Distributed Renewable Energies for Off-Grid Communities: Empowering a Sustainable, Competitive, and Secure Twenty-First Century [2 ed.] 0128216050, 9780128216057

Distributed Renewable Energies for Off-Grid Communities: Empowering a Sustainable, Competitive, and Secure Twenty-First

631 9 47MB

English Pages 594 [568] Year 2021

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Distributed Renewable Energies for Off-Grid Communities: Empowering a Sustainable, Competitive, and Secure Twenty-First Century [2 ed.]
 0128216050, 9780128216057

Table of contents :
Front-Matter_2021_Distributed-Renewable-Energies-for-Off-Grid-Communities
Distributed Renewable Energies for Off-Grid Communities
Copyright_2021_Distributed-Renewable-Energies-for-Off-Grid-Communities
Copyright
Dedication_2021_Distributed-Renewable-Energies-for-Off-Grid-Communities
Dedication
Citations_2021_Distributed-Renewable-Energies-for-Off-Grid-Communities
Citations
Foreword_2021_Distributed-Renewable-Energies-for-Off-Grid-Communities
Foreword
Preface_2021_Distributed-Renewable-Energies-for-Off-Grid-Communities
Preface
Acknowledgments_2021_Distributed-Renewable-Energies-for-Off-Grid-Communities
Acknowledgments
Chapter-One---What-Kind-of-Energy-D_2021_Distributed-Renewable-Energies-for-
One . What Kind of Energy Does the World Need?
1.1 . Distributed renewable energy
1.1.1 What kind of energy does the world need?
1.1.1.1 Introduction
1.1.2 Distributed renewable energy for energy access
References
1.2 . Using distributed energy resources to meet the trilemma challenges
1.2.1 Energy trilemma index
1.2.2 Dimensions
1.2.3 Monitoring the sustainability of national energy systems
Reference
1.3 . Scope of the book
1.3.1 Distribution
1.3.2 Distributed energy generation
1.3.3 Distributed energy supply
1.3.4 Community power
1.3.5 Off-grid systems
1.3.6 Concluding remarks
Further reading
Chapter-Two---Restructuring-future-ene_2021_Distributed-Renewable-Energies-f
Two . Restructuring future energy generation and supply
2.1 Basic challenges
2.2 Current and future energy supplies
2.3 Peak oil
2.4 Availability of alternative resources
2.5 Outlook
References
Further reading
Chapter-three---Road-map-of-distributed_2021_Distributed-Renewable-Energies-
three . Road map of distributed renewable energy communities
3.1 Energy and sustainable development
3.2 Community involvement
3.3 Facing the challenges
3.4 The concept of the food and agriculture organization, an integrated energy community
3.5 Global approach
3.5.1 Basic elements of energy demand
3.5.1.1 Heat
3.5.1.2 Electric power
3.5.1.3 Water
3.5.1.4 Lighting
3.5.1.5 Cooking
3.5.1.6 Health and sanitation
3.5.1.7 Communications
3.5.1.8 Mobility
3.5.1.9 Agriculture
3.5.1.10 Maintenance workshops and small marketsindustries
3.6 Basic and extended needs
3.6.1 Typical electricity demands
3.6.2 Single- and multiple-phase island grid
3.6.2.1 Version 1: single-phase island grid
3.6.2.2 Version 2: three-phase island grid
3.6.2.3 Version 3: three-phase island grid and parallel operation of the Sunny Island inverter
3.6.3 System solution for island grids
3.7 Representative energy plant species for different climate regions
3.7.1 Temperate climate
3.7.2 Representative energy plant species for different climate regions (arid and semiarid climate)
3.7.3 Representative energy plant species for different climate regions (tropical and subtropical climate)
3.8 Regional implementation
3.9 Opportunities driven by energy sector coupling
3.9.1 Demand-side flexibility programs
References
Further reading
Chapter-four---Planning-of-integrated_2021_Distributed-Renewable-Energies-fo
four . Planning of integrated renewable communities
4.1 Introduction
4.2 Scenario 1
4.3 Scenario 2
4.4 Case study I: implementation of integrated energy farm under climatic conditions of central Europe
4.4.1 Specifications
4.4.2 Distribution of farm area
4.4.3 Farm production
4.4.4 Energy requirement
4.4.4.1 Administration and household
4.4.4.2 Agricultural activities
4.4.4.3 Site energy production
4.4.4.4 Origin of biomass
4.4.4.5 Contribution of different renewable energy sources
4.4.4.6 Investment requirement
4.5 Case study II: arid and semiarid regions
4.5.1 Specifications
4.5.2 Farm production
4.5.3 Energy requirement
4.5.3.1 Administration and household
4.5.3.2 Agricultural activities
4.5.3.3 Energy production on the farm
4.5.3.4 Origin of biomass
4.5.3.5 Contribution of different renewable energy sources
4.5.3.6 Investment requirement
Further reading
Chapter-Five---The-water-energy_2021_Distributed-Renewable-Energies-for-Off-
Five . The water–energy–food nexus
5.1 Determination of community requirements for energy, water, and food
5.1.1 Definitions
5.2 Modeling approaches
5.2.1 Scenario 1
5.2.2 Scenario 2
5.3 Data acquisition
5.4 Determination of energy and food requirements
5.4.1 Agricultural activities
5.4.2 Households
5.4.2.1 Heat energy
5.4.2.2 Electricity
5.4.3 Food requirement
5.5 Energy potential analysis
5.5.1 Solar energy
5.5.2 Exploitation of solar energy
5.5.3 Solar thermal system
5.5.4 Solar photovoltaics
5.6 Data collection and processing for energy use
5.6.1 Water and space heating
5.6.2 Drying of agricultural produce
5.7 Wind energy
5.8 Biomass
5.8.1 Energetic use of biomass
5.8.1.1 Combustion
5.8.1.2 Extraction
5.8.2 Biogas production
References
Further reading
Chapter-Six---Energy-bas_2021_Distributed-Renewable-Energies-for-Off-Grid-Co
Six . Energy basics
6.1 Basics of energy
6.1.1 Energy rating
6.1.2 Energy consumption
6.2 Special topics relating to electricity
6.2.1 Gross and net electricity production
6.2.2 Electricity sales
6.2.3 Efficiency for fossil fuel and nuclear sources
6.2.4 Energy equivalent for non–fossil fuel sources
6.2.5 Energy generation
6.3 Global contribution
6.4 Resources and applications
References
Further reading
Chapter-Seven---Solar-energy--Techn_2021_Distributed-Renewable-Energies-for-
Seven . Solar energy: technologies and options
7.1 Worldwide installed capacities
7.2 Photovoltaic
7.3 Global PV market
7.4 Applications
7.4.1 National Renewable Energy Laboratory design new solar cell with efficiency
7.4.2 High-concentration photovoltaics
7.5 Accumulation of soiling on solar energy systems
7.6 Concentrating solar thermal power
7.7 Solar thermal collectors
7.8 Solar cookers and solar ovens
7.8.1 Advantages and disadvantages of solar cookers
7.8.1.1 Advantages
7.8.1.2 Disadvantages
References
Further reading
Chapter-Eight---Wind-ene_2021_Distributed-Renewable-Energies-for-Off-Grid-Co
Eight . Wind energy
8.1 Wind power and wind energy
8.2 Types of wind turbines
8.2.1 Horizontal-axis wind turbines
8.2.2 Vertical-axis design
8.2.2.1 Darrieus wind turbine
8.2.2.2 Giromill
8.2.2.3 Savonius wind turbine
8.2.2.4 Twisted Savonius
8.3 Global market
8.4 Offshore wind farm Dogger Bank
8.5 Small wind turbines
8.5.1 Market overview of small wind turbine
References
Further reading
Chapter-Nine---Energy-resources--global-_2021_Distributed-Renewable-Energies
Nine . Energy resources, global contribution, and applications
9.1 Introduction
9.2 Bioenergy and biofuels: innovation and technology progress
9.3 Characteristics and potentials
9.4 Solid biofuels
9.4.1 Charcoal
9.4.2 Solid biomass fuels in Austria
9.4.2.1 Boilers and stoves
9.4.3 Briquettes
9.4.4 Pellets
9.5 Biogas and biomethane
9.5.1 Ethanol
9.5.2 Bio-oils
9.6 Conversion systems to heat, power, and electricity
9.6.1 Combined heat and power
9.6.1.1 Heat
9.6.1.2 Electricity
9.6.2 Steam technology
9.6.3 Gasification
9.6.4 Biomass stoves
9.6.5 Pyrolysis
9.6.6 Methanol
9.6.7 Synthetic oil
9.6.8 Fuel cells
9.6.9 The Stirling engine
9.6.10 Algae
9.6.10.1 Algae bioreactors
9.6.11 Hydrogen
9.7 Outlook
References
Further reading
Chapter-Ten---Hydropow_2021_Distributed-Renewable-Energies-for-Off-Grid-Comm
Ten . Hydropower
10.1 Introduction
10.2 Global production of hydropower energy
10.3 Types of hydropower plants
10.3.1 Impoundment plants
10.3.2 Diversion plants
10.3.3 Pumped storage plants
10.3.4 Sizes of hydroelectric power plants
10.3.4.1 Large hydropower
10.3.4.2 Small hydropower
10.3.4.3 Microhydropower
10.4 Types of turbines
10.4.1 Modern turbine types
10.4.1.1 Pelton, Cross-flow and Turgo turbines
10.4.1.2 Kaplan and Francis turbines
10.4.1.3 Archimedes’ screw and Waterwheel turbines
10.5 Relative efficiencies
10.6 Assessment of hydropower potential
10.7 Impact of climate change on hydropower generation
References
Chapter-Eleven---Marine-e_2021_Distributed-Renewable-Energies-for-Off-Grid-C
Eleven . Marine energy
11.1 Introduction
11.2 Ocean thermal energy conversion
11.2.1 Ocean thermal energy conversion systems technology
11.2.1.1 Closed cycle
11.2.1.2 Open cycle
11.2.1.3 Hybrid cycle
11.3 Advantages and disadvantages
11.3.1 Advantages
11.3.2 Disadvantages
11.4 Ocean tidal power
11.5 Ocean wave power
11.5.1 Offshore systems
11.5.2 Onshore systems
11.6 Environmental and economic challenges
References
Further reading
Chapter-Twelve---Geothermal_2021_Distributed-Renewable-Energies-for-Off-Grid
Twelve . Geothermal energy
12.1 Introduction
12.2 The history of geothermal energy
12.3 Geothermal heat pumps
12.4 Geothermal electricity
12.5 Environmental effects, benefits, and economic costs
12.6 The future of geothermal energy
References
Further reading
Chapter-Thirteen---Energy-storage--smar_2021_Distributed-Renewable-Energies-
Thirteen . Energy storage, smart grids, and electric vehicles
13.1 Energy storage
13.1.1 Batteries and hydrogen technology: keys for a clean energy future
13.1.2 Storage methods
13.1.3 Technologies for upregulation and downregulation
13.2 Smart grids
13.2.1 Definition and importance
13.2.2 Smart meters
13.2.3 United States version
13.2.3.1 Challenges
13.2.4 European strategies
13.2.5 Korean version
13.3 Electric vehicles
13.3.1 Current developments
13.3.2 Types of electric vehicles
13.3.3 Global battery electric vehicle and plug-in hybrid electric vehicle sales
13.3.4 Types of electric vehicles
13.3.4.1 Battery electric vehicles
13.3.4.2 Plug-in hybrid electric vehicles
13.3.4.3 Hybrid electric vehicles
13.4 Future developments
References
Further reading
Chapter-Fourteen---Current-distributed-rene_2021_Distributed-Renewable-Energ
Fourteen . Current distributed renewable energy in rural and urban communities
14.1 Thisted, Denmark: 100% renewable energy community
14.1.1 Implementation
14.2 Samsø island
14.3 Energy island of VindØ
14.4 Kampala, Uganda taxi-bike drivers move to electric bikes
14.5 Rural community of Jühnde
14.6 Containerized solar minigrid, Fanidiama village, Mali
14.7 Decentralized desalination systems powered by solar energy in Maasai, Tanzania
14.8 Road map to renewable energy in remote communities in Australia
14.9 Iraq Dream Homes
14.10 Renewables in Africa
14.10.1 Hydropower
14.10.2 Biomass
14.10.3 Geothermal
14.10.4 Wind power
14.10.5 Solar power
14.10.6 Biofuels
14.10.7 Energy efficiency
14.11 Renewables in India
14.12 Distributed renewable energy and solar oases for deserts and arid regions: the DESERTEC concept
14.13 Vatican City
14.13.1 And suddenly there was light!
References
Further reading
Chapter-Fifteen---Ownership--citizens-pa_2021_Distributed-Renewable-Energies
Fifteen . Ownership, citizens participation and economic trends
15.1 Community ownership
15.1.1 Benefits of community energy
15.2 Citizens' participation
15.3 The Danish ownership model
15.4 Integration of the energy supply by public ownership
15.5 Economic impacts
15.6 Socioeconomic benefits and economic impacts of Renewables 2019
Renewable Generation Capacity by Region.
15.7 Actions for broadening the ownership of renewables
15.8 Global investment's in renewables
15.9 Costs of renewables
References
Further reading
Chapter-Sixteen---The-importance-o_2021_Distributed-Renewable-Energies-for-O
Sixteen . The importance of green mobility
16.1 Environmental and social impacts
16.1.1 The CO2 impact of transport
16.1.2 Air pollution and health
16.2 Mobility on the road
16.2.1 Available technologies
16.2.1.1 Conventional fossil fuels
16.2.1.2 Liquefied petroleum gas
16.2.1.3 Natural gas and its alternatives
16.2.1.3.1 Synthetic natural gas
16.2.1.3.2 Unconventional fossil methane
16.2.1.3.3 Biogas and biomethane
16.2.1.3.4 Vehicles run on methane
16.2.1.4 Biofuels
16.2.1.4.1 First-generation biofuels
16.2.1.4.2 Second-generation biofuels
16.2.1.4.3 Third-generation biofuels
16.2.1.4.4 Fourth-generation biofuels
16.2.1.5 Electricity
16.2.1.6 Hydrogen
16.2.1.7 Hybrids
16.2.1.8 Electrofuels
16.2.1.8.1 Ammonia
16.2.1.8.2 Carbon electrofuels
16.2.2 Light-duty transportation
16.2.3 Heavy-duty transportation
16.3 Mobility on the rail
16.4 Mobility on the water
16.5 Mobility in the air
16.6 Rethinking mobility: are there any alternatives to current models?
References
Chapter-Seventeen---Water-desalination--puri_2021_Distributed-Renewable-Ener
Seventeen . Water desalination, purification, irrigation, and wastewater treatment
17.1 Introduction
17.2 Renewable energy and pumps
17.2.1 Selection of pumps operating by renewable energy
17.2.2 Solar pumps
17.2.3 Wind pumps
17.2.3.1 Example
17.2.4 Biomass energy and biofuel pumps
17.2.4.1 Solid biofuel for pumps
17.2.4.2 Liquid biofuel for pumps
17.2.4.3 Biogas fuel for pumps
17.2.4.3.1 Example
17.3 Renewable energy and water purification
17.4 Renewable energy and desalination
17.5 Renewable energy and wastewater treatment
17.6 Renewable energy and farm irrigation
References
Chapter-Eighteen---Technologies-at-t_2021_Distributed-Renewable-Energies-for
Eighteen . Technologies at the experimental stages
18.1 Introduction
18.2 Fusion power
18.3 Antimatter energy
18.4 Atmospheric electricity
18.5 Microalgae
18.6 Osmotic power
18.7 Advanced hydrogen technology
18.8 Outlook
Reference
Further reading
Chapter-Nineteen---Drivers-for-digi_2021_Distributed-Renewable-Energies-for-
Nineteen . Drivers for digitalization of energy
References
Chapter-Twenty---Blockch_2021_Distributed-Renewable-Energies-for-Off-Grid-Co
Twenty . Blockchain
20.1 Characteristics of blockchain
20.2 Blockchain technology background
References
Chapter-Twenty-one---Grid-challenges--Integra_2021_Distributed-Renewable-Ene
Twenty one . Grid challenges: integration of distributed renewables with the national grid
21.1 The electricity distribution grid
21.1.1 Transmission and distribution line
21.2 Siemens to install smart distribution networks in Iraqi Provinces
21.3 Penetration of renewables in the grid
21.4 Development direction, cyberattacks, and outlook
21.4.1 Rapid growth of coordination and control technology for future distributed generation
References
Further reading
Chapter-Twenty-Two---Marshall-plan-for-Empowering_2021_Distributed-Renewable
Twenty Two . Marshall plan for Empowering Urban and Rural Communities: strategies toward poverty and migration reduction
22.1 Introduction
22.1.1 Background
22.2 Integrated energy settlement, Wierthe, Germany
22.3 Desert culture in Al Minya, Egypt
22.3.1 The vital role of deserts
22.3.1.1 Farm benchmark and design
22.3.1.1.1 Animals
22.3.1.2 Cultivated crops in the Minya Farm
22.3.1.3 Energy supply
22.3.1.4 Infrastructure facilities
22.3.1.5 Irrigation water and methods
22.3.1.6 Recommendations for improvement, upgrading, and rehabilitation of the farm
22.3.1.7 Agriculture
22.3.1.7.1 Farm design
22.3.1.8 Desert plant adaptations and survival
22.3.1.8.1 Restructuring of the farm and creating new facilities
22.3.1.8.2 Energy sector
22.3.1.8.3 Energy forms
22.3.1.8.4 Irrigation systems: the following irrigation systems should be compared:
22.3.1.8.5 Research, training, education, and facilities
22.3.1.8.6 Economic and social impacts
22.3.1.8.7 Impact on climate, environment, and desertification
22.3.1.8.8 Financing, funding, and sponsorship sources of the project
22.3.1.8.9 Outlook, recommendations, and procedure for implementation
References
Further reading
Internet sources
22.3 . Integrated energy settlement in Rousse, Bulgaria
22.3.1 Objectives of the project
22.3.2 Adaptation of the integrated energy farm concept for Bulgaria
22.3.3 Working plan
22.3.3.1 Detailed description of the site to plan the integrated energy farm in Bulgaria
22.3.4 Analysis of collected data for the site
22.3.4.1 Current state of energy demand and supply
22.3.4.2 Renewable energy sources of the site
22.3.4.2.1 Solar energy
22.3.4.2.2 Wind energy
22.3.4.2.3 Biomass potential
22.3.5 Project outlines
22.3.5.1 Requirements
22.3.5.1.1 Scope of project (total project activity and supply)
22.3.5.2 Implementation of energy plantation
22.3.5.3 Scope: contribution of local authorities for project
22.3.5.4 Scope: contribution of project (outsourcing)
References
Further reading
22.4 . Sustainable development of village of Kiga, Iran
22.4.1 Integration of renewable energy and fuel conservation for the sustainable development of the village of Kiga, Tehran, Iran
22.4.1.1 Justification for the project site selection of Kiga village
22.4.1.2 Strategic analysis
22.4.1.3 Social and commercial characteristics of Kiga village
22.4.2 Determination and evaluation of energy consumption and needs
22.4.2.1 Modeling approaches
22.4.3 Typical user's appliances in remote areas
22.4.4 Heat supply
22.4.5 Biomass and food
22.4.5.1 Identification of needs and options
22.4.6 Options for the implementation of the project
22.4.6.1 Energy needs and supply
22.4.7 Fuel conservation
22.4.8 Conclusions
Further reading
22.5 . Empowering of three urban cities in Africa (Empowering Urban Cities in Africa)
22.5.1 Project: strengthening urban resilience in African cities
22.5.2 Objectives
22.5.3 The three African urban cities
22.5.3.1 Technical framework of the project
22.5.3.2 Conducting the survey
22.5.3.3 Data acquisition, evaluation, processing, and analysis of the results
22.5.3.4 Recommendations for resilience implementation
22.5.3.5 Presentation of scientific project objective
References
Further reading
Chapter-Twenty-three---Our-vision-for-peace-_2021_Distributed-Renewable-Ener
Twenty three . Our vision for peace via renewables: power, water and food for all
23.1 Key words “solar oases”
23.2 Procedure
23.3 Concluding remarks and outlook
Further reading
Twenty three . . Additional note
The plan
Appendix-One---Glossa_2021_Distributed-Renewable-Energies-for-Off-Grid-Commu
One - Glossary
Regional definitions
Appendix-Two---List-of-energy-abbrev_2021_Distributed-Renewable-Energies-for
Two - List of energy abbreviations and acronyms
A
B
C
D
E
F
G
H
I
J
K
L
M
N
O
P
Q
R
S
T
U
V
W
X
Y
Z
Residential energy acronyms
Key energy acronyms and terms
Acronyms: agencies and organizations
Acronyms and abbreviations
Units of measure
Unit conversion factors
Multiply by to obtain
Abbreviations and acronyms
Appendix-Three---Conversion-_2021_Distributed-Renewable-Energies-for-Off-Gri
Tree - Conversion factors
Units and conversions
General conversion factors for energy:
Conversion factors for mass:
Conversion factors for volume:
Decimal prefixes:
Some conversion factors you may need to assess your site's feasibility:
Energy conversion and related WEC conversions
Conversion factors and energy equivalents
Conversion factors and energy equivalents
Basic energy units
Appendix-Four---Inventory-of-photovoltaic-s_2021_Distributed-Renewable-Energ
Four - Inventory of photovoltaic systems for sustainable rural development
Index_2021_Distributed-Renewable-Energies-for-Off-Grid-Communities
Index
A
B
C
D
E
F
G
H
I
J
K
L
M
N
O
P
R
S
T
U
V
W

Citation preview

DISTRIBUTED RENEWABLE

ENERGIES FOR OFF-GRID COMMUNITIES Empowering a Sustainable, Competitive, and Secure Twenty-First Century Second Edition

Editor

NASIR EL BASSAM Contributors

GREG P. SMESTAD, MARCIA LAWTON SCHLICHTING, THAMER A. MOHAMED, DANIELE PAGANI AND LOTHAR SCHLICHTING

Elsevier Radarweg 29, PO Box 211, 1000 AE Amsterdam, Netherlands The Boulevard, Langford Lane, Kidlington, Oxford OX5 1GB, United Kingdom 50 Hampshire Street, 5th Floor, Cambridge, MA 02139, United States Copyright Ó 2021 Elsevier Inc. All rights reserved. No part of this publication may be reproduced or transmitted in any form or by any means, electronic or mechanical, including photocopying, recording, or any information storage and retrieval system, without permission in writing from the publisher. Details on how to seek permission, further information about the Publisher’s permissions policies and our arrangements with organizations such as the Copyright Clearance Center and the Copyright Licensing Agency, can be found at our website: www.elsevier.com/permissions. This book and the individual contributions contained in it are protected under copyright by the Publisher (other than as may be noted herein). Notices Knowledge and best practice in this field are constantly changing. As new research and experience broaden our understanding, changes in research methods, professional practices, or medical treatment may become necessary. Practitioners and researchers must always rely on their own experience and knowledge in evaluating and using any information, methods, compounds, or experiments described herein. In using such information or methods they should be mindful of their own safety and the safety of others, including parties for whom they have a professional responsibility. To the fullest extent of the law, neither the Publisher nor the authors, contributors, or editors, assume any liability for any injury and/or damage to persons or property as a matter of products liability, negligence or otherwise, or from any use or operation of any methods, products, instructions, or ideas contained in the material herein. Library of Congress Cataloging-in-Publication Data A catalog record for this book is available from the Library of Congress British Library Cataloguing-in-Publication Data A catalogue record for this book is available from the British Library ISBN: 978-0-12-821605-7 For information on all Elsevier publications visit our website at https://www.elsevier.com/books-and-journals

Publisher: Brian Romer Acquisitions Editor: Lisa Reading Editorial Project Manager: Leticia Lima Production Project Manager: Manju Thirumalaivasan Cover Designer: Greg Harris

Typeset by TNQ Technologies

Dedicated to: Preben Maegaard The renewable energy pioneer, founder and director emeritus of Nordic Folkecenter, Denmark, co-author of the first edition

Citations

The human being has three ways to learn: first, by reflection, which is the noblest; second, by imitation, which is the easiest; and third, by experience, which is the bitterest. Confucius

Off-grid renewable energy systems have transformed our ability to deliver secure, affordable electricity to rural communities all over the world, and are playing a vital role in breaking a cycle of energy poverty that has held back socio-economic progress for hundreds of millions of people. Adnan Amin, Director-General, International Renewable Energy Agency

Coming together is a beginning, keeping together is progress, and working together is success. Henry Ford

vii

j

Foreword Since the first publication of the book in 2013, much has changed regarding distributed renewable energy technologies for off-grid communities. With the third decade of the 21st century presenting humanity with serious and profound challenges that connect health, food, water, energy, and interwoven international economies, the second edition is a welcome and much needed addition to essential reading for students, policy-makers, and an informed public that wants to be involved in the ongoing energy transformation and transition. The book builds on the strengths of its first edition, which include a clear presentation of the various renewable energy technologies themselves, together with graphics and tables that connect the easily comprehended theoretical framework to many practical and tangible examples. These have been detailed and expanded to include multiple case studies that allow the reader to see the case in point; these technologies are rapidly being deployed throughout the world and have an enormous impact on people’s occupations and lives. I had the good fortune of being able to use Distributed Renewable Energies for Off-Grid Communities in the graduate program of Santa Clara University’s School of Engineering. The course, “Distributed and Renewable Energy for the Developing World,” took the book as its text and inspiration. As such, and with the generous guidance and advice of the book’s primary author and editor, Professor Dr. Nasir El Bassam, it surveyed energy engineering and entrepreneurship in emerging market countries, with an emphasis on strategies for coping with the absence of a grid. It analyzed strategies for energy generation, transmission, and storage at the household, community, and regional scales, drawing from sector and case studies in the developing world. In reading the second edition of book, one finds that it continues to connect and explain concepts using a unique and easily comprehended framework. This guiding principle is described near the introduction to the book and is called the energy trilemma. Briefly put, it is the nexus among environmental sustainability, energy, equity, and energy security. These are the key concepts for providing economic development and prosperity for more than 7.6 billion people as the economy of the world is reconstructed and configured to be more resilient. Make no mistake, the book presents a road map

xvii

j

xviii

Foreword

for basic needs, as well as work that has been done and the work that remains. It connects both energy generation and use with food production and economic development so that engineers, scientists, planners, and financial institutions can help to determine a suitable mix of appropriate technologies and policy approaches for a given location. The locations presented include those in India, Iraq, Vatican City, Germany, and Tunisia. From its beginning, after describing the trilemma challenge, the book carefully and systematically describes the necessary restructuring for energy generation and supply as well as demand, and then transitions to present the road map for renewable energy communities, as well as planning and scenarios that include integrated renewable energy communities and farm production. Both arid and semiarid regions are considered. There and throughout the book, energy and food requirements are connected and described, together with modeling approaches. As is typical for each chapter, a healthy set of references is provided so that the reader can continue to explore those topics of particular interest further. The energy potential analysis from renewable energy resource availability is discussed with connections to various forms of solar energy. Wind energy is described (at small, medium, and large scales). A detailed description follows of the various types of biomass and bioenergy systems, all tied to distributed and small-scale production. Hydrogen generation (as an energy carrier) is then described, as well as hydropower and water power for both landbased and marine energy-generating systems. Of course, geothermal, energy storage, smart grids, and electric vehicles are covered in subsequent chapters. A feature much appreciated by students and policy-makers alike is the comprehensive coverage of the whole range of renewable energy options that are available. Throughout the book and in several appendices, case studies are presented for various technological approaches suitable for particular countries and regions. Another unique part of the second edition of the book is the consideration of mobility and transportation in fuels, infrastructure, and approaches. It is also refreshing to see the discussion on rethinking mobility and how it ties to urban planning. Yet another addition to the second edition is water purification in all of its various forms. The book’s final chapters complete the work by describing the connections among energy resiliency, digitization, blockchain, and the energy sector. The author presents a Marshall Plan for empowering urban and rural communities and the transition to an achievable vision for regional and world peace. Because the book contains a glossary as well as abbreviations

Foreword

xix

and acronyms, and considering the style and format of its presentation, readers from a variety of backgrounds will easily read and understand it. It achieves a holistic approach, yet it is detailed. A multitude of distributed energy systems throughout the world are preferred over central power plant generation schemes because of such concerns. In addition, they allow for improved access, social equity, and rapid scale-up through incremental additions to existing projects that are viable and cost-effective. This provides investors and the financial community with the confidence that they need to use the concepts for distributed renewable energies to empower off-grid communities in the 21st century to create sustainable, competitive, secure, and prosperous societies. Greg P. Smestad, PhD San José, California

Preface Energy is directly related to the most critical economic and social issues that affect sustainable development: mobility, job creation, income levels and access to social services, gender and racial disparity, population growth, food production, climate change, environmental quality, industry, communications, and regional and global security issues. Many of the crises on our planet arise from the desire to secure supplies of raw materials, particularly energy sources, at low prices. The International Energy Agency forecasts that the world primary energy demand will grow by 1.6%/year on average up to 2030. Current approaches to energy are unsustainable and nonrenewable. Today, the world’s energy supply is largely based on fossil fuels, nuclear power, hydro, and others (IEA).

International Energy Agency, global annual average change in energy production by fuel, 1971e2017. These sources of energy will not last forever and have proven to be contributors to our environmental problems. In less than three centuries xxi

j

xxii

Preface

since the industrial revolution, humanity has already burned roughly half of the fossil fuels that accumulated under the earth’s surface over hundreds of millions of years. Nuclear power is also based on a limited resource (uranium), and the use of nuclear power creates such incalculable risks that nuclear power plants cannot be insured. After 50 years of intensive research, no single, safe, long-term disposal site for radioactive waste has been found. Renewable energy offers our planet a chance to reduce carbon emissions, clean the air, and put our civilization on a more sustainable footing. Renewable sources of energy are an essential part of an overall strategy of sustainable development. They help reduce dependence on energy imports, thereby ensuring a sustainable supply and climate protection. Furthermore, renewable energy sources can help improve the competitiveness of industries over the long run and have a positive impact on regional development and employment. Renewable energies will provide a more diversified, balanced, and stable pool of energy sources. The main targets of this book will be a comprehensive and solid contribution to enlighten the vital role of developing decentralized and distributed renewable energy production and supply for off-grid communities along with their technical feasibilities to meet the growing demand for energy, and to face current and future challenges of limited fossil and nuclear fuel reserves, global climate change, and financial crises. It presents various options and case studies related to the potential of renewable energies and future transition options along with their environmental, economic, and social dimensions. With rapid and continued growth in the world, it is no longer a question of when we will incorporate various renewable energy sources into the mix, but how fast the transition can be managed. The impact of COVID-19 on renewable energy; how economic stimulus packages need to be built around renewable energy and energy efficiency; how to continue informing, influencing, and debating online to advance the use of renewable energy; and perhaps most important, how any crisis is an opportunity to step back, learn, adjust, and change. The momentum behind COVID-19 is enormous. We have an unprecedented opportunity to accelerate much needed change! For both climate and sustainable development reasons, we need to question the way we are doing things: how we produce, consume and finance production of goods, how we move those goods and provide services, how we trade and share resources, and how we create a more robust and resilient infrastructure. COVID-19 raises the same fundamental questions but with an urgency

Preface

xxiii

that, unfortunately, those of us in the climate and development community have been unable to communicate successfully (Rana Adib, Executive Director, REN21, 2020). This book is an attempt to outline the necessary information and concepts so that we can, as many are calling it, “Build Back Better” https://en.wikipedia.org/wiki/Build_Back_Better. I hope that this book offers a platform and resource for planning to foster an improvement in energy generation and supply. We wish it to contribute to enlightening and understanding of the vital economic and social roles that distributed renewable energy can provide in meeting the growing demand for energy and facing current and future challenges of limited fossil fuel reserves, global climate change, and equitable economic development for all. IEA, Global annual average change in energy production by fuel, 1971-2017, IEA, Paris https://www.iea.org/data-and-statistics/charts/global-annualaverage-change-in-energy-production-by-fuel-1971-2017. The current century will witness a major transformation in how energy is acquired, stored, and used globally. At this point, nearly one-fifth of the way through the 21st century, changes are clearly discernible, but more profound ones are still to come. The challenges we face in carrying out these transformations range from scientific and technological to societal, cultural, and economic involving how we live, work, and play. The impetus for these changes comes from the deep impacts that both developed and developing societies have had on our planet’s environment during the past century and projections going forward regarding what will happen globally if we do not act. Real and projected urbanization together with growing global population make it clear that we must act now. The transition to a climate-neutral society and carbon-free energy generation is both an urgent challenge and an opportunity to build a better future for all. Nasir El Bassam, Ph.D. International Research Centre for Renewable Energy, www.ifeed.org. Chairperson, WCRE, World Council for Renewable Energy www. wcre.org. Scientific Advisory Board, Federal Association of Regenerative Mobility, Berlin www.brm-ev.de.

Acknowledgments Most grateful thanks are due to Marcia Lawton Schlichting, who did the most arduous and time-consuming work of preparing the manuscript. I would also like to thank Greg P. Smestad, Thamer Ahmed Mohamed, Daniele Pagani, and Lothar Schlichting for their contributions to the book. The editor wishes to thank the supporting team at Elsevier, the Acquisitions Editor, Lisa Reading; Editorial Project Manager, Letícia Lima; Production Project Manager, Manju Thirumalaivasan; and others for the substantial assistance provided.

xxv

j

CHAPTER ONE

What Kind of Energy Does the World Need? N. El Bassam International Research Centre for Renewable Energy, IFFED.org, Germany

Contents Abstract 1.1.1 What kind of energy does the world need? 1.1.1.1 Introduction 1.1.2 Distributed renewable energy for energy access References Abstract 1.2.1 Energy trilemma index 1.2.2 Dimensions 1.2.3 Monitoring the sustainability of national energy systems Reference Abstract 1.3.1 Distribution 1.3.2 Distributed energy generation 1.3.3 Distributed energy supply 1.3.4 Community power 1.3.5 Off-grid systems 1.3.6 Concluding remarks Further reading

6 6 6 7 11 12 12 13 15 19 20 21 22 22 23 24 24 26

Along with the developments of the past three-quarters of a century have come disparities, energy injustice, and major environmental threats, in particular climate change. So, what should governments do to quell the civil unrest and growing populism resulting from these inequalities? Certainly, in our view, one clear answer is to transform our energy system to a renewables-based distributed system resulting in much greater economic opportunities for all people, energy justice, and environmental recovery and improvement (David Renné, Sunburst ISES Newsletter, 2019). There were also side events where other important publications were presented. For example, Rana Abib, Executive Director of REN 21, Distributed Renewable Energies for Off-Grid Communities, Volume 2 ISBN: 978-0-12-821605-7 https://doi.org/10.1016/B978-0-12-821605-7.00028-3

© 2021 Elsevier Inc. All rights reserved.

1

j

2

N. El Bassam

announced two new publications: “Perspectives on the Global Renewable Energy Transition” and “Asia and the Pacific Renewable Energy Status Report.” The Perspectives report provides key takeaways from the 2019 Global Status Report, highlighting important facts such as that renewables accounted for 64% of all new electricity generation in 2018; and that same year, nine countries generated more than 20% of electricity with wind and solar photovoltaics (PV). Furthermore, cities are taking a leading role in adopting some of the most ambitious targets globally, and at least 100 cities worldwide now use 70% or more renewable electricity. However, the report also notes that slow growth in the use of renewables in heating and cooling needs to be addressed to achieve decarburization in all of our energy sectors. The point I would like to make here is that the current global unrest that we are seeing, which at times has become violent, especially when economic and political factors come into play, has in part been caused by, but also can be solved by, the way we produce and use energy to fuel our economic system. We are already seeing massive global demonstrations against the way in which we use energy, which has led to the climate crisis, such as the peaceful demonstrations by FridaysforFuture protesting the lack of government action on climate change, and climate demonstrations adopting civil disobedience tactics such as those of the Extinction Rebellion (BBC, 2019, https://www.bbc.com/news/uk-48607989). The movement is reported critically in part because many of its protests are not legal. For this reason, arrests have been made worldwide. On Oct. 14, 2019, the London police issued a demonstration ban for the movement; it lifted the ban 4 days later because the measure was no longer necessary because the wave of protests had ended. The British High Court of Justice brought an action against the legality of the ban on Nov. 6, 2019. ISES is one of many key like-minded organizations working hard to communicate how the renewable energy transformation will create immense and more equitable economic benefits, energy access and security, and environmental recovery. The installed capacity of solar PV systems has grown from 23 GW at the start of 2010 to around 600 GW currently and solar thermal systems from 203 GW-thermal to over 500 GW-thermal during that period. Over the same period, wind energy installed capacity grew from 159 to over 600 GW and concentrating solar thermal power grew from less than 1 GW to over 5.5 GW. Today, the global power sector is powered by over 26% renewable energy, and renewables surpass traditional energy sources for new installed power capacity around the world. Government policies and national targets for renewable energy deployments have grown

What Kind of Energy Does the World Need?

3

significantly through the decade, and utility acceptance of variable renewable energy supply has expanded in many countries. Clearly, the 2010s saw great strides in a global clean energy transformation; for all purposes, it could be called the decade of renewable energy. However, during this decade, the scope of our challenge in addressing key energy-related issues such as environmental impacts, energy security, and access to finance also became paramount. At the start of the decade, there was hope that annual global CO2 emissions had peaked at around 30 Gt and would start to decrease owing to all of the clean-energy initiatives and growing acknowledgment of the need to combat climate change. Nevertheless, despite the signing of the Paris Climate Agreement in the middecade, annual global CO2 concentrations actually increased and are now around 33.1 Gt, much of this still resulting from coal-fired power generation and the use of fossil fuels in the transportation sector. We see these challenges as indicators that ISES’s work is nowhere near completed: despite the impressive growth of solar technologies from laboratory experiments to commercial success that has happened since the early days of ISES, dating back to the middle of the past century, much work remains to urge governments and civil society to be more ambitious in addressing climate change, and to articulate the multitude of environmental, economic, and energy security benefits of a 100% renewable energy system (Renné, D., Sunburst ISES Newsletter 2019). These numbers are derived from the REN21 2019 Global Status Report and the IEA Global Energy and CO2 Status Report. A FERC report confirmed the rise of renewables above coal, gas, oil, and nuclear combined. According to a review by the SUN DAY Campaign of data issued by FERC, the mix of renewable energy sources (i.e., biomass, geothermal, hydropower, solar, and wind) provided 57.26% of new US electrical generating capacity added in 2019, swamping that provided by coal, natural gas, oil, and nuclear power combined (Kenneth Bossong, 2020). FERC’s latest monthly Energy Infrastructure Update report (with data through Dec. 31, 2019) revealed that renewable sources (i.e., biomass, geothermal, hydropower, solar, and wind) accounted for 11,857 MW of new generating capacity by the end of the year. That is a third more (33.97%) capacity than that of natural gas (8557 MW), nuclear (155 MW), oil (77 MW), and coal (62 MW) combined. Renewables have also surpassed 22% (i.e., 22.06%) of the nation’s total available installed generating capacity, further expanding their lead over

4

N. El Bassam

coal capacity (20.89%). Among renewables, wind can boast the largest installed electrical generating capacity: 8.51% of the US total, followed by hydropower (8.41%), solar (3.49%), biomass (1.33%), and geothermal (0.32%). Thus, wind and solar combined account for 12.0% of the nation’s electrical generating capacity. Moreover, FERC foresees renewables dramatically expanding their lead over fossil fuels and nuclear power in terms of new capacity additions in coming years. Net generating capacity additions (i.e., proposed additions under construction minus proposed retirements) for renewable sources total 48,254 MW: wind: 26,403 MW; solar: 19,973 MW; hydropower: 1460 MW; biomass: 240 MW; and geothermal: 178 MW. By comparison, net additions for natural gas total 21,090 MW whereas the installed capacities for coal, nuclear, and oil are projected to drop by 18,857, 3391, and 3085 MW respectively. In fact, FERC reported no new coal capacity in the pipeline over coming years. Thus, although net new renewable energy capacity is projected to be nearly 50,000 MW greater within the next few years, that of fossil fuels and nuclear power combined will decline by over 4200 MW. New wind capacity alone will be greater than that of natural gas, whereas that of wind and solar combined will more than double new gas capacity. Moreover, if FERC’s data prove correct, renewable sources will account for more than a quarter of the nation’s total available installed generating capacity (25.16%) whereas coal will drop to 18.63% and that of nuclear and oil will decrease to 8.29 and 2.95%, respectively. Natural gas will increase its share, but only slightly, from 44.67 to 44.78%. As the executive director of the SUN DAY Campaign, I believe that the rapid growth of renewables and the corresponding drop in electrical production by coal and oil provides a glimmer of hope for slowing the pace of climate change. In addition, renewables’ continued expansion in the near future, as forecast by FERC, suggests that with supportive governmental policies, these technologies could provide an even greater share of total US electrical generation (https://www.renewableenergyworld.com/2020/ 03/09/new-ferc-report-confirms-the-rise-of-renewables-above-coal-gas-oil-a nd-nuclear-combined/?utm_medium¼email&utm_campaign¼rew_weekly_ newsletter&utm_source¼enl&utm_content¼2020-03-11). On Mar. 3, 2020, the Government of the Australian state of Tasmania announced a long-term strategy for the island, in which it set not a 100% but a 200% renewable energy target for 2040. The announcement, made by Tasmanian Premier Peter Gutwein, follows a previous commitment to 100% renewable energy by 2022.

What Kind of Energy Does the World Need?

5

The new strategy aims to cover Tasmania’s domestic energy supply, but also to export renewable energy to other parts of Australia and potentially to other countries. The main technologies, which will achieve this goal and make Tasmania a green powerhouse, will be hydropower, wind power, and hydrogen. Gutwein declared that a detailed Renewable Energy Action Plan would be released in April and that hydrogen production for domestic use and for export latest by 2027 would be an important part of the plan. The government intended to boost the rollout of the state’s hydrogen economy with $50 million in public funds. The Tasmanian government expected that the new plan would lead not only to new jobs and a substantial contribution to reducing Australia’s greenhouse gas emissions, but potentially to a combined investment of $7.1 billion into the Tasmanian economy (http://www.premier.tas.gov. au/releases/state_of_the_the_state_address). Tackling poverty, which affects one-third of the world’s population, and serving the needs of the unserved should be our priorities. We have the knowledge and technologies to achieve these goals. What is needed is for all to be honest, faithful, and credibledto us as well as to others, to live in peace and dignity: not only for part of the world, but for all. Lack of sufficient energy supply leads to a lack of development. In countries and regions with energy shortages, populations suffer the most. It is imperative that with the era of fossil fuel coming to an end, future initiatives for energy supply be based on renewable energy. With this book, we pledge to use our knowledge, voices, and determination to: - Persevere, each in our own way, nurtured by the cultural wellsprings that are our heritage, whether from Asia, Africa, Europe, South and North America, or elsewhere; and. - Join hands and work together, inspired, r-energized, and committed TO DO ALL WE CAN . TO MAKE REAL THE WORLD OF OUR DREAMS!

6

N. El Bassam

CHAPTER 1.1

Distributed renewable energy Strategies toward achieving energy security, equity, and the environmental sustainability of energy systems throughout the transition process

Abstract We have reached our limits as the result of the excessive use of fossil fuels and related technologies that benefit a few financially but leave the rest to cope with the consequences. Developing the energy supply would automatically improve the major issues of sustainable development. Keywords: Distributed renewable energy, Types of energy access

1.1.1 What kind of energy does the world need? 1.1.1.1 Introduction It is no secret that we have reached our limits as a result of the excessive use of fossil fuels and related technologies that benefit a few financially but leave the rest to cope with the consequences. Among these unfortunate outcomes are health hazards, security issues, dwindling public services, and restricted access to education and job opportunities. Developing the energy supply would automatically improve the major issues of sustainable development: poverty, job creation, income levels, and access to social and economic services, gender disparity, population growth, agricultural production, climate change, the environment, security issues, and migration. Today, around two billion people still lack access to a reliable supply of electricity. Our challenge in the 21st century will be to provide energy for a further five to seven billion people while cutting emissions by half. By 2050, humanity will need three earths to supply enough resources to meet the growing demands for energy. We cannot continue to manage our resources in such a negligent manner. This option does not exist. We have to consider the needs of future generations.

What Kind of Energy Does the World Need?

7

With rapid and continued population growth in the world, depletion of natural resources, and climate change, it is no longer a question of when we will incorporate various renewable energy sources into the mix, but how fast the transition can be managed. In less than three centuries since the industrial revolution, humanity has already burned roughly half of the fossil fuels that accumulated under the earth’s surface over hundreds of millions of years. Nuclear power is also based on a limited resource (uranium). Although some fossil energy resources might last a little longer than predicted, especially if additional reserves are discovered, the main problem of scarcity will remain, and this represents the greatest challenge to humanity. Renewable energy offers our planet a chance to reduce carbon emissions, clean the air, and put our civilization on a more sustainable footing. Renewable sources of energy are an essential part of an overall strategy of sustainable development. They help reduce dependence on energy imports, ensuring a sustainable supply and climate protection. Furthermore, renewable energy sources can help improve the competitiveness of industries over the long run and have a positive impact on regional development and employment. Renewable energies will provide a more diversified, balanced, and stable pool of energy sources. Energy cannot be created; it can be converted from one form to other by technical, biological and chemical means, such as gas, oil, coal, solar, and wind energy, into heat and power energy, biomass into heat, electricity or biofuels, and so on.

1.1.2 Distributed renewable energy for energy access According to the 20th-century model of energy distribution, large power plants fueled by coal, hydro, or gas generated electricity that was distributed through a centralized grid. The picture has changed. Advancing technology has diversified the grid, adding new sources of energy generation and two-way power flows. Utility-scale wind and solar farms are supplying an increasing proportion of power. Enter distributed energy resources, known as distributed energy resources (DER): small-scale units of local generation connected to the grid at the distribution level. DERs can include behind-the-meter renewable and nonrenewable generation, energy storage, inverters (electronic devices that change DC to AC), electric vehicles, and

8

N. El Bassam

other controlled loads (separately metered appliances such as hot water systems). DER is also composed of new technology such as smart meters and data services (ARENA, 2020). DER penetration is growing every year. Increased demands on the nation’s electrical power systems and incidences of electricity shortages, power quality problems, rolling blackouts, and electricity price spikes have caused many utility customers to seek other sources of high-quality, reliable electricity. DERs, small-scale power generation sources located close to where electricity is used (e.g., a home or business), provide an alternative to or enhancement of the traditional electric power grid. A limiting factor is hosting capacity, or the amount of DER that can be connected to a distribution network and operated within its technical limits. DERs can be incorporated into the grid where no threats to safety, reliability, or other operational features exist and no infrastructure upgrades are required. In many cases, however, grid modernization is necessary to integrate DERs safely into the network. DERs are a faster, less expensive option to the construction of large, central power plants and high-voltage transmission lines. They offer consumers the potential for lower cost, higher service reliability, high-power quality, increased energy efficiency, and energy independence. The use of renewable distributed energy generation technologies and green power such as wind, PV, geothermal, biomass, and hydroelectric power can also provide a significant environmental benefit. DER deployment is increasing in the developing world despite limited financial support. Approximately 1.2 billion people (about 16% of the global population) live without electricity, and about 2.7 billion people (38% of the global population) are without clean cooking facilities. The vast majority of people without access to both electricity and clean cooking facilities are in subSaharan Africa and the Oceania region; most live in rural areas. The old paradigm of energy access through grid extension alone is becoming obsolete, as ground-level customer demand is motivating hundreds of millions of households to generate their own energy to feed off-grid units or community-scale minigrids. Mobile technology Pay-asYou-Go (PAYG) business models, the availability of microloans, the viability of microgrids, and falling technology prices continue to support DER deployment worldwide. The most popular business models within

What Kind of Energy Does the World Need?

9

the DER sector in 2016 were distributed energy service companies for mini/ micro/picogrids, the PAYG model for stand-alone systems, and microfinance and microcredit. In 2018, global energy demand increased an estimated 2.3%, the greatest rise in a decade. This was the result of strong global economic growth (3.7%) and higher heating and cooling demands in some regions. China, the United States, and India together accounted for almost 70% of the total increase in demand. Because of a rise in fossil fuel consumption, global energy-related CO2 emissions grew an estimated 1.7% during the year. As of 2017, renewable energy accounted for an estimated 18.1% of total final energy consumption (TFEC). Modern renewables supplied 10.6% of TFEC, with an estimated 4.4% growth in demand compared with 2016. Traditional use of biomass for cooking and heating in developing countries accounted for the remaining share. The greatest portion of the modern renewable share was renewable thermal energy (an estimated 4.2% of TFEC), followed by hydropower (3.6%), other renewable power sources including wind power and solar PV (2%), and transport biofuels (about 1%) (Figures 1.1.1 and 1.1.2). The best route to go is renewable energy: solar energy, wind power, hydro power, bioenergy, hydrogen and fuel cells, geothermal power, and other forms of energy such as that from tides, the oceans, and hot hydrogen

Figure 1.1.1 Estimated renewable energy share of total energy consumption for 2017. Courtesy of REN21 (2019).

10

N. El Bassam

Figure 1.1.2 Renewable energy today in a global context. PV, photovoltaics. From REN 21 GSR (2019).

fusion. Before the development of coal in the mid-19th century, nearly all energy used was renewable. As far back as 400,000 years ago, humans used biomass to light fires; we sailed the seas with the power of wind and we used waterpower to crush grains. However, there is a certain disadvantage to renewable energy sources: storage capacity. Other countries such as Denmark, Costa Rica, Nicaragua, and Sweden are also on the way to 100%. Although the path is sometimes rocky, you still have to walk it if you want success in the long run.

What Kind of Energy Does the World Need?

11

However, the expansion of power plants for green energy alone is not enough. Time after time, grid operators must compensate for imbalances in electricity consumption and switch on conventional power plants to cope with consumption peaks and prevent total blackouts. The problems are that the sun usually does not shine exactly when the most electricity is needed, or that there is no wind when the whole country is watching football on TV. Gas power plants, for example, must supply electricity as needed, which drives up costs. The biggest problem is how to store energy that is produced in a green and sustainable way. Of course, storing energy in batteries is obvious. Solar systems on the roof of a family home can feed a battery in the basement, which can be used when needed. Depending on the size of the solar panels and battery, as well as the average hours of sunshine at a given location, the energy consumption of a typical household could even be covered completely off-grid. So, which kind of batteries are we talking about? There are different types of batteries: lead, lithium-ion, which are used in smartphones and ecars, and redox-flow, which are used primarily for stationary applications such as wind power plants. However, there is a problem with batteries: prices will rise with demand. Because batteries depend on rare raw materials, they are also a precious goods and unsuitable for mass production in a way that would be needed to power an entire society. Nevertheless, some providers have set themselves the goal of supplying the market with such batteries and creating a decentralized power grid. Pioneers in this field are companies such as Tesla.

References Australian Renewable Energy Agency (ARENA), 2020. Estimated Renewable Energy Share of Total Energy Consumption for 2017, REN21, 2019. Renewable Energy Today in Global Context, 2019. REN 21, GSR.

12

N. El Bassam

CHAPTER 1.2

Using distributed energy resources to meet the trilemma challenges Abstract Distributed generation is also called on-site, dispersed, embedded, or decentralized generation. Decentralized energy, or distributed energy, generates electricity, heat, and fuel from many small energy sources. In the future, when planning a new power and heating or transportation fuel system on a clean sheet of paper, there will be no big fossil fuelebased power stations or large high-voltage transmission lines. This book illustrates the future of the off-grid power supply, because there is no advantage to having international and/or interregional grid structures. Keywords: Community power, Distributed generation, Off-grid systems

1.2.1 Energy trilemma index The world is undergoing an unprecedented energy transition from a system based on carbon-intensive fossil fuels to a one based on lowcarbon, renewable energy, driven by the twin imperatives of mitigating climate change and generating economic prosperity. The speed of change and the effectiveness of individual governments to develop and implement policies to deliver energy sustainability vary across countries and geographies. The World Energy Council recognizes the value of adopting a whole energy systems approach in providing the benefits of sustainable energy to all. This energy transition is a connected policy challenge. Success involves managing the three core dimensions: energy security, energy equity, and the environmental sustainability of energy systems throughout the transition process (Worldenergy, 2019). The Energy Trilemma Index, developed in partnership with Oliver Wyman, provides an objective rating of national energy system performance across these three Trilemma dimensions. The Trilemma was created to

What Kind of Energy Does the World Need?

13

support an informed dialogue about improving energy policy to achieve energy sustainability, by providing decision-makers with information about countries’ relative performance. Objectively comparing the success of energy systems around the globe is challenging, but a high-level ranking of performance against a set of benchmark indicators helps to start a conversation about policy coherence and effectiveness. Deeper analysis at the regional and national levels can give policy-makers real insights into trajectories and outlooks, informing future priorities. To provide greater insight, we have evolved the methodology for the 2019 Trilemma and, for the first time, introduced visualization of historical trends to enable the Trilemma performance of individual countries to be tracked back two decades to 2000. The new time-series analysis provides insights into a country’s historical trends, challenges, and opportunities for improvements in meeting energy goals now and in the future. The Index demonstrates the impact of the varying policy pathways that countries have taken in each of the dimensions over the past 20 years. Looking at these trends can inform a dialogue on national energy policy to promote coherence and integration to enable better calibrated energy systems in the context of the global energy transition challenge.

1.2.2 Dimensions The World Energy Trilemma Index has been prepared annually since 2010. It presents a comparative ranking of 128 countries’ energy systems. It provides an assessment of a country’s energy system performance, reflecting balance and robustness in the three Trilemma dimensions. It reflects a nation’s capacity to meet current and future energy demand reliably, and withstand and bounce back swiftly from system shocks; and it assesses a country’s ability to provide universal access to affordable, fairly priced, and abundant energy for domestic and commercial use (Figures 1.2.1 and 1.2.2).

14

N. El Bassam

Figure 1.2.1 The Trilemma dimensions.

Figure 1.2.2 World Energy Trilemma top 10 performers.

What Kind of Energy Does the World Need?

15

1.2.3 Monitoring the sustainability of national energy systems The world is undergoing an unprecedented energy transition from a system based on carbon-intensive fossil fuels to a one based on lowcarbon, renewable energy, driven by the twin imperatives of mitigating climate change and generating economic prosperity. The speed of change and the effectiveness of individual governments to develop and implement policies to deliver energy sustainability vary across countries and geographies. The World Energy Council recognizes the value of adopting a whole energy systems approach in providing the benefits of sustainable energy to all. This energy transition is a connected policy challenge. Success involves managing the three core dimensions: energy security, energy equity, and the environmental sustainability of energy systems throughout the transition process. The Council’s World Energy Trilemma Index, developed in partnership with Oliver Wyman, provides an objective rating of national energy system performance across these three Trilemma dimensions. We have created the Trilemma to support an informed dialogue about improving energy policy to achieve energy sustainability, by providing decision-makers with information on countries’ relative performance. Objectively comparing the success of energy systems around the globe is challenging, but a high-level ranking of performance against a set of benchmark indicators helps start a conversation about policy coherence and effectiveness. Deeper analysis at regional and national levels can give policy-makers real insights into trajectories and outlooks, informing future priorities. To provide greater insight, a methodology was evolved for the 2019 Trilemma; for the first time, it introduced a visualization of historical trends to enable the Trilemma performance of individual countries to be tracked back two decades to 2000. The new time-series analysis provides insights into a country’s historical trends, challenges, and opportunities for improvements in meeting energy goals now and in the future. The Index demonstrates the impact of the varying policy pathways that countries have taken in each of the dimensions over the past 20 years. Looking at these trends can inform a dialogue on national energy policy to promote coherence and integration to enable better calibrated energy systems in the context of the global energy transition challenge. Ten countries achieve the top AAA balance grade in the 2019 World Energy Trilemma Index, representing top quartile performance in every

16

N. El Bassam

dimension. Since 2000, no countries have consistently improved in each dimension every year. Instead, most show historical trends with a variety of peaks and troughs in a general upward direction. Overall Trilemma performance for 119 countries over 20 years has improved; only nine countries saw their overall performance decline. The rate of improvement in overall Trilemma performance also increased as the transition progressed and encouraged countries to improve their energy policies (Figures 1.2.3e1.2.8). Finally, some features that led to the ranking of the 10 states that occupy first places are that: • They all have stable economic and political systems. • They all barely have conflicts with other states or within their own population.

Figure 1.2.3 Middle East and Gulf State region Trilemma balance.

Figure 1.2.4 Latin America and Caribbean region Trilemma balance.

What Kind of Energy Does the World Need?

17

Figure 1.2.5 Europe region Trilemma balance.

Figure 1.2.6 Asia region Trilemma balance.

Figure 1.2.7 Africa region Trilemma balance.

• They all have a high investment volume because of their knowledge and insight, and the conviction that RES is essential for an intact environment. • The state and the respective population were in agreement and started the changeover at an early stage (Figure 1.2.9)

Figure 1.2.8 North American region Trilemma balance.

Figure 1.2.9 (A) World Trilemma ranking for 128 countries, part 1 (ranks 1e64) 2019. (B) World Trilemma ranking for 128 countries, part 2 (ranks 65e128) 2019.

What Kind of Energy Does the World Need?

Figure 1.2.9 (continued).

Reference https://trilemma.worldenergy.org, 2019.

19

20

N. El Bassam

CHAPTER 1.3

Scope of the book Abstract This book offers a platform for planning and fostering improvements in distributed energy generation and supply and contributes to an understanding of the vital economic and social roles of distributed renewable energy to meet the growing demand for energy and face current and future challenges of global climate change, financial crises, and energy poverty in some regions. A multitude of distributed energy systems throughout the world are preferred over central power plant generation schemes because of such concerns. In addition, they allow for improved access, social equity, and rapid scale-up through increasing additions to existing projects that are found to be viable and cost-effective. This provides investors and the financial community with the confidence they need to use the concepts for distributed renewable energies to empower off-grid communities in the 21st century to create sustainable, competitive, secure, and prosperous societies. Keywords: Community power, Distribution, Energy generation, Energy supply, Off-grid systems This book offers a platform for planning and fostering improvements in distributed energy generation and supply and contributes to an understanding of the vital economic and social roles of distributed renewable energy to meet the growing demand for energy and face current and future challenges of global climate change, financial crises, and energy poverty in some regions. A multitude of distributed energy systems throughout the world are preferred over central power plant generation schemes because of such concerns. In addition, they allow for improved access, social equity, and rapid scale-up through increasing additions to existing projects that are found to be viable and cost-effective. This provides investors and the financial community with the confidence they need to use the concepts for distributed renewable energies to empower off-grid communities in the 21st century to create sustainable, competitive, secure, and prosperous societies.

21

What Kind of Energy Does the World Need?

1.3.1 Distribution Distribution means the delivery of electricity to the retail customer’s home or business through low-voltage distribution lines. Distributed generation is also called on-site, dispersed, embedded, or decentralized generation. Decentralized energy, or distributed energy, generates electricity, heat and fuels from many small energy sources. It reduces the amount of energy lost in transmitting electricity because the electricity is generated near where it is used, perhaps even in the same building. This also reduces the size and number of power lines that must be constructed. Both electric demand reduction (energy conservation, load management, and so forth) and supply are generated at or near where the power is used. A distributed generation system involves amounts of generation located on a utility’s distribution system for the purpose of meeting local (substation level) peak loads and/or displacing the need to build additional (or upgrade) local distribution lines (Figure 1.3.1).

ENERGY energy system losses

CAPACITY generation capacity transmission & distribution capacity DPV installed capacity

GRID SUPPORT SERVICES reactive supply & voltage control regulation & frequency response energy & generator imbalance synchonized & supplemental operating reserves scheduling, forecasting, and system control & dispatch

GRID SERVICES

FINANCIAL RISK

FINANCIAL

SECURITY RISK

fuel price hedge market price response

reliability & resilience

SECURITY ENVIRONMENTAL ENVIRONMENTAL

SOCIAL

carbon emissions (CO2) criteria air pollutants (SO2, NOx, PM) water land

SOCIAL economic development (jobs and tax revenues)

Figure 1.3.1 Benefit and cost categories for solar distributed generation value analysis. DPV; PM, particulate matter. From https://www.e-education.psu.edu/eme444/node/365, 2020.

22

N. El Bassam

1.3.2 Distributed energy generation Distributed generation is also defined as the installation and operation of small modular power-generating technologies that can be combined with energy management and storage systems. It is used to improve the operations of electricity delivery systems at or near the end user. These systems may or may not be connected to the electric grid. When electricity is generated at or near where it is going to be used (the load center), this is called distributed generation. Solar and wind are both widely used for distributed generation, but so are nonrenewable sources such as diesel generators. The US Environmental Protection Agency defines distributed generation as “a variety of technologies that generate electricity at or near where it will be used, such as solar panels and combined heat and power.”

1.3.3 Distributed energy supply Typical distributed power sources in a feed-in tariff scheme have low maintenance, low pollution, and high efficiencies. In the past, these traits required dedicated operating engineers and large complex plants to reduce pollution. However, modern embedded systems can provide these traits with automated operation and renewables, such as sunlight, wind, and geothermal. This reduces the size of a profitable power plant (Figure 1.3.2). In the future, when planning a new power and heating or transportation fuel system on a clean sheet of paper, there will be no big fossil fuelebased power stations or large high-voltage transmission lines. Each community in this new energy supply structure will have a variety of supply technologies based on solar, wind, biomass, and other locally available sources of energy. Solar and wind will be the primary sources of supply. Biomass will be especially important for complementary power and heat when solar and wind is insufficient. Balancing fluctuating power from solar and wind is necessary; chemical and thermal storage solutions can also be applied. Such future supply systems can be of many sizes. The smallest will be for one-family houses or settlements and the biggest for a region or city. Because small-scale technologies can be mass produced and therefore are cheap, it may well prove preferable and most economical to divide cities into many independent and autonomous decentralized systems. In this book, we will call them off-grid, because there is no advantage to having international

What Kind of Energy Does the World Need?

23

Figure 1.3.2 Relevance of distributed generation. (India Energy Portal, Distributed Generation). From http://www.indiaenergyportal.org/subthemes_link.php?themeid¼ 14&text¼dis_ge.

or interregional grid structures. Up to this point, the off-grid supply has referred to unserved areas in developing countries without a national grid. DER systems are small-scale power generation technologies (typically in the range of 3e10,000 kW) used to provide an alternative to or enhancement of the traditional electric power system. As the cost of fossil fuels increases, the cost of renewable energy technologies declines, and government support increases, investors, utilities, and governments are exploring ways to implement or invest in distributed renewable energy programs, projects, and companies.

1.3.4 Community power Community is a term that has different meanings for different people. In this book, a community is defined as a social group of any size whose members live in a specific place. The term thus relates to geographical proximity, or “communities of locality” (Walker, 2008), such as a neighborhood, town, district, or city. This book focuses on distributed energy that is generated and distributed to consumers within a geographic locality. Distributed

24

N. El Bassam

energy generation can be a continuum of energy generation, from a household or multiple buildings to a larger community. Some energy may be fed back into the electricity grid, but ideally, at least some of the total energy generated is distributed and consumed locally. The book profiles the special topic of community Powerecitizens’ power, referring to the development and ownership of renewable energy projects by local citizens and communities, including farmers, landowners, cooperatives, municipalities, local and regional developers, and utilities.

1.3.5 Off-grid systems Off-grid renewable energy systems have transformed our ability to deliver secure, affordable electricity to rural communities all over the world, and are playing a vital role in breaking a cycle of energy poverty that has held back socioeconomic progress for hundreds of millions of people. Adnan Z. Amin, Director-General International Renewable Energy Agency (2019).

Off-grid systems provide an independently regulated power supply that has at least the same reliability and quality as a public power grid. The term “off-grid” refers to not being connected to a grid. it is mainly used in terms of not being connected to the main or national transmission grid in electricity. Off-grid electrification is an approach to accessing electricity used in countries and areas with little recourse to electricity, owing to a scattered or distant population. It can be any kind of electricity generation. Electrical power can be generated on-site with renewable energy sources such as solar, wind, or geothermal; or with a generator and adequate fuel reserves (Figure 1.3.3). It can also connect to local and national grids to substitute for the energy supply generated by nuclear or other nonrenewable fuel sources; then, it is called green electricity in some industrialized countries (Figures 1.3.4 and 1.3.5).

1.3.6 Concluding remarks This book illustrates the future of the off-grid power supply, because there is no advantage to having international and/or interregional grid structures. So far, off-grid supply has generally referred to communities in developing countries without a national grid. Because of the decentralized character of renewable energies, the future rationale will be to apply

What Kind of Energy Does the World Need?

25

Figure 1.3.3 Stand-alone off-grid systems. From http://www.wholesalesolar.com/ products.folder/systems-folder/OFFGRID.

Figure 1.3.4 Off-grid system that can also be connected to the grid. Reprinted with permission ©2012 Home Power Inc., www.homepower.com.

off-grid technologies to all types of communities, urban and rural, in fossil fueleserved areas in industrialized countries, and for unserved regions in developing countries as well. What has been the exception thus far may change to become mainstream, owing to the transition to decentralized energy forms.

26

N. El Bassam

Figure 1.3.5 Combined power plant. From http://www.blog.thesietch.org/2007/12/30/ germany-going-100-renewable-or-yet-another-reason-why-america-is-falling-behind/. (Accessed 23 March 2010).

Similar fundamental changes have often appeared when considering the long historical perspective. Facing the end of the fossil fuel age and the enormous risks for the ongoing climate change, it is time to prepare for the exit of the fossil fuel era. There are no technological barriers. This is the important message of the book.

Further reading (BBC, 2019, https://www.bbc.com/news/uk-48607989) (https://www.renewableenergyworld.com/2020/03/09/new-ferc-report-confirms-the-rise-of-renewables-above-coalgas-oil-and-nuclear-combined/?utm_medium¼email&utm_campaign¼rew_weekly_ newsletter&utm_source¼enl&utm_content¼2020-03-11 Kenneth Bossong j 2020 (Worldenergy, 2019) Gangwar, R., April 2009. Building community resilience towards climate change adaptation through awareness and education. In: Paper Presented at the Seminar Energy and Climate in Cold Regions of Asia, pp. 21e24. Available from: http://india.geres.eu/ docs/Seminar_proceedings/3-Climate_Change_Impacts_and_Adaptation/Building% 20Community%20Resilience%20towards%20Climate%20Change%20Adaptation% 20through%20Awareness%20&%20Education.pdf. Maruf, N.I., 2016. Optimization of Distributed Energy Resources to Balance Power Supply and Demand in a Smart Grid. https://www.slideshare.net/shahmdalimran/optimizationof-distributed-energy-resources-to-balance-power-supply-and-demand-in-a-smart-grid. Penn State, 2020. College of EMS. Credit: Rocky Mountain Institute, “A Review of Solar PV Benefits & Cost Subsidies.” https://www.e-education.psu.edu/eme444/node/365. Renewable Communities, 2008. Renewable Communities: Moving Towards Community Resiliency [online] Available from: http://renewablecommunities.wordpress.com.

CHAPTER TWO

Restructuring future energy generation and supply N. El Bassam International Research Centre for Renewable Energy, IFFED.org, Germany

Contents 2.1 Basic challenges 2.2 Current and future energy supplies 2.3 Peak oil 2.4 Availability of alternative resources 2.5 Outlook References Further reading

27 29 30 31 33 36 37

2.1 Basic challenges By 2050, humanity will need two to three earths to cover its consumption of resources if we continue to manage our resources as business as usual. The global energy system currently relies mainly on hydrocarbons such as oil, gas, and coal, which together provide nearly 80% of energy resources. Traditional biomass such as wood accounts for 11% and nuclear for 6%, whereas all renewable sources combined contribute just 3%. With the exception of nuclear energy, energy resources are ultimately derived from the sun. Nonrenewable resources such as coal, oil, and gas are the result of a process that takes millions of years to convert sunlight into hydrocarbons. Renewable energy sources convert solar radiation, the rotation of the earth, and geothermal energy into usable energy in a far shorter time. In the International Energy Agency reference scenario, the world primary energy demand will have grown by 1.6% per year on average in 2006e30, from 472 EJ (11,730 million metric tons of oil equivalents [Mtoe]) to just over 714 EJ (17,010 Mtoe). Because of continuing strong economic growth, China and India account for just over half of the increase Distributed Renewable Energies for Off-Grid Communities, Volume 2 ISBN: 978-0-12-821605-7 https://doi.org/10.1016/B978-0-12-821605-7.00029-5

© 2021 Elsevier Inc. All rights reserved.

27

j

28

N. El Bassam

in the world’s primary energy demand between 2006 and 2030. Middle Eastern countries strengthen their position as an important demand center, contributing a further 11% to incremental world demand. Collectively, noneOrganization for Economic Cooperation and Development (OECD) countries account for 87% of the increase. As a result, their share of the world’s primary energy demand rises from 51 to 62%. Their energy consumption overtook that of the OECD in 2005 (OECD/IEA, 2008). The major challenges of today’s energy system are closely related to a wide range of essentials such as welfare, dignity, peace, nature, and sustainable development, including: • Limited oil, gas, and uranium resources • 2No access to electricity, gas, oil, or clean water for almost 2 billion of the world’s population • Increasing import dependency in most industrialized countries, including China and India • Energy prices and volatility (the time of cheap oil and gas is over). • Climate change and other environmental risks (energy accounts for 80% of all greenhouse gas emission). • Geostrategic tensions caused by scarce energy resources • The extracting, transport, processing, and use of fossil and nuclear fuels, which can eventually become threatening to nature and the existence of humanity (i.e., accidents in the Gulf of Mexico and Fukushima) • No single safe long-term storage facility worldwide for highly radioactive nuclear waste • Growing world population (in 2012, 7 billion; in 2050, 9e10 billion). Figure 2.1 effectively demonstrates the finite nature of fossil energy

Figure 2.1 Past, present and future energy sources. (From El Bassam (1992)).

Restructuring future energy generation and supply

29

resources and the vital role of renewable energies in satisfying the needs of present and future generations for adequate and affordable energy to ensure sustainable development.

2.2 Current and future energy supplies Current global energy final supplies are dominated by fossil fuels (388 EJ per year), with much smaller contributions from nuclear power (26 EJ) and hydropower (28 EJ). Biomass provides about 45 10 EJ, which makes it by far the most important renewable energy source used. Renewable energy supplies 18% of the world’s final energy consumption, including traditional biomass, large hydropower, and renewable (small hydro, modern biomass, wind, solar, geothermal, and biofuel energy) (Figure 2.2). Traditional biomass, primarily for cooking and heating, represents about 13% and is growing slowly or even declining in some regions as biomass is used more efficiently or is replaced by more modern forms of energy.

Figure 2.2 World primary energy consumption, 1950e2050. (From: World energy 2018e50: World energy annual report (part 1), minqi li, ron patterson, 2018).

30

N. El Bassam

2.3 Peak oil The peak oil theory states that any finite resource (including oil) will have a beginning, middle, and end of production, and at some point it will reach a level of maximum output. Today, we consume around four times as much oil as we discover. We used to think about peak oil like this: the reserves are finite, we know where they are and how long they will last, and we will start running out soon. However, with technological innovations, we keep finding new oil deposits that are recoverable and a peak may not happen for a decade or more (Association for the Study of Peak Oil & Gas) (Figures 2.3 and 2.4). Peak oil is the hypothetical point in time when the global production of oil reaches its maximum rate, after which production will gradually decline.

Figure 2.3 Renewable energy today in a global context. PV, photovoltaic. (From REN 21, GSR, (2019)).

Restructuring future energy generation and supply

31

Figure 2.4 Peak oil is the hypothetical point of reserves. pngl, PNG Liquefied Natural Gas. (From https://warosu.org/biz/image/8AhlPYG2dLG3xe0-gkEBgg, 2020).

2.4 Availability of alternative resources Energy cannot be created; it can be converted from one form to other by technical, biological, and chemical means, such as solar and wind energy into heat and power energy, biomass into heat, electricity, or biofuels, and so on. The good news is that we have all that we need of energy resources as well as conversion tech nologies to ensure a complete supply of clean and green energy. All renewable energy sources provide 3078 times current global energy needs (Figure 2.5). Renewable energy offers our planet a chance to reduce carbon emissions, clean the air, and put our civilization on a more sustainable footing. Renewable sources of energy are an essential part of an overall strategy of sustainable development. They help reduce dependence on energy imports, ensuring a sustainable supply and climate protection. Furthermore, renewable energy sources can help improve the competitiveness of industries over the long run and have a positive impact on regional development and employment. Renewable energies will provide a more diversified, balanced, and stable pool of energy sources. With rapid growth in Brazil, China, and India, and continued growth in the rest of the world, it is no longer a question of when we will incorporate various renewable energy sources more aggressively into the mix, but how fast.

32

N. El Bassam

Figure 2.5 Physical potential energy resources of the world. All renewable energy sources provide 3078 times current global energy needs. (From https://pakpas.org/ SOLAR.RENEWABLE%20ENERGY/2.AMS%20GREEN%20PV/9-energy-resources-world.png, (2020).

Technically exploitable amounts of energy in the form of electricity, heat, and chemical energy from renewable sources exceed the current world energy consumption by about sixfold (Nitsch, 2007). • Improving cooperation among nations as well as between public and private sectors. Renewable energy and energy efficiency do not represent an alternative to fossil resources; they are the only options that can ensure sustainable development and the survival of humanity and protect our climate. The share of renewable energy in the total energy supply needs to grow by 2% per year to ensure a future energy demand and avoid regional and global crises. Although more than 100 years’ supply of crude oil is left in the ground, for the most part, resources that are cheap and easy to extract have already been discovered and adequate proportions of gas and oil resources should be reserved for human welfare in decades to come.

Restructuring future energy generation and supply

33

2.5 Outlook The BP Energy Outlook considers different aspects of the energy transition and the key issues and uncertainties these raise. In all of the scenarios considered, world gross domestic production more than doubles by 2040, driven by increasing prosperity in fast-growing developing economies. In the evolving transition scenario, this improvement in living standards causes energy demand to increase by around a third over the Outlook, driven by India, China, and other countries in Asia, which together account for two-thirds of the increase. Despite this increase in energy demand, around two-thirds of the world’s population in 2040 still live in countries where average energy consumption per head is relatively low, which highlights the need for more energy. Energy consumed within industry and buildings accounts for around three-quarters of the increase in energy demand (Figure 2.6). Growth in transport demand slows sharply relative to the past as gains in vehicle efficiency accelerate. The share of passenger vehicle kilometers powered by electricity will have increased to around 25% by 2040.

Figure 2.6 (A) One hundred percent renewable energy targets in cities as of mid-2019. (From REN21 Status Report, 2020).

34

N. El Bassam

Figure 2.6 (B) Renewable power in cities, by number of cities and renewable share 2017. (from REN21, renewables in cities 2019 global status report 2020).

This is supported by the growing importance of fully autonomous cars and shared-mobility services (Figure 2.7). Content in this section is taken from BP Energy Outlook 2019, published in Feb. 2019. The Energy Outlook explores forces shaping the global energy transition out to 2040 and key uncertainties surrounding that transition (Figure 2.8). It shows how rising prosperity drives an increase in global energy demand and how that demand will be met over the coming decades through a diverse range of supplies including oil,

Figure 2.7 Electric car global stock (Renewables in cities REN21-global status report, 2020).

Restructuring future energy generation and supply

35

Figure 2.8 Primary energy consumption billion tons of oil equivalents (toe) by region. OECD, Organization for Economic Cooperation and Development. (BP, 2020; https:// www.bp.com/en/global/corporate/energy-economics/energy-outlook/demand-by-region. html).

gas, coal, and renewables (https://www.bp.com/en/global/corporate/ energy-economics/energy-outlook.html). A transition is under way in the global pattern of demand, with the dominance of the developing world increasing (Figures 2.9 and 2.10).

Figure 2.9 Primary energy growth, percentage per annum and regional contributions. OECD, Organization for Economic Cooperation and Development. (BP, 2020; https:// www.bp.com/en/global/corporate/energy-economics/energy-outlook/demand-by-region. html).

36

N. El Bassam

Figure 2.10 Primary energy consumption billion tons of oil equivalents (toe) by fuel. (BP, 2020; https://www.bp.com/en/global/corporate/energy-economics/energy-outlook/ demand-by-region.html).

The transition into distributed and decentralized renewable energy systems has to be associated with multiple measures: • Renewable energies should remain the priority (solar, wind, hydro, biomass, and geothermal) • Improving energy efficiency • Construction of smart grids • Creating power storage facilities • Future-oriented and innovative policy within national, regional, and global contexts • Creating a global climate framework • Intensifying research and education activities

References Nitsch, F., 2007. BMU Documentation. https://warosu.org/biz/image/8AhlPYG2dLG3xe0-gkEBgg 2020.

Restructuring future energy generation and supply

37

Further reading BSU Solar, 2007. German Solar Industry Association. Peak Oil news & message boards, 2011. Exploring Hydrocarbon Depletion [online] available at: http://peakoil.com/what-is-peak-oil/. REN21 Renewables, 2007. reportGlobal Status Report, 2008 Deutsche Gesellschaft for Technische Zusammenarbeit (GTZ) GmbH. www.ren21.net. https://pakpas.org/SOLAR.RENEWABLE%20ENERGY/2.AMS%20GREEN%20PV/ 9-energy-resources-world.png, 2020. https://www.bp.com/en/global/corporate/energy-economics/energy-outlook/demandby-region.html. https://www.ren21.net/wp-content/uploads/2019/05/gsr_2019_full_report_en.pdf.

CHAPTER THREE

Road map of distributed renewable energy communities N. El Bassam International Research Centre for Renewable Energy, IFFED.org, Germany

Contents 3.1 3.2 3.3 3.4

Energy and sustainable development Community involvement Facing the challenges The concept of the food and agriculture organization, an integrated energy community 3.5 Global approach 3.5.1 Basic elements of energy demand 3.5.1.1 3.5.1.2 3.5.1.3 3.5.1.4 3.5.1.5 3.5.1.6 3.5.1.7 3.5.1.8 3.5.1.9 3.5.1.10

Heat Electric power Water Lighting Cooking Health and sanitation Communications Mobility Agriculture Maintenance workshops and small marketsindustries

3.6 Basic and extended needs 3.6.1 Typical electricity demands 3.6.2 Single- and multiple-phase island grid

45 48 48 48 48 49 49 50 50 50 50 50

51 51 52

3.6.2.1 Version 1: single-phase island grid 3.6.2.2 Version 2: three-phase island grid 3.6.2.3 Version 3: three-phase island grid and parallel operation of the Sunny Island inverter

3.6.3 System solution for island grids 3.7 Representative energy plant species for different climate regions 3.7.1 Temperate climate 3.7.2 Representative energy plant species for different climate regions (arid and semiarid climate) 3.7.3 Representative energy plant species for different climate regions (tropical and subtropical climate)

Distributed Renewable Energies for Off-Grid Communities, Volume 2 ISBN: 978-0-12-821605-7 https://doi.org/10.1016/B978-0-12-821605-7.00017-9

40 41 42 43

© 2021 Elsevier Inc. All rights reserved.

52 52 52

53 53 53 55 56

39

j

40

N. El Bassam

3.8 Regional implementation 3.9 Opportunities driven by energy sector coupling 3.9.1 Demand-side flexibility programs References Further reading

56 59 59 60 60

3.1 Energy and sustainable development The world continues to seek energy to satisfy its needs while giving due consideration to the social, environmental, economic, and security impacts. It is now clear that current approaches to energy are unsustainable. It is the responsibility of political institutions to ensure that technologies that enable sustainable development are transferred to the end users. Scientists and individuals bear the responsibility of understanding the earth as an integrated whole and must recognize the impact of our actions on the global environment, to ensure sustainability and avoid disorder in the natural life cycle. Wise policy in a regional and global context requires that demands be satisfied and risks be avoided (Figure 3.1). Current approaches to energy are unsustainable and nonrenewable. Furthermore, energy is directly related to the most critical social issues affecting sustainable development: poverty, jobs, income levels, access to social services, gender disparity, population growth, agricultural production,

Figure 3.1 Sustainability in regional and global context demands, risks, and measures. El Bassam revised (2020).

Road map of distributed renewable energy communities

41

climate change, environmental quality, and economic and security issues. Without adequate attention to the critical importance of energy to all of these aspects, the global social, economic, and environmental goals of sustainability cannot be achieved. Indeed, the magnitude of the change that is needed is immense; it is fundamentally and directly related to the energy produced and consumed nationally and internationally. The key challenges to realizing these targets are to overcome the lack of commitment and to develop the political will to protect people and the natural resource base. Failure to take action will lead to continuing degradation of natural resources, increasing conflicts over scarce resources and widening gaps between rich and poor. We must act while we still have choices. Implementing sustainable energy strategies is one of the most important levers humankind has for creating a sustainable world. More than two billion people, mostly living in rural areas, have no access to modern energy sources. Food and fodder availability is closely related to energy availability.

3.2 Community involvement There are different views regarding what scale of community involvement is advisable or even possible, and a range of models can be designed for distributed renewable energy projects with some form of community involvement. This book includes a range of case studies with different renewable energy types, geographies, ownership models, and types of community participation. For all case studies selected, however, communities are involved in some ongoing way with the project, even if local people were not the initiators of the renewable energy project. Communities also have a powerful role in rapidly reducing greenhouse gas emissions across the globe. Many communities are already leading the response to climate change by establishing renewable energy production in their town or city, at the same time strengthening relationships among community members, creating local jobs, and often generating income for the community. Community energy includes the social process of establishing and distributing renewable energy technology locally, with social and economic benefits to that defined community. Community energy is therefore about the social arrangements regarding how an important technology that contributes to the sustainability solution is implemented and brings benefit to people. The localized distribution of energy has considerable advantages for efficiency and sustainability. However, how communities are involved in

42

N. El Bassam

the initiation, development, and consumption of locally produced energy is also important. If energy is produced on a local scale but does not involve or benefit local people, arguably the full sustainability benefits of community energy will not be achieved. In terms of achieved renewable energy outcomes, research also suggests that there is a lesser acceptance of renewable energy projects if local people are not involved and benefits are not shared among community members. Most energy generation is centralized, involving the production of electricity at a large, central facility, the transmission of high-voltage electricity over long distances through the power grid, and the conversion and distribution of that energy to a large number of consumers. Energy may travel many hundreds of kilometers from where it was produced before it reaches the user of the energy. Most centralized production facilities use nonrenewable sources such as coal, oil, or nuclear material to power their electrical generation. Significant waste occurs during the transmission of high-voltage electricity over large distances: losses through the grid can amount to 7e25% of generation electricity (IEA, 2016). As energy markets are restructured, customers and utilities feel more pressure to control costs and increase operating flexibility. Contributing to this trend is a heightened concern about energy security and the emergence and advanced development of modular renewable generation technologies. The environmental benefits of these distributed power sources exploiting, for example, renewable resources or combined heat and power are substantial. Renewable energy communities involve groups of citizens, social entrepreneurs, public authorities, and community organizations participating directly in the energy transition by jointly investing in, producing, selling, and distributing renewable energy. Beyond the reduction of greenhouse gas emissions, there are many benefits for the communities involved, including economic development, the creation of new jobs, cheaper energy, self-sufficiency, community cohesion, and energy security. Regional authorities can support the emergence of energy communities by providing financing, expertise, and advice, and ensuring that regulatory issues can be easily understood and navigated (Figure 3.2).

3.3 Facing the challenges Some appealing concepts for fostering rural development imply the systematic exploitation of locally available renewable sources of energy.

Road map of distributed renewable energy communities

43

Figure 3.2 Basic elements of optimizing main grids and microgrids and maintaining sustainability. PV, photovoltaics. From Lowitzscha et al., (2019).

Within this perspective, integrated energy farms, or integrated energy settlements, can be thought of as sustainable power centers, supplying local markets comfortably with electricity and fuels while covering their demand for food crops and soft commodities. Different ecological and sociocultural settings require specific types of design for locally adapted integrated energy communities (IECs). To meet challenges, future energy policies should more strongly emphasize developing the potential of energy sources, which should form the foundation of a future global energy structure. Within this context, the Food and Agriculture Organization (FAO) of the United Nations and the Sustainable Rural Energy Network (SREN) have developed the concept of the optimization, evaluation, and implementation of integrated renewable energy farms (IREFs) in rural communities (El Bassam, 1999, updated 2020).

3.4 The concept of the food and agriculture organization, an integrated energy community Renewable energy has the potential to bring power to communities, not only in the literal sense, but by transforming their development prospects. There is tremendous latent demand for small-scale, low-cost, off-the-grid solutions to people’s varying energy requirements.

44

N. El Bassam

People in developing countries understand this only too well. If they were offered new options that would truly meet their needs and engage them in identifying and planning their own provision, success in providing renewable energy services would become a reality. Energy is one of the important inputs to empower people, provided it is made available to people in unserved areas on an equitable basis. Therefore, access to energy should be treated as a fundamental right for everyone. This is possible only if the end users are made the primary stakeholders in the production, operation, and management of the generation of useful energy. Despite much well-intended effort, little progress has been made and a radical new approach is called for, based on the following imperatives: • Rural development in general, and rural energy development specifically, need to be given higher priority by policy-makers. • Rural energy development must be decentralized and local resources managed by rural people. • Rural energy development must be integrated with other aspects of rural development, overcoming the institutional barriers between agriculture infrastructure and education as well as in the social and political spheres. • Because of the problem of noncommitting policies on the government level, different priorities in each country should be addressed. The productivity and health of a third of humanity are diminished by a reliance on traditional fuels and technologies; women and children suffering the most (El Bassam, 2004). Current methods of energy production, distribution, and use worldwide are major contributors to environmental problems, including global warming and degradation of the ecosystem at the local, regional, and global levels. The IEC concept includes farms or decentralized living areas from which the daily necessities (water, food, and energy) can be produced directly on-site with minimal external energy inputs. Energy production and consumption at the IEC has to be environmentally friendly, sustainable, and ultimately based mainly on renewable energy sources. It includes a combination of different possibilities for nonpolluting energy production, such as modern wind and solar electricity production, as well as the production of energy from biomass. It should seek to optimize energetic autonomy and an ecologically semiclosed system, while providing socioeconomic viability and giving due consideration to the newest concepts of landscape and biodiversity management.

Road map of distributed renewable energy communities

45

3.5 Global approach The concept of an IEC or settlement includes four pathways: Economic and social Energy Food Environmental Specifications for an IEC include: • Decentralized, autonomous, and location-based production of energy, food, and innovation • Combined use of different renewable energy sources (biomass, solar, wind, etc.) • Suitability for remote areas: settlements, villages, and islands • Job creation and social, economic, ecological, education, and training opportunities for people in the community Basic data should be available for the verification of an (IREF). Various climatic constraints, water availability, soil conditions, infrastructure, availability of skills and technology, population structure, flora and fauna, common agricultural practices, and economic and administrative facilities in the region should be considered. Renewable energy technologies are available from different natural resources: biomass, geothermal, hydropower, ocean power, solar (photovoltaic [PV] and solar thermal), wind, and hydrogen. Climatic conditions prevailing in a particular region are the major determinants of agricultural production. In addition, other factors such as local and regional needs, availability of resources, and other infrastructure facilities determine the size and product spectrum of the farmland. The same requisites also apply to an IREF. The climate fundamentally determines the selection of plant species and their cultivation intensity for energy production on the farm. Moreover, the climate influences the production of the energy mix (consisting of biomass, wind, and solar energies) essentially at a given location and the type of technology that can be installed also depends decisively on the climatic conditions of the locality in question. For example, cultivation of biomass for power generation is not advisable in arid areas. Instead a larger share can be allocated to solar energy techniques in such areas. Likewise, coastal regions are ideal for wind power installations (Figure 3.3). 1. 2. 3. 4.

46

N. El Bassam

Energy Forest Cereals

Vegetables

Energy Crops Sugar Crops Poultry Farming Sheepkeeping

Fruits

Starch Crops

Energy Crops Energy Apiculture (perennial) Crops Spice and (annual) med. Herbs

Energy Forest

Grazing and Fodder Area Floriculture

Oil Crops Energy Forest

Crops Energy

Figure 3.3 Model for an integrated energy community with farming systems. med., medical. From El Bassam (2001) and (2020).

Taking these circumstances into account, a scenario was made for an energy farm of 100 hectares (ha) (about 247 acres) in the different climatic regions of Northern and Central Europe, Southern Europe, Northern Africa, and the Sahara and Equatorial regions. It was presumed that one unit of this size needs about 200 megawatt hours (MWh) heat and 100 MWh power per annum for its successful operation. A need for fuel of approximately 8000 L per annum was calculated. The possible shares of different renewable energies are presented in Table 3.1. It is evident that throughout Europe, wind and biomass energies contribute the major share to the energy mix, whereas in North Africa and the Sahara, the main emphasis is on solar and wind energies. Equatorial regions offer great possibilities for solar as well as biomass energies; little share is expected from the wind source of energy in these regions. Under these assumptions, in Southern Europe, the Equatorial regions, and North and Central Europe, a farm area of 4.8, 10, and 12 ha (12, 24.7, and 29.6 acres),

Northern and central Europe Southern Europe

North Africa Sahara Equatorial region

Solar Wind Biomass Solar Wind Biomass Solar Wind Biomass Solar Wind Biomass

7 80 60 100 80 40 100 100 20 100 45 70

15

60

12

100 100

36

4.8

65 100

14

1.2

25 100

45

10

Road map of distributed renewable energy communities

Table 3.1 Possible share of different renewable energy sources in diverse climatic zones produced on an energy farm of 100 ha (247 acres). Power production Heat production Biomass area Climate region Energy source (% of total need) (% of total need) Biomass need (t/a) (% of total area)

100

From El Bassam (1998a).

47

48

N. El Bassam

respectively, would be needed to cultivate biomass for energy purposes. This would correspond to an annual production of 36, 45, and 60 tons for the respective regions. In North Africa and the Sahara regions, in addition to wind and solar energy, 14 tons of biomass from 1.2% of the total area would be necessary for energy provisions.

3.5.1 Basic elements of energy demand The IEC concept includes a decentralized living area from which the daily necessities (food and energy) can be produced directly on-site with minimal external energy inputs. The land of an IEC may be divided into compartments for use in growing food crops, fruit trees, annual and perennial energy crops, and short-rotation forests, along with wind and solar energy units within the farm. An IEC system based largely on renewable energy sources would seek to optimize energetic autonomy and an ecologically semiclosed system, while providing socioeconomic viability and giving due consideration to the newest concepts of landscape and biodiversity management. Ideally, it would promote the integration of different renewable energies as well as rural development and contribute to the reduction of greenhouse gas emissions. Energy supply in rural communities has to meet the needs of the people and ensure economic and social development. To generate adequate energy, it is necessary to determine the most appropriate and affordable technologies, equipment, and facilities. (Figure 3.4). 3.5.1.1 Heat Heat can be generated from biomass or solar thermal sources to create both high-temperature steam and low-temperature heat for space heating, domestic and industrial hot water, pool heating, desalination, cooking, and crop drying. 3.5.1.2 Electric power Electric power can be generated by solar PV, solar thermal, biomass, wind, hydro, microhydro, and geothermal, sources. 3.5.1.3 Water Water is an essential resource for which there can be no substitute; it is needed for drinking and irrigation. Renewable energy can have a major role in supplying water in remote areas. Several systems could be adopted for this purpose:

Road map of distributed renewable energy communities

49

Figure 3.4 Basic elements and needs of Integrated Renewable Energy Community. med., medicine. (El Bassam, 2000).

• Solar distillation • Renewable energy operated desalination units • Solar, wind, and biomass operated water pumping and distribution systems 3.5.1.4 Lighting To improve living standards and encourage the spread of education in rural areas, a supply of electricity is vital. Several systems could be adopted to generate electricity for this purpose: • Solar systems (PV and solar thermal) • Wind energy systems • Biomass and biogas systems (engines, fuel cells, and Stirling) 3.5.1.5 Cooking Women in rural communities spend long hours collecting firewood and preparing food. Other methods are more efficient, healthy, and environmentally benign: • Solar cookers and ovens • Biogas cooking systems • Improved biomass stoves using briquettes and pellets • Plant oil and ethanol cookers

50

N. El Bassam

3.5.1.6 Health and sanitation To improve serious health problems among villagers, solar energy from PV, wind, and biogas could be used to operate: • Refrigerators for vaccine and medicine storage • Sterilizers for clinical items • Wastewater treatment units • Ice making 3.5.1.7 Communications Communication systems are essential for rural development. The availability of these systems has a great impact on people’s lives and can advance their developmental process more rapidly. Electricity can be generated from any renewable energy source to operate basic communication needs such as radio, television, weather information systems, and mobile telephone. 3.5.1.8 Mobility Improved transportation in rural areas and villages has a positive effect on the economic situation as well as social relations among people in these areas. Several methods could be adapted for this purpose: • Solar electric vehicles • Vehicles fueled by biodiesel, ethanol, plant oil, biogas, and hydrogen (alternative engines or fuel cells) • Electric cars • Traction animals 3.5.1.9 Agriculture In rural areas, agriculture represents a major energy end use. Mechanization using renewable sources of energy can reduce time spent in labor-intensive processes, freeing time for other income-producing activities. Renewable sources of energy can be applied to: • Soil preparation and harvesting • Husking and milling of grain • Crop drying and preservation • Textile processing 3.5.1.10 Maintenance workshops and small marketsindustries Adequate levels of power and heat should be available for processing and manufacturing.

Road map of distributed renewable energy communities

51

3.6 Basic and extended needs Power supply systems from renewable sources in off-grid operation should be robust, inexpensive, and reliable. Most important, they need to have a modular structure so they can be extended later. PV power supply systems in off-grid operation supply power to small consumers (3e30 kW) far from the public utility grid. An essential component of a modular supply system is a battery inverter, such as the Sunny Island, with a nominal power of 3.3 kW each. The advanced Sunny Island battery inverter is the grid master and the central component of a modular supply system; it enables small-scale island utilities in remote areas. An island grid is easy to plan and install and allows a flexible operation. The device includes an intelligent control able to supply different consumers and feed power from different generators. Such generators are, for example, PV string inverters for grid supply, small wind energy plants, or diesel units, which make the battery inverter useful for any source. The system management handles battery control, enables limited load management, and provides communication interfaces for optional system management units. The required operating modes and parallel switching of current converters can be realized. A battery inverter and a lead acid battery can establish a simple single-phase island grid.

3.6.1 Typical electricity demands To be able to size the power supply properly, the peak power, daily energy consumption, and annual energy growth should be estimated. This is necessary to size both the generating plant and the conductor used in the distribution system. The sustainability of the system will not be guaranteed if the capacity of the system is too small; it leads to consumer dissatisfaction. Alternatively, too much capacity would mean additional investment costs and possibly too high tariffs. It is frequently difficult to make load projections that reflect reality, especially for prospective consumers who have little experience with electrification. The more reliable approach to assessing demand is to survey households in adjoining, already electrified areas or in a region with similar economic activities, demographic characteristics, and so forth. This would assess the average initial loads per household in these areas as well as their historical load growth. The already electrified regions to be surveyed preferably should have a type of service similar to that being proposed in the new community, such as 24-h power or electricity for 4 h each evening.

52

N. El Bassam

Any projections of load and load growth in an area to be electrified using information gathered from already electrified regions should also consider factors such as the difference in the level of disposable income in the two areas, the presence of raw materials or industry, the potential for tourism, and access to outside markets for goods that might be grown or produced locally. For people who are accustomed to using electricity, projections give an overview of the daily electricity demand for these three classes: basic, extended, and normal needs. Because of the high cost of generating electricity, it is important to choose the most efficient appliances. The basis for all assumptions in the table is the use of such appliances. Otherwise, the energy consumption will increase significantly. The peak power is typically three to four times higher than the average power.

3.6.2 Single- and multiple-phase island grid In power generation, PV, wind, and hydroelectronic power plants can be combined. Normally an additional electric generator such as a diesel generator can make the supply more reliable. The controls allow for a power increase by switching to three battery inverters in parallel on one phase. In this example, every third household has a refrigerator and every village consists of 50 households. 3.6.2.1 Version 1: single-phase island grid This is a combination of power generation with parallel operation of inverters. 3.6.2.2 Version 2: three-phase island grid If there is a need to connect three-phase consumers, the design of the island is flexible and extendable. The smallest three-phase system has a nominal output power of 10 kW and consists of three Sunny Island inverters. Three-phase systems also simplify the connection of larger diesel sets and wind energy systems. These are mostly equipped with three-phase generators. 3.6.2.3 Version 3: three-phase island grid and parallel operation of the Sunny Island inverter Several Sunny Islands can be combined to establish a three-phase system up to 30 kW.

Road map of distributed renewable energy communities

53

3.6.3 System solution for island grids • Simple design of island grids resulting from the connection of all components on the AC side • Reliable and safe power supply with utility quality in remote areas • Easy integration of PV plants and wind energy or diesel generators • Power supply for single houses or even small villages • Extendable design (one- or three-phase combinations in parallel operation) • Optimal battery life Moving ahead, to broaden the scope and seek the practical feasibility of such farms, the dependence of local inhabitants (end users) is to be integrated into this system. Roughly 500 persons (125 households) can be integrated into one farm unit. They have to be provided with food as well as energy. As a consequence, the estimated extra requirement of 1900 MWh of heat and 600 MWh of power have to be supplied from alternative sources. Under the assumption that the share of wind and solar energy in the complete energy provision remains at the same level, the production of 450 tons of dry biomass is needed to fulfill the demands of such farm units. To produce this quantity of biomass, 20% of farm area needs to be dedicated to cultivation. In Southern Europe and the Equator, 15% of the land area should be made available for the provision of additional biomass. More than 200 plant species have been identified in different regions of the world to serve as sources for biofuels. A summary of energy plant species that can be grown under various climatic conditions is documented in the following sections (Figures 3.5e3.7).

3.7 Representative energy plant species for different climate regions 3.7.1 Temperate climate These data are from El Bassam (1996), (1998b)., updated (2020): Cordgrass (Spartina spp.) Reed Canary grass (Phalaris arundinacea.) Fibre sorghum (Sorghum bicolor) Rosin weed (Silphium perfoliatum) Giant knotweed (Polygonum sachalinense) Safflower (Carthamus tinctorius) Hemp (Cannabis sativa)

54

N. El Bassam

Generator PV-modules Consumers Sunny Boy

230 v 50/60 Hz Sunny Island 3.3 kW

3.3 kW

Battery 60 V

Figure 3.5 Integration of photovoltaic (PV) plants and diesel set in parallel operation of Sunny Island inverters. From El Bassam and Maegaard (2004).

Wind turbine Generator Set PV-modules

Consumers Sunny Boy

3a/400 V 50/60 Hz Sunny Island 9.9 kW Battery 60 V

Figure 3.6 Integration of photovoltaic (PV) plants and diesel set in parallel operation of Sunny Island inverters. From El Bassam and Maegaard (2004).

55

Road map of distributed renewable energy communities

Hydro Biodiesel Diesel

Wind turbine

Generator Set PV-modules

Consumers

Sunny Boy

3a/400 V 50/60 Hz Sunny Island

Battery 60 V 19.8 kW

Figure 3.7 Parallel operation of Sunny Island inverters suitable for high performance. PV, photovoltaic. From El Bassam and Maegaard (2004).

Soybean (Glycine max) Kenaf (Hibiscus cannabinus) Sugar beet (Beta vulgaris) Linseed (Linum usitatissimum) Sunflower (Helianthus annuus) Miscanthus (Miscanthus x giganteus) Switchgrass (Panicum virgatum) Poplar (Populus spp.) Topinambur (Helianthus tuberosus) Rape (Brassica napus) Willow (Salix spp.)

3.7.2 Representative energy plant species for different climate regions (arid and semiarid climate) Argan tree (Argania spinosa) Olive (Olea europaea) Broom (Genisteae) (Spartium junceum) Poplar (Populus spp.) Cardoon (Cynara cardunculus) Rape (Brassica napus)

56

N. El Bassam

Date palm (Phoenix dactylifera) Safflower (Carthamus tinctorius) Eucalyptus (Eucalyptus spp.) Salicornia (Salicornia bigelovii) Giant reed (Arundo donax) Sesbania (Sesbania spp.) Groundnut (Arachis hypogaea) Soybean (Glycine max) Jojoba (Simmondsia chinensis) Sweet sorghum (Sorghum bicolor)

3.7.3 Representative energy plant species for different climate regions (tropical and subtropical climate) Aleman grass (Echinochloa polystachya) Jatropha (Jatropha curcas) Babassu palm (Orbignya oleifera) Jute (Corchorus spp.) Bamboo (Bambusa spp.) Leucaena (Leucaena leucocephala) Banana (Musa x paradisiaca) Neem tree (Azadirachta indica) Black locust (Robinia pseudoacacia) Oil palm (Elaeis guineensis) Brown beetle grass (Leptochloa fusca) Papaya (Carica papaya) Cassava (Manihot esculenta) Rubber tree (Acacia senegal) Castor oil plant (Ricinus communis) Sisal (Agave sisalana) Coconut palm (Cocos nucifera) Sorghum (Sorghum bicolor) Eucalyptus (Eucalyptus spp.) Soybean (Glycine max) Sugar cane (Saccharum officinarum)

3.8 Regional implementation The International Research Center for Renewable Energy in Germany was contracted by the FAO in 2000 to accomplish planning for

57

Road map of distributed renewable energy communities

the implementation of an IREF in a practical sense at a regional level, considering the climatic and soil conditions. The planning work was started in Dedelstorf (northern Germany). An area of 280 ha (691.6 acres) was earmarked for this farm, which would satisfy the food and energy demands of the 700 participants in the project. For settlement purposes, old military buildings were renovated. It was expected to take 3 years to complete the project. The main elements of heat and power generation would be solar generators and collectors, a wind generator, a biomass combined heat and power generator, a Stirling motor, and a biogas plant. The total energy to be provided was calculated to be as much as 8000 MWh heat and 2000 MWh power energy. The cultivation of food and energy crops would be according to ecological guidelines. The energy plant species foreseen were short-rotation coppice, willow and poplar, miscanthus, polygonum, sweet and fiber sorghum, switchgrass, and reed canary grass and bamboo. Adequate food and fodder crops as well as animal husbandry would be implemented according to the needs of the people and specific environmental conditions of the site. A research, training, and demonstration center would accompany this project (Figure 3.8).

Biogas Plant and Gas Storage WIND

G Biogas

100 KW SUN

G

Solar Cells 21 KWp 200 m 2

Stirling Motor 40 KW

300 m 2 Solar Collectors G

Combined Hear and Power Station (2,5 MW) BIOMASS

Farm & Village

Control panels Heat

Meters

Power Heat Accumulator 2 X 10 m 3

G

Generator

Figure 3.8 Technologies for heat and power production in integrated renewable energy community. From El Bassam and Maegaard (2004).

58

N. El Bassam

In northern Germany, a site near Hanover was identified for the implementation of an IREF (90% biomass, 7% wind, and 3% solar) and as a research center for renewable energy from solar, wind, and biomass as well as their configuration. Special emphasis was on optimizing energetic autonomy in decentralized living areas and promoting regional resource management. The research center (www.ifeed.org) undertakes responsibility in the fields of research, education, transfer of technology, and cooperation with national and international organizations. It also offers trade and industry the opportunity to introduce, demonstrate, and commercialize their products. Cooperation with developing countries on issues of sustainable energy and food production is also one a prime objective of the research center. Providing access to clean, reliable, and affordable energy is important to promoting sustainable economic development. The rapidly falling costs and growing maturity of green energy technologies have substantially reduced the dependence of remote communities on imported hydrocarbons. Yet, socioeconomic barriers have slowed the development and rollout, especially in low-income communities. Smart integrated renewable energy systems could overcome some of these barriers. These systems, which could logistically deliver sustainable energy services to all sections of communities, rely on three pillars: energy sector coupling, demand-side management, and peer-to-peer transactions of energy (Figure 3.9).

Figure 3.9 Schematic showing conceptual design of a typical smart integrated renewable energy system. From IEA (2019).

Road map of distributed renewable energy communities

59

3.9 Opportunities driven by energy sector coupling Energy sector coupling is increasing interconnections among different energy carriers (electricity, heating and cooling, and synthetic fuels such as hydrogen, ammonia, and ethanol) to exploit their synergy. This process helps initiate sustainability in different energy-use sectors and improve efficiency and adequacy. Although there are some technoeconomic challenges regarding energy sector coupling, policy and regulatory requirements pose more serious challenges. Setting up a policy and regulatory framework within utility business operations is much less challenging for first-class energy systems than for independent sector services in metropolitan areas. Therefore, the value chain, from technology to markets, can be properly analyzed and appropriate decisions can be made regarding market designs and procedures of flexible provisions.

3.9.1 Demand-side flexibility programs Through the deployment of demand response, financial viability can also be improved. This process refers to schemes developed by utilities to balance peaks and troughs of electricity demand. Different levels of service reliability can ultimately be tolerated by different strata of society. In addition, the incorporation of demand-side flexibility resources in long-term planning decisions lays the groundwork for establishing new design standards for off-grid renewable energy systems. The new design standards entail varied reliability margins for varying populations and can serve as a buffer against uncertain investments under specific economic conditions of communities. This process of habitual electricity consumption patterns can be transformed into 100%-renewable energy systems. This could pave the way for interactions among different energy carriers. Accordingly, prioritization weights can be personalized by the end users of different energy services subscribed to an integrated energy system. Smart integrated renewable energy systems as a community resource. In addition to encouraging end-customers to be active in supporting grid services, advanced information and communications technologies provide an opportunity to establish peer-to-peer (local) energy marketplaces. A study of local energy exchanges on the Latrobe Valley Microgrid at LO3 Energy in Victoria, Australia reported that the implementation of local energy markets has resulted in 6e12% savings on electricity bills, while

60

N. El Bassam

increasing the revenue of prosumers by 18e37%. This difference of efficiency is between feed-in tariffs and retail rates, which generates financial benefits shared among both parties. Overall, peer-to-peer energy trading is valuable within multienergy systems. It allows charitable prosumers to donate or sell subsidized energy to low-income customers. From the utility’s point of view, such local energy markets may be a useful risk-management tool, reducing the long-term financial risk of investing in renewable energy projects. The energy trilemma, a term that sums up the difficulty in optimizing the trade-off among the reliability, affordability, and sustainability of energy systems, has been recognized for decades. However, the advent of smart integrated renewable energy systems seems to offer promising prospects for a way out of this trilemma. Originally published on renewableenergyworld.com e Written by Soheil Mohseni, Alan Brent and Scott Kelly https://www.power engineeringint.com/renewables/how-smart-integrated-renewables-systemscan-drive-sustainable-economic-development-in-remote-communities/ 2020.

References El Bassam, N., 1996 and 2020. Renewable Energy: Potential Energy Crops for Europe and the Mediterranean Region, REU Technical Series 46. Food and Agriculture Organization of the United Nations, Rome (FAO), 200 S. El Bassam, N., 1998. Biological life support systems under controlled environments. In: Bassam, N El, et al. (Eds.), Sustainable Agriculture for Food, Energy and Industry, vol. 2. James & James Science Publishers, London. El Bassam, N., 1998. Energy Plant Species: Their Use and Impact on Environment and Development. James & James Science Publishers, London. El Bassam, N., 1999. Integrated Energy Farm Feasibility Study. and 2019. SREN-FAO. El Bassam, N., 2001. Renewable energy for rural communities. Renew. Energy 24, 401e408. El Bassam, N., 2004. Integrated renewable energy farms for sustainable development in rural communities. In: Biomass and Agriculture. OECD, Paris, pp. 262e276. www.oecd.org/ agr/env. El Bassam, N., Maegaard, P., 2004. Integrated Renewable Energy for Rural Communities: Planning Guidelines, Technologies and Applications. Elsevier Science, Amsterdam. IEA, 2019. CC BY-NC-ND. J. Lowitzscha, C.E. Hoickab and F.J. van Tuldera 2019.

Further reading Africa Energy Commission (AFREC), 2007. Business Times of Nigeria. February 10, 2008. Chiaramonti, D., Grimm, P., Cendagorta, M., El Bassam, N., 1998. Small energy farm scheme implementation for rescuing deserting land in small Mediterranean islands, coastal areas, having water and agricultural land constraints Feasibility study. In: Proceedings of the 10th International Conference. Biomass for Energy and Industry, WuVrzburg, pp. 1259e1262.

Road map of distributed renewable energy communities

61

Davidson, O.R., Sokona, Y., 2001. Energy and sustainable development: key issues for Africa. In: Wamukonya, N. (Ed.), Proceedings of the African High-Level Regional Meeting on Energy and Sustainable Development. UNEP, Roskilde, pp. 1e20. Lowitzscha, J., Hoickab, C.E., van Tuldera, F.J., 2019. Renewable energy communities under the European Clean Energy Package e governance model for the energy clusters of the future? Else. Renew. Sustain. Energy Rev. 122, 109489. April 2020. Mohseni, S., tand Scot Kelly, A.B., 2020. Originally Published on renewableenergyworld.Com. Written by. www.powerengineeringint.com/renewables/how-smartintegrated-renewables-systems-can-drive-sustainable-economic-development-in-remote-co mmunities/.

CHAPTER FOUR

Planning of integrated renewable communities N. El Bassam International Research Centre for Renewable Energy, IFFED.org, Germany

Contents 4.1 4.2 4.3 4.4

Introduction Scenario 1 Scenario 2 Case study I: implementation of integrated energy farm under climatic conditions of central Europe 4.4.1 Specifications 4.4.2 Distribution of farm area 4.4.3 Farm production 4.4.4 Energy requirement 4.4.4.1 4.4.4.2 4.4.4.3 4.4.4.4 4.4.4.5 4.4.4.6

Administration and household Agricultural activities Site energy production Origin of biomass Contribution of different renewable energy sources Investment requirement

4.5 Case study II: arid and semiarid regions 4.5.1 Specifications 4.5.2 Farm production 4.5.3 Energy requirement 4.5.3.1 4.5.3.2 4.5.3.3 4.5.3.4 4.5.3.5 4.5.3.6

66 67 70 70 70 71 71 73 74 74

75 75 75 75

Administration and household Agricultural activities Energy production on the farm Origin of biomass Contribution of different renewable energy sources Investment requirement

Further reading

Distributed Renewable Energies for Off-Grid Communities, Volume 2 ISBN: 978-0-12-821605-7 https://doi.org/10.1016/B978-0-12-821605-7.00021-0

64 64 65 65

75 77 77 78 80 80

80

© 2021 Elsevier Inc. All rights reserved.

63

j

64

N. El Bassam

4.1 Introduction The process of planning an integrated energy system at the community level should include optimizing farm production in such a way that reserves (land, capital, and human effort) are freed from the field of food production and can be used in an economically feasible way in the process of production of energy and energy raw materials. A comparison of industrialized and developing countries illustrates that the preferences are different. In industrialized countries, agriculture is under pressure to reduce food production, whereas in developing countries, food production is of the highest importance. Therefore, in industrialized countries, the replaced production capacity should be used preferably to produce marketable energy and energy raw materials to create new sources of income for farmers. On the other hand, in developing countries, the additional production reserves of farms should serve only to secure autonomous food production energetically. The planning procedure has been classified into main scenarios.

4.2 Scenario 1 Necessary data such as climatic conditions and soil data, as well as the those concerning the identification of production factors, are normally available for an existing agricultural region. These data should be used as the basis for planning of integrated energy farms (IEFs) and systems. However, the energy demand of an agricultural area cannot be covered solely by the use of a circuit economy: that is, by the energetic use of agricultural waste and residues or through solar and wind energy. In some climatically unfavorable regions, in addition to available resources, we need the production of energy raw materials (biomass) on the farm. However, this may lead to a contentious situation concerning land use between farm products and energy raw materials. The production of the energy crop must not be introduced or expanded at the expense of other farm products such as foods. To plan integrated systems, the following aspects should first be considered: • Intensification and optimization of existing agricultural production, considering site conditions. This must increase the yields, productivity, and income (optimization model).

Planning of integrated renewable communities

65

• Choice of adapted production branches with a consideration of existing demands and an elimination of economically inefficient production branches. • Introduction of sustainable production models with the aim of minimizing input (capital and surface) in the case of unchangeable net profits. • Analysis of additional inputs (capital, land, and human effort) with regard to their use for additional food production or to produce energy raw material (preference model). • Integration of energy into agricultural production with regard to regional preferences, considering: • criteria of economical profitability • technical and economical feasibility • availability of technologies and integration possibilities of different technologies • surface, capital, and human effort requirements • socioeconomic and other regional particularities • Installation of an energy production and user management system

4.3 Scenario 2 An IEF could be created on the basis of the energy and food requirements of a specific number of persons. Assuming the farm area is available, the necessary site data such as climate, soil, and so on should first be determined. This can be recorded from national or regional statistics, or calculated or measured. In arid and semiarid regions, irrigation possibilities must be determined. Here, an analysis of the relevant irrigation system and its economic feasibility is important, taking into consideration regional sociological aspects (Figures. 4.1 and 4.2).

4.4 Case study I: implementation of integrated energy farm under climatic conditions of central Europe The main planning objective of implementing IEF on an existing agricultural farm settlement is that it should be autonomous in its energy supply. Specifications are as follows (Figure 4.3).

66

N. El Bassam

Existing agricultural farm

Step 1

Production optimization Model (POM)

Integrated Energy Farm

Integrated Energy Farm

with energy autonomous food production and low production level of energy for external use

with energy autonomous food production and high production level of energy for external use

Step 4 Available reserves of production factors (area, capital, man power)

Case I

Case II

food › Energy raw materials

food ‹ Energy raw materials

Step 3 Step 2 Analysis of the production factors

Preference Models

Figure 4.1 Planning steps integrated energy farm (Scenario 1).

4.4.1 Specifications Climate

Production

Climatic conditions Northern Germany Farm size: 100 ha (247 acres) Crops, vegetables, horticulture, energy plantation, animal husbandry, and vegetable and fruit production

67

Planning of integrated renewable communities

Figure 4.2 Implementation of integrated energy farm (Scenario 1).

4.4.2 Distribution of farm area Crops and root plants Oil plants Vegetables Fruit trees Grassland Building Total

60 ha (148.2 acres) 10 ha (24.7 acres) 5 ha (12.35 acres) 5 ha (12.35 acres) 17 ha (42 acres) 3 ha (7.41 acres) 100 ha (247 acres)

68

N. El Bassam

Creation of the farm Production area (Ftn) is not limited

estimation of initial conditions of the site estimation of energy potential

food requirement production area required: Ar = (Y ) / (F ) energy requirement agric. production: n Ep = Ep1 + Ep2 + 6 D 1 1 household: Eh (MWh/a) = {Eh1(MWh/a) + Eh2(MWh/a)} x n

number of energy users (n)

EDP

DATA ACQUISITION AND CALCULATION

INPUT

ENERGY

CULTURES

TECHNOLOGIES

PRODUCTION

OUTPUT ENERGY heat, power and fuel THROUGHPUT

Food Production

100% supply with basic food and energy

Target Group persons households

Figure 4.3 Implementation of integrated energy farm (Scenario 2).

Figure 4.4A graphically presents a general system for the power generation of an IEF. The infrastructure of the farm might consist of a residential building with an administration tract, office rooms (area of 400 m2), storage rooms, workshops, greenhouses, and stable facilities. Figure 4.4B shows the system control unit of the power generation system on an IEF.

69

Planning of integrated renewable communities

Figure 4.4 (A) General system complex of power generation of integrated energy generation. Biomass:

energy CHP connection (requirement point) CHP connection (requirement point)

connection of radiator dependent)

Emergency

Vegetable

connection change to public power supply (requirement point)

connection of additional power supply (requirement point)

disconnection of CHP (requirement point)

generator

supply

Figure 4.4 (B) Basic diagram of the system control unit of the power generation system on an integrated energy farm. intern., internal. (From Wolf, (1998); modified by the authors).

70

N. El Bassam

Table 4.1 Agricultural production per year (estimated). Production branch Products Area, ha

Quantity

Crop production

40.0

100 t

10.0 20.0 5.0 5.0

15 t 3000 t 100 t 200 t 75.000 units 2.0 t ¼ life weight

Vegetables Fruit trees Animal Husbandry

Cereals (wheat and barley) Sunflower Potatoes Different products Apples, pears, etc. Hens (eggs) Sheep breeding1

2.0

1

1.0 lamb/y, dt ¼ 0.1 ton.

4.4.3 Farm production Activities of the created IEF consist of crop production (70 ha; 172.9 acres), an energy plantation (15 ha ;37.1 acres), grassland/fodder (2 ha; 4.94 acres), and horticulture: • Vegetables (5 ha; 12.35 acres) • Fruit trees (5 ha; 12.35 acres); as well as Animal husbandry: • Sheep breeding (n ¼ 100) • Chicken (n ¼ 500). Table 4.1 presents production data for the different agricultural activities.

4.4.4 Energy requirement 4.4.4.1 Administration and household The area for the living and administration buildings is projected to be approximately 400 m2. The determination of the heat energy requirement is based on the following criteria: • The farm buildings should be well-insulated so that relatively low energy for heating will be required, corresponding to a 60e70 W/m2 reference area. • Total hours needed for heating will be approximately 1800 h/y (10 h/ d  6 months/y  30 days). Heat requirement kWh ¼ y used area m2  70W ¼ m2  1; 800 h 1 1 400  70  1800  10 50: 4MWh ¼ y 4 4 The hot water requirement is calculated to be 40 L/d per person. The annual demand for energy for hot water from 20 to 60 C is calculated as: 1 40  60  10  4: 19  365  0: 000278 850: 3kWh ¼ y ¼ person 4

71

Planning of integrated renewable communities

Table 4.2 Power requirement in MWh/y. Fields Connection value (kW)

Basic load Household Cooking Administration Workshop Total

2.5 2.5 8.2 1.5 12.1 26.8

Full load h

Consumption (MWh/y)

8760 1100 730 1300 450

21.9 2.75 5.99 1.95 5.45 38.04

For the entire building, a heating load of approximately 28.0 kW should be available; thus, the entire annual energy demand will be 50.4 MWh/y. Table 4.2 illustrates the electricity needed for various activities in the administration and household (six persons). A basic load of 2.5 kW for the entire farm is projected for permanent (24 h/d) power generation for circulating pumps, ventilation, engines, a refrigerator, emergency lighting, and so on. Total electricity consumption amounts to 38.1 MWh/y; this corresponds to a consumption of 104.2 KWh/d. 4.4.4.2 Agricultural activities Energy requirements, subdivided into power, heat, and fuel for the various farming activities, are summarized in Table 4.3. For all farming activities, 824 tractor hours per year are needed, with a total fuel consumption of 6476 L (about 1711 US gallons). Requirements for electricity and heat energy are calculated to be 39.5 and 75.0 MWh/ y, respectively. Summarizing the energy demand of farm production and that of the household and administration space, the following energy is required for the entire farm: Heating energy: 125.4 MWh/y Electricity: 77.6 MWh/y Fuel: 6476.2 L 4.4.4.3 Site energy production Total energy demand calculated for heat was 125.4 MWh/y, and for electricity, 77.6 MWh/y. The required energy should be produced fully on the farm. Surplus electric energy can be sold to the public energy network. The energy farm exploits exclusively renewable energy resources. The concept includes a combined use of solar and wind energy as well as energy from biomass. The energy supply system of the farm consists of the following technologies:

Cereals

Potatoes Oil plants

Vegetables

Fruit trees

1

Th, quarter tractor hour.

Soil Preparation care Harvest/storage Soil Preparation Care Harvest/storage Processing Care Harvest processing Care harvest/transport port storage e Feeding other

40.0

Th1/area

Electricity, MWh

Heat, MWh

144.0 0 72.0 10.0 64.0

1166.4 583.2 81.0 518.4

19.0

10.0

95.0 95.0 91.2 67.0 16.0 20.0

76.5 769.5 738.7 542.7 129.6 162.0

5.5

9.8 7.5

5.0 6.5 10.0

7.4

7.5 6.0

3.5

704.2 120.0 120.0 824.2

33.7 5.8 5.8 39.5

52.7 81.1 e 60.8 48.6

5.0

79.0 5.0 5.0 84.0

Fuel, L

75.0 75.0

5704.2 972.0 972.0 6476.2

N. El Bassam

Subtotal 1 Animal husbandry Subtotal 2 Total

Soil Preparation Care Harvest Straw Transport/recovery

72

Table 4.3 Electricity, heat energy, and fuel requirement on the farm. Branches Activities Area, ha

73

Planning of integrated renewable communities

Table 4.4 Estimated annual energy production on the farm. Heat energy

Photovoltaic plant Thermal solar collector Windmill CHP (cogeneration) Stirling engine (with biogas) Total 1

Electricity

5.0 MWh1 40.0 MWh 120.0 MWh2 125.0 MWh 1.6 MWh

260.0 MWh 4.4 MWh 304.4 MWh

251.6 MWh 2

Calculated on the basis of average global sun radiation of 2.7 KWh/m per d and its energetic use of 10e12%. 2 Wind speed on the site ¼ 4.5 5.0 m/s3) annual operation hours ¼ 2100 h.

• Photovoltaic plant of approximately 5.3 kW (50 modules with an output capacity of 105 Wp/module on a surface of 50 m2) • Thermal solar collectors with a 100-m2 collecting surface • Windmill of 300 kW • Wood/biomass-CHP, 100 kW thermal capacity • Stirling engine with 10 kW electrical capacity and 20 kW thermal capacity operated with biogas • Biogas plant with a production capacity of 750e1000 m3 biogas/y. Table 4.4 shows the contribution of various technologies to produce energy from renewable sources. 4.4.4.4 Origin of biomass The biomass for energy generation is provided on the farm: 15 ha (37 acres) will be cultivated with fast-growing tree species such as eucalyptus and poplars, as well as various energy plants such as giant reed, miscanthus, and different tall grasses. Consequently, the farm uses about 90 tons of biomass annually as energy raw material producing approximately 400 MWh/y. The biogas plant is operated with animal manure, plant waste, and farm residues (estimated quantity is about 40 tons/y) producing 700e750 m3 biogas. A Stirling engine operated with biogas produces 4.4 MWh/y heat and 1.6 MWh/y electricity. The total quantity of energy generation from the produced biomass on the farm amounts to 304.4 MWh/y heat energy.

74

N. El Bassam

Table 4.5 Energy demand and energy generation. Solid Oil Solar Wind biomass plants Biogas energy energy

Total

% of total demand

Heat

444.4

117.8

209.1

194.4

400.0 MWh

4.4 MWh 1.6 MWh

Electricity Fuel

40.0 MWh 7.5 MWh

200.0 MWh

2500 L

2500 L 35.2

4.4.4.5 Contribution of different renewable energy sources Table 4.5 shows the contribution of different energy sources by percentage to cover the total energy requirement of the farm (heat: 377.4 MWh/y; power: 107.6 MWh/y; fuel: 7093.4 L). 4.4.4.6 Investment requirement The investment requirement has been estimated on the basis of information collected from different professional organizations, energy agencies, and energy producers. The total sum of the required investment, including all capital and additional costs such as financing and services costs, amounts to about V918,000 (about US $1,162,188.00). In this case, the estimated total investment requirement consists of the following costs of equipment, installation and service (Box 4.1).

Box 4.1 Estimated total investment costs. Windmill (300 kW) Photovoltaic solar cells for power generation (50 units with capacity of 105 Wp/units) þ installation Solar collectors for heat energy (100 m2) BiomasseCHP Biogas plant with reservoir and generator Oil mill þ tank þ installation Wood cutter þ container Plantation of energy plants Financing costs, planning and service, etc. Total (V)

300,000 38,000 15,000 125,000 75,000 110,000 35,000 125,000 95,000 918,000

75

Planning of integrated renewable communities

4.5 Case study II: arid and semiarid regions The main objectives of the plan are the implementation of a farm settlement that is autonomous in its energy supply including desalination and cooling facilities from renewable energy sources.

4.5.1 Specifications Climate: arid and semiarid regions and islands Farm size: 100 ha Production: crops, vegetables, horticulture, energy plantation, animal husbandry, pisciculture, and apiculture (Box 4.2). The infrastructure of the farm might consist of a residential building with an administration space, office rooms (using an area of 400 m2), storage rooms, workshops, greenhouses, and stable facilities.

4.5.2 Farm production Table 4.6 presents production data for different branches (Box 4.3).

4.5.3 Energy requirement 4.5.3.1 Administration and household The area for the living and administration buildings is projected to be approximately 400 m2. Determination of cooling load requirement is based on the following criteria:

Box 4.2 Distribution of farm area. Crops and root plants Oil plants Vegetables Fruit trees Grassland Building Total

60 ha 10 ha 5 ha 5 ha 17 ha 3 ha 100 ha

76

N. El Bassam

Table 4.6 Agricultural production per year (estimated). Production branch Products Area

Quantity

Crop production

40.0 ha

100 t

10.0 ha 20.0 ha 5.0 ha 5.0 ha e 2.0 ha

15 t 3000 t 100 t 200 t 75,000 units 2.0 t ¼ life weight

Vegetables Fruit trees Animal husbandry

Cereals (wheat and barley) Sunflower Potatoes Different products Apples, pears, etc. Hens (eggs) Sheep breeding1

1

1.0 lamb/y, dt ¼ 0.1 ton.

• The farm buildings should be well-insulated so that relatively low energy for cooling will be required, which corresponds to 1 ton/20 m2 (3.5 kW/ton). • Total hours needed for cooling will be approximately 4320 h/a (18 h/ d  8 months/y  30 d) (Box 4.4). Table 4.7 indicates the electricity needed for various administration and household activities (6 persons). A basic load of 2.5 kW for the entire farm is projected for permanent (24 h/d) power generation for circulating pumps, ventilation, engines, emergency lighting, refrigerators, and so on. Total electricity consumption amounts to 38.1 MWh/y, corresponding to consumption of 104.2 KWh/d. Box 4.3 Activities of the created integrated energy farm.

Crop production Energy plantation Grassland/fodder

70 ha 15 ha 2 ha

Horticulture

· ·

Vegetables Fruit trees:

5 ha 5 ha

Animal husbandry

· ·

Sheep breeding (n) chicken (n)

100 500

77

Planning of integrated renewable communities

Box 4.4 Formula for cooling.

Cooling load requirement (MWh/y) ¼ area (m)  3.5 kW/20  4320 h ¼ 400  3.5/20  4320 ¼ 302.4 MWh/y of heat energy to be used in absorption chillers to produce chilled water for cooling purposes. Hot water requirement is calculated to be 40 L/d per person. The annual demand for energy for hot water from 20 to 60 C is calculated as follows (Box 4.5).

Box 4.5 Formula for hot water. 1 783: 0 kWh ¼ y ¼ person 4 For all buildings, a cooling load of approximately 75.0 kW should be available, so that the entire annual energy demand will be 302.4 MWh/y. 40  ð60  203  4: 19  365  0 : 000; 278

4.5.3.2 Agricultural activities Energy requirement, subdivided into power, heat, and fuel for the various farming activities, is summarized in Table 4.8. For all farming activities 24.2 tractor h/y are needed with a total fuel consumption of 6476.2 L. The requirement of electricity and heat energy is calculated to be 39.5 MWh/a and 75.0 MWh/y, respectively. Summarizing the energy demand of farm production and that of household and administration space, we will have the following energy requirement for the entire farm: heating energy: 377.4 MWh/y; electricity: 107.6 MWh/ y; and fuel: 7093.4 L. 4.5.3.3 Energy production on the farm Total energy demand was calculated for heat at 377.4 MWh/y, and for electricity, 107.6 MWh/y. The required energy should be produced fully on the farm. The surplus electric energy will be used to power the desalination system. Table 4.7 Power requirement in MWh/y. Fields Connection value, KW Full load hours, h Consumption, MWh/y

Basic load 2.5 Household 2.5 Cooking 8.2 Administration 1.5 Workshop 12.1 Total 26.8

8760 1100 730 1300 450 12,340

21.90 2.75 5.99 1.95 5.45 38.04

78

N. El Bassam

Table 4.8 Estimated annual energy production on the farm. Heat energy Electricity

Photovoltaic plant Thermal solar collector Windmill CHP (cogeneration) Stirling engine (with biogas) Total 1

7.5 MWh1 e 200.0 MWh2,3

e 40.0 MWh e 360.0 MWh 4.4 MWh

1.6 MWh

404.4 MWh

208.1 MWh 2

Calculated on the basis of average global sun radiation of 5.0 kWh/m per day and efficiency of 10e12%. 2 Wind speed on the site ¼ 3.5e4.0 m/s. 3 Annual operation hours for CHP ¼ 3100 h.

The energy farm should exploit exclusively renewable energy resources. The energy supply system of the farm consists of the following technologies: a combined use of solar and wind energy as well as energy from biomass: • Photovoltaic plant of approximately 5.3 kW (50 modules with an output capacity of 105 Wp/module on a surface of 50 m2) • Thermal solar collectors with a 100 m2 collecting surface • Windmill of 300 kW • Wood/biomass-CHP with 100 kW thermal capacity • Stirling engine with 10 kW electrical capacity and 20 kW of thermal capacity operated with biogas • Biogas plant with a production capacity of 750e1000 m3 biogas/y. Table 4.9 shows the contribution of various technologies to produce energy from renewable sources. 4.5.3.4 Origin of biomass The biomass for energy generation is provided on the farm; 15 ha will be cultivated with fast-growing tree species such as eucalyptus and poplars, as well as various energy plants such as giant reed, miscanthus, and different tall grasses. Consequently, the farm disposes of about 90 tons of biomass annually as energy raw material producing approximately 400 MWh/y. The biogas plant operates with stable manure, plant waste, and farm residues (estimated quantity of about 40 tons/a) producing 700e750 m3 biogas. A Stirling engine operated with biogas produces 4.4 MWh/y heat and 1.6 MWh/y electricity. The total quantity of energy generation from the produced biomass on the farm amounts to 404,4 MWh/y heat energy.

Heat Electricity Fuel

400.0 MWh

4.4 MWh 1.6 MWh 2500 L

Solar- energy

Wind- energy

Total

% of total demand

40.0 MWh 7.5 MWh

200.0 MWh

444.4 209.1 2500 L

117.8 194.4 35.2

Planning of integrated renewable communities

Table 4.9 Energy demand and energy generation. Solid biomass Oil plants Biogas

79

80

N. El Bassam

4.5.3.5 Contribution of different renewable energy sources Table 4.9 shows the percent contribution of different energy sources to cover the total energy requirement of the farm (heat: 377.4 MWh/y; power: 107.6 MWh/y; fuel: 7093.4 L). 4.5.3.6 Investment requirement The investment requirement has been estimated on the basis of information collected from different professional organizations, institutions of energy supply, and producers. The total sum of the required investment, including all capital and additional costs such as financing and services, amounts to about V1,118,000 (about US $1,415,388) (Box 4.6).

Box 4.6 Estimated total investment requirement consisting of costs of plants, installation, and service.

1. Windmill (300 kW) 2. Photovoltaic solar cells for power generation (50 units with capacity of 105 Wp/units) þ installation 3. Solar collectors for heat energy (100 m2) 4. BiomasseCHP 5. Biogas plant with gas storage and generator 6. Oil press þ tank þ installation 7. Wood cutter þ container 8. Plantation of energy plants 9. Financing costs, planning and service, etc. 10. Reverse osmosis desalination unit (capacity 10 m3/h) 11. Absorption chillers and other air-conditioning equipment 12. Irrigation equipment Total: EUR

V300,000 V38,000 V15,000 V125,000 V75,000 V110,000 V35,000 V125,000 V9000 V100,000 V50,000 V50,000 V1,118,000

Further reading El Bassam, N., Maegaard, P., 2004. Integrated Renewable Energy for Rural Communities: Planning Guidelines, Technologies and Applications. Revised, N. El Bassam 2020. Elsevier Science, Amsterdam, pp. 50e70.

CHAPTER FIVE

The watereenergyefood nexus N. El Bassam International Research Centre for Renewable Energy, IFFED.org, Germany

Contents 5.1 Determination of community requirements for energy, water, and food 5.1.1 Definitions 5.2 Modeling approaches 5.2.1 Scenario 1 5.2.2 Scenario 2 5.3 Data acquisition 5.4 Determination of energy and food requirements 5.4.1 Agricultural activities 5.4.2 Households 5.4.2.1 Heat energy 5.4.2.2 Electricity

82 84 85 86 87 88 88 88 92 92 92

5.4.3 Food requirement 5.5 Energy potential analysis 5.5.1 Solar energy 5.5.2 Exploitation of solar energy 5.5.3 Solar thermal system 5.5.4 Solar photovoltaics 5.6 Data collection and processing for energy use 5.6.1 Water and space heating 5.6.2 Drying of agricultural produce 5.7 Wind energy 5.8 Biomass 5.8.1 Energetic use of biomass

92 94 94 97 97 98 99 100 101 101 102 104

5.8.1.1 Combustion 5.8.1.2 Extraction

104 105

5.8.2 Biogas production References Further reading

106 112 112

Distributed Renewable Energies for Off-Grid Communities, Volume 2 ISBN: 978-0-12-821605-7 https://doi.org/10.1016/B978-0-12-821605-7.00025-8

© 2021 Elsevier Inc. All rights reserved.

81

j

82

N. El Bassam

5.1 Determination of community requirements for energy, water, and food The synergies and trade-offs among the water, energy, and food sectors are represented by the watereenergyefood nexus. The Nexus approach is an integrated decision-making practice that can be used by policy-makers to optimize these synergies and manage trade-offs (Figure 5.1). The nexus among energy, water, food availability, and poverty is conclusive and definite. Adequate provision of food, water, and energy is the basis for human existence and contributes to avoiding poverty and social conflicts and preserving human dignity. Poverty is the trigger for hunger, not vice versa. It is never monocausal but multicausal, not only rural but also urban. Mainly women and children are affected by hunger and poverty. Meaningful long-term alleviation of hunger should be rooted in alleviating energy poverty; reducing poverty requires new strategies and political actions. Isolated and single solutions are cosmetic in nature and can never be the solution. Distributed renewable energy solutions can be designed to meet the needs of rural communities.

Energy Energy-Food Interactions

Energy-Water Interactions

• •



Desalination



Energy generation





Water supply and transmission

WaterEnergy-Food Nexus

Water

Biofuels

Industrialized agriculture Food imports and security

Food Water-Food Interactions

• •

Soil and water quality Irrigation and agriculture



Water security and virtual water imports

Figure 5.1 Watereenergyefood nexus tools and synergy among tools assessing individual nexus areas. © 2017 Elsevier Ltd. All rights reserved. https://doi.org/10.1016/j. envsci.2017.07.007. Kaddoura, S. and El Khatib, S. (2017).

The watereenergyefood nexus

83

The water, energy, and food security nexus, according to the Food and Agriculture Organization of the United Nations (FAO), means that water, energy, and food security are very much linked to one another, meaning that actions in any particular area often have effects in one or both of the other areas. These three sectors (water, energy, and food security nexus) are necessary for the benefit of human well-being, poverty reduction, and sustainable development. As the world population is nearing eight billion, increasing demands for basic services also rise, such as the growing desire for higher living standards and the need for more conscious stewardship of the vital resources required to achieve those services; these desires have become more obvious and urgent. Water, energy, and food are essential for human well-being, poverty reduction, and sustainable development. Global projections indicate that the demand for fresh water, energy, and food will increase significantly over the next decades under the pressure of population growth and mobility, economic development, international trade, urbanization, diversifying diets, cultural and technological change, and climate change. Agriculture accounts for 70% of total global freshwater withdrawals, making it the largest user of water. Water is used for agricultural production, forestry, and fishery along the entire agrofood supply chain; it is used to produce or transport energy in different forms. At the same time, the food production and supply chain consumes about 30% of total energy consumed globally. Energy is required to produce, transport, and distribute food as well as extract, pump, lift, collect, transport, and treat water. Cities, industry, and other users claim increasingly more water, energy, and land resources. They also face the problems of environmental degradation and, in some cases, resource scarcity. This situation is expected to be exacerbated in the near future as 60% more food will need to be produced to feed the world population in 2050. Global energy consumption is projected to grow by up to 50% by 2035. Total global water withdrawals for irrigation are projected to increase by 10% by 2050 (FAO, 2011a, Figure 5.2). More than ever, we are convinced that security in food, energy, and water is interwoven with human, economic, and environmental sustainability, and that this interplay is strengthening under growing natural resource scarcity and climate change. This recognition suggests that policy-making and decision-making for sustainability could benefit from a holistic nexus approach that reduces trade-offs and builds synergies across sectors, and thus helps reduce costs and increase benefits for humans and nature,

84

N. El Bassam

Figure 5.2 Impact of changing trends in resource consumption. Water, food, and energy security can no longer be viewed as an isolated global problem. From https:// www.ibgeographypods.org/b-impacts-of-changing-trends-in-resource-consumption.html, 2020. “The Water-Energy-Food Nexus: A new approach in support of food security and sustainable agriculture”. Food and Agriculture Organization of the United Nations. 2014. Retrieved 2019.

compared with independent approaches to managing water, energy, and food, without compromising the resource basis on which humanity relies (https://www.die-gdi.de/buchveroeffentlichungen/article/sustainability-inthe-water-energy-food-nexus-2/2017).

5.1.1 Definitions Water security is defined in the Millennium Development Goals as “access to safe drinking water and sanitation,” both of which have become a human right. Although not part of most water security definitions, the availability of and access to water for other human and ecosystem uses is also important from a nexus perspective.

The watereenergyefood nexus

85

Energy security has been defined as “access to clean, reliable and affordable energy services for cooking and heating, lighting, communications and productive uses” (United Nations) and as “uninterrupted physical availability [of energy] at a price which is affordable, while respecting environment concerns.” Food security is defined by the FAO as “availability and access to sufficient, safe and nutritious food to meet the dietary needs and food preferences for an active and healthy life.” Adequate food has also been defined as a human right. The emphasis on access in these definitions also implies that security is not so much about average (e.g., annual) availability of resources; it has to encompass variability and extreme situations such as droughts or price shocks, and the psychological resilience of the poor https://en.wikipedia. org/wiki/Water,_energy_and_food_security_nexus.

5.2 Modeling approaches The modeling procedure should include identification and determination of the following parameters: 1. Site conditions: • Climate: temperature, amount, and distribution of precipitation, sunshine duration, and wind velocity (annual mean) • Soil conditions, irrigation possibilities, and so on • Factors of production: capital, machines, building, and agricultural area 2. Energy requirement per year for food production for households 3. Basic food requirement per person per year 4. Number of energy consumers (persons and households) 5. Site energy potential: solar energy, wind energy, and biomass 6. Preparation of a master production schedule for food and energy production 7. Selection and installation of suitable technical tools using the renewable energy resources of the site 8. Energy production and use management 9. Environmental impact 10. Social and economic impact To elaborate, develop, and establish an integrated energy farm, two scenarios were considered.

86

N. El Bassam

5.2.1 Scenario 1 Initial conditions: The farm size (Ftn) is known; that is, an already existing farm, or the agricultural area is limited (Figure 5.3). Objectives: The available area (Ftn) should be managed to achieve a high degree of self-sufficiency for a maximum number of people with basic food (Nx) and energy.

Initial condition Production area (Ftn) is limited

Estimation of initial conditions of the site

Estimation of energy potential

Planning

EDP

-

Data climate soil and water Yield data factors of production renewable energy resources Technology energy resources Land use systems cultivation branches property circumstance

-

Data Socio-economical data food consumption Market structure Population number of persons

INPUT

CULTURES

ENERGY Technologies

PRODUCTION

EDP

OUTPUT

food requirement

Energy heat energy, electricity and fuel

Food production THROUGHPUT

energy requirement

Group of persons supplied with food and energy

Figure 5.3 Flowchart for the modeling approach (Scenario 1).

87

The watereenergyefood nexus

5.2.2 Scenario 2 Initial conditions: The farm size (Ftx) is variable; that is, the size could be adapted according to needs (Figure 5.4). Objectives: High degree of self-sufficiency for a determined number of people with basic food and energy should be achieved.

Data

Initial condition Production area (Ftn) is not limited

Estimation of initial conditions of the site food requirement

soil and water

Estimation of energy potential

energy requirement

EDP

number of energy user

Planning INPUT

Land use systems property circumstance

Data Socio-economical food consumption

ENERGY

CULTURES

Population number of persons

Technologies

PRODUCTION

OUTPUT

production

THROUGHPUT

100% supply of target Group of persons with basic food and energy

Figure 5.4 Flowchart for the modeling approach (Scenario 2).

88

N. El Bassam

5.3 Data acquisition The data shown in Table 5.1 should be identified in detail to plan, model, and implement an integrated energy farm.

5.4 Determination of energy and food requirements 5.4.1 Agricultural activities Agriculture is itself an energy conversion process: the conversion of solar energy through photosynthesis to food energy for humans and feed for animals. Primitive agriculture involved little more than scattering seeds

Table 5.1 Data acquisition (overview). To be measured To be calculated

To be recorded

External data: climate data:

Precipitation and distribution Temperature Annual temperature variations

X

X

X

X X

Socioeconomic data:

Size of population Property structures Age structure of population Education Economy data, number of trade companies, and industry Employment situation Land use: agriculture and forestry Land division: crop production, animal breeding, fruit cultivation, vegetable, pisciculture, forestry area, etc. Number of households

X X X X X

X X

X

X

The watereenergyefood nexus

89

on the land and accepting the scanty yields that resulted. Modern agriculture requires energy input at all stages of agricultural production, such as the direct use of energy for farm machinery, water management, irrigation, cultivation, and harvesting. Postharvest energy use includes energy for food processing, storage, and transport to markets. In addition, many indirect or sequestered energy inputs are used in agriculture in the form of mineral fertilizers and chemical pesticides, insecticides, and herbicides. Although industrialized countries have benefited from these advances in energy availability for agriculture, developing countries have not been so fortunate. Energizing the food production chain has been an essential feature of agricultural development throughout history and is a prime factor in helping to achieve food security. Developing countries have lagged behind industrialized countries in modernizing their energy inputs for agriculture. Agriculture accounts for only a small proportion of the total final external commercial energy demand in both industrialized and developing countries. In the Organization for Economic Cooperation and Development (OECD) countries, for example, around 3e5% of commercial energy consumption is used directly in the agricultural sector. In developing countries, estimates are more difficult to find, but the equivalent figure is likely to be similar, in the range of 4e8% of total final commercial energy use. Data for nonrenewable energy use in agriculture also exclude the energy required for food processing and transport by agroindustries. Estimates of these activities range at twice the energy reported solely in agriculture. Definitive data do not exist for many of these stages. This is particularly problematic in analyzing developing country energy statistics. In addition, the data conceal how effective these energy inputs are in improving agricultural productivity. The relationships between the amounts and quality of the direct energy inputs to agriculture and the resulting productive output are the most interesting. Looking more closely at energy use in specific crops, comparisons of commercial energy use in agriculture for cereal production in different regions of the world are listed in Table 5.2. The relationship between commercial energy input and cereal output per hectare for the main world regions is also shown in Figure 5.5. These data, although relatively old, indicate that developing countries use less than half the commercial energy input (whether in terms of energy per hectare of arable land or energy per ton of cereal) compared with industrialized countries. However, developing countries are not necessarily more efficient in their use of energy for agricultural production.

90

N. El Bassam

Table 5.2 Commercial energy use and cereal output (1982).

Region

Africa Latin America Far East Near East All developing countries average All industrialized countries average World average

Energy per hectare of arable land (kgoe/ha)

Energy per ton of cereal (kgoe/t)

Energy per agricultural worker (kgoe/person)

18 64 77 120 96

20 32 43 80 48

26 286 72 285 99

312

116

3294

195

85

344

kgoe, kilogram of oil equivalent. From Stout (1990).

Figure 5.5 Cereal yield and energy input per hectare for the world’s main regions.

A comparison of the commercial energy required for rice and maize production by modern methods in the United States and transitional and traditional methods used in the Philippines and in Mexico is shown in Table 5.3. The data show that modern methods give greater productive yields and are much more energy-intensive than transitional and traditional methods. These methods include the use of fertilizer and other chemical inputs, more extensive irrigation, and mechanized equipment. The energy data for production (Ep) depends first on the size of the farm (Sq), the degree of mechanization, and the production activities. There is the requirement for fuel (Ep1) and electricity (Ep2). Fuel is required for activities of soil preparation and cultivation, as well as for harvesting and transportation (Box 5.1).

91

The watereenergyefood nexus

Table 5.3 Rice and maize production by modern, transitional, and traditional methods. Rice production Maize production Modern (United States)

Energy input 64,885 (MJ/ha) Productive 5,800 yield (kg/ ha)

Transitional (Philippines)

Modern Traditional (United (Philippines) States)

6,386

170

2,700

1,250

Traditional (Mexico)

30,034

170

5, f

950

From Stout (1990).

Box 5.1 Fuel requirement (Ep1) is calculated as: Ep1ð1=yÞ¼ working duration=ha  areaðSq1Þfuel consumption=h=Machine X Ep1ðliter=yearÞ ¼ Sh=ha  Sq1  al (5.1) where Sh ¼ working duration per machine and per hectare; Sq1 ¼ size of the field; Al ¼ fuel consumption per machine and per hour; 1 . n ¼ different cultivation branches; and ha ¼ hectare.

The electricity requirement depends on the degree of mechanization. Electricity is required especially for animal production, but also for the storage, cooling, and drying of crops. The energy requirement for production (Ep2) is the sum of all use factors for all electric instruments and machines to be employed on the farm (Boxes 5.2 and 5.3).

Box 5.2 Energy requirement for production (Ep2): Ep2ðMWh=yÞ ¼

X

h=y  all

(5.2)

where h/a ¼ working duration per machine and per year; all ¼ power requirement per machine and per hour; and 1 . n ¼ different cultivation branches.

92

N. El Bassam

Box 5.3 Considering Formulas (5.1) and (5.2) the total energy requirement for agricultural production may be calculated as: Ep ¼ Ep1 þ Ep2 þ

X

a

where a ¼ losses in kWh/a per machine.

5.4.2 Households The energy requirement for households (Eh) is divided into heat energy (for heating, cooling, and hot water preparation) and the power requirement for light, electrical appliances, and cooking. The requirement data are different from region to region, depending on climatic conditions. They should be calculated considering specific site conditions. 5.4.2.1 Heat energy In addition to the site and environmental conditions, the requirement data for heating and cooling loads (Eh1/household) depend on the construction type of the buildings (such as full insulation or half insulation). Hot water requirements are indicated considering possible different uses in liters/day per person (Boxes 5.4 and 5.5). 5.4.2.2 Electricity See Boxes 5.6 and 5.7.

5.4.3 Food requirement On a broad regional basis, there appears to be a correlation between high per-capita modern energy consumption and food production. Figure 5.6 Box 5.4 Heating and cooling loads for buildings are calculated as:   Eh1ðMWh = yÞ ¼ g W = m2  h=y  FðmÞ2

(5.3)

where g(W/m2) ¼ energy requirement per reference area; h/y ¼ year requirement of energy in full load hour; and F (m2) ¼ energy reference area.

93

The watereenergyefood nexus

Box 5.5 Total heat energy requirement for hot water with temperature of 60 C is determined as: Qðkj=yÞ ¼ m  ðt2  t1 Þ  4:19kj=K  365 Qðkj=yÞ  0:000278 ¼ QðkWh=yÞ where Q ¼ heat energy requirement; m ¼ hot water consumption per day (L); t1 ¼ cold water temperature (10e20 C); and t2 ¼ desired temperature (60 C).

Box 5.6 The total requirement (Eh2/household) results from the use data of all electrical devices existing in a household for different uses: light, communication, cooking, and cooling: Eh2ðMWh = yÞ ¼

X

bðkWÞ  h þ aðkWhÞ

(5.4)

where b(kW) ¼ connected load of the device; h ¼ full load hours in the year; and a ¼ losses in kWh.

Box 5.7 Considering Eqs. (5.3) and (5.4), the total energy requirement of the household is calculated as: EhðMWh = yÞ ¼ fEh1ðMWh = yÞ þ Eh2ðMWh = yÞ  ng where n ¼ number of households.

Figure 5.6 Modern energy consumption and food intake. kgoe, kilogram of oil equivalent.

94

N. El Bassam

shows data for daily food intake per capita and the annual commercial energy consumption per capita in seven world regions (FAO, 1995). (The regions are Sub-Saharan Africa Asia, East Asia/Pacific, Latin America/ Caribbean, Middle East/North Africa, Europe, and OECD.) Although broad data on a regional basis conceal many differences among countries, crop types, and urban and rural areas, the correlation is strong in developing countries, where higher inputs of modern energy can be assumed to have a positive impact on agricultural output and food production level. The correlation is less strong in industrialized regions where food production is near or above required levels and changes in production levels may reflect changes in diet and food fashion rather than advantages gained from an increased supply of modern energy. Basic food requirement (person/year) is divided into requirement data for carbohydrates, proteins, vitamins, and fats. They are different from country to country. To create an integrated energy farm and prepare a land use plan, regional and national consumer data should be used. Table 5.4 lists basic foods and their resources from agricultural production. The data for food consumption are normally known and must be used while planning. Which kinds of food will finally be produced on the farm depends on the climatic and soil conditions of the site (Box 5.8).

5.5 Energy potential analysis 5.5.1 Solar energy The total solar radiation that strikes the earth’s surface amounts to 1018 kWh/a, which is many times greater than the current global energy demand. In the case of vertical incidence (solar altitude of 90 degrees), the radiation intensity of global radiation can reach a level of 1100 W/m2. The daily sum of global radiation (horizontal surface) on a sunny day in the vicinity of the equator is estimated to be 6e8 kWh/m2 per day. Table 5.4 Basic foods and their sources. Food components Sources

Carbohydrates Protein Vitamin Fats

Cereals Other crops Animal Plant Fruits Vegetables Oil plants Animal

Indicators

kg/person per year kg/person per year Number of animals/farm kg/person per year kg/person per year kg/person per year kg/person per year Number of animals/farm

95

The watereenergyefood nexus

Box 5.8 To determine the production area for each agricultural branch, the following calculation formula can be used: Ar1.n ¼ ðY1.n Þ=ðF1.n Þ where Ar ¼ space requirement in hectares; Y ¼ yield/area; F ¼ requirement/ head per year; and 1 . n ¼ different basic food.

It is possible to estimate the proportion of global radiation represented by diffuse radiation (important for the use of solar collectors). It is a function of the solar altitude and the degree of cloudiness and ranges from 10 to 85%. Available data regarding regional solar radiation come from measurements on horizontal surfaces. To increase their efficiency, solar collectors are usually mounted at a tilt angle. Solar radiation is divided into two components: direct and diffuse radiation. Conversion of direct radiation is relatively simple. However, specific assumptions must be made to convert the diffuse component, because this radiation component is extremely dependent on site conditions and technical facilities. Generally, three estimating procedures are possible: (Boxes 5.9e5.11) 1. The first assumes that sky radiation becomes the predominant part of direct solar proximity. This can be possible only on clear days. 2. The second assumes that sky radiation is distributed uniformly via the entire sky vault. This is an approximation for cloudy days. 3. The third procedure represents a middle ground between the two extremes. One assumes here that, because of the inclination of the absorber, it sees only a part of the sky vault (indeed ½ (1 cos n), but it receives an additional diffuse part of the radiation in the form of ground reflection from the collector environment.

Box 5.9 Total radiation falling on tilt surface GG,g is then:   GG;g ¼ RGG;h W = m2 where R ¼ ratio of the direct radiation on a tilt surface (Go,g ¼ Go cos 4) to the radiation value on a horizontal surface.

96

N. El Bassam

Box 5.10 Entire radiation falling on a tilt surface results in:   GG;g ¼ RGD;h þ RGH;h W = m2 where GD,h ¼ direct radiation on a horizontal surface; and GH,h ¼ sky radiation on a horizontal surface.

Box 5.11 Total radiation falling on a tilt surface consists of three parts: direct component, diffuse component, and reflected part:   GG;g ¼ RGD;h þ ½1 = 2ð1 þ cosnÞGH;h þ sB GG;h W = m2 where n ¼ inclination angle of the receiving face (degree); and sB ¼ reflection coefficient of the surrounding ground.

With these procedures performed, the solar energy potential of a site can be estimated. Extensive local measurements are necessary to evaluate the potential use of solar radiation energy in a particular region. Solar installation sites must be carefully selected. The primary energy supply and the presumed energy demand are the decisive factors in determining the economic feasibility of a particular site. Local measurements should include the following quantities: • global radiation G • direct solar radiation S • diffuse sky radiation H • number of hours of sunshine SD • degree of cloudiness N • air temperature TA • wind direction D • wind intensity F If possible, the measurement should be conducted over a relatively long period (several years). Transformation of the sun’s energy to electricity by photovoltaic (PV) panels and heat energy by solar thermal collectors depends on the type and model of collectors. Therefore, the efficiency of the chosen collector should be considered when calculating the site’s solar energy potential.

The watereenergyefood nexus

97

5.5.2 Exploitation of solar energy Today, solar energy is being used in many ways on various scales. On a small scale, it is used at the household level in goods such as watches, cookers, and heaters. Medium-scale uses, such as in solar architecture houses, include water heating and irrigation. At the community level, it can be used for water pumping, water desalination, purification, and rural electrification. On an industrial scale, solar energy is used for power generation, detoxification, municipal water heating, and telecommunications. In general, there are two basic ways to use solar energy.

5.5.3 Solar thermal system While heat from the sun (over 300 C) is used on a large scale to generate electricity, it can also be used in small- to medium-scale heating, cooling, cooking, and drying equipment. Solar therm-electric technologies employ energy from the sun in the form of heat to generate electricity. The sun evaporates a fluid from which heat transfer systems may be used to operate an engine that drives a power conversion system. In a solar thermoelectric system, sunlight is concentrated with mirrors or lenses to attain a high temperature sufficient for power generation. Parabolic trough systems, centralreceiver systems, parabolic dish systems, and solar ponds are among those used. The basic components of a solar thermoelectric system (Figure 5.7) are a collector system (panels that collect solar radiation), a receiver system, a transport storage system (mainly in the form of fluid that transfers heat between systems) and, finally, a power conversion system converting energy from one form to another. Solar water heaters are relatively simple solar thermal applications that transform solar radiation into heat used to warm water for heating, washing, cooking, and cleaning. Solar water heaters consist of glass-covered collectors with a dark-colored or specially coated absorber panel inside. Water (used as the heat transfer fluid) is warmed by the sun and can be stored in insulated tanks for later use. There are two main components of a typical solar water heating system: a flat plate solar collector and a hot water storage tank. Flat plate collectors absorb solar radiation and conduct the heat to water that circulates through the collector in pipes. Another solar thermal application is a solar drier. There are often two stages to the process. First, solar radiation is captured and used to heat air; then comes actual drying, during which heated air moves through, warms,

98

N. El Bassam

Figure 5.7 Main components of a solar thermoelectric system. Adapted from De Laquil et al. (1993). Table 5.5 Various applications of solar thermal systems on an agricultural farm. Temperature range Applications

Low-grade thermal energy 25 m)

Thamer Mohamed

Gravity turbines Low heads (1e5 m)

Medium flows (0.3e1.5 m3/s)

Source for configuration is Pixabay

Francis turbines (for older turbines)

Reaction turbine Low to medium heads (1.5e20 m)

Medium flows (0.5e4 m3/s)

No longer commonly used except in large storage hydropower systems, although lots of older, smaller turbines are in existence and can be restored. Source for configuration is Hydro-Québec

Hydropower

Waterwheels

223

224

Thamer Mohamed

10.4.1.1 Pelton, Cross-flow and Turgo turbines The Pelton turbine consists of a wheel with a series of split buckets set around its rim; a high velocity jet of water is directed tangentially at the wheel. The jet hits each bucket and is split in half, so that each half is turned and deflected back by an angle of 180 degrees. Nearly all of the energy of the water goes into propelling the bucket, and the deflected water falls into a discharge channel below. The Turgo turbine is similar to the Pelton, but the jet strikes the plane of the runner at an angle (typically 20e25 degrees) so that the water enters the runner on one side and exits on the other. Therefore, the flow rate is not limited by the discharged water interfering with the incoming jet (as is the case with Pelton turbines). As a consequence, a Turgo turbine can have a smaller diameter runner and rotate faster than a Pelton for an equivalent flow rate. The Cross-flow turbine has a drum-like rotor with a solid disk at each end and gutter-shaped slats joining the two disks. A jet of water enters the top of the rotor through the curved blades, emerging on the far side of the rotor by passing through the blades a second time. The shape of the blades is such that on each passage through the periphery of the rotor, the water transfers some of its momentum before falling away with little residual energy. 10.4.1.2 Kaplan and Francis turbines These turbines exploit the oncoming flow of water to generate hydrodynamic lift forces to propel the runner blades. They are distinguished from the impulse type by having a runner that always functions within a completely water-filled casing. In addition, the turbines have a diffuser known as a draft tube below the runner through which the water discharges. The draft tube slows the discharged water and creates suction below the runner, which increases the effective head. Propeller-type turbines are similar in principle to the propeller of a ship, but operate in reversed mode. A set of inlet guide vanes admits the flow to the propeller. These are often adjustable so as to allow the flow passing through the machine to be varied. In some cases, the blades of the runner can also be adjusted, in which case the turbine is called a Kaplan. The mechanics for adjusting turbine blades and guide vanes can be costly and tend to be more affordable for large systems, but they can greatly improve efficiency over a wide range of flows. The Francis turbine is essentially a modified form of propeller turbine in which water flows radially inward into the runner and is turned to emerge

Hydropower

225

axially. For medium-head schemes, the runner is most commonly mounted in a spiral casing with internal adjustable guide vanes. Because the Cross-flow turbine is a less costly (although less efficient) alternative to the spiral-case Francis, it is rare for these turbines to be used on sites with less than 100 kW output. The Pit-Francis turbine was originally designed as a low-head machine, installed in an open chamber (or pit) without a spiral casing. Thousands of such machines were installed in the Europe from the 1920s to the 1960s. Although it was an efficient turbine, it was eventually superseded by the propeller turbine, which is more compact and faster-running for the same head and flow conditions. However, many of these open-flume or wall plate Francis turbines are still in place; hence, this technology is still relevant for refurbishment schemes. 10.4.1.3 Archimedes’ screw and Waterwheel turbines The Archimedes’ screw has been used as a pump for centuries but has only recently been used in reverse as a turbine. Its principle of operation is the same as the overshot waterwheel, but the clever shape of the helix allows the turbine to rotate faster than the equivalent waterwheel and with high efficiency of power conversion (over 80%). However, these are still slowrunning machines that require a multistage gearbox to drive a standard generator. A key advantage of the screw is that it avoids the need for a fine screen and automatic screen cleaner because most debris can pass safely through the turbine. The Archimedes’ screw is proven to be a fish-friendly turbine. The approximate ranges of head and flow of turbines are shown in Table 10.1, whereas power applicable to the different turbine types can be determined from Figure 10.7 (up to 500 kW power). These are approximate and depend on the precise design of each manufacturer.

10.5 Relative efficiencies A water turbine running at a certain speed will draw a particular flow. If there is insufficient flow in the river to meet this demand, the turbine could start to drain the river and its performance rapidly degrades. Therefore, it has to shut down or to change its internal geometry, a process known as regulation. Regulated turbines can move their inlet guide vanes and/or runner blades to increase or reduce the amount of flow they draw. The efficiency of the different turbines will inevitably reduce as they draw less flow. The typical variation is shown subsequently. Therefore, a significant

226

Thamer Mohamed

Figure 10.7 Chart for determination of turbine type. (From Waterturbines).

factor in comparing different turbine types is their relative efficiencies at both their design point and reduced flows. For example, Pelton and Kaplan turbines retain high efficiencies when running below design flow, whereas the efficiency of Cross-flow and Francis turbines falls away more rapidly when run at below half their normal flow (GEI, 2020).

10.6 Assessment of hydropower potential The hydropower obtained from a plant can be calculated using the formula (Eq. 10.1): P ¼ gQH

(10.1)

where P is the power in watts; g is the specific weight of water, usually taken to be 9810 N/m3; H is the head in m (Figure 10.1), and Q is the flow rate or discharge in m3/s following through the turbine blades. Let ht be the efficiency of turbines, hg the efficiency of generator, and h the overall efficiency. The hydropower, P, generated by the plant is given by (Eq. 10.2):  P ¼ ðgÞðht Þ hg ðQÞðHÞ (10.2)

227

Hydropower

The overall efficiency of the plant is equal to the product of the turbine efficiency and the generator efficiency (Eq. 10.3): h ¼ ht  hg

(10.3)

The average value of h is usually taken as 0.8 (Eq. 10.4): P ¼ ðgÞðhÞðQÞðHÞ (10.4) The firm power is the amount of power the plant can deliver throughout the year or 100% of the time. It is the power that will be available when the flow in the river is minimal for a run of river plant. However, the firm power is considerably increased if a storage reservoir is provided. The firm power can be guaranteed to consumers because it will be always available. The installed capacity of the plant is the maximum power that can be developed by all generators of the plant at the normal head and with full flow. Generally, the installed capacity is kept to 1.1 of the peak load. The load factor is the ratio of the average load during a period to the peak load during the same period. It can be written as (Eq. 10.5): Average Load (10.5) Peak Load The load is therefore for a specific period. Depending on the period specified, it can be a daily, weekly, monthly, or yearly load factor if the period is respectively a day, week, month, or year. The load factor will be equal to the capacity factor if the peak load is equal to the plant capacity. Gross theoretical hydropower potential is defined as the annual energy that is potentially available if all natural runoff at all locations can be harnessed down to sea level (or to the border of a region when calculating regional potential) with no energy losses. In this study, the gross hydropower potential (GHP) is calculated in each grid globally using the following (Yuyu et al., 2015) (Eq. 10.6): Load Factor ¼

Egi ¼ ðmi ÞðgÞðDHi Þ

(10.6)

where Egi is the GHP (watt hours) in grid i, mi is the mass of monthly water runoff water in grid i (kg), g is the acceleration due to gravity (m/s2), and DHi is the elevation difference between the grid cell i and the lowest grid cell in each country (m). The elevation data can be derived from hydrological data and maps based on shuttle elevation derivatives at multiple scales (HydroSHEDS). The gross potential of hydropower in each grid is theoretical, not the hydropower that could actually be generated in that grid (Yuyu et al., 2015).

228

Thamer Mohamed

The economic potential of hydropower is defined as the annual energy that can be developed at competitive costs compared with other energy sources. It requires an estimate of the cost to generate electricity from hydropower in each grid. The cost of hydropower generation can be calculated by considering costs for licensing, environmental mitigation, operation, and maintenance for the turbine and generator. The equations for the cost can be written as (Yuyu et al., 2015) (Eqs. 10.7e10.9): C development þ C om Et    c development ¼ ð B ÞðvÞ pqdesign Dhbi ðS a Þ C oe ¼

(10.7) (10.8)

   c om ¼ ðvÞ pqdesign Dhbi ðS a Þ

(10.9)

where Cdevelopment and Com are the cost in US dollars for the hydropower development and annual operation and maintenance; f is the fixed charge rate (unitless), Et is the technical hydropower potential (watt hours); Pdesign is the design capacity, calculated based on a 30% chance that the monthly stream flow will exceed the turbine design capacity and will not be used for power production; Dhi is the head (m); S is the generator rotational speed (rpm); and f, q, b, and a are the empirical parameters (Table 10.2).

Table 10.2 Values of cost parameters (Yuyu et al., 2015). Cost details f

Licensing Construction Fish and wildlife mitigation Recreation mitigation Historical and archeological mitigation Water quality monitoring Fish passage mitigation Turbine upgrade Francis turbine Turbine upgrade Kaplan turbine Turbine upgrade Pelton turbine Turbine upgrade Low head turbine Generator upgrade Fixed operation and maintenance Variable operation and maintenance

þ5

6.1  10 3.3  10þ6 3.1  10þ5 2.4  10þ5 1.0  10þ5 4.0  10þ5 1.3  10þ6 3.0 10þ6 4.0  10þ6 2.4  10þ6 6.0  10þ6 3.0  10þ6 2.4  10þ4 2.4  10þ4

q

b

a

0.70 0.90 0.96 0.97 0.72 0.44 0.56 0.71 0.72 0.71 0.86 0.65 0.75 0.80

0 0 0 0 0 0 0 0.42 0.38 0.42 0.63 0 0 0

0 0 0 0 0 0 0 0 0 0 0 0.38 0 0

Hydropower

229

10.7 Impact of climate change on hydropower generation Hydropower, which is the dominant component of renewable energy, is also under the threat of climate change. Climate change has a large impact on water resources and thus on hydropower. Hydropower generation is closely linked to the regional hydrological conditions of a watershed and reacts sensitively to seasonal changes in water quantity. This shows that the impact of climate change is different in various regions. The impact of climate change on hydropower generation can be quantified by modeling. Based on different modeling results, the impact of climate change on selected hydropower projects in Europe, South East Asia, and North America are demonstrated in this section. The development of hydroelectric power generation in the Upper Danube basin was modeled by Franziska et al. (2011) for two future decades: 2021e30 and 2051e60. By considering 16 climate scenarios, the modeling results show a slight to severe decline in hydroelectric power generation. In general, with these climate trends, mean annual hydroelectric power generation will decline to a range of 17e18 TWh in 2021e30 and a range of 15e17 TWh in 2051e60. Hydropower is a valuable renewable energy resource in India and helps to meet increasing energy demands. The crucial role of climate change in hydropower production in India was studied by using Multimodel (Ali et al., 2018). Modeling results showed that according to the future climate, projected hydropower production will increase up to þ25%. Hydropower is an important renewable energy source in China, but it is sensitive to climate change because the changing climate may alter hydrological conditions (e.g., river flow and reservoir storage). Future changes and associated uncertainties in China’s GHP and developed hydropower potential (DHP) were projected using simulation models by Xingcai et al. (2016). Modeling results showed that the projected annual GHP will change by 1.7% to 2% in the near future (2020e50) and will increase by 3e6% in the late 21st century (2070e99). The projected annual DHP will change by 2.2% to 5.4% (0.7e1.7% of total installed hydropower capacity [IHC]) and 1.3% to 4% (0.4e1.3% of total IHC) for 2020e50 and 2070e99, respectively. The results of the modeling chain conducted by Vinod (2019) on the Steephill Falls hydroelectric project, constructed on Magpie River, Northern Ontario, Canada, showed that annual hydropower generation is not

230

Thamer Mohamed

considerably affected by climate change, but there is a significant seasonal redistribution of energy production. Changes in hydropower revenues compared with the current level for the four seasons (winter, spring, summer, and autumn) are estimated to be 21.1%, 18.4%, 13.4%, and 15.9%, respectively, for midcentury and 23.1%, 19.5%, 20.1%, and 22.9%, respectively, for end-century scenarios. To reduce the vulnerability of hydropower systems to climate change and consequently mitigate the impacts, it will be profitable to adapt suitable measures such as providing additional live storage by optimizing reservoir operations, which also reduces the vulnerability of the system to climate change by 24%. The seasonal alteration in energy production will require modifications in power purchase and sharing agreements with buyers.

References Ali, S.A., Aadhar, S., Shah, H.L., 2018. Projected increase in hydropower production in India under climate change. Sci. Rep. 8, 12450. https://doi.org/10.1038/s41598-018-30489-4. Energypedia, 2018. Hydro Power Basics. https://energypedia.info/wiki/Hydro_Power_ Basics. (Accessed 24 March 2020). Franziska, K., Monika, P., Heike, B., Wolfram, M., Florian, A., Markus, W., 2011. How will hydroelectric power generation develop under climate change scenarios? A case study in the upper Danube basin. Energies 4 (10), 1508e1541. GreenBug Energy Inc., GEI, 2020. Types of Turbines. http://greenbugenergy.com/geteducated-knowledge/types-of-turbines. (Accessed 10 March 2020). Hydro-Québec. Turbines. http://www.hydroquebec.com/learning/hydroelectricite/typesturbines.html. (Accessed 15 March 2020). International Hydropower Association, Iha, 2019. 2019 Hydropower Status Report. https:// www.hydropower.org/statusreport. (Accessed 23 March 2020). International Renewable Energy Agency, IREA, 2019. Hydropower. https://www.irena. org/hydropower. (Accessed 23 March 2020). Pixabay. Waterwheel. https://pixabay.com/images/search/waterwheel/. (Accessed 30 March 2020). Pumpfundamentals. Micro-hydro Installation Sizing (Cross-flow Turbine). https://www. pumpfundamentals.com/micro-hydro-banki.htm. (Accessed 30 March 2020). Shutterstock. Archimedes Screw Turbines Images. https://www.shutterstock.com/search/ archimedesþscrewþturbines. (Accessed 28 March 2020). Vinod, C., 2019. Impacts of Climate Change on Hydropower Generation and Developing Adaptation Measures Through Hydrologic Modeling and Multi-objective Optimization. A Ph.D. Dissertation. University of Windsor, Windsor, Ontario, Canada. Water Power Technologies Office, WPTO. The U.S. Department of Energy’s (DOE). https://www.energy.gov/eere/water/types-hydropower-plants. (Accessed 22 March 2020). Waterturbines. Types of Water Turbines. http://waterturbines.wikidot.com/main:types-ofwater-turbines. (Accessed 20 March 2020). Xingcai, L., Qiuhong, T., Nathalie, V., Huijuan, C., 2016. Projected impacts of climate change on hydropower potential in China. Hydrol. Earth Syst. Sci. 20, 3343e3359. Yuyu, Z., Mohamad, H., Steven, J.S., Allison, T., 2015. A comprehensive view of global potential for hydro-generated electricity. Energy Environ. Sci. https://doi.org/ 10.1039/c5ee00888c.

CHAPTER ELEVEN

Marine energy Thamer Mohamed Department of Water Resources Engineering, College of Engineering, University of Baghdad, Baghdad, Iraq

Contents 11.1 Introduction 11.2 Ocean thermal energy conversion 11.2.1 Ocean thermal energy conversion systems technology

231 233 234

11.2.1.1 Closed cycle 11.2.1.2 Open cycle 11.2.1.3 Hybrid cycle

235 236 236

11.3 Advantages and disadvantages 11.3.1 Advantages 11.3.2 Disadvantages 11.4 Ocean tidal power 11.5 Ocean wave power 11.5.1 Offshore systems 11.5.2 Onshore systems 11.6 Environmental and economic challenges References Further reading

237 237 238 239 240 242 242 243 244 245

11.1 Introduction The gravitational pull of the moon and sun along with the rotation of the earth cause the tides. In some places, tides cause water levels near the shore to vary up to 12 m. People in Europe harnessed this movement of water to operate grain mills more than 1000 years ago. Today, tidal energy systems generate electricity. Producing tidal energy economically requires a tidal range of at least 3 m. Therefore, marine energy is an old practice; however, modern technologies are used in the production of marine energy. Energy demand is rapidly growing on a global scale and more energy than ever before is needed. Marine energy could have a more significant role in meeting global energy demands.

Distributed Renewable Energies for Off-Grid Communities, Volume 2 ISBN: 978-0-12-821605-7 https://doi.org/10.1016/B978-0-12-821605-7.00012-X

© 2021 Elsevier Inc. All rights reserved.

231

j

232

Thamer Mohamed

Marine energy is sometimes called marine power, ocean energy, or ocean power; it is a seawater-based renewable type of energy in which the kinetic and potential energies in tides, ocean waves, and river currents are used to drive turbines and produce electricity. The production of such clean energy will reduce dependence on fossil fuels. In addition, differences in salinity (salt levels) and temperature that occur in water bodies create dynamic forces that can be used to produce power. These different forms of marine renewable energy will be available for as long as the tides continue to ebb and flow and the rivers continue to run. Marine energy provides a significant contribution to the production of low-carbon renewable energy around the world. The use of natural resources in a sustainable energy supply contributes to the world’s future. It can contribute to reduce carbon emissions while minimizing impacts on the marine environment. The energy density of various ocean energies is relatively low in general. The maximum tidal range of tidal energy in the world is about 17 m, and the maximum range in China is 9.3 m. The average wave height of wave energy for the largest single station in the world is 2 m, and that for the largest single station in China is 1.6 m. The maximum flow rate of ocean currents is 2.5 m/s, and the maximum in China is 1.5 m/s. For energy created from temperature differences, the maximum temperature difference between surface seawater and deep seawater in the world is 24 C, comparable to the value in China. Energy created from salinity differences has the greatest energy density among all ocean energies, with an osmotic pressure of 24 atm, equivalent to a head of 240 m and also comparable to the value in China. Statistics show that the theoretically developable capacity of ocean energies is 76,600 GW, including a technologically developable capacity of 6400 GW (Table 11.1). Tidal power generation is a relatively welldeveloped ocean power generation technology. As at the end of 2013, the installed capacity of ocean power generation in the world had reached approximately 530 MW, and the world’s largest ocean power station was the 254 MW tidal power generation station in South Korea (Zhenya, 2016). Ocean energy has the potential to provide a substantial amount of new renewable energy around the world. Table 11.1 shows theoretical estimated wave energy in various regions in the world, whereas Figure 11.1 shows the world’s installed marine energy.

233

Marine energy

Table 11.1 Regional theoretical potential of wave energy (Matthew et al., 2016). Region Wave energy TWh/y

Western and Northern Europe Mediterranean Sea and Atlantic Archipelagos (Azores, Cape Verde, and Canaries) North America and Greenland Central America South America Africa Asia Australia, New Zealand, and Pacific Islands Total

2800 1300 4000 1500 4600 3500 6200 5600 29,500

600

Installed Capacity (MW)

500 400 300 200 100 0 2010 2011 2012 2013 2014 2015 2016 2017 2018 2019 Year

Figure 11.1 Installed capacity for marine energy in the world. (From IRENA (2019)).

Marine energy provides significant contribution to the production of low-carbon renewable energy around the world. This energy contributes to the world’s future. It can reduce carbon emissions while minimizing impacts on the marine environment.

11.2 Ocean thermal energy conversion Oceans cover more than 70% of the earth’s surface. As the world’s largest solar collectors, oceans generate thermal energy from the sun. Although the sun affects all ocean activity, the gravitational pull of the

234

Thamer Mohamed

moon primarily drives the tides and wind powers the ocean waves. A process called ocean thermal energy conversion (OTEC) uses heat energy stored in the earth’s oceans to generate electricity. OTEC works best when the temperature difference between the warmer top layer of the ocean and the colder deep ocean water (DOW) is about 20 C (36 F). These conditions exist in tropical coastal areas, roughly between the Tropic of Capricorn and the Tropic of Cancer. To bring cold water to the surface, OTEC plants require an expensive large-diameter intake pipe, which is submerged a mile or more into the ocean’s depths. Some energy experts believe that if it could become cost-competitive with conventional power technologies, OTEC could produce billions of watts of electrical power. OTEC technology is not new. In 1881, Jacques Arsene d’Arsonval, a French physicist, proposed tapping the thermal energy of the ocean. However, it was d’Arsonval’s student, Georges Claude, who in 1930 actually built the first OTEC plant in Cuba. The system produced 22 kW of electricity with a low-pressure turbine. In 1935, Claude constructed another plant aboard a 10,000-ton cargo vessel moored off the coast of Brazil. Weather and waves destroyed both plants before they became net power generators. (Net power is the amount of power generated after subtracting power needed to run the system.) In 1956, French scientists designed another 3MW OTEC plant for Abidjan, Ivory Coast, West Africa. The plant was never completed, however, because it was too expensive. The United States became involved in OTEC research in 1974 with the establishment of the Natural Energy Laboratory of Hawaii Authority. The laboratory has become one of the world’s leading test facilities for OTEC technology.

11.2.1 Ocean thermal energy conversion systems technology OTEC is an energy generation technology that uses cold DOW and warm surface water to produce electricity. Active development of OTEC was started in the 1970s with Hawaii as a major research and development center. In the several decades that followed, small pilot-scale closed-cycle and open-cycle OTEC plants were successfully designed, constructed, and tested. An open-cycle OTEC plant produces both electricity and freshwater. Besides its low temperature, DOW is nutrient-rich and free from pathogenic bacteria. DOW-enhanced open ocean mariculture can significantly increase the world fish catch and induce an air to water transfer of greenhouse gas CO2. Therefore, an integrated development of DOW as a natural resource

Marine energy

235

is the center of a blue revolution that has the potential to solve four of the most urgent world problems: energy, freshwater, food, and global warming (Liu, 2018). 11.2.1.1 Closed cycle In a closed-cycle OTEC system, warm seawater vaporizes a working fluid such as ammonia flowing through an evaporator. The vapor expands at moderate pressures and turns a turbine coupled to a generator that produces electricity. The vapor is then condensed in a condenser using cold DOW pumped from the ocean’s depth through a cold-water pipe. The condensed working fluid is pumped back to the evaporator to repeat the cycle. The working fluid remains in a closed system and circulates continuously, as shown in Figure 11.2A (Liu, 2018). In 1979, the Natural Energy Laboratory and several private-sector partners developed a mini OTEC experiment, which achieved the first successful at-sea production of net electrical power from closed-cycle OTEC. The mini OTEC vessel was moored 1.5 miles (2.4 km) off the Hawaiian coast and produced enough net electricity to illuminate the ship’s light bulbs and run its computers and televisions. In 1999, the Natural Energy Laboratory tested a 250-kW pilot OTEC closed-cycle plant, the largest such plant ever put into operation.

Figure 11.2 Ocean thermal energy conversion systems: (A) closed-cycle, and (B) open-cycle. DOW, deep ocean water. (From Liu (2018)).

236

Thamer Mohamed

11.2.1.2 Open cycle An open-cycle OTEC system uses warm seawater as the working fluid. The pressure above the warm water is lowered sufficiently for the water to boil and vaporize at the ambient temperature of about 25 C. This vapor is used to drive the turbine and DOW is used to condense the water vapor exiting from the turbine. An open-cycle system requires a vacuum to be created and maintained to boil and vaporize the working fluid; thus, it is less efficient than a closed-cycle OTEC system. However, it has the advantage of producing freshwater as a by-product, as shown in Figure 11.2B (Liu, 2018). In 1984, the Solar Energy Research Institute (now the National Renewable Energy Laboratory) developed a vertical-spout evaporator to convert warm seawater into low-pressure steam for open-cycle plants; energy conversion efficiencies as high as 97% were achieved. In May 1993, an open-cycle OTEC plant at Keahole Point, Hawaii, produced 50,000 W of electricity during a net power-producing experiment. 11.2.1.3 Hybrid cycle The hybrid cycle combines the best characteristics of the open cycle and closed cycle and the drinkable water generation capabilities of open cycles with the potential for large electricity production capabilities offered by the closed cycles (Figure 11.3).

Vacuum pump

Ammonia Vapor Condenser Vapor

Gases (not condensable)

Spout

Grid

Desalinated

Electric Energy

Ammonia Turbine

Hot seawater

Generator

Ammonia Generator Cold seawater

Liquid Ammonia pump

Cold seawater

Figure 11.3 Diagram for hybrid system. (From Kresala (2018)).

Marine energy

237

In this cycle, warm seawater (which comes from the surface of the ocean) enters a vacuum chamber, where it is flash-evaporated. After that, this steam arrives at a heat exchanger, where it will have the role of a warm fluid and will be used to warm the working fluid, which works in a closed loop. Usually, ammonia is used as a working fluid because it has good transport properties, it is easily available, and it is a low-cost fluid. On the other hand, it is toxic and flammable. Other options are fluorinated carbons such as chlorofluorocarbons (CFCs) and hydrochlorofluorocarbons (HCFCs) or hydrocarbons. CFCs and HCFCs are fully or partly halogenated paraffin hydrocarbons that contain only carbon (C), hydrogen (H), chlorine (Cl), and fluorine (F). They are produced as volatile derivatives of methane, ethane, and propane. They are also commonly known by the DuPont brand name Freon. An interesting part of the process is that as water evaporates, it leaves behind all impurities and salt, so the consequent steam is drinkable. Once heated and evaporated, the working fluid vapor flows through a closedcycle power loop. At this point, the ammonia is used to turn a turbine that is connected to an electricity generator, supplying customers with electric power. After that, the ammonia is condensed using cold seawater from the deepest ocean waters. The noncondensables are then compressed and discharged into the atmosphere. It can be concluded that the role of seawater is double: on the one hand, it works as an intermediary heat transfer medium; on the other, it produces potable water after the condensation process.

11.3 Advantages and disadvantages 11.3.1 Advantages The main advantages of OTEC are (Woodford, 2020): 1. It is clean, green renewable energy that does not involve burning fossil fuels, producing large amounts of greenhouse gases, or releasing toxic air pollution. By helping to reduce our dependence on fuels such as petroleum, OTEC could also help to reduce dependence on oil and reduce marine pollution from oil tanker spills. 2. It could also provide a useful source of power for tropical island states that lack their own energy resources, effectively making them self-sufficient. OTEC can also be used to produce fuels such as hydrogen; the electricity it generates can be used to power an electrolysis plant that would split seawater into hydrogen and oxygen, which could be bottled or piped ashore and then used to power such things as fuel cells in electric cars.

238

Thamer Mohamed

3. Open-cycle OTEC can have a useful part in providing pure, usable water from ocean water. 4. The waste cooling water used by an OTEC plant can also be used for aquaculture (growing fish and other marine food such as algae under controlled conditions), refrigeration, and air-conditioning.

11.3.2 Disadvantages There are main disadvantages of OTEC (Woodford, 2020): 1. The biggest problem with OTEC is that it is relatively inefficient. The laws of physics say that any practical heat engine must operate at less than 100% efficiency; most operate well below, and OTEC plants, which use a relatively small temperature difference between their hot and cold fluids, have among the lowest efficiency of all, typically just a few percent. For that reason, OTEC plants have to pump huge amounts of water to produce even modest amounts of electricity, which causes two problems. First, it means a significant amount of electricity generated (typically about a third) has to be used to operate the system (pumping the water in and out). Second, it implies that OTEC plants have to be constructed on a relatively large scale, which makes them expensive investments. Large-scale onshore OTEC plants could have a considerable environmental impact on shorelines, which are often home to fragile, already threatened ecosystems such as mangroves and coral reefs. 2. Although OTEC plants are suitable only for tropical seas with relatively large temperature gradients, OTEC could theoretically operate in 29 different sovereign territories (including warmer, southern parts of the United States) and 66 developing nations. Temperate parts of the world that cannot operate OTEC most likely have alternative forms of ocean power they could exploit, including offshore wind turbines, tidal barrages, and wave power. 3. Although OTEC produces no chemical pollution, it involves human intervention in the temperature balance of the sea, which could have localized environmental impacts that would need to be assessed. One important (and often overlooked) impact of OTEC is that pumping cold water from the deep ocean to the surfaces releases carbon dioxide, the greenhouse gas currently most responsible for global warming. The amount released of carbon dioxide is only 10% of that produced by a fossil-fueled power plant.

239

Marine energy

11.4 Ocean tidal power Some of the oldest ocean energy technologies use tidal power. All coastal areas consistently experience two high and two low tides over a period of slightly greater than 24 h. For those tidal differences to be harnessed into electricity, the difference between high and low tides must be at 5 m or more. There are only about 40 sites on the earth with tidal ranges of this magnitude. Currently, there are no tidal power plants in the United States. However, conditions are good for tidal power generation in both the Pacific Northwest and the Atlantic Northeast regions of the country. For tidal power technologies: 1. Tide barrage is a barrage typically used to convert tidal energy into electricity by forcing the water through turbines, activating a generator, as shown in Figure 11.4. Gates and turbines are installed along the barrage. When the tides produce an adequate difference in the level of the water on opposite sides of the dam, the gates are opened. The water then flows through the turbines. The turbines turn an electric generator to produce electricity. 2. Tidal turbines look like wind turbines. They are arrayed underwater in rows, as in some wind farms. The turbines function best where coastal currents run between 3.6 and 4.9 knots (4 and 5.5 miles/h [mph]). In currents of that speed, a 15-m (49.2-ft)-diameter tidal turbine can generate as much energy as a 60-m (197-ft)-diameter wind turbine. Ideal (A) Turbine and generator

Tidal Barrage sea

Basin

Tide coming from the sea to the basin

(B)

Tide going out from the basin to the sea

Figure 11.4 Concept of operation for tidal barrage. (From NASA (2020)).

240

Thamer Mohamed

Figure 11.5 The tidal turbine. (From Getty Images (2020)).

locations for tidal turbine farms are close to shore in water depths of 20e30 m (65.5e98.5 ft) s shown in Figure 11.5 (U.S. Energy Information Administration, EIA, 2019). 3. A tidal fence is a type of tidal power system that has vertical axis turbines mounted in a fence or row placed on the sea bed, similar to tidal turbines. Water passing through the turbines generates electricity. As of the end of 2018, no tidal fence projects were operating. Tidal fences look like giant turnstiles across channels between small islands or across straits between the mainland and an island. The turnstiles spin via tidal currents typical of coastal waters. Some of these currents run at 5.6e9 mph and generate as much energy as winds of much higher velocity. Because seawater has a much higher density than air, ocean currents carry significantly more energy than air currents (wind). Ocean Energy’s device, called an OE Buoy (Figure 11.6), uses the oscillating water column principle. As waves enter a subsea chamber, they force air through a turbine on the surface, generating electricity. The technology has only one moving part, minimizing maintenance costs. However, there is a suite of rival designs racing to harness ocean energy.

11.5 Ocean wave power Wave energy (or wave power) is the transport and capture of energy by ocean surface waves (Figure 11.7). The energy captured is then employed for all different kinds of useful work, including electricity generation, water

Marine energy

241

Figure 11.6 The Ocean Energy buoy. (From Ocean Energy (2019)).

Figure 11.7 General view of wave power conversion. (From U.S. Department of Energy (2015)).

desalination, and pumping of water. Wave energy is also a type of renewable energy and is the largest estimated global resource form of ocean energy. Wave powererich areas of the world include the western coasts of Scotland, northern Canada, southern Africa, Australia, and the northeastern and northwestern coasts of the United States. In the Pacific Northwest alone, it is feasible that wave energy could produce 40e70 kW per meter (3.3 ft)

242

Thamer Mohamed

of western coastline. The West Coast of the United States is more than 1000 miles long. Wave energy can be converted into electricity through both offshore and onshore systems.

11.5.1 Offshore systems Offshore systems are situated in deep water, typically more than 40 m (131 ft) deep, as shown in Figure 11.8. Sophisticated mechanisms such as the Salter’s Duck use the bobbing motion of the waves to power a pump that creates electricity. Other offshore devices use hoses connected to floats that ride the waves. The rise and fall of the float stretch and relax the hose, which pressurizes the water. This in turn rotates a turbine. Specially built seagoing vessels can also capture the energy of offshore waves. These floating platforms create electricity by funneling waves through internal turbines and then back into the sea.

11.5.2 Onshore systems Built along shorelines, onshore wave power systems extract the energy in breaking waves. For onshore system technologies: 1. The oscillating water column consists of a partially submerged concrete or steel structure that has an opening to the sea below the waterline. It encloses a column of air above a column of water. As waves enter the air column, they cause the water column to rise and fall. This alternately

Figure 11.8 Offshore system. OTEC, ocean thermal energy conversion. (From MOE (2020)).

Marine energy

243

compresses and depressurizes the air column. As the wave retreats, the air is drawn back through the turbine as a result of the reduced air pressure on the ocean side of the turbine. 2. The tapered channel wave energy system consists of a tapered channel that feeds into a reservoir constructed on cliffs above sea level. Narrowing of the channel causes the waves to increase in height as they move toward the cliff face. The waves spill over the walls of the channel into the reservoir and the stored water is then fed through a turbine. 3. The pendulous wave power device consists of a rectangular box that is open to the sea at one end. A flap is hinged over the opening and the action of the waves causes the flap to swing back and forth. The motion powers a hydraulic pump and a generator.

11.6 Environmental and economic challenges Marine energy will assist in reducing carbon emissions worldwide. The large-scale development of marine energy projects will have uncertain environmental impacts, most of which have not been adequately evaluated. The ecological effects of the marine energy, such as wave, tidal, ocean current, and thermal gradient, mainly cause habitat and community changes. This necessitates renewable energy developers, regulators, scientists, engineers, and ocean stakeholders to work together to achieve the common dual objectives of clean renewable energy and a healthy marine environment (George and Andrew, 2015). In general, careful site selection is the key to keeping the environmental impacts of wave power systems to a minimum. Wave energy system planners can choose sites that preserve scenic shorefronts. They also can avoid areas where wave energy systems can significantly alter flow patterns of sediment on the ocean floor. Economically, wave power systems have a hard time competing with traditional power sources. However, the costs of producing wave energy are falling. Some European experts predict that wave power devices will find lucrative niche markets. Once built, they will have low operation and maintenance costs because the fuel they use (seawater) is free (U.S. Department of Energy [DOE], 2011a). An Irish wave energy developer is working with Cornwall’s Wave Hub with a view to deploying the scheme’s first device. Tidal power plants that dam estuaries can impede sea life migration, and silt buildups behind such facilities can affect local ecosystems. Tidal fences may also disturb sea life migration. Newly developed tidal turbines may prove ultimately to be the

244

Thamer Mohamed

least environmentally damaging of the tidal power technologies because they do not block migratory paths. It does not cost much to operate tidal power plants, but their construction costs are high and lengthen payback periods. As a result, the cost per kilowatt-hour of tidal power is not competitive with conventional fossil fuel power. OTEC power plants require substantial capital investment upfront. OTEC researchers believe private sector firms probably will be unwilling to make the enormous initial investment required to build large-scale plants until the price of fossil fuels dramatically increases or national governments provide financial incentives. Another factor hindering the commercialization of OTEC is that there are only a few hundred land-based sites in the tropics where DOW is close enough to shore to make OTEC plants feasible (U.S. DOE, 2011b).

References George, W.B., Andrew, B.G., 2015. Environmental and ecological effects of ocean renewable energy development: a current synthesis. Oceanography. https://doi.org/10.5670/ oceanog.2010.46. Getty Images, 2020. Tidal Power. https://www.gettyimages.com/photos/tidal-power? mediatype¼photography&phrase¼tidal%20power&sort¼mostpopular. (Accessed 12 April 2020). International Renewable Energy Agency, IRENA, 2019. Ocean Energy Data. https:// www.irena.org/ocean. (Accessed 10 April 2020). Kresala, 2018. Alternative Marine Energy Resources. Higher School of Engineers of the University of Navarra, Spain. https://kresalaenergia.wordpress.com/author/ juncal903087/. (Accessed 10 April 2020). Liu, C.K., 2018. Ocean thermal energy conversion and open ocean mariculture: the prospect of mainland-Taiwan Collaborative research and development. Sustain. Environ. Res. 28 (6), 257e472. Makai Ocean Engineering, MOE, 2020. Ocean Thermal Energy Conversion. https:// oceanexplorer.noaa.gov/edu/learning/player/lesson11/l11la1.html. (Accessed 12 April 2020). Matthew, J.H., John, G., Angus, V., Max, C., Stuart, B., Richard, B., Stephen, W., 2016. Report on World Energy Resources, Marine Energy. World Energy Council. https://doi.org/10.13140/RG.2.2.24836.73607. NASA, 2020. Huge Machine Harnesses the Tides. https://climatekids.nasa.gov/tidalenergy/. (Accessed 12 April 2020). Ocean Energy, 2019. What is OE Buoy. https://oceanenergy.ie/oe-buoy/. (Accessed 12 April 2020). U.S. Department of Energy, 2015. Wave Energy. https://openei.org/wiki/Wave_Energy. (Accessed 12 April 2020). U.S. Department of Energy (DOE) EERE, 2011a. Ocean Wave Power Accessed. http:// www.energysavers.gov/renewable_energy/ocean/index.cfm/mytopic ¼ 50009. (Accessed 12 December 2019). U.S. Department of Energy (DOE) EERE, 2011b. Ocean Thermal Energy Conversion. http://www.eere.energy.gov/basics/renewable_energy/ocean_thermal_ energy_conv. html. (Accessed 12 December 2019).

Marine energy

245

U.S. Energy Information Adminstration, EIA, 2019. Hydropower Explained Tidal Power. https://www.eia.gov/energyexplained/hydropower/tidal-power.php. (Accessed 12 April 2020). Woodford, C., 2020. Ocean Thermal Energy Conversion. The Free Online Science and Technology Book. https://www.explainthatstuff.com/how-otec-works.html. (Accessed 11 April 2020). Zhenya, L., 2016. Global Energy Interconnection. Elsevier B.V., Science Direct. ISBN 9780-12-804405-6.

Further reading Pinterest. Ocean Thermal Energy Conversion. https://www.pinterest.com/sinnedcity/ otec/. (Accessed 12 April 2020).

CHAPTER TWELVE

Geothermal energy Thamer Mohamed Department of Water Resources Engineering, College of Engineering, University of Baghdad, Baghdad, Iraq

Contents 12.1 Introduction 12.2 The history of geothermal energy 12.3 Geothermal heat pumps 12.4 Geothermal electricity 12.5 Environmental effects, benefits, and economic costs 12.6 The future of geothermal energy References Further reading

247 249 250 252 257 258 260 261

12.1 Introduction Geothermal energy is thermal energy generated and stored in the earth. Thermal energy is the energy that determines the temperature of matter. Earth’s geothermal energy originates from the original formation of the planet (20%) and from the radioactive decay of minerals (80%) (Turcotte and Schubert, 2002). The geothermal gradient, which is the difference in temperature between the core of the planet and its surface, drives the continuous conduction of thermal energy in the form of heat from the core to the surface (Figure 12.1). The heat that is used for geothermal energy can be stored deep within the earth, 4000 miles down to the earth’s core. At the core, temperatures may reach over 9000 F (5000 C). Heat conducts from the core to surrounding rock. Extreme high temperature and pressure cause some rock to melt, commonly known as magma. Magma convects upward because it is lighter than solid rock. This magma then heats rock and water in the crust, sometimes to 700 F (370 C) (Nemzer, 2012). From hot springs, geothermal energy has been used for bathing since Paleolithic times, and for space heating since ancient Roman times, but it is now better known for the generation of electricity. Worldwide, about

Distributed Renewable Energies for Off-Grid Communities, Volume 2 ISBN: 978-0-12-821605-7 https://doi.org/10.1016/B978-0-12-821605-7.00022-2

© 2021 Elsevier Inc. All rights reserved.

247

j

248

Thamer Mohamed

Figure 12.1 Temperatures in the earth. (From Our Energy (2012)).

Figure 12.2 Steam rising from the Nesjavellir Geothermal Power Station in Iceland. (From Gretar, Ivarsson, 2007).

10,715 MW of geothermal power is online in 24 countries. An additional 28 GW of direct geothermal heating capacity is installed for district heating, space heating, spas, industrial processes, desalination, and agricultural applications (Fridleifsson, 2008). Geothermal power is cost-effective, reliable, sustainable, and environmentally friendly (Glassley, 2010), but has historically it has been limited to areas near tectonic plate boundaries (Figure 12.2). Technological advances have expanded this range, especially for applications such as home heating. Geothermal wells release greenhouse gases trapped deep within the earth, but these emissions are much lower per energy unit than those of fossil fuels. Therefore, geothermal power has the potential to help mitigate global warming.

Geothermal energy

249

The earth’s geothermal resources are theoretically more than adequate to supply humanity’s energy needs, but only a small fraction may be profitably exploited. Drilling and exploration for deep resources is expensive. Forecasts for the future of geothermal power depend on assumptions about technology, energy prices, subsidies, and interest rates (Cothran et al., 2002).

12.2 The history of geothermal energy According to John (2020), geothermal energy from natural pools and hot springs has long been used for cooking, bathing, and warmth. There is evidence that Native Americans used geothermal energy for cooking as early as 10,000 years ago. In ancient times, baths heated by hot springs were used by the Greeks and Romans, and examples of geothermal space heating date back at least as far as the Roman city of Pompeii during the first century CE. Such uses of geothermal energy were initially limited to sites where hot water and steam were accessible. Although the world’s first district heating system was installed at Chaudes-Aigues, France, in the 14th century, it was not until the late 19th century that other cities, as well as industries, began to realize the economic potential of geothermal resources. Geothermal heat was delivered to the first residences in the United States in 1892, to Warm Springs Avenue in Boise, Idaho, and most of the city used geothermal heat by 1970. The largest and most famous geothermal district heating system is in Reykjavík, Iceland, where 99% of the city received geothermal water for space heating starting in the 1930s. Early industrial direct-use applications included the extraction of borate compounds from geothermal fluids at Larderello, Italy, during the early 19th century. The first geothermal electric power generation also took place in Larderello, with the development of an experimental plant in 1904. The first commercial use of that technology occurred there in 1913 with the construction of a plant that produced 250 kW. Geothermal power plants were commissioned in New Zealand starting in 1958 and at the Geysers in Northern California in 1960. The Italian and American plants were dry steam facilities, in which low-permeability reservoirs produced only steam. In New Zealand, however, high-temperature and high-pressure water emerges naturally as a mixture made of 80% superheated water and 20% steam. The steam coming directly from the ground is immediately used for power generation. It is sent to the power plant through pipes. In contrast, the superheated water from the ground is

250

Thamer Mohamed

separated from the mixture and flashed into steam. Most current geothermal plants are of this latter wet steam type. By 2015, more than 80 countries were using geothermal energy, either directly or in conjunction with geothermal heat pumps (GHPs); the leaders are China, Turkey, Iceland, Japan, Hungary, and the United States. The total worldwide installed capacity for direct use in 2015 was about 73,290 MW thermal (MWt) using about 163,273 GW-h/y (587,786 TJ/y), producing an annual utilization factor, the annual energy produced by the plant (in megawatt-hours) divided by the installed capacity of the plant (in megawatts) multiplied by 8760 hdof 28% in the heating mode. Geothermal energy was used to produce electricity in 24 countries in the early 21st century. The leaders were the United States, the Philippines, Indonesia, Mexico, New Zealand, and Italy. In 2016, the total worldwide installed capacity for electrical power generation was about 13,400 MW, producing about 75,000 GW-h/y for a usage factor of 71% (equivalent to 6220 full-load operating hours annually). Many geothermal fields have usage factors of around 95% (equivalent to 8322 full-load operating hours annually), the highest for any form of renewable energy. The waste fluid from the power plant is often used for lower-temperature applications, such as the bottom cycle in a binary-cycle plant, before it is injected back into the reservoir. Such cascaded uses can be found in the United States, Iceland, and Germany.

12.3 Geothermal heat pumps GHPs take advantage of relatively stable moderate temperature conditions that occur within the first 300 m (1000 ft) of the surface to heat buildings in the winter and cool them in the summer. In that part of the lithosphere, rocks and groundwater occur at temperatures between 5 and 30 C (41 and 86 F). At shallower depths, where most GHPs are found, such as within 6 m (about 20 ft) of the earth’s surface, the temperature of the ground maintains a near-constant temperature of 10e16 C (50e60 F). Consequently, that heat can be used to help warm buildings during the colder months of the year, when the air temperature falls below that of the ground. Similarly, during the warmer months of the year, warm air can be drawn from a building and circulated underground, where it loses much of its heat and is returned, as shown in Figure 12.3.

251

Geothermal energy

Winter Heang

heang system

heat transferred from pipe to ground during winter

Summer Heang

cooling system

heat transferred from pipe to ground during summer

Underground loops of pipe in contact with soil and rock

Figure 12.3 Residential heat pumps operating for winter heating and summer cooling. (From Encyclopedia Britannica, Inc., 2012).

A GHP system is composed of a heat exchanger (a loop of pipes buried in the ground) and a pump. The heat exchanger transfers heat energy between the ground and air at the surface by means of a fluid that circulates through the pipes; the fluid used is often water or a combination of water and antifreeze. During warmer months, heat from warm air is transferred to the heat exchanger and into the fluid. As it moves through the pipes, the heat is dispersed to the rocks, soil, and groundwater. The pump is reversed during the colder months. Heat energy stored in the relatively warm ground raises the temperature of the fluid. The fluid then transfers this energy to the heat pump, which warms the air inside the building. GHPs have several advantages over more conventional heating and airconditioning systems. They are efficient, using 25e50% less electricity than comparable conventional heating and cooling systems, and they produce less pollution. The reduction in energy use associated with GHPs can translate into as much as a 44% decrease in greenhouse gas emissions compared with air-source heat pumps (which transfer heat between indoor and outdoor air). In addition, compared with electric resistance heating systems (which convert electricity to heat) coupled with standard air-conditioning systems, GHPs can produce up to 72% less greenhouse gas emissions ( John, 2020).

252

Thamer Mohamed

12.4 Geothermal electricity Heat from earth’s geothermal energy heats water that has seeped into underground reservoirs. These reservoirs can be tapped for a variety of uses, depending on the temperature of the water. The energy from hightemperature reservoirs (225 to 600 F) can be used to produce electricity. Technologies in use include dry steam, flash steam, and binary cycle power plants. Geothermal electricity generation is currently used in 24 countries (Geothermal Energy Association, 2010), whereas geothermal heating is in use in 70 countries (Fridleifsson, 2008). Estimates of the electricity-generating potential of geothermal energy vary from 35 to 2000 GW. Current worldwide installed capacity is 10,715 MW (Geothermal Energy Association, 2010), with the largest capacity in the United States (3086 MW), the Philippines, and Indonesia. Geothermal power is considered to be sustainable because the heat extraction is small compared with the earth’s heat content (Rybach, 2007). The emission intensity of existing geothermal electric plants is on average 122 kg of CO2 per megawatt-hour of electricity, about oneeighth that of a conventional coal fire plant (Bertani and Thain, 2002). The International Geothermal Association (IGA) reported in 2010 that 10,715 MW of geothermal power in 24 countries were online and were expected to generate 67,246 GWh of electricity. This represented a 20% increase in online capacity since 2005. IGA projected growth to 18,500 MW by 2015, owing to projects currently under consideration, often in areas previously assumed to have few exploitable resources (Geothermal Energy Association, 2010). In 2010, the United States led the world in geothermal electricity production with 3086 MW of installed capacity from 77 power plants (Geothermal Energy Association, 2010). The largest group of geothermal power plants in the world is located at the Geysers, a geothermal field in California (Khan, 2007). The Philippines is the second highest producer, with 1904 MW of capacity online (Figure 12.4). Geothermal power makes up approximately 18% of the country’s electricity generation (Geothermal Energy Association, 2010). There are currently three types of geothermal power plants: 1. Dry steam plants use steam from underground wells to rotate a turbine, which activates a generator to produce electricity (Figure 12.5). There are only two known underground resources of steam in the United States: the Geysers in Northern California and Old Faithful in

253

Geothermal energy

Figure 12.4 Palinpinon Geothermal power plant in Sitio Nasuji, Barangay Puhagan, Valencia, Negros Oriental, Philippines. (From Mike Gonzalez, 2010). Dry steam power plant turbine

electrical generator

load

steam rising

cooled water

production well

injection well rock layers

© 2012 Encyclopaedia Britannica,Inc.

Figure 12.5 Dry steam geothermal power plant. (From Encyclopedia Britannica, Inc., 2012).

Yellowstone National Park. Because Yellowstone is protected from development, the power plants at the Geysers are the only dry steam plants in the country. 2. Flash steam plants, the most common type of geothermal power plant, use water at temperatures greater than 360 F. As this hot water flows up through wells in the ground, the decrease in pressure causes some of the water to boil into steam. The steam is then used to power a generator; any leftover water and condensed steam is returned to the reservoir (Figure 12.6).

254

Thamer Mohamed

Flash steam power plant

turbine

electrical generator load

flash tank steam

hot water

cooled water

injection well

production well

rock layers

© 2012 Encyclopaedia Britannica,Inc.

Figure 12.6 Flash steam geothermal power plant. (From Encyclopedia Britannica, Inc., 2012). Binary cycle power plant

turbine

electrical generator load

heat exchanger with working fluid

hot water

production well

cooled water

injection well rock layers

© 2012 Encyclopaedia Britannica,Inc.

Figure 12.7 Binary cycle geothermal power plant. (From Encyclopedia Britannica, Inc., 2012).

3. Binary cycle plants use the heat from lower-temperature reservoirs (225 to 360 F) to boil a working fluid, which is then vaporized in a heat exchanger and used to power a generator. The water, which never comes into direct contact with the working fluid, is then injected back into the ground to be reheated (Figure 12.7).

Geothermal energy

255

The choice of which design to use is determined by the resource. If the water comes out of the well as steam, it can be used directly, as in the first design. If it is hot water of a high enough temperature, a flash system can be used; otherwise it must go through a heat exchanger. Because there are more hot water resources than pure steam or high-temperature water sources, there is more growth potential in the heat exchanger design. Geothermal reservoirs of low to moderate temperature water (68 to 302 F [20 to 150 C]) provide direct heat for residential, industrial, and commercial use. This resource is widespread in the United States and is used to heat homes and offices, commercial greenhouses, fish farms, food processing facilities, gold mining operations, and a variety of other applications. In addition, spent fluids from geothermal electric plants can be subsequently used for direct-use applications in so-called cascaded operations. Direct use of geothermal energy in homes and commercial operations is much less expensive than using traditional fuels. Savings can be as much as 80% over fossil fuels. Direct use is also clean, producing only a small percentage of (and in many cases no) air pollutants emitted by burning fossil fuels. In the geothermal industry, low temperature means temperatures of 300 F (149 C) or less. Low-temperature geothermal resources are typically used in direct-use applications, such as district heating, greenhouses, fisheries, mineral recovery, and industrial process heating. However, some low-temperature resources can generate electricity using binary cycle electricity-generating technology. Approximately 70 countries made direct use of 270 PJ of geothermal heating in 2004. More than half went for space heating, and another third for heated pools. The remainder supported industrial and agricultural applications. Global installed capacity was 28 GW, but capacity factors tend to be low (30% on average) because heat is mostly needed in winter. These figures are dominated by 88 PJ of space heating extracted by an estimated 1.3 million GHPs with a total capacity of 15 GW. Heat pumps for home heating are the fastest-growing means of exploiting geothermal energy, with a global annual growth rate of 30% in energy production. Direct heating is far more efficient than electricity generation and places less demanding temperature requirements on the heat resource. Heat may come from cogeneration via a geothermal electrical plant or from smaller wells or heat exchangers buried in shallow ground. As a result, geothermal heating is economic at many more sites than geothermal electricity generation. Where natural hot springs are available, the heated water can be piped directly into radiators. If the ground is hot but dry, earth tubes or down-hole

256

Thamer Mohamed

heat exchangers can collect the heat. However, even in areas where the ground is colder than room temperature, heat can be extracted with a GHP more cost-effectively and cleanly than by conventional furnaces. These devices draw on much shallower and colder resources than traditional geothermal techniques, and they frequently combine a variety of functions, including air-conditioning, seasonal energy storage, solar energy collection, and electric heating. GHPs can be used for space heating essentially anywhere. Geothermal heat supports many applications. District heating applications use networks of piped hot water to heat many buildings across entire communities. In Reykjavík, Iceland, spent water from the district heating system is piped below pavement and sidewalks to melt snow. Geothermal desalination has been demonstrated. The largest geothermal system currently in operation is a steam-driven plant in an area called the Geysers, north of San Francisco, California (Figure 12.8). Despite the name, there are actually no geysers there, and the heat that is used for energy is all steam, not hot water. Although the area was known for its hot springs as far back as the mid-1800s, the first well for power production was drilled in 1924. Deeper wells were drilled in the 1950s, but real development did not occur until the 1970s and 1980s. By 1990, 26 power plants had been built, for a capacity of more than 2000 MW (Union of Concerned Scientists, 2009).

Figure 12.8 Geothermal power plant at the Geysers near Santa Rosa, California. (From Julie Donnelly-Nolan, U.S. Geological Survey, 2009).

Geothermal energy

257

12.5 Environmental effects, benefits, and economic costs Geothermal energy is a renewable energy source that causes little damage to the environment. Geothermal power plants do not burn fuels to generate electricity as do fossil fuel plants. Geothermal power plants release less than 1e4% of the amount of CO2 emitted by coal plants. Geothermal power plants, on the other hand, emit only about 1e3% of the sulfur compounds that coal and oil-fired power plants do. Welldesigned binary cycle power plants have no emissions at all. The environmental effects of geothermal development and power generation include the changes in land use associated with exploration and plant construction, noise and sight pollution, the discharge of water and gases, the production of foul odors, and soil subsidence. Geothermal steam and hot water contain naturally occurring traces of hydrogen sulfide (a gas that smells like rotten eggs) and other gases and chemicals that can be harmful in high concentrations. However, most of those effects, can be mitigated with current technology so that geothermal uses have no more than a minimal impact on the environment. To mitigate the environmental effects, geothermal power plants use scrubber systems to clean the air of hydrogen sulfide and the other gases. Sometimes the gases are converted into marketable products, such as liquid fertilizer. Newer geothermal power plants can even inject these gases back into the geothermal wells. In addition, GHPs have a minimal effect on the environment because they use shallow geothermal resources within 100 m of the surface. GHPs cause only small temperature changes to the groundwater or rocks and soil in the ground. In closed-loop systems, the ground temperature around the vertical boreholes is slightly increased or decreased; the direction of the temperature change is governed by whether the system is dominated by heating (which would be the case in colder regions) or cooling (which would be the case in warmer regions). With balanced heating and cooling loads, the ground temperatures will remain stable. Likewise, open-loop systems using groundwater or lake water would have little effect on temperature, especially in regions characterized by high groundwater flows. Comparing the benefits of geothermal energy with other renewable energy sources, the main advantage of geothermal energy is that its base load is available 24 h/d, 7 d/wk, whereas solar and wind are available only about one-third of the time. Also, geothermal power plants are compatible with many environments. They have been built in deserts, in

258

Thamer Mohamed

the middle of crops, and in mountain forests. Geothermal development is often allowed in any country public land because it does not harm the environment significantly. Before permission is granted, however, studies must be made to determine what effect geothermal development may have on the environment. In addition, the cost of geothermal energy varies between 5 and 10 cents per kilowatt-hour, which can be competitive with other energy sources, such as coal. The main disadvantage of geothermal energy development is the high initial investment cost in constructing the facilities and infrastructure and the high risk of proving the resources. (Geothermal resources in low-permeability rocks are often found, and exploration activities often drill dry holes, or holes that produce steam in amounts too low to be exploited economically.) However, once the resource is proven, the annual cost of fuel (that is, hot water and steam) is low and tends not to escalate in price ( John, 2020).

12.6 The future of geothermal energy Internationally, geothermal energy has a significant potential role in a clean and sustainable energy system. It is one of the few renewable energy technologies that can supply continuous, base load power. In addition, unlike coal and nuclear plants, binary geothermal plants can be used as a flexible source of energy to balance the variable supply of renewable resources such as wind and solar. Figure 12.9 shows the installed geothermal energy in the world.

Figure 12.9 Installed capacity for geothermal energy in the world. (From IRENA, 2019).

Geothermal energy

259

According to IRENA (2019), there was an increase by about 40% in the installed capacity of the geothermal energy plants around the world between 2010 and 2019. This increase can be related to the competitive costs for electricity from geothermal facilities in addition to the future for the direct use of geothermal resources as a heating source in housing and commercial buildings worldwide. However, to tap into the full potential of geothermal energy, two emerging technologies require further development (UCS, 2014): 1. Enhanced geothermal systems (EGS): Geothermal heat occurs everywhere under the surface of the earth, but the conditions that make water circulate to the surface are found in less than 10% of the earth’s land area. An approach to capturing heat in dry areas is known as EGS or hot dry rock. The hot rock reservoirs typically found at greater depths below the surface compared with conventional sources are first broken up by pumping high-pressure water through them. The plants then pump more water through the broken hot rocks, where it heats up, returns to the surface as steam, and powers turbines to generate electricity. The water is then returned to the reservoir through injection wells to complete the circulation loop. Plants that use a closed-loop binary cycle release no fluids or heat-trapping emissions other than water vapor, which may be used for cooling. For example, it is estimated that using EGS technology, 100 GW of electricity can be produced in the United States by 2050 (UCS, 2014). One cause for careful consideration with EGS is the possibility of induced seismic activity that might occur from hot dry rock drilling and development. This risk is similar to that associated with hydraulic fracturing, an increasingly used method of oil and gas drilling, and with carbon dioxide capture and storage in deep saline aquifers. Although it is a potentially serious concern, the risk for an induced EGS-related seismic event that can be felt by the surrounding population or that might cause significant damage currently appears to be low when projects are located an appropriate distance away from major fault lines and properly monitored. Appropriate site selection, assessment, and monitoring of rock fracturing and seismic activity during and after construction, and open, transparent communication with local communities are also critical. 2. Coproduction of geothermal electricity in oil and gas wells: Lowtemperature geothermal energy is derived from geothermal fluid found in the ground at temperatures of 150 C (300 F) or less. These resources are typically used in direct-use applications such as heating buildings, but

260

Thamer Mohamed

they can also be employed to produce electricity through binary cycle geothermal processes. Oil and gas fields already under production represent a large potential source of this type of geothermal energy. In many existing oil and gas reservoirs, a significant amount of hightemperature water or suitable high-pressure conditions are present, which could allow for the coproduction of geothermal electricity along with the extraction of oil and gas resources. In some cases, exploiting these geothermal resources could even enhance the extraction of the oil and gas. It is estimated that by using this method, 10% of baseload electricity in the United States could be supplied by 2050 (UCS, 2014).

References Bertani, R., Thain, I., July 2002. Geothermal Power Generating Plant CO2 Emission Survey, vol. 49. IGA News (International Geothermal Association), pp. 1e3. Retrieved from 2009-05-13. Cothran, H., March 1, 2002. Energy Alternatives. Greenhaven Press, p. 220. Encyclopaedia Britannica, Inc, 2012. Residential Heat Pump Operation for Summer Cooling and Winter Heating. https://www.britannica.com/science/geothermal-energy. (Accessed 13 April 2020). Fridleifsson, I.B., Bertani, R., Huenges, E., Lund, J.W. Geothermal Energy Association, May 2010. Geothermal Energy: International Market Update, pp. 4e6,7. Glassley, W.E., 2010. Geothermal Energy: Renewable Energy and the Environment. CRC Press. John, W.L., April 30, 2020. Geothermal Energy. Research Britannica. Geothermal Energy, Encyclopædia Britannica. Encyclopædia Britannica, Inc. https://www.britannica.com/ science/geothermal-energy. (Accessed 16 April 2020). Khan, M.A., 2007. The Geysers Geothermal Field, an Injection Success Story (pdf). Nemzer, J., 2012. Geothermal Heating and Cooling. Cambridge University Press, Cambridge, England, UK, pp. 136e137. ISBN 978-0-521-66624-4. Our Energy, 2012. Geothermal Energy Facts. Our Energy. Available from: http://www. our-energy.com/energy_facts/geothermal_energy_facts.html. Rybach, L., September 2007. Geothermal sustainability, 3. In: Geo-Heat Centre Quarterly Bulletin, vol. 28. Oregon Institute of Technology, Klamath Falls, Oregon, pp. 2e7. ISSN 0276-1084, retrieved 2009-05-09. Turcotte, D.L., Schubert, G., 2002. “4”, Geodynamics, second ed. Cambridge University Press, Cambridge, England, UK, pp. 136e137. ISBN 978-0-521-66624-4. Union of Concerned Scientist, UCS, 2014. Report on How Geothermal Energy Works. https://www.ucsusa.org/resources/how-geothermal-energy-works. (Accessed 13 April 2020). Union of Concerned Scientists, 2009. How Geothermal Energy Works. Clean Energy. Available from: http://www.ucsusa.org/clean_energy/technology_and_ impacts/ energy_technologies/how-geothermal-energy-works.html.

Geothermal energy

261

Further reading Annual Forum of the Groundwater Protection Council. (Accessed 25 January 2010). Ragnarsson, A., Rybach, L., February 11, 2008. In: Hohmeyer, O., Trittin, T. (Eds.), The Possible Role and Contribution of Geothermal Energy to the Mitigation of Climate Change. Luebeck, Germany, pp. 59e80. Retrieved 2009-04-06. U.S. Department of Energy (DOE), 2012. Geothermal Technologies Program. Energy Efficiency & Renewable Energy. Available from: http://www1.eere.energy. gov/ geothermal/powerplants.html.

CHAPTER THIRTEEN

Energy storage, smart grids, and electric vehicles N. El Bassam International Research Centre for Renewable Energy, IFFED.org, Germany

Contents 13.1 Energy storage 13.1.1 Batteries and hydrogen technology: keys for a clean energy future 13.1.2 Storage methods 13.1.3 Technologies for upregulation and downregulation 13.2 Smart grids 13.2.1 Definition and importance 13.2.2 Smart meters 13.2.3 United States version 13.2.3.1 Challenges

264 265 271 275 276 276 278 278 279

13.2.4 European strategies 13.2.5 Korean version 13.3 Electric vehicles 13.3.1 Current developments 13.3.2 Types of electric vehicles 13.3.3 Global battery electric vehicle and plug-in hybrid electric vehicle sales 13.3.4 Types of electric vehicles 13.3.4.1 Battery electric vehicles 13.3.4.2 Plug-in hybrid electric vehicles 13.3.4.3 Hybrid electric vehicles

280 282 284 284 287 287 290 290 291 292

13.4 Future developments References Further reading

292 294 295

The dynamics of the world are changing, and people prefer low-cost and reliable power throughout the day. The addition of renewable energy to the existing system is one way to provide reliable and cheap electricity. The bottleneck in transmission lines, continuous contamination of the environment owing to heavy reliance on fossil fuels, and the highly fluctuating cost of fossil fuel are reasons for the widespread use of renewable energy Distributed Renewable Energies for Off-Grid Communities, Volume 2 ISBN: 978-0-12-821605-7 https://doi.org/10.1016/B978-0-12-821605-7.00008-8

© 2021 Elsevier Inc. All rights reserved.

263

j

264

N. El Bassam

technology. Energy storage technologies are a need of the time and range from low-capacity mobile storage batteries to high-capacity batteries connected to intermittent renewable energy sources (RES). The selection of different battery types, each of which has distinguished characteristics regarding power and energy, depends on the nature of the power required and delivered. This chapter presents a detailed review of battery energy storage technologies pertaining to the latest technologies, benefits, sizing considerations, efficiency, cost, and recycling. An in-depth analysis in terms of advantages and limitations of the different types of batteries is discussed and compared. In terms of microgrid applications, the economic benefits of battery sizing using optimization and probabilistic methods provide a potential solution during the design stage by considering various factors affecting the sizing of the battery, such as the degradation rate, reliability, and battery placement. Energy storage technologies will have an important position in combining RES in modern electrical power systems and the smart grid. Storage technologies could provide more balancing and flexibility to the power system, providing incorporation of intermittent RES to the smart grid. Energy storage technologies have a critical function of providing ancillary services in the power generation source for the smart grid. This chapter gives a short overview of current energy storage technologies and their available applications as well as the opportunities and challenges the power systems faces for successful integration of RES into the smart grid.

13.1 Energy storage There is a continuous global need for more energy; at the same time, it has to be cleaner than energy produced from traditional generation technologies. This need has facilitated the increasing penetration of distributed generation (DG) technologies, primarily of RES. The extensive use of such energy sources in today’s electricity networks can indisputably minimize the threat of global warming and climate change. However, the power output of these energy sources is not as reliable and easy to adjust to changing demand cycles as the output from traditional power sources. This disadvantage can be effectively overcome only by storing excess power produced by DG-RES. Therefore, for these new sources to become completely reliable as primary sources of energy, energy storage is a crucial factor (https:// www.sciencedirect.com/science/article/abs/pii/S1364032108001664).

265

Energy storage, smart grids, and electric vehicles

13.1.1 Batteries and hydrogen technology: keys for a clean energy future Energy efficiency and renewable energy such as wind and solar photovoltaics (PV), the cornerstones of any clean energy transition, are good places to start. Those industries employ millions of people across their value chains and offer environmentally sustainable ways to create jobs and help revitalize the global economy (Figure 13.1). However, more than just renewables and efficiency will be required to put the world on track to meet climate goals and other sustainability objectives. IEA analysis has repeatedly shown that a broad portfolio of clean energy technologies will be needed to decarbonize all parts of the economy. Batteries and hydrogen-producing electrolyzers stand out as two important technologies owing to their ability to convert electricity into chemical energy, and vice versa. This is why they also deserve a place in any economic stimulus packages being discussed today. Batteries and electrolyzers are small, modular technologies that are potentially well-suited for mass manufacturing. Cost reductions such as those experienced through the large-scale production of solar PV are not inconceivable and in fact are under way. Progress in battery technology is more advanced than that of electrolyzers. The cost of lithium-ion batteries in particular has decreased owing to higher production volumes. The

Methane (Power-to-Gas)

1 Year

Compressed air Storage

1 Day

Hydrogen (Power-to-Gas)

TW h 10

0

TW

h

h 10

TW 1

h 10

0

G

W

G W h

h W G 1

M 0 10

10

h W

h W M 10

M

W

h

h 1

10

0

kW

kW

h kW

h

Flywheel

1 Minute

1

Pumped Storage

Batteries

1 Hour

10

Discharge Time

1 Month

Storage Capacity

Figure 13.1 Available storage technologies, their capacity, and their discharge time. (From D. Williams 1978, own work, CC BY-SA https://commons.wikimedia.org/w/index. php?curid¼68489756).

266

N. El Bassam

scale-up of electrolyzer manufacturing, on the other hand, is at an earlier stage. However, that makes its scope for significant near-term cost reductions even larger. Batteries and electrolyzers apply the same scientific principles of electrochemistry, meaning that they share several components such as electrolyte and membrane materials, as well as key manufacturing processes. The future development of electrolyzers therefore stands to benefit from the experience of manufacturing batteries. Knowledge acquired from batteries should spill over into the scaling-up of electrolyzer production, enabling faster cost reductions. The price of lithium-ion batteries, the key technology for electrifying transport, has declined sharply in recent years after having they were developed for widespread use in consumer electronics. Governments in many countries have adopted policies encouraging increased the deployment of electric cars, further accelerating the decline in battery prices. At the same time, the power sector offers growing opportunities for the use of batteries to support the integration of variable renewables such as wind and solar PV into electricity systems. As such, lithium-ion batteries are now a technology opportunity for the wider energy sector, well beyond just transport. Electrolyzers, devices that split water into hydrogen and oxygen using electrical energy, are a way to produce clean hydrogen from low-carbon electricity. Clean hydrogen and hydrogen-derived fuels could be vital for decarbonizing sectors where emissions are proving particularly hard to reduce, such as shipping, aviation, long-haul trucks, and the iron and steel or chemical industries. These are areas where other clean energy technologies cannot easily be deployed. However, natural gas and coal are currently the primary sources for almost all of the approximately 70 million tons of hydrogen produced each year to make fertilizers and for use in oil refineries. This means that the production and use of hydrogen are associated with more than 800 million tons of carbon dioxide (CO2) emissions today, a staggering amount equivalent to the emissions of the United Kingdom and Indonesia combined. The world’s capacity to make battery cells has expanded rapidly. Today, manufacturing operations globally can produce around 320 GW-hours (GWh) of batteries per year for use in electric cars. This is well above the approximately 100 GWh of batteries required for the 2.1 million electric cars that were sold in 2019. Having sufficient capacity available for battery manufacturing is critical for the continued electrification of road transport. Global production capacity is unevenly distributed. China is the world leader, accounting for around 70% of global capacity, followed by the United States (13%), Korea (7%), Europe (4%), and Japan (3%). The outbreak of the novel coronavirus-19

Energy storage, smart grids, and electric vehicles

267

(Covid-19) epidemic has affected all of China’s battery production hubs, which are located in the provinces of Hubei, Hunan, and Guangdong. Manufacturing has resumed gradually owing to the time it takes to restore the supply chain and return employees to work. There is a need for manufacturing capacity to grow further. Assuming that the global auto industry’s announced targets for electric vehicle (EV) production are met despite the Covid-19 crisis, around 1000 GWh of battery manufacturing capacity would be needed in 2025. This output would require the equivalent of 50 plants, each on the scale of a Tesla Gigafactory. Longer-term targets set by governments around the world, as reflected in the Stated Policies Scenario of the IEA’s World Energy Outlook, could require global annual battery production to reach around 1500 GWh by 2030 for all EVs combined (including cars, buses, and so forth). Moreover, about twice as much production would be needed in 2030 to supply the amount of batteries envisaged in the IEA’s Sustainable Development Scenario, which provides a pathway to meeting long-term sustainability goals. Although such figures are ambitious, they are achievable. Battery manufacturing capacity targets for 2030 announced by companies led by CATL, LG Chem, BYD, Northvolt, and Panasonic stack up to around 2100 GWh/y. Nevertheless, time is of the essence, because building a large-scale battery factory can take anywhere from 2 to 5 years, depending on the country. Electrolyzer production is still in the early stages. Europe, the world leader, has a manufacturing capacity of 1.2 GW/y, enough capacity in theory to power more than half a million fuel cell passenger cars with hydrogen from water. Production capacity is expanding rapidly. The world’s largest electrolyzer plant, which is under construction by the UK’s ITM Power, is expected to produce 1 GW/y. In addition, NEL Hydrogen of Norway announced plans to build a plant with a production capacity of 360 MW/ y and the potential to expand to triple that amount. The deployment of electrolyzers has also picked up in terms of both the number and size of the projects. A decade ago, most projects were smaller than 0.2 MW. More recently, several projects were in the range of 1e5 MW, with the largest at 6 MW. In Japan, a 10-MW project has begun operating, and a 20-MW project in Canada is under construction. Larger projects in the hundreds of megawatts have been announced. As a result, the upcoming years could set new records, with announced projects bringing the global installation of electrolyzer capacity from

268

N. El Bassam

170 MW in 2019 to 730 MW in 2021. To ensure that such momentum continues after the Covid-19 crisis, it will be important for governments to reassure investors about their continued commitment to hydrogen (https://www.iea.org/articles/batteries-and-hydrogen-technology-keysfor-a-clean-energy-future). A battery storage power station uses a group of batteries to store electrical energy. As of 2019, the maximum power of battery storage power plants was an order of magnitude less than pumped storage power plants, the most common form of grid energy storage. In terms of storage capacity, the largest battery power plants are about two orders of magnitude less than pumped hydro-plants (Figure 13.2 and Table 13.1). Battery storage power plants are used for short-term peak power and ancillary services, such as providing a frequency-response reserve to minimize the chance of power outages. Hydrogen storage is a term used for any of several methods used to store hydrogen for later use. These methods include mechanical approaches such as high pressure and low temperature, or chemical compounds that release H2 upon demand. Although large amounts of hydrogen are produced, it is mostly consumed at the site of production, notably for the synthesis of ammonia. Interest in hydrogen storage is driven by the idea that it could be a medium for storing energy (e.g., to compensate for intermittent energy sources). The overarching challenge is the low boiling point of H2: it boils around 20.268 K (252.882 C or 423.188 F). Achieving such low temperatures requires significant energy (Figure 13.3).

Figure 13.2 A battery storage power station in Norwalk, California. (From https://en. wikipedia.org/wiki/Battery_storage_power_station#/media/File:World%E2%80%99s_1st_ Low-Emission_Hybrid_Battery_Storage,_Gas_Turbine_Peaker_System.jpg).

Largest grid batteries Name

Commissioning date

Energy (MWh) Power (MW) Duration (h)

Type

Country

Buzen Substation Rokkasho, Aomori Hornsdale Power Reserve Escondido Substation Pomona Substation Mira Loma Substation Tesla solar plant Stocking Pelham facility Jardelund Minamis oma Substation

Mar. 3, 2016 May 2008 Dec. 1, 2017

300 245 129

50 34 100

6 7

Sodium-sulfur Japan Sodium-sulfur Japan Lithium-ion Australia

Feb. 24, 2017 Jan. 2017 Jan. 30, 2017 Mar. 8, 2017 Jul. 2018 Jun. 2018 Feb. 2016

120 80 80 52 50 50 40

30 20 20 13 50 48 40

4 4 4 5 1 1 1

Lithium-ion Lithium-ion Lithium-ion Lithium-ion Lithium-ion Lithium-ion Lithium-ion

United States United States United States United States United Kingdom Germany Japan

Energy storage, smart grids, and electric vehicles

Table 13.1 The largest battery stations.

Under construction Name

Minety power storage project

Expected Energy (MWh) Power (MW) Duration (hours) Type commissioning date

Country

Planned Nov. 2021 Planned Q4 2020

United States United Kingdom

900 100

409 100

2.25

Lithium-ion Lithium-ion

From https://en.wikipedia.org/wiki/Battery_storage_power_station.

269

270

N. El Bassam

Figure 13.3 Frames Group, Hydrogenious LOHC Technologies, and MAN Energy Solutions have entered a partnership to store hydrogen on a large scale. (From https://www. thechemicalengineer.com/news/partnership-aims-to-build-large-scale-hydrogen-storagesystems/).

https://commons.wikimedia.org/w/index.php?curid¼20074457. Energy storage is accomplished by devices or physical media that store some form of energy to perform some useful operation at a later time. A device that stores energy is sometimes called an accumulator. All forms of energy are either potential energy (e.g., chemical, gravitational, electrical energy) or kinetic energy (e.g., thermal energy) (Wagner, 2007). The general method and specific techniques for storing energy are derived from some primary source in a form convenient for use at a later time when a specific energy demand is to be met, often in a different location. In the past, energy storage on a large scale was limited to the storage of fuels. Now, applications such as hydroelectric dams store energy in a reservoir (gravitational energy), or ice storage tanks store ice (thermal energy) at night to meet peak demand for cooling. On a smaller scale, electric energy is stored in batteries (chemical energy) that power automobile starters and a great variety of portable appliances. In the future, energy storage in many forms is expected to have an increasingly important role in shifting patterns of energy consumption away from scarce to more abundant and renewable primary resources. An example of growing importance is the storage of electric energy generated during the day by solar or wind energy or other renewable power plants to meet peak electric loads during daytime periods. This is achieved by pumped hydroelectric storage, which involves pumping water from a lower to a higher reservoir and reversing this process at night, with the pump then being used as a turbine and the motor as a generator.

Energy storage, smart grids, and electric vehicles

271

Off-peak electric energy can also be converted into mechanical energy by pumping air into a suitable cavern, where it is stored at pressures up to 80 atm (8 MPa). Turbines and generators can then be driven by the air when it is heated and expanded (The Free Dictionary, 2002).

13.1.2 Storage methods Energy storage methods can be generally categorized as: • Chemical • Hydrogen • Biofuels • Liquid nitrogen • Oxyhydrogen • Hydrogen peroxide • Biological • Starch • Glycogen • Electrochemical • Batteries • Flow batteries • Fuel cells • Electrical • Capacitor • Supercapacitor • Superconducting magnetic energy storage • Mechanical • Compressed air energy storage (CAES) • Flywheel energy storage • Hydraulic accumulator • Hydroelectric energy storage • Spring • Gravitational potential energy (device) • Thermal • Ice storage • Molten salt • Cryogenic liquid air or nitrogen • Seasonal thermal store • Solar pond • Hot bricks

272

N. El Bassam

• Graphite accumulator very high temperature • Steam accumulator • Fireless locomotive • Eutectic system Electrolysis has been around for many decades and is widely used to produce oxygen and hydrogen in the chemical and paper industries, in hospitals, and for welding. For energy storage, hydrogen is still in the early stage of development. Initial costs are high owing to the high pressure and diffusion of hydrogen, and conventional gas storage equipment is not suitable. Losses of conversion in the process from electricity back to electricity may be 65e80% accumulated by losses in rectifier, electrolyzer, compression, transmission, and the fuel cell (QuantumSphere Inc., 2006). Several commercially viable energy storage systems are being developed for hybrid EV (HEVs) on the market. The types of devices that hold the most promise for solving energy storage problems are batteries, flywheels, and ultracapacitors. As shown in Figure 14.2, both gasoline and hydrogen have a higher specific energy than the rest of these electrical storage devices (Fuel Cells, 2000, 2008). An advantage of HEVs is that they can use the high specific energy of liquid or gaseous fuels to provide vehicles with long-range capabilities. Conversely, the HEV can use the high specific power of electrical energy storage to provide peak power requirements. Batteries for the storage of electricity are widely used in many applications. For electric cars, a new generation of lithium batteries is being developed in many industrialized countries; they are expected to be gradually available for large-scale storage as well. Another possible technology is ultracapacitors. These devices work by accumulating and separating unlike charges. Their promise lies in the fact that they have no moving parts and that the number of times they can be cycled through their chargeedischarge cycle is high. The energy density of supercapacitors is 100 times higher than that of normal capacitors and the power density is 10 times higher than that of normal batteries, which enables their use in portable electronics and EVs and for the storage of energy generated from renewable sources such as wind and solar power (Wagner, 2008) (Figure 13.4). Electrochemical devices called fuel cells were invented about the same time as the battery in the 19th century. However, for many reasons, fuel cells were not well-developed until the advent of manned spaceflight (such as the Gemini program in the United States), when lightweight, nonthermal (and

Energy storage, smart grids, and electric vehicles

273

Figure 13.4 Flywheel module, National Aeronautics and Space Administration. (From the NASA Aerospace Flywheel Technology Program).

therefore efficient) sources of electricity were required in spacecraft. Fuel cell development has increased owing to an attempt to increase the conversion efficiency of chemical energy stored in hydrocarbon or hydrogen fuels into electricity (Wagner, 2007). Several other technologies have also been investigated: compressed air storage that can be pumped into underground caverns and abandoned mines (Wild, 2010), and a method used at the Solar Project and the Solar Tres Power Tower, which uses molten salt to store solar power and then dispatch that power as needed. The system pumps molten salt through a tower heated by the sun’s rays. Insulated containers store the hot salt solution; when needed, water is used to create steam that is fed to turbines to generate electricity. It can be used alone or combined with wind energy in utility-size installations of 50 MW or bigger, as demonstrated in southern Spain and the United States. With operational temperatures of up to 400 C, the storage medium can produce steam for conventional steam turbines combined with the production of electricity. The process heat can be distributed by a district heating network for heating and for cooling by absorption chillers (NREL, 2011). CAES is a way to store energy generated at one time for use at another time; it has been in operation for several years in the United States and Germany. Off-peak (low-cost) electrical power compresses air into an underground air-storage vessel (Figure 14.4) and later the air feeds a gas-fired turbine generator complex to generate electricity during on-peak (highprice) times (Wild, 2010).

274

N. El Bassam

Excess fluctuating electricity is used to compress atmospheric air into deep underground caverns of a type similar to natural gas storage. During consumption, the process is reversed and the air drives a conventional type turbine that, instead of natural gas or steam, uses compressed air connected to a generator. During compression, heat is produced, whereas the reversal process occurs with decompression and the air expands so that the system can deliver cooled air. The electric efficiency is around 50%; the overall efficiency can be improved if the heating and cooling potential is employed. A similar concept uses wind-powered air compressors (Pockley, 2008). Water pumped storages are installed in many countries to compensate for fluctuations in demand for power (Figure 14.5). Pump storages have dual purposes. A pumped-storage plant is designed with two reservoirs: upper and lower. Like every other hydroelectric plant, a pumped-storage plant generates electricity by allowing water to fall through a turbine generator. However, unlike conventional hydroelectric plants, once the pumpedstorage plant generates electricity, it can then pump that water from its lower reservoir back to the upper reservoir. This is done during the off-peak hours, using electricity from another source to run the plant’s pumps, in effect storing that off-peak electricity (Duke Energy, 2012). Their general application is limited by the topography; in Europe, most potential sites for pump storage have already been developed. Other storage solutions can be mentioned as well. Molten salt is employed for concentrated solar power storage. It can be used alone or combined with wind energy in utility-size installations of 50 MW or larger, as demonstrated in southern Spain and the United States. With operational temperatures of up to 400 C, the storage medium can produce steam for conventional steam turbines combined with the production of electricity. The process heat can be distributed by a district heating network for heating and cooling by absorption chillers (Mancini, 2006). In Denmark, several hundred hot water storage tanks are installed at local CHP; sizes range from 10 m3 up to 30,000 m3. The criteria for dimensions often cover the CHP station’s need for supply to the district heating network during the low peak period during weekends. Energy storage has critical roles in securing our energy future (Figure 13.5): • serving as an electricity reserve, much like the national petroleum reserve; • stabilizing electricity markets;

275

Energy storage, smart grids, and electric vehicles

Pumped-Storage Plant

Switchyard

Visitors Center

Reservoir Intake

Elevator

Main Access Tunnel

Discharge

Surge Chamber

Power Plant Chamber Breakers Transformer Vault

Figure 13.5 Conceptual representation of compressed-air energy storage concept. (From Tennessee Valley Authority (TVA) (2004). http://www.tva.gov/power/pumpstorart. htm).

• stabilizing the transmission and distribution grid; • enabling more efficient use of existing generation assets; and • making renewable energy economically viable (Maegaard, 2011).

13.1.3 Technologies for upregulation and downregulation Many RES (most notably solar and wind) produce intermittent power. Wherever intermittent power sources reach high levels of grid penetration, energy storage becomes an option to provide reliable energy supplies. Other options include a recourse to peaking power plants, methane storage (excess renewable electricity to hydrogen via electrolysis, combining with CO2 [low to neutral CO2 system] to produce methane [synthetic natural gas Sabatier process] with stockage in the natural gas network), and smart grids with advanced energy demand management. The latter involves bringing prices to devices (i.e., making electrical equipment and appliances able to adjust their operation to seek the lowest spot price of electricity). On a grid with a high penetration of renewables, low spot prices would correspond to times of high availability of wind and/or sunshine. For the economic evaluation of large-scale applications, such as pumped hydro-storage and compressed air, potential benefits are the avoidance of wind curtailment and grid congestion avoidance, price arbitrage, and carbon-free energy delivery.

276

N. El Bassam

Solar and wind alone cannot ensure a continuous supply of electricity. In a future supply scenario dominated by fluctuating energy forms completely based on renewables, three typical situations may occur to meet the actual demand for electricity: Wind power and PV power are the cornerstones of future renewable integrated energy supply structures. However, they fluctuate, which required consumers to adapt and to need backup from other supply solutions or storage. A number of storage solutions are available. They are of a different character regarding technology, medium, and cost. Flexibility and response time are important requirements that the various types of energy storage will meet differently, to match well with the integrated supply of power, heat, and cooling. A storage solution that converts residual power from wind and solar may be chemical, gravity, heat, compression, and so on, and may be suitable for either conversion back to electricity or heating/cooling. The need for storage may be for seconds, minutes, hours, or a few days. Seasonal storage is well-developed for hot water, but it is rare for large-scale electricity storage (Maegaard, 2011) (Figure 13.6).

13.2 Smart grids 13.2.1 Definition and importance Smart grids are considered a key requirement for the transition to renewables. Smart grids use modern communication technology to combine different elements of the energy system, such as generation and demand,

Figure 13.6 Storage of electricity. Est., Estimated. (From Maegaard, Power (2004)).

Energy storage, smart grids, and electric vehicles

277

thus ensuring a balance between sides. They help feed renewable energy into the grid more easily and allow for the grid’s capacity to be used optimally (https://www.bmwi-energiewende.de/EWD/Redaktion/EN/Newsletter/ 2019/05/Meldung/direkt-account.html). A smart grid is a digitally enabled electrical grid that gathers, distributes, and acts on information about the behavior of all participants (suppliers and consumers) to improve the efficiency, importance, reliability, economics, and sustainability of electricity services (U.S. DOE, 2012). The term smart grid is an umbrella term that covers the modernization of both the transmission and distribution grids. Modernization is directed at a disparate set of goals including facilitating greater competition among providers, enabling the greater use of variable energy sources, establishing the automation and monitoring capabilities needed for bulk transmission at crosscontinent distances, and enabling the use of market forces to drive energy conservation. Another element of the fault tolerance of traditional and smart grids is decentralized power generation. DG allows individual consumers to generate power on-site, using whatever generation method they find appropriate. This allows individual loads to tailor their generation directly to the load, making them independent of grid power failures. Classic grids were designed for the one-way flow of electricity, but if a local subnetwork generates more power than it consumes, the reverse flow can raise safety and reliability issues. A smart grid can manage these situations, but utilities routinely manage this type of situation in the existing grid. An electrical grid is not a single entity but an aggregate of multiple networks and multiple power generation companies with multiple operators employing varying levels of communication and coordination, most of which is manually controlled. Smart grids increase the connectivity, automation, and coordination among these suppliers, consumers, and networks that perform either long-distance transmission or local distribution tasks: • Transmission networks move electricity in bulk over medium to long distances. They are actively managed and generally operate from 345 to 800 kV over AC and DC lines. • Local networks traditionally moved power in one direction, distributing bulk power to consumers and businesses via lines operating at 132 kV and lower. This paradigm is changing as businesses and homes begin generating more wind and solar electricity, enabling them to sell surplus energy back to the utilities. Modernization is necessary for energy consumption

278

N. El Bassam

efficiency and the real-time management of power flows and to provide the bidirectional metering needed to compensate local producers of power. Although transmission networks are already controlled in real time, many in the United States and European countries are antiquated by world standards and are unable to handle modern challenges such as those posed by the intermittent nature of alternative electricity generation, or continental-scale bulk energy transmission (Kaplan, 2009).

13.2.2 Smart meters Smart meters are the core element of a smart grid. They ensure that consumers, grids, and utilities are linked in a way that is particularly secure. Smart meters determine how much electricity is used, encrypt this information, and send it to the utility, which calculates the price. Customers have full transparency of their electricity consumption. Grid operators obtain important information about the situation on the ground, which helps them to control the grid better. Smart meters need to comply with strict data protection rules. In addition, the Federal Office for Information Security carries out extensive cybersecurity checks that need to be passed before a smart meter can be used. The security standards that need to be met can be compared with those applicable to banks.

13.2.3 United States version A smart meter consists of a digital meter and a communications unit, the smart meter gateway. Furnished with the mark of the Federal Office for Information Security (BSI), the smart meter gateway permits meters to be connected to the smart grid in a way that meets data protection and data security standards. Support for smart grids became federal policy in the United States with passage of the Energy Independence and Security Act of 2007. The law, Title XIII, set out $100 million in funding per fiscal year from 2008 to 12, establishing a matching program to states, utilities, and consumers to build smart grid capabilities, and created a Grid Modernization Commission to assess the benefits of demand response and recommend needed protocol standards. The Energy Independence and Security Act of 2007 directed the National Institute of Standards and Technology to coordinate the development of smart grid standards, which the Federal Energy Regulatory Commission (FERC) would then promulgate through official rule-making.

Energy storage, smart grids, and electric vehicles

279

Smart grids received further support with the passage of the American Recovery and Reinvestment Act of 2009, which set aside $11 billion for the creation of a smart grid (Wikipedia, 2011). Section 1301 of Title XIII (2007) reads: It is the policy of the United States to support the modernization of the Nation’s electricity transmission and distribution system to maintain a reliable and secure electricity infrastructure that can meet future demand growth and to achieve each of the following, which together characterize a Smart Grid: (1) Increased use of digital information and controls technology to improve reliability, security, and efficiency of the electric grid (2) Dynamic optimization of grid operations and resources, with full cybersecurity (3) Deployment and integration of distributed resources and generation, including renewable resources (4) Development and incorporation of demand response, demand-side resources, and energy-efficiency resources (5) Deployment of smart technologies (real-time, automated, interactive technologies that optimize the physical operation of appliances and consumer devices) for metering, communications concerning grid operations and status, and distribution automation (6) Integration of smart appliances and consumer devices (7) Deployment and integration of advanced electricity storage and peakshaving technologies, including plug-in EVs (PEVs) and HEVs, and thermal-storage air-conditioning (8) Provision to consumers of timely information and control options (9) Development of standards for communication and interoperability of appliances and equipment connected to the electric grid, including the infrastructure serving the grid (10) Identification and lowering of unreasonable or unnecessary barriers to adoption of smart grid technologies, practices, and services 13.2.3.1 Challenges (1) To develop advanced techniques for measuring peak load reductions and energy-efficiency savings from smart metering, demand response, DG, and electricity storage systems; (2) to investigate means for demand response, DG, and storage to provide ancillary services;

280

N. El Bassam

(3) to conduct research to advance the use of wide-area measurement and control networks, including data mining, visualization, advanced computing, and secure and dependable communications in a highly distributed environment; (4) to test new reliability technologies, including those concerning communications network capabilities, in a grid control room environment against a representative set of local outage and wide-area blackout scenarios; (5) to identify communications network capacity needed to implement advanced technologies; (6) to investigate the feasibility of a transition to time-of-use and real-time electricity pricing; (7) to develop algorithms for use in electric transmission system software applications; (8) to promote the use of underemployed electricity generation capacity in any substitution of electricity for liquid fuels in the transportation system of the United States; and (9) in consultation with FERC, to propose interconnection protocols to enable electric utilities to access electricity stored in vehicles to help meet peak demand loads.

13.2.4 European strategies The Smart Grids Platform was started by the European Commission Directorate General for Research of the European Commission in 2005. The European Technology Platform for Electricity Networks of the Future, also called Smart Grids European Technology Platform (ETP), is the key European forum for the crystallization of policy and technology research and development pathways for the smart grids sector. It is also the link between European Unionelevel related initiatives. Development of smart grid technologies is part of the ETP initiative. Its aim is to formulate and promote a vision for the future development of European electricity networks. Smart grid deployment is traditionally based on improving utility operations at both the transmission and distribution grid levels. Since 2010, we have seen accelerated deployments of advanced metering infrastructure, systems to improve voltage and outage management, and synchrophasor technology to enhance situational awareness. However, we are now witnessing the rapid adoption of distributed technologies such as PV systems, and

Energy storage, smart grids, and electric vehicles

281

increasing ownership of these distributed assets by utility customers and third-party merchants. The proliferation of distributed devices is driven largely by state policies, lowering technology costs, and changing customer expectations, and is not occurring consistently across the country. Where it is happening, the rise in the number of distributed technologies and their ownership by entities other than utilities significantly increases the complexity of grid operations and poses challenges to traditional approaches for grid planning and market designs. Addressing the emerging complexity will require the deployment of advanced grid capabilities based largely on the application of smart grid technology. This will include the continued development of a variety of technologies and improved strategies for grid sensing, information management, communications, control, and coordination. In this effort, we will also need to ensure the affordability, reliability, resilience, and security of the electric grid. The department will continue to work closely with the electric utility industry and federal and state agencies to determine prudent approaches for deploying smart grid technologies. We have witnessed the accelerated deployment of technologies meant to improve the reliability and efficiency of utility operations, including the deployment of systems and practices to engage utility customers better in managing energy. This has included increased deployments of advanced metering infrastructure, systems to improve voltage and outage management, and synchrophasor technology to enhance situational awareness. However, more recently, we have witnessed the rapid adoption of distributed technologies, such as PV systems, and increasing ownership of these distributed assets by utility customers and third-party merchants. The effective integration of the grid with distributed assets presents a more complex and potentially transformative situation that will require the deployment of advanced grid capabilities based largely on the application of smart grid technology. The smart grid is enabled by digital technology applied in devices and systems that allows for enhanced sensing and control of grid elements, more widespread information sharing and communication, more powerful computing, and finer control. Integration of the digital structure with the physical structure of the grid is evolving rapidly owing to the enhanced performance and declining costs of digital technology. Digital networks will eventually lead to greater levels of information exchange between utilities and their customers, as well as the convergence of the electric

282

N. El Bassam

grid with other infrastructures such as buildings, transportation, and telecommunications. The ETP for the Electricity Networks of the Future (Smart Grids) is a European Commission initiative that aims to boost the competitive situation of the European Union in the field of electricity networks, especially smart power grids. The ETP represents all European stakeholders. The establishment of an ETP in this field was first suggested by industrial stakeholders and the research community at the first International Conference on the Integration of Renewable Energy Sources and Distributed Energy Resources, held in Dec. 2004 (European Commission, 2006). In so-called E-Energy projects, several German utilities are creating the first nucleolus in six independent model regions. A technology competition identified these model regions to carry out research and development activities with the main objective of creating an Internet of Energy(Federation of German Industries, 2010) (Figure 13.7).

13.2.5 Korean version Smart grids in South Korea constitute a platform that is reimagining electricity grids, equipping them with technology that allows more capability,

Figure 13.7 The internet of energy integrates all elements in the energy supply chain to create an interactive system. (From Federation of German Industries (2010)).

Energy storage, smart grids, and electric vehicles

283

particularly in addressing the demands of the 21st century and forward. This process follows a modular approach to grid construction and focuses on the development of Internet technology enabling its electric power generation system. The country views smart grids, along with so-called new energy industries, as an emergent pillar of the Korean economy (https://en.wikipedia. org/wiki/Smart_grids_in_South_Korea). According to Korea Electric Power Corporation, a leaders of the initiative, “smart grids would help the country use more RES and cut overall energy consumption.” The “smart” in a grid is achieved through installed software rather than hardware, banking more on the element of intelligence for more consistent upgrades, pattern learning, and timely response to new technologies. South Korean smart grids include the components: Smart power: The intelligent monitoring of demand, a high level of fault tolerance, and fast restoration in case of failures; Smart service: The provision of domestic, commercial, and industrial customers with electricity tariffs and services customized according to their needs; Smart place: The use of intelligence at home (e.g., smart appliances), real-time pricing, and demand management; Smart transport: Installation of sophisticated systems to manage the connections of EVs to the smart grid effectively; and Smart renewables: The connection and use of large and diverse sources of power to the grid to ensure stability. The Korean government has launched a $65 million pilot program on Jeju Island with major players in the industry. The program consists of a fully integrated smart grid system for 6000 households; wind farms and four distribution lines are included in the pilot program. This demonstrates the extent of Korea’s commitment to an environmentally viable future. Korea plans to slash overall energy consumption by 3% and cut total electric energy consumption by 10% before 2030. The government also plans to reduce greenhouse gas emissions by 41 million tons by that time. The government announced that it will undertake nationwide smart grid implementation by 2030 (Korea Smart Grid Institute). A full-scale test bed for smart grid technologies will be introduced in Jeju Island, South Korea (http://wallpapershighresolution.blogspot.com/2012/ 07/jeju-island-wallpapers.html). Consumers were placed at the core of the South Korean smart grid project, with an emphasis on empowering grid users by informing them about how much energy they use at any given time, while encouraging them to

284

N. El Bassam

adjust their consumption patterns to the amount of energy available. Up to 2000 households on Jeju Island (located south of the Korean peninsula) were chosen to test the project. Each home was fitted with a smart meter sending real-time information on its energy use to the grid operator, which in turn implemented a real-time pricing scheme to encourage users to adjust consumption patterns (Figure 13.8). In more practical terms, 84 new technologies were tested as part of the experiment, among which are home appliance monitoring systems such as smart dishwashers and smart refrigerators, which aim to perform the most energy-intensive tasks during off-peak hours, thus shedding load on the grid at peak times. Besides automated monitoring and control, systems feature a phone app that enables users to operate their smart appliances remotely when they are not at home (https://www.google.com/search? client¼firefox-b-d&q¼smartþgridþinþkorea).

13.3 Electric vehicles 13.3.1 Current developments EVs first appeared in the mid-19th century. An EV held the vehicular land speed record until around 1900. The high cost, low top speed, and short range of BEVs, compared with later internal combustion engine

Figure 13.8 Jeju Island, South Korea is one of seven new wonders. Jeju Island is located 130 km off the southern coast of Korea.

Energy storage, smart grids, and electric vehicles

285

vehicles, led to a worldwide decline in their use, although EVs have continued to be employed in the form of electric trains and other niche uses. At the beginning of the 21st century, interest in electric and other alternative fuel vehicles increased owing to growing concern over problems associated with hydrocarbon-fueled vehicles, including damage to the environment caused by their emissions and the sustainability of the current hydrocarbon-based transportation infrastructure as well as improvements in EV technology. Since 2010, combined sales of all-electric cars and utility vans achieved one million units delivered globally in Sep. 2016, and reached 3.3 million units in Dec. 2018. The global ratio of annual sales of battery electric cars and plug-in hybrids went from 56:44 in 2012 to 74:26 in 2019. As of Mar. 2020, the Tesla Model 3 was the world’s all-time best selling plugin electric passenger car, with over 500,000 units. Over the past few decades, the environmental impact of the petroleumbased transportation infrastructure, along with peak oil, has led to renewed interest in an electric transportation infrastructure (Eberle and von Helmolt, 2010). EVs differ from fossil fuel-powered vehicles in that the electricity they consume is generated from a wide range of renewable sources. However it is generated, this energy is then transmitted to the vehicle through the use of overhead lines, wireless energy transfer such as inductive charging, or a direct connection through an electrical cable. The electricity may then be stored onboard the vehicle using a battery, flywheel, or supercapacitors. Vehicles using engines working on the principle of combustion can usually derive energy only from a single or a few sources, usually nonrenewable fossil fuels. A key advantage of electric or HEVs is regenerative braking and suspension, the ability to recover energy normally lost during braking as electricity to be restored to the onboard battery (Levant Power Corp, 2011). Because EVs can be plugged into the electric grid when not in use, there is a potential for battery-powered vehicles to even out the demand for electricity by feeding electricity into the grid from their batteries during peak use periods (such as midafternoon air-conditioning use) while doing most charging at night, when there is unused generating capacity (Pacific Gas and Electric Company, 2007). This vehicle to grid connection has the potential to reduce the need for new power plants. Furthermore, our current electricity infrastructure may need to cope with increasing shares of variable-output power sources such as windmills and PV solar panels. This variability could be addressed by adjusting the

286

N. El Bassam

speed at which EV batteries are charged or possibly even discharged (Figure 13.9). Some concepts are battery exchanges and battery charging stations, much like gas and petrol stations today. Clearly, these will require enormous storage and charging potentials, which could be manipulated to vary the rate of charging, and to output power during shortage periods, much as diesel generators are used for short periods to stabilize some national grids (Andrews, 2006). With electric cars gaining in popularity, AEP Ohio and Walmart premiered the region’s first free, public EV charging station at the Walmart Supercenter/Sam’s Club at 3900 Morse Road, Ohio. The Blink charging station was developed by San Francisco-based ECOtality, Inc., a provider of clean electric transportation and storage technologies. The charging station features two Blink Pedestal units that allow two drivers to charge their EVs simultaneously. According to AEP, a million PEVs were expected to be on the road, which is why AEP launched the demonstration project to determine the best way to integrate PEVs into its electricity distribution system. To analyze the new demand, up to 10 people will drive a PEV as part of their daily routine to assess the amount of power needed for basic travel, effects of PEV charging during peak electricity periods and the quality of charging technology. An EV, also referred to as an electric drive vehicle, is a vehicle which uses one or more electric motors for propulsion. Depending on the type of

Figure 13.9 In 2016, Solar Impulse 2 was the first solar-powered aircraft to complete circumnavigation of the world. (From https://en.wikipedia.org/wiki/Electric_aircraft).

Energy storage, smart grids, and electric vehicles

287

vehicle, motion may be provided by wheels or propellers driven by rotary motors, or in the case of tracked vehicles, by linear motors. EVs can include electric cars, electric trains, electric trucks, electric lorries, electric airplanes, electric boats, electric motorcycles and scooters, and electric spacecraft. There are two basic types of EVs: all-EVs (AEVs) and plug-in HEVs (PHEVs). AEVs include BEVs and fuel cell EVs (FCEVs). In addition to charging from the electrical grid, both types are charged in part by regenerative braking, which generates electricity from some of the energy normally lost when braking. Which type of vehicle will fit your lifestyle depends on your needs and driving habits.

13.3.2 Types of electric vehicles • EVs (also known as PEVs) derive all or part of their power from electricity supplied by the electric grid. They include AEVs and PHEVs. • AEVs are powered by one or more electric motors. They receive electricity by plugging into the grid and store it in batteries. They consume no petroleum-based fuel and produce no tailpipe emissions. AEVs include BEVs and FCEVs. • PHEVs use batteries to power an electric motor, plug into the electric grid to charge, and use a petroleum-based or alternative fuel to power the internal combustion engine. Some types of PHEVs are also called extended-range EVs (Figure 13.10).

13.3.3 Global battery electric vehicle and plug-in hybrid electric vehicle sales Global plug-in vehicle deliveries reached 2,264,400 units in 2019, 9% higher than for 2018. This is a clear departure from the growth rates of the previous 6 years, which were between 46 and 69%. The reason was developments in the two largest markets, China and United States, where sales stagnated in the second half of 2019 and stayed significantly below the sales boom in the second half of 2018. In the United States, sales of most plug-in models decreased compared with the boom in the second half of 2018. In China, further slashing of subsidies, paired with more stringent technical regulation, caused a crash in NEV demand and supply, starting in July (Figure 13.11). Europe became the beacon of 2019 EV sales with 44% growth, accelerating toward the end of the year. The WLTP introduction, together with changes in national vehicle taxation and grants, created more awareness of and demand for EVs. The industry geared up to meet the 95 g CO2/km

288

N. El Bassam

Bloomberg

140

ICE + HEV

NEW ENERGY FINANCE

100%

120

90%

PHEV EV % of new sales

80%

100

70% 60%

80

50% 60

35%

40% 30%

40

% of new car sales

m of vehicles sold per year

BEV

20% 20 0 2015

10% 0% 2020

2025

2030

2035

2040

Figure 13.10 The electric vehicle (EV) market is booming and the future looks good. BEV, battery electric vehicle; HEV, hybrid electric vehicle; ICE, internal combustion engine; PHEV, plug-in hybrid electric vehicle. (From https://cadenzainnovation.com/blog/ electric-vehicle-ev-market-booming-future-looks-good/, 2016).

Figure 13.11 Global light-duty electric vehicle sales, 2010e18. (From https://www. researchgate.net/figure/Global-light-duty-electric-vehicle-sales-2010-2018_fig1_332170448).

target for 2020e21. Over 30 new and improved BEV and PHEV models were introduced in 2019, many of them in the fourth quarter, which pushed EV sales in that year and the next (https://www.ev-volumes.com/) (Figure 13.12). The global BEV and PHEV share for 2019 was 2.5% and the smaller car markets continued to lead EV adoption. The share leader was Norway, as

Energy storage, smart grids, and electric vehicles

289

Figure 13.12 Annual sales of plug-in passenger cars in Europe compared with the world’s top country markets between 2011 and 2019. (From https://en.wikipedia. org/wiki/Plug-in_electric_vehicles_in_Europe#/media/File:Global_plug-in_car_sales_since_ 2011.png).

usual, where 56% of new car sales were plug-ins in 2019. Iceland came in second at 24.5% and the Netherlands third at 15%. Among the larger economies, China led with a plug-in share of 5.2%, the United Kingdom posted 3.2%, Germany 2.9%, France 2.8%, and Canada 2.7%. All other car markets with over one million total sales showed 2% or less for 2019. At the end of 2019, the global fleet of plug-ins was 7.5 million, counting light vehicles. Medium and heavy commercial vehicles added another 700,000 units to the global stock of plug-ins. Their global deliveries were 100,000 units in 2019, 95%of which were in China, mostly large buses. With Covid-19, the outlook for 2020 global EV sales has become more difficult. High growth in Europe was expected throughout the year and high growth in the United States and other markets in the second half of the year, but China might become another disappointment. Preliminary EV sales data for Jan. and Feb. were positive in Europe and encouraging in the United States, but dismal in China, where the total vehicle market was down 80% in Feb. If quarantines and factory closures continued into the second quarter, insufficient parts supply would affect the global car

290

N. El Bassam

industry during a longer period and the lost volumes would be unlikely to be recovered during the year. In China, this has another dimension when dealers remain closed and buyers have to stay at home. The charging station is part of AEP Ohio’s gridSMART initiative to promote energy efficiency and Walmart’s corporate Sustainability 360 effort. The gridSMART initiative is designed to provide customers with greater energy control and improve electric distribution service and performance through rebates and cost incentives, plus tips, tools, and technologies (New Albany Innovation Exchange, 2011).

13.3.4 Types of electric vehicles Types of EVs are BEVs, PHEVs, and HEVs (https://www.evgo.com/whyevs/types-of-electric-vehicles/). There are three main types of EVs, classed by the degree to which electricity is used as the energy source. BEVs, PHEVs, and HEVs. Only BEVs are capable of charging on a level 3 DC fast charge. 13.3.4.1 Battery electric vehicles BEVs, more frequently called EVs, are fully EVs with rechargeable batteries and no gasoline engine. BEVs store electricity onboard with high-capacity battery packs. Their battery power is used to run the electric motor and all onboard electronics. BEVs do not emit harmful emissions and hazards caused by traditional gasoline-powered vehicles. BEVs are charged by electricity from an external source. EV chargers are classified according to the speed with which they recharge an EV battery. The classifications are level 1, 2, and 3 or DC fast charging. 1. Level 1 EV charging uses a standard household (120-V) outlet to plug into the EV; it takes over 8 h to charge an EV for approximately 75e80 miles. Level 1 charging is typically done at home or at the workplace. Level 1 chargers have the ability to charge most EVs on the market. 2. Level 2 charging requires a specialized station that provides power at 240 V. Level 2 chargers are typically found at workplaces and public charging stations and will take about 4 h to charge a battery to 75e80 miles of range. 3. Level 3 charging, DC fast charging, or simply fast charging is currently the fastest charging solution in the EV market. DC fast chargers are found at dedicated EV charging stations and charge a battery up to 90 miles range in approximately 30 min.

Energy storage, smart grids, and electric vehicles

291

BEV examples that can charge on DC level 3 fast chargers are: Tesla Model 3 BMW i3 Chevy Bolt Chevy Spark Nissan LEAF Ford Focus Electric. Hyundai Ioniq Karma Revero Kia Soul Mitsubishi i-MiEV Tesla Model S Tesla X Toyota Rav4 Volkswagen e-Golf 13.3.4.2 Plug-in hybrid electric vehicles PHEVs can recharge the battery through both regenerative braking and plugging into an external source of electrical power. Whereas at low speed standard hybrids can go about 1e2 miles before the gasoline engine turns on, PHEV models can go anywhere from 10 to 40 miles before the gas engines provide assistance. PHEV examples are: Chevy Volt Chrysler Pacifica. Ford C-Max Energ Ford Fusion Energ Mercedes C350e Mercedes S550e Mercedes GLE550e Mini Cooper SE Countryman Audi A3 E-Tron BMW 330e BMW i8 BMW X5 xdrive40e Fiat 500e Hyundai Sonata Kia Optima Porsche Cayenne S E-Hybrid

292

N. El Bassam

Porsche Panamera S E-hybrid Toyota Prius Volvo XC90 T8 13.3.4.3 Hybrid electric vehicles HEVs are powered by both gasoline and electricity. The electric energy is generated by the car’s own braking system to recharge the battery. This is called regenerative braking, a process in which the electric motor helps to slow the vehicle and uses some of the energy normally converted to heat by the brakes. HEVs start off using the electric motor; then, the gasoline engine cuts in as load or speed rises. The two motors are controlled by an internal computer, which ensures the best economy for driving conditions. HEV examples are: Toyota Prius Hybrid Honda Civic Hybrid Toyota Camry Hybrid

13.4 Future developments The ability to power wireless sensor nodes from harvested energy sources allows embedded designers to offer systems with a significantly reduced cost of ownership for the end user as well as benefits to the environment. The cost of replacing batteries housed in out-of-the-way sensor node locations can be significant (Figure 13.13). For example, these wireless sensor nodes can be embedded in structures such as buildings or bridges, or even buried underground. Three key

Figure 13.13 Fast charging station for future electric vehicles. (From Milano Medien GmbH/Siemens AG (2009)).

Energy storage, smart grids, and electric vehicles

293

Figure 13.14 Fast charging station for future electric vehicles embedded in public places such as shopping malls. (From Milano Medien GmbH/Siemens AG (2009)).

enabling technologies needed to create self-sustaining wireless sensor nodes are readily available today: cost-effective energy harvester devices, small and efficient energy storage devices, and single-chip ultralow-power wireless MCUs. Wireless sensor nodes powered by harvested energy sources will soon become commercially viable and commonplace technologies used in our homes, offices, factories, and infrastructure (Silicon Laboratories, 2011). The high cost of electric cars is one of the most common reasons consumers cite for choosing not to own one. For those people, a pair of green tech startups plans to implement a program that would enable EV owners to offset some of the costs by selling electric storage services to help stabilize the electricity grid (Nguyen, 2011). The following images show plans for some future technologies related to EVs (Figures 13.4 and 13.5). Electromagnetic radiation from high-performance electrical motors has been claimed to be associated with some human ailments, but such claims are largely unsubstantiated except for extremely high exposures. Electric motors can be shielded within a metallic Faraday cage, but this reduces efficiency by adding weight to the vehicle, although it is not conclusive

Figure 13.15 Autonomous solar power/electricity/combined heat and power at the microgrid of a family house, including electric car charging facilities. (From Milano Medien GmbH/Siemens AG (2009)).

294

N. El Bassam

that all electromagnetic radiation can be contained (http://www.greenfacts. org/en/power-lines/index.htm).

References Andrews, D., 2006. Senior Technical Consultant, Biwater Energy. January 2006. A Talk Originally Given by as the Energy Manager at Wessex Water at an Open University Conference on Intermittency 24th National Grid’s Use of Emergency. Diesel Standby Generator’s in Dealing with Grid Intermittency and Variability. Potential Contribution in Assisting Renewables. Duke Energy, 2012. Generating Electricity with Pumped-Storage Hydro. http://www. duke-energy.com/about-energy/generating-electricity/pumped-storage-faq.asp. Eberle, U., von Helmolt, R., 2010-05-14. Sustainable Transportation Based on Electric Vehicle Concepts: A Brief Overview. Royal Society of Chemistry. http://pubs.rsc. org/en/Content/ArticleLanding/2010/EE/c001674h. Energy Independence and Security Act, U.S.C. Title XIII Sec. 1301, 2007. Retrieved from: http://energy.gov/sites/prod/files/oeprod/DocumentsandMedia/EISA_Title_XIII_ Smart_Grid.pdf. European Commission, 2006. European SmartGrids Technology Platform: Vision and Strategy for Europe’s Electricity Networks of the Future Luxembourg, EUR 22040. Federation of German Industries/BDI Initiative Internet of Energy, 2010. Internet of Energy ICT for Energy Markets of the Future: The Energy Industry on the Way to the Internet Age (translation of the brochure “Internet der Energie e IKT fuVr EnergiemVarkte der Zukunft” published in Germany, December 2008, to which information about the German government’s E-Energy model projects has been added). Available from: http://www.bdi.eu/bdi_english/download_content/Marketing/Brochure_Internet_ of_Energy.pdf. Fuel Cells, 2008. Fuel Cell Basics, 2000. http://www.fuelcells.org/basics/types.html. Kaplan, S.M., 2009. Electrical Power Transmission: Background and Policy Issues. Smart Grid. The Government Series, pp. 1e42 (Capital.Net). Levant Power Corp, 2011. Revolutionary GenShock Technology. http://www. levantpower.com/technology.html. Maegaard, P., 2004. Wind Energy Development and Application Prospects of Non- GridConnected Wind Power 2004. World Wind Energy Institute. World Renewable Energy Committee, Nordic Folkecenter for Renewable Energy, Denmark. Maegaard, P., June 2011. Integrated Systems to Reduce Global Warming 2011. Springer Science Business Media. LLC 22 PDF. Mancini, T., 2006. Advantages of Using Molten Salt. Sandia National Laboratories. Archived from the original on 2011-07-14. http://www.webcitation.org/60AE7heEZ. New Albany Innovation Exchange, 2011. Ohio’s First Electric Car Charging Station. Posted on September 15, 2011. http://www.innovatenewalbany.org/business/ohio%E2%80% 99s-first-electric-car-charging-station/. Nguyen, T.C., 2011. Electric Cars Can Now Earn Money for Owners. September 29, 2011. Smart planet. Available from: http://www.smartplanet.com/blog/thinking-tech/ electric-cars-can-now-earn-money-for-owners/8756. Pacific Gas and Electric Company, 2007. First Vehicle-To-Grid Demonstration. http:// seekingalpha.com/article/31992-pacific-gas-and-electric-demonstrates-vehicle-to-gridtechnology. Pockley, S., 19/05/2008. Compressed Air Energy Storage (CAES). Prepared for Intro. to Renewable Energy (PDF).

Energy storage, smart grids, and electric vehicles

295

QuantumSphere Inc., 2006. Highly Efficient Hydrogen Generation via Water Electrolysis Using Nanometal Electrodes. http://www.qsinano.com/white_papers/2006_09_15. pdf. Tennessee Valley Authority (TVA), 2004. Energy [WWW]. Available from: http://www. tva.gov/power/pumpstorart.htm. The Free Dictionary, 2002. McGraw-Hill Concise Encyclopedia of Engineering. http:// encyclopedia2.thefreedictionary.com/Energy storage Energy storage. U.S. Department of Energy, 2011. Smart grid/department of energy. Retrieved 2012-06-18. In: The Energy Harvesting Tipping Point for Wireless Sensor Applications [WWW]. Silicon Laboratories. White paper available from: http://www.eetimes.com/electricalengineers/education-training/tech-papers/4217176/Th Energy-Harvesting-TippingPoint-for-Wireless-Sensor-Applications. U.S. Department of Energy, 2012. Smart Grid/Department of Energy. Retrieved 2012-06-18. Wagner, L., December 2007. Overview of Energy Storage Methods. Research report. http://www.moraassociates.com/publications/0712%20Energy%20storage.pdf. Wagner, L., January 2008. Nanotechnology in the Clean Tech Sector. Research report. http://www.moraassociates.com/publications/0801%20Nanotechnology.pdf. Wild, M.L., July 28, 2010. Wind Drives Growing Use of Batteries. New York Times, p. B1.

Further reading Doughty, D., Butler, H., Paul, C., Akhil, A.A., Clark, N.H., Boyes, J.D., 2010. Batteries for Large-Scale Stationary Electrical Energy Storage. The Electro- chemical Society Interface. http://www.electrochem.org/dl/interface/fal/fal10/fal10_ p049-053.pdf. Electrical Storage, 2012. http://www.microchap.info/electrical_storage.htm. Korea Smart Grid Institute (KSGI), 2010. Korea’s Jeju Smart Grid [WWW]. Available from: http://www.smartgrid.or.kr/10eng3-3.php. Pacific Gas & Electric (PG&E), 2011. Smart Grid Definition. Smart Grid Deployment Plan. Available from: http://www.neuralenergy.info/2009/09/pg-e.html. U.S. Dept of Energy (DOE), October 2011. Gemasolar Thermosolar Plant. Concen- Trating Solar Power Projects. National Renewable Energy Laboratory (NREL), 24.

CHAPTER FOURTEEN

Current distributed renewable energy in rural and urban communities N. El Bassam International Research Centre for Renewable Energy, IFFED.org, Germany

Contents 14.1 Thisted, Denmark: 100% renewable energy community 14.1.1 Implementation 14.2 Samsø island 14.3 Energy island of VindØ 14.4 Kampala, Uganda taxi-bike drivers move to electric bikes 14.5 Rural community of J€ uhnde 14.6 Containerized solar minigrid, Fanidiama village, Mali 14.7 Decentralized desalination systems powered by solar energy in Maasai, Tanzania 14.8 Road map to renewable energy in remote communities in Australia 14.9 Iraq Dream Homes 14.10 Renewables in Africa 14.10.1 Hydropower 14.10.2 Biomass 14.10.3 Geothermal 14.10.4 Wind power 14.10.5 Solar power 14.10.6 Biofuels 14.10.7 Energy efficiency 14.11 Renewables in India 14.12 Distributed renewable energy and solar oases for deserts and arid regions: the DESERTEC concept 14.13 Vatican City 14.13.1 And suddenly there was light! References Further reading

Distributed Renewable Energies for Off-Grid Communities, Volume 2 ISBN: 978-0-12-821605-7 https://doi.org/10.1016/B978-0-12-821605-7.00024-6

© 2021 Elsevier Inc. All rights reserved.

298 298 299 301 303 304 306 307 308 312 313 315 315 316 316 317 318 318 319 320 323 323 324 326

297

j

298

N. El Bassam

This chapter reviews some selected implemented and planned renewable energy communities in rural, remote. and urban areas. Many more communities have been established across the continents.

14.1 Thisted, Denmark: 100% renewable energy community 14.1.1 Implementation The municipality of Thisted (population, 43,993; area, 1072 km2) is located in the northwest section of Jutland region in Denmark. Today, it is seen as a pioneer in the grassroots development of renewable energy based on decentralized energy. In fact, the successful transition toward 100% renewable electricity in Thisted was not the result of a political decision; it has its origins in the gradual expansion of private investments into the use of renewable energy since the early 1980s. Today, all but one of the 252 wind turbines surrounding Thisted are privately owned. In total, 80% of Thisted’s total electricity demand is produced by wind power, whereas the remaining 20% are provided from biogas plants. Additional production of energy comes from privately owned solar power and geothermal plants that cover 85% of Thisted’s heating demand with renewables. Thisted is pursuing its energy goal based on an alliance of citizens, grassroots organizations, and local companies, as well as the commitment of the municipal administration (Thisted, Denmarkd100% Renewable Energy Atlas, 2019). The involvement of the community is particular important in providing the finance of the renewable energy systems. For example, many farmers have invested in wind power, and these investments have paid off quickly, generating benefits after 6e7 years. Aside from the surplus energy that is sold and fed into the general grid, agricultural by-products are able to be processed into bioethanol, biogas, or biopellets, constituting an additional energy supply from renewable sources. At the same time, Thisted’s citizens are saving one-third of their energy costs for heating compared with oilbased heating systems. Thisted municipality has inspired innovation and has since established test centers for the development of new approaches in renewable energy production based on wave energy, as well as wind, solar, and biogas.

Current distributed renewable energy in rural and urban communities

299

Moreover, the community plans to build a virtual power plant to be able to manage the decentralized energy supplies more efficiently, such as temporarily shifting the energy supply to the most economical production source: RES: Wind power, geothermal energy, biogas, solar photovoltaics (PV), and district heating. Target: 100% self-sufficiency in electricity supply from renewable sources. Status: Achieved. (https://www.100-percent.org/thisted-denmark/)

14.2 Samsø island The world’s first 100% renewable energy-powered island and 100% CO neutral, Samsø is a Danish island in the Kattegat 15 km (9.3 miles) off the Jutland Peninsula. Samsø is located in Samsø municipality. The community has 3724 inhabitants (2017), and is 114 km2 in area. Owing to its central location, the island was used during the Viking Age as a meeting place. In 1997, Samsø won a government competition to become a model renewable energy community. As a result, 100% of its electricity today comes from offshore and onshore wind power and biomass. It also has several biomass-based district heating systems, so that 70% of heat demand is generated by local resources. The island often exports renewable electricity to the mainland. Although there is not yet an issue with curtailment of renewable generation in Samsø’s energy system, several bottlenecks present opportunities for the better management of locally generated energy. Addressing these issues, by shifting peaks in energy demand for example, can help to stabilize and reduce energy prices for residents and provide a valuable service for the local DSO by helping them to manage and balance the grid overall. Samsø island in Denmark was transformed into a green powerhouse with onshore and offshore wind turbines, biomass boilers, and solar (https:// www.theguardian.com/sustainable-business/2017/feb/24/energy-positivehow-denmarks-sams-island-switched-to-zero-carbon 2020. https://www. google.com/search?client¼firefox-b-d&q¼Samso.dk. https://trustyetverify. wordpress.com/2016/04/23/if-it-sounds-too-good-to-be-true-Samso-aself-sufficient-island/; https://blog.continentalcurrency.ca/Samsø-greenenergy/). 2

300

N. El Bassam

Samsø runs entirely on renewable energy: all electricity is generated by wind turbines and 70% of heating comes from solar plants; small-scale biomass plant settles the difference (Figure 14.1). Samsø has three straw heating plants, one solar and wood chip heating plant, 11 onshore turbines of 1 MW, and 10 offshore turbines of 2.3 MW; in 2013, it had also 12 microturbines. That is a lot for a population of about 4000 and a yearly consumption of about 25,000 MWh/y (https:// trustyetverify.wordpress.com/2016/04/23/if-it-sounds-too-good-to-be-trueSamsø-a-self-sufficient-island/2020). The vision of making Samsø an island free of fossil fuels by 2030 has become part of a Danish project with scenarios developed to reduce the island’s heating demand by installing heat pumps in the district heating networks, as in buildings that are heated by fossil fuel-based systems. Ideas for converting transportation to electric transport means and for shifting current biomass consumption from the heating to the transportation sector are also considered. It is possible to make a 100% renewable energy system at Samsø by 2030 using only local electric power generation by wind turbines and PV systems together with the use of biomass resources. Associated costs will stay similar to current costs, but the system could also contribute to the creation of local jobs, considering that the island is already a destination for international energy tourism, and it may enhance the security of supply (https://www.theguardian.com/sustainable-business/ 2017/feb/24/energy-positive-how-denmarks-sams-island-switched-to-zerocarbon 2020). There are plans to establish a cooperatively run pig farmebased biogas plant. This biogas will be used to produce electricity; the excess heat will be used for heating purposes (Figure 14.2).

Figure 14.1 The plant is run on 75% wood chips and 25% solar thermal collectors.

Current distributed renewable energy in rural and urban communities

301

Figure 14.2 Samsø municipality’s electric cars recharge at a solar-powered charging station under a roof filled with solar panels. (From Stetson Freeman/Staff Melanie).

Information campaigns promote new neighborhood heating systems and individual solutions. Heat pump systems are attractive because the island generates a surplus of electricity. The transportation sector can be partially supplied with canola oil for diesel vehicles, and the island’s gasoline cars can use bioethanol or be converted to hydrogen and electricity. These will be ready when technological innovation and lower prices make large-scale conversion feasible (Birger Jensen/Samsø Tourism https://www.theguardian.com/sustain able business/2017/feb/24/energy-positive-how-denmarks-sams-islandswitched-to-zero-carbon 2020; https://www.google.com/search?client¼ firefox-b-d&q¼Samso.dk; https://trustyetverify.wordpress.com/2016/04/ 23/if-it-sounds-too-good-to-be-true-Samso-a-self-sufficient-island/2020; https://trustyetverify.wordpress.com/2016/04/23/if-it-sounds-too-goodto-be-true-Samsø-a-self-sufficient-island/2020; https://en.wikipedia.org/ wiki/Sams%C3%B8#/media/File:Denmark_location_Samsø.svg).

14.3 Energy island of VindØ The Danish government is proposing to build two energy islands in the North and Baltic Seas by 2030 in its Climate Action Plan (Figure 14.3).VindØ Island will be made up of submersible concrete boxes installed some 100 km (62.1 miles) off the coast in the North Sea. There will be a possibility for the island’s continuous expansion in line with the development of storage and Power-to-X technologies, as well as its connection to other countries in the North Sea area. The project will realize the Danish government’s ambition to create energy islands in the North Sea and can

302

N. El Bassam

Figure 14.3 The vision and future project site. (From https://www.erneuerbareenergien. de/investoren-stehen-bereit-fuer-kuenstliche-nordsee-windstrominsel 2020; TenneT).

be financed entirely without state funds. The consortium said it is a large and demanding project, but it can be implemented using existing technology. The plan includes using the Bornholm island in the Baltic Sea as a 2-GW energy island, and will establish a 2-GW artificial energy island in the North Sea that has a potential capacity of 10 GW (https://www.offshorewind.biz/ 2020/05/22/pensiondanmark-pfa-and-seas-nve-ready-to-finance-vindoenergy-island/). According to the government, the long-term aim is to convert power from the islands into green hydrogen, which can further be processed into fuels, allowing the reduction of greenhouse gases, a concept called Power-to-X. Overall, the plan is expected to enable a greenhouse gas reduction of two million tons by 2030. The plan includes using the Bornholm island in the Baltic Sea as a 2-GW energy island and establishing a 2-GW artificial energy island in the North Sea that has a potential capacity of 10 GW. Besides Denmark, the North Sea Island will be connected to the Netherlands, whereas Bornholm will have a connection to Poland, because the 4-GW capacity expansion tops Danish household annual consumption. In addition to the energy islands, the Climate Action Plan includes investment into green heat, green fuels, and CO2 capture, and increases energy efficiency, transforming the industry and creating a completely green waste sector. Overall, the plan is expected to enable a greenhouse gas reduction of two million tons by 2030.

Current distributed renewable energy in rural and urban communities

303

14.4 Kampala, Uganda taxi-bike drivers move to electric bikes Environmental transition (greener streets and technology) is the trend in Uganda. Motorcycle taxi drivers in the capital Kampala are no exception. They have been testing e-bikes for some time, transporting passengers on electric bicycles powered by solar energy. Electric mobility is still a new phenomenon in Uganda; less than 10% of vehicles on the roads of Kampala, the country’s capital, is powered by electric energy. Although local companies such as Zembo, importer of custommade electric bikes and supplier of solar energy, and Bodawerk, specializing in converting conventional motorcycles to electric ones, are working to make the streets cleaner, much remains to be done (Figure 14.4). Against this backdrop, the International Climate Initiative of the German Environment Ministry and the United Nations Development Program have set up the e-bikes project. The aims of the project are to help Uganda transition to electric mobility and to raise awareness of green technologies. Motorcycle taxi drivers in Kampala are gradually adopting the technology to make their operations more environmentally friendly and lower operating costs. Having more eco-friendly vehicles and devices would limit the production of carbon dioxide and reduce air pollution, especially the emission of fine particles, which are harmful to nature and humanity. The green bicycle project, developed in the East African country, is part of a $34 million (about 133 billion Ugandan shillings) initiative by the United Nations Environment Program and the International Energy Agency to support the transition to electric mobility in 17 developing

Figure 14.4 Anticipated fleet of electric bikes powered by solar energy. (From I. Magoum, Africa 21, https://www.afrik21, 2020; africa/en/uganda-kamapala-taxi-bikedrivers-move-to-electric-bikes/).

304

N. El Bassam

countries. It was implemented by the International Climate Initiative of the German Environment Ministry in partnership with the Ugandan Ministry of Energy and Mineral Development.

€ hnde 14.5 Rural community of Ju The community project of J€ uhnde was inspired by the IEC concept and other institutions. The village of J€ uhnde (750 inhabitants) is located in Southern Lower Saxony in the middle of Germany. It was started in 2001 to become a bioenergy village. With one-third of funds from the German Ministry BMELV and Lower Saxony, it was possible to invest in such a project. The main emphasis is that the whole village is involved, and more than 70% of the households are connected to the hot water grid. The aim of the project is to convert biological material into electrical power and heat. A block-type thermal power station (or heat and power generator) run by biogas has been realized. For additional heating during winter, a wood hogged heating system was implemented (IZNE, 2007). Under anaerobic conditions, microorganisms engage in enzymatic digestion to create biogas. Biogas is obtained during the fermentation process of liquid manure and plant silage in an anaerobic digestion plant. The combustion of biogas in a combined heat and power generator generates enough electricity for the entire village. Biogas also generates heat as a by-product. This heat is used mainly to heat homes and other living spaces, replacing the conventional fossil fuels, oil, and natural gas. A smaller portion of the generated heat is required to fuel the digestion process. The amount of heat generated cannot cover the high demand during winter months in Germany. During this period, an additional heating plant fueled with wood chips is required. Rarely, on extremely cold days, peak demand necessitates a further boiler fueled by oil or biodegradable diesel. The distribution of heat energy to the 140 households in J€ uhnde (750 inhabitants) started in 2005. In 2008, the project produced more than 10,000,000 kWh electricity and saves 3300 tons of CO2 annually. The bioenergy project contributes toward reducing CO2 emissions by 3300 tons/y 60% CO2 reduction per capita per year. It has already reached the CO2ereduction aims of the European Union for 2005 (Figures 14.5 and 14.6).

305

Current distributed renewable energy in rural and urban communities

Figure 14.5 Bioenergy village J€ uhnde, Germany, with energy generation installations. (From www.hna.de/lokales/hann-muenden/dransfeld-ort312904/bioenergie-dorf-juehndean-eam-verkauft-13046085.html 2019). Public electric grid Liquid manure Biogas Electricity Electricity Combined heat and power station (CHP)

Village heating grid

Anaerobic digestion plants

Crops from arable land

Central heating plant Wood chips

Figure 14.6 Technical concept of the project. CHP, combined heat and power generator. (From Ruwisch and Sauer, 2007 and IZNE (2005)).

J€ uhnde is the first bioenergy village in Germany, and even worldwide. Heat and power demand is produced by a renewable energy biogas plant as well with PV technology. It was founded in 2005 and attracted global attention with the unique idea. In Germany alone, 150 villages followed this example. After 10 years of operation, it was time to update the plant with the latest state of technology and make the most of opportunities of the EEG. The company ETS Economic Trading Solutions GmbH developed a sustainable end-to-end concept called J€ uhnde-Bioenergiedorf 2.0. The first step in the overall concept was the installation of the ultrasound disintegration plant from Weber Entec in Sep. 2015 (https://www.energie-und-management. de/nachrichten/erneuerbare/detail/juehnde-ein-vorzeigeprojekt-geraet-infinanznot-128100; https://www.energie-und management.de/filestore/ newsimgorg/.

306

N. El Bassam

14.6 Containerized solar minigrid, Fanidiama village, Mali Fanidiama is a small village located 462 km from Bamako, the Malian capital. It is known for the cultivation of cotton, considered locally to be white gold, but access to electricity is problematic. The national electricity grid does not supply the inhabitants. However, access to electricity is becoming a possibility in this region located on the border with Burkina Faso. Confounded by Thorsten Schreiber and his wife Aïda N’Daye, of Malian origin, the company Africa GreenTec is installing a containerized solar system there. It consists of solar panels installed on a container, capable of producing 45 kW peak (kWp) of electricity. The installation is connected to a battery system with a storage capacity of 60 kWh. Thus, inhabitants of Fanidiama will continue to benefit from electricity after sunset. Similarly, all solar containers are connected to the Internet by satellite, which ensures a good Web connection onsite and better remote management (Afrik2, https://www.afrik21.africa/en/mali-africa-greentec-installs-containerisedsolar-mini-grid-in-fanidiama/).The 54 kWp produced by Fanidiama’s small containerized solar power plant will be sold to residents at an affordable price. The revenues will be used to pay for the operating and maintenance costs of the solar container and to reimburse investment costs (Figure 14.7).

Figure 14.7 Minisolar containerized grid in Fanidiama village, Mali. Equipped with batteries, the containerized minigrid has a capacity of 45 kWp. (From https://www.afrik21. africa/en/mali-africa-greentec-installs-containerised-solar-mini-grid-in-fanidiama/.By Jean Marie Takouleu, 2019).

Current distributed renewable energy in rural and urban communities

307

In the medium term, we will also develop solutions in other countries with the hope of becoming the largest decentralized energy and water supplier in Africa. Owing to its mobility, the containerized solar minigrid is an efficient and fast solution for access to electricity in remote locations in Africa. The possibility of moving the container according to local contingencies (new connections of the village or problems of insecurity) also makes it possible to secure investments and deploy this solution in all kinds of territories. This is an approach increasingly being explored by solar off-grid companies in East Africa in particular. The French company Engie commissioned a 28-kW containerized solar minigrid in the village of Chitandika in southwest Zambia. Several other companies are investing in these off-grid systems. This is the case with the French industrial group Schneider Electric or Aptech Africa, a company based in Kampala, Uganda.

14.7 Decentralized desalination systems powered by solar energy in Maasai, Tanzania In some hard-to-reach areas, the best way to provide freshwater to people is through small, decentralized systems. For example, many companies and organizations are opting to deploy containerized desalination systems powered by solar energy. The small-scale desalination plants are mobile and easy to deploy in rural areas. A German supplier of solutions for autonomous power supply systems, Phaesun, installed two small solar-powered desalination systems in Ndedo, a village in the Maasai community in Tanzania. These small plants provide 2 m3/d of drinking water to the local population. A containerized system was also installed in the municipality of Hessequa, located more than 3 h drive from Cape Town, South Africa. Running on solar energy, the plant was installed in 2018 by French startup Mascara Renewable Water; it has a production capacity of 300 m3/d (Figure 14.8). Mascara Renewable Water intends to spread its solution to the rest of Africa. In 2019, the startup has formed a partnership with Vergnet Hydro, a French company that builds drinking water supply systems in Africa. The two companies want to process brackish water in countries such as Mozambique, Burkina Faso, Guinea, Mali, Mauritania, Niger, the Democratic Republic of Congo (DRC), and Senegal. As a result, it is not absolutely necessary to have access to the sea to desalinate water. Even so, the 35 African countries with a seafront are obviously the first concerned with the desalination option: Algeria, Angola, Benin,

308

N. El Bassam

Figure 14.8 A containerized solar-powered desalination system. ©Mascara Renewable Water By Jean Marie Takouleu. Published in 2019/modified in 2020 (https://www. osmosunwater.solutions/en/).

Cameroon, Cape Verde, Comoros, C^ ote d’Ivoire, DRC, Djibouti, Egypt, Eritrea, Equatorial Guinea, Gabon, Gambia, Ghana, Guinea Bissau, Kenya, Liberia, Libya, Madagascar, Mauritania, Morocco, Mozambique, Namibia, Nigeria, Republic of Congo, Sao Tome and Principe, Senegal, Sierra Leone, Somalia, South Africa, Sudan, Tanzania, Togo, and Tunisia ( Jean Marie Takouleu, https://www.unicef.fr/article/21-milliards-depersonnes-n-ont-pas-acces-l-eau-potable-salubre). Unlike conventional desalination technologies (thermal or reverse osmosis), electrodialysis requires little maintenance and is well-suited to solar power owing to its low energy consumption. Thus, a completely autonomous system that requires no additional infrastructure or power supply could be developed (https://www.ruralelec.org/business-opportunities/mascararenewable-water).

14.8 Road map to renewable energy in remote communities in Australia In late 2009, the Chief Minister of the Northern Territory (NT) formed the Green Energy Taskforce and charged it with developing a road map for the development of the renewable and low emission energy sector and products in the Territory. The NT has a strong history of renewable and low-emissions energy production in urban and remote areas and the ability to expand the number of renewable and low-emission energy systems deployed. NT organizations with experience and expertise in researching and implementing renewable energy include public sector entities, private sector companies, not-for-profit organizations, and academic institutions.

Current distributed renewable energy in rural and urban communities

309

Until this point, a total of 669 renewable energy generating systems had been funded in the NT through the Renewable Remote Power Generation Program (RRPGP), an Australian government program that closed in 2009. Many systems are small in scale but large-scale renewable energy systems have also been deployed across Yuendumu, Hermannsburg, Lajamanu, Kings Canyon, Bulman, and Jilkminggan. Three new PV systems deployed under the RRPGP at Ti Tree, Kalkarindji, and Alpurrurulam in 2011 are more advanced in terms of the dieselePV hybrid configuration and a high level of PV penetration (i.e., replacement of diesel power with renewable solar power). Lessons can be learned from this project to support future PV deployment and broader expansion of the initiatives proposed. Renewable energy technologies currently available or deployable in the NT are PV, wind, and solar thermal systems. Although PV systems are technically proven and are becoming more economic (with falling PV panel prices), the primary issue for solar thermal is that the economic scale is beyond what is being considered for remote communities. The wind resource in the Territory is limited, and although other resources such as geothermal and tidal energy are abundant, the applicable technologies have not yet been commercialized. Biofuels have been commercialized to a limited extent. The NT government, through its Climate Change Policy, has committed to replacing diesel as a primary source of power generation to remote towns and communities, developing a green energy industry, reducing greenhouse gas contributions, and assisting the Power and Water Corporation (PWC) in meeting its Renewable Energy Target (RET) obligations through local sources of renewable energy. To identify how this can occur, the Green Energy Taskforce was established by the Chief Minister to provide expert advice (including their Road Map Report) on strategies, incentives, and pathways to develop renewable and low-emission energy and products in the NT. The first major task identified in the Terms of Reference for the Taskforce was to develop a proposal for substituting a large component of diesel generation with low emissions and renewable energy in remote communities by 2020. However, the Road Map is broader than just electricity generation; it identifies opportunities for skills transfer, training, and management in communities, Indigenous economic development, closing the gap of Indigenous disadvantage, and remote service delivery reform under the Working Futures policy, through the development of a low-emissions and renewable energy industry in the NT.

310

N. El Bassam

The first component is to roll out 10 MW of PV to 46 remote communities. This first step involved integrating solar PV capacity into 46 diesel power stations, to a level that produced significant savings in diesel but did not require expensive modifications to generators or installation of energy storage equipment or infrastructure (e.g., batteries, fly wheels). These communities do not have significant renewable energy systems, nor are there immediate plans to install them. A rollout of 10 MW of PV was proposed to reduce diesel use by 17%, supply approximately 30% of peak demand across the selected communities, and meet 7% of PWC’s cumulative RET obligation by 2030. This level of renewable energy penetration was recommended because it did not require expensive storage or modifications to existing power plants; as such, it represented the least-cost renewable energy option for significant diesel fuel savings. Variability in renewable energy output at this level was within performance thresholds of existing diesel generation plants. The technical and economic viability of other energy sources including pipeline natural gas, liquefied petroleum gas, biodiesel, compressed natural gas, liquefied natural gas, and very high-penetration PV will continue to be monitored by existing internal processes. However, to be successful in achieving 100% diesel substitution and achieve this target both efficiently and effectively will require the integration of a range of fuel sources and technologies (Figure 14.9). The second component is moving toward 100% substitution of diesel generation in NT remote communities. The second component of the Road Map proposes defining options for a more ambitious target of moving toward 100% substitution of diesel generation with renewable energy and low-emissions fuel in remote communities and will be pursued concurrently with the first component (rollout of 10 MW of renewable energy in 46 remote communities). Maximization of the effectiveness and potential of renewable energy in off-grid power stations will involve optimization to suit the characteristics of the renewable resource while ensuring the security of supply. Understanding how to plan and achieve 100% renewable and low-emission energy generation is a prerequisite to building sustainable remote communities in the NT. Understanding how to address immediate targets such as diesel substitution as well as the longer-term ability to secure the Territory’s energy future is the basis for this component of the Road Map. toward 100% substitution of diesel generation with renewable energy and low-emissions fuel in remote communities, These will be pursued concurrently with the first

Current distributed renewable energy in rural and urban communities

311

Figure 14.9 Schematic of conventional power station versus stabilized renewable energy power station. (From Northern Territory Government Australia: Road Map to Renewable and Low Emission Energy in Remote Communities, 2017).

component (rollout of 10 MW of renewable energy in 46 remote communities). Australia has a substantial renewable energy resource potentially capable of providing 500 times the electricity currently used in Australia. It has been estimated that growth in renewables worldwide would see a corresponding growth in Australia’s economy of 1.7% above business as usual, even factoring in a decline in coal exports (NT Road Map to Renewables Report, NT Government, 2017; https://www.isacnt.org.au/NTindustries/renewable-energy).

312

N. El Bassam

Figure 14.10 Sustainable community within Iraq. (From Iraq Dream Homes, LLC (IDH) 2010).

14.9 Iraq Dream Homes Iraq Dream Homes, LLC (IDH) was established in 2010 to target the enormous potential to develop low-cost and affordable housing projects in Iraq. IDH is a project company of One Alliance Partners (OAP), which is composed of several related companies providing an integrated management approach to property development, infrastructure development and renewable energy development projects globally (Figure 14.10). OAP is focused and aligned with specific projects in the Middle East, Southeast Asia, and the United States. Iraq Energy Solutions, LLC (IES), another subsidiary of OAP, launched a renewable energy development initiative to work in concert with IDH’s focus on building sustainable communities within Iraq and will provide efficient energy production through privatized distribution (Figure 14.11). IDH’s managing directors have more than 53 years of professional experience in the fields of project and construction management, architecture, planning and urban design, and real estate development. The multidisciplinary background of the company’s key staff empowers it to provide

Figure 14.11 Solar panels power street lights in Fallujah, Iraq. (From U.S Army, 2010).

Current distributed renewable energy in rural and urban communities

313

excellent services in all aspects of land and real estate development, infrastructure construction, large scale planning, project management, and general construction management (Iraq Dream Homes, LLC (IDH), 2010).

14.10 Renewables in Africa The developing nations of Africa are popular locations for the application of renewable energy technology. Many nations already have smallscale solar, wind, and geothermal devices in operation providing energy to urban and rural populations. These types of energy production are especially useful in remote locations because of the excessive cost of transporting electricity from large-scale power plants. The applications of renewable energy technology has the potential to alleviate many of the problems that face Africans every day, especially when done in a sustainable manner that prioritizes human rights (https://en.wikipedia.org/wiki/Renewable_energy_in_ Africa 2020) (Figures 14.12 and 14.13).

Figure 14.12 Distribution of identified renewable energy potential in Africa. (From International Renewable Energy Agency, 2013, https://www.researchgate.net/figure/ Distribution-of-identified-renewable-energy-potential-in-Africa-Source-International_fig4_ 299418273 based on the Global Atlas).

314

N. El Bassam

Figure 14.13 Powering climate action in Africa with Renewable Energy Africa largescale solar projects tracker Ghana. (From renewablesinafrica.com).

Africa’s rapid economic expansion creates a daunting energy challenge, combined with rising expectations of improved resilience and sustainability. Finding a sustainable way to meet growing energy needs is one of the core development challenges for the continent. Africa is rich in renewable energy sources, including hydro, sun, and wind. The time is right for sound planning to ensure the right energy mix. Decisions made today will shape the continent’s energy sector for decades (https://www.irena.org/newsroom/ articles/2019/Mar/Powering-Climate-Action-in-Africa-with-RenewableEnergy). Access to energy is essential to reduce poverty and promote economic growth. Communication technologies, education, industrialization, agricultural improvement, and expansion of municipal water systems all require abundant, reliable, and cost-effective energy access. For African countries, economic and social development cannot be achieved in the absence of adequate energy supplies. Although reliable energy services are not sufficient in themselves to eradicate extreme poverty, they are necessary to create the conditions for economic growth and improve social equality. Access to modern energy services can help reduce spending; because of the inefficiency of traditional energy forms, the poor often pay higher unit costs for energy than do the rich. In addition, women in most parts of Africa spend large amounts of time fetching wood and water, losing time that they could otherwise devote to schooling and revenue-generating activities. Furthermore, indoor air pollution from low-quality solid cooking fuels (wood, charcoal, dung, and waste) imposes a major health hazard. In view of their modular nature and availability at the local level, renewable energy technologies can contribute to sustainable development by increasing access to modern energy services to the majority in rural areas.

Current distributed renewable energy in rural and urban communities

315

Energy access and quality in sub-Saharan Africa (SSA) is relatively low with high costs: • 560 million people do not have access to electricity (74%) compared with 28% in other developing countries. • 625 million people do not have access to modern fuels (83%) compared with 33% in other developing countries. • Two million deaths per year (globally) are caused by indoor air pollution (fuel burning for cooking and heat), and the burden of disease in Africa is particularly high. • In 2009, South Africa produced 70% of all SSA’s electricity. • Subsidies have failed to bring energy prices down and are not sustainable. There are many renewable energy projects in SSA.

14.10.1 Hydropower Only about 5% of Africa’s hydropower potential of just over 1750 TWh has been exploited. The total hydropower potential for Africa is equivalent to the total electricity consumed in France, Germany, the United Kingdom, and Italy together. The Inga River in the DRC holds great potential for hydropower generation in Africa, with an estimated potential of around 40,000 MW. In fact, the DRC alone accounts for over 50% of Africa’s hydropower potential; other countries with significant hydropower potential include Angola, Cameroon, Egypt, Ethiopia, Gabon, Madagascar, Mozambique, Niger, and Zambia. Despite the low percent use, largescale hydropower to date provides over 50% of the total power supply for 23 countries in Africa. Small hydropower systems (less than 10 MW) can supply energy to remote communities and catalyze development in such communities. Compared with large hydropower systems, smaller systems require significantly lower capital costs. This allows increased local private sector participation and community involvement. Most African countries have a large potential for small hydro-systems and some are already exploiting it with a special focus on rural communities.

14.10.2 Biomass Several agro-based industries in the continent, such as wood-based industries, palm and rice mills, sugar, and paper and pulp, use their waste to produce both process heat and power, which in most cases is used locally. Cogeneration from agriculture waste holds great potential for Africa. Cogeneration contributes as much as 40% of the total electricity generated in Mauritius.

316

N. El Bassam

Dissemination of biogas digesters for household applications has not been successful owing to high capital costs, insufficient feedstock and water, high labor demand, and negative public perception, among other reasons. Countries such as Ghana, Kenya, Niger, Burkina Faso, Mali, Ethiopia, Senegal, and Rwanda have implemented pilot projects aimed at establishing the technical and socioeconomic viability of biogas technology as an alternative source of energy for cooking and decentralized rural electrification. In the case of Ghana, the Appolonia project installed a system that generates 12.5 kW electric power, which is fed into a local grid, supplying electricity at 230 V for domestic use to 21 houses, street lighting, and five social centers in the community. The biogas is produced from cow dung and human excreta. Two diesel engines of 8 kW each were modified to operate on a dual fuel (a mixture of biogas and diesel). Project results show that the dieselebiogas system saves 66% in diesel consumption compared with pure diesel generation. The dissemination of biogas digesters to institutions is promising; the private sector already leads the dissemination of biogas digesters in countries such as Ghana, Rwanda, and Tanzania. Most sub-Saharan households rely primarily on wood fuel for cooking and heating. Wood fuel is the main source of fuel in rural areas whereas charcoal is commonly used in the poorer urban households. However, shortages of alternative energy sources including electricity blackouts and brownouts often force even the better-off households to use charcoal. As a response to fuel wood shortages, improved biomass cook stoves have been promoted throughout Africa.

14.10.3 Geothermal Geothermal energy is an untapped renewable energy source that is abundantly present in many parts of Africa. It has the potential to generate up to 14,000 MW from geothermal sources. However, only a few countries such as Kenya have used it commercially. To date, Kenya has installed up to 127 MW, amounting to about 17% of the national power supply, followed by Ethiopia with a 7-MW installation. Plans to use potential of geothermal energy in Uganda, Tanzania, and Eritrea are at different stages.

14.10.4 Wind power In terms of installed capacity at the beginning of 2008, Africa only had about 476 MW of installed wind energy generation capacity, compared with a global estimate of 93,900 MW. Countries developing large-scale wind energy projects to date include Morocco, Egypt, Tunisia, South Africa, and Ethiopia.

Current distributed renewable energy in rural and urban communities

317

14.10.5 Solar power Large-scale solar energy projects are limited in Africa because of cost constraints. Detailed feasibility studies established that Africa has great potential for concentrated solar thermal power generation from desert areas such as the Sahara and Namib, with competitive power production costs around 4e6 c/kWh. To date, only South Africa operates a solar thermal power system plant, generating 0.5 MW. Egypt planned to install solar thermal plants of 30 MW by 2010 and 300 MW by 2020. Several countries in Northern Africa planned to develop solar thermal plants of varying capacities, buoyed by interest from European countries (Figure 14.14). There are two types of small-scale energy systems. The first category produces electricity based on PV and wind power, for instance, whereas the second produces thermal energy for heating, drying, and cooking. Solar home systems in the household sector are by far the most common application of this kind of system. South Africa and Kenya have some of the highest documented installed capacities of solar PV systems that stand at over 11,000 and 3600 kWp, respectively. Unfortunately, poor households have not benefited as much as high-income households from solar PV systems because of the high upfront costs. Some solar thermal systems have been disseminated for water heating and solar cookers. Currently, solar water heaters are predominantly used in Eastern, Southern, and Northern Africa for household application and in the hospitality industry.

Figure 14.14 Solar water pumping and lighting station in Mali. (From Togola, 2011).

318

N. El Bassam

14.10.6 Biofuels Various studies have estimated that the potential for sustainable biofuel production (production that preserves biodiversity, rainforests, and water resources, and does not endanger food security) in 2050 for SSA ranges from 41 to 410 EJ (Smeets et al., 2006). The lower range of this estimate is more than twice the total amount of energy that was consumed in Africa in 2008, about 19 EJ. About 39 countries in Africa are net oil importers. Therefore, the development of biofuels will reduce dependency on imported fuels. Agriculture is the mainstay of most countries in Africa, so the development of biofuels in Africa, especially in rural areas that have the land to use, could bring many potential benefits including increased access to electricity, transformation of the rural economy because of the availability of reliable energy supplies, and employment opportunities. Zimbabwe, Kenya, and Malawi have ethanol programs in which ethanol produced as a by-product from sugar industries is blended with petrol and used as transport fuel. Zimbabwe is the only one of the three countries, however, to mandate that ethanol be blended with all gasoline sold and produced up to 40 million liters of ethanol per year. The Kenya plant was closed in 1992 owing to a lack of viability. At its peak, it produced up to 45,000 L and some rural employment was achieved. Other efforts to develop large-scale biofuels include palm oil and cassavabased ethanol in West Africa and Jatropha-based biodiesel in Mali, Tanzania. Small-scale biofuel production in Africa is so far primarily based on Jatropha biodiesel. Most countries in Africa have ongoing programs to grow Jatropha in rural areas to produce biodiesel.

14.10.7 Energy efficiency Countries in Africa have considerable scope for increasing energy efficiency on the energy supply side and reducing energy consumption on the demand side without decreasing economic output, lowering the standards of living, or diminishing the quantity and quality of social services provided. Studies by the International Energy Agency show that in Africa, energy intensity (i.e., total energy consumed per gross domestic product) is at least twice the global average. However, energy efficiency continues to be a peripheral issue in overall energy sector planning and development in Africa.

Current distributed renewable energy in rural and urban communities

319

14.11 Renewables in India India is one of the countries with large production of energy from renewable sources. As of Mar. 31, 2020, 35.86% of India’s installed electricity generation capacity was from renewable sources, generating 21.22% of total utility electricity in the country. In the Paris Agreement, India committed to an Intended Nationally Determined Contributions target of achieving 40% of its total electricity generation from nonefossil fuel sources by 2030. The country aimed for an even more ambitious target of 57% of the total electricity capacity from renewable sources by 2027 in the Central Electricity Authority’s strategy blueprint. According to the 2027 blueprint, India aimed to have 275 GW from renewable energy, 72 GW of hydroelectricity, 15 GW of nuclear energy, and nearly 100 GW from other zeroemission sources. In the quarter ending Sep. 2019, India’s total renewable electricity capacity (including large hydro) was 130.68 GW. This represents 35.7% of the total installed electricity generation capacity in the country, which was around 366 GW. The government of India also set a target for installation of Rooftop Solar Projects of 40 GW by 2022, including installations on rooftops of houses (Figure 14.15). As of Oct. 2019, of the 175-GW interim target, 83 GW was operational, 29 was under installation, 30 GW was under bidding, and the remaining 43 GW was under planning. Moreover, the 175-GW interim target was 100 GW of solar, 60 GW of wind, 10 GW of biomass, and 5 GW of small hydro. As of 2019, 35% total power production came from renewable

Figure 14.15 The largest wind farm of India in Muppandal, Tamil Nadu.

320

N. El Bassam

energy, 13% or 45.399 GW of the total from all sources came from large hydro projects, 10% or 36,686.82 GW of the total from all sources came from wind power (the fourth largest in the world), and 8% or 9.1 GW of total power from all sources from biomass power came from biomass combustion, biomass gasification, and bagasse cogeneration (Indian Renewable Energy Industry Report, March 2020, https://www.ibef.org/industry/ renewable-energy.aspx). India was the first country to set up a ministry of nonconventional energy resources (Ministry of New and Renewable Energy [MNRE]), in the early 1980s. Its public sector undertaking, the Solar Energy Corporation of India, is responsible for developing solar energy industry in India. Hydroelectricity is administered separately by the Ministry of Power and is not included in MNRE targets (Figure 14.16). India has a strong manufacturing base in wind power, with 20 manufacturers of 53 different wind turbine models of international quality up to 3 MW in size, and with exports to Europe, the United States, and other countries. Wind or solar PV paired with 4-h battery storage systems are cost-competitive, without subsidy, as a source of dispatchable generation compared with new coal and new gas plants in India. From https://www. google.com/search?client¼firefox-bd&q¼RENEWABLESþINþINDIA.

14.12 Distributed renewable energy and solar oases for deserts and arid regions: the DESERTEC concept The DESERTEC concept of pioneer Gerhard Knies aimed to promote the generation of electricity in Northern Africa, the Middle East, and Europe using solar power plants, wind parks, and the transmission of

Figure 14.16 Solar power plant Telangana II in the state of Telangana, India.

Current distributed renewable energy in rural and urban communities

321

this electricity to consumption centers, promoted by the nonprofit DESERTEC Foundation. Despite its name, DESERTEC’s proposal would see most of the power plants located outside the Sahara Desert itself, in the more accessible southern and northern steppes and woodlands, as well as the relatively moist Atlantic coastal desert. The original and first region for the assessment and application of this concept is the Europe, Middle East, and Northern Africa (EU-MENA) region. Under the DESERTEC proposal, concentrating solar power systems, PV systems, and wind parks would be spread over the desert regions in Northern Africa such as the Sahara Desert. Produced electricity mainly by concentrated thermal power (CSP) and would use high-voltage DC cables. It would provide a considerable part of the electricity demand of MENA countries and provide continental Europe with 15% of its electricity needs by 2050. Investments into solar power plants and transmission lines to Europe would cost a total of V400 billion (Figure 14.17). The project of realizing the DESERTEC concept in EU-MENA was developed by Dii GmbH, a consortium of European and Algerian companies founded in Munich and led by Munich Re. is incorporated under German law. The consortium consists of the DESERTEC Foundation,

Figure 14.17 Possible infrastructure for a sustainable supply of power to Europe, the Middle East, and North Africa (EU-MENA). (From Euro-Supergrid with a EU-MENAConnection Proposed by TREC).

322

N. El Bassam

Munich Re., Deutsche Bank, Siemens, ABB, E.ON, RWE, Abengoa Solar, Cevital, HSH Nordbank, M and W Zander Holding, MAN Solar Millennium, and Schott Solar. Press investigations point to a number of more interested parties, among them ENEL, E’lectricité de France, Red Eléctrica de Espa~ na and companies from Morocco, Tunisia, and Egypt (Wikipedia, 2012). Main scientific data were delivered in three research studies by an international network of scientists, experts, and politicians from the field of renewable energies coordinated by Dr. Franz Trieb, German Aerospace Center (DLR), conducted in three research studies to which the author also contributed: 1. Concentrating Solar Power for the Mediterranean Region 2. Trans-Mediterranean Interconnection 3. Concentrating Solar Power for Seawater Desalination Main findings were: Only 1e2% of the desert area could supply the whole world with electricity, and 1 km2 of desert land using CSP technology can harvest up to: • 250 million kWh/y of electricity (250 L oil/m2) and • 60 million m3/y of desalted water (6000 L water/m2); • Produced electricity would be transmitted to European and African countries by a supergrid of high-voltage direct current cables. • The expected load losses amount to 3% per 1000 km. (Highways for Renewable Electricity (Supergrid) EUMENA e TransMediterranean) (Figure 14.18).

Figure 14.18 Planned highways for the projected renewable electricity supergrid, high-voltage DC grid.

Current distributed renewable energy in rural and urban communities

323

14.13 Vatican City 14.13.1 And suddenly there was light! The least populated sovereign state in the world, with a population of just 800 people, Vatican City has been declared the world’s most environmentally friendly state, with the installation of giant solar power panels. The city reached a record in solar energy production per capita with a figure of 200 W per inhabitant. In 2008, SolarWorld AG completed the first solar power plant for the Vatican right next to St. Peter’s Cathedral. To date, some 2394 solar modules generate electricity on the roof of the Papal audience hall (Solar World AG, 2010). The solar power plant on the roof of the Paolo VI audience hall has a peak total output of 221.59 kW, enough to generate some 300,000 kWh of electricity. This is equivalent to the annual needs of more than 100 households. The generation of this volume of clean energy is designed to avoid the emission of 225,000 kg of CO2. The aesthetically sophisticated plant was blended into the historical ensemble of Vatican City with a great deal of technical and architectural effort. The installation of solar panels on the Paul IV conference hall has saved the Vatican 89.84 tons of oil equivalent. The Vatican is not exactly a large state, so its proposed solar plant will generate enough energy to power all of its 40,000 households. The installation will be located on a 740-acre site near Santa Maria di Galeriad, where the Vatican Radio’s transmission tower is located. The energy that it produces will be far beyond the needs of the entire Vatican, providing enough power to meet the needs of Vatican radio nine times over (Figures 14.19 and 14.20).

Figure 14.19 The Vatican City received its first solar power plant in 2008. (From Jolly, A. 2010. Vatican City Becomes the Greenest State in the World. MSN News).

324

N. El Bassam

Figure 14.20 Powered by sun energy: 2394 modules, 300 MWh/y, and less 225 t/y/of emitted CO2. (From Solarworld Group, 2008).

These are not the only steps the Vatican is taking to reduce its greenhouse emissions. It is contemplating using an electric Pope mobile, the Vatican cafeteria will soon be decked with a solar heating system to provide heating and cooling, and even the Pope’s summer residence is being fitted out to obtain power from methane generated by the horse stables. The director of Vatican Radio has declared: “When looking for inspiration, the Pope clearly defers toward the heavens, but when looking for electricity, the sun is his choice.”

References Allen, C., 2011. German Village Achieves Energy Independence and Then Some [WWW] BioCycle. Available from: http://www.infiniteunknown.net/2011/08/22/germanvillage-wildpoldsried-pop-2600-produces-321-percent-more-energy-than-it-needs-andis-generating-5-7-million-in-annual-revenue/. Bloomberg, L.P., 2012. New Energy Finance [WWW]. Available from: http://www. newenergyfinance.com/PressReleases/view/186. Boston Consulting Group, 2009. Electricity Storage Making Large-Scale Adoption of Wind and Solar Energies a Reality, 106-41977. http://www.bcg.no/expertise_impact/ Industries/Energy_Environment/PublicationDetails.aspx?id tcm. Danish Energy Agency, Heat Supply: Goals and Means Over the Years. http://dbdh.dk/ images/uploads/pdf-diverse/varmeforsyning%20i%20DK%20p%C3%A5%20engelsk.pdf. El Bassam, N., 2010. In: Handbook of Bioenergy Crops: A Complete Reference to Species Development and Applications, second ed. Taylor & Francis Group Ltd, Oxford, Routledge. El Bassam, N., Maegaard, P., 2004. Integrated Renewable Energy for Rural Communities, Planning Guidelines, Technologies and Applications. Elsevier, The Netherlands. Etcheverry, J., 2008. Challenges and Opportunities for Implementing Sustainable Energy Strategies in Coastal Communities of Baja California Sur. University of Toronto, Mexico. Federal Ministry of Food, Agriculture and Consumer Protection, Berlin Office 11055 Berlin [WWW]. Available from: http://www.wege-zum-bioenergiedorf.de/ metanavigation/datenschutz/. Iraq Dream Homes, 2010. LLC (IDH) [WWW]. Available from: http://www.iraq-homes. com/about_us-en.html.

Current distributed renewable energy in rural and urban communities

325

IZNE, 2005. The Bioenergy Village Self-Sufficient Heating and Electricity Supply Using Biomass. www.bioenergiedorf.de. IZNE, 2007. WVarme- und Stromversorgung durch heimische Biomasse April 2007, InterdisziplinVares Zentrum fuVr Nachhaltige Entwicklung (IZNE) der UniversitVat GoVttingen. - Projektgruppe BioenergiedoVrfer. http://www.gar-bw.de/fileadmin/gar/ pdf/Energie_ und_Klima/Juehnde.pdf. Lund, H., 2010. Renewable Energy Systems: The Choice and Modeling of 100% Renewable Solutions. Academic Press. Maegaard, P., 2008. Denmark: Wind Leader in Stand-By. International Sustainable Energy Review. http://xplqa30.ieee.org/xpl/references.jsp?arnumber 5673216. Maegaard, P., 2009. Danish Renewable Energy Policy, Article. WCRE. org Homepage September 2009. Maegaard, P., 2009/2010. Thisted: 100% Renewable Energy Municipality. Powerpoint presentation. Maegaard, P., 2010. Wind energy development and application prospects, of non-gridconnected wind power. In: Proceedings of 2009 World Non-Grid-Connected Wind Power and Energy Conference. IEEE Press. Marshall, M., 2012. New Scientist issue 2850, accessed from. http://www.newscientist.com/ article/mg21328505.000-indias-panel-price-crash-could-spark-solar-revolution.html. Northern Territory Government Australia: Roadmap to Renewable and Low Emission Energy in Remote Communities Report. http://www.greeningnt.nt.gov.au/climate/ docs/Renewable_Energy_Report_FA.pdf 2011. Maegaard, P., June 2011. Integrated Systems to Reduce Global Warming 2011. Springer Science Business Media. LLC 22 PDF. http://www.springer.com/about springer/ locations worldwide?SGWID 4-173904-2052-653447-150. Ruby, J.U., 2006. Continuous Headwind e Pioneering the Transition from Fossil Fuels and Atomic Energy to the Renewable Energies. Hovedland Danish. Ruwisch, V., Sauer, B., 2007. Bioenergy Village JuVhnde: Experiences in Rural SelfSufficiency.I. IZNE. www.bioenergiedorf.de. Scheer, H., 2006. Energy Autonomy: The Economic, Social and Technological Case for Renewable Energy. Earthscan 2006. Smeets, E., Faaj, A., Lewandowski, I., 2006. Progress in energy and combustion science. In: A Quick Scan of Global Bioenergy Potentials to 2050. El Bassam, N., 2009. Solar Oases: Transformation of Deserts into Gardens, www.ifeed.org. SolarWorld Group, 2008. Solar Power for the Vatican: Inauguration of the First Solar Power Plant for the Papal State. Bonn [WWW]. Available from: http://solarworld-usa. com/news-and-resources/news/vatican-inauguration-solar-power-for-papal-state.aspx. Workshop, Strategy, June 2011. Green Settlements in Sub Saharan Africa: Future Land Use and Pathways to Wealth Creation. Loccum Protestant Academy, Germany. The Poul la Cour Foundation, 2009. Wind Power, the Danish Way, from Poul la Cour to Modern Wind Turbines. Self-Published, Denmark. Toke, D., 2007. Evaluation and Recommendations. DESIRE Project e Dissemination Strategy on Electricity Balancing Large Scale Integration of Renewable Energy. Trieb, F., 2005. MED-CSP, Final Reports. German Aerospace Center (DLR). Institute of Technical Thermodynamics, Section Systems Analysis and Technology Assessment, Study Commissioned by: Federal Ministry for the Environment, Nature Conservation and Nuclear Safety,. Institute of Technical Thermodynamics, Section Systems Analysis and Technology Assessment, Germany. www.dlr.de/tt/med-csp. Trieb, F., 2006. TRANS-CSP. Final Reports. German Aerospace Center (DLR). Institute of Technical Thermodynamics, Section Systems Analysis and Technology Assessment, Study Commissioned by: Federal Ministry for the Environment, Nature Conservation

326

N. El Bassam

and Nuclear Safety,. Institute of Technical Thermodynamics, Section Systems Analysis and Technology Assessment, Germany. www.dlr.de/tt/trans-csp. Trieb, F., 2007. AQUA-CSP, Final Reports. German Aerospace Center (DLR). Institute of Technical Thermodynamics, Section Systems Analysis and Technology Assessment, Study Commissioned by: Federal Ministry for the Environment, Nature Conservation and Nuclear Safety. Institute of Technical Ther- modynamics, Section Systems Analysis and Technology Assessment, Germany http://www.dlr. de/tt/aqua-csp.

Further reading Abramsky, K., 2010. Sparking a Worldwide Energy Revolution: Social Struggles in the Transition to a Post-Petrol World. AK Press, Oakland. El Bassam, N., Maegaard, P., 2004. Integrated Renewable Energy for Rural Communities, Planning Guidelines, Technologies and Applications. Elsevier, The Netherlands. Etcheverry, J., 2008. Challenges and Opportunities for Implementing Sustainable Energy Strategies in Coastal Communities of Baja California Sur. University of Toronto, Mexico. Lund, H., 2010. Renewable Energy Systems: The Choice and Modeling of 100% Renewable Solutions. Academic Press. Maegaard, P., 2008. Denmark: Wind Leader in Stand-By. International Sustainable Energy Review. Maegaard, P., 2009. Danish Renewable Energy Policy, Article. WCRE. org Homepage September 2009. Maegaard, P., 2009/2010. Thisted: 100% Renewable Energy Municipality. Powerpoint presentation. Maegaard, P., 2010. Wind energy development and application prospects of non-grid- connected wind power. In: Proceedings of 2009 World Non-Grid-Connected Wind Power and Energy Conference. IEEE Press. Ruby, J.U., 2006. Continuous Headwind e Pioneering the Transition From Fossil Fuels and Atomic Energy to the Renewable Energies. Hovedland (in Danish language). Scheer, H., 2006. Energy Autonomy: The Economic, Social and Technological Case for Renewable Energy. Earthscan 2006. Smeets, E., Faaj, A., Lewandowski, I., 2006. Progress in energy and combustion science. In: A Quick Scan of Global Bioenergy Potentials to 2050.

CHAPTER FIFTEEN

Ownership, citizens participation and economic trends N. El Bassam International Research Centre for Renewable Energy, IFFED.org, Germany

Contents 15.1 Community ownership 15.1.1 Benefits of community energy 15.2 Citizens’ participation 15.3 The Danish ownership model 15.4 Integration of the energy supply by public ownership 15.5 Economic impacts 15.6 Socioeconomic benefits and economic impacts of Renewables 2019 Renewable Generation Capacity by Region. 15.7 Actions for broadening the ownership of renewables 15.8 Global investment’s in renewables 15.9 Costs of renewables References Further reading

327 328 328 330 333 334 335 338 339 340 341 343 343

15.1 Community ownership Community energy is the economic and operational participation and/or ownership by citizens or members of a defined community in a renewable energy project. Community energy is not limited by size; it takes place on both large and small scales. Since 2011, the World Wind Energy Association (WWEA) has used a definition developed by members of its Community Power Working Group (WWEA, 2016) and supported by the Community Energy Group of the Coalition for Action. Community energy is any combination of at least two of the following elements: • Local stakeholders own most or all of a renewable energy project. • Voting control rests with a community-based organization. • Most social and economic benefits are distributed locally. Distributed Renewable Energies for Off-Grid Communities, Volume 2 ISBN: 978-0-12-821605-7 https://doi.org/10.1016/B978-0-12-821605-7.00019-2

© 2021 Elsevier Inc. All rights reserved.

327

j

328

N. El Bassam

However, various definitions of community energy are found worldwide, depending on a government’s intention to steer investment and ownership in renewable energy generation in this direction. Requirements for communities to qualify as a community energy project may be more or less stringent, depending on the respective policy’s actual intention to democratize the energy assets and create a distributed energy system. This makes global stocktaking of community energy projects difficult. Community energy projects may differ in size and scope. They may include minority and partial ownership by a few members of a municipality in a project, the generation assets owned by a cooperative whose shareholders are solely recruited from the project-hosting community, or communities developing their own energy autarkic community centers or municipal and citizen partnerships (WWEA, 2016).

15.1.1 Benefits of community energy A community’s economic and operational participation in renewable energy projects is a key factor for building community acceptance and support for the development of renewable energy projects (REN21, 2017; WWEA, 2016). Additional benefits of community energy can include: • added value for the region through the establishment of a new economic sector, job creation, and a local identity; • an increase in actor diversity resulting in shared decision-making and increased transparency in planning and construction; • integration of citizens into sustainable economic processes; • lower energy prices; • acceleration of energy access and general renewable deployment rates; and • technology and business model innovation.

15.2 Citizens’ participation What’s at stake? Community involvement in building the energy systems of the future is important for an effective and inclusive energy transformation. Local governments have a unique responsibility in facilitating the energy revolution by providing communities with the framework and support to build a local sustainable energy infrastructure. From improved building codes to district heating systems, more efficient lowcarbon public transit, and walkable neighborhoods, municipalities could be the drivers of sustainable and affordable energy supply for their communities. Communities could have a further pivotal role in the democratization

Ownership, citizens participation and economic trends

329

and decentralization of energy systems around the world. Decentralized energy systems are in many cases more resilient to climate change and disaster and more reliable than centralized systems. They also sustain fewer network losses and could create a pride of ownership that has far-reaching behavioral impacts on all sustainability activities. Contrary to the general perception, community-owned companies have also been proven to have the ability not only to provide quality service for their customers, but also to do it at lower energy prices than commercial energy companies (WWEA, 2018). One example is Slagslunde Fjernvarme A.m.b.a., a consumer-owned district heating company located 30 km northwest of Copenhagen, Denmark. Local district heating consumers decided to buy the district heating system from the utility and operate it themselves. The cooperative started to run the system in 2013. Within the following years, energy prices went down considerably (Danish Energy Regulatory Authority, 2018). In 2016, a total of V15.1 billion was invested in renewable energy facilities in Germany. Germans first set up renewable energy installations, such as photovoltaic (PV) panels and wind turbines in the 1990s. In 2000, the Renewable Energy Act was passed, guaranteeing fixed feed-in tariffs for anyone generating renewable power for 20 years. This further encouraged households to install PV panels on their roofs, either feeding the electricity they produced into the grid or consuming it themselves (Figures 15.1 and 15.2).

Figure 15.1 Ownership structure of photovoltaic systems.

330

N. El Bassam

Figure 15.2 Ownership structure in Germany fact sheet (2018). From Julian Wettengel, https://www.cleanenergywire.org/factsheets/citizens-participation-energiewende.

Citizens’ energy is often organized by specialized project developers. Many of these projects are formed on a local level. For example, local citizens investing their savings in a nearby wind farm. However, some are also regional or even interregional operations. These citizen-owned projects may include companies or municipalities among their members, but they are separate from municipal utilities. To distinguish among different types of involvement by citizens in the renewable sector, researchers from Leuphana University and Trend: Research defined different types of citizens’ energy in 2013 (Figure 15.3). In the narrower sense, citizens’ energy means: • private individuals or farmers (jointly or individually) invest in energy facilities; • the investment is made with their own capital, giving actors a certain level of control over the project; • citizens own at least 50% of voting shares; and • citizens have a connection to the region where the facility is operated. More than 90% of energy coop members are private individuals. The rest are companies, municipalities, or farmers.

15.3 The Danish ownership model Denmark already has hundreds of consumer-owned local energy supply companies for combined heat and power (CHP), district heating,

Ownership, citizens participation and economic trends

331

Figure 15.3 Membership structure of energy cooperatives in Germany, 2017. From https://www.cleanenergywire.org/factsheets/citizens-participation-energiewende.

and power distribution. New organizational structures and nonprofit ownership models for the direct benefit of the involved municipalities may prove to be the most realistic long-term solution for community wind power as well, to make it locally acceptable. The absence of financial investors early on made the wind sector in Denmark unique compared with other countries. At the turn of the century, around 150,000 households were co-owners of a local windmill. The ownership model, rather than technology and tariff schemes, was important to the success of wind energy in Denmark. It was the key factor behind the high public acceptance that wind power projects enjoyed during that time. It also enabled a much faster deployment, because large numbers of people were involved in the sector that provided the necessary good will. In 1992, the role of cooperatives shrank owing to a change in planning procedures. In 1998, as a result of liberalization in the sector, the ownership model changed dramatically. The residential criteria for ownership were abolished and everyone was allowed to own as many windmills as possible for which they could acquire permission anywhere in the country. The takeover bids began, resulting in a dramatic decrease in public involvement. As a result, in 2009, about 50,000 households were co-owners of windmills, compared with 150,000 a decad earlier.

332

N. El Bassam

Consequently, the attitude toward wind power suffered a reversal. The erection of a windmill often became a local drama and resulted in bitter conflicts that led to long delays or cancellations. In 2008, a wind turbine neighbor compensation scheme was introduced that also caused local conflicts. The solution seemed to be to normalize the wind energy sector so that wind power, like district heating, CHP and power distribution, changed from investment to public supply, with municipalities and consumer owner local companies as the future wind power owners. This would even make more wind power significantly cheaper. Denmark already has hundreds of consumer-owned local energy supply companies for CHP, district heating, and power distribution. New organizational structures and nonprofit ownership models for the direct benefit of the involved municipalities may prove to be the most realistic long-term solution for community wind power as well to make it locally acceptable again. Nontechnical aspects are also important for a successful transition to renewable energy supply. In the beginning, the absence of financial investors made the wind sector in Denmark unique compared with other countries. At the turn of the century, around 150,000 households were co-owners of a local windmill. The ownership model rather than technology and tariff schemes was important to the success of wind energy in Denmark. It was the key factor behind the high public acceptance that wind power projects enjoyed during that time. It also enabled a much faster deployment, because large numbers of people were involved in the sector that provided the necessary good will. In 1992, the role of cooperatives shrank owing to a change in planning procedures. In 1998, as a result of liberalization in the sector, the ownership model changed dramatically. The residential criteria for ownership were abolished and everyone was allowed to own a number of windmills as possible for which they could acquire building permission anywhere in the country. The takeover bids began, resulting in a dramatic decrease in public involvement. As a result, in 2009, about 50,000 households were co-owners of windmills, compared with 150,000 a decade earlier. Consequently, the attitude toward wind power suffered a reversal. The erection of a windmill often became a local drama and resulted in bitter conflicts that led to long delays or cancellations. In 2008, a wind turbine neighbor compensation scheme was introduced that also caused local conflicts. The solution seemed to be to normalize the wind energy sector so that wind power, like district heating, CHP, and power distribution,

Ownership, citizens participation and economic trends

333

changed from investment to public supply, with municipalities and consumer-owned local companies as the future wind power owners. This made wind power significantly cheaper.

15.4 Integration of the energy supply by public ownership Renewable energy is still young; technology and tariff problems have found reasonable solutions in generally acceptable organizational structures for decentralized ownership for the common good as well as part of the transition process. The ownership and operation of large wind turbines for community supply should be a service provided by the local public authorities, like CHP, the supply of water, central district heating, public transport, and other parts of the public infrastructure. It is the lesson from the past 100 years of practice that the state should promote public regulation in favor of the local and collective ownership of basic public services, such as energy. This approach is in line with the protection and promotion of the common good in most democratic societies. Considering the size of order and complexity of the transition to a 100% renewable energyebased energy system, and its urgency, it is realistic for public administrations to undertake this task. The common-good approach will make wind power a part of the public planning with expropriation of the necessary areas for wind turbines, as is done with power pylons, waterworks, and similar areas of public interest. It is already standard practice to provide monetary compensation when areas are being designated for the common good. This would be a decisive contribution to attain local acceptance and make wind energy more competitive. In Denmark, it could lead to a 30% cost reduction or more than could be expected from more efficient technology. In several countries, new legislation instructs utilities to buy electricity from renewable energy installations at a price determined by the government and guaranteed for 20 years. Such laws are used in the most successful wind energy countries. This would be a serious incentive for the municipalities as well, to be an active part of the development of CHP combined with renewable energy. This can be done through municipal companies, or existing or eventually new local renewable supply companies that can supply a full renewable energy package including wind power. By making the establishment and operation of large wind turbines the responsibility of public supply companies, there will be significant savings

334

N. El Bassam

owing to cheaper areas for wind turbines, saved repowering fees, and cheaper long-term financing. This will make wind energy more attractive for the individual municipality as well as at a national level; it will improve supply security, secure steady energy prices, and be the fulfillment of international agreements concerning CO2 reduction as well.

15.5 Economic impacts Renewable energy delivers on all main pillars of sustainable development: environmental, economic, and even social. Alongside declining costs and steadily improving technologies, the transition to renewables also creates numerous employment opportunities (Renewable Energy and Jobs: Annual Review 2019eIRENA www.irena.org). Renewable energy commercialization involves the deployment of three generations of renewable energy technologies dating back more than 100 years. First-generation technologies, which are already mature and economically competitive, include biomass, hydroelectricity, geothermal power, and heat. Second-generation technologies are market-ready and are being deployed; they include solar heating, PV, wind power, solar thermal power stations, and modern forms of bioenergy. Thirdgeneration technologies require continued research and development efforts to make large contributions on a global scale and include advanced biomass gasification, biorefinery technologies, hot dry rock geothermal power, and ocean energy (International Energy Agency, 2007). PV and solar-thermal plants may meet most of the world’s demand for electricity by 2060 and half of all energy needs, with wind, hydropower, and biomass plants supplying much of the remaining requirements (Figure 15.4). The price of PV modules per megawatt has fallen by 60% since summer 2008, according to Bloomberg New Energy Finance estimates (2011), putting solar power on a competitive footing with the retail price of electricity for the first time in a number of sunny countries. Wind turbine prices have also fallen by 18% per megawatt, reflecting fierce competition in the supply chain, as with solar. Further improvements in the leveled cost of energy for solar, wind, and other technologies lie ahead, posing a growing threat to the dominance of fossil fuel generation sources in the next few years (Renewable Energy World, 2011). The International Solar Energy Society argues that renewable energy technologies and economics will continue to improve with time, and that they are “sufficiently advanced at present to allow for major

Ownership, citizens participation and economic trends

335

Figure 15.4 Clean energy projected growth, 2007e17. From GGByte (2008), from Clean Energy Trends Report.

penetrations of renewable energy into the mainstream energy and societal infrastructures” (Aitken, 2010). As time progresses, renewable energy will generally become cheaper (Analysis, 2009), while fossil fuels will generally be more expensive (Gore, 2009). First, once the renewable infrastructure is built, the fuel is forever free. Unlike carbon-based fuels, the wind, the sun, and the earth itself provide fuel that is free, in amounts that are effectively limitless. Second, although fossil fuel technologies are more mature, renewable energy technologies are rapidly being improved. So, innovation and ingenuity give us the ability to increase the efficiency of renewable energy constantly, and continually reduce its cost. Third, once the world makes a clear commitment to shifting toward renewable energy, the volume of production itself will sharply reduce the cost of each windmill and each solar panel, while adding yet more incentives for research and development to speed the innovation process (Gore, 2009).

15.6 Socioeconomic benefits and economic impacts of Renewables 2019 • The global renewable energy sector employed 11 million people in 2018. This compares with 10.3 million in 2017, based on available information. • Employment remains concentrated in a handful of countries; China, Brazil, the United States, India, and members of the European Union were in the lead. Asian countries’ share remained at 60% of the global total.

336

N. El Bassam

• Several factors, including national deployment and industrial policies, as well as changes in the geographic footprint of supply chains and in trade patterns, and industry consolidation trends, shape how and where jobs are created. • Nonetheless, the increasingly diverse geographic footprint of energygeneration capacities and, to a lesser degree, assembly and manufacturing plants, has created jobs in a rising number of countries. • The solar PV industry retains the top spot, with a third of the total renewable energy workforce. In 2018, PV employment expanded in India, Southeast Asia, and Brazil, whereas China, the United States, Japan, and the European Union lost jobs. • Rising off-grid solar sales are translating into growing numbers of jobs in the context of expanding energy access and spurring economic activities in previously isolated communities. • Rising output pushed biofuel jobs up 6% to 2.1 million. Brazil, Colombia, and Southeast Asia have labor-intensive supply chains, whereas operations in the United States and the European Union are far more mechanized. • Employment in wind power supports 1.2 million jobs. Onshore projects predominate, but the offshore segment is gaining traction and could build on expertise and infrastructure in the offshore oil and gas sector. • Hydropower has the largest installed capacity of all renewables but is expanding slowly. The sector employs 2.1 million people directly, three-quarters of whom are in operations and maintenance. • Although the analysis suggests that there would be job growth in 2018, some of the increase reflected continued improvement and refinement of methodologies that allowed a rising share of employment to be captured in statistics. Global renewable power growth outpaced fossil fuel growth by a factor of 2.6 in 2019, says a report. While ongoing health, political, economic, and social impacts of the coronavirus pandemic continue to preoccupy people across the world, as recovery plans are discussed and taking shape, decision-makers are encouraged to use this as an opportunity to set our societies on a more sustainable development path. The United Nations Environment Program (UNEP) and others are advocating the idea that the novel coronavirus-19 (COVID-19) crisis, which is changing the way we live and work, and perhaps priorities in life and mindsets, can pivot economies toward “building back better,” with, for example, a greater focus on clean energy, green jobs, and sustainable development.

Ownership, citizens participation and economic trends

337

The International Renewable Energy Agency’s (IRENA’s) Renewable Capacity Statistics 2020 shows that new renewable power, principally hydropower, wind, solar, geothermal, and bioenergy, accounted for 72% of all power expansion. Renewable energy expanded by 7.6% in 2019, adding 176 GW of generating capacity globally, marginally lower than the (revised) 179 GW added in 2018. “Renewable energy is the fastest growing source of electricity supply, but it’s important to bear in mind that electricity accounts for only about 20% of energy useddthe rest is mainly fossil fuels: coal, oil and gas,” said UNEP energy expert Mark Radka. “Renewable energy is still challenging in many end-use sectors such as aviation, shipping, industry and heavy transport,” he added. Solar and wind continued to dominate renewable capacity expansion, jointly accounting for 90% of all net renewable additions in 2019. Solar, with 586 GW, increased by 20%, whereas wind, with 623 GW, increased by 10%. China and the United States continued to dominate the increase in wind power, whereas China, India, Japan, the Republic of Korea, and Viet Nam had the highest new solar capacity in 2019 (Figure 15.5). Hydropower accounted for the largest share of the global total, with a capacity of 1190 GW. It increased minimally by 12 GW (up 1% in 2018), possibly because some large projects missed their expected completion dates. China and Brazil accounted for most of the expansion. Other renewables included 124 GW of bioenergy, 14 GW of geothermal, and 0.5 GW of marine energy. Turkey, followed by Indonesia and Kenya, which expanded their use of geothermal energy.

Figure 15.5 A special construction vessel at the Amrumbank West offshore wind park in the North Sea. Photo by Reuters.

338

N. El Bassam

Off-grid capacity grew by 160 MW (up 2%) to reach 8.6 GW in 2019. Bioenergy accounted for 40% of off-grid capacity. China accounted for half of all new capacity in biofuel use.

Renewable Generation Capacity by Region. Asia accounted for 54% of new capacity in 2019 (44% of the global total). Capacity in Europe and North America expanded by 6.6 and 6% respectively. Oceania and the Middle East were the fastest-growing regions (up 18.4 and 12.6%, respectively), although their share of global capacity was small. Africa increased by only 2.0 GW (up 4.3%), to reach 48 GW (Figure 15.6). Nonrenewable capacity (oil, coal, and gas) expansion in 2019 continued to follow long-term trends, with net growth in Asia, the Middle East, and Africa, but with net decommissioning in Europe and North America and little change in other regions (https://www.unenvironment.org/news-andstories/story/uptick-renewable-electricity-generation-2019) (Figure 15.7).

Figure 15.6 Energy consumption by sectors.

Figure 15.7 Economics of renewable energy: global electricity generation. PV, photovoltaics; PNG,. From IRENA et al. (2018); https://www.euislands.eu/document/ community-energy-ownership.

Ownership, citizens participation and economic trends

339

15.7 Actions for broadening the ownership of renewables • Policies could seek to avoid discrimination against smaller investors, especially community-based ones, and could ideally create equal market access for all market participants. In this light, auctions are not the preferred instrument to stimulate community energy deployment (Fell, 2017). Rather, governments could incentivize decentralized, integrated, community-based renewable energy systems and self-consumption and remove any kind of barrier to such approaches. Feed-in tariffs have proven to be more adjustable to the specific needs of smaller-scale investors and community ownership. They also carry a much smaller risk of discrimination against this type of investor (WWEA, 2016). • Where governments wish to include community projects in tendering procedures, they should be aware that to date, community energy investors have not had a substantial role in renewable energy projects developed based on auctions. One way to achieve a higher level of community participation could be to set up specific targets and regulations for community projects (e.g., by reserving a reasonably high capacity for community projects). However, this does not solve the basic problem that most communities cannot absorb the planning risk associated in general with participating in auctions. • When setting up renewable energy support schemes, governments could consider not only the price per kilowatt hour but also the overall macroeconomic costs and socioeconomic benefits, as highlighted earlier. Such an overall analysis could be part of any long-term energy and development plan. • If governments were willing to create special incentives for community energy investments, they must ensure that they target the right group (i.e., the definition of community energy could follow a set of principles similar to the ones outlined earlier). Single indicators such as voting rights alone are insufficient (WWEA, 2018). • The establishment of community energy authorities with the sole purpose of supporting community energy projects by or through providing advisory services and funding opportunities, facilitating stakeholder engagement, and increasing public awareness could significantly accelerate the development of community projects. Such authorities could be established on various levels (e.g., local, regional, national, or international) and could be consolidated into existing institutions (REN21, 2017). Some countries or regions have already established organizations that focus on supporting community energy

340

• •





N. El Bassam

investment, often partly funded by public sources (e.g., Community Energy Scotland and the Community Power Agency in Australia) (Community Energy Scotland, 2018; Community Power Agency, 2018). At the global level, international organizations could include community energy as a priority in their work programs. To overcome the equity gap, particularly in developing countries, governments could contribute to developing alternative business models to encourage financial institutions to dispense loans. Public guarantees can have an important role in this area, especially when given by multilateral finance institutions. An appropriate international finance institution could establish a facility specifically dedicated to financing community energy projects in developing countries. Such a facility could not only provide loan guarantees, it could help to overcome the equity gap. Learning among pioneers and new entrants is one of the keys to community energy’s further development. Often, good practices in one country are replicated in others. Therefore, to foster open innovation, the creation of a regular space for networking or a meeting place to exchange knowledge, experience, and ideas is recommended. Existing networks that could be strengthened, including an increasing number of national community energy organizations and events. On an international level, in addition to the Community Energy Group of the Coalition group, networks such as the WWEA Community Power Working Group or the World Community Power Conference could contribute to these efforts.

15.8 Global investment’s in renewables In 2011, the global investment in renewable power and fuels increased to a new record. Significantly, developing economies made up 35% of this total investment. In addition, the whole period of 2004e17 witnessed a remarkable increase in investments in renewables in different sectors or for different technologies in different countries with different economic systems, as illustrated in Figures 15.8 and 15.9. However, recent years have seen investments in renewable energy in the power sector stagnate. Yet, renewable power generation capacity continued to be installed at record pace mainly thanks to continuously falling costs of technology.

Ownership, citizens participation and economic trends

341

Figure 15.8 Renewable energy investments 2019 by region. Source: http://fs-unepcentre.org/research/report.

Notable trends for 2018 were that investments continued to be geographically more widespread; 29 countries recorded US $1 billion or more in investments (25 countries in 2017) and an additional 14 countries exceeding US $500 million. 2018 also marked the fourth year in a row in which investments in developing countries were higher than in developed countries (https://energypedia.info/images/d/d3/Renewable_energy_investments_ 2019_by_region.jpg). The cost advantage that fossil fuels used to have over renewable energy sources has been decreasing. Some renewable technologies (solar PV, wind, and hydropower) already compete with fossil fuels directly on the financial frontier. Furthermore, renewables’ costs are expected to decline even further, and those of fossil fuels will incline. Figures 15.8 and 15.9 show that although oil prices were on the rise in the 2000s, investments in renewables were also on the rise during that period, reflecting renewables’ competitiveness against oil.

15.9 Costs of renewables According to recent reports on renewable energy technologies from IRENA, the Renewable Energy Policy Network for the 21st Century, and the International Energy Agency, electricity costs from almost all renewable projects that were commissioned in 2017 continued to decline (Table 15.1).

342

N. El Bassam

Figure 15.9 Renewable energy investments by technology, 2019. AMER, Americas; ASOC, Antarctic and Southern Ocean Coalition. From http://fs-unep-centre.org/research/report. Table 15.1 German levelized cost of electricity renewable energy technologies in euros per megawatt-hour. ISE (2013) ISE (2018) Low cost

High cost

Low cost

High cost

Brown coal Hard coal

38 63 75

53 80 98

46 63 78

80 99 100

Onshore wind farms Offshore wind farms Photovoltaic systems

45

107

40

82

119

194

75

138

78

142

37

115

135

250

101

147

Technology

Coal-fired power plants Combined cycle gas turbine power plants Wind power

Solar Biogas power plant

From Fraunhofer ISE (2018) e Stromgestehungskosten erneuerbare Energie. https://en.wikipedia. org/wiki/Cost_of_electricity_by_source.

Projects of bioenergy power, hydropower, geothermal, and onshore wind, which were commissioned in that year, widely fell into the generation costs’ range of fossil-generated electricity; furthermore, some of those projects actually undercut those of fossil fuelebased ones.

Ownership, citizens participation and economic trends

343

“In some ways the COVID-19 crisis is the perfect opportunity for us to pause and to ramp up a just transition to carbon neutral economies, with all the benefits that we will have in terms of health (cleaner air) and mitigating costly climate change impacts,” said UNEP climate change expert Niklas Hagelberg. Nature is in crisis, threatened by biodiversity and habitat loss, global heating and toxic pollution. Failure to act is failing humanity. Addressing the current coronavirus (COVID-19) pandemic and protecting ourselves against future global threats requires sound management of hazardous medical and chemical waste; strong and global stewardship of nature and biodiversity; and a clear commitment to “building back better,” creating green jobs and facilitating the transition to carbon neutral economies. Humanity depends on action now for a resilient and sustainable future. (Niklas. [email protected]).

References Aitken, D.W., January 2010. Transitioning to a Renewable Energy Future. Interna- tional Solar Energy Society, p. 3. Bloomberg New Energy Finance, 2011. Global New Investments in Renewable Energy UNEP SEFI, Frankfurt School, Global Trends in Renewable Energy Investment. Clean Energy Projected Growth 2007e2017. Gore, Al, 2009. Our Choice. Bloomsbury Publishing. PLC, p. 58. ISBN: 0747590990. International Energy Agency (2007). International Renewable Energy Agency (IRENA), 2018. International Energy Agency (IEA). & Renewable Energy Policy Network for the 21st Century (REN21). Renewable Energy Policies in a Time of Transition. Retrieved from: http://www.ren21.net/ wp-content/uploads/2018/04/17-8622_Policy_FullReport_web_.pdf.

Further reading Clean Edge, 2008. Energy Trends Report. GGByte at en.wikipedia 2012. http://en. wikipedia.org/wiki/File:Re_investment_2007-2017.jpg. Frankfurt School & United Nations Environment Programme (FS-UNEP Collaborating Centre), 2018. Global Trends in Renewable Energy Investment 2018. Retrieved from: https://europa.eu/capacity4dev/search?text¼Renewable%20energy%20investment. Gipe, P., 2012a. From data by Unendlich viel energie. Windworks. January 5, 2012 [WWW]. Available from: http://wind-works.org/coopwind/CitizenPowerConferencetobeheld inHistoricChamber.html. Gipe, P., 2012b. Germany renewable energy generation is owned by its own citizens. Windworks [WWW], January 5, 2012. http://wind-works.org/coopwind/Citizen PowerConferencetobeheldinHistoricChamber.html. Gloystein, H., November 23, 2011. Renewable Energy Becoming Cost Competitive, IEA Says. Reuters. Johnson, K., March 25, 2009. Wind shear: GE Wins, Vestas Loses in Wind-Power Market Race. Wall St. J. (Accessed 7 January 2010). National Renewable Energy Laboratory, 2006. Nontechnical Barriers to Solar Energy Use: Review of Recent Literature, p. 30. Technical Report NREL/TP-520e40116, September.

344

N. El Bassam

REN21, 2010. Renewables 2010 Global Status Report, p. 9, 26 & 34. REN21, 2011. Renewables 2011: Global Status Report, pp. 11e13. REN21, 2012. Renewables Global Status Report, 2012, p. 17. Renewables in global energy supply: An IEA facts sheet (PDF) OECD Sills, B., August 29, 2011. Renewables Investment Breaks Records. Renewable Energy World 29, p. 34. Solar May Produce Most of World’s Power by 2060, IEA Says. Bloomberg. Solar Power 50% Cheaper by Year End e Analysis Reuters, November 24, 2009. United Nations Environment Programme and New Energy Finance Ltd, 2007. Global Trends in Sustainable Energy Investment. Analysis of Trends and Issues in the Financing of Renewable Energy and Energy Efficiency in OECD and Developing Countries, 3 (PDF). [ 10.0 10.1 Frankfurt School of Finance & Management e UNEP Collaborating Centre, 2019. Global Trends in Renewable Energy Investment 2019. http://fs-unep-centre. org/research/report. [ 13.0 13.1 International Renewable Energy Agency (IRENA), 2018. International Energy Agency (IEA) & Renewable Energy Policy Network for the 21st Century (REN21). Renewable Energy Policies in a Time of Transition. Retrieved from: http://www. irena.org/-/media/Files/IRENA/Agency/Publication/2018/Apr/IRENA_IEA_REN21_ Policies_2018.pdf https://www.euislands.eu/document/community-energy-. Actions to promote community energy.

CHAPTER SIXTEEN

The importance of green mobility Daniele Pagani Sustainable Mobility Department, Nordic Folkecenter for Renewable Energy, Hurup Thy, Denmark

Contents 16.1 Environmental and social impacts 16.1.1 The CO2 impact of transport 16.1.2 Air pollution and health 16.2 Mobility on the road 16.2.1 Available technologies 16.2.1.1 16.2.1.2 16.2.1.3 16.2.1.4 16.2.1.5 16.2.1.6 16.2.1.7 16.2.1.8

346 346 348 349 349

Conventional fossil fuels Liquefied petroleum gas Natural gas and its alternatives Biofuels Electricity Hydrogen Hybrids Electrofuels

349 350 353 357 362 368 372 376

16.2.2 Light-duty transportation 16.2.3 Heavy-duty transportation 16.3 Mobility on the rail 16.4 Mobility on the water 16.5 Mobility in the air 16.6 Rethinking mobility: are there any alternatives to current models? References

377 379 380 381 382 383 386

Transport has become a major industry and its development allowed the world to become connected. From simple daily commuting to international passenger and freight transport, humanity’s horizons have widened and globalization has developed. Although people and goods have always traveled, this trend was accelerated by the discovery of the internal combustion engine (ICE) and the consequent use of liquid oil to power it. Electric mobility had already existed (the first car capable of reaching 100 km/h was electric), but the discovery of cheap and abundant oil paved the way to the success of ICEs. Capable of achieving decent performance, ICEs were, and still are, less efficient than electric engines, but they allowed Distributed Renewable Energies for Off-Grid Communities, Volume 2 ISBN: 978-0-12-821605-7 https://doi.org/10.1016/B978-0-12-821605-7.00020-9

© 2021 Elsevier Inc. All rights reserved.

345

j

346

Distributed Renewable Energies for Off-Grid Communities

for a longer range and quick refills, features that constituted a limitation for electric vehicles (EVs). Electricity was stored in batteries, the development of which was still at an early stage. A lot of space and weight were necessary to store enough energy to move the vehicles, energy that could be stored in only a few liters of oil. Although these limitations continue to exist, technology has evolved on both sides, allowing batteries to store more energy, but allowing ICEs to operate more efficiently using alternative, more sustainable fuels. This chapter will provide an overview of available technology and developments expected to take place in coming years. First, a general overview will be presented of the transport sector and its impact, after which a deeper analysis will be performed of possibilities available in the different transport sectors. The chapter will conclude with solutions already in place in some realities, the diffusion of which would have a considerable impact on society.

16.1 Environmental and social impacts Moving goods and people does not come free of charge: a considerable amount of energy is used for it and energy demand for transport increased considerably in the last decades. According to Moriarty and Honnery (2016), energy consumption for transport more than doubled compared with 1970 values, resulting in the sector consuming almost a quarter of global energy used. Oil is still the main fuel, and its use is mostly related to road transportation (Statista, 2018).

16.1.1 The CO2 impact of transport Global CO2 has constantly risen in past decades, reaching a record value of 33 gigatons in 2019 (IEA, 2020). Transport contributes considerably to this number; it saw growth over the past five decades (Global Carbon Project, 2019). Within transport, the sector that saw the steepest growth was aviation (þ140% compared with 2000) (IEA, 2019a), but when looking at overall emissions released by this sector, it can be seen that its contribution is still limited in the overall account. Indeed, the major emitter is road transport, which is unsurprising considering that there are an estimated 1.2 billion vehicles in the world (World Health Organization Europe, 2015). The large amount of existing units constitute the major barrier to cleaning the sector and make it the most difficult one to decarbonize. Mass- and large-volume transportation units can lower emissions more easily. The rail sector is by far the easiest to decarbonize because electric trains exist and are widely used around the world. The challenge might

The importance of green mobility

347

be the lack of electrified lines in some areas, but this problem is related to investments. Furthermore, solutions are available for these cases, as will be explained later in the chapter. Maritime transportation is more complex, because a good alternative to fossil-based products has yet to be found, but space and weight allowances on vessels provide some kind of flexibility for different solutions. Aviation is clearly the most challenging sector because of requirements of the industry. A plane has a limited carrying capacity in terms of both space and weight, which considerably limit the possibilities. For example, a battery-powered plane would force the operator to decide whether to fly green or be able to carry more passengers. Some solutions to this problem are being studied, as will be discussed later, but there is no doubt that technological developments within the aviation industry will take years to enter the market. This is also because of high safety requirements in place for airplanes, which, of course, are an additional barrier. Despite these challenges, for the marine and aviation sectors, quick decarbonization is more realistic because they share a common aspect: both are operated by professional entities that aim to be profitable and that have considerable purchasing power. Once a suitable solution for these sectors is available, it will take only a short time for operators to consider it. The propulsion of a large vessel is just a fraction of the whole cost, and an operator who will see an economic benefit in switching to a new technology will not hesitate to invest in it. This is evidence by cruise operators who are investing in hybrid solutions (multiple fuels) (Ship Technology, 2019) and freight companies that are looking with interest into new fuels. On the other hand, airlines purchase plane in blocks, so the replacement of a fleet can be rapid, especially if economic benefits can result from this transition. Pressure on the industry set by society is considerable, and airlines will need to adapt in time if they want to keep their customers. On the contrary, road transportation (and particularly, private transportation) is easy to decarbonize, as far as technology goes (solutions are already available, even if they are not perfect), but the sector as a whole represents a considerable challenge. As mentioned, the large amount of units means that it will take a long time for the whole sector to become green, owing to the limited investment capabilities of the owners, limited interest, and difficulty in checking fleet performance. Excluding electric mobility, some solutions are in place to lower emissions generated by vehicles, but it is difficult to verify that design performance is kept within the parameters for the whole

348

Distributed Renewable Energies for Off-Grid Communities

lifetime of the units. Many countries have periodic inspections, but they are normally placed at long intervals (on the order of years), so there is no control over how the vehicle performs in the meantime, and tests can be faked. Furthermore, not all private users are interested in or able to conduct proper maintenance on the vehicle, unlike professional operators, which poses even more challenges. With these limitations in mind, this chapter will describe, from an objective viewpoint, possible solutions to lower the impact of different sectors of transportation. Which will be the chosen solution will be decided by the market.

16.1.2 Air pollution and health CO2 is the big polluter. All environmental political debates and much of the scientific literature focus on the impact of this gas on society, but CO2 is part of our living cycle. According to Science Focus, a science magazine published by the BBC, an average person releases 1 kg/day of CO2, which, considering the world’s population, results in 2500 million tons per year just as a result of normal breathing. To give a reference parameter, this is equivalent to 7% of yearly emissions of fossil fuel burning (Withers, n.d.). Another scientific study found that an average person emits about 2 tons/year of CO2 by eating (from food production to excretion) (Mu~ noz et al., 2010; FECYT - Spanish Foundation for Science and Technology, 2010). Although they are considerable, these quantities are balanced by the living cycle and by tree absorption, so CO2 per se is not dangerous, but it is part of life on earth. The problem to which scientists refer is therefore unrelated to the gas itself, but to its quantity, which increased considerably over the past century as a result of human activities. Coupled with increased deforestation, this resulted in record values that can be found in the atmosphere. The excessive concentration of the gas creates the so-called greenhouse effect, which increases the heat retention of the planet and thus its temperature, with all the consequences related to it. Although some consequences of climate change are already visible, its real impact will be seen only over the long time (years to decades), whereas the consequences of emissions from the transport sector can be perceived on a daily basis. Air quality in some cities is so bad that local authorities are forced to stop or regulate traffic when some parameters are exceeded. This has resulted in sickness in the population and, in some extreme cases,

The importance of green mobility

349

premature death. According to the World Health Organization, 600,000 premature deaths in Europe were related to pollution, which cost the society about $1.6 trillion/year in medical care. To understand how large this impact is, the amount is equivalent to one-tenth the European Union’s (EU’s) gross domestic product in 2013 (World Health Organization Europe, 2015). These consequences are the result of the fact that transport emits CO2 but also a variety of gases (CO, Hydrocarbons (HC), NOx, particulate matter [PM], and NH3 are some of the most common) (IEA - AMF, 2015), which can have a considerable impact on both the environment and human health. Because it is unrealistic to renounce transportation and its connected benefits, it is crucial to find solutions that can eliminate or at least reduce the emissions related to it. This chapter will attempt this task, presenting concepts and technologies that could mitigate the impact of transport on society.

16.2 Mobility on the road As discussed in the previous section, road mobility is by far the main contributor of CO2 emissions within the transport sector, which is why a large part of this chapter will focus on this topic.

16.2.1 Available technologies The section will describe current technologies used in the transport sector to mitigate its impact and those that have the potential to be adopted in the near future. The review will present the advantages and disadvantages of available solutions, bearing in mind that not necessarily the entire transport sector should not switch to a single technology. The key to obtaining considerable reductions in emissions is to apply the right technology to the right purpose so that multiple solutions might coexist within the same sector. 16.2.1.1 Conventional fossil fuels This open approach should be applied to conventional fossil fuels as well. The ideal situation would be to eliminate them from the market, but the reality is that many of these solutions might not be applicable to the whole world. There are countries and regions that will not be able to use the

350

Distributed Renewable Energies for Off-Grid Communities

same technologies as those in richer and more developed areas owing to a lack of investment capacity or poor infrastructure. However, there is a solution, which consists of upgrading the fleet of these areas to more modern and more efficient units. Technology has developed massively over past decades, reaching relatively low emissions and improved efficiency even in vehicles powered by conventional fuels. In 1992, the EU established the EUROX label with the purpose of reducing pollution generated by vehicles. This label, in which the X is substituted by an Arabic or roman number (depending on the applications), sets maximum levels of emissions allowed for vehicles sold in the region. The label has a legal value, so a manufacturer cannot sell vehicles that do not comply with current emissions standards in any member countries. This limitation forces manufacturers to invest in research and development, with the result that cleaner vehicles are developed. The standard, which is currently in its sixth version, has caused a significant reduction in maximum tailpipe emissions produced by a vehicle. Thus, simply switching to a newer vehicle can massively reduce the impact of transport on society. Similar quality requirements have begun to be applied to fuels, which in some regions of the world (e.g., EU) need to include a percentage of greener alternatives. Specifically, ethanol and biodiesel are added to gasoline and diesel, respectively, to lower emissions generated by conventional fuels. 16.2.1.2 Liquefied petroleum gas Discovered in 1911, liquefied petroleum gas (LPG) is derived mostly from fossil fuels. It consists of a mixture of propane and butane, the ratio of which varies based on the country of use (myLPG.eu, n.d. b). The name comes from the fact that at ambient temperature and pressure, the element is in a gaseous stage, but it is sufficient to provide modest pressure or cooling to turn it into liquid, an aspect that makes it convenient for transport applications. LPG used in the automotive world is known as autogas and it is currently the third most used transportation fuel (Alternative Fuels Data Center, n.d.); over 26 million vehicles are powered with this fuel. Despite its popularity, its use is restricted to a limited amount of countries; Turkey holds the largest fleet (4.6 million vehicles in 2017) and South Korea has the largest consumption (3.3 million tons in 2017) (Autogas.net, 2019). Although most of its production is derived from natural gas and oil production, autogas is considered a green fuel because of its lower emissions compared with conventional fossil fuels. A study conducted in 2003 compared

The importance of green mobility

351

emissions produced by different fuels (Verbeek, 2008). Although the results of the study were considerably old and outdated, they showed that under equivalent technologies, LPG could considerably reduce most harmful emissions generated by gasoline and diesel engines, which opens possibilities to convert old vehicles when no other options are available. Retrofit do not to deliver the same performance as dedicated designs (Verbeek, 2008), but it will improve the situation when an old fleet is available. Throughout the years, technology developed and so did engine performance. Current EURO 6 emissions are 60 and 96% lower than those of EURO 1 for gasoline and diesel, respectively (The International Coucil on Clean Transportation, 2016). Autogas has been used in the automotive industry since 1980, first to convert gasoline vehicles and then as a ready-made solution provided directly by vehicle manufacturers. The latter is the most efficient option and normally is combined with gasoline operations, so the vehicle can run one fuel or the other (bifuel vehicle). This flexibility allows the user not to depend on the LPG infrastructure, which in many countries is still considerably limited. The most advanced system to drive with autogas is liquid phase direct injection (LPdi). Compared with the previous five generations of autogas systems, LPdi injects LPG directly into the combustion chamber (without the need for a converter), which allows the fuel to vaporize instantly and thus improves the performance of the engine and reduces emissions. This is done by means of a high-pressure pump, such as the ones used for gasoline direct injection (Autogas.net, 2017). Gasoline-autogas operability is the most common, but solutions also exist for diesel-autogas operations. This design reduces smoke emissions generated by diesel engines while keeping good levels of torque. However, operations at low and middle loads are not optimal, resulting in higher fuel consumption and higher emissions of HC and CO. To operate efficiently, diesel-autogas should use diesel only at low and medium loads and combine the fuels at high loads, which will reduce emissions and still have good mileage. In addition, because diesel engines are designed differently, more modifications are necessary to accommodate both fuels. The introduction of spark plugs, different compression ratios, and stresses related to different rpm operations are some parameters that often make them uneconomical although the conversions are technically possible (Go-LPG, n.d.).

352

Distributed Renewable Energies for Off-Grid Communities

Regarding the filling procedure, there are currently more than 46,000 filling stations in 62 countries (myLPG.eu, n.d. c). They use six type of connectors, the most common of which is the dish connector. Other configurations used in Europe (and in most of the countries) are the bayonet, Euro, and ACME connectors. There are also the Asian and Eurasian types, but their use is limited to certain countries: Japan and Korea for the first type and Russia, Kazakhstan, and Uzbekistan for the second type. Adapters are available, but it advisable to use them only one at a time and not stack them in series, because doing so risk leaks that can increase exponentially. Some vehicles are also equipped with removable connectors, which increases the safety level (myLPG.eu, n. d. a). Regarding safety, many people may be scared by the risk for explosion connected with transporting autogas. This risk is present but it is minimized by the safety of the tanks, which are engineered to lower the chance of this happening, and by long experience gained in the sector. Nevertheless, countries might draft additional regulations to ensure the safe operation of autogas vehicles. This is the case of directive UN/ECE 67/01, which allows autogas vehicles to be parked underground provided they do not go below the first floor. The industry is undergoing some innovations that could further increase the use of autogas. One is related to the fuel itself, which can be produced from crops and waste feedstocks instead of fossil fuels, making it more sustainable. A number of companies are working in this direction, so the transition to a green LPG might be close (World LPG Association, 2018). Another innovation is related to the vehicles. Because of the similarity of engines using gasoline and autogas and because most hybrid vehicles in circulation run on the combination of gasoline and electricity, it did not take long for manufacturers to develop hybrids running on autogas and electricity. According to the World LPG Association, these solutions will provide larger environmental benefits compared with conventional hybrids (World LPG Association, 2019). The technology is still not widespread, but there are positive signals along that line. In 2017, Toyota launched the JPN Taxi, an autogas hybrid vehicle available for order in Japan (Autogas.net, n.d.), so other manufacturers might follow the example. The market will ultimately decide whether these innovations have a future, and if so, how large an impact they can make.

The importance of green mobility

353

16.2.1.3 Natural gas and its alternatives Natural gas has been used to power vehicles since the 1860s. Because it is also used for energy and heating applications, much experience with it is available. Most people consider natural gas and methane to be synonyms, but they are two different gases, although methane remains the main contributor (70e90%). Natural gas also contains ethane, propane, and butane (up to 20%), CO2 (up to 8%), and other gases at smaller percentages (NaturalGas.org, 2013). Most applications that use the gas are designed for methane use, so the name “natural gas” simply states where the methane come from: in this case, from underground, where it was stored for millions of years. Despite its fossil origins, burning natural gas is more sustainable than using other fossil fuels. This is because less carbon is present in the molecule of methane, because for every atom of carbon there are four atoms of hydrogen. On the other hand, methane has 21 times the greenhouse effect of CO2 over 100 years and it has a lifetime of 12.4 years in the atmosphere, so it will take a long time before the gas is dissolved (Thiruvengadam et al., 2018). This poses particular challenges in containing unwanted emissions generated from leaks throughout the production and distribution chain, because a small spill of methane could easily have an important effect on the environment. Besides natural gas, which is extracted, and which based on the location, can contain methane purities close to 100%, a number of gases are used in the industry to substitute for the natural version. These are listed in the following sections. 16.2.1.3.1 Synthetic natural gas

Synthetic natural gas (SNG) is an artificial gas with properties and compositions as similar as possible to the natural version. It can be obtained by coal, peat, oil, or (waste) biomass by means of gasification. The process, which converts solid mass in gaseous components, takes place under high pressures and temperatures and under the limited presence of oxygen. The output is a gas that contains various elements (H2, CO, CO2, H2O, and CH4) afterward purified in the gas-cleaning and gas-conditioning process. The first process removes impurities such as ammonia and sulfur, whereas the second process converts parts of the produced gas so that it matches needed composition requirements. Finally, the gas undergoes water shift reactions, which combine CO and steam to create CO2 and hydrogen. The first is left out through

354

Distributed Renewable Energies for Off-Grid Communities

scrubbing, whereas hydrogen is coupled with CO, which, owing to a nickel catalyst, produces methane (Energy Education, 2018). If the process is carried out with biomass, the final output is called bio-SNG. 16.2.1.3.2 Unconventional fossil methane

Unconventional fossil methane is a gas that is similar in composition to natural gas, but it has been extracted from unconventional reservoirs such as shales. The procedure for extracting this gas consists of injecting fracturing fluids (water, sand, and a small amount of chemical substances) in geological formations that contain hydrocarbons. The high pressure used breaks the rocks and enlarges fractures through which the gas can flow (European Commission, 2019). To the most, this process is called fracking; it is at the center of a major debate regarding its impact on the environment. A survey conducted on a fracking area in the United Kingdom showed that 62% of people interviewed were worried about destruction of the natural environment (Energy Voice, 2019), an aspect that can lead to limitations enacted by governments regarding this practice. 16.2.1.3.3 Biogas and biomethane

Biogas is gas that originates from the fermentation of organic materials. It is an optimal candidate for circular economy principles, because its concept focuses on producing energy out of products that otherwise would be wasted. The process consists of placing organic matter into a closed container (the digester), where anaerobic bacteria turn part of the biomass into gas. Animal manure is normally used as the baseload, owing to its low cost and high availability, to which other organic types of waste are added. Depending on the configuration of the plant, the expected output, and the biomass inserted, digestion can occur at 35e40 C or 55e60 C, if mesophilic or thermophilic bacteria are used, respectively. Retention times also vary based on the process chosen, but they are on the order of 15e35 days; the thermophilic process has lower retention times. The output of a biogas plant consists of digested slurries, which go back to the fields as fertilizers, and biogas, which normally is composed of a mixture of methane (50e65%), and CO2, with some traces of H2S. The obtained biogas can be used in engines, but some adjustment of equipment might be needed. Instead, it is preferable to upgrade it to higher percentages

The importance of green mobility

355

of methane, so that virtually no differences between that and natural gas can be found. In such a case, we refer to the gas as biomethane, to differentiate its origin. Biogas and biomethane are fully sustainable because they are part of a natural process. If organic matter is not put into a digester, it will still decompose, with the consequent release of methane into the atmosphere. Despite its low impact, it is still important to pay attention to methane spills, because the greenhouse potential is unaltered. Although the upgrading practice adds costs and complexity to the process, benefits are that the gas can become part of the refilling infrastructure, and it can be used to substitution for or be combined with natural gas, thus simplifying gas-powered vehicles. 16.2.1.3.4 Vehicles run on methane

As mentioned, running vehicles on gas had been achieved in the 1860s, but it was also popular during WWI and up to the 1940s, when gas bag vehicles were used. Because of the high volume occupied by the gas and the large amount needed to allow mobility, vehicles were equipped with gas bags installed either on the roof or in a trailer following the vehicle. The gas was first stored at ambient pressure; in the 1940s, pressures of 10 bar could be achieved (Lampinen, n.d.). To have a good range, it was necessary to compress the gas further; currently, compression rates are 20e30 times higher than those of the 1940s (NGVA Europe, n.d. a; Energuide.be, n. d.). In this case, we talk about compressed natural gas (CNG) vehicles (Figure 16.1). Natural gas and its alternatives power more than 20 million vehicles in the world (Energuide.be, n.d.), from small cars to large heavy-duty vehicles (HDVs). The technology is well-known and allows for standardization and cost reductions, as well as flexibility in operation. Users can choose among the monofuel, bifuel, and dual-fuel options based on their needs. Monofuel runs exclusively on gas, but it often has a limited range, so it is suitable for applications that allow for frequent refilling. An example is city buses, which drive a constant route and are never far from filling stations. Europe has approximately 16,000 CNG buses and 10,000 CNG trucks operating on its roads, which proves the reliability of the technology (NGVA Europe, 2018). Bifuel vehicles enable the choice of gas or another fuel, normally gasoline. This flexibility makes up for the lack of infrastructure. Although Europe has almost 3800 gas filling stations (NGVA Europe, n.d. b), this

356

Distributed Renewable Energies for Off-Grid Communities

Figure 16.1 A biogas-powered bus in Bonn, Germany.

number constitutes less than 0.2% of the total refilling network (Statista, 2019a). Most bifuel vehicles are meant for light-duty operations (cars and vans), and it is possible both to convert a gasoline car and buy a specifically designed model. The second option is normally preferable because conversion can be excessively expensive compared with the value of the car, and because it normally implies a loss of space in the luggage compartment. Finally, the dual-fuel option takes the best of both fuels used, but it is mostly meant for HDVs (diesel plus gas) and is not widespread. To allow the spread of the technology and properly service CNG users, the EU issued a directive (2014/94/EU) suggesting the installation of filling stations every 150 km (Osorio-Tejada et al., 2017), which is a good step ahead. Furthermore, fillings stations are being installed on main transport corridors. An alternative to the short range offered by CNG vehicles is represented by liquid natural gas (LNG). Here, gas is cooled to 162 C, turning it into a liquid. At this stage, the gas is at 1/600th of its initial volume, which allows smaller tanks to cover the same mileage. Studies show that an LNG vehicle can drive about 2.4 times the distance of a CNG with the same tank size (Osorio-Tejada et al., 2017). Nevertheless, this technology requires special tanks and systems, which make it beneficial only for long-range HDVs.

The importance of green mobility

357

According to the Natural Gas Vehicle Association, there are 261 LNG stations in Europe (NGVA Europe, n.d. b), mostly located in Italy, Spain, France, the Netherlands, Belgium, Sweden, and the United Kingdom. An overview of the gas stations in Europe can be found on https://www. ngva.eu/stations-map/. Finally, an additional aspect to consider regarding gas vehicles is that they can represent a good transition to hydrogen-powered units, because some of the equipment can be also used in this technology. 16.2.1.4 Biofuels Biofuels are an interesting technology for energy transition in the transport sector. Originating from biomass, they can be used in conventional engines with minor or no modification. The use of this resource is probably the easiest transition the transport industry could have, because it would not require massive investments in infrastructure, since the current one could be easily adapted to the new fuel. From the user’s viewpoint, few or no changes in behavior would be needed, allowing the vehicles to be used as always. In addition, biofuels are believed to be a good solution for the environment, because carbon released while burning the fuel is said to be equivalent to that absorbed by plants during their lifetimes (carbon neutral). Nevertheless, many discussions are related to the sustainability of biofuels. Plants absorb carbon whereas fossil fuels emit carbon that was first stored underground, but biofuel production is not limited to growing crops. Plants need fertilizers, which require energy and ammonia, a component mostly obtained by natural gas. Furthermore, aspects such as water consumption and land use should be considered. According to a study carried out by the World Resource Institute, the world uses three-quarters of vegetated land to cover food demands and forest products. The study calculated that there was a food gap of 70% between calories available in 2006 and those expected to be needed in 2050, and that this gap might rise to 90% if land were used for biofuel production. However, the “battle for land” is not the only problem. That study concluded that the energy output of crops for biofuels is low. For example, sugarcane grown on highly fertile land can convert only 0.5% of solar radiation to sugar and only 0.2% to ethanol. The equivalent land would be considerably more efficient if photovoltaic panels were installed, because solar energy conversion would be 100

358

Distributed Renewable Energies for Off-Grid Communities

times higher than that for ethanol production (Searchinger and Heimlich, 2015). However, this does not mean that biofuels should not be used, just that their production should be smart. Land devoted to growing should not compete with other uses, whether for food production or natural areas, and waste products of other industries should be used so that the final impact can be reduced. Based on this concept, biofuels are divided into four generations, showing the development that the industry had over the past years. 16.2.1.4.1 First-generation biofuels

In this category are biofuels, the origin of which is linked to edible biomass (e.g., corn and sugar) and the exploitation of which leads to competition between energy and food, as discussed earlier. First-generation biofuels are produced at a rate of 50 billion L/year (Alalwan et al., 2019) and are blended or used pure in vehicle engines. Bioethanol is mostly produced from feedstocks; it is synthetized by mean of fermentation. Normally blended with gasoline at a ratio of 10% (E10), it is used to reduce the emissions of fossil fuel. At this concentration, it can be used directly in conventional vehicles with no modification. The benefits of using this biofuel have to be compared with its environmental and social impacts. A study found that crops most commonly used to produce it are sugarcane, maize, wheat, sugar beet, and sorghum; these crops could have fed 200 million people if they were not used to produce energy (Alalwan et al., 2019). However, there are ways to reduce this impact: choosing the right growing locations that do not compete with other uses and using sustainable fertilizers (Alalwan et al., 2019) (e.g., output from biogas plants) would not make the crops more sustainable, but it might increase the amount of carbon extracted from the atmosphere. According to Mahmudul et al., the production of biodiesel mostly depends on oil crops, and most of its costs are related to feedstock production (Mahmudul et al., 2017). More than 350 species of plants, both edible and not, can be used for this purpose (Sajjadi et al., 2016). The processes normally used to obtain the oil are pyrolysis, microemulsion, and transesterification. Like bioethanol, biodiesel can be used pure or blended. The latter reduces emissions generated by diesel and keep performance similar to that for which the engine was designed. This practice is in place in the EU, which aimed for 10% of transport fuel coming from renewable sources by 2020 (European Commission, n.d.). According to a study by Buyukkaya

The importance of green mobility

359

Figure 16.2 Local use of pure plant oil (PPO) can provide a considerable environmental benefit, because emissions related to the extraction, processing, and transport of the fuel are eliminated. A car converted by Nordic Folkecenter, Denmark is filled up in a rapeseed field with PPO from the same crop.

on HDVs, a blend of 20% would be optimal to obtain a good balance between performance and emissions (Buyukkaya, 2010). A separate discussion should be reserved for pure plant oil (PPO) (Figure 16.2). Unlikely what most people believe, biodiesel and PPO are the same, although their origin is similar. Biodiesel is plant oil that has to undergo a refinery process (like conventional diesel), but it does not require modification to the engine. On the other hand, PPO is just oil pressed directly from the plant; it can be used in modified diesel engines. Although PPO shares the same negative aspects of other first-generation biofuels, it can considerably reduce emissions. Normal oil is extracted, refined, transformed into diesel, and transported around the world, where vehicles as well as agricultural machinery use it. All of this generates a considerable amount of emissions, including the tailpipe emissions of vehicles. Agricultural machinery is the key to reducing these emissions. Tractors and other machinery used in agriculture are essential for food production, and sustainable alternatives (e.g., electric) are still not ready for that market. By using locally grown PPO, emissions generated to power the machinery are reduced to tailpipe emissions, for which technological solutions are available. Of course, this is also true for biodiesel, but it still has a higher impact because the refinery is operated at specific locations and because of the

360

Distributed Renewable Energies for Off-Grid Communities

transport involved. PPO, on the other hand, can be produced by the farmer, which allows the tractor to be filled with a “0 km” fuel. Thus, first-generation biofuel have considerable benefits, provided the right approach is taken. 16.2.1.4.2 Second-generation biofuels

Second-generation biofuels make the production of fuel more sustainable, so plants used are not edible and thus are not in direct competition with food production. Furthermore, the goal for this generation is to have a product for which the net carbon (the ratio between produced and consumed CO2) is neutral or negative. Agricultural waste, straw, grass, and wood are used for this purpose (Alalwan et al., 2019). Lignocellulosic biomass is the base for producing bioethanol, which is obtained by hydrolysis and subsequent fermentation, but it can also go through thermochemical processes (such gasification), after which either fermentation or catalyzed reactions are used. The downside of these processes is their complexity, owing to the difficulty of biomass breakdown and the unwanted release of sugars, which require fermentation. Together with the cost of collecting and storing low-density biomass, this makes the process expensive (Alalwan et al., 2019). Second-generation biodiesel is produced from energy crops and agricultural and wood waste as well as waste from cooking oils and animal fats. Animal fats are actually a good source because they produce fuel with a high octane number, which is not corrosive. Drawbacks of second-generation biodiesels are low performance in cold temperatures and the risk for spreading sickness from contaminated feedstocks (Shah et al., 2018). An additional fuel belonging to the second generation is butanol, which consists mainly of carbon and hydrogen; therefore, it can easily be blended with conventional hydrocarbon fuels. Produced by the fermentation of biomass, butanol has more heat energy and is less corrosive than ethanol, whereas it is safer to handle than gasoline and ethanol (Michalski et al., 2017). When an ICE engine is run with butanol, only CO2 and water are released, which makes it a more sustainable choice (Nigam and Singh, 2011); however, production is not sufficiently cost-effective (Höning et al., 2014). In general, second-generation biofuels are an interesting solution, but their availability and production methods limit their diffusion.

The importance of green mobility

361

16.2.1.4.3 Third-generation biofuels

In the third generation, crops are replaced by algae; based on their size and morphology, algae can be divided in macroalgae and microalgae. Kelp is one of the most commonly used macroalgae and its morphology resembles plants on earth. Microalgae can be autotrophic, heterotrophic, and mixotrophic microscopic organisms, found in both marine and fresh waters. Autotrophic algae use inorganic carbon, heterotrophic use organic sources, and mixotrophic can use both, so their production (and adaptability) increases, but their need for glucose drives up the cost. An alternative is to use sugar from agricultural waste and crude glycerol from the biodiesel industry, so production costs can be made more accessible. Microalgae are believed to be a good solution to biofuel production because they are simple to grow (light, CO2, and nutrients are the only ingredients needed) and they do not need large spaces. Furthermore, hydrogen, bioethanol, and methane can be produced, making an even stronger case for the business. Their carbon sequestration potential is interesting. According to a study, every kilogram of microalgae can subtract 1.8 kg of CO2 from the atmosphere (Zhou et al., 2014), which makes it carbon net negative. Depending on the type of algae, the oil content can range from 15 to 80% dry weight (Shah et al., 2018), whereas conventional biomass for oil production has a content of 50e60%. The production rate is also favorable, allowing microalgae to produce up to 100,000 L/ha/year, which is 15 to 100 times the capacity of conventional biomass (Voloshin et al., 2016). Furthermore, the generated oil is compatible with current engines and it has properties similar to those of fossil fuels, so no additional modifications are needed (Pragya et al., 2013). An optimal solution is the coupling of microalgae growth with wastewater treatments. The sludge under treatment can be better disposed and can provide nutrients needed for the algae while CO2 concentrations can be reduced (Granada et al., 2016). In this way, a winewin situation is obtained. Bioethanol can also be produced by means of microalgae, independently using any of the three types mentioned earlier. Similarly to algaegenerated biodiesel, the production of bioethanol according to this principle is also higher compared with the conventional way, yielding

362

Distributed Renewable Energies for Off-Grid Communities

6000 gallons/acre/year compared with 400 gallons/acre/year (John et al., 2011). Major limitations in microalgae use for biofuel production are related to the high cost and the fact that the generated oil is less stable than conventional oil, and especially at high temperatures, it is likely to degrade (Shah et al., 2018). An additional minor drawback is that if biodiesel is produced, there might be slightly lower operating performance in the engine (Alalwan et al., 2019). 16.2.1.4.4 Fourth-generation biofuels

This is the most advanced generation of biofuels. It genetically modifies microorganisms such as microalgae, yeasts, fungi, and cyanobacteria to increase oil production and CO2 sequestration capacities. Technologies such as pyrolysis, gasification, upgrading, and solar to fuel pathways are also used (Cuellar-Bermudez et al., 2015), but those solutions are still at an early stage of development (Sikarwar et al., 2017). Biofuels have a considerable potential to help the transport sector in transition, but further studies are needed to optimize the yields, production processes, and economy so that biofuels can solve environmental problems, not add to them. 16.2.1.5 Electricity When thinking about how to decarbonize the transport sector, the answer seems easy: just turn electric! However, this statement, which is true, does not consider challenges that such a transition would bring. The good news is that there are many solutions that could tackle these challenges. Invented in 1834, the electric engine is a developed concept: our lives are surrounded by electric engines, from microscopic (e.g., tools at the dentist) to large units powering trains and even bigger equipment. The reason for this massive development is the characteristics of this type of engine: silent operation, high efficiency, scalability, and simplicity. An electric engine produces considerably less noise than an equivalent combustion engine. Its efficiency can be as much as three times higher and it is possible to find it at both the miniscale and megascale. Unlikely its combustion opponent, which has hundreds of moving parts, the electric engine is composed of a rotor and a stator; the rotor is the only moving part, so fewer things can go wrong. This reflects on the maintenance, which is limited.

The importance of green mobility

363

The whole industry uses electric engines for different purposes. The first cars were also electric. The first car capable of reaching 100 km/h was powered by electricity! However, as in every technology, there is a major drawback limiting its expansion in the mobility sector: electricity should be provided in some way, and because, by definition, a vehicle moves, it was impossible to use a simple electric cable. Therefore, the choice went to batteries, but their energy capacity was (and still is) insufficient to compete with cheap oil fossil fuel resources. To give an idea of the problem, if a combustion car needs to drive 100 km, a simple tank of 5 L might be sufficient. The low volume, and the fact that oil is a liquid and can take any shape, makes it convenient to install this solution on the vehicle. On the other hand, an equivalent battery-driven car would need hundreds of kilograms of batteries (with consequent space occupation) to cover the same range. Furthermore, the added weight makes the problem worse. According to a study conducted by Sandy Thomas, “To double the range of a [battery electric vehicle] from 161 km to 322 km might require the addition of 800 kg of batteries” (Sandy Thomas, 2012). Such added weight (equivalent to eight passengers) requires stronger suspension, stronger brakes, and a stronger engine, which consumes more energy. The added weight might translate to financial losses if professional vehicles are considered. Vans and truck are more sensitive to this issue because of their function, which is to carry goods, Carrying 800 kg of batteries (or 2e3 tons in the case of trucks) would reduce the available load capacity and therefore generate less income. Originally, batteries that were used were conventional lead acid, but although their price is convenient, the energy stored is limited. The industry therefore began to use lithium batteries, because the element is the lightest metal in the periodic table and also the most electropositive. This makes it a suitable candidate, because relatively high energy density can be achieved (theoretical: 380 Wh/kg; on the market: 150e210 Wh/kg) at a reasonable weight (G€ ur, 2018). Although the price has dropped in past years owing to wide use in the industry, for mobility applications it is still considerable (a car battery can easily costs thousands of euros). Furthermore, lithium is mostly located in a few locations in the world, which would switch energy dependency to different countries.

364

Distributed Renewable Energies for Off-Grid Communities

To solve these problems, research is looking for alternatives that do not involve that element. Sodium (Na), potassium (K), calcium (Ca), magnesium (Mg), and aluminum (Al) are good candidates because their properties are similar to lithium, but they are better distributed in the world. The theoretical energy density expected can reach 2840 Wh/kg with the Mg-air battery, which is still not as high as the expected density of Li-air batteries (3458 Wh/kg), but the cost of the metal is considerably lower (2.75 $/kg against 68 $/kg) (G€ ur, 2018). Flow batteries also represent a possible solution; they promise higher flexibility than conventional ones, but research is ongoing (G€ ur, 2018). Another problem to consider, which is still not resolved, is what happens to batteries when their lifetime ends. Although technologies exist for proper treatment, there is always the danger that used batteries may end up hidden underground or in some developing countries. It is therefore crucial for governments and international authorities to act strongly to prevent such practices by regulating and controlling the disposal of used batteries. Luckily, some solutions may increase the spread of electric mobility: some are technological, some logistic. Starting from the logistic, it is important for a sufficient network of charging points to be available. The average range of EVs (light duty) is 120e250 km, more than enough for most users. Studies showed that the average driving distance per person is 40ekm, depending on the country (European Commission, 2012). Therefore, the available range would theoretically be sufficient. However, other parameters should be considered in this evaluation, such as the weather and the heating and cooling equipment used, which can take 15e20% of the battery but also the cycling of the batteries. Similar to smartphones, the battery lifetime is measured in cycles, which are reduced if the battery charge goes below 20%. Considering these parameters, the available range can become excessively short, creating the so-called fear of the empty tank. A distributed infrastructure will lower this fear and make users more confident that they will reach their destination. This approach is still not followed unanimously; some regions are still not sufficiently covered. A map showing the locations of charging points around the world is available on https://map.openchargemap.io. Although the map is user-based (and therefore might not include the full list), it gives an idea on how stations are distributed around the world. Europe is investing heavily in this and passed from 122,178 normal chargers (1 kW)

Nepal, India, Peru, Trinidad and Tobago, Mexico Bangladesh (Grameen Shakti project), Africa

Bangladesh, Costa Rica, Chile West Bank (Greenstar project)

Applications for basic social services

Health clinics

Potable water pumping

Water purification

Water desalination

150- to 200-Wp electronics, deep-cycle batteries, small refrigerator/freezer 1- to 4-kWp electronics, pump, reservoir (generally no batteries needed) PV to power UV or ozone water purifiers (0.2e0.3 Wh/L) 1e2 kWp needed to power reverse osmosis or other water desalination units for 1 m3/day

Many countries (World Health Organization standards) Many countries, e.g., large project in Sahelian countries (European Union project) Many countries (e.g., China, Honduras, Mexico, West Bank) Italy, Japan, United States, Australia, Saudi United Arab Emirates

(Continued)

556

Appendix Four: Inventory of photovoltaic systems for sustainable rural development

Type of PV application

Typical system design

Existing examples

Applications in agricultural sector

Internet server for telemedicine Schools and training centers Streetlight

PV, photovoltaic.

Integrated in multifunctional solar facility (>1 kW) PV systems for powering lights, TV/VCR, and PCs 35/70 Wp, electronics, battery, one or two compact fluorescent lamps

West Bank (Greenstar project) Many countries: China, Honduras, Mexico, Philippines, Mali India, Indonesia, Philippines, Brazil, Mali, Iraq

Index Note: ‘Page numbers followed by “f ” indicate figures and “t” indicate tables.’

A ABB, 321e322 Abengoa Solar, 321e322 Accumulator, 136f Acetonebutanol-ethanol (ABE) fermentation, 187 Administration energy requirement, 70e71, 75e76 Advanced hydrogen technology, 440e441, 441f Africa, renewables in, 313e318, 313fe314f biofuels, 318 biomass, 315e316 energy efficiency, 318 geothermal energy, 316 hydropower, 315 solar power, 317, 317f wind power, 316 Agricultural energy requirement, 71, 77, 77t Agricultural produce, drying of, 101 Agriculture, 50 Air collectors, 139e140 Algae, 204e206 bioreactors, 206 Alternative resources, availability of, 31e32 American Recovery and Reinvestment Act of 2009, 279 American Wind Energy Association (AWEA), 160e161 Antimatter energy, 437 AQUA-CSP study, of DESERTEC concept, 320e322 Arid regions, Integrated Energy Farm in, 75e80 activities of, 76b energy requirement, 75e80 administration and household, 75e76 agricultural activities, 77, 78t biomass, origin of, 78

cooling load, 77b investment requirement, 80, 80b renewable energy sources, contribution of, 80 site energy production, 71e73 farm area, distribution of, 75b farm production, 75 specifications, 75 Atmospheric electricity, 437e438 Australia climate change policy, 309 closing the gap of indigenous disadvantage, 309 green energy taskforce, 309 renewable energy, in remote communities, 308e311 Renewable Remote Power Generation Program (RRPGP), 309 working futures policy, 309 Autogas, 350e351

B Barrage, tidal power generation from, 239 Basic/extended needs, of power supply, 51e53 Batch solar water heaters, 138e139 Batteries, 265e271 digital quantum, 265e266 electric vehicle (EV), 267 hydrogen storage, 268 hydrogen technology, 265e271 lithium, 265e266 photovoltaics (PV), 265 power station, Norwalk, 268f stations, 269t storage technologies, 265f Benz, Daimler, 203 Binary cycle plants, 254 Bioenergy characteristics of, 171e173, 173f potentials of, 171e173, 173f Bioethanol, 187e188. See also Ethanol

557

j

558 Biofuels production, in Africa, 318 pumps, 409e419 solid, 173e180, 174f Biogas, 106e111, 304e305, 316 biodigesters, 180 biomethane, 180e191 fuel, 416e419 landfill gas recovery systems, 180 plant, 73, 185f production of, 106e111, 107te108t road mobility, 349e380 wastewater treatment plants, 181 Biomass, 202 characteristics of, 171e173 energetic use of, 104e106 combustion, 104e105 extraction, 105e106 energy, 409e419, 410f feedstock, 166, 171 resource, exploitation of, 104 forms of, 103t origin of, 73 potentials of, 171e173 production, in Africa, 315e316 stoves, 199 Biomethane animal manure, 182 crop residues, 182 organic fraction, 183 road mobility, 349e380 thermal gasification, 181e182 upgrading biogas, 181 wastewater sludge, 183 Bio-oils, 169, 189e191 BlinkÒ charging station, 286 Blockchain better security, 449 characteristics, 448e449 decentralization, 448 decentralized autonomous organizations (DAO), 449 definition, 447 distributed generation, 449e450 efficiency, 449 immutability, 448 technology background, 449e450

Index

transparency, 448e449 Briquettes, 179 Buildings, heating and cooling loads for, 92b

C Central Europe, Integrated Energy Farm in, 65e74 energy requirement, 70e74 administration and household, 70e71 agricultural activities, 71 biomass, origin of, 73 investment requirement, 74 renewable energy sources, contribution of, 74 site energy production, 71e73 farm area, distribution of, 67e68 farm production, 70 specifications, 66 Cevital, 321e322 Chapin, Daryl, 125 Charcoal, 175e176 Chemical energy, 32, 168, 202e203, 265, 270, 272e273 Citizens’ participation, 328e330 Claude, Georges, 234 Closed-cycle ocean thermal energy conversion, 235. See also Ocean thermal energy conversion Cogeneration, 180, 204, 315 applications of, 315 combined heat and power, advantages of, 194e196 district heating networks, 273e274 Combined heat and power (CHP), 194e196 electricity, 195e196 heat, 194e195 wood/biomass, 73, 78 Combustion, 104e105 Community energy, 82e85, 327. See also Community energy, determination of food, 82e85 involvement, 23e24 ownership, of renewable energy generation, 327e328

559

Index

power, 23e24 water, 82e85 Community energy, determination of, 327 agricultural activities, 88e91 benefits of, 328 biogas production, 106e111 biomass, 102e111 energetic use of, 104e106 as energy resource, exploitation of, 104, 104t forms of, 103t citizens’ participation, 328e330 community ownership, 327e328 Danish ownership model, 330e333 data acquisition, 88 economic impacts, 334e338 energy potential analysis, 94e99 solar energy, 94e96 solar energy, exploitation of, 97 solar photovoltaic, 98e99, 99f solar thermal system, 97e98, 98f, 98t energy supply integration, public ownership, 333e334 energy utilization, data collection and processing for, 189e190 agricultural produce, drying of, 101 space heating, 100 water heating, 100 food requirement, 92e94 household energy requirement, 92 electricity, 92, 93b heat energy, 92, 92be93b modeling approaches, 85e87 photovoltaic (PV) systems, 329, 329fe330f socioeconomic benefits, 335e338 wind energy, 101e102 World Wind Energy Association (WWEA), 327 Compressed air energy storage (CAES), 274e275, 275f Compressed natural gas (CNG) vehicles, 355e356

Concentrated solar power (CSP), 125, 508 dish/engine system, 135 linear concentrator systems, 135 power tower system, 135 for seawater desalination, 322 storage, 274 Concentrating Solar Power for the Mediterranean Region (MED-CSP), study of DESERTEC concept, 322 Concentrating sunlight, 143 Containerized solar minigrid, Fanidiama village, Mali, 306e307, 306f Control technology, 455 Conventional fossil fuels, 349e350 Conventional power station vs. stabilized renewable energy power station, 311f Conversion factors, 42, 115e116 Cooking, 8 Cooling load requirement, 75e76, 77b Cyberattacks, 454e456

D Dam, tidal power generation from, 239 Darrieus wind turbine, 154e155 d’Arsonval, Jacques Arsene, 234 Data acquisition, 88 Decentralized autonomous organizations (DAO), 449 Decentralized desalination systems, 307e308 Deere, John, 191 Demand-side flexibility programs, 59 Denmark distributed integrated energy systems cogeneration, 180, 185e186, 194e195, 255e256, 315 combined heat and power, 194e196 hot water storage, 274 ownership model, 330e333 wind energy, 331e332 implementation, 298e299 ownership model, 327e328, 330e333 energy supply by public ownership, integration of, 333e334

560 Desert culture agriculture, 473e474 Al Minya, Egypt, 467e481 climate, 478 desertification, 478 desert plant adaptations, 474e481 economic/social impacts, 478 education, 478 energy forms, 476e477 energy sector, 476 energy supply, 470e471 environment, 478 farm benchmark, 468e470 farm design, 468e470 farm improvement, 471e473 farm rehabilitation, 471e473 farm upgrading, 471e473 financing sources, 479e480 funding sources, 479e480 infrastructure facilities, 471 irrigation systems, 477 irrigation water and methods, 471 minya farm cultivated crops, 470 research, 478 role of, 467e481 sponsorship sources, 479e480 training, 478 DESERTEC concept, 320e322, 507 MED-CSP study of, 322 solar oases, 320e322 Deutsche Bank, 321e322 Digitalization definition, 443 distributed renewable energy (DER), 445 exponential growth, 444 geographical information systems (GIS), 445 interconnection planning, 445 Dii GmbH, 321e322 Dish/engine system, 135 Distributed energy generation, 22, 449e450 supply, 22e23 trilemma challenges, 12e13 Distributed renewable energy (DER), 6e7, 445 challenges faced by, 42e43

Index

community involvement, 41e42 containerized solar minigrid, Fanidiama village, Mali, 306e307, 306f decentralized desalination systems, 307e308 Denmark, 298e299 DESERTEC concept, 320e322 energy and sustainable development, 40e41 Food and Agriculture Organization, 43e44 global approach to, 45e50 integrated energy communities, 43e44 energy demand, basic elements of, 48e50 with farming systems, 46f global approach to, 45e50 Iraq Dream Homes, 312e313 remote communities, Australia, 308e311 renewable energy, regional implementation of, 56e58 renewables, Africa, 313e318, 313fe314f biofuels, 318 biomass, 315e316 energy efficiency, 318 geothermal, 316 hydropower, 315 solar power, 317 wind power, 316 renewables, India, 319e320, 319f rural community of J€ uhnde, 304e305, 305f Samsø island, 299e301 solar energy, Maasai, Tanzania, 307e308 solar oases, 320e322 sustainable community, Iraq, 312f Uganda taxi-bike drivers, 303e304 VindØ energy island, 301e302, 302f renewable energy, regional implementation of, 56e58 Distribution line, 452 District heating networks, 300. See also Heat(ing) Down-regulation energy storage technologies, 275e276

Index

Dry steam plants, 252e253 Duothermic combustion system, 191

E Economic trends, of energy ownership, 334e335 E-Energy project, 282 E’lectricité de France, 321e322 Electricity biofuels, 362e368 distribution grid application, 454 control technology, 455e456 coordination, 455e456 coordination and control technology, 455 cyberattacks, 454e456 definition, 451 development direction, 454e456 distribution line, 452 innovations, 454e455 renewables penetration, 453e454 smart distribution networks, Iraqi provinces, 452e453 transmission line, 452 energy equivalent, 115e116 energy generation, 116e117 fossil fuel, 115 gross electricity production, 114 hydroelectricity, 320 net electricity production, 114 nonefossil fuel sources, 115e116 nuclear sources, 115 sales, 114e115 Electricity Networks of the Future, European Technology Platform for, 280 Electric power, 8 Electric vehicles, 284e292 battery electric vehicles, 290e291 current developments, 284e287 future developments, 292e294 global battery electric vehicle, 287e290, 288f hybrid electric vehicles, 292 plug-in hybrid electric vehicles (PHEV), 287e292

561 types, 287, 290e292 Electrofuels, 376e377 ammonia, 376 carbon electrofuels, 376e377 Electrolysis, 272 Elsbett engine, 191 Empowering urban cities, Africa Agona Swedru, Ghana, 502 data acquisition, 503 evaluation, 503 Khartoum, Sudan, 502 Marondera, Zimbabwe, 502 objectives, 501 processing, 503 resilience implementation, 503e504 results analysis, 503 scientific project objective presentation, 504e505 survey, 503 technical framework, 502e503 urban resilience strengthening, 500e501, 500f ENEL, 321e322 Energy bioenergy characteristics of, 171e173 potentials of, 171e173 chemical, 168 community. See Community energy, determination of consumption, 114 conversion factors for, 115e116 defined, 114 distributed, 22 generation, 116e117 supply, 22e23 efficiency, in Africa, 313e318 equivalents, conversion factors and, 115 generation, 116e117 geothermal. See Geothermal energy global contribution of, 117e120 gravitational, 270 kinetic, 149, 152, 270 marine. See Marine energy plant species arid climate, 55e56 semiarid climate, 55e56

562 Energy (Continued ) subtropical climate, 56 tropical climate, 56 potential, 94e99 rating, 114 renewable. See Renewable energy solar. See Solar energy supplies, current, 29 thermal, 270 wind. See Wind energy Energy demand basic elements of, 48e50 agriculture, 50 communications, 50 cooking, 49 electric power, 48 health, 50 heat, 48 lighting, 49 maintenance workshops, 50 mobility, 50 sanitation, 50 small markets industries, 50 water, 48e49 typical, 51e52 Energy Independence and Security Act of 2007, 278 Energy ownership, 339e340 community ownership, 327e328 Danish ownership model, 330e333 economic trends of, 334e335 energy supply by public ownership, integration of, 333e334 Energy potential analysis, 94e99 solar energy, 94e96 exploitation of, 97 solar photovoltaic, 98e99 solar thermal system, 97e98 Energy storage, 264e276 down-regulation, technologies for, 275e276 methods of, 271e275 up-regulation, technologies for, 275e276 E.ON, 321e322 Ethanol, 186e189 bioethanol, 188 definition, 186

Index

ethyl tert-butyl ether (ETBE), 187e188 methyl tert-butyl ether (MTBE), 187e188 oxygenated fuel feedstock, 188e189 Ethyl alcohol. See Ethanol Ethyl tertio butyl ether (ETBE), 187e188 European Organization for Nuclear Research (CERN), 437 European Technology Platform (ETP), for Electricity Networks of the Future, 282 European Union, Middle East and North Africa (EU-MENA) DESERTEC concept, 321f, 322 solar oases, 320e322 Evacuated tube solar collectors, 139 Extraction, 105e106 oil, 106f

F Farm area, distribution of, 67e68, 75b Farm irrigation, 429e430 Farm production, 70, 75 Feed-in Tariff (FIT), 22 First-generation biofuels, 358e360 Flash steam plants, 253, 254f Flat plate collector, 97, 138e139 unglazed, 139 Flywheel energy storage system, 271 Food and Agriculture Organization (FAO), 43, 83, 84f Food requirement, 50 Fourth-generation biofuels, 362 Francis turbines, 224e226 Fritts, Charles, 125 Fuel cell vehicles (FCVs), 202e203 Fuel cells, 202e203 Fuel requirement, 72t, 91b Fuller, Calvin, 125 Fusion power International Thermonuclear Experimental Reactor (ITER), 436, 436f nuclear fusion, 436 Future energy generation and supply, restructuring, 27e37

563

Index

alternative resources, availability of, 31e32 challenges to, 27e29 current and futureenergy supplies, 29 peak oil, 30, 31f

G Gasification, 167e168, 181e182, 197e199, 307, 319e320, 334 biomass stoves, 199 Geographical information systems (GIS), 445 Geothermal energy, 247e261, 253f applications of, 255 benefits, 257e258 definition, 247e249 direct use of, 255 economic costs, 257e258 enhanced geothermal systems (EGS), 259 environmental effects, 257e258 future of, 258e260 geothermal electricity, 252e256, 253f geothermal heat pumps (GHP), 250e251, 251f history of, 249e250 production, in Africa, 313e318. See also Thermal energy temperature, 248f Geothermal heat pumps, 250e251. See also Heat(ing) Geothermal power plants, 249e250, 252e254, 257 binary cycle plants, 254 dry steam plants, 252e253 flash steam plants, 253 Germany energy production process, 298, 308 J€ uhnde, 304e305, 305f Solar Park Finsterwalde, 128e129 Giromill, 155 Global contribution, of energy consumption, 117e120 Global market, 155e157 Gravitational energy, 270 Green electricity, 24 Greenhouse effect, 143

Green mobility (transport sector) air pollution and health, 348e349 aviation, 347 biofuels, 357e362 electricity, 362e368 first-generation biofuels, 358e360, 359f fourth-generation biofuels, 362 second-generation biofuels, 360 third-generation biofuels, 361e362 CO2 impact, 346e348 electricity, 345e346 electrofuels, 376e377 ammonia, 376 carbon electrofuels, 376e377 environmental impacts, 346e349 heavy-duty transportation, 379e380 hybrids, 372e376, 373f, 375f hydrogen, 368e372, 370f, 372f internal combustion engine (ICE), 345 light-duty transportation, 377e379 rail mobility, 380e381 rethinking mobility, 383e386, 385f road mobility, 349e380 autogas, 351 biogas, 354e355 biomethane, 354e355 compressed natural gas (CNG) vehicles, 355e356 conventional fossil fuels, 349e350 liquefied petroleum gas (LPG), 350e352 methane, vehicles running on, 355e357, 356f synthetic natural gas, 353e354 technologies available, 349e377 unconventional fossil methane, 354 social impacts, 346e349 water mobility, 381e382, 382f Grid Modernization Commission, 278 GridSMART, 290

H Health, 50 Heat(ing), 113, 115, 118e119, 121, 124e125, 139, 194e195 converting light to, 143

564 Heat(ing) (Continued ) district, 171, 249, 274, 299e300, 329 energy, 70e71, 77 geothermal heat pumps, 250e251 space, 121, 139 trapping, 143 water, 121 Heavy-duty transportation, 379e380 Horizontal-axis wind turbines (HAWT), 153. See also Wind turbine Hot water storage, 97, 139, 274 Household, 92 electricity, 92 heat energy, 92, 92be93b HSH Nordbank, 321e322 Hybrid electric vehicles (HEVs), 292 Hybrid ocean thermal energy conversion, 287e290. See also Ocean thermal energy conversion Hydroelectricity, 320 pumped-storage, 274, 275f Hydroelectric power plants large hydropower, 220 microhydropower, 220 sizes of, 220 small hydropower, 220 Hydrogen, 206e207 Hydropower, 213e230 assessment, 226e228 climate change, 229e230 diversion plants, 218 hydroelectric power plants, 220 hydropower energy, 215e217 impoundment plants, 217, 217f pumped storage plants, 218e219, 219f turbine types, 220e225

I Immutability, 448 India, renewables in, 319e320, 319fe320f Innovations, 454e455 Insolation, 126 Integral collector-storage systems. See Batch solar water heaters Integrated Energy Communities (IEC), 43e44

Index

energy demand, basic elements of, 48e50, 49f global approach to, 45e50, 46f pathways of, 45 Integrated Energy Farm (IEF) implementation of, 65 in arid regions. See Arid regions, Integrated Energy Farm in in Central Europe. See Central Europe, Integrated Energy Farm in in semi-arid regions. See Semi-arid regions, Integrated Energy Farm in planning of, 65, 66f scenarios, 64e65 Integrated energy settlement Rousse, Bulgaria biomass potential, 488 collected data analysis, 486e488 energy demand, 487 energy plantation implementation, 489 energy supply, 487 integrated energy farm concept, 483e485 local authorities contribution, 490 objectives, 483 project contribution, 490 project outlines, 488e490 requirements, 489 site description, 485e486 solar energy, 487 Sustainable Rural Environment and Energy Network (SREN), 482 wind energy, 487 working plan, 485e486 Wierthe, Germany, 460e467 education and training facilities, 465f electric cars, 464f medium and small-sized enterprises, 465f rural community, 463f solar panels sheep grazing, 464f Solar park, 464f United Nations concept, 461f Integrated renewable energy farm (IREF), 43, 49f, 57f Interconnection planning, 445 Internal combustion engine (ICE), 345

565

Index

Internal combustion engine vehicle (ICEV), 202e203, 284e285, 287 International Renewable Energy Agency, 394e395 International Solar Energy Society, 334e335 International Thermonuclear Experimental Reactor (ITER), 436, 436f Internet of Energy, 282, 282f Investment requirement, for integrated energy farm, 74, 74b, 80, 80b Iraq Dream Homes, 312e313, 312f Iraq Energy Solutions, LLC (IES), 312 Island grid single, 52 system solution for, 53 arid climate regions, 55e56 semi-arid climate regions, 55e56 sub-tropical climate regions, 56 temperate climate regions, 53e55 tropical climate regions, 53e55 three-phase, 52 Sunny Island, 52

J J€ uhnde-Bioenergiedorf 2.0, 305 J€ uhnde, Germany, 304e305, 305f

K Kaplan turbine, 225e226 Kinetic energy, 270 Knies, Gerhard, 320e321

L Lange, Bruno, 125 Large hydropower, 220 Light-duty transportation, 377e379 Light to heat, converting, 143 Lighting, 49 Linear concentrator systems, 135 Liquefied petroleum gas (LPG), 350e352 Liquid biofuel, 412e416 Lithium batteries, 272. See also Batteries

M Maintenance workshops, 50

MAN Solar Millennium, 321e322 Marine energy, 231e233, 233f, 233t economic challenges to, 243e244 environmental challenges to, 243e244 ocean thermal energy conversion, 233e237 advantages and disadvantages of, 237e238 closed-cycle, 235, 235f hybrid, 236e237, 236f open-cycle, 236 ocean tidal power, 239e240, 239fe241f ocean wave power, 240e243, 241f offshore systems, 242, 242f onshore systems, 242e243 Marine power. See Marine energy Maxwell, James Clerk, 125 Methanol, 202 Methyl tert-butyl ether (MTBE), 187e188 Microalgae, 438e439 Microhydropower, 220 Mobility, 50 Molten salt, 274 Multifunctional Oven (MFO), 199, 200f Munich Re, 321e322 M and W Zander Holding, 321e322

N National Renewable Energy Laboratory, 129, 130f Natural Energy Laboratory of Hawaii Authority, 234 Nuclear fusion, 436

O Ocean energy. See Marine energy Ocean thermal energy conversion (OTEC), 233e237 advantages and disadvantages of, 237e238 closed-cycle, 235, 235f hybrid, 236e237, 236f open-cycle, 236 Ocean tidal power tidal fence, 240 tidal turbines, 239e240, 240f

566 Ocean tidal power (Continued ) tide barrage, 239, 239f Ocean wave power, 240e242, 241f offshore systems, 242, 242f onshore systems, 242e243 OE Buoy, 240, 241f Off-grid electrification, 24 supply, 22e25 systems, 24, 25fe26f Off-peak electric energy, storage of, 271 Offshore wave-power systems, 242 Offshore wind farm Dogger Bank, 157e158 Oil extraction, 105e106, 106f. See also Extraction One Alliance Partners (OAP) Iraq Dream Homes, 312e313, 312f Iraq Energy Solutions, 312 Onshore wave-power systems, 242e243 Open-cycle ocean thermal energy conversion, 236. See also Ocean thermal energy conversion Organic dry matter, concentration of, 109e111, 111b Oscillating water column, 242e243 Osmotic power, 439e440, 439f Ownership community energy, 327e328 renewable energy, 339e340

P Peak oil, 30, 30fe31f Pearson, Gerald, 125 Pellets, 180 Pelton wheel, 224 Photovoltaic (PV) systems, 73, 78, 125e127, 309, 329, 329fe330f, 394e395 applications of, 128e131 solar, 98e99, 99fe100f Plug-in electric vehicles (PEV), 279, 286 Potential energy, 213e214 Power generation, 121 Power tower system, 135 Production energy requirement for, 91, 91b

Index

optimization model, 64 preference model, 65 Propeller turbines, 224e225 Proton exchange membrane (PEM) fuel cell, 202e203 Pumped-storage hydroelectricity, 152 Pyrolysis, 200e202, 201f

R Rail mobility, 380e381 Reactor volume, 109e111, 111b Red Eléctrica de Espa~ na, 321e322 Regional implementation, 56e58 Remote communities, Australia, 308e311 Renewable energy, 31 in Africa. See Africa, renewables in in Australia, 308e311, 311f community ownership of, 327e328 costs of, 341e343, 342t defined, 114 in Denmark. See Denmark economic impacts, 335e338 farm irrigation, 429e430 generation, 338, 396f global contribution of, 118fe120f, 340e341, 342f in India. See India, renewables in International Renewable Energy Agency, 394e395 ownership, 339e340 photovoltaic (PV) systems, 394e395 physical potential of, 31, 32f power plants, 393e394 power station vs. conventional power station, 310, 311f pumps, 397e419 biofuel pumps, 409e419 biogas fuel, 416e419 biomass energy, 409e419, 410f example, 408e409, 418e419 liquid biofuel, 412e416 selection, 397e399, 398f solar pumps, 399e403, 400f solid biofuel, 411e412 system head curve, 398f wind pumps, 403e409

Index

regional implementation of, 56e58, 57fe58f socioeconomic benefits, 335e338 sources of, 31, 50 uses of, 394f wastewater treatment, 427e429, 427f water desalination, 422e426, 423te425t water projects, 395 water purification, 419e422, 420f Renewable Remote Power Generation Program (RRPGP) (Australia), 309 Renewables, Africa, 313e318, 313fe314f biofuels, 318 biomass, 315e316 energy efficiency, 318 geothermal, 316 hydropower, 315 solar power, 317 wind power, 316 Renewables, India, 319e320, 319f Rethinking mobility, 383e386 Reverse Archimedes’ screw, 220 Road mobility, 349e380 Run-of-the-river systems, 217 Rural community, J€ uhnde, 304e305, 305f

S Samsø island, 299e301 Sanitation, 50 Savonius wind turbine, 155 Schott Solar, 321e322 Seawater desalination, concentrating solar power for, 322 Second-generation biofuels, 360 Semi-arid regions, Integrated Energy Farm in activities of, 75, 76b energy requirement administration and household, 75e76, 77t, 77b agricultural activities, 77, 78t biomass, origin of, 78 cooling load, 76, 77b farm, energy production, 77e78, 79t investment requirement, 80, 80b

567 renewable energy sources, contribution of, 80 farm area, distribution of, 75, 75b farm production, 75, 76t, 76b specifications, 75, 75b Siemens, 321e322 Siemens, Ernst Werner von, 125 Silicon solar cell, 125 Single island grid. See Island grid Site energy production, 71e73, 73t Small hydropower plants, 220 Small markets industries, 50 Small wind turbines, 158e162, 159fe160f market growth, 161e162, 162f Smart distribution networks, Iraqi provinces, 452e453 Smart grids defined, 276e278 European strategy, 280e282, 282f importance of, 276e278 Korean version, 282e284, 284f smart meters, 278 United States version, 278e280 Smart meters, 278 Solar cookers advantages, 143e145 disadvantages, 143e145 Solar drier, 97e98 Solar energy concentrated solar power (CSP), 125, 274 exploitation of, 97 Maasai, Tanzania, 307e308 photovoltaic systems, 125e127, 127fe128f solar cooker, 141e146 solar thermal collectors, 136e141, 140fe142f Solar Energy Research Institute. See National Renewable Energy Laboratory Solar hot water (SHW) systems, 136e141 Solar oases, 320e322, 321fe322f, 508e509 Solar owen. See Solar cooker Solar photovoltaics, 98e99, 99fe100f. See also Photovoltaic system Solar power, production of in Africa, 317, 317f

568 Solar power, production of (Continued ) in Vatican City, 323e324, 323fe324f Solar pumps, 399e403, 400f Solar thermal collectors, 136e141, 140fe142f Solar thermal system, 97e98, 98f, 98t Solar water heating (SWH) system, 136e141, 140fe142f SolarWorld AG, 323 Solid biofuels, 174f, 411e412 in Austria, 176e179 briquettes, 179 charcoal, 175e176 pellets, 180 Space heating, 100. See also Heat(ing) Space load, 109e111, 111b Steam technology, 196e197 Stirling engine, 73, 78, 204 dishes, 125 Sustainable community, Iraq, 312f Sustainable development, 40e41 Kiga village, Iran, 492e495 biomass and food, 497 commercial characteristics, 495 fuel conservation, 499 heat supply, 496e497 identification, 497 modeling approaches, 495 project implementation, 497e499 social characteristics, 495 strategic analysis, 493e495 typical user’s appliances, 495e496 Sustainable Rural Energy Network (SREN), 43 Synthetic natural gas, 353e354 Synthetic oil, 202 System head curve, 398f

T Temperate climate, 53e55 Thermal energy, 234, 247. See also Geothermal energy Thermal solar collectors, 73, 78 Third-generation biofuels, 361e362 Three-phase island grid, 52. See also Island grid Tidal fence, 240

Index

Tidal turbine, 239e240 Tilt surface, radiation falling on, 95, 95b Title XIII (2007), Section 1301 of, 279 Transesterification, 189e190 Transmission line, 452 Transport fuels, 121 Trapping heat, 143. See also Heat(ing) Trieb, Franz, 322 Turbines Archimedes’ screw, 225 cross-flow turbines, 224 Francis turbines, 224e225 gravity turbine, 220 impulse turbine, 220 Kaplan turbines, 224e225 pelton turbine, 224 reaction turbine, 220 relative efficiencies, 225e226 turgo turbines, 224 types, 220e225, 221te223t waterwheel turbines, 225 wind, 153e155 Twisted Savonius, 155

U Uganda taxi-bike drivers, 303e304 Ultra-capacitor, 272 Unconventional fossil methane, 354 Unglazed flat-plate collectors, 139. See also Flat plate collector United Nations Environment Program (UNEP), 336 Upregulation energy storage, technologies for, 275e276, 276f

V Vatican City, solar power plant in, 323e324, 323fe324f Veba process, 191 Vegetable oils, as fuels, 105e106 Vertical-axis wind turbines, 153e154 Darrieus, 154e155 Giromill, 155 Savonius, 155 Twisted Savonius, 155. See also Wind turbine VindØ energy island, 301e302, 302f

Index

W Wankelmotor, 191 Wastewater treatment, 427e429, 427f Water, 48e49 desalination, 422e426, 423te425t heating, 121 mobility, 381e382, 382f projects, 395 pumped energy storage, 274 purification, 419e422, 420f Watereenergyefood nexus, 82f community requirements, 82e85 definitions, 84e85 Food and Agriculture Organization of the United Nations (FAO), 83 global projections, 83 modeling approaches, 85e87 Wind energy, 101e102, 102b, 149e163. See also Wind turbine annual value of, 101, 102b generation, community ownership of, 327e328

569 production, in Africa, 316 Wind generator, total efficiency of, 101, 102b Wind pumps, 403e409 Wind turbine conversion of, 150 global market for, 155e157, 156t, 157fe158f horizontal-axis, 153 small, 158e162, 159fe160f vertical-axis, 153e154 Darrieus, 154e155 Giromill, 155 Savonius, 155 Twisted Savonius, 155 wind power, 149e152 Windmill, 102, 152, 332 Wireless sensor nodes, 292e293 Wood Alcohol. See Methanol World Wind Energy Association (WWEA), 327