services. The vertical width of each bar in the diagram is proportional to the
area of land use or the amount of carbon per annum associated with a particular segment[3].

Slice 1
Potential
natural vegetation
[M km2 ]

Slice 2
Actual
land use
[M km2 ]

Unit conversion
area to carbon

Built-up
[0.7]

Slice 3
Land productivity
[Pg C/y]

Wood harvest
[2.2]

[30.5Pg C/y]

Cropland
[15.0]

Savanna
[19.2]
Grassland/steppe
[14.3]
Dense shrubland
[6.0]
Open shrubland
[11.9]
Tundra
[15.3]
Desert, polar desert
Rock/ice
[8.2]

Crop harvest
[4.5]

720
gC/m2

Temperate forests
[24.5]
Boreal forests
[8.2]

Pasture
[28.1]

Unmanaged
dryland
[14.8]

Slice 6
Final
services
[Pg C/y]

Fibre [0.6]

400
gC/m2

Fuel [1.2]

[0.84]
[1.8]

Food [0.8]

Meat

Livestock feed
[3.7]

Grazed biomass
[1.8]

Losses

stock
1530Pg C

Food waste

Livestock
respiration [1.8]
Losses

[1.1]

stock
80Pg C

Pollination

Tropical forests

Pest control

NPP0 loss
430
gC/m2

490
gC/m2

50
2
gC/m

Air quality
Water cycling

Temperate forests

[12.2Pg C/y]
stock
310Pg C

NPP0 loss
stock
220Pg C

Fire
[1.4Pg C/y ]

Carbon sink [2.6]


Waste assim.

Fire
[6.0Pg C/y]

[7.2Pg C/y]
Tundra and
desert
[28.9]

Slice 5
Modifications
[Pg C/y]

NPP0 loss

Tropical forests
[22.6]
Forest
[42.4]

Slice 4
Harvesting
[Pg C/y]

Heterotrophic
respiration
Boreal forests
[47.0]

Cropland residual

ECOSYSTEM
SERVICES

Water quality
Flood control
Erosion control
Soil formation

Grasslands

Nutrient cycling
Recreation

Shrublands

Scientific value

NPP0 loss

2 Important concepts | 23

One can see that more than a third of land has been
actively transformed from its natural state to cropland,
pasture and built-up areas. Cropland and pasture are found
in several biomes, and most crop production is harvested
every year. While crop productivity can be quite high, there
are numerous routes for losses in productivity due to
disease, fire, nutrient and climatic variability during
harvest and even waste after harvest. These losses, and the
relatively high fraction of harvest that goes to the relatively
energy-inefficient production of livestock, results in only a
very small fraction of the total productivity providing food,
fuel and fibre. Otherwise unattributed production is
allocated to various worldwide ecosystems.

There are disagreements in the literature as to how land


types should be classified. Nomenclature varies by country
and according to intended purpose. As a result, there are no
internationally standardized definitions of land types and
the quality of land-use data varies between countries. Land
considered by some to be abandoned may actually be used
informally for fuel and grazing. Some even conclude that
no land is really available, as any choice in using land for
a different purpose will ultimately affect the current use
and output of that land, regardless of whether that use is
deliberate. Despite these caveats, it is nevertheless useful
to have a sense of the different land types and the
terminology used to describe their potential productivity.

Land types
The US Department of Agriculture (USDA) describes
eight land classes, based on soil types rated according
to their relative ability to support common agricultural
crops[4]. These are tempered by four sub-categories that
recognize topographic problems that may limit production,
including waterlogging, shallowness of the soil, erodibility
and climatic limitation (such as extreme cold). The
following categories condense these classes (and similar
classifications by other national agricultural services) into
five categories. These are not mutually exclusive: specific
sites may fit one or more of these five categories.

Prime
Prime land can produce the highest yields of major
commodity crops with proper agronomic management.
It is easily accessed and cultivated, and is well suited to
a range of crops, including food, feed, forage, fibre and
oilseed. Despite its great potential, incorrect management
(using techniques that reduce soil organic content or
allow wind or water erosion) of prime land can degrade
productivity. Prime agricultural land is usually rainfed,
rather than irrigated, but irrigation can allow access to
good soils climatically limited by inadequate rainfall. In
this case, careful management is necessary to avoid excess
salt deposition or exhaustion of water sources.

Marginal
Marginal land gives lower yields of annual grain crops than
prime land, or has only limited potential for agricultural
production. It may also be fertile land that is susceptible to
erosion. Even though soil quality may be high, for example,
cultivation of such soil on steep slopes can cause rapid soil
loss unless terraces are economically feasible. Land that
is marginal for conventional row-crop agriculture might
support high yields of biomass crops: sloping land may
support perennial bioenergy feedstocks (given sufficient
water) while minimizing the risk of erosion relative to
other land uses. Low productivity pasture, supporting
one to two head of cattle per hectare, may also be in this
category. In Brazil, with appropriate amelioration of soil
nutrient deficiencies, such land supports significant yields
of sugarcane.

24 | 2 Important concepts

Degraded
Degraded land can support only low productivity of
conventional commodity crops. Degradation is usually the
consequence of intensive management, usually associated
with agricultural or forestry practices that result in the loss
of organic matter, the production of laterites (hard,
clay-like materials) in tropical soils, and actual loss of
bulk soil by water or wind erosion. The dustbowl of the
southern central US, where cultivation led to serious wind
erosion, is a classic example. Such land, however, may
provide viable yields of deep-rooted crops that bind the soil
and have the potential to restore soil carbon. Land may also
be degraded by excessive salt deposition through irrigation
with low-quality water or pollution with industrial
effluents or mining wastes. In some cases, land degradation
is so severe that further agricultural use with traditional
crops is not possible and the land is abandoned.

Abandoned
Land may be abandoned for various reasons, such as
increasingly unreliable rainfall, competition from higher
production elsewhere, or a collapse of local markets. In
the eastern coastal states of the US, for example, much
production became uneconomic following the Civil
War, while the better soils of the mid-western US began
to deliver grain at lower prices, making production
progressively more uneconomic in the north. As a result
much land in the eastern US has dropped out of cultivation
in the past 150 years. This pattern of loss continues to this
day. More recently, increasingly productive agriculture and
the break-up of collective farms in Eastern Europe have
resulted in abandonment.
Land previously used for crops or pasture is classified
as abandoned if it is now unused and has not been
converted to forest or urban use. If the period of
abandonment has been short, such land may be converted
back to agricultural use with relatively low economic
and environmental cost. Long-abandoned land can,
however, through the process of ecological succession,
become host to diverse plant and animal communities.
Re-established native communities often store large
amounts of carbon (both in soils and above-ground
biomass) and may also provide valuable ecosystem
services such as protection of water catchment areas,
wildlife conservation and recreation.

It should be noted that there is no widely agreed definition


of what constitutes abandoned land; indeed many can see
social or environmental benefits in land that others would
deem abandoned. Generally, abandoned agricultural lands
acquire more value as native ecosystems as time since
abandonment goes on. Thus, the most available lands for
agricultural development are those that have been
recently abandoned.

Reclaimed
Reclaimed land includes previously abandoned or
degraded land that is brought back into agricultural use.
It includes large spoil heaps (refuse materials from
mining), which can be contoured and planted with
appropriate pioneer species. In some cases, land can be
reclaimed from the sea, as in the Netherlands.

Agricultural inputs and practices


Agriculture involves management that varies according
to crop species, soil conditions, climate, topography
and culture. Some of the most common practices for
cultivating crops are explained below.
Tillage
Tillage is the preparation of land to receive crops, and is
practised in many forms. For annual crops, it must be
done every year (unless reduced or zero-tillage methods,
as described below, are used). Traditionally, it involves
turning the soil with a plough to bury weeds or crop
residues, followed by harrowing to produce the fine
tilth necessary for good seed germination. It is a major
activity in agriculture, requiring either considerable
labour or mechanization (with its concomitant cost in
GHG emissions). Mechanized tillage has enabled huge
increases in productivity by increasing the speed of
ground preparation and by allowing more land to come
under cultivation.
Tillage improves the conditions of certain soils
unfavourable for agricultural production and is
effective in helping to control annual weeds, but if used
improperly can cause serious land degradation (erosion).
Reduced- or zero-tillage methods, where seed is sown
(using direct-drilling equipment) into unploughed land
that has been cleared (by hand or using herbicides),
are widely used to reduce cost, maintain soil carbon
levels and reduce erosion. Many perennial bioenergy
crops (such as sugarcane and miscanthus) are not sown
as seeds, but the soil must still be prepared to accept
the cuttings or propagules (see below) that are used
to establish new plantings. Once perennial crops are
established, no yearly tilling is required.
Propagation
Crops can be established by various means. Annual
crops (see below) are typically grown from seed, while
many perennial crops are grown from stem cuttings
(e.g. poplar and sugarcane); by dividing roots or rhizomes
(e.g. miscanthus); or from vegetatively produced
plantlets (e.g. agave). Propagation can require significant
labour. Mechanization can improve efficiency and
potential yield, and reduce the amount of labour
required; however, mechanization has a cost in terms of
increased GHG emissions and soil compaction. Tissue
culture provides a means for standardized propagation of
plants with uniform genetic material, but this technique
requires significant investments of time and specialized
labour.

Fertilizer
Fertilizers provide nutrients needed by plants, and
farmers apply them to maximize crop productivity. The
three major plant nutrients are nitrogen (N), phosphorus
(P) and potassium (K). Nitrogen is the primary nutrient
limiting crop production, and the use of synthetic
fertilizers providing nitrogen has been necessary to
support the great increase in crop production enabled
by improved genetics and mechanization. However,
the manufacture and application of synthetic fertilizers
cause significant GHG emissions, partly as a result
of the energy required to produce them and partly
because soil emissions of nitrous oxide (N2O) a potent
GHG and other nitrogen oxides (NOx) are higher after
application. In addition, nitrogen that is not taken up by
the crop can be leached as nitrate into waterways, which
can reduce water quality and cause eutrophication
(a decrease in dissolved oxygen caused by algal blooms
that are detrimental to oxygen-requiring organisms).
The amount of leaching depends on the crop, soil and
climate, and the fertilizer application rate and method.
Irrigation
Irrigation is the application of water to growing
plants to eliminate water deficits. Without sufficient
water, photosynthesis slows down, secondary
physiological effects impede growth, and crop yields
can be dramatically reduced. There are many ways of
providing extra water, but the most efficient methods
supply water very close to plant roots by using buried
drip tubes, thus reducing water loss by evaporation.
Well-designed irrigation systems provide water at the
optimum times during crop growth, and avoid both
under- and over-irrigation. Agriculture accounts for
70% of water withdrawals worldwide and water can,
in some areas, be the major limitation to productivity.

2 Important concepts | 25

Pesticides and herbicides

Intercropping

Pesticides are chemical compounds used to protect


crops from weeds and pests, including insects and other
herbivorous organisms, microbial diseases and viral
infections. Herbicides are pesticides used to control
weeds that compete with a crop for light, water and
nutrients.
Although producing and applying these chemicals
requires energy, these energy costs are much lower than
for tillage and fertilizer application. Also, protecting
crop yields by using these chemicals avoids the need to
plant additional acreage. Some pesticides and herbicides
can be toxic to humans and other organisms, so they
must be used with care. Attempting to control the
growth of pest organisms at a large scale invariably gives
rise to the selection of variants that are resistant to the
control method. As a result, continuing innovation
and improved pest control management schemes are
required to provide effective pest control. Many of the
more modern pesticides and herbicides have much
lower usage rates and much lower toxicity than earlier
chemicals. The advent of genetically modified (GM)
crops that enable the use of modern herbicides or that
mitigate the use of some of the older, more persistent
pesticides, have also contributed to more effective pest
management.

Intercropping is the practice of growing two or more


crops together. It may be used, for example, to provide
protection for a crop that takes time to establish. Crops
must be chosen carefully to reduce competition and
minimize management costs. Owing to the complexity
and costs involved, intercropping has had only limited
application in modern agriculture.

Crop rotation
The practice of periodically alternating the type of crop
grown on a field is known as crop rotation. Farmers
have long known that land maintains its fertility better
if crops are alternated, as plants have different nutrient
requirements and some (such as legumes) can increase
soil nitrogen levels and reduce the need for synthetic
fertilizer. Alternating crops with different root systems
can improve soil structure and tilth. Crop rotation can
also help prevent the build-up of pests and pathogens
that can occur with continuous planting of a single crop.
The annual rotation of maize and soybeans, for example,
has been practised extensively and helps to control corn
rootworm, though it has been replaced in some areas
by the advent of rootworm-resistant corn varieties. Soil
nitrogen and soil organic matter can be enhanced by
planting crops such as clover, and incorporating them
into the soil prior to planting follow-up crops such
as wheat. Traditional crop rotations can be complex,
involving numerous crops and periods of fallow (idle,
non-planted fields) over several years, but high-intensity
agriculture has tended to replace these extensive
rotation systems with new agricultural practices,
fertilizers and improved crop varieties.

26 | 2 Important concepts

Harvesting
To accomodate the multi-year growth period required
in plantation crops such as trees, harvesting can be
staggered by location so that any given plot of land
might not be harvested for many years. When crop
rotation is practised, differential timing of harvesting
will take place depending on the rotation scheme, e.g.
one season may be fallow with no crops harvested.
Otherwise, harvesting takes place one or more times
per year, depending on growth cycle and environment.
Some crops, such as sugarcane, are harvested fresh
and processed immediately. More typically, crops are
allowed to senesce, cease production and dry (partially
or almost completely) in the field prior to harvesting.
This minimizes the energy costs of harvest by reducing
the weight of water transported from the fields while
maintaining the solid contents of the plants. A number
of plants considered for energy production are perennial
grasses with rhizomes (subterranean plant stems). For
these plants, field-drying also allows nutrients such as
nitrogen to be translocated into the rhizomes where
they will be stored over the dormant season and used
again in subsequent growing seasons. This reduces the
need for fertilizers.
Coppicing
Coppicing is a method of producing many stems from
a single tree, by cutting young trees repeatedly to near
ground level every three to eight years. A coppiced
tree produces numerous stems that can be harvested
repeatedly without replanting. Numerous variants exist
that allow for differing intensities of harvest.
Machinery
Modern, high-yield agriculture is highly automated
and requires expensive machinery for tilling, planting,
irrigating, fertilizing, applying pesticide and harvesting.
Increasingly, so-called precision agricultural practices
are being used. These involve global positioning devices
and sensors to guide machinery able to apply fertilizer
and pesticides in amounts tailored to each crop and
soil, at resolutions of a few meters. Machinery allows
significant labour productivity: a large, modern combine
harvester can harvest maize at the rate of almost 10
hectares per hour. Co-evolution of genetics, agricultural
practices and machinery have driven agricultural
productivity to its current state of development, and
there is still scope for improvement.

Plant functional features


Plants can be grouped in various ways, including by
morphology and physiology. Some useful groupings
are discussed below. They are not mutually exclusive,
but characterize different aspects of a plants growth,
metabolism or constitution. For instance, soybean is an
annual, herbaceous, nitrogen-fixing plant that uses the
C3 metabolic pathway for photosynthesis; switchgrass
is a perennial, herbaceous, non-nitrogen-fixing plant
that uses the C4 metabolic pathway.

Life cycle
Annual
Annual plants complete their life cycle within a single
12-month period. They are therefore sown, and later
harvested, every year. This requires recurrent ground
preparation, fertilizer application, and weed and pest
management. With some annual crops, more than one
cycle of growth can be accomplished within a year in
suitable climates.
Examples maize, rice, wheat
Input levels high
Biennial
The least frequent type of crop, biennial plants normally
take two years to complete their life cycle. The first year is
devoted to growing leaves, roots and stems. When colder
months arrive, the plant becomes dormant. Some biennials
require this period of cold before they can flower. During
the following spring and summer, the biennial plant grows
rapidly, ultimately producing flowers and seeds before
dying off.

Examples sugarbeet, onion


Input levels high

Perennial
Perennial crops live for a number of years even for
centuries (many tree species) or occasionally millennia
(bristlecone pine). In moist tropical environments, growth
is possible throughout the year but in colder or drier places
perennial plants use various strategies to survive winter
or the dry season. Trees have woody aerial parts (buds
on branches and stems) from which growth continues
when conditions are suitable. Other plants die back to
underground organs (roots or underground stems) during
the off-season. As perennial crops are only replanted after
a number of years (the exact time varies widely between
species), the cost and overall environmental impact of
tillage and planting are lower than for annual crops.

Examples sugarcane, oil palm, jatropha, willow,


miscanthus, switchgrass
Input levels generally lower than for annual crops

Not all plants can be exclusively categorized as annual,


biennial or perennial, as the life cycle of some species is
dependent on climatic conditions. Breeding strategies
have also been employed to alter the life cycle of some
species. There are wild soybean lines that are perennial, for
example, but soybean is grown as an annual in commercial
agriculture.

Photosynthetic carbon fixation pathways


All plants grow by capturing atmospheric carbon dioxide.
Energy from sunlight is harnessed to convert carbon dioxide
and water into carbohydrates (a store of chemical energy)
and structural components. The annual cycles of carbon
dioxide exchange into and out of the oceans and land are
enormous: approximately 770Pg (billion tonnes) of CO2 are
released to the atmosphere from the land and oceans, and an
additional 29Pg of CO2 from fossil-fuel burning and manmade land-use changes. These releases are not quite
compensated by the approximate 790Pg of CO2 that are
absorbed by vegetation and the oceans[5] (see Figure 2.2).
The yearly excess of CO2 introduced into the atmosphere
has given rise to increases in atmospheric CO2
concentrations from about 315 parts per million (ppm) in
1960 to almost 400ppm today, with worrying implications
for global warming. The importance of photosynthesis
can be readily seen in the atmospheric record as yearly
oscillations of the CO2 concentration. It reaches a
minimum in the northern spring and summer and a
maximum in northern fall and winter: as vegetation, which
is predominantly located in the northern hemisphere, starts
growing and fixing CO2 in the spring, and then annual
vegetation dies back and releases CO2 via decomposition in
the fall and winter [6]. Agriculture contributes a noticeable
(Figure 2.1) and growing (Figure 1.10) proportion to CO2
exchange dynamics.
There are three metabolic pathways for carbon fixation
that provide different responses to the environment (solar
radiation, moisture, temperature, atmospheric CO2
concentrations). As a result, plants can adapt to a wide
range of growing conditions throughout the world.
C3
This is by far the most common carbon fixation pathway,
found particularly (but not exclusively) in plants of
temperate regions. It is called C3 because the initial
molecule formed on capture of CO2 has three carbon
atoms. During periods of light, plants absorb CO2 directly
through open stomata (leaf pores); the open stomata also
allow water loss. As a result, these plants can have lower
water-use efficiencies compared with plants that use
different photosynthetic mechanisms.
Examples rice, wheat, potato, soybean, most trees
C4
The C4 pathway may have evolved as a mechanism to
help plants survive in conditions of drought or high
temperature. In these plants, the first product of CO2
fixation has four carbon atoms, and the mechanism of
this fixation is more efficient than for C3 plants. In effect,
C4 plants concentrate CO2 internally using energy to
do so and then incorporate it into other products using
mechanisms similar to C3 plants. As a result, C4 plants
tend to have higher rates of photosynthesis and wateruse efficiency than C3 plants in warm climates, and can
assimilate CO2 at significant rates while losing less water
than C3 plants. This group includes some of the most
productive tropical crops.
Examples maize, sugarcane, sorghum, miscanthus
2 Important concepts | 27

Figure 9 (5 September 2013)


Draft produced by ON Communication

Figure 2.2
Schematic diagram of the carbon cycle.
Carbon dioxide in the atmosphere is
taken up by plants and converted, using
solar energy through the process of
photosynthesis, into organic compounds.
Some of these organic compounds are
then used as food by herbivores and
humans, whose respiration returns
CO2 to the atmosphere. CO2 is also
returned to the atmosphere when carbon
compounds are burned as fuel. Fossil fuels
were formed as a result of photosynthesis
millions of years ago[7].

Carbon
dioxide

Sunlight

Auto and
factory
emissions

Animal
Soil
Plant
respiration respiration respiration

Photosynthesis

Ocean
exchange

Organic carbon

Decay
organisms

Dead organisms
and waste products

Fossils and fossil fuels

CAM

 Figure 9
The Carbon Cycle

Some plants living in arid environments have evolved


a photosynthetic pathway called crassulacean acid
metabolism (CAM). This metabolism operates in four
stages and involves some enzymatic pathways that
are similar to C4 metabolism. The difference is that,
to conserve water, the stomata of CAM plants remain
closed during the heat of the day. They open during the
night, which allows carbon dioxide to diffuse into the
leaves, to be converted into a four-carbon acid, and then
stored until daybreak. Then, during the day, the fourcarbon acid is broken down and CO2 is released internally
to be converted to useful products while light energy
is available. CAM plants can have greater water-use
efficiency than either C3 or C4 plants.
Examples agave, cacti, pineapple

Nitrogen fixation
Some plants have a symbiotic relationship with bacteria
that transform atmospheric nitrogen into a form that
plants can use. Legumes (which include the bean family)
are widely grown N-fixing agricultural crops, and are
often used in crop rotations. Nitrogen fixation also
occurs in free-living bacteria called diazotrophs that are
most commonly found in soil, but are sometimes found
in plant tissue.
Examples soybeans, trees from the genera
Alnus and Acacia

Plant cell walls


Most of the bulk of biomass is actually plant cell walls,
which serve as structural material. Other components of
plant cells used for bioenergy, such as sugars, starch, oils
and so on are storage reserves for the plant. Details vary by
species, but in general, mature plant cell walls consist of

28 | 2 Important concepts

Root
respiration

Ocean
uptake

a middle lamella (the area between adjacent plant cells), a


primary wall and a secondary wall. The major components
of the cell walls, as shown in Figure 2.3, are carbohydrates
(cellulosic microfibrils and hemicelluloses) and a matrix
of cementing material usually pectin in primary walls
and lignin (a complex polymer that gives strength and
rigidity) in secondary walls. (These compounds are
described below.) The secondary cell walls only develop
once the cells have stopped dividing and expanding, and
are thus seen in non-growing tissues like the wood of trees.
Secondary cell walls make up the majority of biomass,
with cellulose, hemicellulose and lignin the predominant
constituents.
Woody plants
Woody plants trees and shrubs are generally richer in
lignin and hence lower in carbohydrate than other plants.
At the end of the growing season the trunk (composed
of cellulose, hemicellulose and lignin) persists. After
winter, or at the end of the dry season, new growth is
initiated from the over-wintering tissues (buds) within
the persistent aerial stem and branches. Woody plants
grow prolifically over large areas of the earth and have
traditionally been and continue to be a significant
source of biomass.
Examples poplar, willow
Herbaceous plants
Herbaceous plants are generally richer in cellulose than in
lignin. At the end of the growing season, their aerial parts
partially or completely decay. Herbaceous plants can be
annual, biennial or perennial. In the case of biennials and
perennials, the root system overwinters and produces new
shoots when conditions are suitable for growth.
Examples switchgrass, maize

Figure 2.3

Plant composition and products

Schematic diagram representing the main components


of plant cell walls[8].

The various biomass feedstock crops provide a range of raw


materials from different parts of the plant. These materials
may be used by the plant to store and transport energy
internally, or as food stores (for seedling growth) in seeds,
or as structural components (giving plants their stature,
shape and resilience).

Bioenergy crop
Plant cells

Starch from grain crops


Starch is a polymer of glucose, a six-carbon sugar, found
in high abundance in grain crops. It has a wide variety of
food and material uses, and is particularly important for
fermentation into bioethanol. Starch is relatively easily
converted to glucose in an enzymatic process, and the
glucose is subsequently fermented by yeast into ethanol.
Corn starch, for example, is the primary source of ethanol
produced in the US.

Plant cell wall

Oil from seeds

Cellulose
microfibril

Lignin
Hemicellulose
Cellulose

Many crops have oil-rich seeds containing large quantities of


oil in the form of triglycerides, which are glycerol molecules
with three fatty acids (each typically having 1618 carbon
atoms) attached. These oils (also known as lipids or fats)
are liberated by crushing the seeds, and their extraction is
enhanced by the use of solvents. The triglycerides are very
energy rich approximating to the energy content found in
fossil fuels and can be converted into biodiesel molecules
by reaction with alcohols or into renewable diesel (a form of
biodiesel) by treatment with hydrogen.
Soluble carbohydrates
Soluble sugars are small, usually sweet, molecules found in
plants. They include fructose, sucrose, glucose, mannose and
a large variety of less abundant sugars. They have long been
used in the production of alcoholic beverages and are readily
fermented into bioethanol. Sucrose (a major component of
sugarcane), for example, is obtained at a processing plant by
crushing the sugarcane stem and washing it successively
with hot water. The resulting sucrose solution is then
fermented by yeast to produce an aqueous solution of
ethanol that is in turn distilled to provide nearly pure
ethanol. Sucrose is known as a disaccharide, because it is
composed of two sugar molecules (one of glucose and one
of fructose) linked together.

Sugar
molecules

Structural polymers

Glucose

The stems of herbaceous plants and the trunks of trees are


strengthened by cellulose, hemicellulose and lignin. These
large molecules interact and intertwine with each other to
form lignocellulose. It is this lignocellulose that provides
plants with their structural integrity, and which is the major
source of carbon used for combustion (for example when
burning wood). It is also a major target for technological
approaches that seek to deconstruct lignin into its
components for other uses in fuels and chemicals.
Cellulose
Cellulose, found in the primary cell walls of all
green plants, is one of the most abundant organic
compounds on earth. It is a large structural
biopolymer composed of glucose molecules; it differs
from starch in the way the molecules are linked. This
configuration results in a tight crystalline structure
(due to extensive hydrogen bonding), which makes it
ideal for paper manufacture; vast quantities of it are
used in the global pulp and paper industry.

2 Important concepts | 29

It is also the primary constituent in cotton. However,


because of its structure and close association with
hemicellulose and lignin, it is more difficult to
decompose into its constituent glucose molecules
to facilitate fermentation than is starch. It can be
decomposed by chemical treatment or by enzymatic
action, usually using enzymes derived from fungi that
live on cellulose in nature.
Hemicellulose
Hemicellulose is another constituent of plant cell
walls. It is a cross-linked polymer made up of a variety
of sugars, predominantly xylose, a five-carbon sugar.
While cellulose is a complex of a single type of sixcarbon sugar (glucose), hemicellulose consists of both
five- and six-carbon sugars. The balance of these sugars
varies with the particular species, and hemicellulose
can contain a variety of different five- and six-carbon
sugars. When rendered into its constituent sugars, the

resulting mixture of sugars is more difficult to ferment


than the glucose derived from cellulose, and a number
of thermochemical and enzymatic approaches are being
tested to facilitate the process.
Lignin
Lignin is a significant constituent of both woody
and herbaceous plants, conferring strength, water
resistance and resistance to rot. Lignins properties are
derived from its complex molecular structure (it is an
amorphous, cross-linked, hydrophobic biopolymer). It
is typically found with cellulose and hemicellulose in
plant cell walls, where it helps to bind cellulosic fibres
together and provides structure. It also defends the plant
physically against microbial and fungal attack.
It can be converted by thermochemical means to other
compounds, or can be burned to generate heat and
electricity.

Metrics of biomass productivity


There are various ways of measuring the productivity of
crops and their impacts on ecosystems.
Net primary productivity, or NPP, is the projected total annual amount of plant growth. It is a net rather than a gross
figure because plants use some of the captured energy for
their own metabolism. NPP is measured in terms of mass
per area per time unit, commonly grammes of carbon per
square metre per year (gC/m2/yr). NPP can be used to estimate the potential biomass productivity of, for example,
abandoned crop or pastureland. Estimated above-ground
NPP for world ecosystems are useful in estimating overall
biomass potential on a global scale. As such, NPP is often
used by ecologists, and is inherent in the land productivity
estimates in Figure 2.1
Yield is a term commonly used in agriculture. It is usually
given as the amount of a crop end-product produced in
a defined area, usually for a period of one year or one
growing season. This is easily defined for an annual crop,
but perennial crops can increase in yield to reach a stable
plateau, often followed by gradual decline. In these cases,
it is useful to consider an annualized yield based on
production throughout the growing season(s).

Many units are used throughout the world for measuring


yield: bushels (bu), tonnes, tons, kilograms and pounds
for weight; hectares and acres for area. It is also common
to report yields at particular levels of dryness. Fresh
weight (as harvested), for example, is commonly used for
sugarcane yields (reported in tonnes per hectare); maize
yields are usually reported in bushels (56 pounds) of corn
at 15.5% moisture content per acre; and biomass yields
are commonly referenced as bone dry (or oven dry) tonnes
per hectare. It is extremely important to be very clear
regarding units of yield, as apparently similar units can
give very misleading results. As the dry matter content
is most useful for consideration of biomass potential
and uses, tonnes per hectare (t/ha) of dry matter will be
used in this handbook as the standard yield metric unless
otherwise specified.

Some useful units and conversions for yield


1 hectare = 10,000m2 = 2.47 acres
1 tonne = 1 metric ton = 1Mg = 1,000kg = 1.1023 tons = 2,205 pounds
1 litre = 0.26 US gallons or 0.22 imperial (UK) gallons
Examples
1 bushel corn (15.5% moisture) = 21.5kg at 0% moisture
160bu/acre (15.5% moisture) = 8.5 tonne/ha (dry weight)

= 6.1 tonne starch/ha (at 72% starch content)

1 tonne of sugarcane (70% moisture) = 300kg at 0% moisture


70 tonnes/ha (fresh weight) = 21 tonne/ha (dry weight)

= 10.5 tonne sucrose/ha (at 50% sucrose content)

30 | 2 Important concepts

Energy issues and greenhouse gas accounting


requires other energy inputs. Energy inputs for bioenergy
production include the fuel needed for tillage, harvesting,
processing and transporting the crop. The energy needed
to manufacture fertilizer, herbicides and pesticides should
also be taken into account, as should the energy needed to
convert the harvest to biofuel or bioenergy. Energy inputs
will vary depending on the crop. Representative energy
needs for maize agricultural operations are shown as
examples in Table 2.1.
Most harvesting operations consume significant
amounts of energy; chipping wood is a particularly energyintensive operation. Compared with the amount of carbon
and energy harvested (e.g., the energy content of the maize
harvested at an average yield of 5.2 tonnes/ha is greater
than 90,000 MJ/ha), each operation has a relatively small
impact, but taken together can aggregate to 1015% of the
total energy harvested. While the application of fertilizer
does not require much energy, its production requires
significant energy inputs.

A primary motivation for using biomass for energy is that


plants fix carbon dioxide from the air. Although this
CO2 is subsequently returned to the atmosphere when
the biomass is used as fuel, there is zero net GHG impact.
This simple fact, however, is complicated by the many
different energy requirements for agriculture, all of which
can contribute to energy use and GHG emissions; it is
therefore important to look at the broader picture.

Greenhouse gases*
The earth is covered with a blanket of atmospheric gases.
Some of the gas molecules in the atmosphere absorb
outgoing long-wave infrared radiation and re-radiate
this heat back to the surface, like the glass panels of a
greenhouse (hence the term greenhouse gases or GHGs).
This keeps the earths surface warmer than it would
be without such a protective layer. Important GHGs
include carbon dioxide (CO2), methane (CH4), nitrous
oxide (N2O), water vapour and fluorinated gases (CFC,
HFC, PFC and SF6). These gases with the exception of
fluorinated gases exist naturally, but human activities
(such as burning fossil fuels, deforestation, and largescale agriculture and animal husbandry) have increased
GHG emissions. As a result, the concentration of
atmospheric GHGs is increasing at an unprecedented
rate. In terms of bioenergy crop production, the crop
plants take up atmospheric CO2 as they photosynthesize
and grow. Some CO2 is released through plant and soil
respiration, from the soil during tillage, and when energy
is used for farming operations. In the case of perennial
feedstocks, carbon removed from the atmosphere can
be stored in the short and long term, because belowground biomass is not disturbed by annual tillage and can
instead be incorporated into soil carbon pools.

Life cycle analysis*


This type of analysis was devised to assess holistically the
impacts (often in terms of energy use and GHG emissions)
associated with all stages of a products life. As applied
to biofuel and bioenergy crops, this analysis helps to
ensure that the total GHG emissions associated with the
production and use of the bioenergy crop are lower than
from an equivalent amount of energy derived from fossil
sources.

Energy units
There are many units used to describe energy. These
include calories, British thermal units, kilowatt-hours etc.
The standard international unit for energy is the joule.
This is a very small amount of energy, and so it is often
useful to refer to large quantities, such as 1 million joules,
or 1MJ. Some standard conversions and relative numbers
are provided on pages 6 and 7.

Energy requirements for crop production


Bioenergy crops capture solar energy that can be processed
for use as fuel; however, producing such crops also
Agricultural operation

Energy
(MJ/ha)

CO2 emissions
(kg CO2eq/ha)

Tillage

Ploughing
Disc harrowing
Seedbed preparation
Cultivation

780
210
200
200

62
17
16
16

Planting

Corn planting
Grain drill planting

180
140

14
11

Chemical inputs

Nitrogen fertilizer manufacturing


Fertilizer application
Pesticide application
Lime application

8,700
60
100
60

Harvesting

Corn grain
Soybean
Switchgrass (initial year)
Switchgrass (after establishment)
Trees (felling and skidding)
Trees (chipping)

1,280
1,210
260
330
9,150
16,700

100
98
21
26
740
1,300

Post-harvest

Grain drying
Stover mowing
Stover baling

2,810
200
140

190
17
11

* For further reading, please see references on page 32:


Greenhouses gases [11,12], Life cycle analysis [13,14].

480
4.8
8.1
4.8

Table 2.1
Representative energy needs in megajoule
per hectare (MJ/ha) and CO2 emissions in
kilograms of carbon dioxide equivalent
per hectare (kg CO2eq/ha) for agricultural
operations based on maize agriculture
and for operations with other crops. Not
all operations will be used for any given
production strategy. CO2 equivalents
represent all GHGs involved, taking into
account their different global warming
potential and different atmospheric
residence times, and substituting the
amount of CO2 that would correspond
to the same impact [9, 10]. Nitrogen fertilizer
energy need is based on 150kg/ha fertilizer
application.

2 Important concepts | 31

Chapter references
[1] Ramankutty, N. & Foley J. A. (2010), ISLSCP II
historical croplands cover, 1700-1992, in Hall,
Forest G., G. Collatz, B. Meeson, S. Los, E. Brown
de Colstoun & D. Landis (eds), ISLSCP initiative II
collection. Data set. Oak Ridge National Laboratory
Distributed Active Archive Center, Oak Ridge, TN,
US. Available from: http://daac.ornl.gov//ISLSCP_II/
guides/historic_cropland_xdeg.html [accessed
February 2014].

[8] US Department of Energy, Genomic Science Program.


Available from: http://genomicscience.energy.gov
[accessed February 2014].

[2] Haberl, H. et al. (2007), Quantifying and mapping


the human appropriation of net primary production
in earths terrestrial ecosystems, Proceedings of the
National Academy of Sciences, vol. 104, no. 3,
pp. 1294212947.

[10] West, T. O. & Marland, G. (2002), A synthesis of


carbon sequestration, carbon emissions, and net
carbon flux in agriculture: comparing tillage practices
in the United States, Agriculture, Ecosystems
& Environment, vol. 91, no. 1, pp. 217232. Available
from: http://web.ornl.gov/info/ornlreview/v40_3_07/
documents/article17web_West_Marland_ag_net_
flux.pdf [accessed February 2014].

[3] Bajzelj, B. (2013), University of Cambridge, Sankey


diagram based on underlying data published by
Haberl, H. et al. (2007), Quantifying and mapping
the human appropriation of net primary production
in earths terrestrial ecosystems, Proceedings of the
National Academy of Sciences, vol. 104, no. 3,
pp. 1294212947; and by Ramankutty, N. et al. (2008),
Farming the planet: 1. Geographic distribution of
global agricultural lands in the year 2000, Global
Biogeochemical Cycles, vol. 22, no. 1, pp.119.
[4] US Department of Agriculture, Natural
Resources Conservation Service, National
soil survey handbook, title 430-VI, Part 622
Interpretative groups. Available from: http://
www.nrcs.usda.gov/wps/portal/nrcs/detail/soils/
survey/?cid=nrcs142p2_054242 [accessed July 2013].
[5] Denman, K. L., et al. (2007), Couplings between
changes in the climate system and biogeochemistry,
in Solomon, S. et al (eds.) Climate change 2007: the
physical science basis. Contribution of Working
Group I to the Fourth Assessment Report of the
Intergovernmental Panel on Climate Change.
Cambridge University Press, Cambridge, UK and
New York, US. Available from: http://www.ipcc.ch/
publications_and_data/ar4/wg1/en/ch7.html [accessed
July 2014].
[6] Tans, P., NOAA/ESRL ( www.esrl.noaa.gov/gmd/
ccgg/trends ) and Keeling, R., Scripps Institution of
Oceanography ( scrippsco2.ucsd.edu ).
[7] Adapted from the University Corporation
for Atmospheric Research, Kids crossing,
Cycles of the earth, Carbon cycle. Available from:
http://eo.ucar.edu/kids/green/cycles6.htm
[accessed July 2013].

32 | 2 Important concepts

[9] Adler, P. R., Del Grosso, S.J. & Parton, W.J. (2007),
Life cycle assessment of net greenhouse gas flux for
bioenergy cropping systems, Ecological Applications,
vol. 17, no. 3, pp. 675691. Available from: http://
www.esajournals.org/doi/pdf/10.1890/05-2018
[accessed February 2014].

[11] US Environmental Protection Agency, Overview


of greenhouse gases. Available from: http://www.
epa.gov/climatechange/ghgemissions/gases.html
[accessed July 2014].
[12] Solomon, S. et al (eds.), Climate change 2007: the
physical science basis. Contribution of Working
Group I to the Fourth Assessment Report of the
Intergovernmental Panel on Climate Change.
Cambridge University Press, Cambridge, UK and
New York, US. Available from: http://www.ipcc.ch/
publications_and_data/ar4/wg1/en/contents.html
[accessed July 2014].
[13] US Environmental Protection Agency, Life cycle
assessment (LCA). Available from: http://www.epa.
gov/nrmrl/std/lca/lca.html [accessed July 2014].
[14] International Organization for Standardization (2006),
ISO 14040:2006 Environmental management Life
cycle assessment Principles and framework. ISO,
Geneva.

Bioenergy currently provides around


10% of the total global energy supply.
Liquid biofuels account for less than
3% of transport fuel globally, but new
technologies are being developed to
convert more types of biomass to
liquid fuels, driven partly by
government policies to increase
the use of renewable energy.

Estimates of potential annual bioenergy

production range widely due to differing


assumptions used in an inherently
complex analysis. Carefully weighing
the assumptions, estimates suggest that
150200EJ/yr from biomass could be
available by 2050 which would
represent a significant energy source
and a significant increase from the
current biomass utilization of
about 50EJ/yr.
To realize the potential of bioenergy,

the productivity of energy crops will


have to increase. This will require
improvements in both agricultural
techniques and plant breeding.
Given the great improvements that
have been made to important food
crops such as corn and soybean, it is
reasonable to expect similar progress
with biomass crops.
Improvements will also be needed in

the processes used to convert biomass


into fuel. Important current research
is being conducted on a wide array of
approaches for converting abundant but
complex lignocellulosic materials from
biomass into more useful fuels.

34

3 Bioenergy potential
This chapter discussesthe current and possible future bioenergy
production levels and provides information on the technological
developments needed for realizing the potentials.

Current bioenergy production


Bioenergy currently provides around 10% of the total
global energy supply, equivalent to around 53EJ/yr as
shown in Table 3.1. Oil provides almost a third of global
energy, followed by coal and natural gas. Bioenergy supply
is greater than the total of other non-fossil energy supplies
and, as discussed earlier, is expected to grow at greater
than 3% per year to 2035. Global estimates of final energy
consumption are shown in Table 3.2. Almost half of the
total consumption goes to heat. The significant differences
between supply and consumption in Tables 3.1 and 3.2
are due to inefficiencies in converting the primary energy
supplies into heat, transport fuels and, most particularly,
electricity.
Table 3.1
Global energy supply in 2010 shown in EJ[1]
Primary energy supply 2010
Oil
Coal
Natural gas
Bioenergy
Nuclear
Other*

532EJ
172
146
114
53
30
17

Share of total supply


32%
27 %
21%
10%
6%
3%

* Other includes hydro, geothermal, solar, wind and marine.

Table 3.2
Uses of energy supply in 2010 shown in EJ[1]
Total final consumption

363EJ

Heat
Transport
Electricity
Other*

166
100
64
33

Share of total final


consumption
46%
27%
18%
9%

*Other covers those fuels that are used as raw materials and feedstocks
in different sectors, and are not consumed as a fuel, or transformed
into another fuel.

Table 3.3 shows a breakdown of bioenergy use by sector


in 2010. Traditional use (open fires for cooking and
heating in developing countries) accounts for almost 60%
of bioenergy use, though there is a shortage of reliable
statistics.
In Europe, where heat and power production constitute
large uses of bioenergy feedstocks, biomass accounted
for two thirds of primary renewable energy production in
2009. The bioenergy imported by Europe includes wood
biomass sourced through forestry or dedicated woody
crops, ethanol and biodiesel. Globally, biomass is the third
largest contributor to renewable electricity after wind and
hydro-electric power.

Transport accounts for around 100EJ of final energy


consumption, most of it in liquid form, which constitutes
the major fraction of overall liquid fuel use. There is
a widely recognized need to develop low-carbon fuels
(including biofuels) for transport. Biofuels currently
account for less than 3% of transport fuel globally, but
that percentage has grown over recent years.
Table 3.3
Bioenergy use by sector in 2010 shown in EJ [1, 2]
Bioenergy use by
sector in 2010 (EJ)
Traditional
31.4
Heat
12.0
Transport
2.5
Power

4.6

Other*
Total

3.0
53.5

Notes
Small-scale use for heat and cooking
Approximately 100 billion litres of biofuels
were produced in 2010
331TWh (1.19EJ) of electricity was
produced from the 4.6EJ of bioenergy used
in power generation

*Other covers the energy used in production and processing industries,


e.g. biomass farms, ethanol and biodiesel plants etc. and thelosses in
converting primary energy into final consumption form, e.g. gasification plants,
ethanol and biodiesel plants etc.

There are several routes by which biomass can be


converted into usable energy sources, as shown in the
figure inside the front cover of this handbook. These routes
have varying degrees of efficiency and are the subject of
continuing research and development. Ethanol from sugar
and starch, for instance, is produced at >90% of its
theoretical efficiency, and esterification and hydrogenation
of fats and oils are similarly efficient. Conversion of
cellulosic materials to liquid fuels, however, is currently
less efficient and has higher capital costs. Continued
improvement in the efficiency and cost of alternative
conversion methods will be required for these uses of
biomass to be commercially relevant.
North America is currently the largest producer and
user of biofuel. This development has been driven by
the use of corn for ethanol production as a transport fuel
and has allowed the US to overtake Brazil as the largest
global producer of liquid biofuels. Still, in 2010, biofuels
accounted for only 5% of the energy used for road transport
in the US, while in Brazil this was closer to 23% [2].

3 Bioenergy potential | 35

recently mandated greater biofuel production, and the


United Nations has recognized the economic and
environmental potential of new biomass agriculture and
bioenergy industries.
Multinational developments in bioenergy have led
to expansion in international markets as shown in
Figure 3.1. Solid biomass and vegetable oils are being
shipped from North and South America, Africa and
South-East Asia to Europe, which is currently the largest
importer of bioenergy products (for use in many sectors).

Many of the developments in the use of bioenergy have


been driven by government policies promoting renewable
domestic energy for environmental and political reasons.
In Europe, the use of bioenergy was driven by the European
Commissions Renewable Energy Directive (RED) until
late 2012; in the US, government policy mandates the use
of ethanol in gasoline under the Renewable Fuels Standard
(RFS). A combination of subsidies for ethanol producers
and tax incentives for flex-fuel vehicle purchases has
encouraged bioethanol use in Brazil. China has also
BP Biomass Handbook
Figure 3.2 (18 November 2013)
Figure 3.1
Draft produced by ON Communication

Map of world biomass shipping routes in 2011. Routes for ethanol, wood pellets and vegetable oils and
biodiesel converge significantly on Western Europe [3].

Canada
Western
Europe

United States

Eastern Europe
and Russia
Japan

from
US
from
South
America

Middle East
Malaysia and
Indonesia
Brazil
South Africa

Australia

Argentina

Ethanol

Wood pellets

Vegetable oils and biodiesel

 Figure 3.2

Global potential bioenergy production


Biomass potential is defined as the energy available in
harvested biomass. It therefore depends on both the level
of biomass production (the gross yield of dry matter per
unit area) and the amount of energy assimilated (the
energy value of that dry matter) during plant growth.
There is no single method or accepted approach for
assessing biomass potential. Determining biomass
potential depends on the level of environmental impact,
land-use change, water availability, population and
dietary changes assumed in an individual study. The more
limitations accepted in a study, the lower the predicted
biomass potential. A general hierarchy of biomass
potential opportunity is often used: theoretical, technical,
economic and realistic as illustrated in Figure 3.2.
The theoretical potential of bioenergy whether used
for heat, power or transportation depends on just two
fundamental factors: the primary energy production of the

36 | 3 Bioenergy potential

biomass source and the efficiency of converting this into


usable energy. Theoretical potential can therefore ignore
competing land use and socioeconomic constraints, and
gives estimates that are not particularly helpful for policy
decisions. The technical potential is the harvestable
biomass as limited by ecological, land area, topographic
and agro-technological constraints. The economic
potential is available biomass that meets the demand
for biomass feedstock markets at a commercial price.
Realistic potential is the amount of biomass that could
be produced without negative social, environmental or
economic impacts. Realistic potential also takes into
account technological and market development issues.
Figure 3.2, on the next page, depicts the effect of these
increasing constraints when trying to calculate potential
biomass production.

Figure 11 (5 September 2013)


Draft produced by ON Communication
Figure 3.2
Diagrammatic representation of the
hierarchy of theoretical, technical, economic
and realistic potential of biomass. The large
theoretical potential shown at the base of
the pyramid reduces as more constraints are
included in calculations. Realistic potential
is many times smaller than theoretical
potential, and is difficult to quantify because
many of the constraints are either abstract
or hard to measure.

Social impacts

Available technology

Environmental impacts
Realistic potential
Minimizes impacts.

Economic impacts

Market developments

Economic potential
Available biomass that meets market
demands at the intersection of supply
and demand.
Technical potential
Harvestable biomass limited by ecological, land-area,
agro-technological and topographical constraints.
Theoretical potential
Ignores competing land use and socioeconomic or political constraints.
Can result in large potential values that are not helpful for policy decisions.

Mid (low) band (100300EJ/yr by 2050)


The IEA predicts that the worlds total energy demand
will reach somewhere between 600 and 1,000EJ by 2050.
These studies generally assume that increases in
Different scenarios for bioenergy availability predict that
agricultural yields keep pace with population growth
biomass could supply none of this energy, or all of this
and increasing demand for meat (although there are
energy, depending on assumptions relating to regional
various ways of combining these factors to achieve the
diets, agricultural management, the type of land used
same overall result). Some studies assume conversion of
for energy crops, land conservation policies and more.
grassland areas to bioenergy plantations. Biomass residues
Examples of estimated land areas that could be used for
are also included.
biomass production are shown in Table 3.4, below.
Mid (high) band (300600EJ/yr by 2050)
A recent report from the UK Energy Research
Centre [4] classified these varying estimates according
These studies generally make similar assumptions to
to the following groups:
those described in the mid (low) band description, but also
 Figure 11
assume
that increasing
can freepotential
up areas of
previously
Hierarchy of theoretical,
technical,
economicyields
and realistic
biomass
Low band (0100EJ/yr by 2050)
used for food production for bioenergy crops, with
These studies assume little or no availability of land for
~5001,500 million ha used for bioenergy production.
energy crop production (less than 400 million ha globally),
High band (>600EJ/yr by 2050)
together with a large population eating a meat-rich diet.
Environmental protection measures are seen to limit the
These are typically based on theoretical potentials and
growth of cropland, and in some cases biomass residues
are only achieved with relatively low food demands
are not considered. Agricultural systems are assumed to
(e.g. dietary shifts, increased agricultural production
be generally low-input and extensive, with little or no
and/or low population growth). These estimates are
increase in yields.
also characterized by high levels of agricultural inputs
It is important to note that current bioenergy
(fertilizer, irrigation, energy) and the use of native and
production, which includes wastes and residues, is in the
previously uncultivated areas for bioenergy production.
middle of this range. Therefore, even the higher level of
Residues suitable for bioenergy production are also
availability suggested by the scenarios in this band are
included. These studies assume that more than 2,500
likely to be achievable in practice with modest increase in
million ha are used for bioenergy production globally.
agricultural yields in areas with the greatest potential.
High-band estimates are also associated with the
largest negative impacts on ecosystem services (such as
biodiversity and soil and water quality) and are generally
not considered the best path to sustainable energy. Highband estimates provide upper bounds for estimates of
bioenergy potentials, but mid- and low-band estimates are
more widely viewed as reasonable ranges for bioenergy
production.
Table 3.4
Estimates of the amount of land that
could be used for biomass in the future.
Low estimates (0 100EJ/yr) envisage less
than 400 million ha under biomass crops;
mid-estimates (100600EJ/yr) envisage
new plantations or the use of abandoned
agricultural land; high estimates
(>600EJ/yr) envisage 4,000 million ha
converted from pasture used for growing
biomass crops.

Band

Example study

Estimated land use


(million ha)

Band estimate
(EJ)

Low

Hoogwijk et al. 2003 [5]

390 abandoned crop land

0 100

Mid (low)

Beringer et al. 2011 [6]

140450 from new plantations

100 300

Mid (high)

Hoogwijk et al. 2005 [7]

1,300 abandoned agricultural land

300 600

High

Smeets et al. 2007 [8]

4,000 converted from pasture

>600

3 Bioenergy potential | 37

Figure 3.3
Total
technicalHandbook
bioenergy production potential in 2050 in EJ/yr, by Smeets et al. (2007) . Four different scenarios are modelled and represent
BP Biomass
the
following
assumptions
Figure
3.4 (18
November(from
2013)left to right): rainfed mixed agricultural systems (including pastoral practice) with modern technological
production,
irrigated
mixed
agricultural systems with modern technological production, irrigated cropland with only confined livestock and
Draft produced
by ON
Communication
modern technological production, and irrigated cropland with only confined livestock and newly innovated technology for crop production.
All scenarios assume high feed conversion efficiency[8].
1,548

269
223
111
83
204
168

1,273

39

24 29
5 13

Eastern
Europe

30
13 19 25

75

Western
Europe

North America

31 39

Middle East and


North Africa
347

281
234

282

CIS and
Baltic States

194
158

22 28
2

East Asia

Japan

23 26 31 37

South Asia

162

610

89

117
49

Caribbean and
Latin America

367

World

Sub-Saharan
Africa

114
93

Oceania

Dedicated woody bioenergy crops on surplus agricultural land


Agricultural and forestry wastes and residues
Surplus forest growth

 Figure
3.4 in Figure 3.3, there is a wide range of
As
illustrated
estimates for the potential of bioenergy, depending on
many different factors including crop yields, management,
diet, population growth and conversion efficiencies. All
of these factors will affect the amount of land available for
biomass for energy production. In addition, the impacts
of climate change on overall plant yields and the amount
of land required for food are complex, poorly understood
and are the subject of numerous investigations. A recent
assessment of numerous global agro-economic models
showed that there were large uncertainties in estimates
and large differences between models and between
different geographies. However, in general, these studies
tend to indicate a negative impact on crop production
and an increase in land use required for food crops under
conditions of increased climate change. At the same time,
proper choices of where and how to grow bioenergy crops
produced model results indicating that impacts on food
prices were much smaller for scenarios that included up to
100EJ of production from lignocellulosic biomass than for
those scenarios where this energy was derived from high
carbon fossil fuels [911].
It is clear that biomass offers an opportunity for energy
production but there are many social, political, economic
and environmental conditions that affect the scale of
this production. Because none of these conditions are
static, there is unlikely to be a definitive calculation for
the amount that can be produced. Still, the international
consensus summarized in the IPCC Special Report on
Renewable Energy Sources and Climate Change Mitigation
(SRREN [12]) was that 100300EJ/yr could be achieved from

38 | 3 Bioenergy potential

40 55

biomass by 2050. This takes into account factors such


as population, economic, technology, food, fodder and
fibre demand, climate change, and nature preservation
requirements. The 164 scenarios reviewed in SRREN point
to the most likely range of 80190EJ/yr in 2050.
An analysis completed as part of the Global Energy
Assessment (GEA [13]) (conducted by the International
Institute for Applied Systems Analysis) reports global
technical resource potential in 2050 (160270EJ/yr)
stressing stricter constraints related to possible competing
land demands, problems posed by possible deforestation,
and water availability. Deployment levels for year 2050
were 145170EJ/yr in the GEA.
Taken together, these estimates of approximately
150200EJ/yr from biomass by 2050 would represent a
significant fraction of the total projected energy usage
of 6001,000EJ/yr and a significant increase in current
biomass utilization of approximately 50EJ/yr.
There is a wide range of dispersion and uncertainty
in the complex analyses undertaken to date regarding
global potential of biomass. Uncertain and variable
impacts of climate change on weather patterns make
any projections additionally uncertain and subject to
controversy. As a result, while the total amount of
energy that may be ultimately derivable from biomass
remains difficult to estimate, the estimates that we
do have suggest significant potential. This provides
an incentive to develop additional bioenergy crops
and, while expanding their use, to further develop the
precision of our outlooks with empirical evidence to
help better understand their ultimate potential[14].

How might this global potential be realized?


Figure 3.4
Because of the large number of potential species,
BP Biomass
adapted to a wide range of different conditions,
LandHandbook
use in Brazil, in 2013. Sugarcane utilized approximately
Figure 3.5 (21 November 2013)
9Mha of the 60Mha devoted to agriculture, or about 1% of the
bioenergy crops can be grown across most areas of the
Draft produced by ON Communication
total land area[15].
globe that are capable of supporting plant life. Details
of which crops are suitable for production in different
regions can be found in Chapter 6.
Ultimately, land is used for a variety of purposes and
bioenergy crops will need to compete and coexist with
these multiple uses. Brazil is a country that produces
Forests
a large fraction of its liquid fuels from sugarcane, and
354Mha
demonstrates the types of choices that need to be made.
42%
Brazilian land use in 2013 is shown in Figure 3.4. It has
grown its sugarcane crop from a small base to 9Mha today.
Non-cane
In doing so, choices have been made to intensify cattle
agriculture
51Mha
production on pasture, to modify requirements for growing
6%
sugarcane to be more environmentally sound, to require
setting aside additional land whenever sugarcane acreage
Sugarcane
is increased, and to take steps to preserve forested areas.
9Mha
1%
Sugarcane is an example of a crop that has use as both
Cerrado and other non-forest vegetation
food and fuel. Approximately half the sugarcane that
Urban areas
200Mha
and other uses
is harvested is used to produce sugar, the other half to
24%
38Mha
produce ethanol. Countries with significant acreage in
4%
multiple-use crops such as corn, wheat and sugarcane
have significant ability to respond to unforeseen problems
with production by shifting the ultimate use of these
Pastures
commodities.
198Mha
23%
In many parts of the world, agricultural production
does not meet its full potential as illustrated in Figure
3.5, which shows estimated yield gaps throughout
the world for maize, wheat and rice. Yield gaps can be
quite extensive over large areas of the world; often less
In 2010, for example, average corn (maize) yields were
than 20% of the potential yield is produced due to local
9.6t/ha in North America and 6.0t/ha in Europe; while
conditions. Constraints include access to technology, the
Africa produced only 2.6t/ha[16]. Additionally, many of
most suitable cultivars or best practice; the cost of inputs; Figure
3.5crops most suitable for energy production have had
the
suboptimal growing conditions; lack of transportation
scant attention paid to them, so there is every reason to
infrastructure; and, in many cases, a lack of established
expect significant improvements in yields with diligent
markets to drive demand.
application of modern practices.
By growing a wider range of improved food crops
While it is difficult to predict future crop productivity
with improved management, yields could be boosted
with any certainly, predictions can be made using a variety
substantially and the land area required for food
of approaches. As with potential land areas, yields can
production reduced. This approach, implemented by
be theoretically, technically, economically or practically
region according to the climate, soil and local economies,
feasible. Unlike land, however, which is clearly limited
would also help address the inequity of global food
and whose area cannot be significantly increased, the
distribution. Figure 3.5 provides a view of the substantial
limits of crop productivity are far higher than current
gains that could be derived from improved practices. Those
yields and are amenable to technological innovation.
areas with the lowest achieved yields could benefit most
Intensive agriculture, focused on a few important crops,
from adoption of modern agricultural techniques and strict
has seen dramatic improvements through combinations
attention to optimum nutrient and water utilization. The
of agricultural practices, genetic advances and modern
large areas associated with large yield gaps suggest that
technology. An example of such improvement is shown in
significant improvements in overall production can be
Figure 3.6, which plots US corn yields through time.
obtained.
Different crops in different parts of the world will
have different potentials for yield increases. For example
high-yielding crops such as sugarcane, grown in optimal
conditions in parts of the southern hemisphere, are
already producing significantly higher yields than
before; further yield increases are likely to be gradual.
In contrast, the production of biomass and grain crops
in the subtropics has greater potential, as production
is generally still far below any theoretical maximum
production or productivity thresholds.

3 Bioenergy potential | 39

Figure 3.5
Estimates of yield gaps in terms of percentage of attainable yield achieved throughout the world for maize, wheat
and rice. Gaps were estimated by comparing observed yields to those from areas with high yields within zones of
similar climate using data from around the year 2000[17].

Major cereals: attainable yields achieved (%)

0%

10%

20%

30%

40%

50%

60%

70%

80%

90%

100%

Figure 3.6
Graph showing average US corn grain yields from 1866 in bushels/acre and kg/ha. Important developments in plant
breeding provided improved yield potential, by allowing the switch from open pollinated varieties to hybrids with
increasing sophistication. These improvements allowed increased planting densities, which, coupled with the use
of nitrogen
fertilizer,
crop protection chemicals and mechanization, drove large yield improvements. More recently,
BP Biomass
Handbook
[18]
genetic
modification
has continued
Figure
3.6 (21 November
2013) to extend the yield limits. N.B. 1bu/acre (15.5% moisture) = 53kg (dry-weight)/ha .
Draft produced by ON Communication
180

11,000

160

In
m cre
ec as
ha in
ni g u
za s
tio e o
n fn
an it
d ro
cr ge
op n
pr fer
ot til
ec ize
tio r,
n
ch
e

ic

al

10,000

Grain yield (kg/ha)

8,000
7,000
6,000
5,000
4,000

140
120
100
80
60

3,000

40

2,000

20

1,000
0
1860

Grain yield (bu/ac)

9,000

0
1880

1900

Open pollinated

 Figure 3.6

40 | 3 Bioenergy potential

1920

1940

First generation
hybrids

1960

1980

2000

Second generation
hybrids

2020
Biotech (GM)

A summary of crop production technologies


Land use can be minimized by growing the crop with the
highest economic yield for a particular location. Some
of the ways in which crop yields can be increased are
explained briefly below.
Irrigation
With competition for water resources set to increase,
particularly because of population growth and climate
change, irrigation of substantial areas devoted to energy
crops poses additional risk. Prudent measures to avoid
such risk restrict energy crop production to areas of high
water availability or where precipitation provides the large
majority of needed water.
As mentioned previously, although overall agriculture
accounts for almost 70% of total worldwide water withdrawals, the amounts withdrawn for liquid biofuels production accounts for about 0.5% owing to choices made
regarding crop types and growing locations that do not
tend to use significant irrigation. Significant expansion of
crops for bioenergy production will be most sustainable in
locations that do not require significant irrigation.
Nutrition
Plants need adequate nutrition to achieve their full yield
potential. Both the manufacturing and application of
fertilizer use significant amounts of energy and produce
GHG emissions, which must be factored into the life cycle
analyses of bioenergy production. Some energy crops,
particularly perennials, are highly efficient at recycling
nutrients and have low input requirements but, in other
cases, there are large numbers of nutrient inputs required.
The net benefit of these inputs for increasing energy
yield and displacing fossil fuel emissions can be greatly
reduced by the energy and GHG costs of manufacturing
and applying large numbers of nutrients. There are also
issues surrounding the short- to medium-term availability
of potash (potassium carbonate) and phosphate due to
political factors in areas with key reserves, such as parts of
North Africa [19].
It is perhaps more important that any expansion
in energy-crop production is carried out in a way that
protects nutrients in the soil by reducing erosion during
establishment and cultivation. Returning nutrients to the
crop from processing waste (for example using bottom ash
from biomass combustion as a fertilizer or applying the
digestate from anaerobic digestion) can reduce the inputs
needed.
Crop protection (pesticides and herbicides)
Weeds, pests and diseases are estimated to reduce global
crop yields by as much as a third. The use of pesticides
and herbicides is currently estimated to account for
some 2153% of theoretical crop yield and be some
52% efficient at increasing actual crop yields on food
and fibre crops. Work is now beginning to investigate
potential issues with dedicated bioenergy crops, such
as rust caused by Melampsora in willow, for which
many resistance genes have been identified and used in
breeding. In fact, the introduction of resistant willow
varieties has led to an almost doubling of yields,
illustrating the importance of effective crop protection,
whether by chemical or genetical means.

The use of pesticides and herbicides can be expensive and


have unwelcome environmental costs. As a result, crop
management is generally focused on maximizing value
and minimizing use. There are currently large differences
in the use of chemical protection between bioenergy crops:
crops such as corn and soybean have relatively high rates
of use, while biomass crops such as miscanthus and
willow have relatively low usage rates, although the latter
require greater inputs for weed control at establishment.
An advantage of many energy crops, however, is that they
are relatively undomesticated. Great potential therefore
exists to breed in resistances to pest and diseases from the
wide germplasm available.
Tillage
For annual energy crops, such as cereals and many oilseed
crops, moving from conventional to minimum- or zerotillage systems can reduce energy inputs and increase soil
carbon levels. However, establishment requirements differ
between crops, and best practice must be decided on a
case-by-case basis.
Perennial and woody crops in temperate areas are often
used to increase soil carbon stocks. This is in part because
these cropping systems do not require annual tillage
and the crops turnover roots during the growing season,
adding to the organic carbon in the soil. Provided that
high-carbon-stock land is not converted to plantations
(for example the conversion of tropical peatland for
oil palm production in parts of South-East Asia) and
cultivation methods are carefully chosen, perennial crops
could present a significant opportunity to re-carbonize
agricultural soils in some areas.
Mechanization
Improving the efficiency of planting, nutrient, pest
management and harvesting operations can increase
the harvested yields of biomass and reduce overall cost.
Mechanization may also allow potential sources of
biomass, which would otherwise simply be disposed of
(such as exhausted oil and rubber plantations in West
Africa), to be used for bioenergy.
In general, the large areas of land that will be required
for significant production of bioenergy crops will require
extensive mechanization to be effective and economical.
Genetic improvement
There is considerable scope to increase biomass yields
through the use of plant breeding and genetic modification
(GM). The potential for such increases depends both on
the techniques used and the level of domestication of
the crop species. The practice of selection and breeding
for the most favourable traits in plants has been used
since the beginning of agriculture. GM is a more recent
development for crop improvement, and essentially allows
one to accelerate the process of selection and breeding
by directly increasing the number of gene copies that are
responsible for desirable traits and reducing or eliminating
genes that code for undesirable traits.

3 Bioenergy potential | 41

Many breeding programmes for bioenergy crops


are in their infancy, so there is the potential for large
increases in yield from conventional breeding and genetic
modification. Established agricultural crops such as maize
and wheat are expected to have a lower (though still
significant) potential for yield increases, as they have
already been subject to intensive genetic improvement.
Nevertheless, further improvements could be expected
in crops currently used to produce food by focusing on
attributes that are important for conversion to bioenergy.
GM has been proposed as a key method of improving
bioenergy crop yields and enabling species to withstand
stresses such as drought, salinity, extreme temperatures
or attack by pests and diseases. There is also the
possibility that GM could allow more efficient and costeffective biofuel production, for example by increasing
photosynthetic efficiency or improving the breakdown
of lignin during processing.
The use of GM crops has already been widely adopted
around the world but there is strong resistance to GM
organisms in some markets. The use of the technology
in non-food crops may be less of an issue in certain
circumstances, but costs for developing GM crops and
gaining regulatory approval are high and can prohibit
their widespread use. As previously indicated, many of
the potential bioenergy crops have not been widely used
in agriculture, so there is expected to be broad scope for
their agronomic improvement through both improved
management and conventional breeding strategies.

Potential from new crops


Estimated yields for bioenergy can also be expected
to increase with the introduction of new bioenergy
crops. As conversion technologies mature, there will be
greater flexibility for accommodating different feedstock
compositions. As a result, biomass sources can be
diversified temporally (different harvest dates) and across
local landscapes, regions and the world. There is the
potential to develop new sources of bioenergy both from
species of plants that are not currently commercial crops
and from crops that are currently grown for other purposes.
There are many proven and potential bioenergy crop
species, ranging from widely grown agricultural crops to
niche species promoted as being academically interesting
but that have not been proven in agronomic practice or
commercially. By matching potential biomass crops with
optimum locations, biomass yields could be improved,
the landscape could be diversified, and ecosystem services
might in turn also be improved.
While achieving a diversified landscape of bioenergy
crops is a potential benefit of developing new biomass
crops, it will not be without challenges. There are many
difficulties in taking a wild or semi-wild species and
trying to adapt it rapidly for commercial production, as
has been proved in recent years with numerous attempts
to commercialize Jatropha curcas (jatropha) as an oilseed
feedstock for biodiesel. While companies continue to
investigate the potential of jatropha, early efforts revealed
that, while the shrub would grow in semi-arid areas
with little or no agronomic input, it would not produce
commercially viable yields of oil under these conditions.

42 | 3 Bioenergy potential

Another consideration is the risk of introducing highly


invasive non-native species to an ecosystem. In the past,
such introductions have often escaped from ornamental
horticulture; some species (such as kudzu) now proposed
as potential feedstocks for bioenergy production, are
themselves prolific weeds. In many respects, the properties
of an ideal biomass crop vigorous, perennial, fast
growing and an efficient user of nutrients mirror those of
weedy plants, and care will be needed to ensure that risks
to native communities are minimized. With an awareness
of this risk, diversified novel biomass crops should include
more native species, selected by regional biome types.
The literature describes almost 200 plant species
currently used in some form for the production of
bioenergy, or which have the potential to do so in the
future. We provide an abbreviated list at the end of the
Chapter 6 (see Additional species on page 100) to illustrate
the variety of species currently being considered for
bioenergy. These examples of novel bioenergy crops
species have acquired different levels of scientific interest
and are at varied stages of commercial development. It is
far from being a definitive list of candidate varieties for
bioenergy production.
With the diversity of potential biomass crops, it is
difficult to project the potential production of bioenergy in
the future. This is in part because yields will no doubt be
improved as agricultural practices are tested and developed
for each new crop. We need only look to the history of
current commodity crops. Soybean, for example, is now
grown across both temperate and tropical regions all over
the world. Yet the modern varieties are derived from a wild
plant that was selected for cultivation (see The soybean
story on the next page).

Data and research needs


In order to facilitate the development of bioenergy and
feedstock crops, further research and data are needed.
Genetics
There is a need for better understanding of the genetics and
traits of improved varieties, particularly for new crops that
have not been fully commercialized.
Agronomy
An understanding of the best ways to sustainably produce
economic yields from many of these potential crops
is required. There is also a need to find out how best
management practices can be integrated with existing
agricultural systems and other land uses.
Below-ground biomass and soils
There is currently limited knowledge on the below-ground
biomass (roots and rhizomes) of different cropping systems
and of how soil carbon levels change under different landcover types. These data are key to identifying where best to
grow different bioenergy crops, to understanding land use,
and to evaluating their potential impacts for sustainability
and climate-change mitigation.

Life cycle analysis

Land-use change

While guidelines for life cycle analysis (LCA) have been


developed, there are still discrepancies between the way
LCA is applied to different energy forms and land uses.
There is a need to integrate data collection with life cycle
impacts in a way that is comparable across energy sources
and regions.

Understanding the impacts of land-use change


is important. While complex, it is relatively
straightforward to determine the direct effects caused
by a change in land use from its previous (uncultivated,
pasture etc.) condition to one of bioenergy crop
cultivation (direct land-use change or DLUC). There is
also interest, however, in determining the magnitude
of any indirect effects of using land for a new purpose
such as growing bioenergy crops and there is no
scientific or political consensus on how to treat such
indirect land-use change (ILUC).

The soybean story: an example of crop development


As agricultural systems develop and crop science
improves, so crops can be adapted from their original
use to meet a number of different market requirements.
This can be illustrated by the example of soybean,
which was originally identified as a source of protein
for human food, and is now used for a wide variety of
products, including as a bioenergy feedstock.
Soybean (Glycine max (L.) Merrill) is currently grown
on around 90 million hectares globally, having been
successively improved and distributed from the wild
relative G. soja in China some 4,500 years ago.
For the first half of the 20th century China was still the
leading global producer of soy, used as a food and animal
feed crop. Although the crop was first introduced into
the US in 1760, large-scale production only started in
the 1950s but grew rapidly so that, today, the US is the
worlds largest producer followed by another relative
newcomer, Brazil, where farmers adopted the crop only
in the 1970s.
From Brazil, soybean production spread south to
Argentina (now the third largest producer globally)
and back across the Pacific to India. There is also some
production in Africa.
Crop development has mainly focused on the
selection of plants with large seeds with the first formal
breeding programme originating in China in 1913.
In the late 1920s the US sent crop scientists to the
country to learn about the crop and collect germplasm.
The breeding programme in Brazil was particularly
successful, adapting the crop and its associated nitrogenfixing bacteria to new soils and different photoperiods.

In line with many other crops, the majority of soybeans


are now a product of targeted breeding, including genetic
engineering. Genetic modification to date has sought to
improve yields primarily through herbicide tolerance
and insect-resistance traits. Though mainly grown for
its protein content, new soybean varieties are also being
developed with modified oils, suitable for different food
and bioenergy uses.
As the crop has expanded, global markets have
developed. High-income countries account for almost
half the total consumption. Soy is traded on futures
and options markets around the world, with the main
use as animal feed and cooking oil. Worldwide roughly
20% of soybean oil is used for biodiesel production
despite the low yield relative to other bioenergy crops.
However, illustrating the flexibility of the crop, a
wide range of products are produced around the world:
these include fermented products, juice and fresh
beans. There is no sign that soybean yields are about to
plateau, suggesting it will have a key role to play as a
major crop for some time.
Many bioenergy crops that are already established
in agronomic practice, or that are close to becoming
so, require developed markets in order to drive the
necessary technical development and investment. In
most situations where the widespread production of
crops for bioenergy has been undertaken by farmers,
government initiatives have boosted demand through
programmes such as those that support the use of
ethanol in Brazil and the US. Debates around the
benefits of bioenergy and changes in government policy
can act as a strong disincentive to such developments at
all levels of the supply chain.

3 Bioenergy potential | 43

Developments in biomass conversion technologies


Biomass is often highly variable and diffuse when
harvested, making it expensive to transport and less
predictable as a feedstock than conventional fuels. There
are several processes to prepare biomass into convenient
and usable energy sources, including techniques that
improve the handling and logistical properties of biomass
feedstocks before conversion. Processes currently in use
are outlined in Table 3.5.
This table provides a brief description and key
characteristics of the major stages in processing biomass.
In practice, these are stages that would be included in a
bioenergy production chain, either in the same facility
(such as during the production of biomass-to-liquid (BTL)
fuels) or to facilitate logistics (such as pelletizing solid
biomass fuels before transport). Many of these different
technologies can be used together in different process
variants. Additionally, these processes are at very different
stages of their technological evolution, with combustion,
for example, being very highly refined and extensively
practised whereas others, such as lignocellulosic
hydrolysis and fermentation, are just starting to appear
commercially.
To achieve full commercial potential and continuing
investment in the sector, current efficiencies and
technologies will need to be improved.

44 | 3 Bioenergy potential

The two most important trends in biomass utilization


are greater electrical power production and initiatives to
develop cellulosic and advanced biofuels. Both of these
developments will increase demand for lignocellulosic
feedstocks. At the same time, new technologies are likely
to increase conversion efficiency, meaning that more
energy can be produced from a given quantity of biomass.
This will reduce the inputs and land requirements to
produce a given amount of energy or fuel.
Currently, bioenergy is used to produce around 1.2EJ of
electrical power. New installations and the conversion of
coal-fired power stations to biomass fuel could increase
this to 5.4EJ by 2035[2]. Much of this activity is expected to
occur in the US and Germany, although other parts of the
European Union (EU) and Asia are also developing projects
to produce electricity from bioenergy.
By the end of 2013 there were numerous cellulosic
biofuel production projects around the world, ranging
from small demonstration projects to full-scale facilities
just commencing commercial operations. Though
technological difficulties in developing robust, economical
processes have resulted in a slower-than-expected
evolution, currently initiated commercial operations
should substantially increase the production of cellulosic
biofuels in the near future. Most of these plants are
planned for the Americas, although Europe and Asia are
also key locations for companies involved in the sector.
These include oil producers, current biofuel producers,
biotechnology companies, and dedicated lignocellulosic
and advanced biofuel companies. Most are looking to
produce bioethanol or other alcohol-based fuels from
cellulosic materials, with diesel, intermediate feedstocks
and other chemicals representing the other outputs.

3 Bioenergy potential | 45

After various pretreatments to liberate


carbohydrates, numerous organism and
process variants are being investigated for
their ability to convert cellulosic feedstocks
into alcohol-based fuels.

Lignocellulosic hydrolysis
and fermentation

Biological route to biogas

Thermochemical routes
to fuels or power

Anaerobic digestion

Pyrolysis

A biological process where consortia of


microbes break down solid biomass into
methane and CO2 via successive processes
of metabolism.

A thermochemical process where the


feedstock is subjected to high temperatures
(~475490oC) in the absence of oxygen to
produce a liquid bio-oil, a light syngas or a
solid biocharcoal (biochar).

Thermochemical process using high


temperatures (6001,100oC) to turn biomass
into a syngas in the absence of oxygen.

A bio-chemical process (usually aided


by physical/chemical preparation of the
feedstock) where sugars are fermented
into products (particularly ethanol). Where
starch-based feedstocks are used, an
additional stage saccharification is
required before fermentation where the
starches are converted into simpler sugar
molecules by enzymatic hydrolysis.

Hydrolysis and
fermentation

Gasification

The chemical process of adding hydrogen


to vegetable oils to create hydrocarbon
chains.

Hydrogenation

Biological routes
to bioethanol from
carbohydrates

A chemical process where an alcohol (such


as methanol) is combined with a vegetable
oil to produce a fatty acid alkyl ester.

Transesterification

Chemical routes to
biodiesel from oils

Thermochemical process where biomass is


burnt in the presence of oxygen to produce
heat.

Combustion/co-firing

Thermochemical process where water and


other volatile compounds in the biomass
are removed in the absence of oxygen at
temperatures of 200320oC.

Torrefaction

Description

Physical process.

Pelletization

Technology / technique

Conversion to heat
and/or power

Pre-processing

Most of these processes are designed to


produce alcohol-based fuels.

Usually high-sugar or high-starch


feedstocks such as sugarcane, maize and
wheat.

Cellulosic biomass.

Fully commercial with significant research


and development activities under way to
produce biofuels other than ethanol.

Some pilot plants exist but there is little


commercial-scale activity and the first
commercial facilities are just beginning to
operate.

Fully commercial.

Pyrolysis of industrial waste has been


practised commercially but pyrolysis
using biomass is just beginning to be
commercialized.

Most effective on wet biomass sources


including green energy crops, manures,
sewage and food wastes.

Any biomass, including wastes.

Generally cellulosic biomass.

After fermentation the ethanol is generally


distilled to remove surplus water before
it can be used as a transport fuel. Further
molecules for gasoline, diesel and industrial
chemical uses are at the development and
early commercialization stages.

Vegetable oils and waste oils from cooking.

Commercial and growing.

Some trial gasification plants have been


established to produce electricity or liquid
biofuels, although there have been few
large-scale commercial successes.

The resulting hydrocarbon is suitable for


use as a transport fuel in compression
ignition engines.

Vegetable oils from crops such as


rapeseed, soya and oil palm and waste oils
from cooking.

Fully commercial.

Biogas may be burnt to generate heat and


power.
It can also be processed for use as a
transport fuel (CNG) or cleaned prior to
injecting into gas distribution networks.

Pyrolysis oils and other products may be


turned into liquid transport fuels via further
processing or may be burnt to produce heat
and/or electrical power.
Biochar may be burnt or, in some process
variants, used as an amendment to
improve soil.

The resulting syngas can be burnt to


produce heat and/or power. Alternatively
it can be turned into a liquid fuel (or other
renewable chemicals) using the
FischerTropsch catalysis methods.

The resulting alkyl ester is suitable for use


as a transport fuel in compression ignition
engines.

Can be used to generate heat or, when


combined with steam or Organic Rankine
Cycle generators, can be used to produce
electricity.

Most dry biomass.

Fully commercial.

Produces a robust densified fuel suitable


for later combustion. Torrefied fuels have a
greater energy density than the materials
from which they were derived and are
hydrophobic.

Produces a solid fuel for further processing


(generally via combustion or other thermal
treatment).

Energy outputs

Generally woody biomass is used, but could


be applied to other cellulosic biomass
crops.

Any dry millable biomass, including wood


(once turned into sawdust), cellulosic
biomass and cereals.

Suitable feedstocks

Technically proven and some pilot-scale


plants in existence.

Commercially viable and widely carried out.

Current status

Table summarizing current biomass conversion technologies. These range from pre-processing through conversion to heat
and power, conversion of oils to fuels, biological routes, thermochemical routes and the conversion of biomass to gas.

Table 3.5

Chapter references
[1] International Energy Agency (2012), Key world energy
statistics 2012. OECD/IEA, Paris.
[2] International Energy Agency (2012), World Energy
Outlook 2012. International Energy Agency (IEA),
Paris.
[3] International Energy Agency (2011), adapted from
IEA technology roadmap: biofuels for transport, p. 30.
OECD/IEA, Paris.
[4] Slade, R., Saunders, R., Gross, R. & Bauen, A. (2011),
Energy from biomass: the size of the global resource.
Imperial College Centre for Energy Policy and
Technology and UK Energy Research Centre, London.
Available from: http://www.ukerc.ac.uk/support/Ener
gy+from+biomass%3A+the+size+of+the+global+res
ource [accessed February 2014].
[5] Hoogwijk, M., Faaij, A., van Den Broeka, R.,
Berndes, G., Gielen, D. & Andturkenburg, W. (2003),
Exploration of the ranges of the global potential of
biomass for energy, Biomass and Bioenergy, vol. 25,
pp. 119133.
[6] Beringer, T., Lucht, W. & Schaphoff, S. (2011),
Bioenergy production potential of global biomass
plantations under environmental and agricultural
constraints, GCB Bioenergy, vol. 3, pp. 299312.
[7] Hoogwijk, M., Faaij, A. & Eickhout, B. (2005),
Potential of biomass energy out to 2100, for four IPCC
SRES land-use scenarios, Biomass and Bioenergy, vol.
29, pp. 225257.
[8] Smeets, E. M., Faaij, A. P., Lewandowski, I. M. &
Turkenburg, W. C. (2007), A bottom-up assessment
and review of global bio-energy potentials to 2050,
Progress in Energy and Combustion Science, vol. 33,
no. 1, pp. 56106.
[9] Muller, C. & Robertson, R. D. (2014), Projecting
future crop productivity for global economic
modeling, Agricultural Economics, vol. 45, no. 1,
pp. 3750.
[10] Schmitz, C. et al. (2014), Land-use change trajectories
up to 2050: insights from a global agro economic
model comparison, Agricultural Economics, vol. 45,
no. 1, pp. 6984.
[11] Lotze Campen, H. et al. (2014), Impacts of increased
bioenergy demand on global food markets: an AgMIP
economic model intercomparison, Agricultural
Economics, vol. 45, no. 1, pp. 3116..

46 | 3 Bioenergy potential

[12] Edenhofer, O., Pichs-Madruga, R., Sokona, Y. &


Seyboth, K. (eds) (2011), Special report on renewable
energy sources and climate change mitigation. IPCC,
Cambridge University Press. Available from: http://
www.ipcc.ch/pdf/special-reports/srren/SRREN_Full_
Report.pdf [accessed October 2013].
[13] International Institute for Applied Systems
Analysis (2012), Global energy assessment: toward
a sustainable future. Cambridge University Press,
Cambridge, UK and New York, US, and the IIASA,
Laxenburg, Austria. Available from: http://www.iiasa.
ac.at/web/home/research/Flagship-Projects/GlobalEnergy-Assessment/Home-GEA.en.html [accessed
October 2013].
[14] Slade, R., Bauen, A., Gross, R. (2014), Global
bioenergy resources, Nature Climate Change, vol. 4,
pp. 99105.
[15] Nassar, A. M., Moreira, M. (2013), Evidences on
sugarcane expansion and agricultural land use
changes in Brazil, ICONE (Institute for International
Trade Negotiations). Available from: http://www.
iconebrasil.com.br/publication/study/details/568
[accessed March 2014].
[16] Food and Agricultural Organization of the United
Nations. FAOSTAT database. Available from: http://
faostat3.fao.org/home/index.html#HOME [accessed
February 2014].
[17] Mueller, N. et al (2012), Closing yield gaps through
nutrient and water management, Nature, vol. 490,
pp. 254257.
[18] Nielsen, R.L. (2013), Historical corn grain yields
for Indiana and the US, Crop Science, vol. 46,
pp. 528543. Available from: http://www.kingcorn.
org/news/timeless/YieldTrends.html [accessed
February 2014].
[19] Zepf V., Reller A., Achzet B., University of
Augsburg, Rennie C., BP, Ashfield M. & Simmons
J., ON Communication (2014), Materials
critical to the energy industry an introduction,
2nd edition. Available from: www.bp.com/
energysustainabilitychallenge [accessed
February 2014].

48

Significant expansion of bioenergy will


require continuing improvement in the
cost basis of biomass production and
its conversion to fuel. Major financial
investments are needed to develop new
technology and to improve the variety
and performance of biomass crops.

If biomass is to make a significant


contribution to world energy needs,
it will have to be produced on a
large scale and without detriment to
the environment. This will require
reconciling competing demands for
food, water, energy and other resources.

Management choices regarding which


crops to grow in which locations can
completely determine whether biomass
production is beneficial or detrimental
to GHG emissions.

Supportive and consistent government


policies are needed to provide a degree
of certainty for industry to undertake
the serious financial investments
required to further develop biomass
potential.

4 Economics, the environment and politics


This chapter provides an overview of the main issues
associated with large-scale use of biomass from
socio-economic, environmental and political perspectives.

The socio-economic drivers and impacts of bioenergy


Biomass provides energy to millions of people in the form
of cooking fuel, heating, electricity and transportation fuel.
The potential benefits are many and varied as outlined in
Table 4.1. Bioenergy contributes to individual households
and communities as well as national and international
energy portfolios.
There are several key factors critical to understanding the
socio-economic impacts of bioenergy:

With the exception of municipal waste, biomass is


largely a rural resource.
Land and water resources that are needed to support
biomass production may be privately or publicly owned
and may therefore be subject to different governance and
governance related to other uses.
Biomass can be expensive to transport but bioenergy
products such as fuel and electricity can be exported and
traded regionally and internationally.

Because biomass is bulky and relatively expensive to


transport, use for bioenergy tends to be a local activity
(generally within 100km). This means that bioenergy
opportunities are location dependent, as described in other
chapters of this handbook. Because most biomass resources
are in rural areas with lower economic opportunities on
average, bioenergy systems can have large positive effects,
including improved standard of living through access to
energy and jobs.
In many areas, bioenergy systems contribute to
maintaining or creating rural employment. New jobs may

arise directly in the production and processing


of bioenergy crops, and also indirectly in crop research
and development. Increased demand for agricultural
and forestry residues can increase incomes in these
industries, while bioenergy production can generate
more employment and a wider range of jobs per unit of
energy generated than other energy sources. Of course,
the number and type of jobs, the skill level required, and
the local economic impact will be different with different
biomass sources and will vary by region.
The local nature of biomass also means that smaller
sources of bioenergy can have very important impacts.
This is especially true in communities that do not have
access to conventional energy sources. These potential
improvements in rural economies may also help to avoid
the abandonment of land, while encouraging agricultural
competitiveness and providing diversified income streams
for farmers. Although inefficient compared with largescale schemes, small-scale schemes can significantly
benefit local populations.
Unlike biomass, biofuels can be transported with
relatively low cost. There are, however, cases where
densified or pretreated biomass may be exported from
local sources for conversion elsewhere. One example
is the shipment of wood pellets from North America
to EU member states such as the UK, which has been
incentivized by renewable energy policy, as mentioned
in Chapter 3 of this handbook. Whether it is biomass or
biofuels that are traded, the social acceptability of sourcing
and production must be accounted for at both ends of the
supply chain.

Table 4.1.
Some of the potential socio-economic benefits associated with bioenergy [1].
Social aspects
Increased standard of living
Environment.
Health.
Education.
Community wealth and infrastructure.

Social cohesion and stability


Mitigating rural depopulation.
Regional development.
Rural diversification.
Support of traditional agricultural or forestry culture.

Economic aspects
Macro-economics
Security of supply/risk diversification.
Economic growth through business expansion or employment.
Export potential.
Import substitution (effects on GDP).

Demand-side economics
Employment.

Income and wealth generation.


Induced investment.
Support of related industries.

Supply-side economics
Increased productivity.
Enhanced competitiveness.
Improved infrastructure.
Labour and population mobility.

4 Economics, the environment and politics | 49

The demand for resources


With the global population predicted to grow to 9 billion
by 2050 and large numbers of people becoming more
affluent and demanding of consumer goods and services,
the demand for energy, food, water and other resources
is rising fast. Of all the competing demands for resources
that overlap with bioenergy, the demand for food is both
the most serious and the most visible. Current world
food production is sufficient to meet global demand, but
economic factors, distribution failures and waste prevent
that demand from being met. So far, any impacts of
biofuels and bioenergy on food prices or availability appear
to be modest[2]; however, the effect of future expansion of
bioenergy on food production remains a concern and is
shaping bioenergy-related policies in many regions.
The social acceptability of biomass use, includingmodes
and locations of biomass production and conversion
technology choices, is complex and variable. Some groups
oppose recovering energy from the organic material in
municipal solid waste, contending that it will create
a market for waste and interfere with waste reduction
efforts. Others object unequivocally to using forest
biomass (especially from publically owned lands) or food
or feed products for bioenergy. Although proponents of
bioenergy can supply good evidence for well-managed
sustainable use of all of these biomass sources for
bioenergy, these objections are important as they shape
policy and have profound impacts on the availability and
type of biomass and biofuels that will receive investment.

Economic challenges
Unleashing the full potential of bioenergy will depend
largely on developing new conversion technologies
that are cheaper, more efficient and use a wider range
of feedstocks. The commercial development of such
technology at a meaningful scale requires significant

capital investments, with the magnitude and risk


varying according to the particular conversion pathway.
Substantial investment is also required in feedstock
(crop) improvement and cultivation methods. Adopting
new crops is risky for farmers, especially when
commercial biofuel plants, utilizing new crops, are not yet
economically viable without subsidy. Without incentives
to offset some or all of this risk, investors are unlikely to
support the construction of commercial facilities lacking
some level of guaranteed return, and farmers are unlikely
to commit to planting a large acreage of energy crops when
greater returns are available elsewhere in the marketplace.
In any case, there will always be intrinsic competition for
land use from a variety of economic sectors, depending
heavily on location. Ultimately, the rate of economic
return will determine land use.
As well as investment challenges, there are practical
issues to overcome to commercialize large-scale bioenergy
projects. These include building supply chains and
logistical facilities capable of aggregating large quantities
of feedstock while minimizing GHG emissions of the
resulting energy product.
Once the initial hurdles of commercializing an
integrated supply chain are overcome, investors also need
to be aware of the volatility of natural commodity markets.
Managing the risk associated with yield variability from
year to year, caused by seasonal weather variations or
commercial factors, is a key consideration. Climate
variability can also lead to inconsistent feedstock quality
(due to the physiological responses of the plants); this can
affect the processing requirements and economic yield.
While the risks are significant, the rewards can be
significant too: when appropriate technology, feedstock
availability and policy are aligned, substantial investment
is possible. Examples include the Brazilian sugarcane
ethanol industry, the US corn ethanol industry, and the
growing use of waste wood for electric power generation
in the UK.

Environmental sustainability
If bioenergy is to be truly sustainable in economic, social
and environmental terms, it will be important not to
disrupt native ecosystems and world food resources;
careful management will be required with regard to the
production of feedstock and siting of conversion facilities.
Environmental sustainability requires attention to
land, air, water and biodiversity. The state of resources
before bioenergy development and the changes in those
resources caused by bioenergy development are both key
factors for understanding sustainability. Much of the
information addressed in this handbook has been directly
related to land-use issues, but embedded in every landmanagement decision is a potential impact on water, air
and biodiversity.
Water availability (quality and quantity) is of increasing
concern globally, not only in relation to agricultural use,
but also in relation to the rising demands for industrial and
domestic water. Because agriculture is responsible for
70% of worldwide freshwater withdrawal[3], expansion
of agriculture for energy production warrants special
attention. Different bioenergy crops have different water
requirements, so choosing appropriate crops for each
geographic location is critical. While rainfed cultivation
50 | 4 Economics, the environment and politics

does not require water withdrawal from lakes, rivers


and groundwater, it can still affect downstream users
by redirecting precipitation and altering groundwater
recharge and river flows when high yielding biomass crops
utilize more of the available water than natural vegetation.
There are clearly areas around the world where ample
water availability can support very high-yielding energy
crop production, but in other areas water-use efficiency
and downstream consequences will prove to be more
important selection criteria than yields.
Water quality is reduced in many regions by large-scale
agriculture devoted to intensively managed annual crops
(because of their high requirements for fertilizer and
pesticides). Crops with lower input requirements (such as
perennial crops) can lead to water-quality improvements
if they replace these agricultural systems. Some of the
plants considered for bioenergy, such as willows and
switchgrass, have historically been used along waterways
to help protect water resources from pollution, stabilize
the waterway bank, and improve aquatic and wildlife
habitat. In contrast, approaches that displace native
wetlands, forests and grasslands could result in a
degradation of water quality.

Some crops currently used for bioenergy production,


particularly arable crops such as maize, wheat, soybean
and oilseed rape, have significant requirements for
fertilizer and/or chemical crop protection. Not only does
this counteract the reduction of GHG emissions, but
poor nutrient management can also lead to lower soil
quality. Where excess fertilizer is applied, there is a risk
to water quality (as described above) and air quality as a
result of nitrous oxide and other volatile emissions that
can be produced. Crops should be carefully selected and
appropriate management should include nutrient analysis,
restricted timing of nutrient application, and the use of
conservation drainage as required.
Annual bioenergy crops have the potential to
degrade soils through poor cultivation practices, as has
been demonstrated in some cases by the commercial
agricultural management of many commodity crops.
While all crops can be poorly managed, well-managed
perennial and long-rotation crops have significant
potential to increase soil carbon stocks and improve
soil health, bringing previously degraded land back into
economic production or managing the environmental
impacts of intensified food-crop production. Introducing
well-managed perennial systems to degraded soils can,
over time, lead to more resilient agricultural landscapes
capable of supporting both biomass and food crops with
greater yields.
Biodiversity is most at risk if native ecosystems are
displaced by cultivated systems. Biodiversity can also be
affected by land-use change from one type of agriculture
to another, but properly selected and managed bioenergy
crops can enhance biodiversity in some cases. Diversified
landscapes can host a wider range of insects, birds, small
mammals, reptiles and amphibians than monocultures.
On an overall acreage basis, deliberately selecting higher
yielding crops can also limit the acreage under cultivation
and improve habitat availability. It is important to note
that in some areas (such as North America and Western
Europe) the use of certain perennial energy crops (such
as short rotation coppice or miscanthus) could actually
provide greater biodiversity than typical annual crops
such as wheat or maize.

Energy use, GHG emissions and the impact


of land-use changes
Bioenergy projects today are increasingly motivated by
demands for national energy security and decarbonization
of energy. With respect to the latter, the choice of biomass
source and type of processing will have a major impact.
Overall emissions associated with biofuels production
arise during production, distribution and processing of the
feedstock, as well as the burning of the fuel by consumers.
Discounting the carbon dioxide released during
bioenergy use (combustion), the agricultural production
phase generally accounts for most of the total life cycle
GHG emissions. These arise from the manufacturing of
chemical inputs (such as fertilizer and pesticides), the use
of energy in farming operations (including soil cultivation
and harvest), and changes in the levels of carbon and
nitrogen in the crop, soil and the atmosphere caused by
cultivation. A major portion of the overall GHG emissions
can be associated with fertilizer production (and use)
alone. This varies, however, according to the crop: while
maize may require 200kg N per hectare, miscanthus has
been grown and harvested at some locations with no
N input for 15 years with no loss of yield.
Extensive modelling and measurement is under way
to understand factors influencing the GHG performance
of bioenergy. Some legislation (such as the European
Renewable Energy Directive) now requires clear limits
on GHG emissions associated with the production of
bioenergy, often together with additional sustainability
criteria. While the techniques of life cycle analysis are
well established, its use in complex emerging fields such
as bioenergy is less well developed. It is therefore crucial
that the scope and boundaries of the analysis are applied
consistently and transparently to allow meaningful
comparisons between alternative energy pathways.
Data supporting the inclusion of GHG and other effects
associated with land-use change is sparse and an important
cause of diverging results from different studies. Especially
when considering indirect land-use change, data are often
entirely lacking.
From the perspective of the global carbon cycle, the
best bioenergy systems can be considered near-neutral in
the sense that biomass carbon emitted during the biomass
processing and final use was earlier sequestered during
biomass growth. If bioenergy production causes (directly
and/or indirectly) changes in land cover and use, there may
be a net gain or net loss in terrestrial carbon stocks (i.e.
net withdrawal or addition of atmospheric carbon). Some
bioenergy systems (such as biomass extraction from forests
managed with long rotation periods) are not carbon-neutral
over shorter time scales, because carbon sequestration
and emissions are temporally out of balance, though over
longer time scales sequestration and emissions come into
balance.

4 Economics, the environment and politics | 51

While biofuels generally have lower GHG emissions than


fossil fuels, careless choices regarding the production
pathway, feedstocks used and additional fossil energy
process inputs can actually result in increased emissions.
Table 4.2 outlines the ways in which factors in
bioenergy production can influence emissions. Table 4.3
provides examples of how a single management decision
can completely reverse the life cycle GHG emissions from
a bioenergy production system.

As can be seen in Table 4.3, one of the most critical factors


is land use and the increase or reduction of GHG emissions
associated with land-use change. Converting carbon-rich
wetlands into biofuel plantations, for example, can result
in large GHG emissions, while planting perennial grasses
on soils previously used for annual arable crops is likely to
reduce GHG emissions.

Table 4.2
Key factors determining bioenergy GHG emissions, with brief examples and explanations.
In most cases there are many more examples for each factor.
Land-use change

Direct land-use changes and associated GHG


emissions (e.g. oxidation and volatilization
of nitrogen and carbon) can entirely negate
the GHG benefits of biofuel use. In other
situations, for example a change from
annual cropping to perennial crops for
bioenergy, the GHG benefits may increase.

Feedstock type

Some feedstocks (such as maize and


soybean) have high nutrient- and
crop-protection requirements. This may
result in higher direct and indirect GHG
emissions compared with low-input
perennial crops (such as miscanthus or
short-rotation coppice and co-products).

Agricultural
practices

Energy is a key agricultural input, so any


reduction in cultivation and other energyintensive operations will reduce the GHG.
Certain practices (such as minimum tillage
cultivation) can also reduce GHG emissions
that arise from soil disturbance.
Increasing yield also has a significant
effect on reducing overall GHG emissions.
GHG emissions associated with the
production of nitrogen fertilizers generally
dominate the GHG emissions from
agricultural practices.

Manner of
conversion

Like feedstock yield, the overall efficiency


of the conversion process has a key role in
reducing the overall GHG emissions.

Supply-chain
logistics

Transporting biomass over long distances


can place significant burdens on the
economics of a bioenergy enterprise, and
can also contribute to GHG emissions.
Bioenergy producers will pay close attention
to supply-chain logistics. Bulk ocean
transport results in significantly fewer GHG
emissions per unit of biomass or fuel than
road transport, for example.

Figure 4.1, on the next page, deconstructs the source of


emissions that occur as a result of different land uses in
a tropical forest or tropical grassland biome. The initial
storage of carbon depends on the condition of the land.
There is an ongoing exchange of GHG (including CO2,
CH4 and N2O) between the terrestrial ecosystem and
the atmosphere. The GHG flux, combined with physical
characteristics of an ecosystem, determines the net
radiative forcing, or potential to warm the atmosphere,

* For further reading, please see reference on page 56:


Indirect land-use change [4].

52 | 4 Economics, the environment and politics

Type of energy
replaced

Where bioenergy displaces fossil fuels (such


as coal), the potential to deliver higher GHG
savings is greater than where it displaces a
lower GHG fuel (such as natural gas). In the
process chain, GHG emissions are strongly
determined by the fuel that is used for
process energy generation.

End use of fuel

The use of a unit of solid biomass for


combined heat and power generation will
have a greater efficiency (in strict energetic
terms) than pure electrical generation, pure
heating use or conversion to liquid fuels.
The uses of bioenergy, however, are also
dependent on energy demand, competition
with other sources and the best energy
carriers to fulfil this. In contrast with other
renewable energy sources (such as solar
or wind), biomass is well-suited for the
production of liquid transport fuels.

Indirect land-use
change (ILUC)*

The concept of ILUC posits that using land


for one purpose (such as the production
of bioenergy crops) renders that land
unavailable for other uses (such as food
production). Food production, for instance,
will therefore be displaced to other areas,
possibly causing the conversion of natural
environments with high carbon stocks and
biodiversity.
Because it is an indirect effect, there
is much debate about how metrics can
be designed to account for ILUC. To date,
studies have assumed widely varying levels
of emissions due to ILUC for the same fuel
pathways. In general, since the first attempts
to calculate ILUC impacts, estimates of
emissions arising from ILUC have been
substantially reduced. We now have a better
understanding and can carry out more
complex modelling, but there continues to
be a significant lack of consensus about how
to estimate ILUC and establish causality.

caused by that ecosystem. Negative radiative forcing, as


indicated on the horizontal axis, are net emissions to the
atmosphere and positive radiative forcing are net uptake
by the ecosystem. The net GHG benefit of a tropical forest
or grassland is the result of carbon storage in the plants and
soil whereas the net GHG benefit of growing sugarcane
for fuels largely results from displacing the emissions
associated with fossil fuel use, shown with max values
in Figure 4.1.

Table 4.3
Effects of a single management change on life cycle GHG emissions
per unit of energy in biomass-based production systems[5].
Different management decisions are shown for different species.
The assumptions used for each life cycle assessment vary across
the species, e.g. palm and maize life cycles include co-products and
end-use emissions whereas miscanthus and sugarcane life cycles
include fossil-fuel displacement in their assessments. Therefore,
comparisons should be made only within a species. Positive values
indicate a net source (emission) of GHGs into the atmosphere and
negative numbers indicate a net sink or reduction of atmospheric
GHG. Note the large swings associated with land-use change in the
palm and miscanthus life cycles. GHG emissions for comparative
purposes: gasoline, 94g CO2eq/MJ and diesel, 85g CO2eq/MJ[6].

Crop

Management decision

Palm
(Indonesia)

Prior land use: rainforest

Outcome
GHG emissions
g CO2eq/MJ
energy produced
500

Prior land use: palm

28

Prior land use: palm and best


management practices adopted

Much theoretical work has been carried out to assess


the relative GHG emissions of different crop and energy
pathways, as well as the impact of different agricultural
management regimes (such as different tillage and seedbed
preparation techniques, the use of cover crops, harvesting
techniques etc.), but actual measurements of the GHG
balance of different feedstock production systems
especially in commercially realistic settings are rare.
Post-harvest processing and transport also influence
the total GHG emissions associated with any particular
energy type.
The interactions of these various factors with local
conditions mean that biofuels have differing energy
efficiencies and climate, social and environmental
impacts. The balance of beneficial GHG reductions
with other considerations such as national security, and
economic, environmental and social impacts that in
combination are acceptable to societies around the world,
will be a key determinant of the extent to which bioenergy
is adopted.

-190

Miscanthus
(UK)

Prior land use: forest

120

Prior land use: crop

-45

Sugarcane
(Brazil)

Traditional pre-harvest burning

Maize
(US)

Traditional tillage

25

No-till practised

-4

32

No pre-harvest burning

-13

Figure 4.1
Graphical representation of the sources of GHG emissions occurring as a result of different land uses in a tropical forest and
tropical grassland biome. The greenhouse gas value (GHGV) in the right panel is the sum of the GHG mitigation potential
associated with the system prior to change (initial storage) and the change in GHG mitigation potential (ongoing exchange) of
the terrestrial system over a 30-year period. Blue represents CO2, green represents CH4, red represents N2O, and black is the
sum of all three gases in CO2 equivalents. Here, positive numbers indicate a net sink or retention of GHG, and negative numbers
indicate emissions of GHG or increased radiative forcing that contribute to climate warming. The values shown as max in the
figure include an estimate of fossil-fuel displacement in the final GHGV value. The ones shown as eco are just the soil/ecosystem
biogeochemistry.The values for sugarcane are for crops in a variety of landscapes with varying amounts of forest cover [7,8].
Initial storage

Ongoing exchange

Total GHGV

Tropical
peat forest

Tropical
forest

Tropical
cropland

Sugarcane
(not burnt eco)

Sugarcane
(not burnt max)

500

1000

1500

2000

2500

200 400

500 1000 1500 2000 2500

GHGV (Mg CO2-eq ha-1 30yr-1)

4 Economics, the environment and politics | 53

The politics of biomass


Policies designed to boost the development of renewable and
low-carbon forms of energy need to be carefully designed.
They must provide adequate support and certainty for the
industry (actively encouraging the positive benefits) at the
same time as addressing a number of potential challenges
(preventing the development of unsustainable forms of
energy). The intersection of biomass production with energy
and agricultural sectors presents unique challenges for
the early stages of this industry. This chapter provides an
overview of some key issues.

Energy security
Bioenergy can potentially increase the energy security
of a nation by using indigenous biomass grown within
its borders. Such indigenous sources of bioenergy could
provide other benefits to the national economy and reduce
the risk of supply disruption.
The Global Bioenergy Partnership (GBEP) concluded in a
2008 study that, Most of the bioenergy consumed in G8 +5
Countries is produced locally. Even as biomass becomes
traded globally, there will be many more countries able to
supply biomass energy than those with fossil fuel reserves.
By not using energy resources (fossil fuels) that would
otherwise be consumed, the development of a new energy
resource (bioenergy) could improve energy resilience and
security. Investment in local energy co-provision (such
as electricity co-generated from a mix of fuels) can boost
rural electrification programmes with all the concomitant
development and health benefits.

Regulation for sustainability


As described above, not all forms of bioenergy perform
equally in terms of their environmental footprint or social
acceptability. Government policy and regulations are
needed, but issues relevant to the efficacy of regulations are
complex, so policies will need to be tailored both locally
and nationally to protect human rights and promote other
economic or social objectives.
Bioenergy production is an extension of several
ongoing activities in other arenas including commercial
agriculture, commercial forestry and waste management.
All the activities have environmental regulations
that vary regionally across the globe. Industry groups
and third parties have developed a range of voluntary
sustainability standards, some of which are relevant to
bioenergy production. Examples include the Roundtable
on Sustainable Soy, the Roundtable on Sustainable Palm,
the Sustainable Forestry Initiative, the Forest Stewardship
Council, the Sustainable Sugarcane Initiative, and
Solidaridad. In addition, many countries and states are
moving environmental regulation of agriculture, forestry
and waste toward more inclusive sustainability standards
to which biomass for energy will also be subject. Several
key policies, however, have emerged in the past five years
that include sustainability criteria specific to bioenergy
production.

54 | 4 Economics, the environment and politics

In Brazil, the expansion of sugarcane for ethanol has


been subject to new policies that include agroecological
zoning[9], a process that designates land appropriate
for development without interfering with long-term
sustainability goals. Amendments to the forest code now
require 10% of land in sugarcane production to be set
aside for native forests. In addition, special protections for
riverbank areas are being codified.
In the US, biofuels must meet criteria regarding
GHG emissions and land-use criteria to be considered
as renewable or advanced and eligible to meet the
Renewable Fuel Standard[10] mandates for 36 billion
gallons of renewable fuel by 2022. To limit unsustainable
land clearing for biofuels, for example, purpose-grown
energy crops are only eligible for specific incentives when
restricted to existing agricultural land that was cleared or
cultivated prior to enactment of the Energy Independence
and Security Act of 2007. Also, for wood residues to qualify
they must be from planted trees derived from actively
managed tree plantations on non-federal land cleared at
any time prior to 19 December 2007.
The EUs Renewable Energy Directive, which sets a goal
for 10% renewable energy by 2020, also contains specific
factors relating to the sustainability of biofuels. Criteria
prohibiting use of lands with high carbon and/or high
biodiversity and limiting GHG emissions were codified
in 2010[11].
Indicators of sustainability for bioenergy have been
outlined by the Global Bioenergy Partnership, a part of
the Food and Agriculture Organization of the United
Nations. Criteria for sustainable biofuels proposed by the
Roundtable on Sustainable Biofuels, an independent multidisciplinary organization, have also recently been approved
as certification standards.

Commercialization of new technologies


It is generally acknowledged that the most efficient
policies are technology-neutral and do not attempt to
pick winners. The rapidly developing science around
bioenergy is sometimes at odds with slower-moving
political and economic landscapes. Recent bioenergy
policies, however, appear to be ahead of the science and
technology, particularly in the arenas of consequential life
cycle analysis and next-generation biofuel conversions.
In order to provide investor certainty, stable medium- and
long-term policies to support markets for bioenergy are
required.
Typically, government support for bioenergy takes the
form of incentives or mandates, often linked to other
energy or agricultural policies. Incentives can include
feed-in tariffs, tax reductions or rebates, grants, loan
guarantees, construction incentives or other supportive
policies or fiscal instruments. Mandates may include
obligations to use a percentage of bioenergy (or other
renewables) in fuel or energy generation. These vary
considerably from country to country, as Table 4.4
demonstrates, and are subject to politically imposed
change.

Table 4.4
Voluntary and mandatory bioenergy targets for electricity, heat and transport fuels (as stated
in country summaries and key policy documents) in place in different countries in 2008. These
policies change with time in terms of both the types of bioenergy being legislated and the absolute
values of the various targets[12].
Country

Targets
Electricity

Heat

Transport fuels

Brazil

Inclusion of 3.3GW of renewable


energy (wind, biomass, small
hydroelectric) into National Energy
Grid by 2007. Once met, renewable
sources to provide 10% of total energy
consumption within 20 years[13].

No targets.

blend of 2025%
anhydrous ethanol with gasoline;
minimum blending of 2% (B2)
biodiesel to diesel by 2008 and 5%
(B5) by 2013.

China

China is finalizing a revised target


for 16% of primary energy from
renewables by 2020, including large
hydro; plans include a target for
30GW of biomass power for 2020.

No targets.

15% of its transportation energy


needs through use of biofuels in 2020.

India

No targets.

No targets.

A 5% blending mandate for ethanol


will be established before end of
2007, and Planning Commission
has proposed to raise mandate to
10%. On biodiesel, the Committee
for the Development of Biofuels has
decided 20% of diesel consumption as
blending target for 201112.

Mexico

>1GW of Renewable Energy Sources


(RES) by 2006. Mandatory

No targets currently (targets under


consideration).

No targets currently (targets under


consideration).

South Africa

4% by 2013.

No targets.

Up to 8% by 2006 (10% target under


consideration).

Canada

No targets.

No targets.

5% renewable content in gasoline by


2010 and 2% renewable content in
diesel fuel by 2012.

France

21% RES by 2010. Mandatory

50% increase from 2004 until 2010 in


heat from RES. Mandatory

5.75% by 2008, 7% by 2010,


10% by 2015, 10% by 2020.
Mandatory EU target

12.5% by 2010, 20% by 2020.

No targets.

6.75% by 2010, which is set to rise to


8% by 2015; 10% by 2020.
Mandatory EU target

Mandatory

Mandatory

Germany

Mandatory

Italy

25% by 2010.

No targets currently (targets under


consideration).

5.75% by 2010.
10% by 2020.
Mandatory EU target
Mandatory

Japan

Biomass power generation and waste


power generation in the amount of
5.86 billion litres, as converted to
crude oil, by 2010.

Biomass thermal utilization in the


0.5 billion litres, as converted to crude
amount of 3.08 billion litres (this
oil, by 2010.
amount includes biomass-derived fuel,
0.5 billion litres, for transportation), as
converted to crude oil, by 2010.

Russia

No targets in place, but considering a


7% RES target by 2020.

No targets.

No targets.

10% by 2010, 15.4% by 2016.

>1GWe of installed combined heat


and power (CHP) capacity by 2010
with >15% of government buildings
using CHP.

5% biofuels by 2010. Mandatory


10% by 2020. Mandatory EU target

UK

Mandatory

US

No national targets but individual


states are pursuing a variety of
incentive programmes.

No targets.

A Mandatory Renewable Fuel


Standard (RFS) of 34 billion litres of
renewable fuel in 2008, progressively
increasing to 136 billion litres in 2022
with 79 billion litres coming from
advanced biofuels such as cellulosic
ethanol. Non-renewable fuels are not
considered in RFS.

EU

20% RES as overall target for all


Member States by 2020. Mandatory

No targets.

10% by 2020. Mandatory EU target

4 Economics, the environment and politics | 55

The science, politics and economics of bioenergy are


developing rapidly, as is our understanding of the
requirements for producing bioenergy and the effects of
expanding commercial implementation at local, national
and global levels. It is now clear that there is no one size
fits all solution for bioenergy development. Different
options will be appropriate at different scales and different
locations, depending on available infrastructure as well as
cultural and developmental contexts. Given this situation,
precision and details matter when teasing out the benefits
and potential advantages of biomass from the potential
risks.
Overall, agriculture fixes a prodigious amount of carbon
for use as food, feed, materials and energy. Given finite
land resources, the most important lever for enhanced
use of agriculturally derived materials is improved
agricultural productivity regardless of the ultimate use
of the agricultural products. Finding mechanisms that

allow for improved productivity, while limiting negative


impacts of highly productive agriculture, are important
environmental, regulatory and policy considerations.
Increased production and use of biomass has the
potential to reduce GHG emissions from energy
production, reverse degradation caused by intensive
agricultural systems, and create sustainable industries
in rural areas. In order to contribute materially to
global energy, food and climate security, bioenergy
provisionmust expand only with careful and contextsensitive implementation. Otherwise, bioenergy
developments may lead to unintendedconflict and
competition.
There is strong demand for bioenergy, and new
technology is being developed to use biomass resources.
Consistent policies that promote environmental,
economic and social sustainability are essential to fulfil
the potentialof bioenergy development.

Chapter references
[1] Domac, J., Richards, K., & Risovic, S. (2005),
Socioeconomic drivers in implementing bioenergy,
Biomass and Bioenergy, vol. 28, pp. 97106.

[8] Anderson-Teixeira, K. J. & DeLucia, E. H. (2011), The


greenhouse gas value of ecosystems, Global Change
Biology, vol. 17, pp. 425438.

[2] Baffes, J. & Dennis, A. (2013), Long-term drivers of


food prices. The World Bank Development Prospects
Group and Poverty Reduction and Economic
Management Network Trade Department report.

[9] Manzatto, C. V., Assad, E. D., Bacca, J. F. M., Zaroni,


M. J. & Pereira, S. E. M. (2009), Zoneamento
agroecolgico da cana-de-acar: expandir a
produo, preservar a vida, garantir o futuro.
Documentos 110. Embrapa Solos, Rio de Janeiro.

[3] Williams, E.D. & Simmons, J.E. (2013), Water


in the energy industry: an introduction. BP,
London, UK. Available from: www.bp.com/
energysustainabilitychallenge [accessed February
2014].
[4] Khanna M. & Crago C. L. (2012), Measuring indirect
land use change with biofuels: implications for policy,
Annual Review of Resource Economics, vol. 4, pp.
161184.
[5] Davis, S.C., Boddey, R.M., Alves, B.J.R., Cowie,
A.L., George, B.H., Ogle, S.M., Smith, P., van
Noordwijk, M. & van Wijk, M.T. (2013), Management
swing potential for bioenergy crops, GCB Bioenergy,
vol. 5, pp. 623638.
[6] US Energy Information Administration (2013),
Levelized cost of new generation resources in the
annual energy outlook 2013. Available from: http://
www.eia.gov/forecasts/aeo/er/pdf/electricity_
generation.pdf [accessed February 2014].
[7] Buckeridge, M.S., De Souza, A.P., Arundale, R.A.,
Anderson-Teixeira, K.J. & DeLucia, E.H. (2012),
Ethanol from sugarcane in Brazil: a midway strategy
for increasing ethanol production while maximizing
environmental benefits, GCB Bioenergy, vol. 4,
pp. 119126.

56 | 4 Economics, the environment and politics

[10] US Environmental Protection Agency (2012),


Regulation of fuels and fuel additives: 2012 renewable
fuel standards; Final rule (January 9), Federal Register,
vol. 77, no. 5, pp.13201358.
[11] Directive 2009/28/EC of the European Parliament and
of the Council, articles 17, 18 and 19. 2009. Official
Journal of the European Union, L 140/16.
[12] Global Bioenergy Partnership (GBEP) (2007), A review
of the current state of bioenergy development in G8
+ 5 countries, GBEP, Rome. Available at: http://www.
globalbioenergy.org/fileadmin/user_upload/gbep/
docs/BIOENERGY_INFO/0805_GBEP_Report.pdf
[accessed February 2014].
[13] World Resource Institute-SD-PAMs Database,
Programme of Incentives for Alternative Electricity
Sources (PROINFA), Available from http://projects.
wri.org/sd-pams-database/brazil/programmeincentives-alternative-electricity-sources-proinfa
[accessed August 2014].

58

Climate and soils determine which


plants will grow in a particular place, so
energy crops should be chosen for the
growing region (biome) to which they
are best suited.

One way of integrating biomass


production into agriculture is to
find areas that are marginal for food
production but that will support the
different needs of the most suitable
energy crops.

Figures for rainfall, temperature and


growing season the main parameters
for choosing crops are given for
the global biomes that will support
significant plant life.

5 Where can biomass feedstocks be grown?


Growing regions (biomes)
This chapter provides detail on the fundamental
biophysical constraints imposed on agriculture. In order
to understand the potential growing regions for biomass
crops, we must consider the biophysical and climatic
parameters such as temperature, precipitation, soils
and land area that all vary across the globe. This chapter
introduces regional categories (biomes) that are defined
based on these ecological parameters. It provides the
background and context for bioenergy crops in the global
landscape. A summary table is provided at the end of
the chapter as a quick guide to biomes that might host
bioenergy crops.
Biomes are distinguished by climate, soil and vegetation
in any classification scheme; they do not respect political
borders. Here we present the biome categories recognized
by the Food and Agriculture Organization of the United
Nations (FAO) according to Udvardys Ecoregions[1],
although they are named differently by various groups.
Please note that we have included only the broad
classification categories in the following pages, and not
100% of the land surface is depicted in the various biomes
for this reason.
A biome covers a wide geographical area, so conditions
within a biome can vary broadly. The rainfall and
temperatures described in this chapter reflect annual
average ranges. Within each biome, there will be variation
and anomalies beyond the average figures. Temperature
and rainfall patterns are affected by elevation, aspect, the
prevailing wind direction and proximity to the sea or other
large bodies of water.
Growing-season lengths, defined as the number of days
from plant emergence or bud break until senescence (when
leaves turn brown and dry down), are often determined by
temperature and also vary widely across biomes. In warm
climates, where plant growth may persist year-round, the
growing season is sometimes defined by the occurrence
of sufficient rainfall. It should be noted that rainfall
distribution patterns are important some climates have a
single rainy season, others have a bimodal pattern while
tropical coastal climates are wet enough to allow growth
throughout the year.

Soils are highly spatially variable, but the major soil


classifications[2] in each biome are identified because these
have an impact on vegetation and ecosystem productivity.
Climate and soil, together, are key constraints on the
growth of ecosystems (including agricultural ecosystems).
Just as the flora and fauna of a biome are adapted to their
native conditions, crop species can be more suited to
particular locations, but growing-season length, soil
quality and water availability place certain limits on
productivity.
The potential for bioenergy production is also
constrained by the conditions of a biome. Biomass
crops are diverse, so optimal biomass production will be
achieved by selecting regionally appropriate species and
identifying land opportunities that will not displace native
ecosystems or disrupt other agricultural commodities.
This chapter provides context for the global landscape in
which biomass production must fit.
Figure 5.1, on the next page, maps the general suitability
of land globally for agriculture. The map shows suitability
of currently available land for pasture and rain-fed crops
based on the general requirements of conventional
food-cropping systems as defined by the FAO[3]. Assumed
environmental conditions and constraints include thermal
climate, length of growing period, climate variability, soil
types and qualities, and terrain slope classes. Based on
these constraints, the suitability classes are mapped at
a 5 arc-minute spatial resolution. Lands classified as
well-suited for agriculture are dominated by prime
agricultural lands, and lands classified as suitable can be
considered marginal, both defined in Chapter 2. Lands
classified as unsuitable or poorly suited for agriculture are
less likely to support agricultural systems because of
climate and/or soil conditions that are not tolerated by
conventional food crops (e.g. low precipitation, extreme
temperatures). Lands that are urban (>25% of resolved grid
space in urban development), protected (according to the
World Conservation Monitoring Centre), closed forest
(>75% of resolved grid space in forest), or irrigated (>50%
of resolved grid space irrigated) are classified separately, as
these are lands that would be considered unavailable for
agriculture[3]. Irrigated lands, which are used for growing
food in some places of the world should generally be
considered unsuitable because of the management
intensity that would be required to maintain and expand
bioenergy crops in these places.
Biomes described in the following pages are overlaid
with areas of agricultural suitability. The ecological
constraints on plant growth are also defined for each
region to provide context for the specific crops detailed
in Chapter 6.

5Where can biomass feedstocks be grown? | 59

60

Inland water bodies

n
n Irrigated area
n Land prime or well suited for agriculture

n Closed forest

Agricultural suitability of land across the globe

n Land suited for agriculture


n Land unsuited to poorly suited for agriculture
n Protected area
n Urban area

Suitability of land with appropriate levels of inputs for pasture and rainfed crops[3].

Figure 5.1

Subtropical humid forest


These are potentially some of the most productive areas
on earth because of the long growing season, moderate
temperatures and abundant rainfall. South-east Brazil,
southern China and New Zealand enjoy moderate
temperatures, plentiful but not excessive rainfall and a
growing season that lasts at least three-quarters of the
year. Depending on the underlying geology, soils range
from fertile volcanic and clay-based soils to weathered,
infertile and acidic clay-rich soils. Globally abundant
soils that have no distinct layers and soils that are only
moderately developed also occur. The weathered soils
need careful management to maintain their structure
under cultivation. The US Department of Agriculture
offers a useful guide to soil classification[2].

Current land use


As its name suggests, the natural vegetation of this biome
is forest, both broadleaved and coniferous. But much of the
forest has been replaced by smallholder cultivation and
intensive mixed agriculture featuring both livestock and
cereals. With correct fertilizer and liming where necessary
to counteract soil acidity, a wide range of crops can be
grown (including many forms of bioenergy crops).

Useful numbers
Rainfall: 1,000 2,500mm per year
Temperature: 10 30C
Growing season: 270 365 days

Agricultural suitability of land in subtropical humid forest

n Closed forest

n Inland water bodies


n Irrigated area
n Land prime or well suited for agriculture
n Land suited for agriculture
n Land unsuited to poorly suited for agriculture
n Protected area
n Urban area

D
C

E
A

5Where can biomass feedstocks be grown? | 61

Heading broadleaved forest


Temperate
After tropical forests, the temperate hardwood forests
are the most biodiverse forests of the world. This biome
includes the eastern US, much of Western Europe and
the eastern half of China. The natural vegetation of
this biome has largely disappeared on some continents
because of the natural productivity of these regions for
agriculture, but on others this ecosystem has regenerated
over the past century (such as the temperate forests of
the eastern US). Winters can be moderately severe but
summer temperatures are suitable for a wide range of
crops. A range of soils is found, including those rich in
organic material and having formed from limestone,
wind-blown sediment or sand. Soils resulting from the
weathering of clays, and partially developed soils, are also
common in the eastern US.

Current land use


Large-scale commercial agriculture is well established.
All major crop groups are grown, particularly cereals and
vegetables. Livestock farming forms part of the landuse mosaic and a variety of forestry practices are used to
manage the land that is still in forest, much of it secondary
and having regenerated after previous logging events.

Useful numbers
Rainfall: 350 1,500mm per year
Temperature: -10 30C
Growing season: 90 365 days

Agricultural suitability of land in temperate broadleaved forest

n Closed forest

n Inland water bodies


n Irrigated area
n Land prime or well suited for agriculture
n Land suited for agriculture
n Land unsuited to poorly suited for agriculture
n Protected area
n Urban area

A
B

62 | 5Where can biomass feedstocks be grown?

Temperate coniferous forest


Great swathes of the northern hemisphere, lying roughly
between latitudes 50 and 70 North, are naturally
covered by coniferous forest. This biome includes most of
Canada and Russia, and the northern parts of Scandinavia.
Conditions can be harsh in winter and the growing
season is short, but coniferous trees are adapted to these
conditions. In the coldest parts, soils overlay permanently
frozen layers of ground (permafrost) so rooting depth is
restricted. Typically soils under coniferous trees have been
formed by weathering, which has stripped organic matter
and aluminium from the surface and deposited them in the
subsoil. The hard layer that results from such weathering
and their free-draining nature render them poor for
agriculture. Peaty soils have developed in poorly drained
areas, where the lack of oxygen slows the decomposition
of organic matter. These peat deposits are important global
carbon sinks.

Agricultural suitability of land in


temperate coniferous forest

n Closed forest

n Inland water bodies


n Irrigated area
n Land prime or well suited for agriculture
n Land suited for agriculture
n Land unsuited to poorly suited for agriculture
n Protected area
n Urban area

This is a very large biome (17% of global land area), a great


deal of which is not easily cultivated for crops. With an
average maximum temperature at the minimum required
for growth of most crops, productivity is severely limited
by the cold.

Current land use


Forestry still dominates the land use in the northern parts,
which are inhospitable and sparsely populated. Some crops
can be grown, including hardy cereals such as oats and
barley, some tough vegetables and root crops such as beets
and potatoes. Further south, on the better soils, large-scale
growers produce cereals.

Useful numbers
Rainfall: 100 1,500mm per year
Temperature: -30 5C
Growing season: 30 180 days
The northwest coast of the US and Canada is a distinct
area of coniferous forest that receives up to 5,000mm
of annual rainfall, has an average annual temperature
of 10C, and a growing season that lasts up to
300 days in some places; this area is often classified
as temperate rainforest.

5Where can biomass feedstocks be grown? | 63

Heading grassland
Temperate
Natural grassland is known variously as prairie (North
America), steppe (Eurasia) and pampas (South America).
Trees are scarce, restricted by the low rainfall or soil
moisture and risk of fire. Historically, these areas were
home to grazing animals, such as the North American
bison, that moved freely in search of fresh grass following
rain or fire. Some grassland soils are rich in organic
matter, optimal for cultivation, and some are derived from
weathered clay minerals but, in drier areas, they may be
very low in organic matter with cemented subsoils or
salty surface accumulations. Soils, based on shrinkable
montmorillonite clays, have a heavy texture but, when
irrigated, are suitable for arable crops.
In the prairie and steppe regions, winters are extremely
cold, with freezing temperatures. The high altitudes of
the great basin region of the US and China are arid as
well as cold, and have little to no vegetative growth in
some places.

Current land use


Large-scale commercial agriculture is successful in much
of these areas, but erosion can be problematic. Irrigation
extends the range of crops that can be grown, and no-tillage
methods protect soil organic matter and reduce the risk of
erosion. The vast majority of native temperate grassland
has been displaced by agriculture that includes both crops
and livestock.

Useful numbers
Rainfall: 50 1,000mm per year
Temperature: -10 30C
Growing season: 0 320 days

Agricultural suitability of land in temperate grassland

n Closed forest

n Inland water bodies


n Irrigated area
n Land prime or well suited for agriculture
n Land suited for agriculture
n Land unsuited to poorly suited for agriculture
n Protected area
n Urban area

E
A

64 | 5Where can biomass feedstocks be grown?

Tropical dry forest


Much of sub-Saharan Africa, apart from the tropical humid
equatorial forest, is naturally covered with this type
of vegetation. It is also found across much of India and
highland South-East Asia, as well as in north-east Brazil.
More sparse woodland than high forest, it differs from
tropical grassland only in having slightly higher rainfall
and having trees as the dominant vegetation. Soils vary
widely in this biome from productive, moisture-retaining
soils to highly weathered, nutrient-depleted soils, and
others with heavy clay-rich sub-soils. Some clay soils have
the capacity to shrink and expand, affecting water holding
capacity and stability at the surface.

Current land use


Much of this area is cultivated by smallholders, but it also
supports large commercial enterprises growing a wide
range of crops. Total amounts and temporal distribution
of rainfall limit what can be grown unless land can be
irrigated. Highland Kenya, for instance, can expect two
distinct rainy seasons per year, whereas further south, such
as in Zambia and Malawi, rainfall is restricted to a single
season. The weathered soils are often deficient in major
and minor plant nutrients. Rice is widely grown in this
biome, as found in India and Asia. In semi-arid parts of this
biome, succulent crops such as sisal (Agave spp.) for fibre
and jatropha for biodiesel can grow well.

Useful numbers
Rainfall: 700 2,500mm per year
Temperature: 15 30C
Growing season: 30 300 days

Agricultural suitability of land in tropical dry forest

n Closed forest

n Inland water bodies


n Irrigated area
n Land prime or well suited for agriculture
n Land suited for agriculture
n Land unsuited to poorly suited for agriculture
n Protected area
n Urban area

A
B

C
E

D
B

E
C

5Where can biomass feedstocks be grown? | 65

Headinggrassland
Tropical
Found in Colombia and Venezuela, central Brazil and
northern Australia, tropical grassland is similar to tropical
dry forest. It is the slightly lower rainfall that restricts the
growth of trees in this biome. Soils range from dark clay
soils to weathered, infertile and acidic soils.

Current land use


Land in this biome is often farmed by smallholders and
large commercial enterprises. A wide range of crops can be
grown, similar to those described in the tropical dry forest
section, and these crops are restricted primarily by local
rainfall patterns or the possibility of irrigation.

Useful numbers
Rainfall: 500 2,500mm per year
Temperature: 15 30C
Growing season: 30 300 days

Agricultural suitability of land in tropical grassland

n Closed forest

n Inland water bodies


n Irrigated area
n Land prime or well suited for agriculture
n Land suited for agriculture
n Land unsuited to poorly suited for agriculture
n Protected area
n Urban area

C
B

66
66||5Where
5Wherecan
canbiomass
biomassfeedstocks
feedstocksbe
begrown?
grown?

Tropical humid rainforest


Current land use

The most biologically diverse terrestrial ecosystem,


harbouring vast carbon stocks, tropical rainforests are
almost constantly in the news. More than 1,000 plant
species can be found in a single square kilometre of many
tropical rainforests, a quarter of which can be unique to
a given area. Illegal logging and unregulated clearance for
agriculture causes justified alarm in environmental circles.
These forests occur fairly close to the equator, where day
length varies only slightly, and temperature and humidity
are constantly high. They include the Amazon and Congo
basins, Indonesia, Bangladesh, Burma and coastal strips
of western India, Thailand, Vietnam and north-eastern
Australia. In many tropical rainforests, the nutrients are
recycled so quickly that the soils are considered very
poor despite the rich nutrients and biodiversity of the
vegetation. Peat soils are, however, found in some tropical
rainforests of the eastern hemisphere, with very rich
organic material high in both carbon and nutrients. Both
soil types are very sensitive to land-use change.

Large areas of this biome are protected. The dense natural


vegetation led to an early belief that rainforest soils would
be very productive. Although the rainforest system is
very efficient at recycling nutrients, cleared land quickly
loses fertility as the high rainfall leaches minerals from
the topsoil. The land needs careful management to
conserve nutrients and minimize erosion. Smallholders
cultivate plots, growing a range of crops, and commercial
plantations, like oil palm, are also grown in this biome.

Useful numbers
Rainfall: 1,500 5,000mm per year
Island systems depicted on the map may have a mix of
humid rainforest and dry tropical forest with rainfall
as low as 700mm in some areas.
Temperature: 25 30C
Growing season: 300 365 days

Agricultural suitability of land in tropical humid forest

n Closed forest

n Inland water bodies


n Irrigated area
n Land prime or well suited for agriculture
n Land suited for agriculture
n Land unsuited to poorly suited for agriculture
n Protected area
n Urban area

B
E
C

5Where can biomass feedstocks be grown? | 67

Heading
Desert
are suited to near-desert conditions. Semi-arid lands
at the margin of true deserts are as large as the deserts
themselves, totalling roughly 18% of the globe, and are
considered more arable.

True desert covers a significant land area equal to


roughly 15% of the globe. Parts of the south-western
US and Mexico, much of Northern Africa (and parts
of south-western Africa), the Persian Gulf and most
of Australia suffer from infrequent and unpredictable
rainfall, and extremes of heat. Soils are poor, sometimes
merely shifting sand. Some deserts actually have cold
climates, and growth is restricted in these regions
primarily by water availability.

Useful numbers

Current land use


Desert areas are unproductive unless they can be
irrigated, although some plants are adapted to these
harsh conditions. Where irrigation is possible, the warm
temperatures encourage productive crops think of the
strip of land bordering the Nile. Succulent crops such
as sisal (Agave spp.) for fibre and jatropha for biodiesel

Rainfall:
Warm desert 0 350mm per year
Cold desert 0 1,000mm per year
Temperature:
Warm desert 10 30C
Cold desert -14 18C
Growing season:
Warm desert 0 30 days
Cold desert 0 210 days
Although some species may have evolved locally
in order to survive the extremes of climate, most
bioenergy crops will not tolerate these conditions,
leading to zero growing days.

Agricultural suitability of land in deserts

n Closed forest

n Inland water bodies


n Irrigated area
n Land prime or well suited for agriculture
n Land suited for agriculture
n Land unsuited to poorly suited for agriculture
n Protected area
n Urban area

68
68 || 5Where
5Where can
can biomass
biomass feedstocks
feedstocks be
be grown?
grown?

Regional characteristics: comparison table


Rainfall
mm/yr

Temp*
oC

Growing Crops
days**

Possible biofuel crops***

1,000 2,500

10 30

270 365 Cereals, fibres, oil crops,


pulses, roots/tubers,
coffee, tea, sugar crops,
fruit, vegetables, timber.

Maize, sugarcane, soybean,


sorghum, pine, eucalyptus.

Temperate broadleaved forest


Large commercial and smallholder: tree
crops, forest-based livestock, large-scale
cereal and vegetables, cereal/livestock.

350 1,500

-10 30

90 365 Cereals, fibres, oil crops,


pulses, roots/tubers,
coffee, tea, sugar crops,
fruit, vegetables, timber.

Maize, switchgrass,
miscanthus, soybean,
wheat, cordgrass, oilseed
rape, pine, willow/poplar,
sorghum.

Temperate coniferous forest


Forestry, large commercial and smallholder:
cereals/roots, forest-based livestock.

100 1,500

-30 5

30 180 Timber, cereals, roots,


tubers.

Willow/poplar, switchgrass,
miscanthus, cordgrass.

Temperate grassland
Large commercial and smallholder: irrigated
mixed agriculture, small-scale cereal/
livestock.

50 1,000

-10 30

Tropical dry forest


Large commercial and smallholder: tree
crops, rice, cereals/roots.

700 2,500

15 30

30 300 Cereals, fibres, oil crops,


pulses, roots/tubers, tea,
coffee, sugar crops, fruit,
vegetables, timber.

Maize, sugarcane, soybean,


sorghum, energy cane,
jatropha.

Tropical grassland
Large commercial and smallholder:
extensive mixed cropping, cereal/livestock.

500 2,500

15 30

30 300 Cereals, fibres, oil crops,


pulses, roots/tubers, tea,
coffee, sugar crops, fruit,
vegetables, timber.

Maize, sugarcane, soybean,


wheat, sorghum, energy
cane, jatropha.

1,500 5,000

25 30

300 365 Cereals, fibres, oil crops,


pulses, roots/tubers, tea,
coffee, sugar crops, fruit,
vegetables, timber.

0 350

10 40

Biome and type of agriculture


Subtropical/temperate humid forest
Large commercial and smallholder: intensive
mixed agriculture, cereals and livestock, tree
crops.

Tropical humid rainforest


Large commercial and smallholder:
subsistence agriculture, livestock, tree crop,
root crop, large proportions are protected
land.
Desert
Subsistence pastoralism.

0 320 Cereals, fibres, oil crops,


pulses, roots/tubers, sugar
crops, fruit, vegetables,
timber.

0 30 Succulents.

Maize, sugarcane,
switchgrass, miscanthus,
soybean, wheat, cordgrass,
oilseed rape, sorghum.

Oil palm, sugarcane,


sorghum, energy cane.

Agave, jatropha.

Notes
* Average annual temperature, based on FAO GeoNetwork[1]. Note regions, that would be distinct from
the ranges given in this table, are identified and described in the temperate coniferous forest, tropical
humid rainforest and desert pages.
** In general, growth is limited by rainfall (or water availability) in tropical climates and by temperature
in temperate climates. Although species might have evolved locally in order to survive the extremes
of climate, some crops may not, leading to zero growing days. Crop selection and management can
potentially extend the growing season in other cases.
*** Within a biome, the suitability of a site for a particular crop depends on a range of factors, including
altitude, aspect, rainfall and soil type. Crops listed here are examples and are not intended to be a
comprehensive list.
Cereals are generally of the gramineous family and refer to crops harvested for dry grain only

(specifically wheat, rice paddy, barley, maize, popcorn, rye, oats, millets, sorghum, buckwheat,
quinoa, fonio, triticale, canary seed, mixed grain, cereals nes).

Chapter references
[1] Food and Agriculture Organization of the United
Nations, GeoNetwork (2001), Udvardys ecoregions.
[2] US Department of Agriculture, Natural Resources
Conservation Service, National soil survey handbook,
title 430-VI, Part 622 Interpretative Groups. Available
from: http://www.nrcs.usda.gov/wps/portal/nrcs/detail/
soils/survey/?cid=nrcs142p2_054242 [accessed July
2013].

[3] van Velthuizen, H. et al. (2006), Mapping biophysical


factors that influence agricultural production and rural
vulnerability, Environmental and Natural Resources,
series 11. FAO & IIASA, Rome.

5Where can biomass feedstocks be grown? | 69

70

Energy crop species have variable


characteristics, making some more
suited to certain conditions than others.

A range of examples of bioenergy crops,


some already commercially produced,
is provided here to illustrate the
variation in crops grown across different
geographic distributions and under
different managements.

Current production levels of various


crops provide perspective on the
potential for biomass development.

6 Biomass feedstock crops


Introduction to the selected crops
The following pages provide key data about crop species
and biomass types that are already in production or are
being researched for biomass. This is not a comprehensive
review of all species considered for bioenergy, as there are
many others. Instead, this collection includes at least one
representative of each major plant functional group or
biomass type. For each crop page that highlights a species,
there is a box that cites comparable plants or biomass
sources that are also widely recognized.

Yield and bioenergy deliverables


The diagrams on the crop pages indicate worldwide average
annual yield per hectare (ha) and the highest recorded
Each crop page contains the following sections.
yield per hectare (1ha = 100m 100m or 2.47 acres) with
equivalent fuel volumes and energy yields estimated. Yield
Plant
characteristics
Plant
Plant
types
types
data are based on commercial yields for widely grown crops,
Annual
Annual
and on best available data for less widely grown crops.
Icons
It should be cautioned that the yield of each crop varies
These are pictorial guides to the type of plant,
its
photoPerennial
Perennial
dramatically across growing regions according to soil type,
synthetic
the
current
major energy use (more
Woody
Woody and
Grain
Grain
ororseed
seed
Herbaceous
Herbaceous pathway
than one classification in each category sometimes applies). climate and management and, over time, in response to
dynamic climate conditions. Ranges are included in the
Photosynthetic
Photosyntheticpathway
pathway
text, and are often very large because of these variables.
Plant
types
Plant
Plant
Plant
Plant
types
types
types
types
Detailed information about yields in a specific area of
C 3C 3
CPlant
C
types
4 4
Annual
Annual
Annual
Annual
interest are provided at the end of this chapter in the
Annual
Annual
complete reference list on page 104. Symbols are used to
C3C3
C4C4
Perennial
Perennial
Perennial
Perennial
represent the following values per hectare:
Perennial
Perennial

pes

us

The statistics for the area of planting, producing countries


and total production are based on FAOSTAT data for
maize (corn), sugarcane, oil palm and soybean. For the
other crops described in this chapter, data are not available
from the FAO and specific reviews or meta analyses
published in peer-reviewed literature have been used.

Woody

Woody
Woody
Woody
Woody
Grain
Grain
Grain
orGrain
seed
ororseed
seed
or seed
Herbaceous
Herbaceous
Herbaceous
Herbaceous
Grain or seed Herbaceous
Woody Grain or seed

Photosynthetic
Photosynthetic
Photosynthetic
Photosynthetic
pathway
pathway
pathway
pathway
Propagation
Propagation
method
method
Propagation
method
ynthetic pathway
Photosynthetic pathway
Plant
Plant
C types
CCtypes
C
C CC C
C4
C4

33

C3 CC3 3 C3
Seed
Seed
Herbaceous
Herbaceous

44
C3

C4

Annual
Annual

C4 CC4 4 C4
C3
C
Stem
Stem
Rhizome
Rhizome
Micropropagation
Micropropagation
Perennial
Perennial
4
cutting
cutting Grain
ororor
root
root
cuttings
cuttings
Woody
Woody
Grain
or
seed
seed

Dry crop mass in tonnes (t) weight symbol.


Liquid fuel derivable in litres (l) barrel symbol.
Electricity that can be generated at 30% overall efficiency
kilowatt hours (kWh) light bulbs.
Practical energy use equivalent (per hectare) car or
house symbols.

Next to each car symbol is the estimated distance that can


be travelled by an average family car using biomass from
one hectare of land. This is estimated two ways: (1) based
gation method
Propagation method
on the average annual yield of liquid fuel derived from
C 3C 3
C 4C 4
CAM
one hectare; and (2) based on the greatest recorded yield
per hectare. The family car category of vehicles does not
EE
DD
CC
C C3
Seed
Seed 4 4Stem
Stem
Stem
Stem Rhizome
Rhizome
Rhizome
Rhizome
Micropropagation
Micropropagation
Micropropagation
Micropropagation
assume the greatest achievable gas mileage (miles per gallon
Plant 3Seed
lifeSeed
cycle
Bioethanol
Biodiesel
Heat
Biogas
Bioethanol Rhizome
Biodiesel Micropropagation
Heatand
andpower
powerStem
Biogas
Seed
ed
Stem
Rhizome
Micropropagation
cutting
cutting
cutting
cutting
or root
ororroot
root
or
cuttings
root
cuttings
cuttings
cuttings
or litres per kilometre), but it does reflect modern vehicle
cutting
cutting Annual
or root cuttings
or root cuttings
Perennial
efficiencies.
Other
Other
Current
Current
Current
Current
dominant
dominant
dominant
dominant
energy
energy
energy
energy
use
use
use
use
Propagation
Propagation
method
method
t dominant energy
use
dominant energy use
Primary
energyCurrent
use
Parameters used in this calculation are:

ol

Photosynthetic
pathway
Current
Current
dominant
dominant
energy
energy
use
use
Propagation
Propagation
Propagation
Propagation
method
method
method
method
Photosynthetic
Photosynthetic
pathway
pathway

Weight
CarD DD D
Weight
E EE E Car
E
D

Barrel
Barrel

Seed
Seed Biodiesel
Stem
Stem
Rhizome
Rhizome
Micropropagation
Micropropagation
Bioethanol
Biodiesel
Heat
Heat
Heat
and
Heat
and
and
power
and
power
power
power
Biogas
Biogas
Biogas
Bioethanol
Bioethanol
Bioethanol
Biodiesel
Biodiesel
Biogas
cutting
cutting Biodiesel
or or
root
root
cuttings
cuttings
Biodiesel Heat and power Bioethanol
Biogas
Heat
and power Biogas
Icons shown in grey indicate pre-commercial stages of adoption.
Other
Other
Other
Other
These
iconsdominant
are
repeated
onenergy
a fold out
on the back cover.
Current
Current
dominant
use
use
Other energy

Growth and production


Car

Car
Car
CarCar Weight
Weight
Weight
Weight Barrel
Barrel
Barrel
Barrel
EE
DD
Barrel
Car
Weight
Barrel
MapWeight
Bioethanol
Bioethanol Biodiesel
Biodiesel Heat
Heat
and
and
power
power Biogas
Biogas
Indicates
latitude
limits
for
commercial
growth
and
the
Power
Powerusage
usage
fiveOther
countries
with
the
greatest
production
of
the
crop.
Other
The political boundaries of large countries will cross
the latitude limit in some cases whereas, in general, the
growing zones do not.
CarCar
Weight
Weight
Barrel
Barrel
Global production
diagram
Crop-specific data for global production of liquid fuels,
Power
Power
Power
Power
usage
usage
usage
usage
heat or power are variable
and
difficult
to substantiate.
Power usage
Power usage
Where a crop is produced as a major product in a country,
these data are routinely recorded at country level.

Consumption gasoline 8 litres/100km and diesel


6 litres/100km.
Energy density of biofuels bioethanol, 67% of gasoline;
biodiesel, 91% of typical fossil-derived diesel blend.
Density of fuels bioethanol 0.79kg/litre and biodiesel
0.87kg/litre.
Biodiesel conversion factor 97% of crop oil (bioethanol
yield varies per crop).

Next to each house symbol is the estimated number of


months an average household energy budget could be
supported using biomass from one hectare of land.
The household energy budget assumed is for a typical UK
household with total of 20,000kWh/yr energy consumption.
Normally, around 16,500kWh of the total is related to
heating requirements, which is delivered by gas. In this
calculation, the household uses electricity for all energy
requirements including heating and cooking as well as
lighting and cooling. The power generation efficiency
assumed in the calculation is 30%.
6 Biomass feedstock crops | 71

Power
Power
usage
usage

Propagation method
Photosynthetic pathway
Plant
C 3 types

Maize (corn)
C3

Zea mays
L.
Herbaceous

Propagation method

C4

Annual

C4
Seed

Stem
cutting
Grain or seed

Woody

Rhizome
Micropropagation
Seed
Stem
Rhizome
Micropropagation
or root cuttingsPerennial
cutting
or root cuttings

Current dominant energyCurrent


use
dominant energy use
Propagation method
Photosynthetic
pathway
Plant types
Annual
C4

C3
Woody

Grain or Herbaceous
seed

hetic pathway

Perennial
Perennial
CSeed
C4 Stem
Woody
Grain
or Biodiesel
seed
3 Bioethanol
Heat
and
power
Biogas
Rhizome
Micropropagation
Bioethanol
Biodiesel Heat and power
cutting
or root cuttings

Biogas

Photosynthetic
Type of fuel: ethanol
(grain),
pathway

C4
C4

biogas (whole plant).


Other
Other
Stage of
adoption:
extensive
commercial
Current
dominant
energy
use
Propagation method
production
ofCethanol,
leading source of biofuel in
C3
4
the world, commercial biogas production.
C3

C4

Seed
Bioethanol

Car

Weight

Barrel Car

Weight

Barrel

Rhizome
BiodieselStem
Heat and power
BiogasMicropropagation

cutting
or root cuttings
Propagation method
Other
Current dominant energy use

on method

Stem
cutting

Annual

Annual

Weight Micropropagation
Barrel
Seed Micropropagation
Stem Car
Rhizome
Rhizome
E
D
cutting
or root cuttings
or root cuttings
usage Biogas
Bioethanol
Biodiesel Heat Power
and power

ominant energy use


Current dominant energy use
Other

In the US ~44% of maize was used for ethanol production


in 2012, almost equivalent to the amount used for animal
feed. In some European countries maize is the major
biogas crop.

Power usage

Plant characteristics

Maize, known
as corn
E
D in the US, was first domesticated
10,000 years ago in Central America. Its distinctive lush
Carand power
Weight
Barrel
Biodiesel Heatfoliage
and Bioethanol
power
Biodiesel
Heat
Biogas
and Biogas
tasseled
flowers
are now a common
sight
throughout the temperate and tropical world. With end
Other
Power usage
products ranging from sweet corn to silage, maize is an
important human and animal food, as well as providing
biomaterials and fuel. Despite having been introduced to
Africa onlyBarrel
a few hundred
years
is now considered
Car
Weight
Car
Weight ago, itBarrel
a vital staple there and in many areas has replaced
traditional sources of starch such as cassava, sorghum and
millet. The amount of maize consumed globally as food is
Power
dwarfed by the amount used to
feedusage
cattle.
D

Maize is a tall, fast-growing annual grass reaching 2 or 3


metres at maturity in a single growing season. Using the
highly productive C4 photosynthetic pathway, it produces
corncobs that grow from nodes on the stem. The male
flowers appear as tassels at the top of the plant.
Selective breeding of the original species focused on
developing larger kernels and creating varieties with
differing sugar levels. Thousands of maize varieties are
now available, from sweet corn and baby corn (eaten
as vegetables) to starchier cultivars for various end uses.
Green stems of maize are also used to make silage for
cattle or biogas production. There are tens of GM hybrids,
the first of which was created to give resistance to the

World map with latitude limits for growth and five top producing countries
Power usage

Power usage

54N
US

China

Top producing
countries

Million
tonnes*

US
China
Brazil
Mexico
Argentina

320
180
55
23
22

*grain

Mexico

Brazil

36S
Argentina

Global planting 160 million hectares in 2010, with a 10-year average of 150 million hectares.
Global production of maize was 850 million tonnes grain (dry weight) in 2010. Worldwide average annual grain
yield was 5.2 tonnes/ha from 2008 to 2010, but yields vary widely (with grain yields of up to 10 tonnes/ha in the
US and total biomass ranging worldwide from less than a tonne to 28 tonnes/ha).

72 | 6 Biomass feedstock crops

potentially devastating European corn borer (Ostrinia


nubilalis). Use of GM hybrids is linked to yield increases
and reduced pesticide use. Additional transgenic hybrids
have been developed for stress tolerance and resistance
to certain pests and herbicides. Innovations in crop
genetics, mechanization, fertilization, and pest and crop
management have resulted in a fourfold increase in maize
yields over the past 70 years in the US.

Where to grow it
Maize needs a frost-free growing season with optimal
growing temperatures of 24 30C. Plants grow best in
warm climates but are damaged by temperatures above
45C. It is grown successfully over a broad latitudinal
range, from 54 North to 34 South. In temperate climates
it is sown in spring and harvested in autumn; in climates
with a pronounced wet season it is planted with the first
rains and harvested as rainfall tails off. Maize is well
adapted to medium-textured (0.25 0.5mm) soils. It can be
grown in nearly every biome described in this handbook,
but is most common in temperate grassland and temperate
broadleaved forest biomes.

How to grow it
Large-scale mechanized production is the norm in
commercial farms, although smallholders in developing
countries grow and harvest maize entirely by hand.
Reduced-tillage methods such as direct drilling protect the
soil and improve soil organic carbon sequestration. Maize
has higher yield when planted in a two-year rotation with
a nitrogen-fixing crop such as soybeans. This is the most
common rotation, but maize is also sometimes grown
without rotation, in a three-year rotation with wheat
and soybean, or in a three-year rotation that follows the
sequence maizemaizesoybean. Intercropping is also
practised over a small area in some places, with a variety
of crops that can be planted between the rows of maize.

Maize average annual yield


per hectare

11

5.2 tonnes
What distance could you drive?
World

18,000km
11,000 miles

Best recorded 33,000km


20,000 miles
0

10

20

30

40 50 60 70 80 90 100
Thousand km

Average yield (tonnes/ha seed)

Dry weight in tonnes

Fertilizer
Commercial crops of maize are boosted by the application
of nitrogen, phosphorus and potassium (NPK)
fertilizer. Nitrogen is needed in relatively large quantities
(140 200kg per hectare), especially on light sandy soils.
This can lead to undesirable leaching of nitrates into
groundwater if application rates are not carefully matched
to soil type and rainfall or irrigation levels.

Pests and diseases


Maize is host to a range of diseases, including types of
mildew (Peronoscleraspora sorghi up to 80% yield loss),
leaf spot (Cercospora zeamaydis average 30% yield
loss), and blight (Exserohilum turcicum up to 70% yield
loss). Fungal diseases sometimes gain entry to plants via
holes made by maize borers, which are the most damaging
insect pests wherever maize is grown, causing grain yield
losses of 36%.
Globally, maize yields can be reduced by almost
one-third due to weeds and pests: weeds (11%), herbivores
(10%), pathogens (9%) and viruses (3%).

Defences
Herbicides
Many herbicides are used to control weeds on non-GM
varieties. Glyphosate is widely used on GM maize bred for
resistance to this herbicide. To help prevent infestation by
Striga, a parasitic plant that causes extensive damage in
Africa, seeds can be given a special chemical coating.

Global average yield


Argentina
Brazil

10

gasoline equivalent of 1,400 litres

Water
The crop needs adequate rainfall, in the range
of 670 790mm during the growing season. Most maize
agriculture in the US is rainfed, but the shallow-rooted
crop is sensitive to water limitation in drier regions where
rainfall is supplemented with irrigation.

Maize yields global average


and top producing countries

Litres of bioethanol fuel

2,000 litres

Inputs required

China
Mexico
US

8
7
6
5
4
3
2
1
0
1960

1970

1980

1990

2000

2010

6 Biomass feedstock crops | 73

Pesticides
Many different pesticides are used on maize, the type and
amount depending on the region of the world. The use of
GM maize resistant to European corn borer has reduced the
need for pesticides against this pest. A range of pesticides
is available for use against other pests such as armyworms.
Non-chemical push-pull strategies that used trap crops
to attract pests to the outside of the field in addition to
intercropping with plants that repel pests (and/or inhibit
Striga) have also been very successful.

Invasion risk
Maize is generally not considered an invasive species, but
risk of invasion with or without transgene dispersal into
the Mexican landraces (varieties) from where modern
maize arose is cited as a potential risk associated with the
large, nearby US corn crop. This risk is currently debated
in the scientific literature.

Key references

Harvesting
Mechanical harvesting is the norm in commercial
production to ensure harvest at the optimal time. As with
other grain crops, the seed must have dried sufficiently for
efficient combine harvesting. For maize kernels, optimum
moisture to avoid seed damage during harvest is usually
22%. Moisture above 30% will result in poor kernel
separation, and below 15% will result in a large portion of
the kernels being cracked and broken.
Labour-intensive manual weed control and harvesting
are still widespread in much of Africa. This involves
cutting each stem by hand and stacking the maize into
large stacks to dry further, before the kernels are removed
from the cobs. For biogas production, maize is harvested
with a chopper when a dry matter content of 25 30% is
reached. In a process called ensiling, the whole crop can
also be cut and chopped when relatively fresh, optionally
inoculated with acid-forming bacteria, and stored for later
use as animal food (silage).

Yield

Doebley, J. (2004), The genetics of maize evolution,


Annual Review of Genetics, vol. 38, pp. 37 59.
Kim, S. & Dale, B. E. (2005), Environmental aspects
of ethanol derived from no-tilled corn grain: non
renewable energy consumption and greenhouse gas
emissions, Biomass and Bioenergy, vol. 28, no.5,
pp. 475 489.
Leakey, A. D. B. et al. (2004), Will photosynthesis of
maize (Zea mays) in the US corn belt increase in future
[CO2] rich atmospheres? An analysis of diurnal courses
Plant types
of CO2 uptake under free-air concentration enrichment
(FACE), Global Change Biology, vol. 10, no. 6,
pp. 951 962.
Oerke, E. C. (2005), Crop losses to pests, The Journal of
Woody Gr
Herbaceous
Agricultural Science, vol. 144, no. 1, pp. 31 43.
US Department of Agriculture, Economic Research
Photosynthetic pathwa
Service, corn background data.
C3
Plant types
This box highlights another international commodity
C3
crop that is harvested for grain.

Wheat (Triticum
aestivum)
Woody
Herbaceous

Worldwide average annual grain yield was just over


5 tonnes/ha, but yields vary hugely (from one-fifth of a
tonne to 28 tonnes/ha if the whole plant is harvested).
Plant types
High yielding fields in the US can exceed 10 tonnes/ha
of grain.
Ethanol yield from maize is ~390 litres per tonne of grain.

Alternative markets

Herbaceous

Woody

Propagation method

Photosynthetic pathway
C 4 types
Plant

C3

Annual
C4

A
Stem
cutting
Pe

Seed

Perennial
Herbaceous

Grain or seed

Perennial

Grain or seed

C3

C4
Annual
C4

Woody

Grain or seed

Current dominant ene


Propagation method
Photosynthetic pathway
Photosynthetic
pathway Plant types
Plant
types
Maize is grown for food and animal feed, with the
Annual
Annual
proportion allocated to these markets heavily dependent
Annual
C3
C4
C3
C4
E
D
on the region of the world. Turned into high-fructose corn
Perennial
Perennial
syrup, it is an ingredient in many processed
foods.
C3 Stem
Grain orHerbaceous
seed
Woody Grain
or seed C4 RhizomeBioethanol
Herbaceous
Seed
Micropropagation
C3 It is Woody
C4
Biodiesel He
Photo in production: fromcutting
Midwestern
also used to make starch and cereals, and fermented to
or rootUS.
cuttings

Why
is
it
similar?
Also
one
of
the
top
grain
Photosynthetic
pathway
Photosynthetic
pathway
make alcoholic drinks such as bourbon. Corn oil is another
Other
Current
dominant
energy
use harvested
commodities
in the
world; grain
of plant
important foodstuff that is essentially a by-product of
Propagation method
Propagation
method
for food,
C3
C 3 feed and
C4
C 4 fuel; intensive inputs; often grown
making animal feed. Maize is also used as a biomaterial
for
in rotation with other crops.
packaging and disposable cups.
C3
C4
C
WhatCmakes
it different?
Uses C3 photosynthesis;
3
Car
E 4
D
can grow in colder climates than tolerated by corn;
Co-products
Bioethanol
Biodiesel
Seed Heat and power
Stem Biogas Rhizome
Microprop
lower
yielding
than corn.
Seed
Stem
Rhizome
Micropropagation
root cuttings
From the crop
cutting Growing
or
root cuttings
regions:
Top five producing cutting
countriesorare
Propagation method
Propagation
method
Other
Maize stover and cobs (the crop residue after the kernels
China, India, the US, the Russian Federation and
Current dominant energy use
Current
dominant
energy
use
have been extracted) has potential use as bioenergy
France.
feedstock for cellulosic ethanol production.

Weight Micropropagation
Barrel
From conversion to fuel
Seed
Stem
Seed Micropropagation
Stem Car
Rhizome
Rhizome
E
D
E
D cutting
cutting
or root cuttings
or root cuttings
The main co-product of fermentation is sold as highBioethanol
Biodiesel Heat and power Biogas
Bioethanol
Biodiesel Heat and power Biogas
protein animal feed called dry distillers grains
and solubles.
Current dominant energy use
Current dominant energy use
Power
This co-product is a major commodity that adds value to
Other
Other
the corn grain ethanol supply chain. Carbon dioxide from
fermentation is also often collected and used to make
E
D
E
D
carbonated beverages.
Bioethanol

74 | 6 Biomass feedstock crops

Biodiesel
HeatWeight
and Bioethanol
power Biogas
Biodiesel Heat and power Car
Biogas
Car
Barrel

Other

Power usage

Other

Car

Weight

Barrel Car

Weight

Barrel

Weight

Barrel

Propagation method
Propagation method

Photosynthetic pathway

C3
C3
Woody

Plant
types
C4

Sugarcane

Annual

Annual

C4

Seed

Stem
Seed
cutting

Perennial
Saccharum
officinarum
L.
Woody Grain or seed
Herbaceous

Rhizome
Stem Micropropagation
Rhizome
Micropropagation
or rootcutting
cuttings or root cuttings
Perennial

Grain or seed

Current dominant
energy
use energy use
Current
dominant
Propagation
method
Photosynthetic
pathway
hetic pathway
Plant types
Annual

C4
C4

C3

C4

Perennial

Perennial
C3 Woody Grain
C4 orBioethanol
seed Micropropagation
Herbaceous Stem
Biodiesel
Heat and
power Biogas
Bioethanol
Biodiesel
Heat and power
Seed
Rhizome
cutting
or
root
cuttings
Type of fuel: solid combustion of residues (bagasse)

Photosynthetic pathway Other


Other
for heat and electricity, ethanol.
Current dominant
energy use
Propagation
method
on method Stage
of adoption:
extensive commercial
C3
C4
production, leading source of biofuel in the world
before
the recent
rise of corn ethanol.Car
C3
C4
WeightCar
E
D

StemBiogas
Rhizome
Micropropagation
Stem
Rhizome Seed
Micropropagation
Bioethanol
Biodiesel
Heat
and power
or root cuttings
cutting Propagation
or root cuttingsmethod cutting

Other
ominant energy use Current dominant energy use

Barrel
Weight

Biogas

Barrel

Plant characteristics

Sugarcane is a giant perennial grass, growing stems up to


4 metres high and 5 centimetres in diameter. It uses a C4
Seed
Stem
Rhizome
Micropropagation
photosynthetic
pathway that is particularly efficient in
Car E
Weight
Barrel
D
D
cutting
or root cuttings
concentrating sucrose into multiple stems that grow from
Power usage
Bioethanol
Powerthe
usage
Biodiesel Heat and power
Biogas Biodiesel Heat and power Biogas
plantbase.
Current dominant energy use
At the end of the 19th century, breeders settled on S.
Other
officinarum as the variety of choice for yield and disease
resistance. Today, there are around 70 cultivars of S.
E
D
officinarium in commercial use, with 58 of them being
Once the instant gratification of chewing sugarcane was
hybrids, often crossed with S. spontaneum. At the time of
Car
Weight
Barrel
Bioethanol the
Biodiesel
and power
Biogasits origins in the
Car
Weight
Barrel
discovered,
spreadHeat
of the
plant from
writing, there are no commercial GM cultivars although
island of New Guinea and India was rapid. Carried by Arab
field trials are ongoing.
Other
Power usage
traders, sugarcane
surged across the Middle East and North
Africa, reaching Spain by 715AD. Sugarcane crossed the
Where to grow it
Atlantic in 1493 with Columbus who introduced it to the
Sugarcane does not survive frosts and it grows best in areas
Caribbean island
the Spanish carried it
Carof Hispanolia,
Weight andBarrel
with long, warm growing seasons followed by slightly
on to South America. The popularity of processed sugar in
cooler and drier but sunny ripening and harvesting season.
Europe, in the 18th century, led directly to the huge rise in
This restricts the commercial raising of sugarcane to
plantations across European colonies in the Caribbean and
Power usage
land below 1,500 metres from 36 North to 31 South,
Brazil.
Powerin
usage
World map with latitude limits for growth and five top producing countries
Power usage

Top producing
countries
Brazil
India
China
Thailand
36N Mexico

China
Mexico

India

Million
tonnes*
720
290
110
69
50

*fresh weight

Thailand
Brazil

31S

Global planting 24 million hectares


Global production 1,700 million tonnes (fresh weight)

6 Biomass feedstock crops | 75

primarily in the tropical and subtropical biomes but also


in the southern edges of the temperate broadleaved forest
biome. The plants are tolerant of a wide variety of soil
types. The soils must be well drained, however, if planted
on heavy clays.

How to grow it
Sugarcane is grown as a monoculture and, although
it produces seeds, is harvested before seed production
occurs to maximize sugar concentrations of the stems.
Commercial cultivation is accomplished by planting
sections of stems. These produce new stalks known as
ratoons, which need between nine and 24 months to
reach a harvestable state. Plants can last up to 10 years
but productivity decreases after five to seven ratoon
cycles, so they are normally re-planted then to maximize
yield. Stems are now often machine planted and the soils
are typically fully tilled, although minimum tillage is
being practised in some regions of Brazil.

Inputs required
Water
Sugarcane is sensitive to drought: 1,500 to 2,000mm
of rain is required to avoid irrigation. In water-limiting
conditions, research has shown that 10mm of water
raises yields by 1 tonne of cane per hectare. Vinasse, an
organic liquid containing high amounts of organic carbon
(6,00023,000mg C per litre) generated as a by-product
of fermentation, is also sometimes recycled to the
sugarcane fields to supplement water needs; 10 15 litres
of vinasse is generated for every litre of ethanol.
Fertilizer
To maximize yield, commercial growers use significant
amounts of nitrogen, phosphorus and potassium (NPK)
fertilizer as well as lime, with nitrogen application
having the greatest environmental impact. The most
effective method of delivering these elements is through

Sugarcane average annual yield


per hectare

90

71 tonnes
What distance could you drive?
49,000km
30,000 miles

World
Best recorded
0

10

20

30

84,000km
52,000 miles

40 50 60 70 80 90 100
Thousand km

76 | 6 Biomass feedstock crops

Average yield (tonnes/ha fresh mass)

Wet weight in tonnes (6373% water)

Sugarcane is affected by many diseases caused by fungal,


bacterial and viral attacks with an estimate, in India, of
15% crop loss due to disease. The red rot fungal disease,
which is common globally, severely damages yields
and has wiped out one sugarcane variety in India, while
another fungal disease, smut, can cause crop losses of 70%.
The plants are also commonly vulnerable to attack
from insects that bore into roots and stalks, as well as
nematodes in the soil that eat roots. The most serious pest
group are white grubs of the beetle family Coleoptera,
which can reduce sugar yield by up to 39%. Some
mammals are also a threat, with rats known to cause crop
loss, again of 39%, in many growing regions.
The effects of pests and diseases are always a risk, but
can often be effectively controlled.

Defences
Herbicides
A range of herbicides, including glyphosate, metalochlor,
alachlor and paraquat, are commonly used prior to
planting and to aid plant maturation. Use varies by region
and by legislation, with some herbicides banned in certain
legislations.
Pesticides
A wide range of pesticides is used against threatening
species, including root borers and white grubs. Pesticide
application is widely variable and difficult to generalize.

Global average yield


Brazil
China

80
gasoline equivalent of 3,900 litres

Pests and diseases

Sugarcane yields global average


and top producing countries

Litres of bioethanol fuel

5,800 litres

fertigation, where the nutrients are added to irrigation


water and via application of vinasse. Reported rates for
fertilizer application vary widely within a range of
45300kg per hectare.
Although still debated in scientific literature, there
is some evidence that biological nitrogen fixation (via
nitrogen-fixing bacteria) supplements the nitrogen
requirements of sugarcane in rare cases.

India
Mexico
Thailand

70
60
50
40
30
20
1960

1970

1980

1990

2000

2010

Harvesting

Invasion risk

The crop can be harvested by hand or mechanically. When


harvested manually, two techniques can be employed
depending on local practice and custom. Mature green
canes can be cut on ripening or on the day after the fields
have been burned. Burning removes poisonous snakes and
non-useful plant material, and improves worker safety
by removing sharp leaves, so the yield per worker is more
than twice as high as for green cane: five tonnes per day
compared with two. Mechanical harvesting is carried
out on unburned crops and, although the productivity
is much higher, it leaves long stumps and residual cane,
which promote pests and disease as well as causing soil
compaction and splitting of the cane stumps that inhibits
ratoon growth. The practice of burning is being phased
out of commercial operations in response to concerns over
local air quality, the wasting of energy contained in the
leaf material, and improvements in harvesting machinery.

Modern cultivars of sugarcane are not considered invasive


and typically fail to persist without human assistance.
Most of the competitive and invasive traits found in the
original species have been lost during plant breeding.
Hybrid cultivars do not produce rhizomes or produce
vigorous seedlings, although after a ratoon cycle some
volunteer plants may appear in the next crop. There are
certain sugarcane cultivars, in particular S. spontaneum,
that are considered invasive due to rhizome production.

Key references

Yield
Sugarcane worldwide yields average 71 tonnes fresh
wet weight per hectare, with wide variation between
countries (0.9122 tonnes). In Brazil, converting sugarcane
to ethanol produces on average 82 litres per tonne. Brazil
produced 21 billion litres in 2011, some 25% of global
bioethanol production.

Hartemink, A. (2008), Sugarcane for bioethanol: soil and


environmental issues, Advances in Agronomy,
vol.
99,
Plant
types
pp. 125 182.
James, G. (ed.) (2004), Sugarcane, 2nd edition, Blackwell
Publishing Ltd, Oxford, UK.
Lisboa, C. C. et al. (2011), Bioethanol production
from
Woody Grai
Herbaceous
sugarcane and emissions of greenhouse gases known
Photosynthetic
pathway
and unknowns, GCB Bioenergy, vol. 3, pp. 277
292.
Renewable Fuels Association (2012), Accelerating
industry innovation: 2012 ethanol industry outlook.
C3
C4
RFA, Washington DC.
C3

This box highlights another important crop that is also


harvested for high soluble sugar concentrations.
Propagation method

Alternative markets
Although sugarcane has been used for large-scale
bioethanol production since 1975, the primary
Plantmarket
types is
sugar, and sugarcane provides 70% of the worlds sugar.
Many sugarcane mills serve both markets simultaneously.

Co-products

Herbaceous

Plant types

Annual

Annual

Woody

Grain or seed

Herbaceous

Seed

Perennial
Woody

Grain or seed

Photosynthetic pathway
Photosynthetic
pathway Plant types
Plant
types
Crop residue
Annual
Leaves and other plant debris, often called trash that is
Annual
C3
C4
C 3 animal C 4
often left in the field, are sometimes collected for
Perennial
feed or composted and returned to the fields.
Herbaceous
C3

C4

Woody
C4

Grain orHerbaceous
seed

C3 Woody

C4 or seed
Grain

Stem
cutting
Perennial

Current dominant energ


Annual

Perennial

Bioethanol

Biodiesel Heat

Sweet sorghum (Sorghum bicolor)


Bagasse
Photosynthetic pathway Photosynthetic pathway
Photo in production: Brazilian cerrado.
After the cane is crushed and the juice extracted, the
Other
similar? Alsomethod
uses C4 photosynthesis;
resulting fibrous material, called bagasse, isPropagation
used for a method Why is it Propagation
C3
C3
C4
C4
high sugar
concentration
in stem.
variety of purposes. Its original use to fuel the process
What makes it different? Grows in colder climates;
continues, both for the extraction of juice and for the
C3
C3
C4
C4
Car
wide genetic
variation;
typically lower yielding
conversion to bioethanol, with the added benefit
of
than sugarcane; produces grain that is a primary
lowering the overall carbon footprint of the resultant
Seed
Stem
Rhizome
Micropropagation
Seed
Stem
Rhizome
Micropropagation
product.
biofuel product. Some plants produce enough energy
or root cuttings
cutting
orPropagation
root cuttings method cutting
Propagation
method
Growing region: Top 10 producing countries are
from burning bagasse to power the entire process, with
Current
dominant
energy
useArgentina,
the
US, India,
Mexico,
China,
Nigeria,
surplus to export as electricity. This generation
is rapidly
Current
dominant energy
use
Sudan, Ethiopia, Australia, Burkina Faso.
increasing, enabling Brazil to meet 6% of its electricity
demand in 2011, with the potential of generating far more
from the bagasse available across the country. Bagasse
is
Seed
Stem
Seed Micropropagation
Stem
Rhizome
Rhizome
Micropropagation
E
D
E
D cutting
cutting
or root cuttings
or root cuttings
also used to make paper and board.
Bioethanol

Biodiesel Heat and power Bioethanol


Biogas

Biodiesel Heat and power

Biogas

Vinasse
Current dominant energy use
Current dominant energy use
Other
A liquid rich in plant nutrients, vinasse is a Other
significant byproduct of the distillation process (10 litres of vinasse are
produced for every litre of ethanol). If carelessly disposed
E
D
E
D
of it can cause environmental pollution by lowering the
Car and power
Weight
Bioethanol
HeatWeight
and Bioethanol
power Biogas
Biodiesel Heat
Biogas Barrel
Car
Barrel
pH value of the soil and water, but it is often
amendedBiodiesel
and
returned to sugarcane fields as a form of irrigation and
Other
Other
fertilization.

Car

Weight

Power usage

Barrel

Car

Weight

Barrel

6 Biomass feedstock crops | 77

Power usage

Power us

Propagation method

Propagation
Photosynthetic
pathway method
C 3types
Plant

C4

Switchgrass
Woody

Annual

Annual
C4 Seed

C3

Stem
cutting
Perennial
Woody Grain or seed

Panicum virgatum L

Herbaceous
Grain or seed

Stem
Rhizome Seed
Micropropagation
cutting
or root Perennial
cuttings

Rhizome
Micropropagation
or root cuttings

Current dominant energy use Current dominant energy use


Propagation method
Photosynthetic
pathway
hetic pathway
Plant types
C3

C4
C4

Herbaceous

C4

CSeed
3Woody

Perennial

Annual

Perennial
C4 Stem
Grain
or seed Biodiesel
Rhizome
Micropropagation
Bioethanol
Heat and
power
Biogas
Bioethanol
cutting
or root cuttings

Type of fuel:
solid
combustion for heat and
Photosynthetic
pathway
Other
Other
electricity,
cellulosic energy
ethanol.
Current dominant
use
Propagation
method
ion method CStage
of
adoption:
early
commercial
stage
for
heat
C4
3
and electricity, pre-commercial stage for liquid
C3ethanol. C4
Car
Weight
Barrel
E
D

D
Biodiesel Heat and power

Car

Weight

Biogas

Barrel

Bioethanol
BiodieselStem
Heat and power
BiogasMicropropagation
SeedMicropropagation
Rhizome
Rhizome
cutting
or root cuttings
or root cuttings

Stem
cutting Propagation method

Other
dominant energy useCurrent dominant energy use

Weight
Barrel
StemCar
Rhizome
Micropropagation
D
cutting
or root cuttings
PowerBiogas
usage
Bioethanol
Biodiesel Heat and power
Biogas Biodiesel Heat and power

Seed

Current dominant energy use


Other

Car

E
D
Switchgrass is a native wild grass found in a range of
Car and power
Weight
Barrel
Bioethanol
Biodiesel
Heat
Biogas America.
Weight from
Barrel
habitats
eastern
Canada to
Central
Its
deep roots and tolerance of poor soils make it useful for
Power usage
Other
preventing erosion, and its ornamental habit is valued
by designers aiming to create prairie-style naturalistic
gardens. Low fertilizer requirements and resistance to
insect attack give
potential
Car switchgrass
Weight as-yet-unfulfilled
Barrel
as a feedstock for bioenergy.

Power usage

Power usage

Plant characteristics
Switchgrass is a clump-forming C4 perennial that grows
up to 1.5 metres tall. It spreads slightly by stout rhizomes,
especially
wetter conditions, and the delicate flower
Powerin
usage
heads and seeds are attractive in late summer. There are
distinct upland and lowland varieties of switchgrass
(about 25 cultivars in total), and several naturally
occurring strains of the plant. These have been selected
for improved yield or ornamental qualities. The lowland
varieties have greater potential yields than upland
varieties, but are more susceptible to cold damage and are
thus less suited to higher latitudes.
In 1992 the US Department of Energy began a research
programme to develop switchgrass as a bioenergy
feedstock. Research is ongoing into GM varieties and there
is now evidence that modifications to lignin content,
for example, can reduce pretreatment costs and increase
fermentation yields.

World map with latitude limits for growth


Power usage

55N

17N

78 | 6 Biomass feedstock crops

Where to grow it
The wide natural range of switchgrass reflects the genetic
variability of the plant useful to the grower in that
there are varieties to suit different microclimates and
soil conditions. The genetic variability also provides a
useful starting place for plant breeders seeking to improve
productivity. Switchgrass has in the past been grown in
latitudes between 55 and 17 North, but could also be
grown in the southern hemisphere. Switchgrass will not
thrive in extremely acid or alkaline soils, or where the soil
temperature reaches 40C, and is most common in the
temperate broadleaved and temperate grassland biomes.

How to grow it
Switchgrass is usually grown as a monoculture, but
because it is an out-crossing species, the monocultures are
genetically diverse populations. It is typically established
by seed drilled (or spread and raked) directly into clean
ground. Seed priming can sometimes improve germination
rates. Prior tillage is not necessary, but a firm seedbed
is required. Growth is initially slow and there is no
harvest in the first year. But by the second year biomass
production can reach 60% of a mature stand, and the crop
remains productive for more than a decade.

Inputs required

Fertilizer
Because switchgrass evolved under low-nutrient
conditions, it grows well without additions of phosphate
and potassium, although it may require supplements over
the long term. Some additional nitrogen can be beneficial,
although not in the first year, as this encourages weed
growth. Too much nitrogen causes switchgrass to lodge
(grow luxuriantly and collapse particularly under heavy
rain or strong winds), which hampers harvesting.

Pests and diseases


Insects that chew leaves and roots can be pests of both
seedbeds and established stands. Insect damage is
generally slight, however, and does not require pesticide
use. Fungal diseases and the barley yellow dwarf virus
have been reported. Yield losses greater than 50% were
reported from smut disease.

Defences
Herbicides
Various broad spectrum herbicides are widely used to clear
the ground of broadleaved weeds before drilling seed.
Pesticides
Not normally necessary.

Harvesting

Plant types
Plant types

Annual Annual
Water
Essentially a form of haymaking, harvesting switchgrass
Perennial
Perennial
Switchgrass is found in the wildHerbaceous
in Herbaceous
damp
areas,
but it
Woody Woody
Grain orGrain
seed or seed
involved cutting the grass using a mechanical mower.
is considered to be reasonably drought-tolerant
under
Photosynthetic
Photosynthetic
pathway
pathway
The dry grass is then baled, or sometimes chopped and
cultivation. The genetic variability mentioned above
C
C
C
C
in a module building system that compresses the
Plant types
Plant
Plant
types
Plant types
gives a wide range of average water
usetypes
efficiencies,
from Annual Annualcollected
Annual Annual
C
C
C
C
chopped material into a denser mass for transportation.
2 103kg/ha/mm. Irrigation is not normally used but can
Perennial
Perennial
Perennial
Perennial
WoodyHerbaceous
Woody
Grain Woody
orGrain
seed Woody
or Grain
seed orGrain
seed or seed
Herbaceous
Herbaceous
Herbaceous
boost production in drought conditions.
Propagation
Propagation
method
method
3

Photosynthetic
Photosynthetic
Photosynthetic
pathway
Photosynthetic
pathway
pathway
pathway
C3

C3

C4 C3

C 4C 3

C4

C4

Seed Seed Stem Stem RhizomeRhizome


Micropropagation
Micropropagation
C3
C3
C4 C3 C4C3
C
C
cutting cutting
or4 root cuttings
or4 root cuttings

Current
Current
dominant
dominant
energyenergy
use use
Propagation
Propagation
Propagation
method
Propagation
method
method
method
E

Bioethanol
Biodiesel
Biodiesel
Heat
andHeat
power
and power
Biogas Biogas
Switchgrass average annualBioethanol
yield
per
hectare
Seed Seed Seed
Stem Seed
Stem Rhizome
Stem Rhizome
Stem
Micropropagation
Rhizome
Micropropagation
Rhizome
Micropropagation
Micropropagation
cutting cutting
cutting
cutting
or root
cuttings
or root
cuttings
or root cuttings
or root cuttings

Other Other

Current
Current
dominant
Current
dominant
Current
energy
dominant
energy
use
dominant
energy
use energy
use use

Litres of bioethanol fuel

Car

Effective energy
DE D
D
available

Car Weight Weight Barrel Barrel

DE

Bioethanol
Bioethanol
Biodiesel
Bioethanol
Biodiesel
Bioethanol
HeatBiodiesel
andHeat
power
Biodiesel
andHeat
power
Biogas
andHeat
power
Biogas
and power
Biogas Biogas

Other Other Other Other

World

2,900 litres
gasoline equivalent of 2,000 litres

Car

Car Weight
Car Weight
CarBarrel
WeightBarrel
Weight Barrel Barrel
Power usage
Power usage

20,000kWh

Dry weight in tonnes

Best recorded

How long could


you supply energy
for a house?
World

12

months

Best recorded

20

14 tonnes

Power usage
Power usage
Power usage
Power usage

33,000kWh

months

What distance could you drive?


World

25,000km
15,000 miles

40,000km
Best
recorded 25,000 miles
0

10

20

30

40 50 60 70 80 90 100
Thousand km

6 Biomass feedstock crops | 79

Yield

Key references

Historically, switchgrass has been grown for forage but


its potential as a source of liquid biofuels has reached
an advanced stage of research. Because its potential as a
lignocellulosic feedstock is still being investigated, there
are no reported data on this crop from agencies such as the
FAO. In North America and Europe, switchgrass yields
range from 10 to 23 tonnes/ha (dry weight) and conversion
efficiencies to ethanol range from 180 to 240 litres/tonne.
In terms of thermal energy, switchgrass has a calorific
value of 17GJ/tonne.

Alternative markets
Switchgrass has traditionally been used as fodder for
cattle, and it is often planted to stabilize soils at risk of
erosion and to increase biodiversity as part of conservation
projects. Like other lignocellulosic materials, it can be
burned in power stations to produce electricity.

Co-products
From the crop
Currently none.

McLaughlin, S. & Kiniry, J. (2006), Projecting yield and


utilization potential of switchgrass as an energy crop,
Advances in Agronomy, vol. 90, pp. 267 297.
Monti, A. (ed.)
(2012),
Switchgrass. Springer,Plant
London.
Plant
types
types
Annual
Vogel, K. P. (2004), Switchgrass, in Moser L.E. et al.
(eds) Warm-season (C4) grasses. American Society of
Agronomy, Crop Science Society of America and Soil
Perennial
Science Society
of America,Woody
Madison,
WI,
Grain
or US.
seed
Woody G
Herbaceous
Herbaceous
Lewandowski, I., Scurlock, J. M. O., Lindvall, E. &
pathwayand current
Photosynthetic pathw
Christou, M. Photosynthetic
(2003), The development
status of perennial rhizomatous grasses as energy crops
C 3 25,
C 4 and Bioenergy, vol.
in Europe and the CUS,
Biomass
3
types
no. 4,Plant
pp. 335
361.
C3

Annual

C3

C4

C4
C4

This box highlights another temperate perennial grass


Perennial
Woody Grain
or seed
Herbaceous
species
with wide tolerance
ranges.
Propagation method
Propagation method
Photosynthetic pathway

Plant types

PlantCtypes
3

C4

C3

Seed
C4

Annual

Annual

Stem
cutting

Perennial

Stem
RhizomeSeedMicropropagation
cutting
or root cuttings
Perennial

From conversion to fuel


Woody Grain or seed
Herbaceous
Woody Grain or seed
Herbaceous
Current dominant energy useCurrent dominant en
Wastes after ethanol production could be burned as fuel.
Propagation method
Photosynthetic pathway
Photosynthetic pathway
Plant types

Invasion risk

Annual

C3

Switchgrass can easily become an invasive species and


displace other native species if not properly managed.
C3
It can grow well on marginal lands, and interspecific
competition can affect water and nitrogen availability.

C4
C4

C3

C4

Perennial

Photosynthetic pathway
Other
Other
Photo in production:
Current dominantIllinois,
energyUS.
use
Propagation
is it similar?method
Also has wide genetic variation;
Propagation method Why
C3
C4
growing regions
very large and diverse; C4
perennial;
limited
commercial production;
C3
C4
E a dedicated
D
Car crop. Weight
Barrel
potentially
energy
What
makes
it Biodiesel
different?
Can
grow
in Biogas
saturated
Bioethanol
Heat
and
power
Seed
Stem
Rhizome
Micropropagation
Seed
Stem
Rhizome
Micropropagation
and
salty soils.
or root cuttings
cutting Propagation
or root cuttings
method cutting
Other
Growing region: Limited commercial production,
Current
dominant
energyacross
use all of North
but
the
native
range spreads
Current dominant energy
use
America, and parts of South America, Africa and
Europe.
E

Bioethanol

Weight
Barrel
StemCar
Rhizome
Micropropagation
D
cutting
or root cuttings
Bioethanol
Biodiesel Heat and power
Biogas Biodiesel Heat and power Biogas

Seed

E
Car

Bioethanol
Weight

Car
WeightBiogas Barrel
Biodiesel
Heat and power
Barrel
Power usage

Car

Weight

Power usage

Power usage

80 | 6 Biomass feedstock crops

Power usage

Other

Power usage

Car

Current dominant energy use


Other

Other

Perennial
CGrain
Woody
orBiodiesel
seed
Herbaceous C3Seed
Rhizome
Micropropagation
4 Stem
Bioethanol
Heat and power
Bioethanol
Biogas Biodiesel
cutting
or root cuttings
Cordgrass (Spartina spp.)

Barrel

Powe

synthetic pathway
C4

C4
Woody

Propagation method

Propagation method
Plant types

Miscanthus

Annual

Annual
Seed

Stem
cutting
Perennial
Woody Grain or seed

Stem
RhizomeSeed
Micropropagation
cutting
or root Perennial
cuttings

Miscanthus x giganteus Greef et Deuter

Grain or seedHerbaceous

Rhizome
Micropropagation
or root cuttings

Current dominant energy use Current dominant energy use


agation method
Photosynthetic pathway
hetic pathway
Plant types
Annual

C3

C4

Seed

C4

Perennial

Perennial
Stem
Rhizome
C3Woody Micropropagation
C4 or seed Biodiesel Heat and power Biogas
Grain
Herbaceous
Bioethanol
Bioethanol
cutting
or root cuttings

C4

D
Biodiesel Heat and power

Biogas

Type of fuel:
solid
combustion for heat and
Photosynthetic
pathway
Other
Other
ent dominant energy
use cellulosic ethanol.
electricity,
Propagation
method
on method CStage of adoption: commercial use for heat and
C4
3
electricity, developmental stage for liquid biofuel.

anol

C3

C4

Car

Biodiesel Heat and power Biogas


Stem
Stem
RhizomeSeedMicropropagation
cutting Propagation
or root cuttings
method cutting

Weight

Barrel

Car

Weight

Barrel

Rhizome
Micropropagation
or root cuttings

ominant energy use Current dominant energy use


Car

Weight
Seed

Barrel
Stem
Rhizome
Micropropagation
E
D
D
cutting
or root cuttings
PowerBiogas
usage
Bioethanol
Biodiesel Heat and power
Biogas Biodiesel Heat and power

Current dominant energy use


Other

Car

Originating in South-East Asia, varieties of miscanthus


have Elong beenD
valued in gardens for their tall, upright
stature
and elegant
But
in theBarrel
past 30 years,
Car and power
Weight
Bioethanol
Biodiesel
Heat
Biogas
Weight
Barrel flowerheads.
as its potential to become a biomass crop was realized,
Power usage
Other efforts have focused on productivity. Natural
breeding
hybridization between two different species (M. sinensis
and M. sacchariflorus) has created the giant miscanthus
(Miscanthus x giganteus), a sterile hybrid that produces
Weight
Barrel
large quantitiesCarof biomass
very efficiently.
Giant
miscanthus is the only genotype commercially grown
for biomass.
Power usage

Power usage

With its healthy, vigorous growth and low management


requirements, miscanthus is easy to grow. There are
challenges, however, in producing the rhizomes required
for establishment on a large commercial scale.
Power usage

Plant characteristics
There are more than a dozen species of miscanthus, all
of them perennial grasses with spreading rhizomes (an
underground organ that produces both shoots and roots).
Some grow as tall as 4 metres. Most of the leaves are
firmly attached to the stems, although some fall as litter
in autumn and winter. The tough woody stems persist
through the winter. New shoots emerge from the rhizome
as temperatures rise in the spring. Giant miscanthus is a
sterile hybrid that does not produce seed, making it more
useful than similarly productive non-hybrid cultivars that
have invasive tendencies through seed dispersal. Like
maize, sugarcane and switchgrass, it photosynthesizes
using the C4 metabolic pathway.

World map with latitude limits for growth


Power usage

56N
37N

It is estimated that 30,000ha is grown in Europe for co-firing in the UK


and for heat in Austria, Switzerland and Germany.

6 Biomass feedstock crops | 81

Where to grow it

Fertilizer
Regarded as a low-input crop, miscanthus has in some
sites been grown successfully for more than a decade with
no nitrogen inputs. The autumn leaf-fall returns some
nutrients to the soil but, more importantly, the perennial
crop has an efficient nutrient cycle, where in autumn
nutrients are relocated from the above-ground shoots to
the underground rhizomes. Recent studies show little
response to nitrogen application in the first five years and
a small response in subsequent years (0.02 tonnes/ha
increasing in yield for every kilogram of nitrogen applied
per hectare) although this is likely to be site dependent.
There are indications that there may be nitrogen-fixing
bacteria associated with miscanthus in some places.
Potassium and phosphorus fertilizer should be applied
according to withdrawal, which is about 0.5kg P and
1kg K per tonne of dry biomass.

Miscanthus is adaptable to a range of climates and trials


have been successful in latitudes from 56 to 37 North.
Its tolerance of cold is unusual in C4 plants and it can
withstand freezing winter temperatures once established.
The best yields are achieved in soils that do not dry
out. Interestingly, miscanthus also does well on sandy
soils as long as water supply is sufficient, because weed
competition is less in such conditions.

How to grow it
As the crop will remain in the ground for two decades or
more, good establishment is crucial, for which thorough
ground preparation is vital. After removing weeds using
a broad-spectrum herbicide, ground is ploughed and a
good tilth created. If no herbicide is applied, other weed
control methods are typically required. Miscanthus is
usually established by planting small pieces of rhizomes,
ideally using specialized machinery. It is important to use
fresh, vigorous rhizomes. Planting is done in spring once
the soil has warmed and there is still plenty of moisture
available. A seeded variety has recently been developed
for Miscanthus x giganteus but it has not yet been broadly
tested.

Inputs required

Pests and diseases


Miscanthus is regarded as a crop with few pests and
diseases, but it is not immune to damage: the corn borer
(Diatraea grandiosella) can reduce dry matter yield by
up to 30%. Sixty-nine pathogens have been recorded that
affect miscanthus, although no severe damage has been
reported.

Plant types
Plant Plant
types types

Defences

Annual Annual Annual

Water
Perennial
Perennial
Perennial
Herbicides
Woody Woody
GrainWoody
orGrain
seed orGrain
seed or seed
Herbaceous
Herbaceous
Herbaceous
Miscanthus crops are drought-tolerant except in the
Photosynthetic
Photosynthetic
Photosynthetic
pathway
pathway
pathway
Broad-spectrum herbicides are needed to control weeds at
early stages of crop establishment, when it may require
intervals during the first two years. After that, the roots
C
C
CC
C
C
Plantrates.
types
Plant types
Plant
types
Plant Plant
typestypes
Plant types
irrigation to improve establishment
Miscanthus
Annualand
Annual canopy
Annual Annual Annual
of Annual
the grass suppress weed growth.
C
C
CC
C
x giganteus requires a minimum of 450mm
of Cwater
per
Perennial
Perennial
Perennial
Perennial
Perennial
Perennial
year, but will attain greater yields Herbaceous
with
higher
rainfall.
WoodyHerbaceous
Woody
Grain
Woody
orGrain
seed
Woody
or Grain
seedWoody
orGrain
seedWoody
orGrain
seedorPesticides
Grain
seed or seed
Herbaceous
Herbaceous
Herbaceous
Herbaceous
3

4 3

4 3

Propagation
Propagation
Propagation
method
method
method
Photosynthetic
Photosynthetic
Photosynthetic
pathway
Photosynthetic
pathway
Photosynthetic
pathway
Photosynthetic
pathway
pathway
pathway

Not generally required.

C3

C3

C4 C3

C 4C 3

C 4C 3

C 4C 3

C4

C4

Seed Seed Seed


Stem Stem Rhizome
Stem Rhizome
Micropropagation
Rhizome
Micropropagation
Micropropagation
C3
C3
C4 C3 C4C3
CC3
CC3
C4
C4
cutting cutting
cutting
or4 root
cuttings
or4 root
cuttings
or root cuttings

Current
Current
dominant
Current
dominant
energy
dominant
energy
use energy
use use
Propagation
Propagation
Propagation
method
Propagation
method
Propagation
method
Propagation
method
method
method
E

DE

Bioethanol
Bioethanol
Biodiesel
Bioethanol
Biodiesel
HeatBiodiesel
andHeat
power
and
Heat
power
Biogas
and power
Biogas Biogas
Seed Seed Seed
Stem Seed
Stem Rhizome
Seed
Stem Seed
Stem
StemMicropropagation
StemMicropropagation
Rhizome
Micropropagation
Rhizome
Rhizome
Rhizome
Micropropagation
Rhizome
Micropropagation
Micropropagation
cutting cutting
cutting
cutting
cutting
cutting
or root
cuttings
or root
cuttings
or root
cuttings
or root
cuttings
or root cuttings
or root cuttings
Other OtherOther

Miscanthus average annual yield per hectare

Current
Current
dominant
Current
dominant
Current
energy
dominant
Current
energy
use
dominant
Current
energy
dominant
use energy
dominant
use energy
use energy
use use

Litres of bioethanol fuel

How long could


you supply energy
Bioethanol
Bioethanol
Biodiesel
Bioethanol
Biodiesel
Bioethanol
HeatBiodiesel
Bioethanol
andHeat
power
Biodiesel
Bioethanol
andHeat
power
Biogas
Biodiesel
andHeat
power
Biogas
Biodiesel
and
Heat
power
Biogas
andHeat
power
Biogas
and power
Biogas Biogas
for a house?
Effective energy
D E DE DE D
D
available

Car

Car Weight
Car Weight Barrel
Weight Barrel Barrel

DE

Other Other Other OtherOther Other

Europe

5,600 litres Europe

gasoline equivalent of 3,700 litres

Car

Europe

Car Weight
Car Weight
CarBarrel
Weight
CarBarrel
Weight
Car Barrel
Weight Barrel
Weight Barrel Barrel
Power usage
Power usage
Power usage

27,000kWh

Midwest winter US

16

months

Midwest winter US

34

Power usage
Power usage
Power usage
Power usage
Power usage
Power usage

57,000kWh

12,000 litres Midwest winter US

months

gasoline equivalent of 7,900 litres

What distance could you drive?

Dry weight in tonnes


18 tonnes
Europe

47,000km
29,000 miles

Europe
Midwest winter US

38 tonnes
Midwest
winter US

82 | 6 Biomass feedstock crops

10

20

30

99,000km
61,000 miles

40 50 60 70 80 90 100
Thousand km

Harvesting

Invasion risk

Annual harvesting starts in year two or three, with


biomass increasing by 80% from the first to the second
year after planting. The crop is usually harvested in late
winter to early spring after the woody stems have dried
and before new shoots appear. This maximizes the amount
of nutrients returned to the soil but reduces the overall
yield by 30 50%, as biomass is lost to litterfall during
winter. The grass is cut and baled, or collected with a
forage harvester.

Miscanthus x giganteus itself has a very low risk of


invasion because it is a sterile hybrid that does not
produce seed. This biomass variety is a relative of several
species, such as Miscanthus sinensis, that are considered
highly invasive.

Key references

Clifton-Brown, J. et al. (2001), Performance of


15 miscanthus genotypes at five sites in Europe,
Agronomy Journal, vol. 93, pp. 1013 1019.
Yield
Davis, S. C. et al. (2012), Impact of second-generation
Plant types
Plantintypes
biofuel agriculture
on greenhouse-gas emissions
the
As the potential of miscanthus as a lignocellulosic
Annual
corn-growing regions of the US, Frontiers in Ecology
feedstock is still being investigated, there are no reported
and the Environment, vol. 10, pp. 69 74.
data on this crop from agencies such as the FAO. Most of
the yield data are sourced from scientific investigations.
Heaton, E. A. et al. (2008), Meeting US biofuel goals
Perennial
Grain or seed
Woody G
Herbaceous
Herbaceous
Yields per hectare are impressive, reaching well over
with less land:
the potentialWoody
of miscanthus,
Global
40 tonnes (dry weight) per hectare at best. Averages depend
Change Biology, vol. 14, no. 9, pp. 2000 2014.
Photosynthetic pathway
Photosynthetic pathw
on time of harvest, as mentioned above. Autumn (fall)
Lewandowski, I. & Clifton-Brown, J. (2000),
yields an average of 25 tonnes/ha in Europe and
Miscanthus: European experience with a novel energy
C3
C 4vol. 19, pp. 209 227.
C4
47 tonnes/ha in the US Midwest. Winter yields are lower
crop, Biomass andC 3Bioenergy,
Plant
types
(18 in Europe and 38 in the Midwest).
FP7 OPTIMISC (2012), Uses of miscanthus.
Annual
C4
C4
C3
C3
In terms of thermal energy, miscanthus has a calorific
value of 18GJ/tonne.
Perennial
This box highlights another high yielding
Research into the conversion of miscanthus to ethanol
Woody
Grain or seed
Herbaceous
perennial
grass.
has a reported conversion efficiency of 310 litres/tonne.
Propagation method
Propagation method
Photosynthetic pathway

Alternative markets

Plant
C3
typesto
C 4 types
Miscanthus is becoming more widely usedPlant
in Europe
Annual
Annual
augment coal in power stations and thus reduce GHG
Seed
Stem
Stem
RhizomeSeedMicropropagation
C3
C4
emissions. It can also be used for pressed particle board.
cutting
cutting
or root cuttings
Perennial
Perennial
In Asia it is a traditional roofing material, and
in Europe Woody
is
Woody Grain or seed
Herbaceous
Grain or seed
Herbaceous
Current dominant energy useCurrent dominant ene
popular as horse bedding.
PropagationPhotosynthetic
method
pathway
Photosynthetic pathway
Plant types

Co-products
From the crop
Currently none.

Annual

C3

C4

C3

C4

C3

C4

Perennial

Perennial
CGrain
orBiodiesel
seed Micropropagation
Herbaceous CStem
3 Woody
4Rhizome
Seed
Bioethanol
Heat and power
Bioethanol
Biogas
Energy cane
(Saccharum
officinarum variety bred for
cutting
or root cuttings

From conversion to fuel


Photosynthetic
pathway
fibre instead of Other
sugar)
Other
Wastes from ethanol production could be burned as fuel Current dominant energy use
Propagation
method
Photo
in
production:
Florida,
US.
Propagation
method
for heat and/or power.
C3
C4
Why is it similar? Also a high yielding C4 grass with
low
nutrientCrequirements.
C3
4
Car
Weight
Barrel
in warmer,
E WhatDmakes it different? Grows
subtropical
climates,
south
of
the
latitudes
where
Seed
StemBiogas Rhizome
Micropropagation
Bioethanol
Biodiesel
Heat
and power
Seed
Stem
Rhizome
Micropropagation
Miscanthus
is most productive.
or root cuttings
cutting Propagation
or root cuttings
method cutting
Other Growing region: Currently there is very limited
Current production,
dominant energy
use
commercial
but plantations
are
Current dominant energy
use
developing in the southern US.
E
Bioethanol

Stem
Rhizome
Micropropagation
Weight
Barrel
D
cutting
or root cuttings
Bioethanol
Biodiesel Heat and power
Biogas Biodiesel Heat and power Biogas

E
Car

Car

Seed
Car E

Current dominant energy use


Other

Other

Biodiesel He

Bioethanol
Weight

Other

Power usage

D
Car
WeightBiogas Barrel
Biodiesel
Heat and power
Barrel
Power usage

Car

Weight

Barrel

6 Biomass feedstock crops | 83


Power usage

Power usage

Power

Photosynthetic pathway

es

C3

Plant
C 4 types

C3

Annual
C4

Oil palm

Propagation method

Annual
Seed

Perennial and E. oleifera


Elaeis guineensis
Woody Grain or seed
Herbaceous

Woody

Grain or seed

Stem
cutting

Rhizome
Micropropagation
Perennial
or root
cuttings

Current dominant energy use


Propagation
method
Photosynthetic
pathway

nthetic
pathway
s
Plant
types

Annual

Annual

C3

C4
Woody
Grain or seed
Woody
Herbaceous
C4

Perennial
C3 Stem
GrainSeed
or seed
cutting

C4
Perennial

Perennial
C4 Bioethanol
Rhizome
Micropropagation
Biodiesel Heat and power
or root cuttings

Biogas

Car

Barrel

of fuel: biodiesel.
hetic Photosynthetic
pathway Type
pathway
Other
Stage ofCurrent
adoption:
extensive
commercial
dominant
energy
use
Propagation
method
tion method
production.
C3
C
C
4

C4

C3

C4

Weight

Seed Heat and power


Stem Biogas Rhizome
Micropropagation
Biodiesel
RhizomeBioethanol
Micropropagation
cutting
or root cuttings

Stem

cutting
or method
root cuttings
on method
Propagation
Other
dominant energy use

Current dominant energy use

and, even though the oil palm acts as a carbon sink, it can
Car Micropropagation
Weight
Barrel
Stem Micropropagation
Rhizome
Rhizome
E
D
take 40 years or more to achieve carbon payback.
cutting
or root cuttings
or root cuttings
If oil palm does not displace native forests, plantations
Bioethanol
Biodiesel
Heat
and
power
Biogas
Power
usage
Biodiesel Heat and power Biogas
StemSeed

Dcutting

ominant
Current
energy
dominant
use
energy use

can result in a net carbon sink and efforts are under way
Other
to establish guidelines for more sustainable growing
Although oil palm originated in West Africa and tropical
techniques. For example, there is a Roundtable on
parts of Central and South America, the biggest oil palm
Sustainable Palm Oil that has developed a certification
D
E plantations
D are now in Indonesia and Malaysia. Palm oil
system for sustainable oil palm producers.
was
traditionally
used
for cooking
making soap,
Weight but Barrel
Biodiesel
Bioethanol
Heat
and
power
BiodieselBiogas
Heat and
power
Biogas andCar
Car
Weight
Barrel
it has become an important raw material in processed
Plant characteristics
Power
usage
Other food. Biodiesel from palm oil has
been
in production for a
quarter of a century.
Oil palms have stout single trunks and leaves that can
Oil palm now produces 32% of world vegetable oil,
reach 5 metres in length. Bunches of fruit are formed at
despite covering only 4% of land area devoted to oil crops.
the top of the trunk and it is the fruit pulp that provides
Car
Weight
Car Barrel
Weight controversy
Barrel for decades.
The crop
has raised
palm oil. Palm trees bear fruits starting in the fourth year.
Producers love its high yields but critics point to serious
Selective breeding has created a hybrid (the Tenera variety)
environmental damage caused by large-scale plantations
with a particularly high percentage of oil in the fruit for
that displace native rainforest. Changing land
use
from
commercial production. A complete breeding programme
Power
usage
Power usage
forest or peatland releases large amounts of carbon dioxide
using controlled pollination can take eight to 10 years.
World map with latitude limits for growth and five top producing countries
Power usage

Power usage

Thailand
Colombia

Top producing
countries

Million
tonnes*

Indonesia
Malaysia
Nigeria
Thailand
Colombia

90
88
9
8
3

15N *fruit

Malaysia

Nigeria
Indonesia

12S

Global planting 15 million hectares.


Global production 220 million tonnes of palm fruit, 44 million tonnes of oil.

84 | 6 Biomass feedstock crops

Where to grow it

Pests and diseases

Oil palm is a tropical crop that needs constant warmth


and humidity. Optimum temperatures are in the range
of 24 28C and commercial production is not possible if
seasonal mean temperatures vary by more than 6C.
It grows best in well-drained sandy soils and is limited
to low altitudes and a fairly narrow equatorial belt
(15 North to 12 South) that corresponds to the tropical
humid rainforest biome.

The heat and humidity typically needed by oil palm


are also ideal conditions for plant diseases to flourish.
Decaying palm fronds left to protect the soil and replace
some nutrients provide shelter to a range of pests. Not
surprisingly therefore, oil palms are prey to a range of
diseases including trunk rot (Ganoderma sp.) and dry
basal rot (Ceratocystis sp.), which can reduce yields by
20 60%. Grasshoppers can defoliate young palms or
severely depress yield by up to 40% in older plantations.
Some of the worst pest attacks are thought to have been
exacerbated by overuse of broad-spectrum pesticides that
killed beneficial insects along with problematic species. In
many plantations owls are kept to control mice that can
damage the crop.

How to grow it
Palm trees can be productive for many years, but they are
typically replanted after about 20 years when they become
too tall for efficient harvesting. Ground for planting may
be cleared mechanically or by burning, followed by tillage
or herbicide treatment. Leguminous ground cover is often
established before the young palm trees are planted, with
the aim of preventing erosion and increasing the nitrogen
content of the soil. Palm leaves are also left between the
rows of trees for the same reasons.

Inputs required
Water
Oil palm needs a constant water supply and high humidity
for optimum production with daily evapotranspiration
rate of 5 6mm of water. While research shows that
irrigation in dry climates can increase yields by
approximately 36%, irrigation is unnecessary in the very
wet climates where oil palm is often cultivated.
Fertilizer
High levels of fertilizer, particularly nitrogen and
potassium, are required throughout the life of an oil palm.
Plantations on peaty soil are especially prone to potassium
deficiency.

Oil palm average annual yield


per hectare

25

Fruit yield in tonnes

Oil yield in tonnes

14 tonnes

2.9 tonnes

What distance could you drive?


46,000km
29,000 miles

World
Best recorded
0

10

20

30

75,000km
46,000 miles

40 50 60 70 80 90 100
Thousand km

Average yield (tonnes/ha oil palm fruit)

mineral diesel equivalent of 2,800 litres

Herbicides
The use of herbicides can be reduced by using leguminous
ground cover to suppress weeds between young palm trees.
Herbicides are still needed before planting, and paraquat is
commonly used, despite its toxicity to humans.
Pesticides
Broad range pesticides are not recommended because they
can exacerbate a pest outbreak. A range of pesticides has
been used, but some are very toxic.

Harvesting
The heavy bunches of fruit are cut by hand, sometimes
using a chainsaw. This requires strength and dexterity,
as the leaf fronds around the fruit are armed with tough
spikes. Removed leaves are left on the ground. Bunches of
fruit are then taken to a processing mill for stripping.

Oil palm yields global average


and top producing countries

Litres of biodiesel fuel

3,000 litres

Defences

Global average yield


Columbia
Indonesia

20

Malaysia
Nigeria
Thailand

15

10

0
1960

1970

1980

1990

2000

2010

6 Biomass feedstock crops | 85

Plant types

Yield
The fresh-fruit yield from the palms average
14 tonnes/ha, with an average oil yield of 2.9 tonnes/ha
per year. Maximum yields have been reported in Malaysia
(with up to 23 tonnes fruit and 4.5 tonnes oil
pertypes
hectare)
Plant
and Colombia (with up to 19 tonnes fruit and 4.7 tonnes
oil per hectare).

Alternative markets

Herbaceous

Woody

C3

C4

C3

C4

This box highlights another woody plant that


Woody Grain or seed
Herbaceous
is grown for
oil.

Annual
Perennial

Photosynthetic pathway Propagation method


C3
C3

Plant
C 4 types
Annual
C4
Perennial
Herbaceous

Grain or seed

Seed
Woody

Stem
cutting
Grain or seed

Rhizo
or root c

Current dominant energy use


PropagationPhotosynthetic
method
pathway
Palm oil is processed and refined to produce
a range
Photosynthetic
pathway
Plant
typesof
Plant
types
Annual
Annual
products, many of them used in food production and
C4
CAM C 3
Perennial
as cooking fats and cosmetics. Fruit bunches must
C 3 be
C4
E
D
sterilized and then threshed to separate the palm fruit,
Perennial
Perennial
C
C
Woody
Grain
or
seed
Woody
Grain
or
seed
Herbaceous
Herbaceous
3
4
Seed
Stem
Rhizome
Micropropagation
C
C
which is then pressed to extract the crude oil. Several
Bioethanol
Biodiesel Heat and power
3
4
or root cuttings
Jatropha (Jatropha curcas) cutting
refining stages are then required to produce oil suitable
Photosynthetic pathway
Photosynthetic pathway
Other
for further processing. The kernels are separated from the
Photo in Current
production:
Indonesia.
dominant
energy use
Propagation
method
fruit bunches, and are often processed separately
in
a
mill
Propagation
method
C3
C 3 Why Cis4 it similar? Also a woody plant that grows as
C4
with other oilseeds.
a shrub or tree with oil-rich seeded fruits (inedible
in this
C4
C4 case); grows in tropical regions of the world.
C3
C3
Car
Weight
E
D
Co-products
What makes it different? Not a food commodity
because
the
products
are toxic;
inStem
very
dry Rhizome Micropr
Seed
Bioethanol
Biodiesel
Heatgrows
and power
Biogas
Seed
Stem
Rhizome
Micropropagation
From the crop
or root cuttings
tropical
conditions as well as the moistcutting
tropics and
cutting
or
root
cuttings
Propagation
method
Propagation
method
Plant biomass remaining after oil extraction can be used
Other than palm oil. Without irrigation,
is lower yielding
to make paper, fibreboard or used as solid fuel.
Current dominant energy use
Current dominant energy
use of the oil-rich seeds
yields
are little more than
From conversion to fuel
2 tonnes/ha in well-established plantations. With
irrigation and fertilizer, this figure increases to
Residues such as the sludge from milling (decanter cake)
Car
Weight
Barrel
Seed
Stem Micropropagation
Rhizome
Rhizome
or more.
Jatropha
usesDboth the C3
E Micropropagation
and empty fruit bunches have high nutrient content
and D StemSeed 12 tonnes/ha
E
cutting
cutting
or root cuttings
or root cuttings
and CAM photosynthetic pathways, depending on
make good fertilizer. Glycerol can be used in cosmetics.
Bioethanol
Biodiesel Heat and power
Power Biogas
usage
Bioethanol
Biodiesel Heat and
power Biogas
water
Palm oil mill effluents are often used for biogas
Current dominant
Current
energydominant
use availability.
energy use
Growing region: Although
production.
Otherthis plant grows
Other
throughout the tropical and subtropical regions of
the world and is considered invasive in some regions,
Invasion risk
E
D
E
D
commercial
production is very limited. Small-scale
Oil palm is generally not considered an invasive species
Weight
Barrel
plantations
have
in Car
Africa and
India.
Bioethanol
Biodiesel
HeatWeight
and power
BiodieselBiogas
Heat
andbeen
powerdeveloped
Biogas
CarBioethanol
Barrel
and typically requires a large amount of plantation
It has also been introduced in South-East Asia with
maintenance to prevent weed competition.Other
Elaeis
Power
usage
Other differing levels of production
efficiency,
and work
guineensis (African oil palm), however, has been observed
continues to improve the genetics and agronomy of
to be potentially invasive in Micronesia.
the crop around the world.

Key references

Car

Weight

Corley, R. H. V. & Tinker, P. B. (2003), The oil palm, 3rd


edition, Blackwell Publishing Ltd, Oxford, UK.
Lim, S. & Teong, L. K. (2010), Recent trends,
Power usage
opportunities and challenges of biodiesel in Malaysia:
an overview, Renewable and Sustainable Energy
Reviews, vol. 14, pp. 938 954.
Yee, K. F. et al. (2009), Life cycle assessment of palm
biodiesel: revealing facts and benefits for sustainability,
Power usage
Applied Energy, vol. 86, pp. S189 S196.

86 | 6 Biomass feedstock crops

Car Barrel Weight

Barrel

Power usage

Power usage

Photosynthetic pathway
C3

es

Propagation method

Plant
C 4 types

Soybean (soyabean, Ta dou, daidzu)


Annual
C4

C3

Woody

Glycine max Perennial


Herbaceous
Grain or seed

Annual

Seed

Woody

Stem
cutting
Grain or seed

Rhizome
Micropropagation
or rootPerennial
cuttings

Current dominant energy use


PropagationPhotosynthetic
method
pathway
thetic pathway Plant types
Annual

Annual

C3

C4
Woody
C4

Grain orHerbaceous
seed

Perennial
Woody
Grain Cor3Stem
seed
Seed
cutting

C4
Annual

Perennial
C4 Rhizome
Micropropagation
Bioethanol
Biodiesel
Heat and power
or root cuttings

Biogas

Type of fuel: biodiesel.


hetic pathway Photosynthetic
pathway
Other
Stage of Current
adoption:
extensive
commercial
dominant
energy
use
Propagation
method
ion method
production.
C3
C4
C4
C4

C3
Stem

C4

Car

SeedHeat and power


Stem Biogas Rhizome
Biodiesel
RhizomeBioethanol
Micropropagation

Weight

Barrel

Micropropagation

cutting
or root cuttings
or root cuttings method
Propagation
Other
There is broad global knowledge of soybean agriculture
Current dominant energy use
dominant energy use
because of the long history in cultivation. It has a low
yield per unit area, and requires a larger land footprint than
many other bioenergy crops, but it produces high amounts
Weight
Barrel
Stem
Seed Micropropagation
Stem Car
Rhizome
of protein, significant amounts of oil, and the conversion
E Rhizome D Micropropagation
Dcutting
cutting
or root cuttings
or root cuttings
of the oil to biodiesel produces little waste.

cutting
on method

Biodiesel Heat and power

Biogas

Bioethanol

Biodiesel Heat and power


Biogas
Power
usage

ominant energy use


Current dominant energy use
Plant characteristics
Other
Soybean was domesticated in China references
Soybean is an annual leguminous plant with
to shu appear in ancient literature dating back
clover-like leaves that uses the C3 photosynthetic
D
E and its
D cultivation spread to Korea, Japan
4,500 years
pathway. The beans are produced in short, slightly furry
and
other
parts of
South-East Asia roughly
years ago. Barrel
Car 3,000 Weight
Biodiesel
and Bioethanol
power Biogas
Biodiesel Heat and power Biogas
Car Heat Weight
Barrel
pods. Its long domestication is reflected in the 45,000
The domesticated plant is shorter than its wild relatives
Asian varieties (landraces) stored in international
and has
much larger seed. MorePower
recently
usage the plant has
Other
seedbanks there are 23,000 varieties in the Chinese
been grown for high-protein animal feed, with oil as a
gene bank alone. Breeding now involves genetic
co-product. Seventy-five per cent of all cultivated soybean
modification, and glyphosate-resistant soybean is widely
is GM, primarily to impart herbicide resistance.
grown. Modifications have also been made to increase
Car
Weight
Barrel Car
Weight
Barrel
the production of particular fatty acids (e.g. the monounsaturated 18-carbon oleic acid) in the bean.
Power usage

Power usage

World map with latitude limits for growth and five top producing countries
Power usage

Power usage

52N
US

China
India

Top producing
countries

Million
tonnes*

US
Brazil
Argentina
China
India

91
69
53
15
13

*seed

Brazil

Argentina

39S

Global planting 99 million hectares.


Global production 240 million tonnes seed.

6 Biomass feedstock crops | 87

Where to grow it

Pests and diseases

Considered a warm-season crop, soybean grows between


10C and 40C, at altitudes below 1,000 metres from
52 North to 39 South. It is widely planted across the
subtropical humid forest, tropical humid rainforest and all
temperate biomes. It needs a free-draining soil and is most
productive on clay loam. Soybean is moderately salt-tolerant.

Soybean is prey to a host of serious pests and


diseases. They include a pod borer specific to soybean
(Leguminivora glycinivorella 20 50% yield losses), a
late-season fungal disease (Macropomina phaseolina up
to 70% yield losses) that can destroy more than two-thirds
of the crop, and soybean rust (Phakopsora meibomiae
and P. pachyrhizi) able to cause up to 80% loss in yield.
Soybean cyst nematode (Heterodera glycines) occurs
widely in regions where the crop is grown and causes
considerable losses.

How to grow it
Soybean is an annual crop, grown from seed planted in early
to late spring. Although it is often grown as a monoculture,
yields are better when grown in a two-year rotation with
maize. Zero-tillage (direct-drilled) systems also give
better long-term results than traditional land preparation
(ploughing and harrowing to make a seedbed).

Inputs required

Defences
Herbicides
The advantage of glyphosate-resistant soybean is
that glyphosate can be used to control weeds without
damaging the crop. Because glyphosate is broken down
on contact with soil, this makes it a better choice from an
environmental perspective than more persistent or more
toxic herbicides (also used in soybean cultivation). Choice
of herbicide and concentration should be tailored to the
type of soil, climatic conditions and developmental stage
of the plants.

Water
Soybean has moderate water requirements average
irrigation of 140mm (it loses roughly 600mm in
evapotranspiration in a growing season) and increases in
yield as a result of irrigation have been documented.
Water requirements depend on soil, rainfall and temperature,
Pesticides
and should be calculated for each site and meteorological
conditions.
Various pesticides are available, with control regimes
varying regionally. Estimates of attainable yield protection
Fertilizer
range from 25% in Central Africa to 43% in southern
Being a legume, soybean has nitrogen-fixing bacteria in
Europe.
root nodules, although soybean uses more nitrogen than it
can acquire through this symbiotic relationship. Nitrogen
Harvesting
fixation is inhibited when nitrogen fertilizer is used, and
early studies showed no increased yield from applying
Most soybean is harvested with combines once the crop
nitrogen. But more recent studies suggest that modest
has reached a moisture content of 13 15%. Delaying the
applications of nitrogen (less than 50kg per hectare) can
harvest after this time is detrimental, as overripe pods
increase yield by 0.6 tonnes/ha. Applications of phosphorus shed their beans and over-dried beans (less than 12%
and potassium are also required. Matching fertilizer
moisture) tend to shatter during harvest. Losses after the
requirements to each specific site reduces GHG emissions.
optimum date can amount to 11kg per hectare per day.

Soybean average annual yield


per hectare

Soybean yields global average


and top producing countries
3

480 litres

mineral diesel equivalent of 430 litres

Grain yield
in tonnes

Oil yield
in tonnes

2.4 tonnes

0.44 tonnes

Average yield (tonnes/ha seed)

Litres of biodiesel fuel

Global average yield


Argentina
Brazil

China
India
US

What distance could you drive?


World
7,200km
4,500 miles
Best recorded 9.000km
5,600 miles
0

10

20

30

40 50 60 70 80 90 100
Thousand km

88 | 6 Biomass feedstock crops

0
1960

1970

1980

1990

2000

2010

Plant types

Yield
The yield of seeds on average varies between 1.5 and
3 tonnes/ha with a worldwide average of 2.4 tonnes/ha.
The oil yield represents between 17% and 20% of the seed
mass, resulting in 0.41 0.48 tonnes of oil per hectare
Plant types
per year.

C3

C4

C3

C4

Annual

This box highlights another important oil crop that is Perennial


Woody Grain or seed
Herbaceous
produced commercially.
Propagation method
Photosynthetic pathway

Alternative markets

C3

C 4 types
Plant

C3

Annual
C4

Seed

Perennial

Stem
cutting
Grain or seed

A
Rhizom
or root Pe
cu

Woody
Herbaceous
Woody Grain or seed
Herbaceous
The main products from soybean are the whole
beans, the
Propagation method Current dominant energy use
protein-containing defatted soybean meal, and oil that
Photosynthetic pathway
Photosynthetic
pathway Plant types
Plant
types
are used for food and feed. Many processed foods include
Annual
Annual
soya products, and soybean meal is vital feedstuff for meat
Annual
C3
C4
C3
C4
production.
E
D
Perennial
Perennial

Co-products

Herbaceous
C3

Woody
C4

Grain orHerbaceous
seed

Woody
Seed

Grain
seed
Stem
Micropropagation
C3or
C4 Rhizome Biodiesel
Bioethanol
Heat and power
cutting
or root cuttings

Oilseed rape or canola (Brassica napus)


Photosynthetic pathway Photosynthetic pathway
Other
From the crop
Photo in production:
Canada.
Current dominant
energy
use
Propagation
method
Propagation
method
Why is
it
similar?
Also
an
oil-rich
C3
plant, grown
During soybean processing, the defatted soybean
meal,
the
C
C
C3
C4
3
4
for both food and biodiesel, and although the
major product (about two-thirds) from soybean processing,
highest
crop in temperate regions,
is primarily used as animal feed, with a small portion
as C4
C3 yieldingCoil
C3
4
E yielding
Drelative to other biomass
Car
Weight
rapeseed is low
protein ingredients for the food industry. The soybean oil
feedstocks.
Bioethanol
Biodiesel
Heat and power
Biogas Rhizome
can also be used in various other biomaterials, including
Seed
Stem
Microprop
Seed
Stem
Rhizome
Micropropagation
different? Much greater
cold
cutting
or root cuttings
printing ink, candles and waxes, low-temperature
cutting What
or
rootmakes
cuttings itmethod
Propagation
Propagation method
tolerance;Other
salt tolerance; greater fertilization rates
lubricants, industrial cleaners and metalworking fluids,
Current
dominant nitrogenenergy use
Current
dominant
energy
use
recommended;
does not
host symbiotic
lotions and paint additives.
fixing bacteria like soybean.
From conversion to fuel
Growing region: Canada is the largest producer,
Car
Weight Micropropagation
Barrel
Seedby China,
Stem
Rhizome
The glycerol that is a co-product of biodiesel conversion
Seed
Stem
Rhizome
Micropropagation
followed
India, Germany
and France.
D
E
D
cutting E or root cuttings
or root cuttings
can replace glycerol made from petroleum. Glycerol has cutting
Together, more than 15% of the worlds vegetable
Bioethanol
Biodiesel Heat and power Biogas
many uses in cosmetics, food, oral care, tobacco,
synthetic
Bioethanol
Biodiesel Heat and power
Biogasby canola.
is supplied
Current
dominant energy use
Current dominant energyoil
use
Power usage
polymer production and synthetic resins.
Other
Other

Invasion risk
E
D
E
D
Soybean is typically considered non-invasive, but
Biodiesel Heat and power Car
Biogas
Biodiesel
HeatWeight
and Bioethanol
power Biogas
cultivated soybean can hybridize with wildBioethanol
strains. The
Car
Barrel
use of specific no-planting zones is a simple and effective
Power usage
Other
way to avoid transgene dispersal in the caseOther
of GM
soybean.

Key references

Car

Weight

Hymowitz, T. & Newell, C. (1981), Taxonomy of the


genus glycine, domestication and uses of soybeans,
Economic Botany, vol. 35, pp. 272288.
Singh, G. (ed.) (2010), The soybean: botany, production Power usage
and uses. CAB International, London, UK.
Kim, S. & Dale, B. E. (2009), Regional variations in
greenhouse gas emissions of biobased products in the
United Statescorn-based ethanol and soybean oil, The
International Journal of Life Cycle Assessment, vol. 14, Power usage
pp. 540546.

Barrel Car

Weight

Weight

Barrel

Power usage

Power usage

6 Biomass feedstock crops | 89

Barrel

Propagation method

Propagation method

Photosynthetic pathway

C3

C4

C3

C4

Plant types

Willow and
woody species)
Seedhybrid
Stem poplar
Rhizome (Short-rotation
Micropropagation
Seed
Stem
Rhizome
Micropropagation

Woody

Annual

Annual

cutting

Perennial spp
Salix and Populus
Woody
Herbaceous

Grain or seed

or root cuttings

Grain or seed

cutting
Perennial

or root cuttings

Current dominant energy useCurrent dominant energy use

Propagation method Photosynthetic pathway


thetic
pathway
Plant
types
Annual

C3

C4

Perennial
C4

Perennial
C4
Grain
or seed C3Rhizome
Bioethanol
Stem
Micropropagation
Bioethanol
Biodiesel
Heat and power
Biogas Biodiesel
or combustion
root cuttings
Type
ofcutting
fuel: Solid
for heat and
Photosynthetic
pathway
Woody
Herbaceous
C4
Seed

Other
Other biofuels.
electricity, cellulosic
Current dominant energy
use
Propagation
method
ion method
CStage of adoption: developmental stage for liquid
C3
4
biofuel, extensive commercial use for heat and
Celectricity.
C3
4
E
D
Car
Weight
Barrel
Stem
Bioethanol

Seed Biogas Stem


Rhizome
Micropropagation
Biodiesel Heat
and power
cutting

cutting
or root cuttings
Propagation
method
Other
dominant energy use

Car

Heat and power

Weight

Biogas

Barrel

Rhizome
Micropropagation
to 11 eleven years (called short-rotation forestry or SRF).
or root cuttings

Woody crops grown in this way are already widely used as


fuels for heating and in thermal power stations.

Current dominant energy use

Plant characteristics
D

Seed

Stem
Car
cutting

Biodiesel Heat and power

Rhizome
Micropropagation
Weight
BarrelD
E
or root cuttings
Biodiesel Heat and power
Biogas Bioethanol

Current dominant energy use


Other

Biogas

Power usage

Willow
Hybrid poplar
D
The graceful form of weeping willow trees reflected in
Car marks
Weight
Barrel
Bioethanol
Biodiesel
and power Biogas
Car
Weight
Barrel
water
and Heat
unmistakable
exclamation
of Lombardy
poplars are familiar sights in many temperate landscapes.
Other
In fact, thesePower
treesusage
have diverse forms, with the smallest
willow being a creeping arctic species. But most share the
useful attribute of growing readily from cuttings which
merely
pushingBarrel
a length of stem into the ground.
Car requires
Weight
This ease of propagation and fast growth make the trees
invaluable producers of biomass, especially in soils that do
not dry out. Moreover, they can be grown in rapid cycles in
usage
roughly
every three to
Powerwhich
usage they are cut back or coppiced Power
five years (called short-rotation coppice or SRC) or seven
E

There are 300 500 willow species that have a wide range
of genetic variability. There are fewer poplar species but an
estimated
elite poplar cultivars. Most species are trees
Power125
usage
or large shrubs with questing roots, although when grown
as coppice most roots are found within the first 30cm of
the soil profile. Willows show great diversity of form,
but the shrub willows (roughly 2 3 metres tall) are most
suitable for biomass.
Native poplar grows as a single-trunked deciduous
tree with a wide variety of growth rates, heights and
leaf structures among species. Some species of poplar,
however, can also be grown as short-rotation coppice.
Breeding has focused on crossing species to give hybrids of
increased height and vigour, and there are many varieties
available commercially.

World map with latitude limits for growth and five top producing countries
Power usage

75N
Sweden

France
Italy

Top producing Thousand


countries
hectares
China
India
France
Turkey
Italy

Romania
Turkey

Poplar

China
India

4,900
1,000
240
130
120

Willow
Top producing Thousand
countries
hectares
New Zealand
Argentina

Key
nPoplar
nWillow

90 | 6 Biomass feedstock crops

34S

Argentina
Romania
New Zealand
Sweden

46
24
20
15

Where to grow it

minimum of 600mm rainfall annually. Although these


crops are tolerant of wet sites, the ground must be firm
enough at times to allow the use of planting and harvesting
machinery.

Willow and poplar are native to temperate broadleaf and


conifer biomes in the northern hemisphere. Poplar grows
from 66 North to as far south as 30 North; willow can
grow from 75 North to 34 South. They grow best where
there is regular rainfall and in water-retentive soils; they
are slow to establish on heavy clay. Some varieties can
withstand extremes of heat and cold (4050C), and all are
frost-hardy when dormant over winter.

Fertilizer
Fertilizer use in commercial plantations is dictated by
deficiencies noted in the leaves. The removal of large
quantities of biomass over several years depletes the soil
of nitrogen, so nitrogen is added as necessary, with a typical
range of 20 80kg N per hectare added to willow (usually
after harvest) and up to 200kg N per hectare to poplar
annually when nutrient depletion is evident.

How to grow it
Cultivation techniques exploit the ability of many
broadleaved tree species to regrow (coppice) strongly when
all their aerial growth is removed. The cut stumps quickly
produce vigorous new shoots, nourished by the established
root system, and multiple harvests can be taken from a
single planting.
Poplar and willow can be established mechanically, by
planting rods cut from year-old shoots into clean ground
in the spring. For good establishment, the soil should
be ploughed and harrowed before planting. Plantations
can remain productive for more than two decades. Crops
are grown both as monocultures and in polycultures
with multiple varieties. The benefits of mixed planting
(reducing the spread of pests and diseases) are sometimes
outweighed by increased costs of managing and harvesting
a non-uniform crop.

Pests and diseases


Various leaf and stem diseases, particularly rusts but also
spots and cankers, can decrease growth rate and yield from
10% to 100% depending on the conditions. Stem borers
and beetles are also prevalent, causing 50 100% damage
to populations. Pests and diseases are generally more
serious in young crops. Weeds compete with newly planted
cuttings for light and nutrients so weed suppression is
necessary. Control may also be needed after coppicing,
when increased light levels lead to a spurt in weed growth.

Defences
Herbicides
Weed control in early establishment is essential: a range of
herbicides is used to clear the ground during establishment
in the first year and after coppicing. An alternative to
initial herbicide use is to plant cuttings through weedsuppressant horticultural membrane.

Inputs required

Water
The availability of water is more important than soil type
when selecting growing sites. Willow and poplar are both
Pesticides
Plant types
Plant types
Plant types
thirsty crops, often
needing
more than 100kg of water Plant
pertypes
Annual
Annual
Annual Annual
Some resistance has been bred into certain varieties
day during the growing
season. Poplar water use ranges
Perennial
Perennial
Perennial
Perennial
of Woody
willow
Woody
Woody
Grain50
orGrain
seed
Woody
Grain orGrain
seed orand
seed poplar. There is no standard pesticide
Herbaceous
Herbaceous
Herbaceous
Herbaceous
from
toor seed
110kg per day; willow species use about 105kg
treatment,
but a range of pesticides of varying
Photosynthetic
Photosynthetic
pathway
pathway
Photosynthetic
pathway
pathway
per
day,
although wide genetic variation for water use Photosynthetic
environmental
toxicity can be used to control insect
C
C
C
C
Cis found
C
C
C
in
both
trees.
Generally,
both
species
require
a
Plant types
Plant types
Plant types
Plant types
Plant types
Plant types
Plant types
Plant types
Plant types
damage
when infestations
arise.
Annual Annual Annual Annual Annual
Annual Annual Annual
Annual
C
C
C
C
C
C
C
C
3

Perennial
Perennial
Perennial
Perennial
Perennial
WoodyHerbaceous
Woody
GrainHerbaceous
Woody
orGrain
seed Woody
or Grain
seed Woody
orGrain
seed or Grain
seed or seed
Herbaceous
Herbaceous
Herbaceous

Poplar average annual yield per hectare

Propagation
Propagation
method
method
Photosynthetic
Photosynthetic
Photosynthetic
pathway
Photosynthetic
pathway
Photosynthetic
pathway
pathway
pathway
C3

C3

C4 C3

C 4C 3

C 4C 3

C4

Dry weight in tonnes

C3

C4 C3

C 4C 3

C4

C4

Seed Seed Stem Stem RhizomeRhizome


Micropropagation
Micropropagation
C3
C3
C4 C3 C4C3
C
C
cutting cutting
or4 root cuttings
or4 root cuttings

Dry weight in tonnes

Current
Current
dominant
dominant
energyenergy
use use
Propagation
Propagation
Propagation
method
Propagation
method
Propagation
method
method
method
E

Willow average annual yield per hectare

C3

C4

Seed Seed Stem Stem RhizomeRhizome


Micropropagation
Micropropagation
C3
C3
C4 C3 C4C3
C C 3 C4
C4
cutting cutting
or4 root
cuttings
or root cuttings

Perennial
Perennial
Perennial
Perennial
WoodyHerbaceous
Woody
Grain Woody
orGrain
seed Woody
or Grain
seed orGrain
seed or seed
Herbaceous
Herbaceous
Herbaceous

Propagation
Propagation
method
method
Photosynthetic
Photosynthetic
Photosynthetic
pathway
Photosynthetic
pathway
pathway
pathway

Current
Current
dominant
dominant
energyenergy
use use
Propagation
Propagation
Propagation
method
Propagation
method
method
method

Bioethanol
Bioethanol
BiodieselBiodiesel
Heat andHeat
power
and power
Biogas Biogas
Seed Seed Seed
Stem Seed
Stem Rhizome
Seed
Stem Rhizome
Stem
Stem
Micropropagation
Rhizome
Micropropagation
Rhizome
Micropropagation
Rhizome
Micropropagation
Micropropagation
cutting cutting
cutting
cutting
cutting
or root
cuttings
or root
cuttings
or root
cuttings
or root cuttings
or root cuttings
Other Other

Bioethanol
Bioethanol
BiodieselBiodiesel
Heat andHeat
power
and power
Biogas Biogas
Seed Seed Seed
Stem Seed
Stem Rhizome
Stem Rhizome
Stem
Micropropagation
Rhizome
Micropropagation
Rhizome
Micropropagation
Micropropagation
cutting cutting
cutting
cutting
or root
cuttings
or root
cuttings
or root cuttings
or root cuttings
Other Other

Current
Current
dominant
Current
dominant
Current
energy
dominant
Current
energy
use
dominant
energy
use
dominant
energy
use energy
use use

Current
Current
dominant
Current
dominant
Current
energy
dominant
energy
use
dominant
energy
use energy
use use

7.1 tonnes

7.3 tonnes

How long could


you supply energy
Bioethanol
Bioethanol
Biodiesel
Bioethanol
Biodiesel
Bioethanol
HeatBiodiesel
and
Bioethanol
Heat
power
Biodiesel
andHeat
power
Biogas
Biodiesel
andHeat
power
Biogas
andHeat
power
Biogas
and power
Biogas Biogas
for a house?
Car

Effective energy
DE DE D
D
available

Car Weight Weight Barrel Barrel

DE

Other Other Other Other Other

World

Car

11,000kWh

Best recorded

Effective energy
DE D
D
available

Car Weight Weight Barrel Barrel

DE

Bioethanol
Bioethanol
Biodiesel
Bioethanol
Biodiesel
Bioethanol
HeatBiodiesel
andHeat
power
Biodiesel
andHeat
power
Biogas
andHeat
power
Biogas
and power
Biogas Biogas

Other Other Other Other

World

World

Car Weight
Car Weight
CarBarrel
Weight
CarBarrel
Weight Barrel
WeightBarrel Barrel
Power usage
Power usage

Car

months

Car

Car Weight
Car Weight
CarBarrel
WeightBarrel
Weight Barrel Barrel
Power usage
Power usage

12,000kWh

Best recorded

Best recorded

33

Power usage
Power usage
Power usage
Power usage
Power usage

55,000kWh

months

Global planting
5.3 million hectares in plantations plus 3.9 million
hectares in agroforestry and conservation
(for soil and water protection).
Global production
38 million tonnes/yr.

How long could


you supply energy
for a house?
World

months

Best recorded

24

Power usage
Power usage
Power usage
Power usage

39,000kWh

months

Global planting
90,000 hectares for wood production plus
86,000 hectares for land reclamation and
conservation.
Global production
660,000 tonnes/yr.

91

Harvesting

Key references

The plants are cut off close to the ground at harvest. The
first years crop is small, but is removed to stimulate
growth. Subsequent harvests are made after leaf-fall
at intervals of three to 11 years, depending on variety,
growing conditions and harvest technology available.
Harvesting is highly mechanized and consists of either
cutting and storing whole stems, or chipping the stems
immediately after harvest. The latter approach requires
drying equipment, as a mass of moist chips will rapidly
heat up (as happens in a compost heap). The harvested
wood has a moisture content of around 50%.

Yield
The yield for these trees has to be considered over a
harvest cycle of three to five years for SRC (willow or
poplar) or around seven to 11 years for SRF (poplar grown
as single stems). Worldwide average annual yields of both
willow and poplar are about 7 tonnes/ha, but there is huge
variation between sites. Improved varieties produce higher
yields, with many commercial growers now achieving
1014 tonnes/ha.
Biofuel (liquid) yield is not yet known because
the process is still under research. Thermochemical
Plant types
conversion pathways have been shown to successfully
yield liquid fuels. In terms of thermal energy, willow
has a calorific value of 20GJ/tonne and poplar 19GJ/tonne.

Alternative markets

Herbaceous

Karp, A., Hanley, S. J., Trybush, S. O., Macalpine, W.,


Pei, M. & Shield, I. (2011), Genetic improvement of
willow for bioenergy and biofuels, Journal of Integrative
Plant Biology, vol. 53, no. 2, pp. 151165.
Fischer, G. et al. (2005), Biomass potentials of
miscanthus, willow and poplar: results and policy
implications
for Eastern
Plant
types Europe, Northern and Central
Asia, Biomass and Bioenergy, vol. 28, pp. 119 132.
Annual
Plant
types
Davis, S. C. et al. (2012), Harvesting carbon from
eastern
US forests: opportunities and impacts of an expanding
Perennia
bioenergy industry,
Forests,
vol. 3, Grain
pp. 370
397.
Woody
or seed
Herbaceous
Jansson, S., Bhalerao, R. & Groover, A. (eds) (2010),
Woody G
Herbaceous
Genetics and
genomics of Populus.
Springer, New York,
Photosynthetic
pathway
NY, US.
Photosynthetic pathw
Lantmannen Agroenergi
(noCdate), Manual for SRC
C3
4
willow growers. York, UK and Orebro, Sweden. C
C4
3
Plant types
C3

C4

Annual

This box highlights another group of woody plantsC3that


can be managed as short-rotation forestry systems.Perennial
Woody Grain or seed
Propagation
method

C4

Herbaceous

Propagation method

Photosynthetic pathway
C3

PlantCtypes
4
Seed
C

C3

Woody

Grain or seed

Annual

Stem
cutting

Perennial
Woody
Herbaceous

Annual

Rhizome
Micropropagation
Seed
Stem
or root cuttings
cutting
Grain or seed

Perennial

energy use Current dominant en


PineCurrent dominant
Eucalyptus
Propagation method
Photosynthetic pathway
Photosynthetic
pathway
Plant
types
Biomass from willow and poplar is already used as fuel
Annual
in thermal power stations. Poplar grown in single stems
Perennial
C3
C4
C
C
3
4
(rather than coppiced) has various uses including plywood
E
D Perennial
E
D
and pulp for paper. The trees hardiness and tolerance of
C
C
Woody
Grain
or
seed
Herbaceous
Seed
Rhizome
Micropropagation
Bioethanol
Bioethanol3 Stem
Biodiesel4 Heat
and power
Biogas Biodiesel H
waterlogging makes them useful as windbreaksC3and in C4
cutting
or root cuttings
Pine,
eucalyptus
(Pinus
spp.,
Eucalyptus
spp.)
bioengineering to stabilize soils. Some poplars and willows
Photosynthetic pathway
Other
Photo
in production:
Australia
and Brazil. Other
are also effective in phytoremediation of soils with heavy
Current
dominant
energy use
Propagation
method
Propagation
method

Why
is
it
similar?
Softwood
(e.g.
pine) and
metals and have facilitated the reduction of nutrient run- C 3
C4
hardwood (e.g. eucalyptus) plantations also
off from agricultural landscapes.
comprise
C3 trees that are intensively managed to
C3
C4
Car
D
produce Ewood biomass;
harvest
rotation
Car
WeightlengthsBarrel
Co-products
range
from
five
to
25
years.
Eucalyptus
coppices
Bioethanol Micropropagation
Biodiesel
power Biogas
Seed Heat and Stem
Rhizome
Micropropagation
Seed
Stem
Rhizome
From the crop
cuttingand are
or root
cuttings
regrow
vigorously from cut stumps
often
cutting
or root cuttings
Propagation
method
Otherlike poplar and willow.
Currently none but poplars and willows is rich in
managed
dominant
energypine
use trees
Current as
dominant energy
use makes itCurrent
secondary metabolites (e.g. salicin, later developed
What
different?
After harvest,
aspirin, was first identified in willow).
must be reseeded (eucalyptus can also be grown
from seed rather than relying on coppice regrowth);
From conversion to fuel
Car
Weight
Barrel
Seed
Stemrates of Rhizome
Micropropagation
greater
soilEdegradation
D occur in this system
E
D
When burned as fuel for heating or in power stations,
cutting
or root cuttings
relative to coppicing woody crops.
only ash residues remain. If used to make bioethanol,
Bioethanol
Biodiesel Heat and power Biogas
Bioethanol the
Biodiesel Heat
power Biogas
and
Growing
region:
Powe
Current dominant
energy
useWood plantations exist all over
wastes after hydrolysis and fermentation could be burned
the world, with
the
largest
production
currently in
Power
usage
Other
Other
for heat.
the US followed by Brazil, the Russian Federation,
Canada, China and northern European nations.
Invasion risk
E
D produces the largest amount of wood that is
India
intended
exclusively
for fuel.
Some species naturally shed branches, which can then Bioethanol
Car
Weight
Barrel
Biodiesel
Heat
and power Biogas
Car
Weight
Barrel
re-root, and both poplar and willow can be potentially
Power usage
Other
invasive in some environments. This has been historically
observed in Australia and New Zealand, where Salix
fragilis L. and Salix cinerea L. were found to rapidly
colonize along streams. Future improvements in genetic
Car
Weight
Barrel
engineering may increase productivity but also plant
invasiveness.
Power usage

Power usage

92 | 6 Biomass feedstock crops


Power usage

Seed

Wood residues

Stem
cutting

Rhizome
Micropropagation
Seed
Stem
or root cuttings
cutting

Rhizome
Micropropagation
or root cuttings

Current dominant energy use Current dominant energy use


Plant types

Annual

E
Herbaceous

Woody

C3

C4

DPerennial

Grain or seed

Bioethanol
Bioethanol
Biodiesel Heat and power
Biogas Biodiesel Heat and power Biogas
Type
of fuel: Solid combustion for heat and
Photosynthetic
pathway
Other
Other biofuel.
electricity, cellulosic
C
Stage
of
adoption:
early
commercial
production.
C3
4

The main use of wood is now industrial


sawn
Car
Weighttimber,Barrel
poles, manufactured panels, plus pulp and paper. Although
tree-breeding programmes aim to reduce waste at source
Propagation method
by producing trees with minimally tapering trunks and
few branches, residues arise at all stages of production,
from harvesting though to the final product. The many
by-products of timber production therefore represent a
Seed significant
Stemsource of
Rhizome
Micropropagation
woody biomass,
which can be burned
or root cuttings
for heatcutting
or used in
thermal power stations.
As timber
processing
will be a major industry for the
Current dominant
energy
use
foreseeable future, wood wastes are Power
a reliable
source of
usage
biofuel feedstock. Improvements in timber harvesting
techniques will reduce the amount of wastage during
E
D but this represents only a small fraction
logging,
of
the
whole.
Bioethanol
Biodiesel
Heat and power Biogas
Other

Characteristics
Wood residues take many forms, including logging
waste (tree tops and branches), sawdust and shavings
fromCar
sawmills,
pulping liquor
Weight
Barrel from paper production,
and waste wood from construction and demolition
sites. The material is far from uniform. As well as the
obvious physical difference between, say, sawdust and
waste lumber, tree species differ in their carbon and
lignin contents. Hardwoods tend to have a higher carbon
content than softwoods (55% versus 46% on average).

Car

Weight

Barrel

The moisture
content of woody waste depends on species,
Power usage
time of harvest and source. Poplar and willow can have
moisture contents of more than 50% and primary mill
residues less than 20%.

Accessibility
Logging, wood processing and production industries
provide a ready supply of wood residues. The largest source
of uniform-format material comes from forest product
industry waste. But there is also a large amount of woody
waste (termed brash) that is left on the ground after
forestry operations because it is currently uneconomic
to retrieve; it also supports recycling nutrients in some
places (not all residues should be removed). Collecting
timber waste from demolition and construction sites,
and from urban tree management, would provide another
source of woody residues.

World map showing five top producing countries


Power usage

Russia Federation

Canada
US

China

Top producing
countries

Million
Tonnes

US
Canada
Russia Federation
Brazil
China

284
139
136
128
102

Brazil

Global planting n/a.


Global production 240 million tonnes of wood chips and particles
plus 31 million tonnes of mechanical wood pulp (dry weight).

6 Biomass feedstock crops | 93

Alternative markets
Wood residues are used in pulp production, to make
particle board for construction, as well as pellets for fuel.

Key references

Potential uses after conversion to fuel


Wastes after hydrolysis and fermentation can be burned
for heat.

94 | 6 Biomass feedstock crops

Food and Agriculture Organization of the


United Nations, FAOSTAT Forestry database.
Haberl, H. et al. (2010), The global technical potential
of bio-energy in 2050 considering sustainability
constraints, Current Opinion in Environmental
Sustainability, vol. 2, pp. 394 403.
Hoekman, S. K. (2009), Biofuels in the U.S. challenges
and opportunities, Renewable Energy, vol. 34,
pp. 14 22.
US Department of Energy, (2011), U.S. billion-ton
update: biomass supply for a bioenergy and bioproducts
industry. Oak Ridge National Laboratory, Oak Ridge,
TN, US.

Seed

Woody

Annual
Stem
Stem
cutting
cutting
Perennial

Grain or seed

C4

or root cuttings

Current
dominantenergy
energy
use
Current dominant
use

hetic pathway
Plant types
C4

Rhizome
Micropropagation
Seed
Stem
Rhizome
Micropropagation
or root cuttings
cutting

Herbaceous crop residues


Seed

Rhizome
Micropropagation
or root cuttings

Current dominant energy use


Annual

DD

Perennial
Woody Grain
or seedHeat and power Biogas Bioethanol
Herbaceous Bioethanol
Biodiesel
Bioethanol
Biodiesel
Heat and power Biogas

Type of fuel:
Cellulosic biofuel, solid combustion
Photosynthetic
Otherpathway
Other
Other
for heat
and electricity, biogas production.
on method CStage of adoption: extensive commercial
C4
3
production for other uses, pre-commercial scale
biofuel.C4
Cfor
Car
Weight
Barrel
3
Car
Weight
Barrel

D
Biodiesel Heat and power

Car

Weight

Biogas

Barrel

Stem
Rhizome
Micropropagation
Characteristics
cutting Propagation
or root cuttings
method

Mostly bulky, with variable handling requirements,


ominant energy use
agricultural wastes have different carbon (50 70%) and
lignin (7 19%) contents depending on the crop. Moisture
contents usually vary between 10% and 30%. Most crop
Seed
Stem
Rhizome
Micropropagation
residues would becutting
collectedorin
the
field, although residues
D
root
cuttings
are sometimes available at the
site
of pre-processing for
Power
usage
Biodiesel Heat and power Biogas
Current
dominant
energy
use
fuel,
as with
sugarcane
bagasse.
Power usage

Accessibility
CropEwastes areDusually cheap and widely available
wherever
is large-scale
commercial
Bioethanol there
Biodiesel
Heat and power
Biogas agriculture.
Car
Weight
Barrel
Global availability mirrors crop production patterns
the
largest source of uniform-format material is rice straw,
Other
followed by wheat straw, maize stover and sugarcane
bagasse. These four feedstocks together have been
estimated to have the potential to annually produce
Car of bioethanol.
Weight
Barrel for the potential
418 billion litres
Estimates
of all crop residues suggest that 1,200 billion litres of
bioethanol could be produced annually.
While sugarcane bagasse is already collected and a
Power usage
readily available by-product of processing, the other
agricultural residues are dispersed across fields and would

require additional collection and handling. The biggest


risk of using this low-cost, lightweight and plentiful
material is the depletion of soil organic matter and
increased erosion of agricultural soils. To guard against
this, soil organic matter must be regularly monitored and
it is suggested that only a proportion of available residues
Power usage
are collected.
Reduced or zero-tillage systems allow a
greater proportion of crop waste to be removed without
increasing erosion risk.

Alternative markets
Crop wastes have, for millennia, been used as soil
conditioners, low-grade animal feed and animal bedding.
Left on the soil they are particularly valued to prevent
erosion, improve soil texture and provide some degree of
nutrient recycling (although nitrogen losses can increase
in the short term).

Potential uses after conversion to fuel


Wastes after hydrolysis and fermentation can be
burned for heat.

World map showing five top producing continents


Power usage

Europe

North
America

Africa

Asia

Top producing
continents

Million
tonnes*

Asia
North America
Europe
South America
Africa

920
220
220
110
38

*dry weight

South
America

Global planting n/a.


Global production 1.6 billion tonnes (dry weight).

6 Biomass feedstock crops | 95

Herbaceous crop residues average annual yield


Current fuel yields vary according to crop, as shown in the table below.

Residue/crop ratio

Estimated ethanol yield


(L kg-1 dry mass)

Energy yield
(MJ kg-1 dry mass)

Maize stover

1.0

0.29

98

Barley straw

1.2

0.31

100

Oat straw

1.3

0.26

88

Rice straw

1.4

0.28

94

Wheat straw

1.3

0.29

98

Sorghum straw

1.3

0.27

91

Sugarcane bagasse

0.6

0.28

94

Mean

1.2

0.28

95

Key references

Cherubini, F. & Ulgiati, S. (2010), Crop residues as raw


materials for biorefinery systems a LCA case study.
Applied Energy, vol. 87. pp. 47 57.
Kim, S. & Dale, B. E. (2004), Global potential bioethanol
production from wasted crops and crop residues,
Biomass and Bioenergy, vol. 26. pp. 361 375.
Lal, R. (2005), World crop residues production and
implications of its use as a biofuel, Environment
International, vol. 31, pp. 575 84.

96 | 6 Biomass feedstock crops

Propagation method

Annual

Seed
Stem
Rhizome
Micropropagation
Seed
Stem
Rhizome
Micropropagation
Miscellaneous
wastes
cutting
or root cuttings
cutting
or root cuttings
Seed

Woody

Grain or seed

hetic pathway
C4

Bioethanol

on method

Stem
cutting

Rhizome
Micropropagation
or root cuttings

Current
dominant
energyCurrent
use dominant energy use
Current dominant
energy
use

C4

Stem

Perennial
cutting

D E

Biodiesel
Heat and
power Biogas
Bioethanol
Biodiesel
Bioethanol
Biodiesel
Heat and
power Biogas

Type of fuel: methane gas for heat and electricity,


Other
Other
ethanol, biodiesel. Other
Stage of adoption: early commercial production of
methane, pre-commercial stage for ethanol, early
Car of biodiesel
Weight
Barrel
commercial production
from
used
Car
Weight
cooking oil and animal fats.

BarrelCar

Heat and power

Weight

Biogas

Barrel

Rhizome
Micropropagation
or root cuttings

Characteristics

ominant energy use


Miscellaneous wastes include municipal organic waste,
sewage and animal manure as well as fatty oil and cooking
oil wastes. These wastes have different properties,
D
Power
usage
although there are similarities
between
animal manure
and
and between fatty waste and waste cooking
Biodiesel Heat
andsewage,
power Biogas
Power usage
oil. Moisture content varies from 5% to 80%, and most
wastes need to be dried before further processing can take
place. Given the variation in source material in this broad
category, it is not surprising that they give rise to a wide
range of fermentation products, chemical extracts and
Car
Weight
Barrel
processing residues.
Using waste materials to generate energy has great
environmental benefits, can reduce methane emissions
and avoid the need to expand the area of landfill sites
(and dealing with the environmental issues that this
practice raises). Special equipment is needed to prevent
environmental contamination and minimize public
nuisance in their processing.
PowerAccessibility
usage

Unlike wastes from crops, which have seasonal


availability, miscellaneous wastes are produced
throughout the year. This makes them reliable feedstocks
for fuel production. The largest source of uniform-format
material is animal manure, from which it is estimated that
9 25EJ/yr could be recovered globally.

Power usage

Alternative markets
Organic waste is often incinerated to produce heat or
co-generate electricity. Animal manures have long been
used as fertilizer, sometimes after a period of composting.
Animal manures, municipal waste and sewage can all be
digested anaerobically to produce biogas, which has the
added benefit of reducing carbon dioxide and methane
emissions. Some of these waste materials can be processed
into animal feed or biomaterials.

Potential uses after conversion to fuel


Wastes are highly variable and the by-products from
conversion to fuel depend on the source material and fuel
processing technology.

Key references

Yield
Fuel yields depend on source material, with a range of
114 gigajoules per tonne. Municipal organic waste yields
more than animal manure, on average.
Although an estimated 50EJ of energy is expected to be
available from wastes by 2050, estimates of sustainably
available energy from all waste residues, worldwide, range
between 10 and 29EJ/yr.

Cherubini, F. et al. (2009), Life cycle assessment (LCA)


of waste management strategies: landfilling, sorting
plant and incineration, Energy, vol. 34, pp. 2116 2123.
Hoogwijk, M. (2003), Exploration of the ranges of the
global potential of biomass for energy, Biomass and
Bioenergy, vol. 25, pp. 119 133.
Mnster, M. & Lund, H. (2009), Use of waste for heat,
electricity and transportchallenges when performing
energy system analysis, Energy, vol. 34, pp. 636 644.

6 Biomass feedstock crops | 97

Woody

Grain or seed

ic pathway

Propagation method
Propagation method

Perennial

Plant types

Annual

Stem Seed Rhizome Stem


Micropropagation
Rhizome
Micropropagation
Photosynthetic Seed
algaecutting
cutting
or root cuttings
or root cuttings

C4
C4

Woody

C3

C4

Perennial

Grain or seed

Current dominant Current


energy use
dominant energy use
Photosynthetic pathway

method

Stem
cutting

Herbaceous

Rhizome
Micropropagation C3
or root cuttings
Type of fuel: biodiesel,

minant energy use

C4
Bioethanol

Biodiesel Bioethanol
Heat and power
Biodiesel
BiogasHeat and power

Biogas

ethanol.
Other
Other
Stage of adoption: pre-commercial
scale production.
Propagation method

Representatives of this vast group of organisms are found


in almost all global habitats, reproducing rapidly
Car to create
Weight
CarBarrel Weight
Barrel
odiesel Heat and power Biogas
oils, carbohydrates and proteins within their single cells.
Seed producing
Stem
Rhizome
Micropropagation
They do not waste energy
the infrastructure
cutting
or root cuttings
found in more complex organisms. Microalgae are already in
commercial production
for dominant
nutritionalenergy
supplements
but the
Current
use
of biofuel is as yet unfulfilled.
Car
Weightpromise
Barrel
Despite a 20-year USDOE-funded research programme
into producing biodiesel from algae, the process is not yet
Inputs required
E less than
D 76,000 litres being
commercially viable, with
Water
produced in 2010. Nevertheless,
the fact
that
Bioethanol
Biodiesel
Heat
andalgal
powerponds
Biogas
can tolerate a wide range of salinity and have a
Power
usage
Power usage
can be established on otherwise unproductive land, plus the Microalgae
large
water
requirement.
Other
high potential yields, suggest that microalgae may become
more important as production and harvesting techniques
Fertilizer
Power usage are further refined. Small-scale experimental results
For optimum growth, algae need nitrogen, phosphorus
indicate a possible yield of 13,000 litres per hectare, with
and silicate (diatoms need this to produce their silicaceous
Car
Weight
Barrel
some suggesting that much higher yields may be achievable. coats) as well as a range of micronutrients. In closed
It is also possible to ferment algae biomass to produce
systems, CO2 must also be supplied.
ethanol, although this has rarely been done.
D

Plant characteristics
There are somewhere between several hundred thousand
and tens of millions of different algal species, all singlecelled, although some form colonies. Research into biofuels
has focused on the photosynthetic green
Power algae
usage and diatoms,
because the lipids they manufacture can be converted to
biofuels very efficiently.

Where to grow it
Microalgae grow spontaneously wherever there is water and
sunlight. Commercial production (which is currently done
to produce high-value products but not fuels) can therefore
be done almost anywhere that water temperatures can be
maintained between 16C and 27C. Optimum growth
for many species is 2024C. High-volume production
of low-cost biofuels requires access to moderate to high
temperatures, sufficient sunlight and water and, in most
cases, a source of CO2.

Contamination
Algal cultures are easily contaminated, so sterilization and
hygiene techniques are important.

Harvesting
The removal and dewatering of the algae uses a great deal
of energy. During the transesterification route of algae
to biodiesel, the algae are first separated from the water
by centrifuge or filtration and then dried, sometimes
using solar energy. Algal lipids are then removed and
transformed through reaction with an alcohol (such as
ethanol or methanol) into biodiesel and glycerol.

Alternative markets
Pilot plants researching energy feedstock generate income
by selling microalgae for animal feed or as specialist
human nutritional supplements. Algae are also used in
cosmetics and in the treatment of wastewater.

How to grow it

Co-products

There are two methods in current use: open circulating


ponds or enclosed photobioreactors that produce biomass
of similar quality. Ponds are cheaper and simpler,
relying on a paddlewheel to recirculate and mix the
water. But they are open to the air, so there is the risk
of contamination by dust, bacteria and other strains of
algae. Photobioreactors maximize sunlight capture by
containing the algal mix in small-diameter tubes and
protect the algae from contamination. Closed systems
present their own challenges, however, such as providing
for sufficient CO2 transfer and making sure that heat and
oxygen do not build up.
98
98 | 6 Biomass feedstock crops

From the crop


Currently none, as production is still in the research stage.
From conversion to fuel
Potential uses of microalgae residues after conversion to
biodiesel are animal feed or to produce methane. Glycerol
is a co-product of biodiesel conversion and can replace
glycerol made from petroleum. Glycerol has many uses
in cosmetics, food, oral care, tobacco, synthetic polymer
production and synthetic resins.

Key references

Brennan, L. & Owende, P. (2010), Biofuels from


microalgaea review of technologies for production,
processing, and extractions of biofuels and co-products,
Renewable and Sustainable Energy Reviews, vol. 14,
pp. 557 577.
Chisti, Y. (2007), Biodiesel from microalgae,
Biotechnology Advances, vol. 25, pp. 294 306.
Daroch, M. et al. (2013), Recent advances in liquid
biofuel production from algal feedstocks, Applied
Energy, vol. 102, pp. 1371 1381.
Schenk, P. M. et al. (2008), Second generation biofuels:
high-efficiency microalgae for biodiesel production,
BioEnergy Research, vol. 1, pp. 20 43.
Macroalgae are also researched for their bioenergy
potential, but are very different from microalgae
and require very different production pathways.
Commonly known as kelp or seaweed, these
organisms are grown today for food use, most
commonly by tethering the strands of growing
macroalgae in open seawater during their growing
period and harvesting by collection. The macroalgae
can be quite productive, but have a high content of
water and salts, which degrade their fuel value. This
research is in the early stages and the viability of
macroalgae as a feedstock for fuel is uncertain.

6 Biomass feedstock crops | 99

Additional species
Representative list of additional species of plants used (or
with potential) for the production of bioenergy:
Agave (Agave spp.)
Agave is a genus of succulent plants that grows in
semi-arid and arid regions. These plants use the CAM
photosynthetic pathway, fixing carbon during the night
while closing their stomata (essentially pores in the
leaves) to minimize water loss during the day, so have a
much lower water requirement than many other crops.
Yields vary according to species and location, but are
reportedly between 1 and 34 tonnes/ha/yr. An additional
potential advantage of agave is the availability of land
formerly intended for agave fibre production, which is
estimated to have been about 600,000 hectares worldwide.
Agave species occur naturally in the south-western US,
Central America and parts of South America. Varieties are
commercially cultivated in Central and South America
and Africa, and experimentally cultivated in Australia.
Agave thus has the potential to be grown in many
semi-arid regions of the world, specifically on land not
otherwise used for food production.
Bamboo
A family of perennial giant woody grasses of the subfamily
Bambusoideae, these plants are native to equatorial and
tropical regions around the globe. While bamboo is
cultivated for a wide range of purposes, only small
amounts are used for firewood. Annual yields of common
native species range from 1.5 to 14 tonnes/ha. The stems
have been found to have an energy content of ~17GJ/tonne
dry matter and contain around 40% cellulose. Different
species of bamboo mean that the potential crop species can
be grown over a wide geographical area including southern
Russia, Europe, North America and central Asia. Recent
work in India has indicated that clear felling of the crop
may open the potential for mechanization. A major
concern with bamboo, however, is invasiveness and
difficulties of removal once the plantation is no longer
in use.
Black locust (Robinia pseudoacacia L.)
A pioneer hardwood species found in North America,
these trees displays rapid juvenile growth. Mature trees
can reach a height of 15 35m and the plant has a high
rate of photosynthesis. In trials, commercial stands have
produced yields of 5 10 tonnes/ha/yr in three- or fouryear-old stands; peak production is seen at seven years. It
burns slowly and has very high energy content, making it
an ideal feedstock. Because of the trees tendency to sprout
from both the stem and roots, it may be suitable for both
SRF and SRC. As a legume, its roots contain nitrogenfixing bacteria, which can help improve soil fertility. A
native of North America, the tree is also successfully
grown in other areas including Europe; however, it has the
potential to become invasive outside its native range.

100 | 6 Biomass feedstock crops

Camelina (Camelina sativa L.)


Although promoted as a potential feedstock for biodiesel
production, the unique properties and relatively low yields
(1.5 3 tonnes of seed per hectare) mean that this plant, a
member of the cabbage family, is likely to be confined to
the specialist chemical market or jet fuels. Camelina can
be widely grown in temperate agricultural systems.
Cassava (Manihot esculenta Crantz.)
There are many cultivars of cassava, a crop that has
been domesticated for more than 2,000 years, some
of which are grown as biofuel feedstock and others
as food. It is one of the most drought-tolerant crops,
can be successfully grown on marginal land, and has
a great ability to produce more useable energy per
unit of land than comparable crops. The cultivation
of cassava as a feedstock for ethanol production is
increasing rapidly with new plants in operation in
South-East Asia, especially Thailand, China and
Vietnam. Development is also under way in other
parts of the world including Panama and Nigeria,
where a Chinese energy company is planning a
110-million-litre-a-year plant. In addition to ethanol
production, cassava residues are used to generate
biogas.
Modern breeding efforts are focused on increasing
the protein content to improve diets in Asia,
South America and Africa, where it is already
successfully grown. As well as having the potential
to produce dedicated bioenergy crops based on the
starchy tuber, conventional cultivation already
produces a large amount of vegetative material
suitable for anaerobic digestion. Soil disturbance
associated with harvesting below-ground biomass
could, however, result in greater terrestrial carbon
emissions than alternative crops.

Diesel tree or Jesuits balm (Copaifera langsdorfii Desf.)


Reaching up to 12m in height, this tropical rainforest
tree produces hydrocarbon-rich oil, which is harvested
by tapping. In tropical areas oil yields can reach
10,000l/ha/yr or more, although plantations can take
15 20 years to reach this level of production. Despite this
slow attainment of full productivity, some farmers
in Australia are investigating production of the tree as a
crop following its introduction from Brazil.
Abyssinian mustard (Crambe abyssinica Hochst.)
This herbaceous crop can be harvested after 100 days with
an oilseed yield of 2.5 3 tonnes/ha, although it may be
more suitable for high-value chemical production rather
than bioenergy. It can be grown in areas with moderate
rainfall such as central Asia and Western Europe.
Field pennycress (Thlaspi arvense L.)
A naturally occurring small brassica, field pennycress is
a low-growing annual weed often found in arable crops. It
has attracted academic interest as a potential bioenergy
feedstock due to its production of oil-rich seeds. It is
widespread in areas of temperate agriculture.
Giant reed (Arundo donax L.)
Found wild in parts of southern Europe, as well as the
US and China, it is believed that this species could adapt
to more temperate conditions. Producing yields of up
to 40 tonnes/ha in the wild, it is regarded as the largest
temperate grass after bamboo. The plant, however, is
listed as an invasive species in the US and in the Global
Invasive Species Database, and is considered a high risk for
waterways.

Napiergrass (Pennisetum purpureum Schumach.)


This is the true elephant grass (a term sometimes
mistakenly used for miscanthus), a robust C4 grass that
can grow to 6m in height. It requires nutrient-rich soils
and yields under cultivated conditions are highly variable
ranging from 2 to 85 tonnes/ha (dry mass). A native of the
African grasslands, it is being investigated as a commercial
species in many temperate and sub-tropical regions, with
trials conducted in Central and South America, Africa and
parts of Asia. It is considered a high-risk invasive species in
the Pacific Islands, Australia and the US.
Prairie cordgrass (Spartina pectinata Bosc.)
There are 16 species of Spartina, and they utilize the C4
photosynthetic pathway. Of the three species that have
been seriously investigated for bioenergy use, S. pectinata
is the most common, although yields in experimental trials
have been variable (516 tonnes/ha), due to wide-ranging
environmental stress factors. Despite this, prairie cordgrass
can be more tolerant to salts and metals than other grass
species. It is found throughout the US and Canada, and
has potential for production in parts of northern Europe
and Russia.
Reed canary grass (Phalaris arundinacea L.)
A perennial C3 plant often found in cool damp climates,
reed canary grass (unlike many other grasses promoted as
bioenergy crops) can be grown from seed and harvested
with existing equipment. There has already been some
commercial breeding in Scandinavia; typical yields in
the region are 10 tonnes/ha. It can be grown throughout
North America and northern Europe and there are an
estimated 10,000ha cultivated for pulp and paper and
solid fuel production in Scandinavia. This plant is listed
as an invasive species in the US and in the Global Invasive
Species Database.

6 Biomass feedstock crops | 101

Biomass crops: comparison table


Sugarcane

Corn

Soybean

Switchgrass

Miscanthus

Type of crop

Perennial
C4
Leaf/stem
Sugar

Annual
C4
Grain
Starch

Annual
C3
Oil seed

Perennial
C4
Leaf/stem
Cellulose

Perennial
C4
Stem
Cellulose

Fuel type
(commercial and
pre-commercial)

Bioethanol,
biopower,
lignocellulosic
ethanol

Bioethanol

Biodiesel

Biopower,
lignocellulosic
ethanol

Biopower,
lignocellulosic ethanol

Latitude

37N31S

54N34S

52N39S

55N17N

56N37N [2]

Suitable soils

Wide range,
requires draining
on heavy clay

Best on mediumtextured soils

Best on clay loam;


moderately salttolerant

Once established,
tolerant of most
soils

Dislikes heavy
clay; best on waterretentive soils; also
good on sand

Water
requirement [1]
(mm)

High
1,5002,500

Moderate
670800

Moderate
600

Moderate
520750

Low-moderate
450 minimum but
will use more when
available

Temp.
(none grows well
above 45C)

Mean temperature
at least 18C

Optimum
2430C

1040C

Germinates above
810C, optimum
2530C

Shoots grow above


7C

Fertilizer
requirement
(kg/ha/yr)

N: 45300
P: 1550
K: as required

N: 145200
P: 26110
K: 25130

N: 070
P: 32155
K: 30320

N: 50168
P: 035
K: 045

N: 092
P: 013
K: 0202

Insect control

Range of pesticides

Range of
pesticides

Range of pesticides Generally not


needed

Generally not needed

Global average
yield
(tonnes/ha/y)

71 wet

5.2 dry

2.9 (seed)
0.44 (oil)

14 dry

18 (Europe) [6]
38 (North America)
dry

Area currently
in cultivation
(million ha)

24

160

99

Unknown

Unknown

Energy equivalent
of current
biomass
(EJ)

3.3

7.5

1.6

Unknown

Unknown

Top three
producing
countries
(or continents)

Brazil
India
China

US
China
Brazil

US
Brazil
Argentina

Unknown

Unknown

102 | 6 Biomass feedstock crops

Oil palm

Poplar/willow

Wood residue

Crop residue

Misc. waste

Algae

Perennial
C3
Oil seed

Perennial
C3
Woody
Cellulose/lignin

Wastes from
timber harvest and
processing etc.

Straw from cereal


crops etc.

Animal, municipal
wastes etc.

Micro-crop
Algal lipids

Biodiesel

Biopower,
lignocellulosic
ethanol

Biopower,
lignocellulosic
ethanol

Biopower,
lignocellulosic
ethanol

Biopower, biogas

Biodiesel

15N12S

Poplar: 66N30N
Willow: 75N34S
[3]

Tolerates wide
range of pH; best
in sandy soils with
good drainage

Wide range, but


establish slowly
on heavy clay

High
2,0002,500

Low-moderate
320450 plus
groundwater

High, esp. in raceway


systems

Optimum
2428C [4]

Wide range: very


tolerant of severe
frost and heat

2030C

N: 114
P: 14
K: 149

N: 20210
P: 1393
K: 25174

Estimated from
molecular formula of
microalgal biomass:
needs only N and P
in ratio 445N:1P

[5]

Range of pesticides Not commonly


used

14 (fruit)
2.9 (oil)

Poplar: 7.1
Willow: 7.3

270 million
tonnes/y

1.6 billion tonnes/y

Unknown

15

Poplar: 5.3
Willow: 0.09

1.6

Poplar: 0.69
Willow: 0.013
(biopower)

4.1EJ/y
(biopower)

150EJ/y [7]
(biopower)
11EJ/y (bioethanol)

1029EJ/y [8]

Indonesia
Malaysia
Nigeria

China
India
France

US
Canada
Russia

Asia
North America
Europe

Unknown

Unknown

Unknown

Notes
[1] Most water requirements are met by precipitation.
[2] This range in successful trials; may well be wider.
[3] The extreme southerly part of range being along water
courses in Argentina.
[4] Commercial production limited to areas with only
a 6C seasonal mean temperature variation.

[5] These figures for 10-year-old stands; 15-year-old stands have


higher requirements for N, P and K (N: 162, P: 21, K: 279).
[6] Winter yields: harvest is usually done in winter,
despite lower yields, to facilitate nutrient retention.
[7] Estimate of global availability of crop residues.
[8] Estimate of current potential sustainably available bioenergy.

6 Biomass feedstock crops | 103

Biomass feedstock crops: complete list of references


Maize (corn)
Ahmed, S. & Rao, M. (1982), Performance of maizesoybean
intercrop combination in the tropics: results of a multi-location
study, Field Crops Research, vol. 5, pp. 147161.
Anderson, E. & Cutler, H. C. (1942), Races of Zea mays: I. their

recognition and classification, Annals of the Missouri Botanical


Garden, vol. 29, pp. 6986.
Anthony, C. G. (1988), Mechanization and maize. Columbia

University Press, New York.


Aubry, C. et al. (1998), Modelling decision-making processes
for annual crop management, Agricultural Systems, vol. 56,
pp. 4565.
Bosque-Prez, N. A. & Mareck, J. H. (1991), Effect of the stem

borer Eldana saccharina (Lepidoptera: Pyralidae) on the yield


of maize, Bulletin of Entomological Research, vol. 81,
pp. 243247.
Bush, M. (1989), A 6,000 year history of Amazonian maize

cultivation, Nature, vol. 340, pp. 303305.


Le Clerc, V. et al. (2006), Indicators to assess temporal genetic

diversity in the French Catalogue: no losses for maize and peas.


Theoretical and Applied Genetics, vol. 113, pp. 11971209.
Daberkow, S. et al. (2000), Nutrient use and management,

in Agricultural resources and environmental indicators.


US Department of Agriculture, Economic Research Service,
Washington DC.
Davis, S. C. et al. (2011), Impact of second-generation biofuel

agriculture on greenhouse-gas emissions in the corn-growing


regions of the US, Frontiers in Ecology and the Environment,
vol. 10, pp. 6974.
Devos, Y. et al. (2008), Environmental impact of herbicide

regimes used with genetically modified herbicide-resistant


maize, Transgenic Research, vol. 17, pp. 105977.
Doebley, J. (2004), The genetics of maize evolution,

Annual Review of Genetics, vol. 38, pp. 3759.


Dowswell, C. R. et al. (1996), Maize in the third world.

Westview Press, Colorado, US.


Duvick, D. (2005), The contribution of breeding to yield

advances in maize (Zea mays L.), Advances in Agronomy,


vol. 86, pp. 83145.
Follett, R. (2001), Soil management concepts and carbon

sequestration in cropland soils, Soil and Tillage Research,


vol. 61, pp. 7792.
Food and Agriculture Organization of the United Nations

(2010), Statistics division: global maize production, FAOSTAT.


Available from: http://faostat.fao.org/ [accessed February 2014].
Furlan, L., Canzi, S., Di Bernardo, A. & Edwards, C. R. (2006),

The ineffectiveness of insecticide seed coatings and plantingtime soil insecticides as Diabrotica virgifera virgifera LeConte
population suppressors, Journal of Applied Entomology,
vol. 130, pp. 485490.
Gentry, L. E., Below, F. E., David, M. B. & Bergerou, J. A. (2001),

Source of the soybean N credit in maize production, Plant and


Soil, vol. 236, pp. 175184.
Glaszmann, J. et al. (1997), Comparative genome analysis

between several tropical grasses, Euphytica, vol. 96, pp. 1321.


Gollehon, N. & Quinby, W. (2006), Irrigation resources and

water costs, in Agricultural resources and environmental


indicators. US Department of Agriculture, Economic Research
Service, Washington DC.
Hertel, T. W. et al. (2010), Effects of US maize ethanol on global

land use and greenhouse gas emissions: estimating marketmediated responses, BioScience, vol. 60, pp. 223231.

104 | 6 Biomass feedstock crops

Hetherington, P. R., Reynolds, T. L., Marshall, G. & Kirkwood,

R. C. (1999), The absorption, translocation and distribution of


the herbicide glyphosate in maize expressing the CP-4
transgene, Journal of Experimental Botany, vol. 50,
pp. 15671576.
Hussain, I. (1999), Impacts of tillage and no-till on production of

maize and soybean on an eroded Illinois silt loam soil, Soil and
Tillage Research, vol. 52, pp. 3749.
Kanampiu, F. K. et al. (2003), Multi-site, multi-season field tests

demonstrate that herbicide seed-coating herbicide-resistance


maize controls Striga spp. and increases yields in several
African countries, Crop Protection, vol. 22 pp. 697706.
Kim, S. & Dale, B. E. (2005), Environmental aspects of ethanol

derived from no-tilled corn grain: nonrenewable energy


consumption and greenhouse gas emissions, Biomass and
Bioenergy, vol. 28, pp. 475489.
Kleter, G. & Bhula, R. (2007), Altered pesticide use on

transgenic crops and the associated general impact from an


environmental perspective, Pest Management Science,
vol. 1115, pp. 11071115.
Lal, R. (2004), Soil carbon sequestration impacts on global

climate change and food security, Science, vol. 304,


pp. 16231627.
Leakey, A. D. B. et al. (2004), Will photosynthesis of maize

(Zea mays) in the US Corn Belt increase in future [CO2] rich


atmospheres? An analysis of diurnal courses of CO2 uptake
under free-air concentration enrichment (FACE), Global
Change Biology, vol. 10, pp. 951962.
Li, L., Sun, J., Zhang, F., Li, X., Yang, S. & Rengel, Z. (2001),

Wheat/maize or wheat/soybean strip intercropping I. yield


advantage and interspecific interactions on nutrients, Field
Crops Research, vol. 71, pp. 123137.
Liska, A. J. et al. (2009), Improvements in life cycle energy

efficiency and greenhouse gas emissions of corn-ethanol,


Journal of Industrial Ecology, vol. 13 pp. 5874.
Maskina, M. S., Baddesha, H. S. & Meelu, O. P. (1988),

Fertilizer requirement of rice-wheat and maize-wheat rotations


on coarse-textured soils amended with farmyard manure,
Fertilizer Research, vol. 17, pp. 153164.
Meja, D. (2003), Maize: post-harvest operation. Food and

Agriculture Organization of the United Nations, AGST.


Morey, R. V., Peart, R. M. & Zachariah, G. L. (1972), Optimal

harvest policies for corn and soybeans, Journal of Agricultural


Engineering Research, vol. 17 pp. 139148.
Musick, J. T. & Dusek, D. A. (1980), Irrigated corn yield

response to water, Transactions of the ASABE, vol. 23,


pp. 9298.
Musser, F. & Shelton, A. (2003), Bt sweet corn and selective

insecticides: impacts on pests and predators, Journal of


Economic Entomology, vol. 96, pp. 7180.
National Corn Growers Association (2012a), Biotech seed

products available for the 2012 planting season.


National Corn Growers Association (2012b), Corn rooted in

human history. Available from: http://www.ncga.com/uploads/


useruploads/woc_2012.pdf [accessed February 2014].
Oberle, S. L. & Keeney, D. R. (1990), Factors influencing corn

fertilizer N requirements in the northern US Corn Belt, Journal


of Production Agriculture, vol. 3, pp. 527534.
Oerke, E. -C. (2005), Crop losses to pests, The Journal of

Agricultural Science, vol. 144, pp. 3143.


Pandey, R. K., Maranville, J. W. & Admou, A. (2000), Deficit

irrigation and nitrogen effects on maize in a Sahelian


environment. 1. Grain yield and yield components,
Agricultural Water Management, vol. 46, pp. 113.

Petrolia, D. R. (2008), The economics of harvesting and

transporting corn stover for conversion to fuel ethanol: a case


study for Minnesota, Biomass and Bioenergy, vol. 32,
pp. 603612.
Pingali, P. & Pandey, S. (2001), Meeting world maize needs:

technological opportunities and priorities for the public sector.


International Maize and Wheat Improvement Center.
Ragauskas, A. J. et al. (2006), The path forward for biofuels and

biomaterials, Science, vol. 311, pp. 484489.


Stewart, D. W., Dwyer, L. M. & Carrigan, L. L. (1998),

Phenological temperature response of maize, Agronomy


Journal, vol. 90, pp. 7379.
Tyner, W. (2008), The US ethanol and biofuels boom: its origins,

Cheavegatti-Gianotto, A. et al. (2011), Sugarcane (Saccharum X

officinarum): a reference study for the regulation of genetically


modified cultivars in Brazil, Tropical Plant Biology,
vol. 4, pp. 6289.
Cheeseman, O. (2004), Environmental impacts of sugar

production. CABI, UK.


Chona, B. L. (1956), Chairmans address, Pathology section,

Proceedings of the International Society of Sugar Cane


Technologists, vol. 9, pp. 975986.
Conpanhia Ambiental do Estado de Sao Paulo (2006), Norma

CETESB P4.231 Vinhaca criterios e procedimentos para


aplicacao no solo agrcola. Sao Paulo.
Dal-Bianco, M. et al. (2012), Sugarcane improvement: how

current status, and future prospects, BioScience, vol. 58,


pp. 646653.

far can we go?, Current Opinion in Biotechnology, vol. 23,


pp. 265270.

US Department of Agriculture (2012a), Conservation plant

Daniels, J. & Roach, B. (1987), Taxonomy and evolution, in

characteristics: Zea mays L.


US Department of Agriculture (2012b), Fertilizer use: corn,

Agricultural Resource Management Survey.


US Department of Agriculture, Economic Research Service,

Corn background data. Available from: http://www.ers.usda.


gov/topics/crops/corn/background.aspx#.U5hBiRZV8ds
[accessed February 2014].
Wang, M. (2005), Updated energy and greenhouse gas emission

results of fuel ethanol. Presented at the 15th International


Symposium on Alcohol Fuels.
Wiatrak, P. & Frederick, J. R., Corn production guide: harvest,

Clemson University, Cooperative Extension.

Sugarcane
Agnihotri, S. & Dutt, D. (2010), Complete characterization of

bagasse of early species of Saccharum officinerum-CO 89003 for


pulp and paper making, BioResources, vol. 2016, pp. 11971214.
Almazan, O. et al. (2001), The sugarcane, its by-products and

co-products, Sugar Cane International, July, pp. 38.


Andrietta, M. et al. (2007), Bioethanol Brazil, 30 years of

Prolcool, International Sugar Journal, vol. 109, pp. 195200.


Anonymous (1997), Herbicide guide. SASEX, Durban.
Bailey, R. A. (2004), Diseases, in James, G. (ed.) Sugarcane.

Blackwell Publishing Ltd, Oxford.


Bailey, R. A. & Bechet, G. R. (1997), Further evidence of the

effects of ratoon stunting disease on production under irrigated


and rainfed conditions, Proceedings of the South African Sugar
Technologists Association, pp. 97101.
Bailey, R. A. & McFarlane, S. A. (1999), The incidence and effects

of ratoon stunting disease of sugarcane in southern and central


Africa, Proceedings of the International Society of Sugar Cane
Technologists, vol. 23, pp. 338346.
Barnes, A. (1974), The sugar cane, 2nd edition. Hill Books,

London.
Bell, M. J. & Garside, A. L. (2005), Shoot and stalk dynamics

and the yield of sugarcane crops in tropical and subtropical


Queensland, Australia, Field Crops Research, vol. 92,
pp. 231248.
Berding, N. & Roach, B. (1987), Germplasm collection,

maintenance, and use, in Heinz, D. J. (ed.) Sugarcane


improvement through breeding. Elsevier Press, Amsterdam.
Black, C. & Chen, T. (1969), Biochemical basis for plant

competition, Weed Science, vol. 17, pp. 338344.


Cardona, C. A. et al. (2010), Production of bioethanol from

sugarcane bagasse: status and perspectives, Bioresource


Technology, vol. 101, pp. 47544766.
Charernsom, K. & Suasa-ard, D. W. (1994), Pest status, biology

and control measures for soil pest of sugar cane in South East
Asia, in Conlong, A. J. M. & Carnegie, D. E. (eds) Proceedings of
the International Society of Sugar Cane Technologists Second
Entomology Workshop. SASA Experiment Station, Durban.

Heinz, D. J. (ed.) Sugarcane improvement through breeding.


Elsevier Press, Amsterdam.
Denmead, O. T. et al. (2010), Emissions of methane and nitrous

oxide from Australian sugarcane soils, Agricultural and Forest


Meteorology, vol. 150, pp. 748756.
Doorenbos, J. & Kassam, A. (1979), Efeito da agua no

rendimento das culturas (Irrigacao e Drenagem, 33). Food and


Agriculture Organization of the United Nations.
Ellis, R. (2004), Sugarcane agriculture, in James, G. (ed.)

Sugarcane. Blackwell Publishing Ltd, Oxford.


Fauconnier, R. & Bassereau, D. (1970), Techniques agricoles

et productions tropicales, in La Canne Sucre. Maisonneuve


et Larose, Paris.
Feigin, A., Letey, J. & Jarrell, W. M. (1982), N utilization

efficiency by drip irrigated celery receiving preplant or water


applied N fertilizer, Agronomy Journal, pp. 978983.
Ferrer, W. F. (1994), Soil-inhabiting insect pests and nematodes

that damage sugar cane in South America: biology and control


measures, in Conlong, A. J. M. & Carnegie, D. E. (eds)
Proceedings of the International Society of Sugar Cane
Technologists Second Entomology Workshop. SASA Experiment
Station, Durban.
Food and Agriculture Organization of the United Nations (2010),

Statistics division: world sugarcane production, FAOSTAT.


Available from: http://faostat.fao.org/ [accessed February 2014].
Goldemberg, J., Coelho, S. T., Nastari, P. M. & Lucon, O. (2004),

Ethanol learning curve the Brazilian experience, Biomass and


Bioenergy, vol. 26, pp. 301304.
Hartemink, A. (2008), Sugarcane for bioethanol: soil and

environmental issues, Advances in Agronomy, vol. 99,


pp. 125182.
Hemwong, S. et al. (2008), Dynamics of residue decomposition

and N2 fixation of grain legumes upon sugarcane residue


retention as an alternative to burning, Soil and Tillage Research,
vol. 99, pp. 8497.
Hotta, C. T. et al. (2010), The biotechnology roadmap for

sugarcane improvement, Tropical Plant Biology, vol. 3,


pp. 7587.
Humbert, R. (1968), The growing of sugar cane, 1st edition.

Elsevier Press, Amsterdam.


Ingram, K. T. & Hilton, H. W. (1986), Nitrogen-potassium

fertilization and soil moisture effects on growth and


development of drip-irrigated sugarcane, Crop Science,
vol. 26, pp. 10341039.
Irvine, J. (2004). Sugarcane agronomy, in James, G. (ed.)

Sugarcane. Blackwell Publishing Ltd, Oxford.


James, G. (2004), An introduction to sugarcane, in James, G.

(ed.) Sugarcane. Blackwell Publishing Ltd, Oxford.


Kira, M. T. & El-Sherif, H. (1971), Estimation of losses in cane

and sugar yields caused by infestations of Chilo agamemnon


Bles, Proceedings of the International Society of Sugar Cane
Technologists, vol. 14, pp. 427428.

6 Biomass feedstock crops | 105

Lefebvre, L. W., Ingram, C. R. & Yang, M. C. (1978), Assessment

Weekes, D. (2004), Harvest management, in James, G. (ed.)

of rat damage to Florida sugarcane in 1975, Proceedings of the


American Sugar Technologists Association, vol. 7, pp. 7580.

Wegener, M. (1990), Analysis of risk in irrigated sugarcane at

Lesie, G. (2004), Pests of sugarcane, in James, G. (ed.)

Sugarcane. Blackwell Publishing Ltd, Oxford.


Lima, E., Boddey, R. M. & Dbereiner, J. (1987), Quantification

of biological nitrogen fixation associated with sugar cane using a


15N aided nitrogen balance, Soil Biology and Biochemistry,
vol. 19, pp. 165170.
Lisboa, C. C. et al. (2011), Bioethanol production from sugarcane

and emissions of greenhouse gases known and unknowns,


GCB Bioenergy, vol. 3, pp. 277292.
Long, W. (1972), Insect pests of sugar cane, Annual Review of

Entomology, vol. 17, pp. 149176.


Macedo, I. C., Seabra, J. E. & Silva, J. E. (2008), Greenhouse gases

emissions in the production and use of ethanol from sugarcane


in Brazil: the 2005/2006 averages and a prediction for 2020,
Biomass and Bioenergy, vol. 32, pp. 582595.
Martinez, A. (1997), in Leslie, G. W. & Keeping, M. G. (eds)

Report on the Third International Society of Sugar Cane


Technologists Entomology Workshop, Culiacan, Mexico. SASA
Experiment Station Internal Report, Durban.
Mascarehas, H. et al. (1994), Efeito residual do adubo aplicado na

soja (Glycine max (L.) Merrill) sobre a cana-de-acar


(Saccharum spp.), Scientia Agricola, vol. 51, no. 2, pp. 264269.
Meade, G. P. & Chen, J. C. (1977), Cane sugar handbook. John

Williamson Ltd, New York/London.


Moreira, J. R. & Goldemberg, J. (1999), The alcohol program,

Energy Policy, vol. 27, pp. 229245.


National Center for Biotechnology Information (2012),

Taxonomy database.
Oliveira, J. et al. (2001), Soil temperature in a sugar-cane crop as

a function of the management system, Plant and Soil, vol. 230,


pp. 6166.
Pandey, A., Soccol, C. R., Nigam, P. & Soccol, V. T. (2000),

Biotechnological potential of agro-industrial residues. I:


sugarcane bagasse, Bioresource Technology, vol. 74, pp. 6980.
Rajabalee, A. (1990), Management of Chilo spp. on sugarcane

with notes on mating disruption studies with the synthetic sex


pheromone of C. sacchariphagus in Mauritius, International
Journal of Tropical Insect Science, vol. 11, Special issue 45,
pp. 825836.
Renewable Fuels Association (2012), Accelerating industry

innovation: 2012 ethanol industry outlook.


Russell, J. (1990), Effect of interactions between available water

and solar radiation on sugarcane productivity and the impact of


climatic change, Proceedings of Australian Society Sugar Cane
Technologists, vol. 12, pp. 2938.
Sosa, O. J. (1994), Soil pests of sugarcane in North America, in

Conlong, A. J. M. & Carnegie, D. E. (eds) Proceedings of the


International Society of Sugar Cane Technologists Second
Entomology Workshop. SASA Experiment Station, Durban.
South African Sugar Association Experimental Station (1979),

Smut and cane yield, in annual report 197879, pp. 71.


Srivastava, S. & Suarez, N. (1992), Sugarcane, in Wichmann, W.

(ed.) World fertilizer use manual. BASF AG, Germany.


Thompson, G. D. (1976), Water use by sugarcane, South African

Sugar Journal, vol. 60, pp. 627635.


United Nations Environmental Programme (2001), Stockholm

convention on persistent organic pollutants. UNEP, Geneva.


Viswanathan, R. & Padmanaban, P. (2008), Hand book on

sugarcane diseases and their management. Sugarcane Breeding


Institute, Coimbatore, India.
Viswanathan, R. & Rao, G. P. (2011), Disease scenario and

management of major sugarcane diseases in India, Sugar Tech,


vol. 13, pp. 336353.

106 | 6 Biomass feedstock crops

Sugarcane. Blackwell Publishing Ltd, Oxford.


Mackay, Proceedings of Australian Society Sugar Cane
Technologists, vol. 12, pp. 4551.

Switchgrass
Anderson, G. Q. A. & Fergusson, M. J. (2006), Energy from

biomass in the UK: sources, processes and biodiversity


implications, Ibis, vol. 148, pp. 180183.
Balan, V. et al. (2012), Biochemcial and thermochemical

conversion of switchgrass to biofuels, in Monti, A. (ed.)


Switchgrass: a valuable biomass crop for energy. Springer,
London.
Brejda, J. J. (2000), Fertilization of native warm-season grasses,

in Moore, K. J. & Anderson, B. E. (eds) Native warm-season


research trends and issues, CSSA Spec. Publ. 30, pp. 177200.
Burns, J. C., Mochrie, R. D. & Timothy, D. H. (1984), Steer

performance from two perennial pennisetum species,


switchgrass, and a fescue coastal bermudagrass system,
Agronomy Journal, vol. 76, pp. 795800.
Casler, M. D. (2012), Switchgrass breeding, genetics, and

genomics, in Monti, A. (ed.) Switchgrass: a valuable biomass


crop for energy. Springer, London.
Davis, S. C. et al. (2010), Comparative biogeochemical cycles of

bioenergy crops reveal nitrogen-fixation and low greenhouse gas


emissions in a Miscanthus giganteus agro-ecosystem,
Ecosystems, vol. 13, pp. 144156.
Davis, S. C. et al. (2012), Impact of second-generation biofuel

agriculture on greenhouse-gas emissions in the corn-growing


regions of the US, Frontiers in Ecology and the Environment,
vol. 10, pp. 6974.
Deevey Jr, E. S. (1949), Biogeography of the Pleistocene: part I:

Europe and North America, Geological Society of America


Bulletin, vol. 60, pp. 13151416.
Duffy, M. D. & Nanhou, V. Y. (2002), Costs of producing

switchgrass for biomass in southern Iowa, in Janick, J. &


Whipkey, A. (eds) Trends in new crops and new uses. ASHS
Press, Alexandria, VA, US.
Fike, J. H. et al. (2006), Switchgrass production for the upper

southeastern USA: influence of cultivar and cutting frequency


on biomass yields, Biomass and Bioenergy, vol. 30, pp. 207213.
Fu, C. et al. (2011), Genetic manipulation of lignin reduces

recalcitrance and improves ethanol production from


switchgrass, Proceedings of the National Academy of Sciences
of the United States of America, vol. 108, pp. 38033808.
Garrett, K. A. et al. (2004), Barley yellow dwarf disease in natural

populations of dominant tallgrass prairie species in Kansas,


Plant Disease, vol. 88, pp. 574.
Gravert, C. E., Tiffany, L. H. & Munkvold, G. P. (2000), Outbreak

of smut caused by Tilletia maclaganii on cultivated switchgrass


in Iowa, Plant Disease, vol. 84, p. 596.
Heaton, E. A., Dohleman, F. G. & Long, S. P. (2008), Meeting US

biofuel goals with less land: the potential of Miscanthus, Global


Change Biology, vol. 14, pp. 20002014.
Hickman, G. C., Vanloocke, A., Dohleman, F. G. & Bernacchi,

C. J. (2010), A comparison of canopy evapotranspiration for


maize and two perennial grasses identified as potential
bioenergy crops, GCB Bioenergy, vol. 2, pp. 157168.
Jimmy Carter Plant Materials Center (2011), Plant fact sheet for

switchgrass (Panicum virgatum L.). US Department of


Agriculture, Natural Resources Conservation Service.
Khor, A. et al. (2007), Straw combustion in a fixed bed

combustor, Fuel, vol. 86, pp. 152160.


Lewandowski, I., Scurlock, J. M. O., Lindvall, E. & Christou, M.

(2003), The development and current status of perennial


rhizomatous grasses as energy crops in Europe and the US,
Biomass and Bioenergy, vol. 25, pp. 335361.

McLaughlin, S. B. & Adams Kszos, L. (2005), Development of

switchgrass (Panicum virgatum) as a bioenergy feedstock in the


United States, Biomass and Bioenergy, vol. 28, pp. 515535.
McLaughlin, S. B. & Kiniry, J. (2006), Projecting yield and

utilization potential of switchgrass as an energy crop, Advances


in Agronomy, vol. 90, pp. 267297.
Meyer, M. H., Paul, J. & Anderson, N. O. (2010), Competitive

ability of invasive Miscanthus biotypes with aggressive


switchgrass, Biological Invasions, vol. 12, pp. 38093816.
Monti, A. (ed.) (2012), Switchgrass: a valuable biomass crop for

energy. Springer, London.


Morris, G. P., Grabowski, P. P. & Borevitz, J. O. (2011), Genomic

diversity in switchgrass (Panicum virgatum): from the


continental scale to a dune landscape, Molecular Ecology,
vol. 20, pp. 49384952.
Muir, J. et al. (2001), Biomass production of Alamo switchgrass

in response to nitrogen, phosphorus, and row spacing, Agronomy


Journal, vol. 901, pp. 896901.
Ocumpaugh, W. et al. (2002), Evaluation of switchgrass cultivars

and cultural methods for biomass production in the south


central US. Consolidated report 2002, prepared by Oak Ridge
National Laboratory.
Parrish, D. J. et al. (2012), The evolution of switchgrass as an

energy crop, in Monti, A. (ed.) Switchgrass: a valuable biomass


crop for energy. Springer, London.
Parrisha, D. J. & Fikea, J. H. (2007), The biology and agronomy of

switchgrass for biofuels, Critical Reviews in Plant Sciences,


vol. 24, pp. 423459.
Popp, M. & Hogan Jr, R. (2007), Assessment of two alternative

switchgrass harvest and transport methods. Farm Foundation


Bioenergy Conference, St Louis, Missouri, US.
Sanderson, M. A., Adler, P. R., Boateng, A. A., Casler, M. D. &

Sarath, G. (2006), Switchgrass as a biofuels feedstock in the USA,


Canadian Journal of Plant Science, vol. 86, pp. 13151325.
Sanderson, M. A. et al. (2012), Switchgrass crop management,

in Monti, A. (ed.) Switchgrass: a valuable biomass crop for


energy. Springer, London.
Sanderson, M. A. & Reed, R. (1996), Switchgrass as a sustainable
bioenergy crop, Bioresource Technology, vol. 56, pp. 8393.
Schmer, M. R., Vogel, K. P., Mitchell, R. B. & Perrin, R. K. (2008),

Net energy of cellulosic ethanol from switchgrass, Proceedings


of the National Academy of Sciences of the United States of
America, vol. 105, pp. 464469.
Smeets, E. M., Lewandowski, I. M. & Faaij, A. P. (2009), The

economical and environmental performance of miscanthus and


switchgrass production and supply chains in a European setting,
Renewable and Sustainable Energy Reviews, vol. 13,
pp. 12301245.
Spatari, S., Zhang, Y. & MacLean, H. L. (2005), Life cycle
assessment of switchgrass- and corn stover-derived ethanolfueled automobiles, Environmental Science & Technology,
vol. 39, pp. 97509758.
Thomason, W. E. et al. (2005), Switchgrass response to harvest

frequency and time and rate of applied nitrogen, Journal of Plant


Nutrition, vol. 27, pp. 11991226.
Tilman, D., Hill, J. & Lehman, C. (2006), Carbon-negative

biofuels from low-input high-diversity grassland biomass,


Science, vol. 314, pp. 15981600.
Vogel, K. P. (2003), Genetic variation among switchgrasses for

agronomic traits, forage quality, and biomass fuel production:


final report for project DE-A105-900R2194. Agricultural
Research Service, US Department of Agriculture.
Vogel, K. P. (2004), Switchgrass, in Moser, L. E. et al. (eds)

Warm-season (C4) grasses. American Society of Agronomy,


Crop Science Society of America and Soil Science Society of
America, Madison, WI.

Vogel, K. P., Brejda, J. J., Walters, D. T. & Buxton, D. R. (2002),

Switchgrass biomass production in the Midwest USA: harvest


and nitrogen management, Agronomy Journal, vol. 94,
pp. 413420.
Wolf, D. D. & Fiske, D. A. (1995), Planting and managing

switchgrass for forage, wildlife, and conservation. Virginia


Cooperative Extension, Virginia Polytechnic Institute and
Virginia State University, Blacksburg, VA, US.
Xi, Y. et al. (2009), Agrobacterium-mediated transformation of

switchgrass and inheritance of the transgenes, BioEnergy


Research, vol. 2, pp. 275283.
Xu, B., Shan, L., Zhang, S., Deng, X. & Li, F. (2008), Evaluation

of switchgrass and sainfoin intercropping under 2:1 rowreplacement in semiarid region, northwest China, African
Journal of Biotechnology, vol. 7, pp. 40564067.
Zan, C. S., Fyles, J. W., Girouard, P. & Samson, R. A. (2001),

Carbon sequestration in perennial bioenergy, annual corn and


uncultivated systems in southern Quebec, Agriculture,
Ecosystems & Environment, vol. 86, pp. 135144.
Zegada-Lizarazu, W., Wullschleger, S. D., Nair, S. S. & Monti, A.

(2012), Crop physiology, in Monti, A. (ed.) Switchgrass: a


valuable biomass crop for energy. Springer, London.

Miscanthus
Ahonsi, M. O. et al. (2010), First report of Pithomyces chartarum

causing a leaf blight of Miscanthus giganteus in Kentucky,


Plant Disease, vol. 94, pp. 480.
Anderson, G. Q. A. & Fergusson, M. J. (2006), Energy from

biomass in the UK: sources, processes and biodiversity


implications, Ibis, vol. 148, pp. 180183.
Angelini, L. G., Ceccarini, L., Nassi o Di Nasso, N. & Bonari, E.

(2009), Comparison of Arundo donax L. and Miscanthus x


giganteus in a long-term field experiment in central Italy:
analysis of productive characteristics and energy balance,
Biomass and Bioenergy, vol. 33, pp. 635643.
Beale, C. & Long, S. (1995), Can perennial C4 grasses attain high

efficiencies of radiant energy conversion in cool climates?,


Plant, Cell & Environment, vol. 18, pp. 641650.
Beale, C. & Long, S. (1997), Seasonal dynamics of nutrient

accumulation and partitioning in the perennial C4-grasses


Miscanthus giganteus and Spartina cynosuroides, Biomass
Bioenergy, vol. 12, pp. 419428.
Bradshaw, J. D., Prasifka, J. R., Steffey, K. L. & Gray, M. E. (2010),

First report of field populations of two potential aphid pests of


the bioenergy crop Miscanthus giganteus, Florida
Entomologist, vol. 93, pp. 135137.
Bullard, M. & Metcalfe, P. (2001), Estimating the energy

requirements and CO2 emissions from production of the


perennial grasses Miscanthus, switchgrass and reed canary
grass. ADAS Consulting Ltd, London.
Christian, D. G. & Haase, E. (2001), Agronomy of Miscanthus,

in Jones, M. B. & Walsh, M. (eds) Miscanthus for energy and


fibre. James & James, London.
Christian, D. G., Bullard, M. J. & Wilkins, C. (1997),

The agronomy of some herbaceous crops grown for energy in


southern England, Aspects of Applied Biology, vol. 49,
pp. 4151.
Christian, D. G., Lamptey, J. N. L., Forde, S. M. D. & Plumb, R.

T. (1994), First report of barley yellow dwarf luteovirus on


Miscanthus in the United Kingdom, European Journal of Plant
Pathology, vol. 100, pp. 167170.
Christian, D. G., Riche, A. B. & Yates, N. E. (2008), Growth,

yield and mineral content of Miscanthus giganteus grown as a


biofuel for 14 successive harvests, Industrial Crops and
Products, vol. 28, pp. 320327.
Clair, S. S., Hillier, J. & Smith, P. (2008), Estimating the pre-

harvest greenhouse gas costs of energy crop production,Biomass


and Bioenergy, vol. 32, pp. 442452.

6 Biomass feedstock crops | 107

Clayton, W. D. & Renvoize, S. A. (1986), Genera graninum:

grasses of the world. HMSO, London.


Clifton-Brown, J. et al. (2001), Performance of 15 Miscanthus

genotypes at five sites in Europe, Agronomy Journal, vol. 93,


pp. 10131019.
Clifton-Brown, J. & Lewandowski, I. (1998), Frosttoleranz der

Rhizome verschiedener Miscanthus Genotypen, Mitteilungen


der Gesellschaft fuer PTanzenbauwissenschaften, vol. 11,
pp. 225226.
Clifton-Brown, J. & Lewandowski, I. (2000), Overwintering

problems of newly established Miscanthus plantations can be


overcome by identifying genotypes with improved rhizome cold
tolerance, New Phytologist, vol. 148, pp. 287294.
Danalatos, N. G., Archontoulis, S. V. & Mitsios, I. (2007),

Potential growth and biomass productivity of Miscanthus x


giganteus as affected by plant density and N-fertilization in
central Greece, Biomass and Bioenergy, vol. 31, pp. 145152.
Davis, S. C. et al. (2010), Comparative biogeochemical cycles of

bioenergy crops reveal nitrogen-fixation and low greenhouse gas


emissions in a Miscanthus giganteus agro-ecosystem,
Ecosystems, vol. 13, pp. 144156.
Davis, S. C. et al. (2012), Impact of second-generation biofuel

agriculture on greenhouse-gas emissions in the corn-growing


regions of the US, Frontiers in Ecology and the Environment,
vol. 10, pp. 6974.
Dohleman, F. G., Heaton, E. A., Arundale, R. A. & Long, S. P.

(2012), Seasonal dynamics of above- and below-ground biomass


and nitrogen partitioning in Miscanthus giganteus and
Panicum virgatum across three growing seasons, GCB
Bioenergy, vol. 4, pp. 534544.
Dubeux, J. et al. (2007), Nutrient cycling in warm-climate

grasslands, Crop Science, vol. 47, pp. 915928.


Engler, D. & Chen, J. (2011), Transformation and trait

modification in Miscanthus species. Patent US2011/0047651


A1. 11.
Eppel-Hotz, A. et al. (1998), Miscanthus: new cultivations and

results of research experiments for improving the establishment


rate, in Kopetz, H. et al. (eds) Biomass for energy and industry:
proceedings of the 10th European conference, Wurzburg,
Germany. Rimpar, Germany.
Ercoli, L. & Mariotti, M. (1999), Effect of irrigation and nitrogen

fertilization on biomass yield and efficiency of energy use in


crop production of Miscanthus, Field Crops Research, vol. 63,
pp. 311.
Farrell, A. D., Clifton-Brown, J. C., Lewandowski, I. & Jones,

M. B. (2006), Genotypic variation in cold tolerance influences


the yield of Miscanthus, Annals of Applied Biology, vol. 149,
pp. 337345.
FP7 OPTIMISC (2012), Uses of Miscanthus. Available from:

http://optimisc.anna-consult.de/index.php?option=com_
content&view=article&id=252:uses-of-miscanthus&catid=62&Itemid=286 [accessed June 2014].
Greef, J. M. & Deuter, M. (1993), Syntaxonomy of Miscanthus x

giganteus Greef et Deu, Angewandte Botanik, vol. 67, pp. 8790.


Hastings, A. et al. (2009), Future energy potential of Miscanthus

in Europe, GCB Bioenergy, vol. 1, pp. 180196.


Heaton, E. A. et al. (2004), Miscanthus for renewable energy

generation: European Union experience and projections for


Illinois, Mitigation and Adaptation Strategies for Global
Change, vol. 9, pp. 433451.
Heaton, E. A., Dohleman, F. G. & Long, S. P. (2008), Meeting US

biofuel goals with less land: the potential of Miscanthus, Global


Change Biology, vol. 14, pp. 20002014.
Hodkinson, T. R., Chase, M. W. & Renovoize, S. A. (2002),

Characterization of a genetic resource collection for Miscanthus


(Saccharinae, Andropogoneae, Poaceae) using AFLP and ISSR
PCR, Annals of Botany, vol. 89, pp. 627636.

Hodkinson, T. R., Renvoize, S. A. & Chase, M. W. (1997),

Systematics of Miscanthus, Aspects of Applied Biology, vol. 49,


pp. 189198.
Joshua, T. & Vicki, W. (2007), Devon Miscanthus and woodfuels

opportunities statement. Report to Devon Wildlife Trust.


Centre for Sustainable Energy.
Lewandowski, I. & Clifton-Brown, J. (2000). Miscanthus:

European experience with a novel energy crop, Biomass and


Bioenergy, vol. 19, pp. 209227.
Lewandowski, I., Scurlock, J. M., Lindvall, E. & Christou, M.

(2003), The development and current status of perennial


rhizomatous grasses as energy crops in the US and Europe,
Biomass and Bioenergy, vol. 25, pp. 335361.
Long, S. P. (1983), C4 photosynthesis at low temperatures, Plant

Cell and Environment, vol. 6, pp. 345363.


Mekete, T. et al. (2011), Distribution and diversity of root-

lession nematode (Pratylenchus spp.) associated with


Miscanthus giganteus and Panicum virgatum used for
biofuels, and species identification in a multiplex polymerase
chain reaction, Nematology, vol. 13, pp. 673686.
Meyer, M. H., Paul, J. & Anderson, N. O. (2010), Competitive

ability of invasive Miscanthus biotypes with aggressive


switchgrass, Biological Invasions, vol. 12, pp. 38093816.
Miguez, F. E., Villamil, M. B., Long, S. P. & Bollero, G. A. (2008),

Meta-analysis of the effects of management factors on


Miscanthus giganteus growth and biomass production,
Agricultural and Forest Meteorology, vol. 148, pp. 12801292.
Nielsen, P. N. (1987), The productivity of the Miscanthus

cultivar Giganteus, Tidsshr Planteavl, vol. 91, pp. 361368.


ONeill, N. R. & Farr, D. F. (1996), Miscanthus blight, a new

foliar disease of ornamental grasses and sugarcane incited by


Leptosphaeria sp. and its anamorphic state Stagonospora sp.,
Plant Disease, vol. 80, pp. 980987.
Ohwi (1964), Flora of Japan. Smithsonian Institute,

Washington DC.
Prasifka, J. R., Bradshaw, J. D. & Gray, M. E. (2012), Potential

biomass reductions to Miscanthus giganteus by stem-boring


caterpillars, Environmental Entomology, vol. 41, pp. 865871.
Remlein-Starosta, D. (2007), Diseases of bioenergy crops,

Progress in Plant Protection, vol. 47, pp. 351357.


Schwarz, K. U., Greef, J. M. & Schnug, E. Untersuchungen zur

Etablierung und Biomassebildung von Miscanthus giganteus


unter verschiedenen Umweltbedingungen. Braunschweig,
Bundesforschungsanstalt fur Landwirtschaft BraunschweigVolkenrode.
Scown, C. D. et al. (2012), Corrigendum: lifecycle greenhouse

gas implications of US national scenarios for cellulosic ethanol


production, Environmental Research Letters, vol. 7, 019502.
Semere, T. & Slater, F. M. . (2007), Invertebrate populations in

miscanthus (Miscanthus giganteus) and reed canary-grass


(Phalaris arundinacea) fields, Biomass and Bioenergy, vol. 31,
pp. 3039.
Smeets, E. M., Lewandowski, I. M. & Faaij, A. P. (2009),

The economical and environmental performance of miscanthus


and switchgrass production and supply chains in a European
setting, Renewable and Sustainable Energy Reviews, vol. 13,
pp. 12301245.
Styles, D., Thorne, F. & Jones, M. B. (2008), Energy crops in

Ireland: an economic comparison of willow and Miscanthus


production with conventional farming systems, Biomass and
Bioenergy, vol. 32, pp. 407421.
US Department of Agriculture (2011), Environmental

assessment: proposed BCAP giant Miscanthus (Miscanthus


giganteus) establishment and production in Arkansas,
Missouri, Ohio, and Pennsylvania. Biomass Crop Assistance
Program.
Venendaal, R. (1997), European energy crops: a synthesis,

Biomass and Bioenergy, vol. 13, pp. 147185.

108 | 6 Biomass feedstock crops

Oil palm
Basiron, Y. & Salmiah, A. (1994), Potential new value-added

products from palm oil and palm kernel oil, in Chee, K. H. (ed.)
Management for enhanced profitability in plantations.
Incorporated Society of Planters, Kuala Lumpur.
Clay, J. (2004), World agriculture and the environment: a

commodity-by-commodity guide to impacts and practices.


Island Press, Washington DC.
Corley, R. H. V. & Tinker, P. B. (2003a), Care and maintenance of

oil palms, in The oil palm. Blackwell Science Ltd, Oxford, UK.
Corley, R. H. V. & Tinker, P. B. (2003b), Diseases and pests of oil

palm, in The oil palm. Blackwell Science Ltd, Oxford, UK.


Corley, R. H. V. & Tinker, P. B. (2003c), Selection and breeding,

in The oil palm. Blackwell Science Ltd, Oxford, UK.


Corley, R. H. V. & Tinker, P. B. (2003d), Site selection

and preparation, in The oil palm. Blackwell Science Ltd,


Oxford, UK.
Corley, R. H. V. & Tinker, P. B. (2003e), The classification and

morphology of the oil palm, in The oil palm. Blackwell Science


Ltd, Oxford, UK.
Corley, R. H. V. & Tinker, P. B. (2003f), The origin and

development of the oil palm industry, in The oil palm.


Blackwell Science Ltd, Oxford, UK.
Food and Agriculture Organization of the United Nations (2010),

Statistics division: world palm oil production, FAOSTAT.


Available from: http://faostat.fao.org/ [accessed February 2014].
Goh, K. J. (2000), Climatic requirements of the oil palm for high

yields, in Goh, K. J. (ed.) Managing oil palm for high yields:


agronomic principles. Malaysian Society of Soil Science and
Param Agricultural Soil Surveys, Kuala Lumpur.
Goh, K. J. et al. (1994), Maximising and maintaining oil palm

yields on commercial scale in Malaysia, in Chee, K. H. (ed.)


Management for enhanced profitability in plantations.
Incorporated Society of Planters, Kuala Lumpur.
Gurmit, S. (1991), Ganoderma: the scourge of oil palms in the

coastal areas, The Planter, vol. 67, pp. 421444.


Hartley, C. W. S. (1988), The oil palm, 3rd edition. Longman,

London.
Hassan O. A., Ishida M., Shukri I. M. & Tajuddin Z. A. (1994),

Oil-palm fronds as a roughage feed source for ruminants in


Malaysia. Malaysia Agriculture Research and Development
Institute, Kuala Lumpur.
Henry, P. (1958), Croissance et dveloppement chez Elaeis

guineensis Jacq. de la germination a la premire floraison,


Revue Gnrale de Botanique, vol. 66, pp. 534.
Henson, I. E. (1999), Comparative ecophysiology of oil palm and

tropical rainforest, in Singh, G. (ed.) Oil palm and the


environment a Malaysian perspective. Malayasian Oil Palm
Growers Council, Kuala Lumpur.
Jalani, B. S. et al. (1999), Systems approach to potential products

from biomass in oil palm. 1999 PORIM International Palm Oil


Conference, Palm Oil Research Institute of Malaysia, Kuala
Lumpur.
Khoo, K. M. (2001), Strategic thrust in addressing current

challenges: the plantation perspective, The Planter, vol. 77,


pp. 405413.
Kushairi, A. & Sambanthamurthi, R. (2006), Breeding and

biotechnology: a powerful value additive for the oil palm


industry. Euro-FedLipid Conference, Madrid, Spain.
Leevijit, T., Tongurai, C., Prateepchaikul, G. &

Wisutmethangoon, W. (2008), Performance test of a 6-stage


continuous reactor for palm methyl ester production,
Bioresource Technology, vol. 99, pp. 21421.
Lim, S. & Teong, L. K. (2010), Recent trends, opportunities and

challenges of biodiesel in Malaysia: an overview, Renewable and


Sustainable Energy Reviews, vol. 14, pp. 938954.

Matthews, G., Wiles, T. & Baleguel, P. (2003), A survey of

pesticide application in Cameroon, Crop Protection, vol. 22,


pp. 707714.
Ng, P. H. C. et al. (1999), Nutrient requirements and

sustainability in mature oil palms an assessment, The Planter,


vol. 75, pp. 331345.
Palat, T. et al. (2000), Irrigation of oil palm in southern Thailand,

Proceedings of the International Planters Conference:


plantation tree crops in the new millennium: the way ahead,
pp. 303315.
Parveez, G. et al. (2000), Transgenic oil palm: production and

projection, Biochemical Society, vol. 2, pp. 969972.


Philippe, R. & Diarrassouba, S. (1979), Method of control of

Coelaenomenodera by introduction of systemic insecticide into


the oil palm trunk, Olagineux, vol. 34, pp. 229233.
Pleanjai, S. & Gheewala, S. H. (2009), Full chain energy analysis

of biodiesel production from palm oil in Thailand, Applied


Energy, vol. 86, pp. S209S214.
Rance, K. A. et al. (2001), Quantitative trait loci for yield

components in oil palm (Elaeis guineensis Jacq.), Theoretical


and Applied Genetics, vol. 103, pp. 13021310.
de Souza, S.P. et al. (2010), Greenhouse gas emissions

and energy balance of palm oil biofuel, Renewable Energy,


vol. 35, pp. 25522561.
Stringfellow, R. (1999), Technological change in the vegetable oil

market: is palm oil being left behind? Oilseeds, oils and meals.
LMC International, Oxford, UK.
Syed, R. A. & Shah, S. (1977), Some important aspects of insect

pest management in oil palm estates in Sabah, Malaysia, in Earp,


D. A. & Newall, W. (eds) International developments in oil
palm. Incorporated Society of Planters, Kuala Lumpur.
Tamunaidu, P. & Bhatia, S. (2007), Catalytic cracking of palm oil

for the production of biofuels: optimization studies, Bioresource


Technology, vol. 98, pp. 35933601.
Thapinta, A. & Hudak, P. F. (1998), Pesticide use and residual

occurrence in Thailand, Environmental Monitoring and


Assessment, vol. 60, pp.103114.
Wood, B. & Corley, R. (1991), The energy balance of oil palm

cultivation. PORIM International Palm Oil Conference


agriculture.
Wood, B. J. & Beattie, T. E. (1981), Processing and marketing of

palm oil, The Planter, vol. 57, pp. 379400.


Yez Angarita, E. E., Silva Lora, E. E., da Costa, R. E. & Torres,

E. A. (2009), The energy balance in the palm oil-derived methyl


ester (PME) life cycle for the cases in Brazil and Colombia,
Renewable Energy, vol. 34, pp. 29052913.
Yee, K. F., Tan, K. T., Abdullah, A. Z. & Lee, K. T. (2009), Life

cycle assessment of palm biodiesel: revealing facts and benefits


for sustainability, Applied Energy, vol. 86, pp. S189S196.
Yusof, B. (2007), Palm oil production through sustainable

plantations, European Journal of Lipid Science and Technology,


vol. 109, pp. 289295.
Yusoff, S. & Hansen, S. B. (2007), Feasibility study of performing

a life cycle assessment on crude palm oil production in


Malaysia, The International Journal of Life Cycle Assessment,
vol. 12, pp. 5058.

Soybean
Adler, P. R., Grosso, S. J. D. & Parton, W. J. (2007), Life-cycle

assessment of net greenhouse-gas flux for bioenergy cropping


systems, Ecological Applications, vol. 17, pp. 675691.
Ahmed, S. & Rao, M. (1982), Performance of maizesoybean

intercrop combination in the tropics: results of a multi-location


study, Field Crops Research, vol. 5, pp. 147161.
Almquist, H. & Mecchi, E. (1942), Soybean protein as a source of

amino acids for the chick, The Journal of Nutrition, vol. 24,
pp. 385392.

6 Biomass feedstock crops | 109

Camp, C. R., Christenbury, G. D. & Doty, C. W. (1988),

Scheduling irrigation for corn and soybean in the southeastern


coastal plain, Transactions of the ASABE, vol. 31, pp. 513518.
Clemente, T. E. (1997), Agrobacterium-mediated

transformation, Journal of Experimental Botany, vol. 48,


pp. 151155.
European Food Safety Authority (2008), Safety and nutritional

assessment of GM plant derived food and feed: the role of animal


feeding trials, Food and Chemical Toxicology, vol. 46,
pp. S2S70.
Fageria, N. K. et al. (2010), Growth and mineral nutrition

of field crops, 2nd edition. CRC Press.


Fangrui, M. & Hanna, M. A. (1999), Biodiesel production:

a review, Bioresource Technology, vol. 70, pp. 115.


Food and Agriculture Organization of the United Nations (2010),

Statistics division: global soybean production, FAOSTAT.


Available from: http://faostat.fao.org/ [accessed February 2014].
Freedman, B. E. H. P., Pryde, E. H. & Mounts, T. L. (1984),

Variables affecting the yields of fatty esters from transesterified


vegetable oils, Journal of the American Oil Chemists Society,
vol. 61, pp. 16381643.
Gentry, L. E., Below, F. E., David, M. B. & Bergerou, J. A. (2001),

Source of the soybean N credit in maize production, Plant and


Soil, vol. 236, pp. 175184.
Ghosh, P. K. & Jayas, D. S. (2010), Storage of soybean, in Singh,

G. (ed.) The soybean: botany, production and uses. CABI, UK.


Graiver, D., Tran, P., Patrick, L., Farminer, K. & Narayan, R.

(2006), Modifications of soybean oil using novel ozone-based


chemistry, ACS Symposium Series, vol. 939, pp. 76100.
Gray, M. & Steffey, K. (1999), Insect pest management for field

and forage crops, in 2008 Illinois Agricultural Pest Management


Handbook. University of Illinois.
Hartman, G. L. & Hill, H. B. (2010), Diseases of soybean and

their management, in Singh, G. (ed.) The soybean: botany,


production and uses. CABI, UK.
Hartman, G. L., Sinclair, J. B. & Rupe, J. C. (1999),

Compendium of soybean diseases, 4th edition. APS Press,


St Paul, Minnesota, US.
Heinemann, A. (2002), Determination of spatial water

requirements at county and regional levels using crop models


and GIS: an example for the State of Parana, Brazil, Agricultural
Water Management, vol. 52, pp. 177196.
Herridge, D. F., Bergersen, F. J. & Peoples, M. B. (1990),

Measurement of nitrogen fixation by soybean in the field using


the ureide and natural N abundance methods, Plant Physiology,
vol. 93, pp. 708716.
Hill, J., Nelson, E., Tilman, D., Polasky, S. & Tiffany, D. (2006),

Environmental, economic, and energetic costs and benefits of


biodiesel and ethanol biofuels, Proceedings of the National
Academy of Sciences of the United States of America, vol. 103,
pp. 1120611210.
Hinchee, M. A. W. et al. (1988), Production of transgenic soybean

plants using Agrobacterium-mediated DNA transfer, Nature


Biotechnology, vol. 6, pp. 915922.
Honary, L. A. T. (1996), An investigation of the use of soybean oil

in hydraulic systems, Bioresource Technology, vol. 56,


pp. 4147.
Hu, Z., Tan, P., Yan, X. & Lou, D. (2008), Life cycle energy,

environment and economic assessment of soybean-based


biodiesel as an alternative automotive fuel in China, Energy,
vol. 33, pp. 16541658.
Huo, H. et al. (2009), Life-cycle assessment of energy use and

greenhouse gas emissions of soybean-derived biodiesel and


renewable fuels, Environmental Science & Technology, vol. 43,
pp. 750756.
Hussain, I. (1999), Impacts of tillage and no-till on production of

maize and soybean on an eroded Illinois silt loam soil, Soil and
Tillage Research, vol. 52, pp. 3749.

110 | 6 Biomass feedstock crops

Hymowitz, T. (1970), On the domestication of the soybean,

Economic Botany, vol. 5, pp. 408421.


Hymowitz, T. & Newell, C. (1981), Taxonomy of the genus

Glycine, domestication and uses of soybeans, Economic Botany,


vol. 35, pp. 272288.
Hyten, D. L. et al. (2006), Impacts of genetic bottlenecks on

soybean genome diversity, Proceedings of the National


Academy of Sciences of the United States of America, vol. 103,
pp. 1666616671.
Islas-Rubio, A. R. & Higuera-Ciapara, I. (2002), Soybeans:

post-harvest operations. Centro de Investigacin y Desarrollo,


A.C. (CIAD), AGSI/FAO.
James, C. (2011), Global status of commercialized biotech/GM

crops: 2011, ISAAA Brief, no. 43.


Johnson, H. W., Borthwick, H. A. & Leffel, R. C. (1960), Effects of

photoperiod and time of planting on rates of development of the


soybean in various stages of the life cycle, Botanical Gazette,
vol. 122, pp. 7795.
Kenar, J. A. (2007), Glycerol as a platform chemical: sweet

opportunities on the horizon?, Lipid Technology, vol. 19,


pp. 249253.
Kikuchi, K. (1999), Use of defatted soybean meal as a substitute

for fish meal in diets of Japanese flounder (Paralichthys


olivaceus), Aquaculture, vol. 179, pp. 311.
Kim, S. & Dale, B. E. (2005), Life cycle assessment of various

cropping systems utilized for producing biofuels: bioethanol


and biodiesel, Biomass and Bioenergy, vol. 29, pp. 426439.
Kim, S. & Dale, B.E. (2009), Regional variations in greenhouse

gas emissions of biobased products in the United States cornbased ethanol and soybean oil, The International Journal of Life
Cycle Assessment, vol. 14, pp. 540546.
Kogan, M. & Turnipseed, S. (1987), Ecology and management of

soybean arthropods, Annual Review of Entomology, vol. 32,


pp. 507538.
Kowalczuk, J. (1996), Mechanization of harvesting and drying of

soybean seeds, Biuletyn Instytutu Hodowlii Aklimatyzacji Ros


lin, vol. 198, pp. 107115.
Kumudini, S. (2010), Soybean growth and development, in

Singh, G. (ed.) The soybean: botany, production and uses.


CABI, UK.
LiJuan, Q. & RuZhen, C. (2010), The origin and history of

soybean, in Singh, G. (ed.) The soybean: botany, production


and uses. CABI, UK.
Mandal, K. et al. (2002), Bioenergy and economic analysis of

soybean-based crop production systems in central India,


Biomass and Bioenergy, vol. 23, pp. 337345.
McLean, E. & Brown, J. (1984), Crop response to lime in the

Midwestern United States, in Adams, F. (ed.) Soil acidity and


liming. American Society of Agronomy, Madison, US.
Mishra, J. S. (2010), Weed management in soybean, in Singh, G.

(ed.) The soybean: botany, production and uses. CABI, UK.


Morey, R. V. et al. (1972), Optimal harvest policies for corn and

soybeans, Journal of Agricultural Engineering Research, vol. 17,


pp. 139148.
Nogueira, L. A. H. (2011), Does biodiesel make sense?, Energy,

vol. 36, pp. 36593666.


Oerke, E. -C. & Dehne, H. -W. (2004), Safeguarding production

losses in major crops and the role of crop protection, Crop


Protection, vol. 23, pp. 275285.
Philbrook, B. D. & Oplinger, E. S. (1989), Soybean field losses as

influenced by harvest delays, Agronomy Journal, vol. 81,


pp. 251258.
Pryor, R. W. et al. (1983), Soybean oil fuel in a small diesel

engine, Transactions of the ASABE, vol. 26, pp. 333342.


Salvagiotti, F. et al. (2008), Nitrogen uptake, fixation and

response to fertilizer N in soybeans: a review, Field Crops


Research, vol. 108, pp. 113.

Shay, E. G. (1993), Diesel fuel from vegetable oils: status and

opportunities, Biomass and Bioenergy, vol. 4, pp. 227242.


Shu, S. Z. et al. (1986), Preliminary study on the evolution of

main traits in soybean, Acta Agronomica Sinica, vol. 4,


pp. 255259.
Singh, G. (ed.) (2010), The soybean: botany, production and uses.

CABI, UK.
Singh, S. & Emden, H. (1979), Insect pests of grain legumes,

Annual Review of Entomology, vol. 81, pp. 107110.


US Department of Agriculture (2004), Argentina soybean

production statistics.
Weber, C. R. (1966), Nodulating and nonnodulating soybean

isolines: II. response to applied nitrogen and modified soil


conditions, Agronomy Journal, vol. 58, pp. 4649.
Wen, Z., Ding, Y., Zhao, T. & Gai, J. (2009), Genetic diversity

and peculiarity of annual wild soybean (G. soja Sieb. et Zucc.)


from various eco-regions in China, Theoretical and Applied
Genetics, vol. 119, pp. 371381.
Whigham, D. K. (1983), Soybean, in Smith, W. H. & Banta, S. J.

(eds) Potential productivity of field crops under different


environments. International Rice Research Institute, Los Baos,
Philippines.
Wood, C. W., Torbert, H. A. & Weaver, D. B. (1993), Nitrogen

Dickmann, D. I. (2001), Poplar culture in North America.

NRC Research Press, Ottawa, Ontario.


Dickmann, D. I. et al. (1996), Effects of irrigation and coppicing

on above-ground growth, physiology, and fine-root dynamics of


two field-grown hybrid poplar clones, Forest Ecology and
Management, vol. 80, pp. 163174.
Djomo, S. N., Kasmioui, O. E., Ceulemans, R. (2011), Energy and

greenhouse gas balance of bioenergy production from poplar and


willow: a review, GCB Bioenergy, vol. 3, pp. 181197.
Eckenwalder, J. E. (1996), Systematics and evolution of

Populus, in Stettler, R. F. et al. (eds) Biology of Populus and its


implications for management and conservation. NRC Research
Press, Ottawa, Ontario.
Edwards, W. R. N. (1986), Precision weighing lysimetry for trees,

using a simplified tared-balance design, Tree Physiology,


vol. 144, pp. 127144.
Fillatti, J. J. et al. (1987), Agrobacterium mediated

transformation and regeneration of Populus, MGG Molecular &


General Genetics, vol. 206, pp. 192199.
Fischer, G., Prieler, S. & van Velthuizen, H. (2005), Biomass

potentials of miscanthus, willow and poplar: results and policy


implications for Eastern Europe, Northern and Central Asia,
Biomass and Bioenergy, vol. 28, pp. 119132.

fertilizer effects on soybean growth, yield, and seed composition,


Journal of Production Agriculture, vol. 6, pp. 354360.

Going, M. (1903), With the trees. Baker & Taylor, New York.

Wrather, J. A. et al. (1994), Soybean disease loss estimates for the

southeastern North America: evidence from the fossil plant


record, Evolution, vol. 18, pp. 571585.

top 10 soybean producing countries in 1994, Plant Disease,


vol. 81, pp. 107110.
Yang, H., Zhou, Y. & Liu, J. (2009), Land and water requirements

of biofuel and implications for food supply and the environment


in China, Energy Policy, vol. 37, pp. 18761885.
Young, B. (2006), Changes in herbicide use patterns and

production practices resulting from glyphosate-resistant crops,


Weed Technology, vol. 20, pp. 301307.

Willow and hybrid poplar


(short-rotation woody species)
Anderson, G. Q. A. & Fergusson, M. J. (2006), Energy from

biomass in the UK: sources, processes and biodiversity


implications, Ibis, vol. 148, pp. 180183.
Argus, G. (1997), Infrageneric classification of Salix (Salicaceae)

in the new world, Systematic Botany Monographs, vol. 52,


pp. 1121.
Ball, J. et al. (2005), Contribution of poplars and willows to

sustainable forestry and rural development, Unasylva, vol. 221,


pp. 39.
Bradshaw, H. D., Ceulemans, R., Davis, J. & Stettler, R. (2000),

Emerging model systems in plant biology: poplar (Populus) as a


model forest tree, Journal of Plant Growth Regulation, vol. 19,
pp. 306313.
Cermk, J., Jenk, J., Kucera, J. & dek, V. (1984), Xylem water

flow in a crack willow tree (Salix fragilis L.) in relation to diurnal


changes of environment, Oecologia, vol. 64, pp. 145151.
Coleman, M. D. et al. (2003), Production of short-rotation

woody crops grown with a range of nutrient and water


availability: establishment report and first-year responses.
US Department of Agriculture.
Davis, S. C. et al. (2012), Harvesting carbon from eastern US

forests: opportunities and impacts of an expanding bioenergy


industry, Forests, vol. 3, pp. 370397.
DeBell, D. S. & Harrington, C. A. (1993), Deploying geno-types

in short-rotation plantations: mixtures and pure cultures of


clones and species, Forestry Chronicle, vol. 69, pp. 705713.
Demeritt Jr, M. E. (1990), Populus L.: poplar hybrids.

US Forestry Service. Available from: http://www.na.fs.fed.us/


pubs/silvics_manual/volume_2/populus/populus.htm [accessed
June 2014].

Graham, A. (1964), Origin and evolution of the biota of

Heilman, P. (1999), Planted forests: poplars, New Forests, vol. 17,

pp. 8993.
Heller, M. (2003), Life cycle assessment of a willow bioenergy

cropping system, Biomass and Bioenergy, vol. 25, pp. 147165.


Hinckley, T. M. et al. (1994), Water flux in a hybrid poplar stand,

Tree Physiology, vol. 14, pp. 10051018.


Jansson, S., Bhalerao, R. & Groover,

A. (eds) (2010), Genetics


and genomics of Populus. Springer, New York.

Kandus, P. & Malvrez, A. (2004), Vegetation patterns and

change analysis in the Lower Delta islands of the Paran River


(Argentina), Wetlands, vol. 24, pp. 620632.
Karp, A., Hanley, S. J., Trybush, S. O., Macalpine, W., Pei, M.

& Shield, I. (2011), Genetic improvement of willow for


bioenergy and biofuels, Journal of Integrative Plant Biology,
vol. 53, pp. 151165.
Kendall, D. & Hunter, T. (1996), Susceptibility of willow clones

(Salix spp.) to herbivory by Phyllodecta vulgatissima (L.) and


Galerucella lineola (Fab.)(Coleoptera, Chrysomelidae), Annals
of Applied Biology, vol. 129, pp. 379390.
Kitin, P. et al. (2010), Tyloses and phenolic deposits in xylem

vessels impede water transport in low-lignin transgenic poplars:


a study by cryo-fluorescence microscopy, Plant Physiology,
vol. 154, pp. 887898.
Kuzovkina, Y. & Quigley, M. (2005), Willows beyond wetlands:

uses of Salix L. species for environmental projects, Water, Air, &


Soil Pollution, vol. 162, pp. 183204.
Kuzovkina, Y. & Weih, M. (2008), Salix: botany and global

horticulture, Horticultural Reviews, vol. 34, pp. 447489.


Labrecque, M. & Teodorescu, T. I. (2005), Field performance and

biomass production of 12 willow and poplar clones in shortrotation coppice in southern Quebec (Canada), Biomass and
Bioenergy, vol. 29, pp. 19.
Lantmannen Agroenergi, Manual for SRC willow growers. York,

UK and Orebro, Sweden. Available from: http://www.


voederbomen.nl/wordpress/wp-content/uploads/2012/08/
ManualSRCWillowGrowers.pdf [accessed June 2014].
Larsson, S. (1997), Commercial breeding of willow for

short rotation coppice, Aspects of Applied Biology, vol. 49,


pp. 215218.

6 Biomass feedstock crops | 111

Ledin, S. (1996), Willow wood properties, production and

economy, Biomass and Bioenergy, vol. 11, pp. 7583.


Manchester, S. R., Judd, W. S. & Handley, B. (2006), Foliage and

fruits of early poplars (Salicaceae: Populus) from the Eocene of


Utah, Colorado, and Wyoming, International Journal of Plant
Science, vol. 167, pp. 897908.
Mashkina, O. S., Tabatskaya, T. M., Gorobets, A. I. &

Shestibratov, K. A. (2010), Method of clonal micropropagation of


different willow species and hybrids, Applied Biochemistry and
Microbiology, vol. 46, pp. 769775.
McCracken, A. R. & Dawson, W. M. (1994), Experiences in the

use of mixed-clonal stands of Salix as a method of reducing the


impact of rust diseases, Norwegian Journal of Agricultural
Sciences, vol. 18, pp. 101109.
Neuhauser, E. et al. (2000), Short-rotation woody crops program

biomass power for rural development technical report.


US Department of Energy.
Perala, D. A. (1990). Populus tremuloides: quaking aspen.

US Forestry Service.
Phillips, D. H. & Burdekin, D. A. (1992), Diseases of forest and

ornamental trees. Macmillan, London.


Pitcher, J. A. & McKnight, J. S. (1990), Salix nigra: black willow.

US Forestry Service.
Ranney, J. W., Wright, L. L. & Layton, P. A. (1987), Hardwood

energy crops: the technology of intensive culture, Journal of


Forestry, vol. 89, pp. 1720, 22, 24, 26.
Sangha, K. S. (2011), Evaluation of management tools for the

control of poplar leaf defoliators (Lepidoptera: notodontidae) in


northwestern India, Journal of Forestry Research, vol. 22,
pp. 7782.
Sartori, F., Lal, R., Ebinger, M. H. & Parrish, D. J. (2006),

Potential soil carbon sequestration and CO2 offset by dedicated


energy crops in the US, Critical Reviews in Plant Sciences,
vol. 25, pp. 441472.
Schoene, W. J. (1907), The poplar and willow borer

(Cryptorhynchus lapathi L.). New York Agricultural Experiment


Station, New York.
Singh, T. & Kostecky, M. (1986), Calorific value variations in

components of 10 Canadian tree species, Canadian Journal of


Forest Research, vol. 16, pp. 13781381.
Stanton, B. J. et al. (2010), Populus breeding: from the classical

to the genomic approach, in Jansson, S. et al. (eds) Genetics and


genomics of Populus. Springer, New York.
Stettler, R. F., Fenn, R. C., Heilman, P. E. & Stanton, B. J. (1988),

Populus trichocarpa x Populus deltoides hybrids for shortrotation culture: variation patterns and 4-year field performance,
Canadian Journal of Forestry, vol. 18, pp. 745753.
Stott, K. G. (2001), Cultivation and use of basket willow.

The Basketmakers Association and IACR Long Ashton Research


Station, UK.
Tuskan, G. (1998), Short-rotation woody crop supply systems in

the United States: what do we know and what do we need to


know?, Biomass and Bioenergy, vol. 14, pp. 307315.
Uliassi, D. & Ruess, R. (2002), Limitations to symbiotic nitrogen

fixation in primary succession on the Tanana River floodplain,


Ecology, vol. 83, pp. 88103.
US Department of Agriculture (2012), Conservation Plant
Characteristics for Salix sepulcralis. National Resources
Conservation Service. Available from: https://plants.usda.gov/
java/charProfile?symbol=SASE10 [accessed June 2014].
Vahala, T., Stabel, P. & Eriksson, T. (1989), Genetic

transformation of willows (Salix spp.) by Agrobacterium


tumefaciens, Plant Cell Reports, vol. 8, pp. 5558.
Van den Driessche, R., Rude, W. & Martens, L. (2003), Effect of

fertilization and irrigation on growth of aspen (Populus


tremuloides Michx.) seedlings over three seasons, Forest Ecology
and Management, vol. 186, pp. 381389.

112 | 6 Biomass feedstock crops

Walsh, M. E., Daniel, G., Shapouri, H. & Slinsky, S. P. (2003),

Bioenergy crop production in the United States: potential


quantities, land use changes, and economic impacts on the
agricultural sector, Environmental and Resource Economics,
vol. 24, pp. 313333.
Weih, M. & Nordh, N. (2002), Characterising willows for

biomass and phytoremediation: growth, nitrogen and water use


of 14 willow clones under different irrigation and fertilisation
regimes, Biomass and Bioenergy, vol. 23, pp. 397413.
Weiland, J. E., Stanosz, J. C. & Stanosz, G. R. (2003), Prediction

of long-term canker disease damage from the responses of


juvenile poplar clones to inoculation with Septoria musiva,
Plant Disease, vol. 87, p. 12.
Wilson, J. (1964), Annual growth of Salix arctica in the high-

arctic, Annals of Botany, vol. 28, pp. 7176.


Wright, L. (1994), Production technology status of woody and

herbaceous crops, Biomass and Bioenergy, vol. 6, pp. 191209.


Wullschleger, S. (1998), A review of whole-plant water use

studies in tree, Tree Physiology, vol. 18, pp. 499512.


Zalesny, R. J. et al. (2011), Woody biomass from short rotation

energy crops, in Junyong (J. Y.) Zhu et al. (eds) Sustainable


production of fuels, chemicals, and fibers from forest biomass,
ACS Symposium Series, vol. 1067.
Zhen-Fu, F. (1987), On the distribution and origin of Salix in the

world, Acta Phytotaxonomica Sinica, vol. 25, pp. 307313.

Wood residues
Adler, P. R., Grosso, S. J. D. & Parton, W. J. (2007), Life-cycle

assessment of net greenhouse-gas flux for bioenergy cropping


systems, Ecological Applications, vol. 17, pp. 675691.
Anderson, G. Q. A. & Fergusson, M. J. (2006), Energy from

biomass in the UK: sources, processes and biodiversity


implications, Ibis, vol. 148, pp. 180183.
Anttila, P., Karjalainen, T. & Asikainen, A. (2009), Global

potential of modern fuelwood. Finnish Forest Research


Institute.
Balan, V., Kumar, S., Bals, B., Chundawat, S., Jin, M. & Dale, B.

(2012), Biochemcial and thermochemical conversion of


switchgrass to biofuels, in Monti, A. (ed.) Switchgrass: a
valuable biomass crop for energy. Springer: London.
Cherubini, F. et al. (2009), Energy- and greenhouse gas-based

LCA of biofuel and bioenergy systems: key issues, ranges and


recommendations, Resources, Conservation and Recycling,
vol. 53, pp. 434447.
Dornburg, V. & Faaij, A. (2001), Efficiency and economy of wood-

fired biomass energy systems in relation to scale regarding heat


and power generation using combustion and gasification
technologies, Biomass and Bioenergy, vol. 21, pp. 91108.
Food and Agriculture Organization of the United Nations (2010),

Global forestry production and trade, FAOSTAT. Available from:


http://faostat.fao.org/ [accessed February 2014].
Food and Agriculture Organization of the United Nations,

FAOSTAT forestry database. Available from: http://faostat3.fao.


org/faostat-gateway/go/to/home/E [accessed June 2014].
Fujino, J., Yamaji, K. & Yamamoto, H. (1999), Biomass-balance

table for evaluating bioenergy resources, Applied Energy, vol. 63,


pp. 7589.
Haberl, H. et al. (2010), The global technical potential of bio-

energy in 2050 considering sustainability constraints, Current


Opinion in Environmental Sustainability, vol. 2, pp. 394403.
Hoekman, S. K. (2009), Biofuels in the US challenges and

opportunities, Renewable Energy, vol. 34, pp. 1422.


Lamlom, S. H. & Savidge, R. A. (2003), A reassessment of carbon

content in wood: variation within and between 41 North


American species, Biomass and Bioenergy, vol. 25, pp. 381388.
McKendry, P. (2002), Energy production from biomass (part 1):

overview of biomass, Bioresource Technology, vol. 83, pp. 3746.

Monte, M. C., Fuente, E., Blanco, A. & Negro, C. (2009), Waste

management from pulp and paper production in the European


Union, Waste Management, vol. 29, pp. 293308.
Parikka, M. (2004), Global biomass fuel resources, Biomass and

Bioenergy, vol. 27, pp. 613620.


Perlack, R. et al. (2005), Biomass as feedstock for a bioenergy

and bioproducts industry: the technical feasibility of a billionton annual supply. Oak Ridge National Laboratory, Oak Ridge,
TN, US.
Perlin, J. (1997), Forest journey: the role of wood in the

development of civilization. Harvard University Press,


Cambridge, US.
US Department of Energy (2011), US billion-ton update:

biomass supply for a bioenergy and bioproducts industry.


Oak Ridge National Laboratory, Oak Ridge, TN, US.
Wihersaari, M. (2005), Greenhouse gas emissions from final

harvest fuel chip production in Finland, Biomass and Bioenergy,


vol. 28, pp. 435443.

Herbaceous crop residues


Adler, P. R., Grosso, S. J. D. & Parton, W. J. (2007), Life-cycle

assessment of net greenhouse-gas flux for bioenergy cropping


systems, Ecological Applications, vol. 17, pp. 675691.
Anderson, G. Q. A. & Fergusson, M. J. (2006), Energy from

biomass in the UK: sources, processes and biodiversity


implications, Ibis, vol. 148, pp. 180183.
Andrews, S. (2006), Crop residue removal for biomass energy

production: effects on soils and recommendations.


US Department of Agriculture, Natural Resources Conservation
Service.
Cherubini, F. & Ulgiati, S. (2010), Crop residues as raw materials

for biorefinery systems a LCA case study, Applied Energy,


vol. 87, pp. 4757.
Dale, B. E. & Kim, S. (2006), Biomass refining global impact

the biobased economy of the 21st century, in Kamm, B., Gruber,


P. R. & Kamm, M. (eds) Biorefineries industrial processes and
products: status quo and future directions. Wiley-VCH Verlag
GmbH & Co. KGaA, Weinheim, Germany.
Dolan, M. S. et al. (2006), Soil organic carbon and nitrogen in a

Minnesota soil as related to tillage, residue and nitrogen


management, Soil and Tillage Research, vol. 89, pp. 221231.
Gallagher, P. W. et al. (2003), Supply and social cost estimates for

biomass from crop residues in the United States, Environmental


and Resource Economics, vol. 24, pp. 335358.
Haberl, H. et al. (2010), The global technical potential of bioenergy in 2050 considering sustainability constraints, Current
Opinion in Environmental Sustainability, vol. 2, pp. 394403.
Hoekman, S. K. (2009), Biofuels in the US challenges and

opportunities, Renewable Energy, vol. 34, pp. 1422.


Kim, S. & Dale, B.E. (2004), Global potential bioethanol

production from wasted crops and crop residues, Biomass and


Bioenergy, vol. 26, pp. 361375.
Lal, R. (2005), World crop residues production and implications

of its use as a biofuel, Environment International, vol. 31,


pp. 575584.
Larson, W. E. (1979), Crop residue: energy production on erosion

control, Journal of Soil and Water Conservation, vol. 34,


pp. 7476.
Ruddiman, W. (2003), The anthropogenic greenhouse era began

thousands of years ago, Climatic Change, vol. 61, pp. 261293.


Sheehan, J. et al. (2003), Energy and environmental aspects of

using corn stover for fuel ethanol, Journal of Industrial Ecology,


vol. 7, pp. 117146.
Spatari, S., Bagley, D. M. & MacLean, H. L. (2010), Life cycle

evaluation of emerging lignocellulosic ethanol conversion


technologies, Bioresource Technology, vol. 101, pp. 654667.
Zadrail, F. (1977), The conversion of straw into feed by

basidiomycetes, Applied Microbiology and Biotechnology,


vol. 4, pp. 273281.

Miscellaneous wastes
Anderson, G. Q. A. & Fergusson, M. J. (2006), Energy from

biomass in the UK: sources, processes and biodiversity


implications, Ibis, vol. 148, pp. 180183.
Behzadi, S. & Farid, M. (2007), Review: examining the use of

different feedstock for the production of biodiesel, Asia-Pacific


Journal of Chemical Engineering, vol. 2, pp. 480486.
Berglund, M. & Brjesson, P. (2006), Assessment of energy

performance in the life-cycle of biogas production, Biomass and


Bioenergy, vol. 30, pp. 254266.
Brjesson, P. & Berglund, M. (2006), Environmental systems

analysis of biogas systems part I: fuel-cycle emissions, Biomass


and Bioenergy, vol. 30, pp. 469485.
Brjesson, P. & Berglund, M. (2007), Environmental systems

analysis of biogas systems part II: the environmental impact of


replacing various reference systems, Biomass and Bioenergy,
vol. 31, pp. 326344.
Cherubini, F., Bargigli, S. & Ulgiati, S. (2009), Life cycle

assessment (LCA) of waste management strategies: landfilling,


sorting plant and incineration, Energy, vol. 34, pp. 21162123.
Danish Energy Authority, ElkraftSystem, Eltra (2005),

Technology data for electricity and heat generating plants.


Copenhagen, Denmark.
Danish Environmental Protection Agency (2003), Status on

organic municipal waste. Copenhagen.


Deublein, D. & Steinhauser, A. (2010), Biogas from waste and

renewable resources, 2nd edition. Wiley-VCH Verlag GmbH &


Co. KGaA, Weinheim, Germany.
Faaij, A. et al. (1997), Characteristics and availability of biomass

waste and residues in the Netherlands for gasification, Biomass


and Bioenergy, vol. 4, pp. 225240.
Haberl, H. et al. (2010), The global technical potential of bio-

energy in 2050 considering sustainability constraints, Current


Opinion in Environmental Sustainability, vol. 2, pp. 394403.
Hoogwijk, M. (2003), Exploration of the ranges of the global

potential of biomass for energy, Biomass and Bioenergy, vol. 25,


pp. 119133.
Imhof, K. (1980), Zum Wert der Schlammfaulung und Hinweise

zum Faulverfahren, Korrespondenz Abwasser.


Janulis, P. (2004), Reduction of energy consumption in biodiesel

fuel life cycle, Renewable Energy, vol. 29, pp. 861871.


Lebedevas, S. et al. (2006), Use of waste fats of animal and

vegetable origin for the production of biodiesel fuel: quality,


motor properties, and emissions of harmful components, Energy
& Fuels, vol. 20, pp. 22742280.
Mahro, B. & Timm, M. (2007), Potential of biowaste from the

food industry as a biomass resource, Engineering in Life


Sciences, vol. 7, pp. 457468.
Mller, H. B., Sommer, S. G. & Ahring, B. K. (2004), Methane

productivity of manure, straw and solid fractions of manure,


Biomass and Bioenergy, vol. 26, pp.485495.
Mnster, M. & Lund, H. (2009), Use of waste for heat, electricity

and transport challenges when performing energy system


analysis, Energy, vol. 34, pp. 636644.

Algae
Amos, R. (ed.) (2004), Handbook of microalgal culture:

biotechnology and applied phycology, 1st edition. Blackwell


Science Ltd.
Belarbi, E. H., Molina, E. & Chisti, Y. (2000), A process for high

yield and scaleable recovery of high purity eicosapentaenoic acid


esters from microalgae and fish oil, Enzyme and Microbial
Technology, vol. 26, pp. 516529.
Brennan, L. & Owende, P. (2010), Biofuels from microalgae

a review of technologies for production, processing, and


extractions of biofuels and co-products, Renewable and
Sustainable Energy Reviews, vol. 14, pp. 557577.

6 Biomass feedstock crops | 113

Brown, M. R., Jeffrey, S. W., Volkman, J. K. & Dunstan, G. A.

(1997), Nutritional properties of microalgae for mariculture,


Aquaculture, vol. 151, pp. 315331.
Chisti, Y. (2007), Biodiesel from microalgae, Biotechnology

Advances, vol. 25, pp. 294306.


Daroch, M., Geng, S. & Wang, G. (2013), Recent advances in

liquid biofuel production from algal feedstocks,Applied


Energy,vol. 102, pp. 13711381.
Gavrilescu, M. & Chisti, Y. (2005), Biotechnology a sustainable

alternative for chemical industry, Biotechnology Advances,


vol. 23, pp. 471499.
Grima, E. M., Fernndez Sevilla, J. M. & Acin Fernndez, F. G.

(2010), Microalgae, mass culture methods, in Flickinger, M. C.


(ed.) Encyclopedia of Industrial Biotechnology: Bioprocess,
Bioseparation, and Cell Technology. John Wiley & Sons, Inc.
Lowrance, R. et al. (2010), Sustainable alternative fuel feedstock

opportunities, challenges and roadmaps for six US regions,


Proceedings of the Sustainable Feedstocks for Advance Biofuels
Workshop.
Metting, F. (1996), Biodiversity and application of microalgae,

Journal of Industrial Microbiology & Biotechnology, vol. 17,


pp. 477489.

114 | 6 Biomass feedstock crops

Pagliaro, M. et al. (2009), Recent advances in the conversion of

bioglycerol into value-added products, European Journal of Lipid


Science and Technology, vol. 111, pp. 788799.
Sander, K. & Murthy, G. S. (2010), Life cycle analysis of algae

biodiesel, The International Journal of Life Cycle Assessment,


vol. 15, pp. 704714.
Schenk, P. M. et al. (2008), Second generation biofuels: high-

efficiency microalgae for biodiesel production, BioEnergy


Research, vol. 1, pp. 2043.
Sheehan, J. (1998), A look back at the US department of energys

aquatic species program: biodiesel from algae. National


Renewable Energy Laboratory
Spolaore, P., Joannis-Cassan, C., Duran, E. & Isambert, A. (2006),

Commercial applications of microalgae, Journal of Bioscience


and Bioengineering, vol. 101, pp. 8796.
Terry, K. L. & Raymond, L. P. (1985), System design for the

autotrophic production of microalgae, Enzyme and Microbial


Technology, vol. 7, pp. 474487.

Glossary
Agroecosystem The organisms (crops, livestock, microflora and -fauna etc.) and environment (soils, water,
climate etc.) of an agricultural area considered as an
ecosystem.
Agroforestry A multi-use form of land management
where trees are grown in association with arable crops
or pasture.
Alcohols A class of organic compounds closely related
to hydrocarbons. They are used in medicine, and in
industry as fuels and solvents. Examples include
methanol and ethanol, the substance that makes beer,
wine and spirits.
Anaerobic fermentation A biological process taking place
in the absence of oxygen, in which sugars are broken
down into alcohols and carbon dioxide.
Annual crops Crops whose life cycle, from seed to
harvest, is complete in less than 12 months.
Bagasse A term often used in relation to sugarcane
production: bagasse is the fibrous residue left after the
sugary juice has been extracted from the crushed cane.
Bioenergy Energy from any renewable biological material
derived from plants or animals.
Biennial plants Plants that produce only vegetative
growth (roots, shoots and leaves) during the first year;
flowering and fruiting (followed by death) occur in the
second year.
Biofuel Liquid fuels derived from biomass, used primarily
for transport, including ethanol, biodiesel and other
liquids.
Biogas A mixture of methane and CO2 produced by
the bacterial decomposition (fermentation) of organic
wastes and used as a fuel.
Biomass The solid matter in living or recently living
organisms.
Biome A major ecological community (ecosystem) type
(such as tropical rain forest, grassland or desert).
Bioplastic Plastics that are derived from biomass
including oils, fats and starches.
C3, C4 and CAM photosynthetic pathways
Green plants use energy from the sun to create
sugars from CO2 and water. This process is called
photosynthesis and has three variants:
C3 The most common photosynthetic pathway, found
particularly in temperate crops including rice and
wheat. Such plants tend to have a lower water-use
efficiency (WUE, see below) than plants with C4 and
CAM metabolism, because the process of CO2 diffusion
into leaf tissues through leaf pores also allows water to
be lost through transpiration.
C4 The C4 pathway may have evolved as a mechanism
to help plants survive drought or high temperatures,
because C4 plants (including maize and sugarcane) are
more common in tropical climates. They tend to have
higher rates of photosynthesis and WUE than C3 plants,
and the group includes some of the most productive
tropical crops.

CAM Some plants, such as agave and pineapple,


living in arid environments have a photosynthetic
pathway called crassulacean acid metabolism (CAM).
To conserve water during the heat of the day, their leaf
pores remain closed. CO2 diffuses into the leaves while
the leaf pores are open during the night and is used for
photosynthesis during the day.

Carbon fixation The incorporation of carbon into organic


compounds (such as simple sugars) by living organisms,
chiefly by photosynthesis in green plants.
Carbon sequestration A process by which CO2 is
removed from the atmosphere and held in solid or liquid
form. It occurs naturally as limestones are produced in
the oceans and peat and coals form. Large amounts of
carbon are also trapped in biomass and in the soil. In
recent years man-made systems have been sequestering
CO2 by injecting the gas under pressure into deeply
buried porous rock formations. It is an important
process, given global concern about rising atmospheric
CO2 levels.
Carbon sink A natural or man-made environmental,
process or system that absorbs and stores carbon
compounds. Often used to refer to the absorption
and storage of CO2.
Cellulose A complex organic polymer (C6H10O5)n
composed of glucose units, cellulose gives strength
to the cell walls of green plants. Cellulose, extracted
mainly from wood, is the basis of the worlds huge paper
industry. The complexity of its molecular structure
makes it difficult to deconstruct prior to fermentation
for liquid fuel production.
Coniferous Mostly evergreen trees and shrubs, usually
have needle-shaped or scale-like leaves. Not all conifers
produce their seed within cones: the yew tree, for
instance, produces berries. Conifers are generally
adapted to withstand drought, and many of them can
withstand very low winter temperatures.
Coppice Many broadleaved trees have the ability to
coppice (to produce vigorous new shoots from the base)
if they are cut down. This means that many crops of
shoots or poles can be produced from each root, over
many years. Few conifers are able to coppice.
Crop residue There are two types of agricultural crop
residues: field residues are materials (including stalks
and stubble (stems), leaves and seed pods) left on
the ground after the crop has been harvested. Good
management of field residues can increase efficiency of
irrigation and help control erosion. Process residues are
those materials (include husks, seeds, bagasse and roots)
left after crop processing. They can be used as animal
fodder, as soil improvers, and in manufacturing.
Cultivar (cultivated variety) A variety or strain of
plant that has been bred deliberately and whose
characteristics persist under cultivation.
Direct land-use change (DLUC) Usually used to describe
the direct and observable effects of change of land
use from undisturbed habitat (such as rainforest) to
agriculture or from one type of agriculture to another.
Glossary | 115

Ecosystem A community of organisms and its


environment functioning as an ecological unit.
Ecosystem services A wide range of services that
together maintain the conditions for life on earth. The
benefits/products of these services include food and
water; nutrient cycling; flood and disease control; and
spiritual, recreational and cultural benefits.
Esterification A reaction of an alcohol with an acid to
produce an ester and water.
Ethanol A type of alcohol with the chemical formula
C2H5OH. It is a volatile, flammable liquid used as a
solvent and in fuel; it is also the intoxicating agent in
alcoholic beverages. Bioethanol is ethanol produced
from biomass.
Evapotranspiration The process of water loss from soil.
This is a combination of evaporation from the soil
surface and transpiration from the plants growing in it.
Fatty acids Organic acids containing a long chain of
carbon atoms.
Feedstock A raw material, such as biomass or coal, used
to supply energy or material to a commercial process.
Fertigation To fertilize and irrigate at the same time, by
adding fertilizers to the water supply.
Genetically modified (GM, of an organism)Containing
genetic material that has been artificially altered so
as to produce a desired characteristic. The process of
genetic modification attracts controversy.
Glycerol Also known as glycerin or glycerine, this is a
sweet, water-absorbing sugaralcohol compound. It is a
by-product of biodiesel production and is widely used in
the pharmaceutical industry.
Greenhouse gas (GHG) A gas that contributes to the
greenhouse effect (by which the earth is kept warm
by an atmospheric blanket) by absorbing infrared
radiation. Carbon dioxide and methane are examples of
greenhouse gases.
HaberBosch process An industrial process invented by
Fritz Haber to react nitrogen and hydrogen, often used
in the production of ammonia used in fertilizers.
Hemicellulose An organic polymer found in plant cell
walls, as is cellulose (see above). Its structure is more
complex than cellulose, though it is more easily broken
down to single sugars (monosaccharides) and shortchain polymers with mild heat treatment.
Hydrolysis A process of decomposition involving the
breaking of chemical bonds by the addition of water.
Hydro-treated vegetable oils (HVO) Diesel fuels produced
by using hydrogen to remove oxygen from triglyceride
components of vegetable oils and animal fats.
Indirect land-use change (ILUC) Used to describe
ancillary or unintended and indirect effects resulting
from changing the use of land for one purpose to
another. If maize acreage in the US were used for fuel
instead of animal feed, for example, and this created a
market signal to plant more maize in Brazil using forest
or pasture land, the impacts of the Brazilian conversion
would constitute an indirect effect of the US action. The
magnitude of any effects of ILUC is substantially more
difficult to evaluate than DLUC.

116 | Glossary

Intercropping Growing two or more crops simultaneously


(such as in alternate rows) on the same plot of land. An
example would be growing beans between maize plants
in smallholder farming systems in parts of Africa.
Life cycle analysis (LCA) Also known as life cycle
assessment, it is a comprehensive interdisciplinary
assessmentthat identifies and attempts to quantify
theenergy,material andwasteflowsof aproduct, and
theirimpacton theenvironment.
Lignin A complex polymer related to cellulose that
provides rigidity and forms the woody cell walls of
plants and the cementing material between them. It
confers strength and disease-resistance on woody plants,
and is difficult to decompose to its constituent parts.
The process by which plants deposit lignin in their cell
walls is called lignification.
Lignocellulosic A word to describe various substances
(consisting of cellulose intimately associated
withlignin) found in the woody cell walls of plants.
Woodchips would be an example of lignocellulosic
biomass.
Lipid A group of naturally occurring molecules that
include fats and waxes. Together with proteins and
carbohydrates, lipids constitute the principal structural
components of living cells.
Loam A soil consisting of a friable mixture of varying
proportions of clay, silt and sand. Loams are productive
agricultural soils.
Metabolic pathway A sequence of chemical reactions in a
living organism.
Methane (CH4) A gaseous hydrocarbon fuel. It is lighter
than air and forms explosive mixtures with air. It is
produced by the decomposition of organic matter and is
used as a fuel and raw material in chemical synthesis.
It is a major greenhouse gas with a more powerful
greenhouse effect than CO2.
Net primary productivity (NPP) The rate at which an
ecosystem accumulates energy or biomass, excluding
the energy it uses for the process of respiration.
Nitrous oxide (N2O) An atmospheric pollutant and
greenhouse gas produced by combustion.
Peat Partially carbonized vegetable substance formed by
incomplete decomposition of plant material in water.
Peat is an important store of carbon, which is released
into the atmosphere when peat is burned (for fuel) or
when peat soils are brought under cultivation.
Perennial crop A crop from plants that live for more than
one year. Examples include sugarcane, woody biomass
and perennial grasses.
Photosynthesis See C3, C4 and CAM photosynthetic
pathways above.
Polymer A substance whose molecular structure is built
up of many similar units bonded together. Examples of
plant polymers include cellulose (composed of glucose
molecules) and starch (also composed of glucose, but
bonded in a different way to give different physical
properties).

Radiative forcing/exchange The difference between the


amount of radiant energy received on earth from the sun
and the amount radiated back into space. This is affected
by, among other factors, the concentration of greenhouse
gases in the atmosphere.
Respiration Energy yielding process found in living
organisms involving the breakdown of nutrients, the use
of oxygen and the production of carbon dioxide.
Rhizome A specialized underground plant stem that
superficially resembles a root. It contains deposits of
reserve food material, and produces both shoots and
roots.
Semi-arid A climate type that has only light rainfall,often
defined as having 2550cm (1020in) of annual
precipitation.
Senescence The final growth phase in a plant or leaf,
during which nutrients are taken back to the root system
before leaf-fall or death.
Silage Fodder converted into succulent feed for livestock
through anaerobic acid fermentation in a silage clamp.
Soil organic matter Theorganic component of soil, which
includes the living biomass of microorganisms, and
fresh and partially decomposed residues. It also includes
well-decomposed, highly stable organic material.
Surface litter is generally not included as part of soil
organic matter but can become part of it if physically
incorporated into the soil. Soil organic matter is of vital
importance for nutrient cycling, erosion protection and
for its water-holding capacity.
Short rotation coppice (SRC) A management regime that
promotes the growth of multiple stems by cutting trees
back quite close to the ground every two to four years.
SRC is often used to produce woody biomass.

Short rotation forestry (SRF) A management regime


under which trees are planted and then felled when they
have reached a size of typically 1020cm diameter at
breast height. Depending on tree species and climate,
this can take between three and 20 years, and is
therefore intermediate in timescale between SRC and
conventional forestry.
Starch A polymer of glucose that is the chief storage form
of carbohydrate in plants. It is an important product for
fermentation into bioethanol via its constituent glucose.
Stover A straw-like substance left over after grain crops
like maize (corn) or sorghum have been harvested.
It is used as feed for livestock and has potential for
conversion to biofuel.
Subsoil The stratum of weathered material that underlies
the surface soil. Subsoil lacks organic matter.
Tillage The preparation of land for growing crops.
Tilth The condition of soil after it has been ploughed and
harrowed to create an even, lump-free surface. A tilth is
required for consistent seed germination.
Topsoil Surface soil in which plants have most of their
roots, and which is turned over during ploughing.
Traditional biomass Woodfuels, agricultural by-products
and dung burned for cooking and heating purposes. Very
widely used in developing countries.
Transpiration Loss of water from a plant, mainly through
the leaf pores (stomata).
Tuber A short, fleshy, usually underground stem bearing
minute-scale leaves each of which is potentially able to
produce a new plant. A potato is a tuber.
Water-use efficiency (WUE) At a crop level, WUE is
defined as the ratio of the crop mass yield to the water
mass lost through evapotranspiration over the growing
season. In practical terms, WUE can be thought of as crop
yield per unit of water.

Image sources and credits


p. 72 Maize close-up from Shutterstock and harvesting
from Tim Mies at the Energy Biosciences Institute.

p. 89 Oilseed rape or canola in production from


Shutterstock.

p. 74 Wheat in production from Shutterstock.

p. 90 Willow and hybrid poplar close-ups, as well as


coppicing, from Tom Voigt at the Energy Biosciences
Institute; short rotation harvest from Shutterstock.

p. 75 Sugarcane close-up and harvesting from


Caio Fortes, BP.
p. 77 Sweet sorghum in production from Caio Fortes, BP.
p. 78 Switchgrass close-up from Tom Voigt and harvesting
from Tim Mies, both at the Energy Biosciences Institute.
p. 80 Cordgrass in production from Tom Voigt at the
Energy Biosciences Institute.
p. 81 Miscanthus close-up from Tom Voigt and harvesting
from Tim Mies, both at the Energy Biosciences Institute.
p. 83 Energy cane in production from Joshua Drake, BP.
p. 84 Oil palm close-up and harvesting from Shutterstock.
p. 86 Jatropha in production from Shutterstock.
p. 87 Soybean close-up from Shutterstock and harvesting
from Tim Mies at the Energy Biosciences Institute.

p. 92 Pine and eucalyptus in production from


Shutterstock.
p. 93 Forest waste being chipped from Nevada Division
of Forestry.
p. 95 Corn stover residue collection from Tom Campbell
at Purdue Agricultural Communication.
p. 97 Biodigester from Shutterstock.
p. 98 Algae pond Cyanotech Corporation 2014
Producers of Nutritional Supplements from Microalgae.
p. 99 Macro algae from Shutterstock.
p. 100 Cassava from Shutterstock.

Glossary | 117

Plant
Planttypes
types
BP Biomass Handbook
Table 3.1 (20 December 2013)
Draft produced by ON Communication

Perennial
Plant characteristics icons in chapter 6Perennial

Herbaceous
Herbaceous

Feedstock

Conversion

Hydrolysis and fermentation

Lignocellulosic biomass

Herbaceous

Chemical process
Pre-process

Oil crops

Fuel for heat


and/or power

Anaerobic digestion

(Wood, straw,
energy crop, etc.)

Liquid fuels,
transport fuels
Biodiesel

Hydrogenation
Bioethanol
Transesterification
Other catalysis

Thermochemical process

(Rape, soy, palm, etc.)

Gasification
Pyrolysis

Other liquids

Woody

C 4C 4

Annual

Perennial
C3C3
C C4
Grain
or seed 4Herbaceous
Woody

Annua

Perenn

Grain or seed

Photosynthetic pathway
Photosynthetic pathway
Propagation
method
Propagation
method
Propagation
method
C3

C4

C3

C4

Plant
Plant
Plant
Plant
types
types
types
types
Plant
Plant
types
types

C3

C4

C3

C4

Annual
Annual
Annual
Annual

Annual
Annual
Seed
Stem
Rhizome
Seed
Stem
Rhizome Micropropagation
Micropropagation
Perennial
Perennial
Perennial
Perennial
cutting
or
cuttings
cutting
orroot
root
cuttings
Woody
Woody
Woody
Woody
Grain
Grain
Grain
orGrain
seed
oror
seed
seed
or seed
Herbaceous
Herbaceous
Herbaceous
Herbaceous
Propagation method
Propagation method
Perennial
Perennial
Woody
Woody
Grain
Grain
or or
seed
seed
Herbaceous
Herbaceous
Current
dominant
energy
use
Current
dominant
energy
use
Photosynthetic
Photosynthetic
Photosynthetic
Photosynthetic
pathway
pathway
pathway
pathway

Photosynthetic pathway

Photosynthetic
Photosynthetic
pathway
pathway

Gaseous fuel
Syngas

C 3 CC33 C 3

Schematic diagram of bioenergy production pathways. Feedstocks on the


left of the diagram are converted via a range of processes to solid, liquid or
gaseous fuels on the right. No attempt is made to show relative scales of
each process

C 4 CC44 C 4

DDC 4C 4Micropropagation
Seed
Stem EE C 3C 3 Rhizome
Rhizome
Micropropagat
CAM Stem
C3orCC3root
C4 CC4 4 C4
3 C3cuttings
cutting
cutting
or root cuttings
Bioethanol
Bioethanol Biodiesel
Biodiesel Heat
Heatand
andpower
power Biogas
Biogas
C3 C3
C4 C 4

Seed

Biogas

Combustion

 Table 3.1
Bioenergy production routes

C 3C 3

Energy

Biochemical process

Woody
Woody Grain
Grainororseed
seed

Plant types pathway


Photosynthetic
Photosynthetic
pathway
Plant types

Plant types

Sugar and starch crops

Annual
Annual

Current dominant energy use


Current dominant energy use
Other
Other
Propagation
Propagation
Propagation
Propagation
method
method
method
Plant life cyclemethod
E
Bioethanol

Propagation
Propagation
method
method
E
D
Car
Weight
Barrel
Car
Weight
Barrel
Biodiesel HeatAnnual
and power Perennial
Bioethanol
Biogas
Biodiesel Heat and power Biogas
D

Other

Car

Seed
Seed
Seed
Seed

Stem
Stem
Stem
Stem

Rhizome
Rhizome
Rhizome
Rhizome
Micropropagation
Micropropagation
Micropropagation
Micropropagation

Seed
Seed

Stem
Stem

Rhizome
Rhizome Micropropagation
Micropropagation

cutting
cutting
cutting
cutting
or root
ororroot
root
or
cuttings
root
cuttings
cuttings
cuttings
Other

cutting
cutting
oruse
or
root
root
cuttings
cuttings
Current
Current
Current
Current
dominant
dominant
dominant
dominant
energy
energy
energy
energy
use
use
use
Primary energy use
Current
Current
dominant
dominant
energy
energy
use
use
Weight

E EE E

Barrel

Car

Weight

D DD D

Bioethanol
Biodiesel
Biodiesel
Heat
Heat
Heat
and
Heat
and
and
power
and
power
power
power
Biogas
Biogas
Biogas
Bioethanol
Bioethanol
Biogas
E EBioethanol
DBiodiesel
DBiodiesel
Power
Powerusage
usage
Bioethanol
Bioethanol
Biodiesel Heat
Heat
and
and
power
power Biogas
Biogas
Other
Other
Other
OtherBiodiesel

Icons shown in grey indicate pre-commercial stages of adoption.

Other
Other
CarCar
CarCar Weight
Weight
Weight
Weight Barrel
Barrel
Barrel
Barrel
Power usage

CarCar

usage
Weight
Weight PowerBarrel
Barrel

Power
Power
Power
Power
usage
usage
usage
usage
Power
Power
usage
usage

Barrel

Biomass in the
energy industry
An introduction

Biomass in the energy industry


An introduction

Supported by BP, as part of the multi-partner


Energy Sustainability Challenge, which explores the
implications for the energy industry of competing
demands for water, land and minerals.

The handbook also provides key data about crops


species and biomass types that are already in
production or are being researched for biomass.
The data includes plant characteristics, suitable
growth conditions, required inputs and agricultural
practices, co-products and alternative markets, as
well as yield and energy productivity indicators.
The handbook offers a valuable guide for policy
makers, businesses and academics on the
characteristics of major biomass crops and the
issues related to sustainable and responsible use
of biomass for energy.
Biomass in the energy industry An introduction
shows:
n

What role biomass plays in the global


energy context.
n What fundamental knowledge is required to
understand bioenergy systems.
n How biomass is converted to energy and what
technological developments are under way.
n Why it is vital to view use of biomass for energy
from socio-economic, environmental and
political perspectives.
n What is the potential for bioenergy and how this
potential can be realized.
n Where can biomass feedstocks be grown and
what are the key characteristics of biomass
crops already in production or being researched
for biomass.
Published by BP p.l.c.
2014 BP p.l.c.
ISBN 978-0-9928387-1-3

9 780992 838713

Biomass in the energy industry An introduction

Biomass in the energy industry An introduction


is a study that provides contextual knowledge
required for assessing the potentials and issues
of using biomass for energy. The book is based
on literature research and review by colleagues
at the Energy Biosciences Institute
(www.energybiosciencesinstitute.org) and
is part of the Energy Sustainability Challenge
(www.bp.com/energysustainabilitychallenge)
series of handbooks. This book addresses the need
for having a holistic view of the benefits and risks
associated with bioenergy by studying the subject
from agricultural, energy, environmental,
technological, socio-economic and political
perspectives. The book emphasizes that realizing
the potential of biomass energy as a major player
in carbon emissions reduction needs careful
consideration of environmental aspects and
competing demands of food, water, energy and
other resources. Clear and consistent supportive
policies are also required to facilitate significant
financial investments for developing biomass
conversion technologies and improving
performance of biomass crops.

You might also like