You are on page 1of 261

Renewable Energy

Technologies and
Water Infrastructure
Advancing Renewable Energy Technologies Committee

Edited by

S. Rao Chitikela, Ph.D., P.E., P.Eng.


Venkata Gullapalli, Ph.D.
William F. Ritter, Ph.D., P.E., D.WRE
Renewable Energy
Technologies and Water
Infrastructure

Prepared by
Advancing Renewable Energy Technologies Committee

Edited by
S. Rao Chitikela, Ph.D., P.E., P.Eng.
Venkata Gullapalli, Ph.D.
William F. Ritter, Ph.D., P.E., D.WRE

Published by the American Society of Civil Engineers


Library of Congress Cataloging-in-Publication Data
Names: Advancing Renewable Energy Technologies Committee, author. | Chitikela, S. Rao,
editor. | Gullapalli, Venkata, editor. | Ritter, William F., editor.
Title: Renewable energy technologies and water infrastructure
Description: First edition. | Reston : American Society of Civil Engineers, 2022. | Includes
bibliographical references and index. | Summary: “Renewable Energy Technologies
and the Water Infrastructure provides an in-depth look at policy, regulation, and the
development and application of renewable energies into existing water infrastructure”--
Provided by publisher.
Identifiers: LCCN 2021022251 | ISBN 9780784415856 (paperback) | ISBN 9780784483664
(ebook)
Subjects: LCSH: Waterworks--Technological innovations--United States. | Waterworks--
Energy conservation--United States. | Waterworks--Environmental aspects--United States. |
Water--Purification--Equipment and supplies--Technological innovations--United States. |
Renewable energy sources.
Classification: LCC TD485 .E58 2021 | DDC 628.10973--dc23
LC record available at https://lccn.loc.gov/2021022251
Published by American Society of Civil Engineers
1801 Alexander Bell Drive
Reston, Virginia 20191-4382
www.asce.org/bookstore|ascelibrary.org
Any statements expressed in these materials are those of the individual authors and do not
necessarily represent the views of ASCE, which takes no responsibility for any statement made
herein. No reference made in this publication to any specific method, product, process, or
service constitutes or implies an endorsement, recommendation, or warranty thereof by ASCE.
The materials are for general information only and do not represent a standard of ASCE, nor are
they intended as a reference in purchase specifications, contracts, regulations, statutes, or any
other legal document. ASCE makes no representation or warranty of any kind, whether express
or implied, concerning the accuracy, completeness, suitability, or utility of any information,
apparatus, product, or process discussed in this publication, and assumes no liability therefor.
The information contained in these materials should not be used without first securing
competent advice with respect to its suitability for any general or specific application. Anyone
utilizing such information assumes all liability arising from such use, including but not limited
to infringement of any patent or patents.
ASCE and American Society of Civil Engineers—Registered in US Patent and Trademark
Office.
Photocopies and permissions. Permission to photocopy or reproduce material from ASCE
publications can be requested by sending an email to permissions@asce.org or by locating a
title in the ASCE Library (https://ascelibrary.org) and using the “Permissions” link.
Errata: Errata, if any, can be found at https://doi.org/10.1061/9780784415856.
Copyright © 2022 by the American Society of Civil Engineers.
All Rights Reserved.
ISBN 978-0-7844-1585-6 (print)
ISBN 978-0-7844-8366-4 (PDF)
Manufactured in the United States of America.
27 26 25 24 23 22    1 2 3 4 5
Contents

Preface...........................................................................................................................................ix

Acknowledgments................................................................................................................. xiii

List of Authors and Reviewers............................................................................................ xv

Chapter 1 US Renewable Energy Policy—Analysis and


Recommendations..................................................................... 1
Alexander Krokus (Deceased)
Introduction.........................................................................................................................1
National Energy Act of 1978 (H.R. 8444 1977-78).........................................2
Energy Policy Act of 2005.....................................................................................2
Energy Independence and Security Act of 2007.........................................3
Food, Conservation, and Energy Act of 2008................................................3
Clean Power Plan of 2015......................................................................................3
Energy Efficiency Resources Standards...........................................................4
Renewable Energy Projects.................................................................................5
Federal Investment Tax Credit............................................................................5
Renewable Production Tax Credit.....................................................................5
Modified Accelerated Cost Recovery System................................................6
Renewable Energy Resources—Review/Recommendations.............................6
Hydropower..............................................................................................................6
Biofuels........................................................................................................................7
Solar Photovoltaic...................................................................................................8
Wind Energy..............................................................................................................8
Future of United States Renewable Energy Policy......................................9
Summary.............................................................................................................................11
References......................................................................................................................... 12

Chapter 2 Renewables and Regulatory Requirements of the


United States............................................................................... 17
S. Rao Chitikela
Introduction.......................................................................................................................17
US-Code, Law, and Act—Renewable Energy........................................................ 19
Renewable Energy Standards..................................................................................... 20
Renewable Portfolio Standards—The State of Connecticut, Example........ 25
Renewable Portfolio Standards Eligibility—The State of California,
Example................................................................................................................... 27

iii
iv Contents

Local Government on Renewable Energy Projects—State of Virginia,


Example................................................................................................................... 28
Solar and Wind Energy Rule—Bureau of Land Management........................ 31
RE Programs—Bureau of Ocean Energy Management..................................... 33
Renewables—Federal Energy Regulatory Commission................................... 34
Summary............................................................................................................................ 35
Disclaimer.......................................................................................................................... 35
References......................................................................................................................... 36

Chapter 3 Biofuels: Ethanol and Biodiesel.............................................. 39


William F. Ritter
Ethanol................................................................................................................................ 39
Introduction........................................................................................................... 39
Legislation............................................................................................................... 40
Classes of Ethanol................................................................................................. 41
Processing of Corn Ethanol............................................................................... 42
Cellulosic Ethanol Processing........................................................................... 43
US Ethanol Production.......................................................................................44
Greenhouse Gases............................................................................................... 46
Water Quality Impacts........................................................................................ 48
Pros and Cons of Ethanol................................................................................... 52
Biodiesel............................................................................................................................. 53
Introduction........................................................................................................... 53
Biodiesel Processing............................................................................................ 54
Classes of Biodiesel Blends............................................................................... 55
US Biodiesel Production.................................................................................... 55
Summary............................................................................................................................ 57
References......................................................................................................................... 57

Chapter 4 Micro-Hydropower: Concept, System Design,


and Innovations........................................................................ 61
Tamim Younos, Juneseok Lee
Introduction...................................................................................................................... 61
Micro-Hydropower Generation: Concept.............................................................. 62
Micro-Hydropower System Design..........................................................................64
Case Study Site......................................................................................................64
Micro-Hydropower System Design Components.................................... 65
Hydraulic Component Design......................................................................... 65
Mechanical Component Design..................................................................... 68
Electrical Component Design.......................................................................... 71
Case Study Project Cost..................................................................................... 75
Cost–Benefit Analysis.......................................................................................... 75
Regulatory Requirements—Case Study Site.............................................. 76
Micro-Hydropower: Viable Technology in Developing Countries................ 76
Contents v

S mall-Stream Micro-Hydropower: Challenges and Limitations..................... 78


Emerging Micro- and Small-Hydropower Technologies.................................. 79
Summary............................................................................................................................ 80
Appendix A. Bill of Materials (Equipment, 2013 Prices)..................................... 81
Appendix B. Bill of Materials (Supplies/Materials, 2013 Prices)....................... 88
Acknowledgments......................................................................................................... 90
Disclaimer.......................................................................................................................... 90
References......................................................................................................................... 90

Chapter 5 Biogas-to-Energy—The Combined Heat and Power (CHP)


Systems...................................................................................... 93
S. Rao Chitikela, William F. Ritter
Introduction...................................................................................................................... 93
Biogas Generation Systems—Theory and Practice............................................ 94
Anaerobic Digestion of Livestock and Poultry Manure.................................... 96
Sludges/Biosolids and Biogas-to-Energy—US Regulations.......................... 101
Biogas-to-Electric Generation and Combined Heat and Power..................102
Economics of Livestock and Poultry Manure Anaerobic Digestion............105
Biogas in the Circular Economy...............................................................................107
Summary.......................................................................................................................... 110
References........................................................................................................................111

Chapter 6 Fuel Cells for Renewable Wastewater Infrastructure.......... 113


Bhuvan Vemuri, Govinda Chilkoor, Jawahar Kalimuthu, Ammi
Amarnath, James E. Kilduff, Venkataramana Gadhamshetty
Introduction.................................................................................................................... 113
Hydrogen......................................................................................................................... 116
Biological Processes for Hydrogen Production....................................... 116
Dark Fermentation..............................................................................................117
Microbial Electrolysis Cells.............................................................................. 118
Photofermentation............................................................................................ 119
Biophotolysis........................................................................................................ 120
Fuel Cells.......................................................................................................................... 121
Hydrogen Fuel Cells........................................................................................... 124
Alkaline Fuel Cell.................................................................................................125
Phosphoric Acid Fuel Cell................................................................................ 126
Hydrogen and Hydrocarbon Fuel Cells...................................................... 127
Molten Carbonate Fuel Cells.......................................................................... 127
Solid Oxide Fuel Cells........................................................................................ 129
Microbial Fuel Cell..............................................................................................130
Summary.......................................................................................................................... 131
Acknowledgments....................................................................................................... 131
References....................................................................................................................... 131
vi Contents

Chapter 7 Sustainable Desalination Using Renewable


Energy Sources....................................................................... 135
Veera Gnaneswar Gude
Introduction.................................................................................................................... 135
Desalination Technologies—Principles of Operation.....................................136
Multistage Flash Desalination........................................................................136
Multieffect Distillation......................................................................................136
Vapor Compression........................................................................................... 137
Energy Efficiency of Thermal Desalination............................................... 137
Reverse Osmosis................................................................................................. 137
Renewable Energy Integration with Desalination Processes........................138
Selection Process of the Desalination Process................................................... 142
Sustainability of Desalination Technologies.......................................................144
Environmental Impacts of Desalination Processes................................144
Brine Disposal......................................................................................................144
Economic Considerations of Renewable Energy-Driven
Desalination Processes.......................................................................146
Regulatory Requirements............................................................................... 147
Social Aspects of Desalination Processes.................................................. 147
Summary.......................................................................................................................... 147
References.......................................................................................................................148

Chapter 8 Geothermal Energy................................................................ 151


Audrey Angelos, Guangdong Zhu
General Description..................................................................................................... 151
Geothermal Resources................................................................................................ 152
Geothermal Resource Applications........................................................................ 152
Geothermal Power Generation................................................................................154
Power Plant Types.........................................................................................................154
Dry Steam..............................................................................................................154
Flash .................................................................................................................... 155
Organic Rankine Cycle/Binary........................................................................156
Current Status.................................................................................................................156
Worldwide Capacity..........................................................................................156
Technological Distribution.............................................................................. 157
Geothermal Direct Use................................................................................................ 157
Direct Heating Applications......................................................................................158
Mineral Recovery...........................................................................................................158
Energy Storage............................................................................................................... 159
Future Direction............................................................................................................ 159
Enhanced Geothermal Systems.................................................................... 159
Hybridization Opportunities with Concentrating Solar Power.........160
Perspective......................................................................................................................160
References....................................................................................................................... 161
Contents vii

Chapter 9 Wind Energy—Increasing Resilience in Water


Infrastructure.......................................................................... 163
Pamela A. Menges
Introduction....................................................................................................................163
Wind Turbine Technologies.......................................................................................165
Wind Technology Powering Mechanical Systems............................................ 167
Efficiency and the Betz Limit.................................................................................... 170
Assessing Power from Wind...................................................................................... 171
Turbine Technology, Applications, and Siting.................................................... 172
Built Environment Wind Turbine............................................................................. 173
Quantifying Turbulence and Turbulence Intensity........................................... 174
Energy and Water.......................................................................................................... 175
Role of Wind Energy in Water Infrastructure Security..................................... 178
Emerging Technologies.............................................................................................. 179
Summary.......................................................................................................................... 181
References....................................................................................................................... 181

Chapter 10 Solar Energy and Water/Wastewater Infrastructure......... 183


Venkata Gullapalli
Introduction....................................................................................................................183
Solar Radiation...............................................................................................................183
Solar Photovoltaics.......................................................................................................184
Solar Thermal..................................................................................................................185
Solar Photovoltaics and Concentrating Solar Power Comparison..............186
Solar in Water Industry................................................................................................187
Desalination....................................................................................................................189
Thermal Technologies......................................................................................189
Membrane Technologies.................................................................................190
Solar in Desalination....................................................................................................190
Solar Water Disinfection............................................................................................. 191
Wastewater Processing............................................................................................... 193
Solar in Wastewater Processing...............................................................................194
Summary..........................................................................................................................195
References.......................................................................................................................196

Chapter 11 Renewable Energy Technologies for Water Quality


Monitoring............................................................................199
Varun K. Kasaraneni
Introduction....................................................................................................................199
Solar-Powered Monitoring Networks....................................................................201
Stormwater Sampling and Monitoring......................................................201
Discharge and Nonpoint Source Monitoring...........................................202
Water Quality Monitoring with Data Buoys..............................................205
Microbial Fuel Cells for Water Quality Monitoring............................................206
viii Contents

Summary.......................................................................................................................... 211
References....................................................................................................................... 211

Chapter 12 Integrating Renewable Energy in Water Infrastructure:


Global Trends and Future Outlook.........................................213
Juneseok Lee, Tamim Younos
Introduction.................................................................................................................... 213
Background..................................................................................................................... 214
Water and Energy Nexus................................................................................. 214
Water and Energy Conservation................................................................... 215
Applications of Renewable Energy in the Water Industry............................. 215
Solar Energy......................................................................................................... 215
Wind Energy......................................................................................................... 217
Outlook............................................................................................................................. 218
Summary..........................................................................................................................220
References.......................................................................................................................221

Appendix......................................................................................................225
List of Acronyms and Abbreviations......................................................................225

Index............................................................................................................. 231
Preface

Renewable Energy Technologies (RETs) Task Committee of the Environmental


and Water Resources Institute (EWRI) was formed in 2016 with the objectives
toward accomplishing on water infrastructure and the application of
sustainability and resilience requirements. Thus, this task committee has been
working on advancement on knowledge of field-proven RETs for the operation
of water infrastructure meeting the triple bottom line. The task committee has
successfully accomplished the selection and writing on various renewable energy
(RE) technologies and proven application to water infrastructure. Following the
task committee’s success, a new full standing committee, called the Advancing
Renewable Energy Technologies Committee (ARETC) of EWRI, has now been
formed.
The planning and inclusion of 12 chapters of this book encompass the
applicable and critical review details on: RE policy and regulatory requirements;
micro-hydro power; biofuels; biogas-to-energy; fuel cells for clean water;
sustainable desalination; geothermal energy; solar and wind energy toward a
resilient water infrastructure; the application of renewables for the monitoring of
water quality; and renewable energy applications to water infrastructure.
We, ARETC, pay our sincere and highest respects to Mr. Alexander
Krokus—who has been an inspiration, provided constant encouragement,
and watched of high expectations for the ARETC—on the sad demise,
and we pray for his great soul rest in peace.
The book chapters put forward knowledge on the application(s) of renewables
toward effectively operating water infrastructure into the future. The authors
converged on the following aspects on invaluable and sound policy, regulation,
science, and engineering with respect to the development and application of
renewables:
• To gear up on renewables, subsidies on fossil-fuel energy applications must be
significantly reduced and an immediate boost to RE funds should be provided;
and this approach would help safeguard humankind from ongoing uncertain
weather patterns and havoc. The federal agency efforts, such as by the Bureau
of Land Management (BLM), Bureau of Ocean Energy Management (BOEM),
and Federal Energy Regulatory Commission (FERC), based on the effective
RE legislative actions, are appreciative (to date) in terms of the production and
consumption of RE at more than 9.5 quintillion J (9.0 quadrillion Btu) and the
resultant significant reduction of greenhouse gases (GHGs). Moreover, the
Energy Independence and Security Act (EISA 2007) has been instrumental
in the production of renewable fuels (Chapters 1 and 2).

ix
x Preface

• Chapter 3 contributed by internationally renowned author highlights that the


use of soybean oil (one of the vegetable oils) stood at 57% (in 2019) to produce
biodiesel, and the market opportunity was rated at 6.8 billion L (1.8 billion
gal.) per year. However, the long-term goal for clean renewable fuels was set
(in 2007) at 136 billion L (36 billion gal.) per year.
• Chapter 4 includes the following recommendations: the small-stream micro-
hydropower generation technologies are “mature technologies,” facilitating
distributed or decentralized RE technology and providing a positive
environmental impact compared with that of large-dam hydropower; and
they must follow the local, state, and national environmental and energy
regulatory guidelines to accomplish potential success in micro-hydropower.
• According to the current and historically well-known sources, Chapter 5
fortifies the effective use of municipal wastewater sludges (the by-product
solids as generated from wastewater to clean water processing), fats–oils–
grease (FOG) feedstocks, vegetable and food wastes, livestock manure(s),
and other (highly) biodegradable municipal and industrial wastes to the
successful operation of biogas-to-energy-distributed RE systems. Also, the
economics of biogas-to-energy systems are proven worldwide and are a
guaranteed positive cash flow (with a short payback time). Most importantly,
biogas-to-energy systems also provide pathways to making environmentally
friendly by-products and are globally required in the “circular economies.”
• Fuel cell development and operation based on wastewater systems has
been rapidly advancing, and current research results are highly utilized
worldwide for effective and immediate applications. Chapter 6 addresses
the key requirements of various fuel-cell technologies, including microbial
fuel cells (MFCs). The challenges—production, storage, and supply—posed
to biohydrogen infrastructure are elucidated, where there is rigorous
research requirement for fermentation technologies, control of fouling, and
the development of effective catalysts and electrode materials for fuel-cell
operation(s). The authors emphasize on the use of MFCs for the potential
NetZeroEnergy operation of wastewater infrastructure.
• Desalination operations are witnessing an unprecedented growth globally
because of increased freshwater demands. Specifically, RO processes or
technologies that dominate this field help make freshwater from brackish,
saline, and clean effluent (to name a few) water resources. In Chapter 7, the
author clearly identifies the need for research on membrane distillation and
adsorption desalination technologies for harvesting solar energy and waste
heat for (fresh) water production. The author also outlines the priority for
developing environmentally responsible and RE-integrated desalination
systems.
• Geothermal energy has been playing a significant role in HVAC systems
globally. In Chapter 8, the scientist authors elucidate on geothermal energy
and its commercial challenges, and the required improvements. It is vital
Preface xi

to note that current improvements focus on implementing enhanced


geothermal systems (EGSs) and binary power plants that will work based on
“lower geothermal resource temperatures,” and the scientists are outspoken
on “geothermal energy has the potential to solve many of the world’s energy
problems.”
• The application of RE systems with resilience is now critical and is a
case for the (immediate) future. This wind energy expert in Chapter 9
unequivocally states, “The capability to implement resilience in the built
infrastructure to tolerate energy loss as well as the outcome of natural and
man-made disasters including highly volatile aspects of climate change, is
key to establishing criteria for resilience in critical water infrastructures.”
The current advancements include hybrid wind energy systems that can well
support microgrids and can also provide real-time data on water security
and availability.
• Need for developing significant level(s) of energy—to operate high standard
water and wastewater treatment trains—owing to polluted resource water and
spent water and the resultant unwanted GHG emissions of fossil-fuel firing
for power generation is critical and needs a clear mandate. It is important
to note the critical role of RE systems in the context of the current climate
patterns, the loss of water due to the ever-increasing impervious areas in
worldwide development activities, and the associated droughts. In Chapter
10, solar energy is detailed for applications to both water and wastewater
systems and infrastructure.
• Current need for environmental quality monitoring is seen to be enormous,
and, so, Chapter 11 addresses the importance of water quality monitoring
through the use of RE systems. It is not an easy task to list out the various
requirements for such monitoring, and this is where the author’s effort in
writing such a chapter becomes commendable. The large wireless monitoring
networks [operating the Internet of Things (IoT) for data collection and
SCADA reporting] for water quality monitoring can make use of in situ
RE sources, such as solar power and MFCs. The emphasis on the use of
RE-based monitoring network systems also comes with a caution on security
requirements.
• Chapter 12 provides real-time project examples of water infrastructure
integrated with solar and/or wind energy systems. It is clearly shown that
RE systems can be used for both demand and supply side operations of water
infrastructure. Based on the proven applications detailed in the chapter, the
authors are insistent on RE integration with water infrastructure so as to
remove the dependence on fossil-fuel-generated power.
Acknowledgments

ARETC and EWRI greatly appreciate the following institutions and firms for
supporting the authors’ and reviewers’ efforts in the preparation of this book:
Central State University, Wilberforce, OH
Electric Power Research Institute, Palo Alto, CA
Gannon University, Erie, PA
Green Water-Infrastructure Academy, Washington, DC
Louisville Parks and Recreation, Louisville, KY
Manhattan College, Riverdale, NY
Mississippi State University, Mississippi State, MS
National Renewable Energy Laboratory (NREL), Golden, CO
Portland State University, Portland, OR
RC-WEE Solutions LLC, Dublin, OH
Rensselaer Polytechnic Institute, Troy, NY
Ritter Engineering, Elkton, MD
Rose-Hulman Institute of Technology, Terre Haute, IN
South Dakota School of Mines and Technology, Rapid City, SD
Star Sailor Energy, Inc., Cincinnati, OH
University of Delaware, Newark, DE
University of Illinois at Urbana-Champaign, Urbana, IL
ARETC honors the Sustainability Committee and Interdisciplinary Council
of EWRI for providing continuous support and appreciates the ARETC members
on successfully completing this book.

xiii
List of Authors and Reviewers

Ammi Amarnath
Ph.D. Candidate
Energy Efficiency and Demand Response Division
Electric Power Research Institute
Palo Alto, CA 94304

Audrey Angelos
Research Scientist
Thermal Sciences Group
National Renewable Energy Laboratory (NREL)
Golden, CO 80401

Govinda Chilkoor
Ph.D. Candidate
Civil and Environmental Engineering
South Dakota School of Mines and Technology
Rapid City, SD 57701

S. Rao Chitikela
Executive, Water, Energy, & EHSs
RC-WEE Solutions LLC
Adjunct Professor & Instructor
Central State University
(1890 Land-Grant Institution)
Dublin, OH 43016
Contact at: Rao.Chitikela@RCWEEsolutions.com

Venkataramana Gadhamshetty
Associate Professor
Civil and Environmental Engineering
South Dakota School of Mines and Technology
Rapid City, SD 57701
Contact at: Venkata.Gadhamshetty@sdsmt.edu

xv
xvi List of Authors and Reviewers

Veera Gnaneswar Gude


Associate Professor
Civil and Environmental Engineering
Mississippi State University
Mississippi State, MS 39762
Contact at: Gude@cee.msstate.edu

Venkata Gullapalli
Engineer II
Louisville Parks and Recreation
City of Louisville
Louisville, KY 40213
Contact at: Venkata.Gullapalli@louisvilleky.gov

Margaret A. Helms
Graduate Student
Environmental Science and Engineering
Gannon University
Erie, PA 16541
Contact at: Helms002@gannon.edu

Jawahar Kalimuthu
Research Assistant II
Civil and Environmental Engineering
South Dakota School of Mines and Technology
Rapid City, SD 57701

Ramanitharan Kandiah
Professor
Center for Water Resources Management
Central State University
Wilberforce, OH 45384
Contact at: RKandiah@centralstate.edu

Varun K. Kasaraneni
Assistant Professor
Environmental Science and Engineering
Gannon University
Erie, PA 16541
Contact at: Kasarane001@gannon.edu
List of Authors and Reviewers xvii

James E. Kilduff
Ph.D. Candidate
Civil and Environmental Engineering
Rensselaer Polytechnic Institute
Troy, NY 12180

Alexander Krokus (Deceased)


Fellow, Institute for Sustainable Solutions
Research Analyst, School of Government
Portland State University
Portland, OR 97207

Juneseok Lee
Associate Professor
Civil and Environmental Engineering
Manhattan College
Riverdale, NY 10471
Contact at: Juneseok.Lee@manhattan.edu

Pamela A. Menges
President
Star Sailor Energy, Inc.
Cincinnati, OH 45224
Contact at: pmenges@starsailorenergy.com

William F. Ritter
Professor Emeritus
University of Delaware
Newark, DE 19702
and
Ritter Engineering
Elkton, MD 21921
Contact at: WRitter@udel.edu

Namita Shrestha
Assistant Professor
Civil and Environmental Engineering
Rose-Hulman Institute of Technology
Terre Haute, IN 47803
Contact at: Shrestha@rose-hulman.edu
xviii List of Authors and Reviewers

Ashlynn S. Stillwell
Associate Professor
Civil and Environmental Engineering
University of Illinois at Urbana-Champaign
Urbana, IL 61801
Contact at: Ashlynn@illinois.edu

Bhuvan Vemuri
Ph.D. Candidate
Civil and Environmental Engineering
South Dakota School of Mines and Technology
Rapid City, SD 57701
Contact at: Bhuvan.Vemuri@mines.sdsmt.edu

Tamim Younos
Founder and President
Green Water-Infrastructure Academy
Washington, DC 20001
Contact at: Tamim.Younos@gwiacademy.org

Guangdong Zhu
Research Scientist
Thermal Sciences Group
National Renewable Energy Laboratory (NREL)
Golden, CO 80401
Contact at: Guangdong.Zhu@nrel.gov
CHAPTER 1
US Renewable Energy
Policy—Analysis and
Recommendations
Alexander Krokus (Deceased)

INTRODUCTION

The first industrial use of hydropower for energy generation that transpired in the
United States occurred in Grand Rapids, Michigan, during 1880, when 16 brush-
arc lamps were powered by a water turbine (Pandey and Karki 2017). By 1920, the
United States had implemented its first federal energy policy regarding renewable
energy production, when it enacted the Federal Water Power Act (FWPA) of 1920
(16 U.S.C. §791a). FWPA promoted the establishment of renewable energy policy
nationally, by creating hydroelectric power plants for energy generation, and
promulgated the formation of the Federal Power Commission, which later, in
1977, became the Federal Energy Regulatory Commission (FERC). This act was
later, in 1935, renamed Federal Power Act (FPA) and increased FERC’s jurisdiction
to encompass all interstate electricity transmission.
On April 18, 1977, ex-US President Jimmy Carter delivered a speech to the
nation that would still be a relevant concern in modern time. “We must not
be selfish or timid if we hope to have a decent world for our children and our
grandchildren … By acting now we can control our future instead of letting the
future control us … (referring to) the oil and natural gas that we rely on for 75%
of our energy … [During 2017, the United States utilized fossil fuels to facilitate
80.9% of all energy consumption (USEIA 2018a).] During the 1950s, people used
twice as much oil as during the 1940s. During the 1960s, we used twice as much
as during the 1950s. And in each of those decades, more oil was consumed than in
all of man’s previous history combined” (Carter 1977). This speech was President
Carter’s precursor to introducing his National Energy Plan to the US Congress,
which led to the establishment of the National Energy Act (NEA) of 1978, which
contained major statutes devoted to harnessing renewable energy resources. NEA
of 1978 promulgated the Energy Tax Act of 1978, Pub. L. No. 95-618, 92 Stat.

1
2 Renewable Energy Technologies and Water Infrastructure

3174, which amended §1954 of the Internal Revenue Code to grant an income
tax credit for individuals utilizing solar, wind, or geothermal energy generation
for their personal residence (HR 5263 1977-78). This act also created residential
energy credit for the tax imposed for energy conservation [26 U.S.C. §44C(a)(1)]
and renewable source expenditures [26 U.S.C. §44C(a)(2)]. §44C(5)(A) states that
energy “installed in connection with a dwelling, transmits or uses (i) solar energy,
energy derived from the geothermal deposits [as defined in Section 613(e) (3)], or
any other form of renewable energy which the Secretary specifies by regulations,
for the purpose of heating or cooling such dwelling or providing hot water for use
within such dwelling, or (ii) wind energy for nonbusiness residential purposes.

National Energy Act of 1978 (H.R. 8444 1977-78)


NEA of 1978 also created the Energy Security Act of 1980, 42 U.S.C. §8701 et seq.
Title I (Synthetic Fuels Corporation Act), §§100-195, which created the Synthetic
Fuels Corporation (Section, Pub. L. 96-294, Title I, § 100, June 30, 1980, 94 Stat.
616, was omitted from the code because of the termination of the United States
Synthetic Fuels Corporation and repealed in 1985.) to collaborate with industry
to develop a market for synthetic liquid fuels. Research and development was
transferred from the US Department of Energy into a public–private partnership
to accelerate achieving successful strategies. Title II (Biomass Energy and Alcohol
Fuels Act) of NEA of 1978, §§201-274, granted loan guarantees for small-scale
biomass energy projects and established the federal Office of Alcohol Fuels and the
Office of Energy from Municipal Waste. Title IV (Renewable Energy Initiatives),
§§401-409, established economic incentives for the utilization of renewable
energy resources. Title V (Solar Energy and Energy Conservation) of NEA of
1978, §§501-597 (Energy Policy Act of 1992, 16 U.S.C. Chapter 46 §2601 et seq.
repealed Title V of the Energy Security Act of 1980.), encouraged the expansion
of solar energy and established the Solar Energy and Energy Conservation Bank,
and 42 U.S.C. §8201, §§241-248, established solar energy improvement loans. Title
VI (Geothermal Energy Act) of NEA of 1978, §§601-644, authorized federal loans
from the Geothermal Resources Development Fund for research associated with
exploration and the economic viability of a geothermal reservoir. It also promoted
the usage of geothermal energy as a feasible method to implement in new federal
buildings. The National Energy Conservation Policy Act (NECPA) of 1978,
Pub. L. No. 95-619, 92 Stat. 3206, replaced the minimum energy performance
standards set forth in the Energy Policy and Conservation Act (EPCA) of 1975,
Pub. L. No. 94-163, 89 Stat. 871, and transformed energy standards from voluntary
to mandatory. The Public Utility Regulatory Policies Act (PURPA) of 1978,
Pub. L. No. 95-617, 92 Stat. 3117, advocated for the usage of renewable energy
implementation and encouraged the creation of cogeneration plants. Regulatory
authority was solely delegated to the states.

Energy Policy Act of 2005


The next monumental piece of US federal energy policy was enacted over two
decades later, when the Energy Policy Act (EPAct) of 2005, 42 U.S.C. §15801
US Renewable Energy Policy—Analysis and Recommendations 3

et seq., became public law. EPAct of 2005 enabled tax incentives for individuals
increasing energy efficiency in their homes and also for consumers to buy or lease
hybrid vehicles. EPAct of 2005 also raised the mandatory percentage of renewable
fuel contained in gasoline. Succeeding the repeal of the Public Utilities Holding
Act of 1935 [15 U.S.C. §§79-79(z)(6)] by the enactment of EPAct of 2005, FERC
permitted immense capital investment into the US oil and gas sector, allowing for
emerging new unconventional oil and gas formations to be exploited, by utilizing
horizontal drilling techniques, enabling the United States, to become the global
leader in oil and gas production in 2018 (USEIA 2018b, British Petroleum 2019).

Energy Independence and Security Act of 2007


The Energy Independence and Security Act of 2007, 42 U.S.C. Ch. 152 §17001 et
seq., increased the Corporate Average Fuel Economy (CAFE) standards to 56 kmph
(35 mpg) for passenger automobiles by 2020. These Renewable Fuel Standards
(RFS) had resulted in amplified biofuel production for 136 billion liters (36 billion
gal.) by 2022, with 79.5 billion liters (21 billion gal.) derived from noncornstarch
products [40 CFR Part 80, FR, Vol. 81, No. 238 (December 12, 2016)].

Food, Conservation, and Energy Act of 2008


The enactment of the Food, Conservation, and Energy Act (Farm Bill) of 2008,
Pub. L. No. 110-234, 122 Stat. 923, provided access to federal loans for biorefineries
and monetary compensation to facilitate the expansion of innovative biofuels
and also expanded the existing Rural Energy for America Program. The Biomass
Crop Assistance Program (BCAP, §9001), initiated by the 2008 Farm Bill, was
reauthorized with modifications by the 2014 Farm Bill (Agricultural Act of 2014,
Pub L. No. 113-79). The revised BCAP provides financial assistance to farmers
and forest landowners who are growing, maintaining, or harvesting biomass that
can be utilized for energy. This assistance includes payments for cultivating new
biomass crops. (The BCAP has the ability to fund up to 50% of costs affiliated
with the establishment of a new perennial energy crop or biomass crop.) Annual
maintenance payments can be used to cultivate biomass crops until they mature,
up to 5 years for an herbaceous crop, or up to 15 years for a woody crop. The BCAP
can also provide monetary support for the cost of sustainable harvesting and the
transportation of agricultural or forest residues to an energy conversion facility.
[Animal waste, bagasse, food and yard waste, and algae are ineligible for retrieval
payments. §9003 “Eliminates grants to assist in paying the costs of development
and construction of demonstration-scale biorefineries to demonstrate the
commercial viability of one or more processes for converting renewable biomass
to advanced biofuels.” §9011 repeals the forest biomass for energy program (§9012
of the Farm Security and Rural Investment Act of 2002, 7 U.S.C. §8112).]

Clean Power Plan of 2015


The Clean Power Plan (CPP) of 2015 (40 CFR Part 60, FR, Vol. 80, No. 64661-65120,
October 23, 2015) was the first comprehensive national strategy to mitigate carbon
4 Renewable Energy Technologies and Water Infrastructure

emissions from existing fossil fuel-fired electric generating units, providing states
flexibility in their methods of implementation. States were permitted the option
to choose either achieving rate-based or mass-based goals, which were calculated
by applying performance rates for fossil fuel power plants versus their entire
energy amalgam. Individual state plans could include various methods to achieve
their goals, including investments for energy conservation, or by implementing
additional wind or solar installations.
On June 19, 2019, EPA repealed the CPP and replaced it with the Affordable
Clean Energy (ACE) rule, amending §111(d) of the Clean Air Act. By 2030, the
ACE rule is expected to reduce CO2 emissions from electric generating units to
35% below 2005 levels (EPA 2019a). During the first session of the 116th Congress,
Senator Tom Udall (NM-3) proposed the Renewable Electricity Standard (RES) Act
of 2019, which attempts to amend Title VI of the PURPA of 1978 to accelerate our
nationwide transition to renewable energy generation. By 2050, §2(2) of the RES
Act of 2019 requires every state to transition to 100% carbon-free electricity but
encourages states to devise their own strategies to obtain these (ambitious) goals.
The RES Act of 2019 does provide an achievable ramp-up approach, requiring
annual percentage increases of renewable energy ranging from 1.5% to 2.5%. By
2035, these reductions, if met, will assist the nation in attaining 50% of renewable
energy generation. The issuance of federal renewable energy credits would be
awarded to states which comply with the national standards set forth in this bill.

Energy Efficiency Resources Standards


Twenty-seven states are presently using Energy Efficiency Resources Standards’
(EERS) policies, which mandate electricity reduction methods. Eighteen states
have EERS policies for natural gas usage. In 1999, Texas was the first state to
implement an EERS. Only 29 states in the United States have adopted Renewable
Portfolio Standards (RPS) to establish renewable energy goals (NCSL 2019).
Renewable energy resources allowable for RPS compliance include wind, solar,
biomass, geothermal, and hydroelectric facilities. Certain states allow for utilizing
landfill gas and tidal energy and for implementing energy efficiency standards. In
2018, Massachusetts, California, and Rhode Island were the top three states leading
in energy efficiency because of the employment of various environmental policy
strategies (ACEEE 2018). Leading by Example (LBE) initiatives, which include
building energy efficiency, nonbuilding energy efficiency, greenhouse gas (GHG)
reduction, green buildings, renewable energy, and sustainable transportation,
have proven to be successful in the State of Massachusetts. The LBE method,
initiated by the 71st Governor of Massachusetts Deval Patrick, proclaims that
“state government has an obligation to lead by example and demonstrate that
large entities such as state colleges and universities, prisons, hospitals and others
can make significant progress in reducing their environmental impacts, thereby
providing a model for businesses and private citizens” (Mass. Exec. Order No.
484, April 18, 2007). In 2012, the LBE partners in Massachusetts achieved their
GHG emission reduction goal of 25% below 2002 levels and, in 2050, plan to
reduce state government GHG emissions by 80% (MA-DER 2019a).
US Renewable Energy Policy—Analysis and Recommendations 5

Renewable Energy Projects


Renewable energy projects immensely vary nationally and are dictated by state
and local policy. The actual renewable potential for a specific form of energy
generation often is not accounted for when formulating legislation. The States
of Massachusetts, California, New Jersey, Arizona, New York, Nevada, Texas,
and Pennsylvania account for 99.5% of the nations installed solar photovoltaic
(PV) capacity; yet, only California, Nevada, and Texas are ranked in the top
10 states possessing the greatest potential for harnessing solar PV energy (EPA
2019b). Despite achieving more than 1 million solar PV installations nationally
in 2016, and 9.8 × 1013 J/h (27.2 GW) of installed capacity (Unger 2016), solar
PV contributed to only 1.4% of the total energy generated (DOE 2017a). In 2017,
national leaders in solar PV installed capacity, such as Massachusetts, generated
7.3 × 1015 J (2,030,879 MWh) of solar PV and 1.2 × 1016 J (3,353,712 MWh) of
wind power. They were able to generate over 4.7 × 1015 J (1.3 million MWh)
more energy utilizing wind, despite possessing 21 times less installed capacity
compared with solar PV (MA-DER 2019b).

Federal Investment Tax Credit


The rise of the solar PV industry has been driven by a federal investment tax
credit (ITC), often referred to as the solar tax credit, which permits a federal
tax deduction of 30% for costs associated with residential and commercial
installations. The solar ITC was a by-product of §1337(a)(A)(i) of EPAct of 2005
and has allowed the national solar PV industry to expand by 10,000% (USEIA
2019). The residential ITC (26 U.S.C. §25D) and the commercial ITC (26 U.S.C.
§48) diminished from 30% in 2019 to 26% in 2020, 22% in 2021, and to 10% after
January 1, 2021. Besides supporting solar investments, the allowable residential
ITCs also apply to small wind energy projects [§(a)(4)] and geothermal heat pump
expenditures [§(a)(5)]. The ITC for renewable energy generation was extended by
the Consolidated Appropriations Act of 2016, Pub. L. 114-113, 129 Stat. 2242 (H.R.
2029, 2015–2016), and currently a new 5 year extension has been introduced by
both chambers of the US Congress. The House bill (Renewable Energy Extension
Act of 2019) is supported by three republican Congressmen, and the Senate bill’s
proponents are 15 democratic Senators. Besides amending §48 of the Internal
Revenue Code of 1986 for solar PV, this proposed policy also extends the ITC for
additional energy technologies about to sunset, which includes fiber-optic solar,
qualified fuel cells, and small wind energy [§2(b)(2)].

Renewable Production Tax Credit


The Consolidated Appropriations Act of 2016 also enabled an extension for the
renewable production tax credit (PTC), which is a per kilowatt hour tax credit
indexed for inflation on energy generated by qualified renewable resources,
differing from ITC, which grants a tax credit based on the monetary assets
invested, not the amount of electricity actually produced (26 U.S.C. §45). With the
exception of wind energy, every form of renewable energy generation PTC expired
6 Renewable Energy Technologies and Water Infrastructure

on January 1, 2018. Yet, projects initiated before the expiration date will have
access to the PTC for their initial 10 years of energy production. This applies to all
qualified facilities, which includes wind, closed-looped biomass, and geothermal
receiving 2.4¢ per kWh and open-loop biomass, small irrigation power, municipal
solid waste, hydropower, and hydrokinetic receiving 1.2¢ per kWh.
The renewable PTC has been extended 11 times since it first became law
subsequent to the enactment of the EPAct of 1992, Pub. L. No. 102-486. The
most recent extension transpired when the Bipartisan Budget Act of 2018, Pub.
L. No. 115-123, provided a retroactive PTC for all nonwind renewable energy
installations conducted before the end of 2017 (HR 1892, 2017-2018). The federal
PTC has played a vital role in promoting wind energy implementation nationally
and has increased wind energy installations substantially, 720 GJ (200 MW) per
state annually (Shrimali et al. 2015). The enactment of a 1 year extension for the
renewable PTC is presently being debated by the US House of Representatives
(H.R. 3301, 2019 to 2020).

Modified Accelerated Cost Recovery System


The Modified Accelerated Cost Recovery System (MACRS) is a complex set of tax
policies that classifies almost every form of renewable energy generation’s assets
as a 5 year property, which allows taxpayers to recover their entire depreciation
allowance in a 5 year period of time [26 U.S.C. §168(e)(3)(B)(vi)(I)]. The Emergency
Economic Stabilization Act of 2008, Pub. L. No. 110-343, 122 Stat. 3765 permitted
bonus depreciation, which provides a 50% first-year bonus depreciation for
qualifying renewable energy investments. The Tax Relief, Unemployment
Insurance Reauthorization, and Job Creation Act of 2010, Pub. L. No. 111-312, 124
Stat. 3296 increased the first-year bonus depreciation to 100%. The Tax Cuts and
Jobs Act of 2017, Pub. L. No. 115-97, 131 Stat. 2054 expanded bonus depreciation
tax relief to cover both new and used equipment to be expensed at 100%, during
the first year of purchase, for all qualified property secured and operational after
September 27, 2017, up until January 1, 2023 (HR 1, 2017). Despite this favorable
federal tax policy, the wind industry has seldom taken advantage of this generous
bonus depreciation option (DOE 2018).

RENEWABLE ENERGY RESOURCES—REVIEW/


RECOMMENDATIONS

Hydropower
The 1986 amendments to FPA of 1920 (Electric Consumers Protection Act of 1986,
Pub. L. 99-495, 100 Stat. 1243) devised protection strategies for specific aquatic
species [16 USC §803(j)1)] [“(j) Fish and wildlife protection, mitigation
and enhancement; consideration of recommendations; findings (1)
That in order to adequately and equitably protect, mitigate damages to, and enhance,
US Renewable Energy Policy—Analysis and Recommendations 7

fish and wildlife (including related spawning grounds and habitat) affected by
the development, operation, and management of the project, each license issued
under this subchapter shall include conditions for such protection, mitigation,
and enhancement. Subject to paragraph (2), such conditions shall be based on
recommendations received pursuant to the Fish and Wildlife Coordination
Act (16 U.S.C. 661 et seq.) from the National Marine Fisheries Service, the
United States Fish and Wildlife Service, and State fish and wildlife agencies.”]
adversely affected by hydroelectric power plants. Despite hydropower’s favorable
sustainability attributes and low carbon footprint, and the fact that it provides the
vast majority of renewable energy generation nationally (USEPA 2019a), this form
of energy may necessitate stricter federal policy measures to mitigate ecological
abnormalities. Almost three decades subsequent to the passage of the Electric
Consumers Protection Act of 1986, scientists compared the impairments with 239
endangered freshwater fish species in the United States, which contained major
threat categories, as follows: “dams/impoundments, invasive/introduced species,
altered hydrologic flow/channelization, overharvesting/overfishing, pollution/
water quality, sedimentation/turbidity/siltation, excess water consumption/
withdrawal, and hybridization”; in addition, it was revealed that a majority of
aquatic species had multiple threats contributing to their decline, and surprisingly,
dams/impoundments are the primary biological inhibitor (McDonald et al. 2012).
Future policies can diminish this adverse effect by mandating that dissolved-
oxygen (DO) monitoring stations are located both upstream and downstream
from the facility while performing near-continuous monitoring. A majority of
freshwater aquatic species necessitate DO levels greater than 5 mg/L for optimum
growth and to avoid chronic effects on survival (Niklitschek and Secor 2009,
Stoklosa et al. 2018, KY-NREPC 2019). An alternate method to conventional
damming is the utilization of instream turbine technology. This strategy is less
environmentally destructive and can reduce unintentional damage to marine life
residing in close proximity to hydropower operations (Wang et al. 2012a, b).

Biofuels
The multiple federal policy definitions for the term biomass have complicated
decisions pertaining to land usage and which feedstock to utilize, especially
when pertaining to RFSs and tax incentives. The US Congress has redefined the
meaning of biomass in 14 separate pieces of federal legislation during the last 15
years (CRS 2019). As a result of the complexity surrounding the legal meaning of
biomass, research and development projects, including the production of biomass
for energy conversion, can encounter unnecessary obstacles in their effort to fully
exploit this emerging method of energy generation. On May 22, 2019, Senator Ron
Wyden (OR-3) introduced a bill (S. 1614) in an attempt to refine the meaning of
renewable biomass under §211(o)(1)(I) of the Clean Air Act [42 USC §7535(o)(1)
(I)], which presently prohibits the usage of biomass derived from federal lands.
This legislative concept permits obtaining biomass from designated federal lands,
necessitating ecological restoration.
8 Renewable Energy Technologies and Water Infrastructure

Certain ethanol–gasoline mixtures produce higher evaporative emissions


from fuel tanks and distribution equipment as compared to gasoline, which
can release toxic substances leading to the formation of harmful smog (USEIA
2019b). Because the lifecycle emissions of ethanol is reliant on the materials used
and the methods employed for processing, future policy measures can lessen the
environmental impact by promoting the conversion of waste into gases, or by
other alternates, such as utilizing algae and additional microorganisms to generate
fuel from water or solar radiation. Biofuels derived from waste products including
municipal and crop waste can eliminate this complication and may alleviate the
formation of fugitive GHG emissions resulting from land-use changes.

Solar Photovoltaic
Solar PV energy has successfully been utilized globally. There is also debate
associated with the amount of GHG emissions yielded during the solar PV
manufacturing process (WNA 2011, Stamford and Azapagic 2014). Factories
located in China currently use nitrogen trifluoride (NF3) and sulfur hexafluoride
(SF6) during the etching process in PV panel manufacturing. Chinese solar
factories primarily manufacture PV panels that use crystalline silicon (c-Si)
cells (Fang et al. 2013). A majority of the remaining panels produced are
either cadmium telluride (CdTe) or copper–indium-gallium–selenide (CISG).
The residual supply of rare earth minerals required to these various forms
of PV manufacturing has been diminishing rapidly. China occupies 97% of
these minerals and has enacted production and export quotas. This immense
uncertainty relating to the future supply of these rare earth minerals poses
a substantial threat to the solar PV industry (Than 2018). Additional federal
funding must be allocated for developing alternate methods for harnessing the
Sun’s energy and to also create an etching gas that is environmentally friendly.
If we could capture 100% of the Sun’s energy reaching the Earth in only the State
of Texas, then it would generate over three-hundred times the aggregate power
of every power plant globally (UTIA 2019).
Future federal policy must amend the Resource Conservation and Recovery
Act, 42 USC §6901 et seq., and include a section specifically pertaining to the
end-of-life disposal for solar PV systems, and establish a reclassification method
for nanomaterials by toxicity, rather than by sheer weight in the EPA TRI (Toxics
Release Inventory) database. Nanomaterials often exhibit transmuted properties
as compared to larger sized particles of similar material and have the potential to
be extremely toxic in diminutive dosages; yet, this is not taken into consideration
when establishing regulations or constructing material safety data sheets (MSDSs).

Wind Energy
According to the United States Department of Energy (DOE), despite a
significant decrease in the costs associated with wind installations, the initial
capital expenditure for wind projects “might not be the most profitable use of
the land,” and wind turbines have the potential to generate “noise and aesthetic
US Renewable Energy Policy—Analysis and Recommendations 9

pollution” and also avian mortalities; yet, a majority of these complications “have
been resolved or greatly reduced” (DOE 2019b). Erickson et al. (2014) performed
a meta-analysis based on 116 prior studies focusing on more than 70 wind energy
facilities in the United States and Canada and estimated that 134,000 to 230,000
small-passerine birds (less than 0.1%), the most abundant bird category in the
United States and Canada, collide with wind turbines annually.
The United States installed 1.9 × 1014 J/h (52,500 MW) of wind energy
capabilities in 2017, raising the national total capacity to 1.9 × 1015 J/h
(539,000 MW). The State of Texas led the nation with 8.1 × 1013 (22,599 MW) of
installed capacity, and the States of Iowa, Kansas, Oklahoma, and South Dakota
used wind energy to supply 30% to 37% of all in-state electricity generation.
Denmark was able to supply 48% of all energy generation nationally by wind
in 2017, and Ireland and Portugal supplied approximately 30% (DOE 2018).
Offshore wind potential in the United States is substantial and can be utilized
if favorable federal and state policies are formed to aid the development of these
facilities. The State of Rhode Island became the first state in the nation to develop
an offshore wind facility lock Island Wind Farm)in the United States, during
2016, and additional projects in neighboring states are being contemplated by
policymakers (USEIA 2019c). This includes a $4.5 million investment by Danish
company Orsted, Fredericia, who plans to install an additional offshore wind farm
in Rhode Island consisting of up to 50 new wind turbines that will have the ability
to power approximately 270,000 residential homes (McDermott 2019). According
to DOE (2015), the “next generation of wind turbines could make reliable, cost-
effective wind power a reality in all 50 states.” This reality can be achieved by
implementing new advanced wind turbines, which utilize taller towers and
longer blades that rely on consistent wind patterns found at higher elevations.
The evolution of wind power generation has made substantial advances in terms
of technological development and is now labeled globally as the “cheapest and
most reliable energy technologies in the market” (GWEC 2015). During 2016, the
wind power sector represented the third largest share of electric power generation
employment nationally. At the beginning of 2017, the US wind sector employed
101,738 individuals, rising 32% in just 1 year (DOE 2017b) and, by 2050, could
facilitate an additional 600,000 jobs (DOE 2019b).

Future of United States Renewable Energy Policy


Future US energy policy must be focused on restricting the supply side of fossil fuel
generation, the production aspects, to reduce the amount extracted rather than
implementing demand-side regulations (such as a cap-and-trade on emissions),
which allows for the continuous expansion of fossil fuel infrastructure. USEIA has
predicted a substantial increase (of 99%) in unconventional oil and gas recovery
to be experienced by 2040 (USEIA 2016). As long as generous federal tax subsidies
endure for the oil and gas industry for the exploration, extraction, and also the
transport of fossil fuels, efforts to shift the United States to more reliance on
renewable energy generation will be compromised.
10 Renewable Energy Technologies and Water Infrastructure

The United Nations Intergovernmental Panel on Climate Change (UN-IPCC)


has alerted the world of the possibility of irreversible negative environmental
impacts occurring in the imminent future if fossil fuel emissions are not
significantly reduced. To avoid a 2°C increase in global warming relative to
preindustrial times, GHGs must become stable at 450 ppm or less (UN-IPCC
2014). The Fifth Assessment Report concluded that “many aspects of climate
change and associated impacts will continue for centuries, even if anthropogenic
emissions of greenhouse gases are stopped. The risks of abrupt or irreversible
changes increase as the magnitude of the warming increases” (UN-IPCC 2014).
At the present time, the Earth’s atmosphere is 408 ppm CO2 and has
been averaging a 2.8 ppm increase for the past decade (JAXA 2019), leaving
us approximately 15 years to find a solution to mitigate unalterable ecological
consequences. A supply-side cap-and-trade system restricting extraction and
production, rather than emissions, would diminish pollution and allow the
renewable energy industry a fairer chance to politically compete with the fossil
fuel industry (Collier and Venables 2014, Lazarus et al. 2015). A global shift from
fossil fuel policies focusing on regional demand to policies dedicated to lowering
global supply, by enacting global quantity constraints, export taxes, and by
taxing interest income earned with a minimum source tax, would diminish GHG
emissions and increase the favorability for renewable energy implementation.
Reducing the amount of federal subsidies for fossil fuels would also encourage
the formation of renewable energy installations. For example, the expensing of
exploration and development costs [26 USC §263(c)] [“(c) Intangible drilling
and development costs in the case of oil and gas wells and geothermal
wells: Notwithstanding subsection (a), and except as provided in subsection (i),
regulations shall be prescribed by the Secretary under this subtitle corresponding to
the regulations which granted the option to deduct as expenses intangible drilling
and development costs in the case of oil and gas wells and which were recognized
and approved by the Congress in House Concurrent Resolution 50, Seventy-ninth
Congress. Such regulations shall also grant the option to deduct as expenses intangible
drilling and development costs in the case of wells drilled for any geothermal deposit
[as defined in section 613(e)(2)] to the same extent and in the same manner as such
expenses are deductible in the case of oil and gas wells. This subsection shall not
apply with respect to any costs to which any deduction is allowed under section
59(e) or 291.”] provision, an over $1 billion tax expenditure annually, permits oil and
natural gas producers to expense exploration and development expenditures (which
include certain intangible drilling and development costs) rather than capitalizing
and depreciating them over time. The aggregate of federal energy subsidies in the
United States for natural gas and oil were over $51 billion in 2016 (up from $35
billion in 2010), 500% higher than the $10 billion for biomass, hydroelectric, wind,
solar, and geothermal combined (USEIA 2018c). Even the US coal industry, despite
a $7 billion decrease in federal assistance since 2010, acquired almost $5 billion more
in tax subsidies than all renewables combined in 2016 (Table 1-1).
The total amount of renewable energy produced in the United States almost
doubled from 2008 to 2018, rising to 2.7 × 1018 J (742 million MWh), facilitating
18% of the total energy generated nationally (USEIA 2019d). This upsurge in
US Renewable Energy Policy—Analysis and Recommendations 11

Table 1-1. Total Energy Subsidies for FY 2010, FY 2013, and FY 2016.

Indicators FY 2010 FY 2013 FY 2016


Total energy subsides and 37,992 29,335 14,983
support (million 2016 dollars)
US energy consumption 96,850 98,655 96,788
US energy production 73,695 81,151 84,833
 US natural gas (dry and liquids) 24,105 28,220 32,652
 US crude oil 11,512 15,370 18,797
 US coal 21,657 20,223 14,807
 US nuclear 8,318 8,099 8,352
 US biomass 4,358 4,680 4,963
 US hydroelectric 2,588 2,582 2,482
 US wind 863 1,557 2,038
 US solar 88 205 533
 US geothermal 207 215 209
Source: USEIA (2018c) with permission (17 U.S.C. §105).

renewable energy production in the United States will almost certainly increase
even more drastically over the next decade. During the last few years, the States of
Hawaii, California, Colorado, Maine, Nevada, New Mexico, New Jersey, New York,
Washington, Connecticut, Rhode Island, and Virginia, including Washington, DC,
and Puerto Rico, have either passed legislation or enacted executive orders committing
to achieving 100% renewable or clean energy generation by 2050 (Podesta et al. 2019,
Fields 2020). Washington, DC, and Rhode Island have devised the most ambitious
renewable energy policy. Washington, DC, is striving to attain 100% renewable energy
generation by 2032 (DC DEE 2019), and Rhode Island is determined to reach this goal
by 2030 (Rhode Island Exec. Order No. 20-01, January 17, 2020).
Virginia became the first Southern state in the nation to join this newly
emerging carbon-free policy movement, mandating the development of a
9.0 × 1012 J/h (2,500 MW) offshore wind facility that will be completed by 2026
and establishing an additional 2.0 × 1013 J/h (5,500 MW) of onshore wind and
solar energy by 2028 (Virginia Exec. Order No. 43, September 16, 2019). The US
transition to 100% renewable energy utilization will hopefully become a reality
in the imminent future.

SUMMARY

The prolongation of the residential and commercial ITC, PTC, and MACRS is an
essential requirement for expanding our nation’s renewable energy capabilities.
Fortunately, there is bipartisan support for these policy measures in the 116th
US Congress. Despite an over 100% increase in federal subsidies for the US solar
(2013 to 2016) and wind (2010 to 2016) industries (USEIA 2018c), we still must
12 Renewable Energy Technologies and Water Infrastructure

significantly decrease the amount of subsidies for the development of new fossil
fuel infrastructure and allocate additional federal funds to aid the renewable
energy technology sector.
Topics covered in the later chapters discuss the advantageous aspects of
establishing micro-hydropower installations utilized for both remote and urban
areas, the benefits of microbial fuel cells (MFCs) that transform wastewater into
electricity, incorporating renewables in desalination technologies, and also the
opportunity for providing energy generation from solar radiation and solar
disinfection that provides wastewater treatment. This publication also includes
the innovative process of anaerobic digestion of wastewater to produce biogas
energy and the formation of decentralized green water-infrastructure systems,
among other methods of sustainable renewable energy generation.
Once additional states pass new laws to eliminate our reliance on fossil fuels,
renewables will thrive, and the probability of encountering extreme weather
events will diminish, safeguarding future generations of humankind from the
unfavorable consequences of anthropogenic climate change.
As stated previously in this chapter, “We must not be selfish or timid if we hope
to have a decent world for our children and our grandchildren … By acting now
we can control our future instead of letting the future control us…” (Carter 1977).

References
40 CFR Part 60, 80 FR, Vol. 80, No. 64661-65120. “Carbon pollution emissions guidelines
for existing stationary sources: Electric utility generating units.” Federal Register.
Accessed October 23, 2015. https://www.govinfo.gov/content/pkg/FR-2015-10-23/
pdf/2015-22842.pdf.
40 CFR Part 80, FR, Vol. 81, No. 238. “Renewable fuel standard program: Standards for
2017 and biomass-based diesel volume for 2018.” Federal Register. Accessed December
12, 2016. https://www.govinfo.gov/content/pkg/FR-2016-12-12/pdf/2016-28879.pdf
ACEEE (American Council for an Energy-Efficient Economy). 2018. “State and local policy
database.” Accessed April 4, 2019. https://database.aceee.org/state/massachusetts.
British Petroleum. 2019. “BP statistical review of world energy.” 67th ed. Accessed April 27,
2019. https://www.bp.com/en/global/corporate/energy-economics/statistical-review-of-
world-energy.html.
Carter, J. 1977. “Address to the nation on energy.” Univ. of Virginia, Miller Center,
Presidential Speeches. Accessed March 2, 2019. https://millercenter.org/the-presidency/
presidential-speeches/april-18-1977-address-nation-energy.
Collier, P., and A. J. Venables. 2014. “Closing coal: Economic and moral incentives.” Oxford
Rev. Econ. Policy 30 (3): 492–512.
CRS (Congressional Research Service). 2019. “Biomass: Comparison of definitions in
legislation.” Accessed July 3, 2019. https://fas.org/sgp/crs/misc/R40529.pdf.
DC DEE (DC Department of Energy and Environment). 2019. “Mayor Bowser signs
historic clean energy bill, calling for 100% renewable electricity by 2032.” Accessed
March 7, 2019. https://doee.​dc.gov/release/mayor-bowser-signs-his​toric-clean-ener​
gy-bill-calling-100-renewable-electricity-2032.
DOE (United States Department of Energy). 2015. Unlocking our nation’s wind potential.
Washington, DC: Office of Energy Efficiency & Renewable Energy.
US Renewable Energy Policy—Analysis and Recommendations 13

DOE. 2017a. “Q4 2016/Q1 2017 presentation—Solar industry update.” Solar Energy
Technologies Office. Accessed April 4, 2019. https://www.energy.gov/eere/solar/
q4-2016q1-2017-presentation-solar-industry-update.
DOE. 2017b. “U.S. energy and employment report.” Accessed April 8, 2019. https://www.
energy.gov/downloads/2017-us-energy-and-employment-report.
DOE. 2018. 2017 wind technologies market report. Washington, DC: Office of Energy
Efficiency & Renewable Energy. Accessed March 5, 2019. https://www.energy.gov/sites/
prod/files/2018/08/f54/2017_wind_technologies_market_report_8.15.18.v2.pdf.
DOE. 2019a. Ethanol fuel basics. Washington, DC: Office of Energy Efficiency & Renewable
Energy. Accessed March 2, 2019. https://afdc.energy.gov/fuels/ethanol_fuel_basics.
html.
DOE. 2019b. Advantages and challenges of wind energy. Washington, DC: Office of Energy
Efficiency & Renewable Energy. Accessed April 8, 2019. https://www.energy.gov/eere/
wind/advantages-and-challenges-wind-energy.
EPA (Environmental Protection Agency). 2019a. “EPA finalizes affordable clean
energy rule, ensuring reliable, diversified energy resources while protecting our
environment.” News Release from Headquarters, Air and Radiation (OAR). Accessed
June 21, 2019. https://www.epa.​gov/​newsreleases/epa-finalizes-affor​d able-clean-ene​
rgy-rule-ensuring-reliable-diversified-energy.
EPA. 2019b. “Energy resources for state and local governments: State renewable energy
resources.” Environmental Topics. Accessed April 4, 2019. https://www.epa.gov/
statelocalenergy/state-renewable-energy-resources#State Policies to Support Renewable
Energy.
Erickson, W. P., M. M. Wolfe, K. J. Bay, D. H. Johnson, and J. L. Gehring. 2014. “A
comprehensive analysis of small-passerine fatalities from collision with turbines at
wind energy facilities.” PLoS One 9 (9): e107491.
Fang, X., X. Hu, G. Janssens-Maenhout, J. Wu, J. Han, S. Su, J. Zhang, and J. Hu. 2013.
“Sulfur hexafluoride (SF6) emission estimates for China: An inventory for 1990–2010
and a projection to 2020.” Environ. Sci. Technol. 47 (8): 3848–3855.
Fields, S. 2020. “100 percent renewable targets.” Energy Sage. Accessed June 15, 2020.
https://news.energysage.com/states-with-100-renewable-targets/.
GWEC (Global Wind Energy Council). 2015. “Wind in numbers.” Accessed March 17,
2019. http://www.gwec.net/globalfigures/windinnumbers/.
H.R. 8444, 95th Congress. 1977–1978. “National Energy Act of 1978.” https://www.
congress.gov/bill/95th-congress/house-bill/8444.
H.R. 5263, 95th Congress. 1977–1978. “Energy Tax Act of 1978.” Pub. L. No. 95-618, 92
Stat. 3174. https://www.govtrack.us/congress/bills/95/hr5263/text.
H.R. 1424, 110th Congress. 2008. “The Emergency Economic Stabilization Act of 2008.”
Pub. L. No. 110-343, 122 Stat. 3765. https://www.congress.gov/110/plaws/publ343/
PLAW-110publ343.pdf.
H.R. 2029, 114th Congress. 2015–2016. “Consolidated Appropriations Act of 2016.” Pub.
L. No. 114-113, 129 Stat. 2424. https://www.congress.gov/114/plaws/publ113/PLAW-
114publ113.pdf.
H.R. 1, 115th Congress. 2017. “The Tax Cuts and Jobs Act of 2017.” Pub. L. No. 115-97, 131
Stat. 2054. https://www.congress.gov/115/bills/hr1/BILLS-115hr1enr.pdf.
H.R. 1892, 115th Congress. 2017–2018. “Bipartisan Budget Act of 2018.” Pub. L. No. 115-
123. https://www.congress.gov/115/plaws/publ123/PLAW-115publ123.pdf
H.R. 3301, 116th Congress. 2019–2020. “Taxpayer Certainty and Disaster Tax Relief Act
of 2019.” https://www.congress.gov/bill/116th-congress/house-bill/3301.
14 Renewable Energy Technologies and Water Infrastructure

IPCC (International Panel on Climate Change). 2014. “Climate change 2014: Synthesis
report.” Contribution of Working Groups I, II and III to the Fifth Assessment Report of
the Intergovernmental Panel on Climate Change. https://ar5-syr.ipcc.ch/.
JAXA (Japan Aerospace Exploration Agency). 2019. “Whole-atmosphere monthly mean
CO2 concentration based on GOSAT observations.” Accessed March 17, 2019. http://
www.gosat.nies.go.jp/en/recent-global-co2.html. 547.
KY-NREPC (Kentucky Natural Resources and Environmental Protection Cabinet). 2019.
“Dissolved oxygen and water quality.” State of Kentucky. Accessed May 4, 2019. http://
www.state.ky.us/nrepc/water/ramp/rmdo2.htm.
Lazarus, M., P. Erickson, and K. Tempest. 2015. Supply-side climate policy: The road less
taken. Stockholm Environment Institute. Accessed April 1, 2019. https://mediamanager.
sei.org/documents/Publications/Climate/SEI-WP-2015-13-Supply-side-climate-policy.
pdf.
MA-DER (Massachusetts Department of Energy Resources). 2019a. Leading by example
initiatives. MA-DER. Accessed April 6, 2019. https://www.mass.gov/service-details/
leading-by-example-initiatives#transportation.
MA-DER. 2019b. Renewable energy snapshot. Massachusetts Department of Energy
Resources. Accessed April 6, 2019. https://www.mass.gov/info-details/renew​able-ene​
rgy-snapshot.
Mass. (MA-USA) Exec. Order No. 484, April 18, 2007 https://www.mass.gov/files/
documents/2016/08/od/eo484.pdf.
McDermott, J. 2019. “Offshore wind developers to invest $4.5M in Rhode Island.” Phys.
org. Accessed May 7, 2019. https://phys.org/news/2019-04-offshore-invest-45m-rhode-
island.html.
McDonald, R. I., J. D. Olden, J. J. Opperman, W. M. Miller, J. Fargione, C. Revenga, et al.
2012. “Energy, water and fish: Biodiversity impacts of energy-sector water demand in
the United States depend on efficiency and policy measures.” PLoS One 7 (11): e50219.
NCSL (National Conference of State Legislators). 2019. “State renewable portfolio
standards and goals.” Accessed April 4, 2019. http://www.ncsl.org/research/energy/
renewable-portfolio-standards.aspx.
Niklitschek, E. J., and D. H. Secor. 2009. “Dissolved oxygen, temperature and salinity
effects on the ecophysiology and survival of juvenile Atlantic Sturgeon in estuarine
waters: I. Laboratory results. model development and testing.” J. Exp. Mar. Biol. Ecol.
381 (Supp-S): S150–S160.
Pandey, B., and A. Karki. 2017. Hydroelectric energy: Renewable energy and the environment.
Boca Raton, FL: CRC Press, Taylor & Francis.
Podesta, J., C. Goldfuss, T. Higgins, B. Bhattacharyya, A. Yu, and K. Costa. 2019. State
fact sheet: A 100 percent clean future. Washington, DC: Center for American Progress.
Rhode Island (USA) Exec. Order No. 20-01, January 17, 2020. https://governor.ri.gov/
documents/orders/Executive-Order-20-01.pdf.
S. 426, 99th Congress. 1985–1986. “Electric Consumers Protection Act of 1986.” Pub. L. No.
99-495, 100 Stat. 1243. https://www.congress.gov/bill/99th-congress/senate-bill/426.
S. 2289, 116th Congress. 2019–2020. “Renewable Energy Extension Act of 2019.” https://
www.congress.gov/bill/116th-congress/senate-bill/2289/text.
SEIA (Solar Energy Industries Association). 2019. Solar investment tax credit (ITC).
Washington, DC: SEIA.
Shrimali, G., M. Lynes, and J. Indvik. 2015. “Wind energy deployment in the U.S.: An
empirical analysis of the role of federal and state policies.” Renewable Sustainable
Energy Rev. 43 (C): 796–806.
US Renewable Energy Policy—Analysis and Recommendations 15

Stamford, L., and A. Azapagic. 2014. “Life cycle environmental impacts of UK shale gas.”
Appl. Energy 134 (1): 506–518.
Stoklosa, A. M., D. H. Keller, R. Marano, and R. J. Horwitz. 2018. A review of dissolved
oxygen requirements for key sensitive species in the Delaware estuary. Philadelphia:
Academy of Natural Sciences of Drexel Univ.
Than, K. 2018. Critical minerals scarcity could threaten renewable energy future. Stanford,
CA: Stanford Univ., School of Earth, Energy & Environmental Sciences.
Unger, D. J. 2016. “America now has 27.2 gigawatts of solar energy: What does that
mean?” Inside Climate News. Accessed April 4, 2019. https://insideclimatenews.org/
news/24052016/solar-energy-27-gigawatts-united-states-one-million-rooftop-panels-
climate-change-china-germany.
USEIA (United States Energy Information Administration). 2016. “Annual energy
outlook 2016.” Accessed April 1, 2019. https://www.eia.gov/outlooks/archive/aeo16/
mt_naturalgas.php#natgasprod_exp.
USEIA. 2018a. Petroleum, natural gas, and coal still dominate U.S. energy consumption.
Washington, DC: Office of Energy Efficiency & Renewable Energy. Accessed April 5,
2019. https://www.eia.gov/todayinenergy/detail.php?id=36612.
USEIA. 2018b. “The United States is now the largest global crude oil producer.” Independent
Statistics & Analysis. Accessed April 27, 2019. https://www.eia.gov/todayinenergy/
detail.php?id=37053.
USEIA. 2018c. “Direct federal financial interventions and subsides in energy in fiscal year
2016.” Independent Statistics & Analysis. Accessed May 16, 2019. https://www.eia.gov/
analysis/requests/subsidy/.
USEIA. 2019a. “What is U.S. electricity generation by energy source?” Independent
Statistics & Analysis. Accessed April 16, 2019. https://www.eia.gov/tools/faqs/faq.
php?id=427&t=3.
USEIA. 2019b. “Biofuels: Ethanol and biodiesel explained, ethanol and the environment.”
Independent Statistics & Analysis. Accessed April 5, 2019. https://www.eia.gov/
energyexplained/index.php?page=biofuel_ethanol_environment.
USEIA. 2019c. “Wind explained, where wind power is harnessed.” Independent Statistics
& Analysis. Accessed April 5, 2019. https://www.eia.gov/energyexplained/index.
php?page=wind_where.
USEIA. 2019d. “U.S. renewable electricity generation has doubled since 2008.” Independent
Statistics & Analysis. Accessed April 16, 2019. https://www.eia.gov/todayinenergy/
detail.php?id=38752.
UTIA (University of Tennessee Institute of Agriculture). 2019. “The Sun’s energy.” Solar
and Sustainable Energy. Accessed May 4, 2019. https://ag.tennessee.edu/solar/Pages/
What%20Is%20Solar%20Energy/Sun%27s%20Energy.aspx.
Virginia (USA) Exec. Order No. 43, September 16, 2019. https://www.governor.virginia.
gov/media/governorvirginiagov/executive-actions/EO-43-Expanding-Access-to-
Clean-Energy-and-Growing-the-Clean-Energy-Jobs-of-the-Future.pdf.
Wang, J. F., and N. Muller. 2012b. “Performance prediction of array arrangement on
ducted composite material marine current turbines (CMMCTs).” Ocean Eng. 41: 21–26.
Wang, J. F., J. Piechna, and N. Muller. 2012a. “A novel design of composite water turbine
using CFD.” J. Hydrodyn. 24 (1): 11–16.
WNA (World Nuclear Association) 2011. Comparison of lifecycle greenhouse gas emissions
of various electricity generation sources. WNA Rep. Accessed April 3, 2019. http://www.
worldnuclear.org/uploadedFiles/org/WNA/Publications/Working_Group_Reports/
comparison_of_lifecycle.pdf.
CHAPTER 2
Renewables and Regulatory
Requirements of the
United States
S. Rao Chitikela

INTRODUCTION

Renewable energy (RE) is that gets renewed in an inexhaustible way on Planet


Earth. The RE sources include (but not limited to) solar, wind, biofuels, geothermal,
tidal wave, hydroelectric, and the anthropogenic landfill gas (LFG). RE generation
and meeting the sustainability requirements is currently the preferred choice
of power generation in the United States and worldwide. The United States has
successfully embarked on incorporating renewables to a significant level into the
applicable statewide energy mix in generation and distribution, via the use of
effective energy policy and rulemaking. In the United States (and worldwide),
RE policies have been developed and implemented with standards to go with and
financial incentives—grants, loans, rebates, and tax incentives.
RE systems are (considered to be) environmentally friendly compared to
the fossil fuel-fired electric systems, could be a preferred distributed system in a
geographical area, and play a significant role in the avoidance or current control
of greenhouse gases (GHGs) worldwide. These RE resources would include (not-
limited-to) solar photovoltaic (PV) or solar thermal; wind; biogas-to-energy CHP
(where CHP is combined heat and power); landfill gas (LFG) CHP; hydroelectric-
power (hydel); geothermal (open- or closed-loop system); biomass; fuel cells,
including the microbial fuel cells (MFCs); and water wave. Net metering of
power usage is possible when RE systems are taken advantage of at the residential,
commercial, or municipal level, where the electricity billing would be based on
customer’s usage or return of power, from or to the grid.
In the United States, the Federal Energy Regulatory Commission (FERC) has
the authority and the jurisdiction over interstate and wholesale electric commerce
and NO jurisdiction over the local distribution of electricity, retail sales, siting,

17
18 Renewable Energy Technologies and Water Infrastructure

construction, environmental matters, and safety requirements, as reported (NREL


2016). The North American Electric Reliability Corporation (NERC) oversees the
reliability of the bulk electric system, where it verifies the system reliability and
standards’ development and enforcement. The independent system operators
(ISOs) and regional transmission operators (RTOs) are identified and utilized
to operate the electrical system under FERC; the electric-grid areas not under
an ISO or RTO are operated by the investor-owned utilities (IOUs), municipal
(cooperatives), or the Federal Power Marketing Administrations. In all, the
balancing authorities—responsible for integrating resource plans, maintaining
balance of generation and load within its area, and supporting interconnection
frequency in real time—and ISOs and RTOs are required to be in compliance with
reliability standards of NERC. Thus, various federal, state, and local regulatory
requirements on RE resources and systems are elucidated (to the extent feasible)
in the following sections.
The global development and operation of renewables is at 23.9% of the
electricity generation and is projected to increase to 29.4% by 2023. Globally, the
electricity accounts for a fifth of energy consumption and, so, a rapid application
and usage of renewables in the transportation and heating sectors is required
(IEA 2019).
Renewable energy production and consumption of the United States since
the year 2000 is shown in Figure 2-1. The total renewable energy consumption
has been (more than) approximately 9,488 quadrillion J (9.0 quadrillion Btu) in
the United States (USEIA 2019).
As seen in Figure 2-1, the types of renewables included are hydroelectric,
geothermal, solar, wind, wood-biomass, ethanol, biodiesel, and waste biomass; a
Btu is approximately 1,054.2 J and can be used for unit conversion.

Figure 2-1. US’ renewable energy—production and consumption.


Source: USEIA (2019).
Renewables and Regulatory Requirements of the United States 19

US-CODE, LAW, AND ACT—RENEWABLE ENERGY

The “US Title 42—The Public Health and Welfare, Chapter 125—Renewable
Energy and Energy Efficiency Technology Competitiveness,” includes the
Congress-finding, national goals for renewable energy, and energy efficiency
authorizations. Under the 42USC, §12001, the Congress finds “… it is in the
national security and economic interest of the United States to foster greater
efficiency in the use of available energy supplies and greater use of renewable
energy technologies.” In addition, the purpose is “… to pursue an aggressive
national program of research, development, demonstration, and commercial
application of renewable energy and energy efficiency technologies in order to
ensure a stable and secure future energy supply …” The national goals (and the
multiyear funding, on selected REs) for wind, PV, solar thermal, alcohol from
biomass, biofuel energy systems, biodiesel energy systems, hydrogen energy
systems, solar building energy systems, ocean energy systems, geothermal
energy systems, low head hydro, and energy storage systems are specified in the
42USC, §12003.
The RE project technologies are included in 42USC, §12005(c)(2): Projects
under this section may include the following technologies:
1. Conversion of cellulosic biomass to liquid fuels.
2. Ethanol and ethanol by-product processes.
3. Direct combustion or gasification of biomass.
4. Biofuel energy systems.
5. Photovoltaics, including utility scale and remote applications.
6. Solar thermal, including solar water heating.
7. Wind energy.
8. High-temperature and low-temperature geothermal energy.
9. Fuel cells, including transportation and stationary applications.
10. Nondefense high-temperature superconducting electricity technology.
11. Source reduction technology.
12. Factory-made housing.
13. Advanced district cooling.
where the term “source reduction” means [as included in 42USC, §12002(5)] any
practice which reduces the amount of any hazardous substance, pollutant, or
contaminant entering any waste stream or otherwise released into the environment,
including fugitive emissions, prior to recycling, treatment, or disposal; and reduces
the hazards posed to the public health and the environment associated with the
release of such substances, pollutants, or contaminants, including equipment or
technology modifications, process or procedure modifications, reformulation
20 Renewable Energy Technologies and Water Infrastructure

or redesign of products, substitution of raw materials, and improvements in


housekeeping, maintenance, training, and inventory control but not including
any practice that alters the physical, chemical, or biological characteristics or the
volume of a hazardous substance, pollutant, or contaminant through a process
or activity which itself is not integral to and necessary for the manufacture of a
product or providing a service (42USC 2019a).
The Energy Policy requirements are included in the 42USC, Chapter 134,
§13201 to 13574; Subchapter V includes the applicable requirements of Renewable
Energy, §13311 to 13317, that explains the renewable energy export technology
training; renewable energy advancement awards; study of tax and rate treatment
of renewable energy projects; data system and energy technology evaluation;
innovative renewable energy technology transfer program; and renewable
energy production incentive. The “42USC, Chapter 149—National Energy Policy
and Programs, Subchapter II—Renewable Energy,” §15851, includes that the
RE Resource Assessment be conducted—“Secretary (of the Department) shall
review the available assessments of renewable energy resources within the United
States, including solar, wind, biomass, ocean (including tidal, wave, current, and
thermal), geothermal, and hydroelectric energy resources, and undertake new
assessments as necessary, taking into account changes in market conditions,
available technologies, and other relevant factors” and the reports be published
(42USC 2019b).
The Energy Policy Act of 2005 (enacted January 4, 2005) includes Title II—
Renewable Energy; Title VIII—Hydrogen; Title IX—Research and Development,
Subtitle C—Renewable Energy; Title XIII—Energy Policy Tax Incentives; Title
XVI—Climate Change; and Title XVIII—Studies, to list a few relevant to further
review of this enacted law-text (US-Congress 2005).

RENEWABLE ENERGY STANDARDS

The renewable energy standards (RES) or the renewable portfolio standards (RPS)
are adopted and established by the states, where the applicable requirements
vary state-to-state. It is envisaged that at 20% or more of the electric-suite, the
application-of or drawing energy via use of renewables would be significant in
that it supports a good control of air pollution (otherwise, facing the air pollution
due to 100% firing of fossil fuels). Thus, RE credits or certificates (RECs) are
also included under the RE regulations. This US renewable energy market
has been estimated at $US 64 billion. Table 2-1 shows the US legislation(s) on
accomplishment of renewables (NCSL 2019).
The procurement of renewable energy is critical in identifying various
elements of capital expenditure (CapEx), operational expenditure (OpEx), and
operation and maintenance (O&M) and in working with the stakeholders on
a long-term basis. The significant participants and elements are as follows: the
Table 2-1. The United States’ Renewable Portfolio or Energy Standards and Goals.

State Year established Renewable energy requirement Statute, code, or order


Alaska 2009–2010 the state receive 50% of its electrical House Bill 306
Legislative generation from renewable energy
session sources by 2025
Arizona 2006 15% by 2025 “Ariz. Admin. Code §14-2-1801 et seq.”
California 2002 44% by 2024; 52% by 2027; 60% by 2030, “Cal. Public Utilities Code §399.11 et seq.;
and also requires 100% clean energy Cal. Public Resources Code §25740 et
by 2045 seq.; Assembly Bill 327 (2013); Senate
Bill 350 (2015); Senate Bill 100 (2018).”
Colorado 2004 30% by 2020 (IOUs); 10% or 20% for Colo. Rev. Stat. §40-2-124; Senate Bill 252
municipalities and electric cooperatives (2013)
depending on size
Connecticut 1998 44% by 2030 (the State website shows 48%) Conn. Gen. Stat. §16-245a et seq.; Conn.
Gen. Stat. §16-1; Senate Bill 9 (2018)
Delaware 2005 25% by 2025–2026 Del. Code Ann. 26 §351 et seq.
Hawaii 2001 30% by 2020; 40% by 2030; 70% by 2040; Hawaii Rev. Stat. §269-91 et seq.; House
100% by 2045 Bill 623 (2015)
Illinois 2001 (voluntary 25% by 2025–2026 Ill. Rev. Stat. ch. 20 §688 (2001); Ill. Rev.
target); 2007 Stat. ch. 20 §3855/1-75 (2007); Senate
(standard). Bill 2814 (2016)
Indiana 2011 10% by 2025 Ind. Code §8-1-37
Iowa 1983 105 MW of generating capacity for IOUs Iowa Code §476.41 et seq.
Renewables and Regulatory Requirements of the United States

(investor-owned utility)
Kansas 2009 (standard); 15% by 2015–2019; 20% by 2020 Kan Stat. Ann. §66-1256 et seq.; Goal:
2015 (goal). Senate Bill 91
21

(Continued)
Table 2-1. The United States’ Renewable Portfolio or Energy Standards and Goals. (Continued)
22

State Year established Renewable energy requirement Statute, code, or order

Maine 1999 40% by 2017 Me. Rev. Stat. Ann. 35-A §3210 et seq.;
Me. Rev. Stat. Ann. 35-A §3401 et seq.
(wind energy)
Maryland 2004 25% by 2020 Md. Public Utilities Code Ann. §7-701 et
seq.; Senate Bill 921; House Bill 1106
(2016 enrolled, 2017 veto override)
Massachusetts 1997 Class I (new sources): 35% by 2030 and an Mass. Gen. Laws Ann. ch. 25A §11F;
additional 1% each year after. Class II: House Bill 4857 (2018)
6.7% by 2020
Michigan 2007 15% by 2021 (standard), 35% by 2025 Mich. Comp. Laws §460.1001 et seq.;
(goal, including energy efficiency and Senate Bill 438 (2016)
demand reduction)
Minnesota 2007 26.5% by 2025 (IOUs), 25% by 2025 (other Minn. Stat. §216B.1691
utilities)
Missouri 2007 15% by 2021 (IOUs) Mo. Rev. Stat. §393.1020 et seq
Montana 2005 15% by 2015 Mont. Code Ann. §69-3-2001 et seq.
Nevada 1997 25% by 2025 Nev. Rev. Stat. §704.7801 et seq.
New 2007 25.2% by 2025 N.H. Rev. Stat. Ann. §362-F
Hampshire
New Jersey 1991 50% by 2030 N.J. Rev. Stat. §48:3-49 et seq.; Assembly
Renewable Energy Technologies and Water Infrastructure

Bill 3723 (2018)


New Mexico 2002 20% by 2020 (IOUs); 10% by 2020 (co-ops) N.M. Stat. Ann. §62-15; N.M. Stat. Ann.
§62-16
(Continued)
Table 2-1. The United States’ Renewable Portfolio or Energy Standards and Goals. (Continued)

State Year established Renewable energy requirement Statute, code, or order

New York 2004 50% by 2030 NY PSC Order Case 03-E-0188; 2015 New
York State Energy Plan.
North Carolina 2007 12.5% by 2021 (IOUs); 10% by 2018 (munis N.C. Gen. Stat. §62-133.8
and coops)
North Dakota 2007 10% by 2015 N.D. Cent. Code §49-02-24 et seq.
Ohio 2008 12.5% by 2026. Senate Bill 310 (2014) Ohio Rev. Code Ann. §4928.64 et seq.
created a 2-year freeze on the state’s
standard while a panel studied the costs
and benefits of the requirement. The
freeze was not extended in 2016
Oklahoma 2010 15% by 2015 Okla. Stat. tit. 17 §801.1 et seq.
Oregon 2007 25% by 2025 (utilities with 3% or more of Or. Rev. Stat. §469a; Senate Bill 1547
the state’s load); 50% by 2040 (utilities (2016)
with 3% or more of the state’s load);
10% by 2025 (utilities with 1.5%–3% of
the state’s load); 5% by 2025 (utilities
with less than 1.5% of the state’s load)
Pennsylvania 2004 18% by 2020–2021 Pa. Cons. Stat. tit. 66 §2814
Rhode Island 2004 14.5% by 2019, with increases of 1.5% R.I. Gen. Laws §39-26-1 et seq.; R.I. Gen.
each year until 38.5% by 2035. Laws §39-26.1 et seq. (contracting
Renewables and Regulatory Requirements of the United States

standard); House Bill 7413a (2016)


South Carolina 2014 2% by 2021 House Bill 1189
South Dakota 2008 10% by 2015 S.D. Codified Laws Ann. §49-34A-94; S.D.
Codified Laws Ann. §49-34A-101 et seq.
23

(Continued)
Table 2-1. The United States’ Renewable Portfolio or Energy Standards and Goals. (Continued)
24

State Year established Renewable energy requirement Statute, code, or order

Texas 1999 5,880 MW by 2015. 10,000 MW by 2025 Tex. Utilities Code Ann. §39.904
(goal; achieved)
Utah 2008 20% by 2025 Utah Code Ann. §54-17-101 et seq.; Utah
Code Ann. §10-19-101 et seq.
Vermont 2005 (voluntary 55% by 2017; 75% by 2032 Vt. Stat. Ann. tit. 30 §8001 et seq.;
target); 2015 Standard: House Bill 40
(standard)
Virginia 2007 15% by 2025 Va. Code §56-585.2
Washington 2006 15% by 2020 Wash. Rev. Code §19.285; Wash. Admin.
Code §480-109; Wash Admin. Code
§194-37
West Virginia Established: 2009; 10% from 2015 to 2019, 15% from 2020 to W. Va. Code §24-2F; Repeal: House
Repealed 2015 2024, 25% by 2025 Bill 2001
Wisconsin 1998 10% by 2015 Wisc. Stat. §196.378
Washington, 2005 20% by 2020, 100% by 2032 D.C. Code §34-1431 et seq., Bill 650
DC (2016); Bill 904 (2018)
Guam 2008 25% by 2035 Guam Public Law §29-62
Northern 2007; goal 20% by 2016 N. M. I. Public Law §15-23; House Bill
Mariana reduced in 165 (2014)
Islands 2014
Renewable Energy Technologies and Water Infrastructure

Puerto Rico 2010 20% by 2035 PR S 1519 (2010); PR H 2610 (2010)


US Virgin 2009 20% by 2015; 25% by 2020; 30% by 2025; VI B 9 (2009)
Islands up to 51% after 2025
Source: NCSL (2019).
Renewables and Regulatory Requirements of the United States 25

public utility commission (PUC)—A regulatory body that oversees regulation


of rates and services of a public utility. It can also be referred to as a utilities
commission, utility regulatory commission, or public service commission; an
independent power producer (IPP) would be “A corporation, person, agency,
authority, other legal entity or instrumentality that owns or operates facilities
for the generation of electricity for use primarily by the public, and that is not
an electric utility”; a power purchase agreement (PPA) would be “A financial
mechanism through which a bulk electricity customer enters into a long-term
electricity supply contract with an IPP. The contract defines all the terms and
conditions of the transaction of electricity sales between the supplier and buyer,
including dates of commencement and termination, terms of payment, schedule
for delivery, penalties, and exclusions.” In addition, the variable renewable energy
(VRE) generator—a renewable electricity generator, such as wind and or solar
power plants—provides variable and no dispatchable power output owing to the
natural variability of the energy resource. The integration of renewable energy
resources or the infrastructure to the existing power grid has also been evolved
via restructuring the wholesale and retail electricity markets. “Restructured
wholesale markets comprise a range of different market-based approaches
to balancing bulk power supply and demand while creating an institutional
separation between transmission operations and generation. Retail restructuring
refers to the introduction of consumer choice among competing suppliers of end-
use electric services” (NREL 2016).

RENEWABLE PORTFOLIO STANDARDS—THE STATE OF


CONNECTICUT, EXAMPLE

The State of Connecticut uses the RPS policy that requires the electric providers to
annually maintain a “specified percentage or amount of the energy they generate
or sell from renewable sources.” One renewable energy certificate (REC) for each
of 3.6 × 109 J (1 MWh) of electricity produced will be issued to the qualified
electric providers. The qualified renewables are categorized into three classes—
Class I, Class II, or Class III. The regulatory information on Classes I, II, and III,
and the “‘required Annual Renewable Energy Percentages” with respect to the
said classes are, as follows (CT-PURA 2019):
Class I renewable energy source, as defined in §16-1(a)(20) of the General
Statutes of Connecticut (Conn. Gen. Stat.), means electricity derived from:
solar power; wind power; fuel cell; geothermal; landfill methane gas, anaerobic
digestion or other biogas derived from biological sources; thermal electric direct
energy conversion from a certified Class I renewable energy source; ocean
thermal power; wave or tidal power; low-emission advanced renewable energy
conversion technologies, including, but not limited to, zero emission low-grade
heat power generation systems based on organic oil-free rankine, kalina, or
26 Renewable Energy Technologies and Water Infrastructure

similar “nonstream” (nonsteam) cycles that use waste heat from an industrial
or commercial process that does not generate electricity; run-of-the-river
hydropower facility that began operation after July 1, 2003, and has a generating
capacity of not more than 30 MW, or a run-of-the-river hydropower facility that
received a new license after January 1, 2018, under the FERC rules pursuant to 18
CFR 16, as amended from time to time, and provided the facility is not based on
a new dam or a dam identified as a candidate for removal; and biomass facility
that uses sustainable biomass fuel, as defined in Conn. Gen. Stat. §16-1(a)(39)
(cultivated and harvested in a sustainable manner). “Sustainable biomass fuel”
does not mean construction and demolition waste, finished biomass products
from sawmills, paper mills, or stud mills; organic refuse fuel derived separately
from municipal solid waste (MSW); or biomass from old growth timber stands,
except where (1) such biomass is used in a biomass gasification plant that received
funding prior to May 1, 2006, from the Clean Energy Fund established pursuant
to Section 16-245n, or (2) the energy derived from such biomass is subject to a
long-term power purchase contract pursuant to Subdivision (2) of Subsection
(j) of Section 16-244c entered into prior to May 1, 2006) and meets certain
emissions requirements; and any electrical generation, including distributed
generation, generated from a Class I renewable energy source, provided, on and
after January 1, 2014, any megawatt hours that are claimed or counted toward
compliance in another province or state, other than Connecticut, shall not be
eligible.
Class II renewable energy source, as defined in Conn. Gen. Stat. §16-1(a)
(21), means electricity derived from a trash-to-energy facility that has obtained a
permit pursuant to Section 22a-208a and Section 22a-174-33 of the regulations of
Connecticut state agencies.
Class III source, as defined in Conn. Gen. Stat. §16-1(a)(38), means
• Electricity output from combined heat and power systems with a minimum
operating efficiency of 50% that are part of customer-side distributed
resources developed at commercial and industrial facilities in Connecticut
on or after January 1, 2006;
• Waste heat recovery systems installed on or after April 1, 2007, that produces
electrical or thermal energy by capturing preexisting waste heat or pressure
from industrial or commercial processes;
• Electricity savings from conservation and load management programs that
started on or after January 1, 2006 (on and after January 1, 2014, programs
supported by ratepayers are not eligible); and
• Any demand-side management project awarded a contract pursuant to
§16-243 m (eligibility is based on the term of the contract).
Table 2-2 shows the State of Connecticut requirements of Classes I, II,
and III renewable energy contributions on an annual basis, since 2018 and up
to 2030.
Renewables and Regulatory Requirements of the United States 27

Table 2-2. Required, Annual Renewable Energy Percentages.

Class I Class II or Class I Class III Total


Year (%) (add’l) (%) (%) (%)
2018 17.0 4.0 4.0 25.0
2019 19.5 4.0 4.0 27.5
2020 21.0 4.0 4.0 29.0
2021 22.5 4.0 4.0 30.5
2022 24 4.0 4.0 32
2023 26 4.0 4.0 34
2024 28 4.0 4.0 36
2025 30 4.0 4.0 38
2026 32 4.0 4.0 40
2027 34 4.0 4.0 42
2028 36 4.0 4.0 44
2029 38 4.0 4.0 46
2030 40 4.0 4.0 48
Note: As shown, the Class I renewables’ requirement is projected to 40%, while
the Classes II and III requirements are no change at 4% each, by Year 2030 in
the State of Connecticut.

RENEWABLE PORTFOLIO STANDARDS ELIGIBILITY—THE STATE


OF CALIFORNIA, EXAMPLE

The eligible electric generation facilities will be included in the RPS procurement
process and, thus, provided with the RPS Certification. Therefore, to qualify
for the RPS Certification, the eligible electric-generating facilities need to use
one or more RE resources that meeting the resource-specific requirements. The
CA Energy Commission publishes the RPS Eligibility Guidebook that must be
followed by the applicants for RPS Certification. Chapter 2 of the Guidebook
provides the RE resources and the eligibility criteria in the State of California,
and the RE resource type and eligibility are as follows (for complete details, verify
the CA Energy Commission’s regulatory requirements) (Green and Crume 2017):

Biodiesel: “A facility may qualify for RPS certification if it generates electricity


using biodiesel derived from biomass feedstock or from an eligible solid waste
conversion process using municipal solid waste.
Biomass: “A facility may qualify for RPS certification if it generates electricity
using a biomass fuel.” A fuel that is resulting from “biomass conversion” is
inclusive.
Biomethane: “A facility may qualify for RPS certification if it generates
electricity using biomethane derived from digester gas and/or landfill gas.
28 Renewable Energy Technologies and Water Infrastructure

Biomethane may be used to generate electricity at a facility that receives


the biomethane in one of four ways: (1) onsite generating facility using a
dedicated pipeline; (2) offsite generating facility using a dedicated pipeline;
(3) offsite generating facility using a fuel container; and (4) offsite generating
facility using a common carrier pipeline.
Fuel cell using renewable fuel: “A facility that uses a fuel cell conversion
technology may qualify for RPS certification if the facility uses either an
RPS-eligible renewable energy resource, qualifying hydrogen gas, or both.
Geothermal: “A facility may qualify for RPS certification if it generates
electricity using a geothermal resource. Only natural heat from within the
earth that is captured for electric power production may be used to create
RPS-eligible geothermal generation.”
Hydroelectric: “The following types of hydroelectric facilities may be RPS
eligible: (1) Small hydroelectric facilities 30 MW or less. (2) Conduit
hydroelectric facilities 30 MW or less. (3) Hydroelectric generation units
40 MW or less and operated as part of a water supply or conveyance system.
(4) Incremental hydroelectric facilities.
Municipal solid waste: “A facility may qualify for RPS certification if it generates
electricity using municipal solid waste (MSW) in either a combustion or
conversion process…”
Ocean thermal: “A facility may qualify for RPS certification if it generates
electricity using an ocean thermal resource, such as the temperature
differences between deep and surface ocean water.”
Ocean wave: “A facility may qualify for RPS certification if it generates electricity
using an ocean wave.”
Solar: “A facility may qualify for RPS certification if it generates electricity using
either a photovoltaic or solar thermal process to produce electricity.”
Tidal current: “A facility may qualify for RPS certification if it generates
electricity using a tidal current.”
Wind: “A facility may qualify for RPS certification if it generates electricity
using a wind resource. Facilities using wind resources can use any method
to capture the naturally occurring wind, convert it to mechanical energy, and
then generate electricity.”

LOCAL GOVERNMENT ON RENEWABLE ENERGY


PROJECTS—STATE OF VIRGINIA, EXAMPLE

The local governments must also include the responsibilities in the development,
installation, and operation of RE projects.
Renewables and Regulatory Requirements of the United States 29

In the State of Virginia (VA), the Local Government Outreach Stakeholder


Group (LOG) holds the main responsibility of siting the RE projects. Thus, with
the guidance of LOG, VA developed model ordinances for wind and solar RE
projects in the state. For example, the Model Utility Scale Wind Ordinance (dated
April 4, 2012) includes purpose; applicability; definitions; type of permitting;
applications and procedures, location, appearance, and operation of a project
site; safety and construction; and decommissioning (VA-RE 2019):
The “purpose”2 statement, as follows—The purpose of this ordinance
is to provide for the siting, development, and decommissioning of
utility-scale wind energy projects in [locality], subject to reasonable
conditions that promote and protect the public health, safety and welfare
of the community while promoting development of renewable energy
resources.3
2Purpose. The statement of purpose is based on similar provisions found
in existing ordinances and models and was acceptable to most LOG
members. The phrase “promoting development of renewable energy
sources” conforms with Virginia’s Energy Policy (specifically, §67-103
of the Code of Virginia). The legal requirements of this Energy Policy
are discussed in the companion document, “Introduction: DEQ’s Local
Government Outreach for Renewable Energy,” which appears on DEQ’s
website along with this model ordinance. One LOG member believed
that including the promotion of renewable energy as part of the Purpose
was going further than necessary. However, a local government chooses
to articulate the Purpose of its wind ordinance, it should keep in mind
the statutory mandate, incumbent on both local and state government
entities, to promote the development of renewable energy.
3Public Health, Safety, and Welfare. This model ordinance addresses
local governments’ traditional areas of responsibility—public health,
safety, and welfare—as they relate to wind energy projects. The model
ordinance does not address protection of natural resources. In Virginia, the
Department of Environmental Quality (DEQ) regulates impacts of wind
projects on wildlife and historic resources pursuant to 9VAC15-40. The
Virginia legislature delegated this authority to DEQ pursuant to the Small
Renewable Energy Projects Act of 2009 (§10.1-1197.5 et seq. of the Code of
Virginia). Other natural resources are regulated via permits administered
by DEQ and other agencies or levels of government (e.g., air, water, waste,
erosion and sediment control) pursuant to other state or federal laws.
A few critical terms of definitions are important in understanding and
interpreting a law, policy and/or regulation(s), as follows (and included in the
model-ordinance):
“Applicant” means the owner or operator who submits an application to the
locality for a permit to install a wind energy project under this ordinance.
30 Renewable Energy Technologies and Water Infrastructure

“Landowner” means the person who owns all or a portion of the real
property on which a wind energy project is constructed.
“Operator” means the person responsible for the overall operation and
management of a wind energy project.
“Owner” means the person who owns all or a portion of a wind energy
project.
“Rated capacity” means the maximum capacity of a wind energy
project based on the sum total of each turbine’s nameplate capacity. The
nameplate capacity is typically specified by the manufacturer with a label
on the turbine equipment.
“Wind energy project, utility-scale”4 means a facility that generates
electricity from wind, and consists of (1) one or more wind turbines
and other accessory structures and buildings, including substations,
post-construction meteorological towers, electrical infrastructure, and
other appurtenant structures and facilities within the boundaries of the
site, and (2) is designed for, or capable of, operation at a rated capacity
greater than 5-MW.5 Two or more wind turbines otherwise spatially
separated but under common ownership or operational control, which
are connected to the electrical grid under a single interconnection
agreement, shall be considered a single utility-scale wind energy project.
4Definition of “Project.” In land use zoning and ordinances, one of
the first issues is how to name and define wind energy installations.
Commonly used terms include windmills, turbines, wind energy
facilities, wind energy systems, and wind energy conversion systems.
Although several planning experts recommended using the term facility,
LOG members recommended using project wherever the term would
fit, in order to be consistent with Virginia’s Small Renewable Energy
Projects Act of 2009 (hereinafter “2009 statute”).
5Rated Capacity of Utility-Scale Wind Project. This model ordinance
utilizes “greater than 5 MW” to define the size project addressed
by utility-scale wind projects in order to coordinate with the 2009
statute; however, local governments may wish to alter this number.
One alternative mentioned was “>5 MW or 2 or more turbines.” In
determining how to define the project sizes addressed by a utility-scale
ordinance, local governments may want to keep in mind the tiers or
levels of rated capacity that could be addressed in community-scale
ordinances. The LOG is also framing a model wind ordinance for
community-scale projects.
“Wind turbine” means a wind energy conversion system that converts
wind energy into electricity through the use of a wind turbine generator
that typically consists of a tower, nacelle, rotor, blades, controller and
associated mechanical and electrical conversion components.
Renewables and Regulatory Requirements of the United States 31

The types of permitting can be based on—by right or permitted use,


accessory uses, special exception/conditional use/special use permitting.
The application and procedures would include: project description, site
plan, documentation of right to use property for the proposed project,
decommissioning plan, liability insurance, etc.
The location, appearance, and operation of a project site would include:
visual appearance, visual impacts, lighting, signage, noise, shadow
flicker, height, setbacks, use of public roads, etc.
The safety and construction would include: design, climb prevention/
locks, warnings, ground clearance, speed controls and brakes, emergency
response plan, signal interference, construction and installation, etc.
The decommissioning details would include: decommissioning plan,
discontinuation or abandonment of project, surety, etc.
Thus, VA has developed model ordinances for community-scale wind,
residential-scale wind, research studies and other resources, evaluating sources
on wind energy, larger and smaller scale solar model(s), and solar tax exemption
model (VA-RE 2019).


SOLAR AND WIND ENERGY RULE—BUREAU OF LAND
MANAGEMENT

The BLM under the US Department of Interior enacted the Solar and Wind
Energy Rule via the amendment of Title V (Rights-of-Way) of the Federal Land
Policy and Management Act (FLPMA), where the regulatory requirements are
included in the Code of Federal Regulations (CFR), under Title 43, Public Lands:
Interior—the 43CFR, Part 2800—Rights-of-Way under the Federal Land Policy
and management Act (43CFR 2019). The Right-of-Way is defined under the act—
“includes an easement, lease, permit, or license to occupy, use, or traverse public
lands granted for the purpose listed in title V of this Act” (DOI/BLM 2016).
A few definitions of terms relevant to RE development under this Rule 43CFR,
Part 2800, §2801.5 are as follows:
Designated leasing area means a parcel of land with specific boundaries
identified by the BLM land use planning process as being a preferred location
for solar or wind energy development that may be offered competitively.
Megawatt (MW) capacity fee means the fee paid in addition to the
acreage rent for solar and wind energy development grants and leases.
The MW capacity fee is the approved MW capacity of the solar or wind
energy grant or lease multiplied by the appropriate MW rate. A grant or
lease may provide for stages of development, and the grantee or lessee
will be charged a fee for each stage by multiplying the MW rate by the
approved MW capacity for the stage of the project.
32 Renewable Energy Technologies and Water Infrastructure

Megawatt rate means the price of each MW of capacity for various solar
and wind energy technologies as determined by the MW rate formula.
Current MW rates are found on the BLM’s MW rate schedule, which can
be obtained at any BLM office or at http://www.blm.gov. The MW rate
is calculated by multiplying the total hours per year by the net capacity
factor, by the MW hour (MWh) price, and by the rate of return, where:
1. Net capacity factor means the average operational time divided by the
average potential operational time of a solar or wind energy development,
multiplied by the current technology efficiency rates. The BLM establishes
net capacity factors for different technology types but may determine
another net capacity factor to be more appropriate, on a case-by-case or
regional basis, to reflect changes in technology, such as a solar or wind
project that employs energy storage technologies, or if a grant or lease
holder or applicant is able to demonstrate that another net capacity factor
is appropriate for a particular project or region. The net capacity factor for
each technology type is:
i. Photovoltaic (PV)—20 percent;
ii. Concentrated photovoltaic (CPV) and concentrated solar power (CSP)—
25 percent;
iii. CSP with storage capacity of 3 hours or more—30 percent; and
iv. Wind energy—35 percent;
2. Megawatt hour (MWh) price means the 5 calendar-year average of the annual
weighted average wholesale prices per MWh for the major trading hubs
serving the 11 western States of the continental United States (US); and
3. Rate of return means the relationship of income (to the property owner)
to revenue generated from authorized solar and wind energy development
facilities based on the 10-year average of the 20-year US Treasury bond yield
rounded to the nearest one-tenth percent.
Screening criteria for solar and wind energy development refers to the policies
and procedures that the BLM uses to prioritize how it processes solar and wind
energy development right-of-way applications to facilitate the environmentally
responsible development of such facilities through the consideration of resource
conflicts, land use plans, and applicable statutory and regulatory requirements.
Applications for projects with lesser resource conflicts are anticipated to be less
costly and time-consuming for the BLM to process and will be prioritized over
those with greater resource conflicts.
The high-priority applications are categorized, according to 43CFR-§2804.35,
that meet the criteria, as follows:
1. Lands specifically identified as appropriate for solar or wind energy
development, other than designated leasing areas (DLAs);
2. Previously disturbed sites or areas adjacent to previously disturbed or
developed sites;
Renewables and Regulatory Requirements of the United States 33

3. Lands currently designated as Visual Resource Management Class IV; or


4. Lands identified as suitable for disposal in BLM land use plans.
The medium- and low-priority applications based on the federal land criteria
are also included in this section. Sections 43CFR-§2806.50 to §2806.68 include the
rents and fees requirements for solar and wind energy rights-of-way.
The rule provides needed operator flexibility, bidding, and RE development
in the DLAs that have “high generation with low resource conflicts.”
The BLM also leases land for geothermal energy projects. As of March 2018, BLM
manages 50 geothermal leases that produce approximately 5.93 BJ/hr (1,648 MW)
(which amounts to more than 40% of US geothermal energy); moreover, these leases
provide alternative heat sources for direct use. The production of geothermal energy
on these federal lands is projected to meet the United States, target of electricity
production from RE sources. The environmental impact statement (EIS) related
to this geothermal leasing and via the Record of Decision, allocated approximately
449,206 km2 (111 million acres) of land, and another approximately 319,705 km2 (79
million acres) of Forest Service land is also available (BLM 2018).

RE PROGRAMS—BUREAU OF OCEAN ENERGY MANAGEMENT

BOEM works under the US Department of Interior, conducts and is responsible


for the RE projects on the outer continental shelf (OCS). Similar to BLM,
BOEM’s RE program authorizes the leases and rights-of-way for offshore RE
projects. The 30CFR—Mineral Resources, Part 585—Renewable Energy and
Alternate Uses of Existing Facilities on the Outer Continental Shelf, includes
the regulatory requirements on developing the RE projects in the OCS. A few
selected definitions of RE terms, as included in the 30CFR, Section §585.112, are
as follows (30CFR 2019):
Renewable energy means energy resources other than oil and gas and
minerals as defined in 30 CFR Part 580. Such resources include, but are
not limited to, wind, solar, and ocean waves, tides, and current.
Right-of-use and easement (RUE) grant means an easement issued by
BOEM under this part that authorizes use of a designated portion of
the OCS to support activities on a lease or other use authorization for
renewable energy activities. The term also means the area covered by the
authorization.
Right-of-way (ROW) grant means an authorization issued by BOEM
under this part to use a portion of the OCS for the construction and
use of a cable or pipeline for the purpose of gathering, transmitting,
distributing, or otherwise transporting electricity or other energy
product generated or produced from renewable energy, but does not
constitute a project easement under this part. The term also means the
area covered by the authorization.
34 Renewable Energy Technologies and Water Infrastructure

Figure 2-2. The BOEM regulatory roadmap for developing a wind energy facility.
Source: BOEM (2019).

BOEM provides a schematic (Figure 2-2) of the Regulatory Roadmap (based on


30CFR, Part 585) on requirements to an offshore wind energy facility (BOEM 2019).
BOEM has developed and provides the national and regional guidelines for
RE projects on the OCS. The good understanding of these regulatory requirements
including the timelines, application, and area-leasing guidance would lead to
a better “RE Project Plan Submittal” and the approvals toward a possible 25+
year(s) of RE project.
For more information on regulatory frameworks (worldwide) for RE projects
in marine environment, the TETHYS (named after the Greek-Titaness of the
Sea) is developed by the Pacific Northwest National Laboratory (PNNL) of the
Department of Energy (DOE) with the primary functions (TETHYS 2019) “To
facilitate the exchange of information and data on the environmental effects of
wind and marine renewable energy technologies; and, To serve as a commons
for wind and marine renewable energy practitioners and therefore enhance the
connectedness of the renewable energy community as a whole.”

RENEWABLES—FEDERAL ENERGY REGULATORY COMMISSION

FERC is an independent federal entity that includes licensing “hydropower


projects.” FERC regulates the transmission and wholesale sales of electricity
in interstate commerce; reviews the siting application for electric transmission
projects under limited circumstances; licenses and inspects private, municipal, and
state hydroelectric projects; monitors and investigates energy markets; enforces
FERC regulatory requirements through imposition of civil penalties and other
means; oversees environmental matters related to natural gas and hydroelectricity
Renewables and Regulatory Requirements of the United States 35

projects and other matters; and administers accounting and financial reporting
regulations and conduct of regulated companies. However, FERC (works-with or)
observes the State Public Utility Commissions, which would be responsible on
areas (as applicable) outside of FERC.
The use of renewable energy resources to generate electricity has the
potential to be a cost-effective means not only to reduce greenhouse
gas emissions, but also to diversify the fuels used to generate electricity.
The Commission will continue to pursue market reforms to allow
all resources, including renewable energy resources, to compete in
jurisdictional markets on a level playing field. These efforts could include
amendments to market rules, the modification or creation of ancillary
services and related policies, or the implementation of operational
tools that support the reliable integration of renewable resources. By
implementing these or other reforms, the Commission’s actions have
the potential to increase the amount of electricity being produced from
renewable energy resources. (FERC 2019).
The other useful entities and/or databases would be the DOE’s Office of Energy
Efficiency and Renewable Energy—https://www.energy.gov/eere/office-energy-
efficiency-renewable-energy (July 22, 2019) and the Database of State Incentives
for Renewables & Efficiency (DSIRE)—https://www.dsireusa.org/ (July 22, 2019).

SUMMARY

The US renewable energy generation technology mix has been expansive, and the
total RE consumption surpassed approximately 9.5 quintillion J (9.0 quadrillion
Btu). The 42USC, Chapters 125, 134, and 149 address the RE legislative actions
and associated policies, and the implemented RE technologies and programs.
The active participation of various states on RE programs has been significant via
the rigorous implementation of RES or RPS (Table 2-1). The BLM implemented
the Solar and Wind Energy Rule via an amendment of FLPMA, to facilitate the
Rights-of-Way under the Public Lands, and those applicable requirements are
available in the 43CFR, Part 2800. Similarly, BOEM established the Rights-of-Way
for offshore RE programs (according to applicable OCS area); the 30CFR, Part 585,
provides or includes the requirements of RE programs on OCS. The FERC also has
a significant role in integrating the RE resources for generation of electricity and
reduction of GHGs. Thus, the effective implementation of legislative actions and
regulatory approach is required to best see the success of RE programs.

DISCLAIMER

The material included in this chapter is as available from those references as


selected and the author’s interpretation only. For all regulatory and permitting
36 Renewable Energy Technologies and Water Infrastructure

verifications and confirmations on the information presented in this chapter,


the federal, state, and/or local regulators shall be contacted (as needed and
appropriate).

References
BLM (Bureau of Land Management). 2018. “Geothermal energy.” Accessed July 20,
2019. https://www.blm.gov/programs/energy-and-minerals/renewable-energy/
geothermal-energy.
BOEM (Bureau of Ocean Energy Management). 2019. “BOEM’s regulatory framework
and guidelines.” Accessed July 21, 2019. https://www.boem.gov/Regulatory-Framework.
30CFR. 2019. “30CFR, Part 585—Renewable energy and alternate uses of existing facilities
on the outer continental shelf.” Accessed July 20, 2019. https://www.ecfr.gov/cgi-bin/
text-idx?SID=e2b4f27d2273002068c5fcaa1409b2c7&mc=true&node=pt30.2.585&rgn
=div5#se30.2.585_1100.
43CFR. 2019. “43CFR, Part 2800—Rights-of-way under the Federal Land Policy and
Management Act.” Accessed July 20, 2019. https://www.ecfr.gov/cgi-bin/text-idx?SI
D=b5d0331dddb9fe6b2e07c47edfd1c8a8&mc=true&node=pt43.2.2800&rgn=div5
#se43.2.2801_15.
CT-PURA. 2019. “The State of Connecticut—Renewable Portfolio Standard.” Accessed
July 21, 2019. https://www.ct.gov/pura/cwp/view.asp?a=3354&q=415186.
DOI/BLM (US Department of the Interior, Bureau of Land Management). 2016. The
Federal Land Policy and Management Act of 1976, as amended. Office of Public Affairs.
Washington, DC: DOI/BLM.
FERC (Federal Energy Regulatory Commission). 2019. “Integration of Renewables.”
Accessed July 22, 2019. https://ferc.gov/industries/electric/indus-act/integration-renew.
asp.
Green, L., and C. Crume. 2017. Renewables Portfolio standard eligibility guidebook. 9th
ed. Publication No. CEC-300-2016-006-ED9-CMFREV. Sacramento, CA: California
Energy Commission.
IEA (International Energy Agency). 2019. “Renewables.” Accessed July 21, 2019. https://
www.iea.org/topics/renewables/.
NCSL (National Conference of State Legislatures). 2019. Accessed June 5, 2019. http://
www.ncsl.org/research/energy/renewable-portfolio-standards.aspx.
NREL (National Renewable Energy Laboratory). 2016. U.S. laws and regulations for
renewable energy grid interconnections. Rep. No. NREL/TP-6A20-66724. Golden, CO:
NREL.
42USC. 2019a. “Title 42 US Code—The Public Health and Welfare, Chapter 125—
Renewable Energy and Energy Efficiency Technology Competitiveness.” Accessed July
21, 2019. https://uscode.house.gov/browse/prelim@title42/chapter125&edition=prelim.
42USC. 2019b. “Title 42 US Code—The Public Health and Welfare, Chapter 134—Energy
Policy, Subchapter V—Renewable Energy; and Chapter 149—National Energy Policy and
Programs, Subchapter II—Renewable Energy.” Accessed July 21, 2019. https://uscode.house.
gov/browse/prelim@title42/chapter134/subchapter5&edition=prelim; https://uscode.
house.gov/browse/prelim@title42/chapter149/subchapter2/partA&edition=prelim.
TETHYS. 2019. “Environmental Effects of Wind and Marine Renewable Energy.”
Accessed September 11, 2021. https://tethys.pnnl.gov/ https://tethys.pnnl.gov/
regulatory-frameworks-marine-renewable-energy
Renewables and Regulatory Requirements of the United States 37

US-Congress. 2005. “ENERGY POLICY ACT OF 2005; PUBLIC LAW 109–58—AUG.


8, 2005.” Accessed September 11, 2021. https://www.congress.gov/109/plaws/publ58/
PLAW-109publ58.pdf
USEIA (US Energy Information Agency). 2019. “US Renewable Energy Production and
Consumption by Source.” Accessed July 21, 2019. https://www.eia.gov/totalenergy/data/
browser/?tbl=T10.01#/?f=M&start=200001.
VA-RE. 2019. “The State of Virginia—RE model ordinances.” Accessed July 21, 2019.
https://www.deq.virginia.gov/Programs/RenewableEnergy/ModelOrdinances.aspx.
CHAPTER 3
Biofuels: Ethanol and
Biodiesel
William F. Ritter

ETHANOL

Introduction

Ethanol was first used in 1826 to power an engine. In 1876, Nicolaus Otto, the
inventor of the modern four-cycle internal combustion engine, used ethanol to
power an early engine. In 1908, Henry Ford used ethanol to power his Model T.
The first use of ethanol blended with gasoline as an octane booster was developed
in the 1920s and 1930s and was in high demand during World War II because of
fuel shortage (Gustafson 2019).
Modern-day ethanol industry began in the 1970s when petroleum-based fuel
became expensive and environmental concerns involving leaded gasoline created
a need for an octane fuel. Corn became the predominant feedstock for ethanol
production because of its abundance and ease of transformation into alcohol.
Federal and state subsidies for ethanol helped keep the fuel in production when
ethanol prices fell with crude oil and gasoline prices in the early 1980s. Ethanol’s
use as an oxygenate to control carbon monoxide emissions encouraged increased
production of the fuel through the decade and in the 1990s.
With the phasing out of methyl tertiary butyl ether (MTBE) as an oxygenate
and a desire to decrease dependence on imported oil and increase the use of
environmentally friendly fuels, ethanol’s demand increased dramatically. In 2005,
the first renewable fuel standard (RFS) became law as part of the United States’
energy policy (DOE 2019). The law allowed for ethanol production of 15.1 billion
L/year (4 billion gal./year) in 2006 and was later amended to increase production
to 28.4 billion L/year (7.5 billion gal./year) by 2012 (DOE 2019). Figure 3-1
shows global ethanol production by country or region from 2007 to 2017. Global
production peaked in 2017 after a dip in 2011 and 2012. The United States is the

39
40 Renewable Energy Technologies and Water Infrastructure

Figure 3-1. Global ethanol production 2007–2017.


Source: DOE (2019).

world’s largest producer of ethanol, having produced over 60 billion L/year (16
billion gal./year) in 2017 alone. Together, the United States and Brazil produce
85% of the world’s ethanol. The vast majority of US ethanol is produced from corn,
whereas Brazil primarily uses sugarcane (DOE 2019).

Legislation
The Energy Policy Act was first passed by Congress in 1992 (PL 102-486).
The Energy Policy Act was revised in 2005 (PL 109-58). It addressed energy
production in the United States including (1) energy efficiency, (2) renewable
energy, (3) oil and gas, (4) coal, (5) tribal energy, (6) nuclear matters and
security, (7) vehicles and motor fuels, (8) hydrogen, (9) electricity, (10) energy tax
incentives, (11) hydropower and geothermal energy, and (12) chemical charge
technology. One of the provisions of the 2005 Act was to increase the amount of
biofuel that must be mixed with gasoline sold in the United States (DOE 2019).
The Energy Independence and Security Act (EISA) (PL 110-140) was passed
and signed by President George Bush on December 19, 2007 (EPA 2019). The
goal of the act was to help the United States achieve greater energy independence
and security and increase the production of clean renewable fuels. The three key
provisions of the act were the corporate average fuel economy standard (CAFES),
the RFS, and the appliance/lighting sufficiency standard.
The RFS was originally under the Energy Policy Act of 2005 which amended
the Clean Air Act (CAA). The EISA of 2007 further amended the CAA by
expanding the RFS program. The RFS is implemented by EPA in consultation
with the US Department of Agriculture and Department of Energy. The RFS
program requires a certain volume of renewable fuel to replace or reduce gasoline
volume. The categories under the RFS are

• Biomass-based diesel
• Cellulose biofuel,
Biofuels: Ethanol and Biodiesel 41

Table 3-1. Volume renewable standards set by EISA.

Cellulose Biomass-based Advanced Total


Year fuel diesel (billion gal.) biofuel biofuel
2009 NA 0.5 0.6 11.1
2010 0.1 0.65 0.95 12.95
2011 0.25 0.8 1.35 13.95
2012 0.5 1.0 2.0 15.2
2013 1.0 1.0 2.75 16.55
2014 1.75 1.0 3.75 18.15
2015 3.0 1.0 5.5 20.5
2016 4.25 1.0 7.25 22.25
2017 5.5 1.0 9.0 24.0
2018 7.0 1.0 11.0 26.0
2019 8.5 1.0 13.0 28.0
2020 10.5 1.0 15.0 30.0
2021 13.5 1.0 13.0 33.0
2022 16.0 1.0 21.0 36.0
Source: EPA (2019).

• Advance biofuel, and


• Total renewable fuels.
The 2007 EISA has set long-term goal of boosting the renewable fuel
production to 136 billion L (36 billion gal.). It extended the yearly volume
requirements out to 2022. The yearly renewable fuel requirements under the EISA
are listed in Table 3-1 (EPA 2019).

Classes of Ethanol
The three general categories of ethanol–gasoline blends are E10, E15, and E85. E10
is gasoline with 10% ethanol content. E15 is gasoline with 15% ethanol content,
and E85 is a fuel that may contain up to 85% ethanol. In the United States, the
ethanol content of most of the motor gasoline sold is 10% by volume. In June
2012, EPA approved E15 for use in flex-fuel vehicles and light-duty trucks, SUVs,
and cars manufactured since 2001. The next month, a Kansas gas station became
the nation’s first to offer E15. Despite EPA’s certification, E15 was denigrated by
many automakers, warning that it could damage engines. Also, distributing the
fuel often required station modifications, that could cost thousands of dollars.
In most US markets, Reid vapor pressure (RVP) volatility restrictions
currently prevent the sale of E15 to flex-fuel vehicles from June 1 to September
15, meaning most vehicles cannot purchase the ethanol blend during the busiest
driving period of the year. Ethanol advocates want this restriction lifted and see
it as the key to E15’s growth. There are over 1,800 stations in 31 states selling E15
blend ethanol (Growth Energy 2019). Most of them are in the Midwest where most
ethanol production capacity is located. There has been an ongoing discussion with
42 Renewable Energy Technologies and Water Infrastructure

EPA about getting the restriction lifted. After over a decade of hard work and
coordination by renewable fuel organizations such as Growth Energy, leading
fuel retailers, congressional champions, rural advocates, and other key industry
stakeholders, on May 31, 2019, EPA took action to remove the regulation and allow
access to E15 year-round. Growth Energy predictions suggest that it will increase
employment in the ethanol industry by 136,000 jobs and create a market for 26.5
billion L (7.0 billion gal.) more ethanol and 810,000 ha (2.0 million acre) of corn
(Growth Energy 2019).
The energy content of ethanol is about 33% less than pure gasoline. The
impact of fuel ethanol on vehicle fuel economy varies depending on the amount
of methanol denaturant that is added to ethanol. The energy content of the
methanol denaturant is almost equal to the energy content of pure gasoline. In
general, vehicle fuel economy may decrease by about 3% when using E10 relative
to gasoline that does not contain fuel ethanol.

Processing of Corn Ethanol


The process of making ethanol from corn is a multistep process. The first step is
milling the corn. Either dry milling or wet milling is used. Figures 3-2 and 3-3
outline the process steps for wet milling and dry milling, respectively. Today most
of the ethanol plants in the United States use dry milling.
In wet milling, the corn kernels are broken down into starch, fiber, corn germ,
and protein by heating in sulfurous acid solution for 2 days. The starch is separated
and can produce ethanol, corn syrup, or food-grade starch. Additional products
that are produced include corn oil, gluten meal, and gluten feed (Dutton 2019).
Dry milling is a simpler process. The main products are ethanol, CO2, and
dried distillers’ grain. The five steps of dry milling are (1) grinding, (2) cooking

Figure 3-2. Wet milling process for corn ethanol.


Source: RFA (2019).
Biofuels: Ethanol and Biodiesel 43

Figure 3-3. Dry milling process for corn ethanol.


Source: RFA (2019).

and liquefication, (3) saccharification, (4) fermentation, and (5) distillation.


A hammer mill or roller is used for grinding. After the corn is broken down,
it is mixed with heated water to form a mash or slurry. The slurry then goes
through cooking and liquefication. Water interacts with the starch granules when
the temperature is more than 60°C. The liquefication step is partial hydrolysis
that breaks down the longer starch chains into smaller chains. The next step is
saccharification, which is further hydrolyzed to glucose. The final step to make
ethanol from starch is fermentation in which yeast is added. In the fermentation
process, one mole of glucose yields two moles of ethanol and two moles of carbon
dioxide (Dutton 2019).
The final phase in ethanol production is distillation. After fermentation, the
concentration of ethanol is 12% to 15% in water. Distillation is used to increase
the ethanol concentration to 95%. The remaining 5% water is removed from the
ethanol by dehydration. A molecular sieve containing zeolite is used to remove
the water.

Cellulosic Ethanol Processing


Cellulosic ethanol is made from biomass. Processing could be anything from corn
stover, recycled newspaper to trees. Farmers can grow energy crops for cellulosic
ethanol, including switch grass and certain trees.
44 Renewable Energy Technologies and Water Infrastructure

Figure 3-4. Cellulose ethanol flow diagram.


Source: DOE (2019).

Cellulosic ethanol can be made either biochemically or thermochemically


(Dutton 2019). The process for biochemical ethanol is shown in Figure 3-4. In
biochemical production, the biomass is first ground into bits. The bits are treated
in hot sulfuric acid where the cellular walls and content dissolve. The acid pushes
lignin out of the way to form hemicelluloses. The hemicelluloses then decompose
into the four sugars, namely, xylose, mannose, arabinose, and galactose.
The second step is called cellulose hydrolysis. The acid is washed off and the
mixture goes to a tank with enzymes called cellulases, which turns cellulose into
glucose. The third step is fermentation where the glucose and four hemicellulose
sugars are converted to ethanol. The sugar concentration and the microbes used
depend on the plant species used in the beginning.
The final step is separation. Everything that is not alcohol settles to the
bottom of the tank is sent for processing and reuse. The alcohol remains on the
top and is sent to distillation to separate fuel-grade ethanol.

US Ethanol Production
The United States has approximately 200 ethanol plants with a total capacity of 59.7
billion L/year (15.8 billion gal./year) (Figures 3-5 and 3-6). Nearly, all of the plants
are located in the Midwest, with Iowa having the largest capacity at 15.9 billion L/
year (4.2 billion gal./year) followed by Nebraska with a capacity of 8.7 billion L/
year (2.3 billion gal./year) and Illinois with a capacity of 7.2 billion L (1.9 billion
gal./year) (DOE 2019). Today, the United States produces more ethanol than it
consumes. With the decrease in fuel consumption, the E10 blend wall has been
reached, which limits the amount of ethanol consumed. When Congress revised
and dramatically expanded the size and scope of the RFS in December 2007, it
established annual mandates to increase biofuel consumption from less than 18.9
billion L/year (5 billion gal./year) in 2007 to 136 billion L/year (36 billion gal./year)
Biofuels: Ethanol and Biodiesel 45

Figure 3-5. US ethanol plants and capacity 1999–2017.


Source: DOE (2019).

Figure 3-6. US ethanol production and consumption 2000–2018.


Source: DOE (2019).

in 2022. Had everything gone according to the schedule set by Congress, by 2022,
according to the mandate, 79.4 billion L/year (21 billion gal./year) would be filled
by so-called advanced biofuels. The remaining 56.7 billion L/year (15 billion gal./
year) was implicitly reserved for corn ethanol. There has been very little about
the RFS that has gone according to plan. Congress assumed that motor vehicle
fuel would continue to keep rising. The bulk of biofuel today was to have been
supplied by cellulosic ethanol, which was to account for the bulk of the advanced
fuel mandate.
There was no commercial cellulosic ethanol production when the
mandates were established, but proponents of the technology were certain that
commercialization would come in response to the mandates. The cellulosic ethanol
mandate went into effect in 2010, when 378 million L/year (100 million gal./year)
of cellulosic ethanol was required to be blended into the fuel supply (EPA 2019).
46 Renewable Energy Technologies and Water Infrastructure

The mandate quickly ramped up to 1.89 billion L/year (0.5 billion gal./year) in
2012, 3.78 billion L/year (1.0 billion gal./year) in 2013, and in 2017 was supposed to
reach 20.8 billion L/year (5.5 billion gal./year). In reality, no commercial cellulosic
ethanol was produced in 2010 or 2011, but in 2012, the first qualifying batch of
cellulosic ethanol was produced. Blue Sugars Corporation produced some 75,860 L
(20,069 gal.) of cellulosic biofuel in April 2012. Following this, no further cellulosic
ethanol was produced in 2012 or 2013, and Blue Sugars declared bankruptcy a year
later. In 2014, several new plants came online. For the most part, these plants were
heavily subsidized by taxpayers, and every gallon of qualifying production also
received subsidies in the form of renewable energy credits. Companies that built
plants to produce cellulosic ethanol included DuPont, Abengoa, INEOS Bio, and
the privately owned POET. Most of these plants have also now gone out of business,
but they did manage to contribute to the production of 2,753,764 L (728,509 gal.)
of cellulosic ethanol in 2014 (Rapier 2018).
The CAA requires EPA to set annual RFS volumes of biofuels that must
be used for transportation for the total, cellulose and biomass, advanced, and
biodiesel biofuel categories. Since 2010, EPA has been adjusting the volumes below
statutory targets because of market realities. For 2019, the volumes EPA set for the
cellulose, biodiesel, advanced, and total biofuel categories are 1.59, 7.94, 18.59, and
75.29 billion L (0.42, 2.10, 4.92, and 19.92 billion gal.), respectively. For 2020, the
volumes for cellulose, biodiesel, advanced, and total biofuel categories are 2.04,
9.19, 19.05, and 75.75 billion L (0.54, 2.43, 5.04, and 20.04 billion gal.), respectively
(EPA 2020).

Greenhouse Gases
In 2010, EPA released a lifecycle analysis of the greenhouse gas (GHG) emissions
for corn ethanol (EPA 2010). They concluded that by 2022, corn ethanol GHG
emissions from a new refinery would be 21% lower than that of an energy
equivalent of gasoline. Over the years, this 21% value has dominated policy
discussion and Federal regulations related to corn ethanol.
The GHG profile of corn ethanol has been controversial. Searchinger et al.
(2008) concluded that GHG emissions associated with its production and
combustion exceeded the emissions associated with producing and combusting
an equivalent quantity of gasoline. They concluded that farmers brought new land
into production for corn and reduced the production of other crops on existing
land to meet the corn ethanol demand. The land-use changes (LUC) resulted in
corn ethanol having a higher GHG profile than gasoline. RFS under the 2007
Energy Independence and Security Act of 2007 mandated EPA to do a full GHG
lifecycle analysis of corn ethanol and to include both direct and significant
indirect sources of emissions. The EPA indirect sources included LUC.
The Congressional Budget Office (CBO 2014) summarizes several studies on
the issue of GHG emissions from corn ethanol. One of the studies by the National
Research Council challenged the EPA figure of about 20% reduction in GHG
emissions from corn ethanol over gasoline saying that there are several scenarios
in which GHG emissions from corn ethanol are much higher than those from
Biofuels: Ethanol and Biodiesel 47

petroleum-based fuels. The CBO concluded that switching to corn ethanol over
gasoline offered only limited potential to reduce overall GHG emissions.
In a recent study using updated data, Flugge et al. (2017) concluded that the
current GHG emissions profile for corn ethanol was 39% to 43% lower than that
for gasoline. Unlike other studies on GHG benefits, which relied on forecasts of
future ethanol production systems and expected impacts on the farm sector, this
study reviewed how the industry and farm sectors performed in the last decade to
assess the current GHG profile of corn-based ethanol. This report found greater
lifecycle GHG benefits from corn ethanol than a number of previous studies,
driven by a variety of improvements in ethanol production, from the cornfield
to the ethanol refinery. Farmers are producing corn more efficiently and using
conservation practices that reduce GHG emissions, including reduced tillage,
cover crops and improved nitrogen management. Corn yields are also improving.
Between 2005 and 2015, US corn yields increased by more than 10%. Between
2005 and 2015, ethanol production in the United States also increased significantly
from 14.7 billion L/year (3.9 billion gal./year) to 55.9 billion L/year (14.8 billion
gal./year). At the same time, advances in ethanol production technologies, such as
using combined heat and power, using landfill gas for energy, and co-producing
biodiesel, helped reduce GHG emissions at ethanol refinery plants.
They also projected two scenarios for corn ethanol in 2022 in which the GHG
emissions are 47% to 70% lower than those for gasoline (Figure 3-7). Figure 3-7
compares the full lifecycle analysis of corn ethanol GHG emissions for the 2014
current conditions, 2022 business as usual (BAU) for corn ethanol production,

Figure 3-7. Full lifecycle corn ethanol GHG emissions for 2014 current conditions,
20022 BAU and 2022 BBS.
Source: Flugge et al. (2017).
48 Renewable Energy Technologies and Water Infrastructure

and 2022 building block scenarios (BBS). The 70% reduction scenario assumes
that refineries contract with farmers to grow corn with low-emission practices
such as reduced tillage, winter cover crops, targeting nitrogen (N) fertilizer
application rates, and using N inhibitors to slow down nitrification rates.
Several reasons are there to find greater lifecycle GHG benefits from corn
ethanol than some previous studies. Previous estimates anticipated that growing
corn to produce ethanol would result in “indirect land-use change.” In other
words, the land would be converted from grasslands and forests to commodity
production as a result of increased demand for corn used in ethanol production.
However, based on new data and research, there is compelling evidence that,
although LUC have occurred, the actual patterns of changes and innovation
within the farm sector have resulted in these indirect emissions being much lower
than previously projected.

Water Quality Impacts


The Mississippi River Basin (MRB) encompasses more than 55% of the US
agricultural land (Goolsby and Battaglin 2000) and more than 75% of the corn,
cotton, rice, sorghum, wheat, and forage area. In 2015, the total value of all crop
production in the conterminous United States was about $184 billion, with field
crops accounting for $143.4 billion, commercial vegetable crops $13.4 billion, and
fruit and nuts $27.1 billion (USDA 2017). The value of field and miscellaneous crop
production (not including commercial vegetables, fruits, or nuts) within the MRB
in 2015 was estimated at $131.6 billion, which represents more than 90% of the
field and miscellaneous crop production value in the United States.
Row crop agriculture in the Midwest has had an impact on water quality in
the Gulf of Mexico over the years. A hypoxic zone in the northern Gulf of Mexico
has been forming each summer since the 1950s. Measurements of the hypoxic
zone started in 1985 (MTF 2008). The Gulf’s hypoxic zone is caused by excessive
nutrient pollution in MRB. The excess nutrients stimulate algae growth, which
sinks and decomposes and lowers oxygen levels below a point that cannot support
most marine life. Every year NOAA predicts the size of the hypoxic zone based
on USGS data. The major factor contributing to the size of the hypoxic zone every
summer is the USGS May nitrogen and phosphorus nutrient load data.
The Mississippi River/Gulf of Mexico Watershed Nutrient Task Force
(MTF) was established in the fall of 1997 to understand the causes and effects
of eutrophication in the Gulf of Mexico, coordinate activities to reduce the size
and duration, and ameliorate the effects of hypoxia. In 2008, the MTF released
the 2008 Action Plan (MTF 2008), which called for the MTF to complete and
implement comprehensive nitrogen and phosphorus reduction strategies for the
states within the MRB. Each state was to develop strategies for reducing nitrogen
and phosphorus loads by 20% and 45%, respectively.
In Illinois, extensive analyses conducted by researchers at the University
of Illinois estimated that point sources and agricultural nonpoint sources
contributed 48% of the total phosphorus (TP) reaching the Mississippi River from
that state. Agriculture was the source of 80% of the nitrate-nitrogen; point sources
Biofuels: Ethanol and Biodiesel 49

contributed about 18%. Urban runoff contributed 4% of the TP and 2% of the


nitrate-nitrogen. The tile-drained areas of central and northern Illinois are the
largest source of nitrate. Sloping, erosive soils in western and southern Illinois are
the largest contributor of nonpoint TP (ILEPA 2014).
The Iowa Department of Agriculture and Land Stewardship, the Iowa
Department of Natural Resources (Iowa DNR), and Iowa State University
developed a science and technology-based framework to assess and reduce
nutrients to Iowa waters and the Gulf of Mexico (IOWA 2019). On an annual
basis, most nutrient loads in Iowa come from nonpoint sources. It is estimated
that 83% of the nitrogen load to Iowa streams comes from nonpoint sources and
7% from point sources and 79% of the phosphorus load comes from nonpoint
sources and 21% from point sources. Annual row crop production, coupled with
usually abundant rainfall, facilitates the vast majority of nitrogen transport to
streams in Iowa with a large majority discharged as nitrates through tile drainage.
The sources of phosphorus include agricultural nonpoint source runoff and
streambank erosion.
As part of Minnesota’s nutrient reduction strategy, the state conducted a
comprehensive science assessment that incorporated nutrient conditions, trends,
sources, and pathways. The nutrient source assessment was based on multiple
Minnesota Pollution Control Agency (MPCA) studies and engaged numerous
local, state, and federal partners. During an average precipitation year, cropland
sources contribute an estimated 78% of the nitrogen load to the Mississippi River
in Minnesota. Cropland nitrogen reaches surface waters through two dominant
pathways: tile drainage transport; and leaching to groundwater and subsequent
flow to surface waters. The primary sources of phosphorus transported to streams
are cropland runoff, permitted wastewater, and streambank erosion (MPCA 2014).
In the major corn-growing states such as Iowa, Illinois, and Minnesota,
the major sources of nitrogen and phosphorus are from agriculture. All three
of these states have extensive tile drainage which transports a large percentage
of the nitrates to streams. The major sources of phosphorus and nitrogen in the
MRB as predicted by the USGS SPARROW model are shown in Figures 3-8 and
3-9 (MTF 2017). The major sources of nitrogen are from fertilizers, whereas the
major sources of phosphorus are from fertilizers and manure.
The annual total nitrogen (TN), nitrate, and TP loads from the MRB to the
Gulf of Mexico from 1980 to 2017 are shown in Figures 3-10, 3-11, and 3-12 (TP is
shown up to 2015), respectively (USGS 2017). Annual TN loads have been below
the average annual baseline load from 1980 to 1996 for most years from 1997 to
2017. There has been no decrease in the annual TP loads.
Best management practices (BMPs) for nitrogen and phosphorus may be
classified as source management or transport management (Sharpley et al.
2001). Some of the source management BMPs that will reduce both nitrogen and
phosphorus losses are

• Application nutrient rates that meet crop needs;


• Method of application;
50 Renewable Energy Technologies and Water Infrastructure

Figure 3-8. Major sources of phosphorus in the MRB.


Source: MTF (2017).

Figure 3-9. Major sources of nitrogen in the MRB.


Source: MTF (2017).
Biofuels: Ethanol and Biodiesel 51

Figure 3-10. Annual total nitrogen loads from the MRB to the Gulf of Mexico 1980–2017.
Source: USGS (2017).

• Timing of application;
• Manure management such as composting, storage ponds, or lagoons; and
• Adding alum as a manure additive.
Some of the transport BMPs that will reduce both nitrogen and phosphorus
losses are
• Winter cover crops,
• Conservation tillage will reduce TN and TP losses but may increase nitrate
losses,
• Grass or forest buffers at the edge of fields next to streams,

Figure 3-11. Annual nitrate plus nitrite loads from the MRB to the Gulf of Mexico
1989–2017.
Source: USGS (2017).
52 Renewable Energy Technologies and Water Infrastructure

Figure 3-12. Annual total phosphorus loads from the MRB to the Gulf of Mexico
1980–2015.
Source: USGS (2017).

• Wetland areas and grassed waterways,


• Terraces,
• Strip cropping,
• Critical source treatment areas, and
• Sediment delivery structures.
Since the late 1990s, the development of methods to reduce losses of nitrogen in
drainage waters has become a primary objective in addressing the environmental
impacts of agricultural drainage for researchers and engineers. Reducing nitrogen
losses is difficult because the nitrate form is mobile in soil solution and may be
readily leached with subsurface drainage water. A number of methods may be
used to reduce losses. They include source reduction by fertilizing at appropriate
rates and times, cover crops, routing drainage water through wetlands, use of
biofilters, saturated buffers, reduced drainage intensity, and drainage water table
management, also known as controlled drainage (CD). CD can reduce nitrate
loads at an average of 30% in the Midwest, although this can range from 15% to
75% (Christianson et al. 2016).

Pros and Cons of Ethanol


Savin (2019) of Alternative Energies has summarized some of the pros and cons
of corn ethanol. Some of the pros of ethanol are
• Ethanol is a biofuel that can lower the level of GHG emissions released by the
transportation sector.
• Ethanol can be produced from cheap raw materials such as corn and sugarcane.
Biofuels: Ethanol and Biodiesel 53

• Ethanol is an energy-balanced fuel that is produced from corn and generates


1.06 units of energy for any 1 unit of energy used.
• Production of ethanol generates useful by-products.
• Ethanol generates lower GHG emissions than fossil fuels and is also
biodegradable.
Some of the cons of ethanol are
• Ethanol is less effective than gasoline.
• Crops used to produce ethanol occupy a large surface of land.
• In the United States, 40% of the corn crop production goes toward ethanol
production instead of being used as food.
• It is an alternate fuel that is heavily subsidized

BIODIESEL

Introduction
The diesel engine was developed in the 1890s by inventor Rudolph Diesel. Today,
the diesel engine has become the engine of choice for power, reliability, and high
fuel economy worldwide. Early experiments on vegetable oil fuels were conducted
by the French government and Dr. Diesel himself. Dr. Diesel envisioned that pure
vegetable oils could power early diesel engines for agriculture in remote areas of
the world, where petroleum was not available at the time. The early diesel engines
were designed to run on many different fuels, from kerosene to coal dust. The first
public demonstration of vegetable oil–based diesel fuel was at the 1900 World’s
Fair, when the French government commissioned the Otto company to build a
diesel engine to run on peanut oil (Pacific Biodiesel 2019).
Shortly after Dr. Diesel’s death in 1913, petroleum became widely available
in a variety of forms, including the class of fuel we know today as “diesel fuel.”
With petroleum being available and cheap, the diesel engine design was changed
to match the properties of petroleum diesel fuel. Owing to the widespread
availability and low cost of petroleum diesel fuel, vegetable oil–based fuels gained
little attention, except in times of high oil prices and shortages. World War II
and the oil crises of the 1970s saw a brief interest in using vegetable oils to fuel
diesel engines. Unfortunately, the newer diesel engine designs could not run on
traditional vegetable oils, because of the much higher viscosity of vegetable oil
compared with that of petroleum diesel fuel. It was a Belgian inventor in 1937,
who first proposed using transesterification to convert vegetable oils into fatty
acid alkyl esters and use them as a diesel fuel replacement. The transesterification
reaction is the basis for the production of modern biodiesel.
Pioneering work in Europe and South Africa by researchers, such as Martin
Mittelbach, furthered the development of the biodiesel fuel industry in the early
54 Renewable Energy Technologies and Water Infrastructure

1990s, with the US industry coming on more slowly, because of lower prices for
petroleum diesel. Pacific Biodiesel became one of the first biodiesel plants in the
United States in 1996 to establish a biodiesel production operation to recycle used
cooking oil into biodiesel on the island Maui in Hawaii. The biodiesel industry
became a household name in the United States, after the terrorist attacks of
9/11/2001 resulted in historically high oil prices and increased awareness of energy
security. By 2005, worldwide biodiesel production had reached 4.2 billion L/year
(1.1 billion gal./year) with most fuel being produced in the European Union.

Biodiesel Processing
Biodiesel can be made from nearly any feedstock that contains free fatty acids,
which are the raw materials that are converted to biodiesel through a process
called transesterification. Most biodiesel in the United States is produced from
vegetable oils. Other feedstocks (raw materials) include waste animal fats from
processing plants and recycled used cooking oil and grease from restaurants. A
flow diagram of the biodiesel production process is shown in Figure 3-13.
In the transesterification process,-fats and oils are converted into biodiesel
and glycerol (Figure 3-14) (Van Gerpen et al. 2004). Fats and oils are reacted with
a short-chain alcohol in the presence of a catalyst, producing fatty acid esters that
are primely the molecules in biodiesel. Methanol is usually the alcohol used in
the process, and, in general, either sodium hydroxide or potassium hydroxide
is used as the catalyst. The methanol is recovered in the process and reused.
Approximately 45.4 kg (100 lb) of oil or fat is reacted with 4.5 kg (10 lb) of a short-
chain alcohol to form 45.4 kg (100 lb) of biodiesel and 4.5 kg (10 lb) of glycerol
(EIA 2020).

Figure 3-13. Biodiesel processing flow diagram.


Source: DOE (2020).
Biofuels: Ethanol and Biodiesel 55

Figure 3-14. Biodiesel transesterification reaction.


Source: Van Gerpen et al. (2004).

The glycerol produced in biodiesel production contains numerous impurities.


There are various outlets for disposal and utilization of the crude glycerol
generated in biodiesel plants. Many large-scale biodiesel producers refine the
glycerol into a pure form and then it can be used in the food, pharmaceutical,
or cosmetics industries. For small-scale producers, however, purification is too
expensive to be performed in their manufacturing sites. Their crude glycerol is
usually sold to large refineries for upgradation. In recent years, however, with the
rapid expansion of the biodiesel industry, the market is flooded with excessive
crude glycerol; thus, producers need to seek new value-added uses for this glycerol
or dispose it. Combustion is the common method for glycerol disposal (EIA 2020).

Classes of Biodiesel Blends


Biodiesel can be blended and used in many different concentrations. The most
common are B5 (up to 5% biodiesel) and B20 (6% to 20% biodiesel). B100 (pure
biodiesel) is typically used as a blend stock to produce lower blends and is rarely used
as a transportation fuel. B20 is the most common biodiesel blend. It represents a good
balance of cost, emission, cold-weather performance, material compatibility, and
ability to act as a solvent. Engines operating on B20 have similar fuel consumption,
horsepower, and torque to engines running on petroleum diesel. Pure biodiesel
contains less energy on a volumetric basis than petroleum diesel. High-level biodiesel
blends can also impact engine warranties, can gel in cold temperatures, and may
present unique storage issues. B100 use could also increase nitrogen oxide emissions,
although it greatly reduces other toxic emissions (EIA 2020).

US Biodiesel Production
Biodiesel production and exports and consumption from 2001 to 2019 are shown
in Figure 3-15 (DOE 2020). The biodiesel diesel industry has grown since 2010,
with commercial production facilities from coast to coast. The industry reached a
56 Renewable Energy Technologies and Water Infrastructure

Figure 3-15. US biodiesel production, exports, and consumption.


Source: DOE (2020).

key milestone in 2011, when it produced 3.78 billion L/year (1.0 billion gal./year)
for the first time. By 2015, the market had doubled to more than 7.6 billion L/
year (2.0 billion gal./year). In 2019, the market was 6.8 billion L/year (1.8 billion
gal./year). The industry’s total production continues to significantly exceed the
biodiesel requirement under the Federal RFS and has been sufficient to fill a
majority of the advanced biofuel requirement.
The various feedstocks used in the United States to produce biodiesel in 2019 are
shown in Figure 3-16. Soybean oil is by far the largest feedstock used (57%). Animal
fats account for 8% of the feedstocks used. A total of (approximately) 5.79 billion
kg (12.75 billion lb) of feedstock was used to produce biodiesel in 2019 (EIA 2020).

Figure 3-16. Feedstocks used for biodiesel production in 2019.


Source: EIA (2020).
Biofuels: Ethanol and Biodiesel 57

For the year 2020, under the RFS, the final volume requirements for cellulosic
biofuel, biodiesel or BBD, advanced biofuel (in total), and renewable fuel (in
total) categories are: 2.23, 9.20, 19.27, and 76.04 billion L [0.59, 2.43 (for the year
2021, as well), 5.09, and 20.09 billion gal.], respectively (note that all values are
ethanol-equivalent on an energy content basis, except for BBD which is biodiesel-
equivalent) (CRS 2020, EPA 2020).

SUMMARY

The Energy Independence and Security Act (EISA) (PL 110-140) was passed
in 2007 (EPA 2019). The goal of the act was to help the United States achieve
greater energy independence and security and increase the production of clean
renewable fuels. By 2022, the 2007 EISA set boosting the long-term goal to 136
billion L (36 billion gal.) of renewable fuel. Had everything gone according to
the schedule set by Congress, by 2022 according to the mandate, 21 billion gal./
year would be filled by so-called advanced biofuels. The remaining 15 billion
gal./year was implicitly reserved for corn ethanol. Very little about the RFS has
gone according to the plan; today, no significant level of cellulosic ethanol has
been produced.
Biodiesel can be made from nearly any feedstock that contains free fatty acids,
which are the raw materials that are converted to biodiesel through a process
called transesterification. Most biodiesel in the United States is produced from
vegetable oils. In 2019, the market for biodiesel in the United States was 6.8 billion
L/year (1.8 billion gal./year). In 2019, the industry used a total of 5.79 billion kg
(12.75 billion lb) of feedstock to produce biodiesel (EIA 2020). Soybean oil (57%)
was the largest feedstock used to produce biodiesel.

References
CBO (Congressional Budget Office). 2014. The renewable fuel standard issue for 2014 and
beyond. Washington, DC: CBO.
Christianson, L. E., J. Frankenberger, C. Hay, M. J. Helmers, and G. Sands. 2016. Ten ways
to reduce nitrogen loads from drained cropland in the Midwest. Pub. C1400. Urbana-
Champaign, IL: Univ. of Illinois Extension.
CRS (Congressional Research Service). 2020. The Renewable Fuel Standard (RFS): An
overview (updated April 14, 2020); the CRS Report #R43325. Accessed December 13,
2020. https://crsreports.congress.gov/product/pdf/R/R43325.
DOE (US Department of Energy). 2020. “Biodiesel production and distribution.”
Alternative Fuels Data Center, DOE. Accessed June 26, 2020. https://afdc.energy.gov/
fuels/biodiesel_production.html and https://afdc.energy.gov/laws/RFS.
DOE. 2019. “Maps and data.” Alternative Fuels Data Center, DOE. Accessed July 26, 2019.
https://afdc.energy.gov/data/.
Dutton, J. A. 2019. Alternate fuels from biomass. EGEE 439 e-Education Institute. University
Park, PA: Pennsylvania State Univ. Accessed June 5, 2019. https://www.e-education.psu.
edu/egee439/node/673.
58 Renewable Energy Technologies and Water Infrastructure

EIA (US Energy Information Administration). 2020. “Biofuels explained biomass-based


diesel facts.” Accessed June 26, 2020. https://www.eia.gov/energyexplained/biofuels/
biodiesel.php.
EPA (US Environmental Protection Agency). 2010. “Renewable fuel standard
program regulatory impact analysis.” EPA-420-R-10-006. Accessed June 26, 2019.
https://nepis.epa.gov/EPA/html/DLwait.htm?url=/Exe/ZyPDF.cgi/P1006DXP.
PDF?Dockey=P1006DXP.PDF.
EPA. 2019. “Renewable fuels program statues.” Accessed June 25, 2019. https://www.epa.
gov/renewable-fuel-standard-program/statutes-renewable-fuel-standard-program.
EPA. 2020. “Final renewable fuel standards for 2020, and the biomass-based diesel volume for
2021.” Accessed June 11, 2020. https://www.epa.gov/renewable-fuel-standard-program/
final-renewable-fuel-standards-2020-and-biomass-based-diesel-volume.
Flugge, M., J. Lewandrowski, J. Rosenfeld, C. Boland, T. Hendrickson, K. Jaglo, et al.
2017. A life-cycle analysis of the greenhouse gas emissions of corn-based ethanol. Report
prepared by ICF under USDA Contract No. AG-3142-D-16-0243. Accessed July 26, 2019.
https://www.usda.gov/media/press-releases/2017/01/12/usda-releases-new-report-life​
cycle-greenhouse-gas-balance-ethanol.
Goolsby, D. A., and W. A. Battaglin. 2000. Nitrogen in the Mississippi basin-estimating
sources and predicting flux to the Gulf of Mexico. US Geological Survey Fact Sheet 135-
00. Accessed July 25, 2019. http://ks.water.usgs.gov/Kansas/pubs/fact-sheets/fs.135-00.
html.
Growth Energy. 2019. Expanding access to biofuels with year-round E15. Washington,
DC: Growth Energy. Accessed July 6, 2019. https://growthenergy.org/policy-priorities/
yearrounde15/.
Gustafson, C. 2019. History of ethanol production and policy. NDSU Extension Fact
Sheet. Fargo, ND. Accessed June 5, 2019. https://www.ag.ndsu.edu/energy/biofuels/
energy-briefs/history-of-ethanol-production-and-policy.
ILEPA (State of Illinois EPA). 2014. “Illinois nutrient loss reduction strategy.” Accessed
July 25, 2019. http://www.epa.illinois.gov/topics/water-quality/watershed-management/
excess-nutrients/index.
IOWA (State of Iowa). 2019. “Iowa nutrient reduction strategy 2017–18 annual report.”
Iowa Dept. of Agriculture and Land Stewardship, Iowa Dept. of Natural Resources, and
Iowa State Univ. College of Agriculture and Life Sciences. Accessed June 12, 2019. http://
www.nutrientstrategy.iastate.edu>documentsNRS2018annualreportdocs.
Mississippi River/Gulf of Mexico Watershed Nutrient Task Force (MTF). 2008. “Gulf
hypoxiation plan 2008 for reducing, mitigating, and controlling hypoxia in the
northern Gulf of Mexico and improving water quality in the Mississippi River
basin.” Washington, DC: US Environmental Protection Agency, Office of Wetlands,
Oceans, and Watersheds, Mississippi River/Gulf of Mexico Watershed Nutrient Task
Force. Accessed July 25, 2019. http://water.epa.gov/type/watersheds/named/msbasin/
bupload/2008_8_28_msbasin_ghap2008_update082608.pdf.
Mississippi River/Gulf of Mexico Watershed Nutrient Task Force (MTF). 2017. “Mississippi
River/Gulf of Mexico watershed nutrient task force 2017 report to Congress.” 2nd Biennial
Report. Washington, DC: US Environmental Protection Agency, Office of Wetlands,
Oceans and Watersheds, Mississippi River/Gulf of Mexico Watershed Nutrient Task Force.
Accessed July 25, 2019. https://www.epa.gov/ms-htf/hypoxia-task-force-reports-congress.
MPCA (Minnesota Pollution Control Agency). 2014. The Minnesota nutrient reduction
strategy. St. Paul, MN: Minnesota Pollution Control Agency. Accessed July 25, 2019.
http://www.pca.state.mn.us/index.php/view-document.html?gid=20213.
Biofuels: Ethanol and Biodiesel 59

Pacific Biodiesel. 2019. “The history of biodiesel fuel.” Accessed June 10, 2020. https://
www.biodiesel.com/history-of-biodiesel-fuel/.
Rapier, R. 2018. “Cellulosic ethanol falling far short of the hype.” Forbes. Accessed June
26, 2019. https://www.forbes.com/sites/rrapier/2018/02/11/cellu ​losic-etha ​nol-fall​
ing-far-short-of-the-hype/#5975fee3505f.
RFA (Renewable Fuels Association). 2019. How is ethanol made. Washington, DC: RFA.
Accessed June 25, 2019. https://ethanolrfa.org/how-ethanol-is-made/.
Savin, M. 2019. “Pros and cons of ethanol”. Alternative Energies. Accessed June 19, 2019.
https://www.alternative-energies.net/what-is-ethanol-pros-and-cons/.
Searchinger, T., R. Heimlich, R. A. Houghton, F. Dong, A. Elobeid, J. Fabiosa, et al. 2008.
“Use of cropland for biofuels increases greenhouse gases through emissions from land-
use change.” Science 319 (#5867): 1238–1240.
Sharpley, A. N., R. W. McDowell, J. L. Weld, and P. J. A. Kleinman. 2001. “Assessing site
vulnerability to phosphorus loss in an agricultural watershed.” J. Environ. Qual. 30:
2026–2036.
USDA (US Department of Agriculture). 2017. “Crop values 2017 summary, February 2018.”
USDA National Agricultural Statistics Service (USDA NASS). Accessed August 2, 2019,
and December 13, 2020. http://usda.mannlib.cornell.edu/usda/current/CropValuSu/
CropValuSu-02-24-2017.pdf and https://www.nass.usda.gov/Publications/Todays_
Reports/reports/cpvl0218.pdf.
USGS. 2017. “Watershed loadings for the Mississippi River and subbasins.” US Dept. of
Interior, USGS. Accessed August 3, 2019, and December 13, 2020. https://nrtwq.usgs.
gov/mississippi_loads/#/ and https://nrtwq.usgs.gov/nwqn/#/.
Van Gerpen, J., B. Shanke, R. Pruszko, D. Clements, and G. Knothe. 2004. Biodiesel
production technology August 2002 to January 2004. Rep. No. NREL/SR-510-362-44.
Golden, CO: National Research Energy Laboratory. Accessed December 13, 2020, and
June 8, 2020. https://www.nrel.gov/docs/fy04osti/36244.pdf and https://www.nrel.
gov>doc>R-510-36244.pdf.
CHAPTER 4
Micro-Hydropower:
Concept, System Design, and
Innovations
Tamim Younos, Juneseok Lee

INTRODUCTION

Hydropower can simply be defined as “flowing water’s energy” that performs a


useful task. Ancient civilizations used hydropower in the form of waterwheel to
grind wheat into flour (Nunez 2019). A well-known hydropower device, which
dates back to the eighteenth century, is the hydraulic ram—a cyclic water pump
that takes advantage of the water hammer effect (Rayner 1995). The pump uses
a unidirectional valve to create a perpetual cycle of water flow in which a small
portion of the inflow water is pumped via a delivery pipe, whereas excess water
is expelled through a waste valve (Figure 4-1). Hydro rams are still popular for
providing water for multiple uses in small communities and remote areas and
other uses, such as operating a water fountain.
The theme of this chapter is using “flowing water’s energy” to generate
electricity. The history of hydropower development for electricity generation in
the United States dates back to the late nineteenth century (Nunez 2019, EERE
2019a). In 1881, using direct current (DC) technology, a dynamo connected to a
turbine in a flour mill provided street lighting at Niagara Falls, New York, and in
1893, the first commercial installation of an alternating current (AC) hydropower
plant at the Redlands Power Plant in California allowed power to be transmitted
longer distances for consumer use (EERE 2019a).
At present, hydropower is the most widely used renewable energy source and
represents about 17% of the total electricity production in the world (IEA 2018).
China is the largest producer of hydroelectricity, followed by Canada, Brazil,
and the United States. Hydropower accounts for nearly 9% of the US electric
generating capacity (USGS 2018).

61
62 Renewable Energy Technologies and Water Infrastructure

Figure 4-1. Basic components of a hydraulic ram: (1) inlet (drive pipe), (2) free flow
at waste valve, (3) outlet (delivery pipe), (4) waste valve, (5) delivery check valve, and
(6) pressure vessel.
Source: Wikipedia https://en.wikipedia.org/wiki/Hydraulic_ram (available at public domain).

Table 4-1. Hydropower Plant Size and Capacity.

Plant size Power-generation capacity


Large hydropower >30 MW
Small hydropower 100 kW–30 MW
Micro-hydropower Up to 100 kW
Source: EERE (2019b).

The most common and traditional type of a hydropower plant uses a dam on
a river that stores water in a reservoir. The water is released from the reservoir—
creating total hydraulic/energy head (i.e., a sum of pressure, elevation, and
velocity head)—and supplied to a turbine—a rotary engine that converts moving
water to mechanical energy—which, in turn, activates a generator—a device that
converts mechanical energy to electrical energy. However, a dam is not necessarily
required to generate power. Electricity can be generated from any stream flow or
pipe flow with enough hydraulic head that can turn a (suitable) turbine. A system
can be put in place with as little as 0.61 m (2 ft) of head with high flow or as little as
0.008 m3/min (2 gal./min) of flow with high head (Alternative Energy News 2018).
Table 4-1 shows a hydropower plant size based on its power generation
capacity, which, in turn, depends on the available water flow rate and hydraulic
head (EERE 2019b). The focus of this chapter is on micro-hydropower generation
(up to 100 kW).

MICRO-HYDROPOWER GENERATION: CONCEPT

A micro-hydropower plant is basically a small-scale decentralized energy generation


infrastructure. In terms of electricity use, a micro-hydropower plant can be configured
in two ways: (1) integrated into the conventional electric grid and (2) a stand-alone
electricity source, when an electric grid is not available. Micro-hydropower is
particularly useful for meeting electricity demand in rural environments, small
farms, and remote communities. Micro-hydropower can also have specific urban
applications, such as generating power for street lighting and other uses.
Micro-Hydropower: Concept, System Design, and Innovations 63

Figure 4-2. Typical schematic of a small-scale hydropower generation system.


Source: DOE (2010).

A micro-hydropower generation plant location is site specific and depends


on the topographic characteristics of the site. For example, a section of the river/
stream that contains a natural waterfall can be a suitable site. A common practice
is to construct a diversion channel from the mainstream to create the hydraulic
head, that is, a vertical distance from the water source to a turbine.
Figure 4-2 shows major elements of the water conveyance system for a micro-
hydropower plant. These include a water intake weir, settling basin, constructed
channel, forebay tank or reservoir, and penstock. The penstock is a pipe that
transports water from the channel to a turbine located in the power house at the
base of a site.
Figure 4-3 shows the technical concept of a micro-hydropower system based
on field conditions. Two major field design factors are the available hydraulic head
(H) and the flow rate (Q).
Equation (4-1) can be used to estimate the potential available power (P)

P =ηρ g H Q (4-1)

where
η = Turbine efficiency,
ρ = Density of water (kg/m3),
g = Acceleration of gravity (m/s2),
H = Hydraulic head (m), and
Q = Flow rate (volume over time).
64 Renewable Energy Technologies and Water Infrastructure

Figure 4-3. The basic concept of a micro-hydropower generation system.


Source: DOE (2010).

MICRO-HYDROPOWER SYSTEM DESIGN

In this section, the system design for a micro-hydropower plant is illustrated by


a case study project. The goal of the case study project illustrated herein was to
design and install a modern micro-hydropower plant for educational purposes.

Case Study Site


The case study site is located in Glen Alton, a historic site located within the
Jefferson National Forest in Giles County, Virginia, where a micro-hydropower
plant was built in the 1940s (Younos 2013). The site, originally private property,
was donated to the Federal government and is maintained by the US Forest
Service as a sustainability education demonstration site.
In the 1940s, the property owner, Lucas, constructed a dam [Figure 4-4(a)] on
the North Fork Creek, a stream that ran through his property, and also a 271.3 m
(890 ft) channel that diverted water from the North Fork Creek [Figure 4-4(b)].

Figure 4-4. (a)The dam, (b) the channel, and (c) original powerhouse.
Source: Student team.
Micro-Hydropower: Concept, System Design, and Innovations 65

The channel supplied water to a turbine that generated electricity for use on his
property [Figure 4-4(c)]. The original power plant at Glen Alton was abandoned
when the property was connected to the electric grid.

Micro-Hydropower System Design Components


Figure 4-5 shows an overview of the case study site at Glen Alton. Red segments in
Figure 4-5 depict components of the new system design described in this section
(Table 4-2). Black segments depict project components, such as the dam and the
overflow creek that already existed.
The electrical component in the powerhouse does not show the battery bank
that was added later in the updated design.
The Glen Alton system is capable of producing 6 to 8 A (power outputs up
to 2 kW) when operating at the design H and Q. The 2 kW rating falls within
the lower range of a micro-hydropower generation capacity. However, the design
principles described herein apply to all power ranges. The generated electricity
is used to power lights and a stand-alone heating/cooling unit at the Glen Alton
lodge. Details of system design components are described subsequently.

Hydraulic Component Design


Hydraulic component design parameters include hydraulic head and flow rate at
the site, penstock design, and intake water filtration (Table 4-2).

Figure 4-5. Overview of the case study project at Glen Alton.


Source: Student team.
66 Renewable Energy Technologies and Water Infrastructure

Table 4-2. System Design Components for a Modern Micro-Hydropower Plant.

Hydraulic component Mechanical component Electrical component


Field conditions: Turbine Electrical generator
hydraulic head (H) (alternator)
and flow rate (Q)
Penstock design Turbine flow control Inverter
Penstock intake filter Turbine–generator Controller
interface
Penstock flow control — Battery bank
Source: Younos (2013).

Hydraulic Head
Standard surveying procedures were applied to determine the design hydraulic
head (H), that is, a change in the elevation (vertical distance) between the water
source channel (forebay tank) and the turbine to be located in the power house.
Figure 4-6 shows the constructed channel (2% slope) and the measured hydraulic
head (H) 6.4 m (20.93 ft) at the case study site.

Flow Rate
A temporary dam is constructed to measure the volumetric flow rate of a small
stream. At the case study site, water discharge at the temporary dam near the
penstock intake was collected to fill a container of known volume [0.0038 m3
(5 gal.)] and the filling time was recorded. Using this method, time to fill known
volume (i.e., flow rate = volume/time to fill), the estimated design flow rate (Q)
was calculated at 0.95 m3/min (250 gpm).

Flow Control to Penstock


It is critical that water supply to penstock meet the design flow criteria and remain
at a constant flow. A 1.22 m (48 in.) deep weir, which spans 3.66 m (12 ft) across

Figure 4-6. Hydraulic head at the Glen Alton micro-hydropower project site (1 ft =
0.3048 m).
Source: Younos (2013).
Micro-Hydropower: Concept, System Design, and Innovations 67

Figure 4-7. Flow control weir at the case study site.


Source: Student team.

the channel, is constructed immediately upstream from the penstock intake


(Figure 4-7). Weir outflow is connected via an underground pipe to the penstock
intake, where the filtration unit is installed (discussed later).
Preliminary investigation detected that segments of the old channel were
leaking. Dolomite was used to fill the leaking segments of the channel.

Penstock Design
Other than hydraulic head and flow control, critical penstock design parameters
include the following: (1) pipe material, which determines the pipe durability and
cost; and (2) penstock pipe length and diameter, which determines the head loss
and flow capacity to the turbine.
Pipe Material
Three decision factors to determine penstock material are durability, pressure
rating, and cost. Polyvinyl chloride (PVC) pipes and high-density polyethylene
(HDPE) pipes are considered suitable pipe materials for the penstock. HDPE pipe
is typically less expensive than PVC, but its pressure rating is lower than PVC.
As shown in Table 4-3, based on a professional engineering judgment, a weight
was assigned to each decision factor to create a selection matrix (durability 35%,
pressure rating 35%, and cost 30%). It was determined that except for the cost,
the normalized score for PVC was higher than that for HDPE. Thus, PVC was
selected as the penstock pipe material. The required pipe length was 9.9 m (32.6 ft)
(Figure 4-5).
68 Renewable Energy Technologies and Water Infrastructure

Table 4-3. Selection Matrix for the Penstock Pipe Material.

PVC HDPE
Weight Normalized Weighted Normalized Weighted
Decision factor (%) score score score score
Durability 35 0.78 27.22 0.56 19.44
Pressure rating 35 0.73 25.67 0.33 11.67
Cost 30 0.60 18.00 0.73 22.00
Total 70.89 53.11
Source: Younos (2013).

Pipe Diameter
As a general rule, a larger pipe diameter yields a lower flow velocity and lower
head loss. Although the larger penstock pipe diameter reduces head loss, it may
result in a higher cost, and, therefore, it is a critical decision factor. Based on
friction loss for schedule 40 PVC pipe [class of 160 psi (4,206 kPa) PVC pipe], it
was determined that an 8 in. (20.3 cm) diameter (inner) PVC pipe was the optimal
diameter pipe for the penstock. For the 8 in. (20.3 cm) pipe diameter and 0.95 m3/
min (250 gpm) flow rate, the estimated head loss was 0.12 ft per 100 ft of the pipe
length (0.12 m per 100 m).

Penstock Intake Filter


The turbine and generator should be protected from objects, such as fish, leaves,
twigs, and rocks. A filter is needed at the penstock water intake to remove these
objects. After a very thorough investigation and evaluation of various filtration
methods, a Coanda effect hydro-shear screen was purchased from Hydroscreen
LLC, Denver, Colorado (Figure 4-8), for the case study site. This type of screen is
relatively new to small-hydropower projects but has had great success in large-
scale applications. The filter efficiently removes objects from flowing water using
a “wedge-wire” design where water flow is accelerated over a wedge-wire screen,
which shears the flow and forces the debris down the inclined screen (Figure 4-8).
The filtration unit is self-cleaning and requires minimal maintenance.

Mechanical Component Design


As shown, in Table 4-2, mechanical components include the turbine, flow control
devices, and turbine–generator interface (connector).
A water turbine is a rotary engine that converts moving water to rotational
mechanical energy—the water pushes a series of blades mounted on a shaft.
The shaft is electromagnetically coupled to an electric generator (alternator).
The turbine selection depends on hydraulic head (H) and design flow (Q) at the
site. After preliminary investigation, the following turbine types available in
the market were found as potential choices: the Wild Nature Solutions LV1500
turbine (Missouri Wind and Solar, Seymour, Missouri), the energy systems and
Micro-Hydropower: Concept, System Design, and Innovations 69

Figure 4-8. Schematic showing operation of a typical Coanda effect screen unit.
Source: US Bureau of Reclamation (2003).

design’s stream-engine turbine, and a custom-built cross-flow turbine. The option


of custom-building a cross-flow turbine involves a significant effort of planning
and construction and, therefore, was not considered.
For micro-hydropower plants, impulse turbines (e.g., Pelton wheel and the
Turgo wheel) are the preferred turbine type (impulse turbines rely on the velocity
of water to turn the turbine wheel (called the runner). The Turgo impulse wheel,
an upgraded version of the Pelton wheel, can move twice as fast (higher efficiency)
and is known to require less maintenance. The Wild Nature Solutions LV1500
turbine requires more maintenance and costs higher than Energy Systems and
Design’s (Sussex, New Brunswick, Canada) Stream Engine turbine. Therefore,
an LV1500 four (4) nozzle high-precision bronze Turgo wheel turbine (Stream-
Engine turbine) (Energy Systems and Design) was selected for the case study site
(Figure 4-9). Other components needed to be compatible with the selected turbine
(Stream Engine Manual 2017).
Turbine Flow Control
The water supply nozzle diameter through which water is supplied from the
penstock to the turbine is an important design factor. Appropriate nozzle diameter
selection is based on the head and flow rate and the turbine type. For the case
study site conditions using the Turgo turbine, four nozzles of 0.022 m (0.875 in.)
diameter (at 925 rpm) are used. The tube connecting the penstock outlet to the
turbine intake is also a critical factor as it impacts head loss and power output.
Flexible braided vinyl tubing (smooth and minimized sharp bends) was selected
to connect the penstock pipe outlet to the nozzle inserts of the turbine. A gate
valve that controls the amount of water supplied to the turbine is installed in each
of the four water lines that connect the penstock to the turbine [Figures 4-9(b)
and 4-10].
70 Renewable Energy Technologies and Water Infrastructure

Figure 4-9. (a) Top view of a turbine-generator system, and (b) LV1500 four-nozzle
turbine.
Source: Energy Systems & Design, n.d.; reproduced with permission.

Figure 4-10. Turbine flow control.


Source: Energy Systems & Design, n.d.; reproduced with permission.

Turbine–Generator Interface
The generator interface is an electromagnetic shaft that connects turbine to
the electric generator (alternator). Critical design parameters for selecting the
turbine–generator connector include connection mode, efficiency, and cost.
Table 4-4 shows the decision matrix for the interface selection process, which
was processed in consultation with electrical experts. Direct connection was
determined as the most appropriate option.
Micro-Hydropower: Concept, System Design, and Innovations 71

Table 4-4. Decision Matrix for Turbine–Generator Interface.


Direct connection Belt connection Rope connection

Decision Weight Normalized Weighted Normalized Weighted Normalized Weighted


factor (%) score score score score score score

Cost 40 0.96 38.4 0.71 28.40 0.82 32.80


Efficiency 60 0.96 57.75 0.685 41.02 0.72 43.20
Total 96.15 69.42 76.0

Source: Younos (2013).

Turbine Protection
To meet the design-head (H) requirement, it is desirable to install the turbine very
close to the surface of North Fork Creek that flows in the base of the power house
(Figure 4-5). It was noted that several times during the year, the North Fork Creek
water level rises up to 1.5 m (5 ft) above the average and causes base-flooding.
Therefore, it is necessary to protect the turbine generator from flooding. After
considering various options, it was determined that a high-walled watertight
caisson turbine housing would best provide the desired protection. Based on cost
and other factors, such as material weight and endurance, a watertight plastic
caisson was selected to house the turbine.

Electrical Component Design


Figure 4-11 shows the electrical components of a modern grid-tied and battery-
based micro-hydropower generation system similar to the case study site (Table
4-2). Details of the electrical components design are discussed in the following.

Electric Generator
A generator (alternator) with an electromagnet converts mechanical energy into
electricity. Based on consultation with experts, a 48VDC (volt direct-current)
turbine–generator system was selected for the case study project (Figure 4-12).
The system is capable of producing between 6 and 8 A (power outputs up to 2 kW)
when running at the designed conditions.

Battery
Monitor

Charge
Turbine Battery Bank Dump Lead
Controller

DC AC Breaker To Household
Inverter
Disconnect Panel Loads

Figure 4-11. Electric components of a modern micro-hydropower generation and


distribution system.
Source: Chapter Author.
72 Renewable Energy Technologies and Water Infrastructure

Figure 4-12. Magnet electric generator directly connected to the Turgo turbine wheel.
Source: Energy Systems & Design, n.d.; reproduced with permission.

Inverter
The inverter converts 48VDC output from the electric generator to standard 120 V
AC electricity for consumer use. The AC load includes a wire connection to the
Glen Alton lodge and to the electric grid. The electric power company provided
a meter that captures electricity delivered and electricity received. This feature
allows the Glen Alton site to send excess power to the electric grid (for credit) or
receive electricity from the grid in situations when micro-hydropower generation
is minimal. Furthermore, the inverter prevents electricity transmission to and
from the electric grid during power outages.
For a utility-interactive electricity generation plant, such as Glen Alton site,
there is a requirement that the inverter be UL1741 compliant. UL1741 refers
to the Standard for Interconnecting Distributed Energy Resources with an
electric power (grid) system. Based on consultation with hydropower experts
and reviewing various vendors, an Outback GTFX3048 DC to AC inverter was
selected for this system (Figure 4-13). In addition, an Outback surge protector
was used to monitor and manage the system’s power output and load fluctuations
and to protect equipment from grid electrical surges and minimize the risk of
electricity being back fed to the system.

Controller
The electrical loads on the system are dependent on the electricity demand. The
controller automatically adjusts the load so that the generator always turns at
exactly the right speed and constantly monitors voltage or frequency. A dump-load
controller is used to dissipate generated electricity that cannot be used or stored.
The dump-load controller installed at the site is an HL-100 Air Heater Dump-Load
controller (Wholesale Solar, Mt. Shasta, California) (Figure 4-14). The stand-alone
heating/cooling unit at the lodge serves as a temperature regulation unit.
Micro-Hydropower: Concept, System Design, and Innovations 73

Figure 4-13. (a) The outback GTFX3048 inverter and (b) surge protector/monitor.
Source: Wholesale SOLAR, n.d.; reproduced with permission.

Figure 4-14. HL-100 air heater dump-load controller.


Source: altEstore, n.d.; reproduced with permission.

In addition, several AC and DC disconnect panels are necessary to properly


isolate the system during maintenance or equipment failure. The electric company
requires a fused, labeled, and lockable AC disconnect panel at the pole where the
meter is installed to isolate the micro-hydropower system from the grid, if needed.
In addition, a DC disconnect panel is installed between the battery bank and the
turbine generator, which will allow the turbine generator to be isolated during
yearly maintenance.
74 Renewable Energy Technologies and Water Infrastructure

Battery Bank
As was noted, the micro-hydropower system at Glen Alton is both grid-tied and
battery based. The electrical output from the generator is connected to a battery
bank that allows for generated electricity to be stored in situations where there is
no AC load at the lodge or electric grid system (Figure 4-15).
At the Glen Alton site, four Universal UB4D 12 V, 200 Ah sealed absorbed
glass mat (AGM) batteries, a variant, and an advanced design of sealed valve-
regulated lead acid batteries (VRLA) were used (Figure 4-16). The system creates
a 48 V battery bank with a storage capacity of 200 Amp-hrs. The design storage
capacity of the battery bank allows the Glen Alton lodge to be powered for 2 days
when there is no electricity generation from the turbine generator.

Battery Overcharge Controller


This device detects when the battery bank is fully charged and prevents overcharging
of the battery bank (shunt load). A MorningStar Tri-45 controller (Wholesale Solar,
Mt. Shasta, California) was recommended to perform this function (Figure 4-17).
In addition, a remote temperature sensor (RTS) is connected to the overcharge

Figure 4-15. Simplified schematic of a battery-based micro-hydropower system.


Source: Energy Systems & Design Manual; reproduced with permission.

Figure 4-16. Universal UB4D 12 V, 200AH sealed AGM battery.


Source: altEstore, n.d.; reproduced with permission.
Micro-Hydropower: Concept, System Design, and Innovations 75

Figure 4-17. MorningStar Tri-45 battery overcharge controller.


Source: Wholesale SOLAR, n.d.; reproduced with permission.

controller and battery bank. This component is used to improve battery charging
in systems that experience temperature variations throughout the year.

Case Study Project Cost


A micro-hydropower plant’s capital cost depends on the plant size and site
conditions. Table 4-5 shows a summary of cost items for the Glen Alton micro-
hydro project. Detailed costs are documented in the Appendixes A and B.

Cost–Benefit Analysis
A brief outline of the cost–benefit analysis for the Glen Alton case study is presented
in the following. As shown in Table 4-5, the initial incurred cost/salvage value for
the project was $9,900. This is the total project cost because the labor cost is $0 as it
was provided by the in-kind support and volunteer work. The Glen Alton system is
capable of producing power outputs up to 2 kW. With the estimated system overall
efficiency of approximately 60%, the system will continuously produce about 1.2 kW.
The yearly operation and maintenance (O&M) was estimated to be $100 to replace
bearings in the turbine every year and other potential expenses. The system was
tied to the electric grid. The electricity (cost) saving ($) can be based on the system

Table 4-5. Micro-Hydropower Cost Summary


at the Glen Alton Site.

Cost item Cost ($)


Equipment (shipping) 9,471 (450)
Supplies and material 420
Construction* —
Total cost 9,900
Source: Younos (2013).
*Construction and labor cost was provided as
in-kind support by the US Forest Service.
76 Renewable Energy Technologies and Water Infrastructure

providing 1,200 W of electricity (credit). The average electricity cost in the United
States is about 0.13 $/kWh. Assuming that the electricity cost will increase 15% every
year, the projected payback period for the case study project would be 15 to 20 years.
In this brief cost–benefit estimation, only the benefits of electricity capture
are considered for quantifying the economic value. Note that this type of micro-
hydropower can be a source of personal pride for the owners/community. Those
who are environmentally aware and motivated can spend a substantial sum of
dollars on installing these types of energy generation equipment as they believe
in the cause of renewable energy and environmental issues and, even though,
financially the costs do not cover the benefits. These types of “warm glow” benefits
or emotional reward are hard to quantify in monetary values and cannot be
included in the analysis. Also, the energy costs in the United States are expected to
increase and the relatively larger-scale micro-hydropower (i.e., within <100 kW)
will contribute to a shorter payback period.

Regulatory Requirements—Case Study Site


Regulatory and permitting requirements for renewable energy systems including
micro-hydropower vary in the United States according to federal, state, and local
regulations.
In the Commonwealth of Virginia, where the case study was performed,
several local and state regulations must be considered for designing a grid-tied
system. These regulations are stated in the Virginia Net Metering Program
Application (VA-DEQ 2015). In the Application, it states that “the program is
designed for customers desiring to install qualifying renewable energy generators
and interconnect them with their electric utility. It is intended for those that are
generating electricity to offset all or part of their own energy requirements and is
not applicable for those desiring to generate power for resale.” Clearly, to qualify
for this program, at least a part of the generated electricity must be used at Glen
Alton property. In addition, the Application requires that the system’s DC to AC
inverter be UL1741 compliant so that it shuts itself down when there is a power
outage. Appalachian Electric Power (AEP), the electricity provider at the case
study site, requires a signature from the inverter vendor to confirm that this safety
standard is met. To meet the aforementioned requirements, the on-site lodge was
designated as the primary use point of the electricity generated at the micro-
power plant. The generated electricity was used to power lights and provide power
to a stand-alone heating/cooling unit at the Glen Alton lodge.

 ICRO-HYDROPOWER: VIABLE TECHNOLOGY


M
IN DEVELOPING COUNTRIES

Micro-hydropower plants are considered a viable technology for providing


clean and a lower cost-based electricity to small and rural communities in the
developing countries for use at domestic as well as agriculture and small industrial
Micro-Hydropower: Concept, System Design, and Innovations 77

applications. Significant ongoing research and development efforts on micro-


hydropower development are in progress in several Asian and African countries.
An overview of a few selected studies that discuss the potential and obstacles of
implementing micro-hydropower projects in developing countries is presented
in the following.
Smith (2001) in a book titled “Motors as Generators for Micro-Hydropower”
emphasized the increasing role and importance of developing micro-hydropower
projects in developing countries and argued that the conventional generators are
difficult to obtain or are unreliable in service. Motors, on the contrary, are always
available locally. Manufacturers and rural development engineers were encouraged
to select a motor and convert it for use as a generator in micro-hydro projects.
The book provides details on the developed technology and its applications as a
collaborative effort between the International Technology Development Group in
the UK and experts in Sri Lanka, Nepal, and Indonesia.
Michael and Jawahar (2017) discussed the history of hydropower use in India
since 1897 and the potential for developing micro-hydropower plant projects
in India. The authors included a discussion on the design of a 15 kW micro-
hydropower plant for rural electrification at Valara in the State of Kerala, India,
where electricity is not available for 120 tribal families. The authors presented
their work on preliminary studies carried out at Valara for the development of a
micro-hydropower plant focused on main topics, such as technical and economic
feasibility studies, design of civil works, selection of electrical and mechanical
components, and financial analysis. They concluded that the site had enormous
potential for developing a micro-hydropower plant, and the proposed 15 kW
micro-hydropower plant was found to be technically and economically viable
and met the energy demands of about 120 local tribal families. The operation
and maintenance of the plant would be managed by the local people or authority.
Upadhayay (2009) evaluated the effectiveness of micro-hydropower projects
in Nepal. According to the Nepal Electricity Authority (2009), 80% of the Nepalese
population remains without electricity. Citing other studies, the author noted that
despite the potential to generate a total of 50 MW utilizing micro-hydropower
plants, Nepal produces only about 6 MW from small-scale hydropower projects.
The author investigated the effectiveness of community-based micro-hydro
projects and the role of public participation in two remote villages in Nepal. The
topics of investigation included the following: (1) opportunities and levels of
decision-making, (2) the degree of local ownership perceived and/or achieved,
(3) satisfaction with the process of public participation, (4) achievement in terms
of long-term goals, (5) the diversity of participants (ethnicity, gender, or age),
and (6) the benefits and challenges of public participation. The author used a
case study methodology with data collection in the form of interviews, surveys,
and document reviews. The study concluded that “as seen by both the micro-
hydro project promoters and the Nepalese government, these projects appear
to be successful because they performed well, when rated by the criteria and
metrics that have been used historically to evaluate these projects (e.g., high
rates of villager participation). However, a closer look at these reveals that many
78 Renewable Energy Technologies and Water Infrastructure

important development issues remain unaddressed. In particular, large amounts


of social and financial investment have been levied into a development project
that cannot meet either current or future demand for electricity. The results of
this study show that the projects examined are, at best, a temporary solution for
these villagers’ electricity needs.”
The European Union Energy Initiative (EUEI) (2010), in collaboration with
other governmental and nongovernmental entities, published a report on policy
and regulatory framework conditions for small hydro power in Sub-Saharan Africa.
By the United Nations Development Programme (UNDP) definition, Sub-Saharan
Africa comprises 45 countries. The report, citing other studies, states that in Sub-
Saharan Africa, about 560 million people (74%) live without access to electricity.
The EUEI report outlines the potential of micro-hydropower in terms of existing
and planned projects in Sub-Saharan Africa and identifies the following main
barriers: (1) a lack of policy and regulatory framework; (2) inadequate financing;
(3) a lack of capacity to plan, build, and operate micro-hydropower plants; and (4)
a lack of long-term hydrological data, especially seasonal and long-term river flow
variations. The report concludes with recommendations on how the gap between
the existing top-down regulations and the regulatory needs (Barriers 1 to 3) of
micro-hydropower projects can be overcome so as to deploy micro-hydropower
projects in Sub-Saharan Africa on a larger scale. Major recommendations include
the following: (1) owing to the small-scale character of micro-hydropower projects,
the micro-hydropower sector needs to incorporate regulations, capacities, and
financing schemes at the local level; and (2) a guidebook should be developed for the
integrated development of micro-hydropower policy and regulation. The authors
recommend the incorporation of the following subjects in such a guidebook:
(a) the incorporation of micro-hydropower development in national policies
and sector strategies; (b) the establishment of a database (existing projects and
pipeline, potential projects, funding options, operators, and service and equipment
providers); (c) an analysis of existing regulatory framework and adjustments to
promote micro-hydropower development; (d) smart subsidies and other incentive
schemes; (e) strategies of how to involve the private sector; (f) best practices for
operating schemes; and (g) capacity building for the involved institutions.

 MALL-STREAM MICRO-HYDROPOWER: CHALLENGES AND


S
LIMITATIONS

Small-stream micro-hydropower projects are site specific, the system’s capacity


is limited by stream flow and topographic conditions, and the system design
is mostly based on site conditions rather than local power needs. Unlike other
renewable energy technologies (e.g., solar and wind), micro-hydropower plants
cannot be expanded after construction. However, similar to other renewable
energy technologies, advances in battery (energy) storage technologies can
address the problem of possible seasonal variations in stream flow that limit
power generation and use.
Micro-Hydropower: Concept, System Design, and Innovations 79

The cost for a micro-hydropower plant is site specific and varies depending
on the plant size and site condition. The cost-effectiveness of micro-hydropower
can be a limiting factor. However, limited information is available on cost-
effectiveness of micro-hydropower projects. The International Renewable Energy
Agency (IRENA 2012) has published guidelines for hydropower cost analysis that
can be applied to micro-hydropower projects at specific locations.
As stated previously, regulatory requirements for micro-hydropower plants
are country specific; therefore, planners and designers must consider local and
national regulatory requirements pertinent to a project location.

EMERGING MICRO- AND SMALL-HYDROPOWER TECHNOLOGIES

There are several emerging technologies that can harness electricity to take
advantage of the water and energy nexus. Unlike small-stream micro-hydropower
projects that have fixed and/or limited capacity owing to their geographic location,
these emerging, futuristic, and decentralized micro-hydropower technologies can
be expanded to meet the high energy demand in the populated areas.
New technologies taking advantage of the water and energy nexus include
power generation from tidal and wave energy (Maryland Clean Energy Center
2017) and energy harnessed from pressurized (550 to 690 kPa [80 to 100 psi])
drinking water distribution lines (Casini 2015). For example, tidal turbine
projects in Scotland and South Korea have tidal turbines with a 1.5 MW
electricity generation capacity (EIA 2018). High-pressure flows in the pipelines
managed by the San Diego Water County Authority, San Diego, allows the
generation of electricity through a 4.5 MW turbine generator (Viccione et al.
2018). Portland, Oregon, generates electricity from turbines installed in city
water pipes and using the so-called LucidPipes that generate an average of
1,100 MWh of electricity per year, which is sufficient to power about 150 homes
(CITYLAB 2018). The LucidPipe Power System in Portland (Figure 4-18) uses
the flow of water inside the Portland Water Bureau pipeline to spin four 42 in.
(1.06 m) diameter turbines that produce electricity for Portland General Electric
(LucidEnergy 2019).

Figure 4-18. Illustration of the Lucid energy pipe system installed inside a drinking
water line in Southeast Portland.
Source: CITYLAB (2018) (available in the public domain).
80 Renewable Energy Technologies and Water Infrastructure

Similar to the aforementioned examples, there has been increasing attention


to a new technology called pump as turbine (PAT) for small-scale power generation
(Binama et al. 2017; García et al. 2018, 2019; Corcoran et al. 2013; Patelis et al.
2016). Similar to turbines, PAT converts kinetic and pressure head of water into
mechanical energy/electrical power when water flows in the reverse direction.
Most researchers regard PAT technology as a positive step toward sustainable vital
practices within water distribution systems. Some studies also suggest using PAT
instead of a pressure-reducing valve to harvest energy and reduce possible water
leaks (Corcoran et al. 2013; Patelis et al. 2016).

SUMMARY

This chapter focuses on micro-hydropower generation (up to 100 kW), a small-


scale decentralized renewable energy generation infrastructure. Topics discussed
in this chapter include the following: (1) the basic design components of a micro-
hydropower generation system based on an illustrative example of design application
at a case study project in Virginia; (2) example applications of micro-hydropower
generation as a viable and sustainable technology in developing countries; and (3)
examples of emerging and futuristic small-scale power-generation technologies.
The chapter concludes that conventional small-stream micro-hydropower
generation technologies are mature technologies. Similar to other renewable
energy sources, advances in micro-hydropower technologies facilitate their
feasibility as a decentralized and clean energy source that can meet at the least a
partial electricity demand particularly in rural and small communities. Clearly,
hydropower has a positive environmental impact in terms of reducing dependency
on the fossil-fuel–based electric grid. Unlike large dam hydropower systems
that are disruptive to ecosystems, the impact of micro-hydropower systems on
the environment is lesser. Thus, where feasible, micro-hydropower should be
considered as a renewable component of a hybrid energy distribution system.
The potential success in the implementation of micro-hydropower projects
at the local level is affected by several factors and this should be considered at the
planning stages of a project. Unlike other renewable energy technologies (e.g.,
solar and wind), micro-hydropower plants cannot be expanded after construction.
However, similar to other renewable energy technologies, advances in battery
technologies for energy storage can address the problem of possible seasonal
variations in stream flows that limit power generation and use. The cost of a
micro-hydropower plant varies depending on the plant size and site conditions.
However, a limited information is available on the cost-effectiveness of micro-
hydropower projects. Regulatory requirements for micro-hydropower plants are
also country specific. Therefore, planners and designers must consider local, state,
and national (or federal) regulatory requirements.
Advances in emerging and futuristic decentralized micro-hydropower
technologies, such as harnessing energy from pressurized water pipes, present
significant global opportunities to expand the micro-hydropower system in both
rural and urban settings.
APPENDIX A. BILL OF MATERIALS (EQUIPMENT, 2013 PRICES)

Cost per Shipping Method


quantity cost Date of Date of Delivery Funding of
Equipment Product description Quantity Preferred vendor (USD) (USD) order delivery location source payment Notes

4 Nozzle 48VDC output 1 Energy Systems & $2,795.00 $125.00 12/14/ 3/12/ USFS ARC Credit ES&D
stream stream engine Design, P.O. Box 2011 2012 Ranger card recommends
engine turbine 4557, Sussex, NB, Station that we build
turbine generator; Canada E4E 5LT 110 a housing
generator comes with a Phone +1(506) South around the
universal nozzle 433-3151 Park system to
that can be cut Fax +1(506)433-6151 Drive, keep the
to 1/8 in. to 1 in. http://www. Blacksburg, system dry
DIA; Nozzle microhydropower. VA 24059 Need Federal
protrusions are com/our-products/ Tax
threaded 1 stream-engine ID# for order
to7/8 in. OD placement
Outback Model: GTFX3048 1 WholeSale Solar $1,776.00 — — 1/18/ USFS USFS USFS- UL 1741
grid-tie DC Item#: 2500348 412 N. Mt. Shasta 2012 Ranger Sheryl compliant
to AC Blvd., Station Lyles
inverter Mt. Shasta, CA 96067 110 South
TollFree: 800 472 Park
1142 Drive,
http://www. Blacksburg,
wholesalesolar. VA 24060
com/products.
folder/inverter-
Micro-Hydropower: Concept, System Design, and Innovations

folder/
outback%20
GTFX3048.html
81

(Continue)
82

Cost per Shipping Method


quantity cost Date of Date of Delivery Funding of
Equipment Product description Quantity Preferred vendor (USD) (USD) order delivery location source payment Notes

Outback MATE Model: MATE 1 WholeSale Solar $206.50 — — 1/18/ USFS USFS USFS- Allows for PC
display Item#: 9578062 412 N. Mt. Shasta 2012 Ranger Sheryl connection
Blvd., Station Lyles and display
Mt. Shasta, CA 96067 110 South of inverter
TollFree: 800 472 Park status
1142 Drive,
http://www. Blacksburg,
wholesalesolar. VA 24060
com/products.
folder/inverter-
folder/mate.html
Outback Item Number: 1 WholeSale Solar $175.00 — — 1/18/ USFS USFS USFS-
FX-SP-ACA 9552133 412 N. Mt. Shasta 2012 Ranger Sheryl
FLEXware Blvd., Station Lyles
surge Mt. Shasta, CA 96067 110 South
protector TollFree: 800 472 Park
1142 Drive,
http://www. Blacksburg,
wholesalesolar. VA 24060
com/products.
folder/inverter-
Renewable Energy Technologies and Water Infrastructure

folder/FW-SP-
ACA.htm
(Continue)
Cost per Shipping Method
quantity cost Date of Date of Delivery Funding of
Equipment Product description Quantity Preferred vendor (USD) (USD) order delivery location source payment Notes

MorningStar Model: TriStar-45 1 WholeSale Solar $145.00 $12.77 12/21/ 1/5/ USFS ARC Credit UL 1741
diversion Item#: 3680302 412 N. Mt. Shasta 2011 2012 Ranger card compliant
load rated for 45 A at Blvd., Station
controller 48 V Mt. Shasta, CA 96067 110 South
TollFree: 800 472 Park
1142 Drive,
http://www. Blacksburg,
wholesalesolar. VA 24060
com/products.
folder/controller-
folder/tristar-45.
html
Morningstar’s Model: TS-M-2 1 WholeSale Solar $87.00 $12.77 12/21/ 1/5/ USFS ARC Credit Digital meter
TriStar Item#: 3611115 412 N. Mt. Shasta 2011 2012 Ranger card interface for
Meter-2 Blvd., Station load
advanced Mt. Shasta, CA 96067 110 South controller
digital TollFree: 800 472 Park
meter 1142 Drive,
http://www. Blacksburg,
wholesalesolar. VA 24060
com/products.
folder/controller-
Micro-Hydropower: Concept, System Design, and Innovations

folder/TS-M-2.
html
(Continue)
83
84

Cost per Shipping Method


quantity cost Date of Date of Delivery Funding of
Equipment Product description Quantity Preferred vendor (USD) (USD) order delivery location source payment Notes

Remote Model: RTS 1 WholeSale Solar $30.00 $12.77 12/21/ 1/5/ USFS ARC Credit Measures the
temperature Item#: 3688102 412 N. Mt. Shasta 2011 2012 Ranger card temperature
sensor Blvd., Station of batteries
Mt. Shasta, CA 96067 110 South
TollFree: 800 472 Park
1142 Drive,
http://www. Blacksburg,
wholesalesolar. VA 24060
com/products.
folder/controller-
folder/
morningstar-rts.
html
Universal Model: Ub4D 12 V, 4 AltE $ 388.00 $105.26 1/12/ 2/2/ USFS ARC Credit https://www.
batteries 12 200 Ah 43 Broad Street, 2012 2012 Ranger card altestore.
V, 200 Ah Sealed Agm Batter Suite A408 Station com/store/
Hudson, MA 110 South Deep-cycle-
01749-2556 Park Batteries/
USA Drive, Batteries-
18778784060 Blacksburg, Sealed-Agm/
See website in notes VA 24060 Universal-
Ub4D-12V-
Renewable Energy Technologies and Water Infrastructure

200Ah-
Sealed-Agm-
Battery/
p2005/
(Continue)
Cost per Shipping Method
quantity cost Date of Date of Delivery Funding of
Equipment Product description Quantity Preferred vendor (USD) (USD) order delivery location source payment Notes

Fused labeled, Square D 2 EckSupply (local) $191.36 — — 2/20/ USFS ARC Credit Must label
lockable AC H221NRB, 2012 Ranger card disconnect
disconnect Service-rated fused Station switch,
panel disconnect 110 South mounted
switch and 1–20 Park outside
A 250VAC fuse Drive, powerhouse;
Blacksburg, approval
VA 24061 needed by
AEP
Fused main DC Square D H361RB 1 EckSupply (local) $260.38 — — 2/20/ USFS ARC Credit
disconnect fused 2012 Ranger card
disconnect Station
switch with 110 South
ground bar and Park
3–20 A (Ferraz Drive,
Shawmut Blacksburg,
Trionic TRS20R) VA 24062
600VAC/300VDC
fuses
Micro-Hydropower: Concept, System Design, and Innovations

(Continue)
85
86

Cost per Shipping Method


quantity cost Date of Date of Delivery Funding of
Equipment Product description Quantity Preferred vendor (USD) (USD) order delivery location source payment Notes

HL-700 air Item code: 1 AltE $230.00 $14.52 12/19/ 2/7/ USFS ARC Credit https://www.
heater ALRHL-100 43 Broad Street, 2011 2012 Ranger card altestore.
diversion Suite A408 Station com/store/
load Hudson, MA 110 South Charge-
01749-2556 Park Controllers/
USA Drive, Dump-Load-
18778784060 Blacksburg, Dump-Load-
See website in notes VA 24060 Controllers/
Diversion-
LoadDump-
Loads/
HL-100-Air-
Heater-
Diversion-
Load/p6150/
Coanda screen 3 ft long, 18 in. 1 Hydroscreen LLC $1,200.00 $140.27 2/8/ 2/16/ USFS ARC Credit www.
filter-pipe high, 4 ft2 of Bob Weir 2012 2012 Ranger card hydroscreen.
mount screen-6 in a 303-333-6071 Station com
pipe unit rkweir@aol.com 110 South
Park
Drive,
Renewable Energy Technologies and Water Infrastructure

Blacksburg,
VA 24060
(Continue)
Cost per Shipping Method
quantity cost Date of Date of Delivery Funding of
Equipment Product description Quantity Preferred vendor (USD) (USD) order delivery location source payment Notes

2 in. ID, 2–5/8 High-pressure PVC 1 McMaster Carr $154.40 $25.69 3/2/ 3/6/ USFS ARC Credit
in. OD tubing food, 2 200 New Canton Way 2012 2012 Ranger card
braided in. ID, 2–31/64 Robbinsville, NJ Station
vinyl tubing in. OD, 15/64 in. 08691-2343 110 South
wall Thk 609-689-3415 Park
Prod#: 52375K21 Drive,
Blacksburg,
VA 24061
Sample tubing See notes 1 McMaster Carr $28.52 — — — USFS ARC Credit Two samples of
200 New Canton Way Ranger card flexible
Robbinsville, NJ Station tubing were
08691-2343 110 South ordered to
609-689-3415 Park determine
Drive, which tube
Blacksburg, would be
VA 24060 best suited
for
application
Subtotal $9,022.52 Total USFS funding $2,365.50
Shipping costs $449.05 Total ARC funding (equipment) $7,314.07
Total cost of equipment $9,471.57 Total ARC funding (supplies/materials) $208.97
Micro-Hydropower: Concept, System Design, and Innovations

Total ARC funding $7,523.04


Source: Younos (2013).
87
88

APPENDIX B. BILL OF MATERIALS (SUPPLIES/MATERIALS, 2013 PRICES)

Method
Supplies/ Product Preferred Cost Shipping Date of Date of Delivery Funding of
materials description Quantity vendor (USD) cost (USD) purchase delivery location source payment Notes

2–3/4 in. Used to attach 10 Local hardware $ 2.00 — NA — USFS Credit


Hose hose to store card
clamps turbine (heaveners)
nozzle
100 psi Device used to 1 Local hardware $10.00 — 4/5/2012 — USFS Ranger ARC Credit Includes
pressure measure the store Station 110 card nipple and
gauge pressure at South Park 90 degree
the outlet of Drive, elbow-to
the penstock Blacksburg, be attached
pipe VA 24061 to penstock
flange
1.5 in. Brass Used to stop 4 Local hardware $26.00 — NA — USFS Ranger ARC/ Credit 1 bought need
gate valve water flow to store Station 110 USFS card 3 more
turbine (heaveners) South Park (USPS)
nozzles in Drive,
emergencies; Blacksburg,
threaded on VA 24061
inside
1.5 in. Used for 12 Local hardware $10.00 — NA — USFS Ranger ARC/ Credit
galvanized connections store Station 110 USFS card
steel from cap to (heaveners) South Park
Renewable Energy Technologies and Water Infrastructure

coupling gate valve Drive,


and from Blacksburg,
gate valve to VA 24062
flexible hose
(Continued)
Method
Supplies/ Product Preferred Cost Shipping Date of Date of Delivery Funding of
materials description Quantity vendor (USD) cost (USD) purchase delivery location source payment Notes

AL nipples for Connection into 4 McMaster-Carr $6.63 $2.75 3/23/2012 3/27/2012 1750 October ARC Credit
back wall watertight Glory Ct, card
caisson Blacksburg,
VA
AL nipples for Connection 1 McMaster-Carr $24.88 $2.75 3/23/2012 3/28/2012 1750 October ARC Credit
water from water Glory Ct, card
catcher catcher back Blacksburg,
to river VA
Multipurpose 1/4 in. thick 12 2 McMaster-Carr $48.26 $9.55 4/15/2012 4/18/2012 1750 October ARC Credit For caisson
aluminum in. width, 3 Glory Ct, card water
(alloy in. length Blacksburg, catcher box
6061) VA
Estimated subtotal $401.92
Shipping costs $15.05
Supplies/materials $416.97

Source: Younos (2013).


Micro-Hydropower: Concept, System Design, and Innovations
89
90 Renewable Energy Technologies and Water Infrastructure

ACKNOWLEDGMENTS

The Glen Alton micro-hydropower project case study was funded by the
Appalachian Regional Commission and in-kind support from the US Forest
Service Regional Office, Blacksburg, Virginia. Project collaborators included the
following: Eugene Brown, Project Co-Director/Advisor, Professor of Mechanical
Engineering, Virginia Tech; Sheryl Lyle, Project Site Manager/Field Director, US
Forest Service, Glen Alton; Justin Garrette, Project Manager/Research Assistant;
and the Student Design Team: Grant Bischof, Douglas Friedman, Theresa Sweeney,
John Thomas, and Gerald Zingraf, Mechanical Engineering Department, Virginia
Tech, Blacksburg, Virginia.

DISCLAIMER

Mention of the product, vendors, and trade names in this chapter are only for
research and education purposes and does not constitute an endorsement by
the authors of this chapter or the ASCE-EWRI Advancing Renewable Energy
Technologies Committee.

References
Alternative Energy News. 2018. “Micro hydropower—pros and cons.” Accessed June 12,
2020. https://www.alternative-energy-news.info/micro-hydro-power-pros-and-cons/.
altEstore. n.d. Accessed on September 19, 2021 https://www.altestore.com/store/
Binama, M., W. T. Su, X. B. Li, F. C. Li, X. Z. Wei, and S. An. 2017. “Investigation on pump
as turbine (PAT) technical aspects for micro hydropower schemes: A state-of-the-art
review.” Renewable Sustainable Energy Rev. 79: 148–179.
Casini, M. 2015. “Harvesting energy from in pipe hydro systems at urban and building
scale.” Int. J. Smart Grid Clean Energy 4 (4): 316–327.
CITYLAB. 2018. “How Portland is sourcing hydropower from its drinking water.” Accessed
June 12, 2020. https://www.citylab.com/environment/2018/01/portlands-drink​ing-wat​
er-is-powering-the-grid/550721/.
Corcoran, L., P. Coughlan, and A. McNabola. 2013. “Energy recovery potential using
micro hydropower in water supply networks in the UK and Ireland.” Water Sci. Technol.
Water Supply 13 (2): 552–560.
DOE. 2010. “Energy efficiency and renewable energy.” Accessed June 12, 2020. http://www.
energysavers.gov/your_home/electricity/index.cfm/mytopic=11110.
DOE. 2010. “Energy efficiency and renewable energy.” Accessed June 12, 2020. http://www.
energysavers.gov/your_home/electricity/index.cfm/mytopic=11110.
EERE (Energy Efficiency & Renewable Energy). 2019a. History of hydropower. Washington,
DC: Office of Energy Efficiency & Renewable Energy (EERE), US Department of Energy
(DOE). Accessed June 12, 2020. https://www.energy.gov/eere/water/history-hydropower.
EERE. 2019b. Types of hydropower plants. Washington, DC: Office of Energy Efficiency &
Renewable Energy (EERE), US Department of Energy (DOE). Accessed June 12, 2020.
https://www.energy.gov/eere/water/types-hydropower-plants.
Micro-Hydropower: Concept, System Design, and Innovations 91

EIA (Energy Information Administration). 2018. Hydropower explained: Tidal power.


Washington, DC: US EIA. Accessed June 12, 2020. https://www.eia.gov/energyexplained/
index.php?page=hydropower_tidal.
Energy Systems & Design Ltd. 2017. “Stream Engine Manual V2.01.” Accessed June 12,
2020. https://cdn.microhydropower.com/wp-content/uploads/2017/08/The-Stream-
Engine-v2.01.pdf.
Energy Systems & Design Ltd. n.d. “Stream Engine & X-Stream Engine.” Accessed
September 19, 2021. https://microhydropower.com/stream-engine/
EUEI (European Union Energy Initiative). 2010. “Policy and regulatory framework
conditions for small hydro power in Sub-Saharan Africa: Discussion paper.” Accessed
June 12, 2020. http://kerea.org/wp-content/uploads/2012/12/Policy-and-regulatory-
framework-conditions-for-small-hydro-power-in-Sub-Saharan-Africa.pdf.
García, I. F., D. Ferras, and A. McNabola. 2018. “Potential micro-hydropower generation
in community-owned rural water supply networks in Ireland.” MDPI Proc. 2 (11): 677.
García, I. F., D. Ferras, and A. McNabola. 2019. “Potential of energy recovery and water
saving using micro-hydropower in rural water distribution networks.” J. Water Resour.
Plann. Manage. 145 (3): 05019001.
IEA (International Energy Agency). 2018. Renewables 2018. Paris: IEA. Accessed June 12,
2020. https://www.iea.org/renewables2018/.
IRENA (International Renewable Energy Agency). 2012. Vol. 1 of Renewable energy
technologies: Cost analysis series. Hydropower. Working Paper. Accessed June 12, 2020.
https://www.iea.org/renewables2018/.
LucidEnergy. 2019. “FAQ/Fast Facts.” Accessed June 12, 2020. https://nwhanew.
memberclicks.net/assets/pdx%20ygh%20-%20lucidenergypdxfactsheet-2.pdf.
Maryland Clean Energy Center. 2017. “Tide-wave-hydro.” Accessed June 12, 2020. https://
mdcleanenergy.org/all-about-energy/advanced-technologies/hydro.
Michael, P. G., and C. P. Jawahar. 2017. “Design of a 15 kW micro hydro power plant for
rural electrification at Valara.” Energy Procedia 117: 163–171.
NEA (Nepal Electricity Authority). 2009. “A Year in Review, Fiscal Year 2008/09.” 9.
Kathmandu: Nepal Electricity.
Nunez, C. 2019. “Hydropower, explained.” National Geographic. Accessed June 12, 2020.
https://www.nationalgeographic.com/environment/global-warming/hydropower/.
Patelis, M., V. Kanakoudis, and K. Gonelas. 2016. “Pressure management and energy
recovery capabilities using PATs.” Procedia Eng. 162: 503–510.
Rayner, R. 1995. “Hydraulic ram: Special effect pumps.” In Pump user handbook. 4th
ed. Accessed June 12, 2020. https://www.sciencedirect.com/topics/engineering/
hydraulic-ram.
Smith, N. 2001. Motors as generators for micro-hydropower. Great Britain: ITDG
Publishing, Russell Press.
Upadhayay, S. 2009. “Evaluating the effectiveness of micro-hydropower projects in Nepal.”
Master’s thesis, Dept. of Environmental Studies, San Jose State Univ.
US Bureau of Reclamation. 2003. Design guidance for coanda-effect screens. Rep. R-03-03.
Denver: US Dept. of the Interior, Bureau of Reclamation, Technical Service Center.
USGS. 2018. Hydroelectric power water use. Water Science School (USGS). Accessed June
12, 2020.
https://www.usgs.gov/special-topic/water-science-school/science/hydroelectric-power-
water-use?qt-science_center_objects=0#qt-science_center_objects.
VA-DEQ (Virginia Department of Environmental Quality). 2015. Virginia net metering
regulation. Richmond, VA: VA-DEQ.
92 Renewable Energy Technologies and Water Infrastructure

Viccione, G., R. Amato, and M. Martucciello. 2018. “Hydropower potential from the
AUSINO Drinking Water System.” MDPI Proc. 2 (11): 688.
Wholesale SOLAR. n.d. Accessed on September 19, 2021. http://www.wholesalesolar.com/
Younos, T. 2013. Micro-hydro power generation demonstration project for community
education, Glen Alton, Virginia. Final project performance narrative. Submitted to
the Appalachian Regional Commission. Contract Number: ARC-VA-16955-2011. The
Cabell Brand Center for Global Poverty and Resource Sustainability Studies.
CHAPTER 5
Biogas-to-Energy—The
Combined Heat and Power
(CHP) Systems
S. Rao Chitikela, William F. Ritter

INTRODUCTION

Biogas-to-energy renewable systems are not new and are well utilized worldwide.
These renewable energy systems are widely seen at the municipal wastewater
processing facilities that biostabilize the sludge solids to produce biogas at dairy
facilities (cow manure to biogas recovery) and at a good number of industrial high-
strength (biodegradable-organic) wastewater or waste management facilities.
Biodegradable organic materials in anaerobic (no-available oxygen)
conditions are converted to organic acids by acid formers, which are in turn
gasified by methanogens to provide methane (CH4)-rich biogas. Other significant
gaseous elements are carbon dioxide (CO2) and hydrogen sulfide (H2S), water
vapor or moisture, and siloxanes. Many US municipal facilities with anaerobic
digesters for the processing of solids practice flaring of the biogas generated
on-site; however, currently, a good number of facilities are taking advantage of
recovering biogas and utilizing a CHP system on-site (where flaring will act as
a backup). The biogas-CHP operations will reduce air contaminant emissions
including greenhouse gases (GHGs).
The methane makeup in a biogas system is, in general, 60% to 65%, by
volume, and more than 30% is carbon dioxide. So, the heat value of biogas is,
in general, around 23,454 kJ/m3 biogas (630 Btu per standard cubic foot) when
compared with approximately 37,973 kJ/m3 natural gas (1,020 Btu per standard
cubic foot of natural gas) (butane-rich). Some municipal wastewater facilities have
been successful at cleaning up the biogas to 100% rich methane (i.e., pipeline
quality) and use it to in the utility gas supplies. However, biogas clean up or
conditioning is required before it is fired in an integrated electric generation or
combustion equipment. Various types of biogas-generating reactors or anaerobic
digesters and electric generation equipment are in use worldwide. Because

93
94 Renewable Energy Technologies and Water Infrastructure

GHGs are a significant concern and methane is more than 20 times potent than
carbon dioxide, the utilization of biogas generated from biodegradable and waste
materials is a productive and prudent activity. Biogas-to-energy comprehensively
fits into the circular economy.
This chapter will provide details of the following: biogas generation system
theory and practice, anaerobic digestion of dairy/cattle manure, applicable US
regulatory requirements, biogas-to-electric or CHP generation system(s), and
application in the circular economy.

BIOGAS GENERATION SYSTEMS—THEORY AND PRACTICE

Anaerobic digestion or fermentation is, in general, practiced (when feasible


on-site and including economically) at the municipal wastewater processing or
the resource recovery facilities (WRRFs), in livestock operations, and at industrial
(high-strength) wastewater or waste management facilities.
The anaerobic digestion of (readily) biodegradable materials, solids, or
wastewater by various mixed cultures of bacteria would include hydrolysis,
acidogenesis, acetogenesis, and methanogenesis. In the digestion process in Phase
I, complex organic materials are converted to low-molecular-weight fatty acids—
acetic, propionic, and butyric to name a few—which are termed volatile acids
(VAs), and but however, the accumulation of excessive VAs (2,000 to 3,000 mg/L
or more) will have a detrimental effect on the AD process by significantly lowering
the pH balance of the system, where the system pH values at less than 6.7 (or so)
are not favorable to methanogenesis in Phase II of this biochemical processing
of organics in ADs (Personal-Experience 2019, Sawyer and McCarty 1978). The
anaerobic biomethane conversion of 0.35 L per g COD (where COD is the chemical
oxygen demand) has been stoichiometrically calculated and reported (Metcalf &
Eddy | AECOM 2014). The process temperature, pH, and organic loading (volatile
solids) and hydraulic or solids retention time (SRT or HRT) are critical in the
operation of any anaerobic digestion or fermentation process. The mesophilic
range of temperatures is practiced for running anaerobic digesters (ADs). Although
the 32°C to 35°C process operating temperatures are good theoretically, process
operating temperatures of 35°C to 38°C [95 (more-of, 98) to 100°F] are utilized
in most of the municipal ADs in the United States, because the heat losses and/
or process-efficiency parameters are to be met in the field conditions, and a solid
retention time (SRT) of approximately 2 weeks (15 days) is, in general, observed
for operating ADs in the mesophilic mode. The thermophilic temperature (greater
than 45°C) range is also practiced at a few (advanced) facilities with more clear
objectives on end products. The pH of AD-sludge contents, in the process operation,
in the range of 6.8 to 7.3 would be optimal. Other process optimization parameters
would be the volatile solid concentration and alkalinity. Multiple units of sludge
digesters are utilized at the facilities based on proper handling of this process and
(readily) biodegradable feedstocks; for example, primary and secondary digesters
Biogas-to-Energy—The Combined Heat and Power (CHP) Systems 95

are used at the small-to-medium-scale municipal ADs for main process and
sludge-storage operations, respectively. The heat loss-causing elements of ADs may
be as follows (Metcalf & Eddy | AECOM 2014, Personal-Experience 2019): plain
concrete walls, either above or below the ground, plain concrete floors, floating
covers, fixed concrete covers, fixed steel covers, and others based on the AD design
and construction or installation; the heat and any other losses should appropriately
be accounted for at all times of the year, because the net biogas (inclusive of excess
biogas in the summer hot days) production calculations are critical for the entire
biogas-to-energy project.
The Ten State Standards, which are US developed and used by professional
engineers, on anaerobic (municipal sludge) digestion are available in Chapter 80,
§84; a few recommendations on ADs are as follows (Ten-State-Standards 2014,
Personal-Experience 2019):
• Standard design of digestion temperature to be in the range of 29°C to 35°C,
40% to 50% of volatile or organic matter in the digested sludge or content
and a periodic removal of digested content, and (according to §84.531) the
“design operating temperature should be in the range of 29°C to 38°C where
optimum mesophilic digestion is required.”
• ADs’ contents should be mixed (at all times) for maintaining the homogeneity
and required biostabilization and gas production, and digester-health
performance.
• Feedstock or the raw-sludge loading of 1.3 kg/(m3 day) for “completely-mixed”
and 0.65 kg/(m3 day) for “moderately-mixed” systems.
• Minimum side water depth of 6.1 m, with a reasonable depth for maintaining
the supernatant liquid.
• All safety precautions in the design and observance are a must.
• External heating as necessary to counteract the heat losses (in the winter
and other low-ambient-temperature conditions) to maintain the temperature
of ADs should be inclusive; the hot-water recirculation and management of
internal heating controls would be important.
• Waste gas destruction (via a flare) and/or an effective, biogas-use.
• Any possible presence of “toxic materials” in feedstocks and that can inhibit
the AD process must be controlled.
• Ease of operation and maintenance would be critical (digester-foaming will
need to be controlled at any time).
• Make sure the comprehensive monitoring of ADs is in place (at all times).
Various types of ADs are in practice worldwide: cylindrical with floating or
fixed cover, conventional German design with reinforced concrete construction
(including the clad in a metal sheath), and the most effective (and capital intensive)
egg-shaped with a steel shell. Heat exchangers or hot-water boilers are used in
maintaining the temperature of the sludge contents in the AD process. Various
96 Renewable Energy Technologies and Water Infrastructure

AD mixing systems would include (not limited to) gas injection (cover-mounted
lances, bottom-mounted diffusers, gas lifters, gas pistons, low-speed turbines,
low-speed mixers, linear motion mixing, internal or external draft tubes, pump
mixing, and so on…). Typical production of biogas would be 0.75 to 1.12 m3/kg
of volatile (organic) matter or solids reduction. The addition or co-digestion of
biodegradable liquid or high-strength wastes, such as food wastes, FOG (fats, oil,
and grease), can result in recovering biogas at enhanced levels of 1.6 m3/kg of unit
solids reduced, with a methane content of more than 70% by volume. The methane
content in the biogas that is less than 60% by volume or CO2 content that is more
than 40% by volume would be the biogas being generated at off-spec, and the AD
process verification must be conducted. Biogas conditioning is required for the
removal of moisture, hydrogen sulfide, and siloxanes (in the case of municipal
operations) (Metcalf & Eddy|AECOM 2014, Personal-Experience 2019).
Grit-free feedstock utilization and a periodic verification of effective digester
volume are the other critical operating parameters. Thus, the operation of ADs to
ensure the required level of organic biostabilization and optimal biogas recovery,
resulting in effectively digested by-product solids would be necessary. The
rehabilitation of ADs once in 10 to 15 years would be a good updating measure (if
not, on a shorter maintenance schedule).

ANAEROBIC DIGESTION OF LIVESTOCK AND POULTRY MANURE

On-farm biogas production has long been a topic of interest for farmers, with
historical records of biogas production dating back to several hundreds of years.
In 1808, methane was produced via a controlled anaerobic digestion of cattle
manure. In 1859, the first digestion plant was built in Bombay (now Mumbai).
In 1895, biogas recovered from a sewage treatment plant in England fueled
streetlamps. In the 1930s, developments in microbiology identified the anaerobic
bacteria and conditions needed to promote methane production. During the
energy crisis of the 1970s, there was great interest in on-farm anaerobic digestion.
Close to 200 digesters were built on dairy farms and more than 80% of them failed
or shut down. In the 1990s, there was an improvement in the design, but there
were still failures. Katers and Holzem (2015) list out the main reasons for the
failure of anaerobic digesters:
• Poor design and equipment selection,
• Lack of appropriate technical expertise,
• Lack of maintenance, and
• Lack of commitment by the operator.
Today, there is renewed interest in anaerobic digestion. Anaerobic digestion
comes with challenges, including the availability of equipment, construction and
startup costs, and equipment failure if systems are not managed or maintained
Biogas-to-Energy—The Combined Heat and Power (CHP) Systems 97

properly. However, it also comes with many of the following economic and
environmental benefits (Lleleji et al. 2008, Cherosky et al. 2011):
• Lower electrical, natural gas, and heating costs;
• Income from selling excess energy back to the grid;
• Enhanced fertilizer value of the digested manure;
• Fewer pathogens in manure;
• Less odor than raw manure;
• More stable nitrogen readily available to crops; and
• Reduced greenhouse gas (GHG) emissions.
Typical farm-based anaerobic digestion configurations are shown in Figure
5-1. The three main designs used for farm-based digesters are the covered
anaerobic lagoon (Figure 5-2), plug-flow (Figure 5-3), and complete mix (or a
continually stirred tank reactor) (Figure 5-4) (Hamilton 2013). The solid content
of the material to be digested is an important criterion in the choice of digester
design. Plug-flow digesters work best at a solid content of 11% to 13%, so they
work well with dairy manure from operations that collect it by scraping or other
methods that do not add much water. Complete-mix digesters work at a wider
range of 2% to 10% solids, which makes them suitable for a greater variety of

Figure 5-1. Typical anaerobic digestion configurations.


Source: Sharvelle and Loetscher (2011).
98 Renewable Energy Technologies and Water Infrastructure

Figure 5-2. Covered anaerobic lagoon.


Source: Hamilton (2013).

Figure 5-3. Plug-flow anaerobic digestor.


Source: Hamilton (2013).
Biogas-to-Energy—The Combined Heat and Power (CHP) Systems 99

Figure 5-4. Complete-mix anaerobic digestors.


Source: Hamilton (2013).

materials, including swine manure and processing wastes as well as dairy manure.
Variations in these three basic designs have been made to enhance biogas output
and/or to deal with varying moisture levels and other digestate characteristics.
In 1994, EPA started the AgSTAR program. AgSTAR is a collaborative
program sponsored by EPA and USDA that promotes the recovery and use
of biogas and, thus, to reduce methane (CH4) emissions from livestock waste.
AgSTAR assists those who want to purchase or implement anaerobic digesters by
identifying project benefits, risks, options, and opportunities. AgSTAR provides
information and participates in events to create a supporting environment for
on-farm anaerobic digester implementation. Today, AgSTAR partners with 11
state agencies, 12 universities, and 10 nongovernment organizations (NGOs)
to support all phases of anaerobic digester projects: planning, deployment, and
long-term success. According to AGSTAR’s database, there were 248 anaerobic
digesters operating on livestock farms in 2018 (Figure 5-5). Of these 248 digesters,
there are 198 dairy, 43 swine, 8 poultry, and 8 beef digesters. The major types of
digesters are plug-flow or complete mix (Figure 5-6).
The end uses of biogas from 2000 through 2018 for farm-operated anaerobic
digesters are depicted in Figure 5-7. CHP is the most common end use, followed
by electricity. The number of CHP and electricity projects has steadily increased
each year from 2000 to 2013. Since then, the electricity project count has become
more stable, whereas the number of CHP projects continues to rise at a slower rate.
100 Renewable Energy Technologies and Water Infrastructure

Figure 5-5. Number of anaerobic digesters on US Farms from 2000 to 2018.


Source: USEPA (2019).

Figure 5-6. Designs for farm-operated anaerobic digesters in 2018.


Source: USEPA (2019).

Boiler and furnace fuel projects (including the CNG/pipeline) have also increased
but at a much slower rate than CHP and electricity projects. Projects that flare the
biogas full time made up approximately 5% of all projects in 2018.
AgSTAR estimates that biogas recovery systems are technically feasible at over
8,000 large dairy and hog operations (EPA 2018). These farms could potentially
generate nearly 57,600 TJ [16 million megawatt hours (MWh)] of energy per year
and displace about 7.24 GJ/h [2,010 MWs)] of fossil fuel-fired generation.
Biogas-to-Energy—The Combined Heat and Power (CHP) Systems 101

Figure 5-7. Uses of anaerobic digestion biogas from 2000 to 2018.


Source: EPA (2019).

SLUDGES/BIOSOLIDS AND BIOGAS-TO-ENERGY—US


REGULATIONS

Biogas-to-energy as a renewable technology and associated regulatory information


is included in Chapter 2. However, the processing of various municipal sludges and
those mixed with other biodegradable feedstocks do have other environmental
regulatory requirements as part of operating municipal WRRFs. The requirements
of 40CFR—Protection of Environment, Part 503—Standards for the Use or
Disposal of Sewage Sludge, may apply, and a few definitions and requirements are
addressed as follows (40CFR 2019a and 40CFR 2019b):
40CFR §503.1(b)—Applicability: (1) This part applies to any person who
prepares sewage sludge, applies sewage sludge to the land, or fires sewage
sludge in a sewage sludge incinerator and to the owner/operator of a
surface disposal site. (2) This part applies to sewage sludge applied to
the land, placed on a surface disposal site, or fired in a sewage sludge
incinerator. (3) This part applies to the exit gas from a sewage sludge
incinerator stack. (4) This part applies to land where sewage sludge is
applied, to a surface disposal site, and to a sewage sludge incinerator.
40CFR §503.33(b)(1)—“Anaerobic digestion is the biochemical
decomposition of organic matter in sewage sludge into methane gas and
carbon dioxide by microorganisms in the absence of air.”
40CFR §503.33(b)(1)—[On the volatile solids reduction in AD
processes,] “The mass of volatile solids in the sewage sludge shall be
reduced by a minimum of 38 percent (see calculation procedures in
102 Renewable Energy Technologies and Water Infrastructure

“Environmental Regulations and Technology—Control of Pathogens


and Vector Attraction in Sewage Sludge,” EPA-625/R-92/013, 1992, U.S.
Environmental Protection Agency, Cincinnati, Ohio 45268).”
40CFR, Part 503, Appendix B(A)(3)—Processes to significantly reduce
pathogens (PSRP), includes the Anaerobic Digestion, “Sewage sludge
is treated in the absence of air for a specific mean cell residence time
at a specific temperature. Values for the mean cell residence time and
temperature shall be between 15 days at 35 to 55 degrees Celsius and 60
days at 20 degrees Celsius.”
In addition, there are applicable 40CFR Parts 60—Standards of Performance
for New Stationary Sources (NSPS), Subpart JJJJ—Standards of Performance
for Stationary Spark Ignition Internal Combustion Engines (and, other, as
applicable); and 63—National Emission Standards for Hazardous Air Pollutants
for Source Categories (MACT-regulations), Subpart ZZZZ—National Emissions
Standards for Hazardous Air Pollutants for Stationary Reciprocating Internal
Combustion Engines (and, other, as applicable), federal [and state and local, as
delegated and applicable; including based on state implementation plans (SIPs)]
regulatory requirements on installation and operation of air-pollutant causing
combustion turbine(s) or and IC-engine(s), and other (as applicable) (40CFR
2019c, d). The air contaminants included (not limited to) for the regulatory review
and verification are particulate matter—PM/PM10/PM2.5, sulfur dioxide (SO2),
nitrogen oxides (NOX ), volatile organic compounds (VOCs), carbon monoxide
(CO), hazardous air pollutants (HAPs), GHGs—CO2 , CH4, N2O, and others
(as applicable). The associated air permitting would include the requirements
on emissions, operational, monitoring, testing, record-keeping, reporting, and
compliance certification.
Therefore, as included previously, a review and verification of environmental
regulatory requirements must be completed before implementing biogas-to-
energy renewable projects.

BIOGAS-TO-ELECTRIC GENERATION AND COMBINED HEAT AND


POWER

The heat value expected of biogas (with 60% or more methane by volume) is
22,400 kJ/m3 (at 600 Btu/ft3) owing to its nature of various gas and particulate
mixtures—methane, carbon dioxide, hydrogen sulfide, nitrogen, particulates,
and water vapor or moisture, whereas at the standard temperature and pressure,
methane would have a lower heating value of 35,800 kJ/m3 (960 Btu/ft 3) and
compared with 37,300 kJ/m3 (1,000 Btu/ft3) of natural gas (which is a mix of butane,
propane, and methane, as applicable). This invaluable and renewable biogas can
be processed not only to generate electricity but also to develop or make newer
sustainable materials or bioproducts (via biogas liquefaction and processing). In
the context provided here, the discussion is limited to turning biogas to electricity,
Biogas-to-Energy—The Combined Heat and Power (CHP) Systems 103

and it is important to develop facilities as “distributed energy” systems. Biogas can


be fired in internal combustion engines, combustion turbines (at a bigger scale),
microturbines, and fuel cells, where electric power generation and heat recovery
efficiencies can range between 26% and 45% and 30% and 52% (based on the type
of equipment), respectively (Metcalf & Eddy | AECOM 2014).
CHP systems provide a higher energy efficiency (70% to 80% overall) because
electricity and heat recovery are stable and accomplished, and the payback period
could be as low as 3 years (Personal-Experience 2019). A feasibility study on
renewable biogas energy, conducted by the Narragansett Bay Commission (NBC),
provided the following comprehensive details: a small-scale CHP system was
considered using a microturbine or reciprocating engine for electric generation,
and the estimated payback period was 5 to 11 years (NBC 2009)
• Wastewater treatment facility, with the average dry weather flow of
0.106 MM m3/day (28 MGD);
• Facilities included the following: primary settling, secondary biological,
biological nutrient removal, UV disinfection of plant effluent, dissolved air
flotation thickening of waste activated sludge (WAS), and anaerobic digestion
of sludge(s) [primary, oil and grease (and scum/floatable materials), and WAS]
in the mesophilic temperature(s), with a monthly average electric demand at
5,040 MMJ/h (1.4 MW);
• 40% to 60% volatile solids reduction (VSred) was reviewed with biogas production
rates of 0.75 to 1.12 scm/kg VSred [12 to 18 standard cubic feet of biogas per
pound of VSred (scft/poundVSred)]; however, based on the author’s experience
with various municipal wastewater AD systems, the biogas production rates
were observed at 0.69 to 0.75 scm/kgVSred (11 to 12 scft/poundVSred), and a
0.94 to 1.12 (up to 1.25) scm/kgVSred [15 to 18 (up to 20) scft/poundVSred] could
be the expected codigestion with FOG and/or other high-strength wastes.
• Routine operations included two-thirds of biogas reuse/firing in the boilers
to maintain the heat to AD contents, and the excess is flared; additional heat
requirements are met by firing natural gas.
• Biogas composition was as follows: methane—>60% by volume; carbon
dioxide—34% to 38% by volume; nitrogen—1% to 2% by volume;
oxygen—<0.5% by volume; hydrogen sulfide—90 to 260 ppmv; chlorinated
VOCs—approximately 7 ppmv; nonchlorinated VOCs—approximately
35 ppmv; siloxanes (siloxanes are silicones used in industrial and personal
care products, with the D4s and D5s being predominant)—2.7 to 41.0 ppmv;
water (saturated at 35°C)—5.16%.
• Heat value of biogas [at the higher heat value (HHV) of dry gas]—23.1 MMJ/m3
(approximately 620 Btu/cft).
• Micro turbines would have the advantages of —compact size, good
cogeneration efficiency, low air emissions, and long maintenance intervals;
however, their disadvantages would be low fuel-electricity efficiency, high
capital cost, and necessitating higher gas cleanup.
104 Renewable Energy Technologies and Water Infrastructure

• Reciprocating engines would have the advantages oflow capital cost, high
reliability, turndown capabilities, quick startup, and a lower level of gas
cleanup; however, their disadvantages would be higher air emissions, high
noise, and frequent maintenance intervals.
• Biogas conditioning or cleanup is required to remove moisture, hydrogen
sulfide, and siloxanes [and carbon dioxide if, pipeline quality gas, higher heat
value (HHV), and/or reduced GHG emission is required—author’s note].
• Selection criteria for a microturbine or reciprocating engine system, the
comparison that was verified, is shown in Table 5-1.
• Grid interconnection and net metering are required, according to the federal,
state, and local regulations. Where the net metering is “a renewable energy
incentive that allows electricity customers that generate on-site renewable
power to interconnect their renewable generator (solar, wind, and so on) to
the grid allowing for excess electricity to be placed onto the grid for use by
other customers. In a net metering situation, the customer’s electric meter
will “spin” backwards, banking excess electricity production for future credit
with the local electricity supplier. Net metering rules and regulations vary
greatly from state to state…”
• Environmental permitting is required on air—nitrogen oxides, carbon
monoxide, carbon dioxide, and particulate-matter emissions, to name a few
(and wastewater and waste discharges, as applicable).
As of completing this chapter, the update on the NBC’s biogas renewable
energy project is as follows: the installation of a 600 kW IC-engine supported
CHP system was completed and the operations began in 2019. It is expected that
electric production would be at 14.4 × 1012 J per year (14.4 TJ or 4,000,000 kWh

Table 5-1. Microturbine and Reciprocating Engine Comparison.

Reciprocating
Operating parameter Microturbine engine
Electrical efficiency 25% (HHV) 35% (HHV)
Thermal efficiency 48% 45%
Design system efficiency 72% 79%
Net power output [MMJ/h (kW)] 1,562.4 (434) 1,994.4 (554)
Installed cost USD 1,625,000 USD 1,783,000
Biogas treatment annual cost USD 80,000 USD 70,000
Annual maintenance cost USD 160,000 USD 244,000
Availability 85% 90%
Average, capacity factor 92.6% 98.5%
Renewable energy credits (RECs) USD 102,372 USD 142,728
Average 10-year [$US/MMJ ($US/kWh)] 0.37 (0.103) 0.31 (0.086)
Source: Adapted from NBC (2009).
Biogas-to-Energy—The Combined Heat and Power (CHP) Systems 105

per year) and the useful heat at 24.25 × 1012 J per year (24.25 TJ or 23,000 MMBtu
per year). The expected reduction in carbon dioxide emissions would be at 1,000
metric tonnes per year (https://www.narrabay.com/about-us/renewable-energy/,
July 28, 2019).

ECONOMICS OF LIVESTOCK AND POULTRY MANURE


ANAEROBIC DIGESTION

The capital requirements to install a digester will vary widely depending on the
digester design chosen, size, and choice of equipment for utilizing the biogas and/
or for separating out the manure fiber. The current capital cost range for complete
digester systems is estimated at USD 1,300 to 2,600 per cow depending on the herd
size, with the cost to maintain an engine-generator set at USD 0.072 to 0.093/MMJ
(USD 0.020 to 0.026/kWh) of electricity generated. An AgSTAR regression of
investments made versus the herd size at 19 recent dairy farm plug-flow digesters
gave a result of “$730,147 + $796 per cow” in 2019 US dollars. Ancillary items that
may be incurred are charges for connecting to the utility grid and equipment to
remove hydrogen sulfide, which could add up to 20% to the base amount. Figuring
the ancillary items at 10%, the investment works out to $1.5 million for a 700-
cow dairy operation, going up to $3.5 million for 2,800 cows (Lazarus 2008). A
similar regression for 13 mixed digesters gave “$320,864 + $563 per cow.” A solids
separator would add up to another 12% to these amounts. There is considerable
interest in digester designs that are economically feasible for smaller farms, but
some digester components are difficult to scale down. A complete-mix digester
with separator installed on a 160-cow Minnesota dairy farm in 2019 US dollars
cost “$598,000, or $3,737/cow.” Another recent study found that the electrical
generation equipment made up on average 36% of the total investment for a
group of 36 digesters, suggesting that substantial cost savings may be possible in
situations where biogas can be used for heating rather than to produce electricity.
Aui and Wright (2017) did a life-cycle cost analysis of anaerobic digestion in
Iowa. Their results indicate that an anaerobic digester attached to a 2,400 head of
cattle operation that is codigested with glycerin and cornhusk has 3,420 MMJ/h
(950 kW) of generation capacity. At a capital cost of USD 3.21 million, it could
achieve an internal rate of return of 4.56% at electricity prices of US¢23.67/MMJ
(US¢6.59/kWh). By replacing cornhusk with rye and wheat, the internal rate of
return is still in the upper range of 4%. The main contributors to the cost include
capital, labor, and operating capacity (Figure 5-8).
Klavon et al. (2013) did a study on the cost of AD for small dairies in the
United States using cost data from 9 existing 100 to 250 cow dairies and 7
theoretical systems. They also reevaluated the minimum size dairy farm needed for
economically feasible ADs in the United States. Cash flow analysis results showed
that, in general, the total capital costs, capital costs per cow, and net costs per cow
decreased with increasing herd size in existing systems. Among existing revenue
106 Renewable Energy Technologies and Water Infrastructure

Figure 5-8. Biodegradable feedstocks, biogas and energy generation, and end use.
Source: OR-DOE (2018).

streams, the use of digested solids for bedding generated the highest revenue (USD
127 cow/year), followed by biogas use for heating and/or electrical generation (USD
64 to USD 89 cow/year) and CO2 credits (USD 9 cow/year). Capital costs ranged
from $US 1,016/cow to $US 3,810/cow. No system had a positive cash flow under the
assumed conditions (8% discount rate, 20-year term). However, six of the sixteen
(16) systems had positive cash flows when 50% cost sharing was included in the
analysis. They concluded that with cost sharing, economically viable AD systems
are possible on 250 cow dairies, and additional revenue streams, such as tipping
fees for food waste, may reduce the minimum size to 100 cow dairies.
Historically, the major benefits of ADs have included odor control and the
potential to generate income from the energy produced and from tipping fees
received for organic waste such as food waste that would be codigested with the
manure. Despite many benefits that ADs offer, they have not been widely adopted
by US dairy farms to date because, in general, the cost to own and operate an
anaerobic digestion system (ADS) exceeds the revenues and direct avoided costs.
From the perspective of the value of renewable energy produced, the price paid
by the utility for the electricity generated does not fully value the GHG reductions
that an ADS is able to deliver, as reported by Wright and Cooch (2017).
Lauer et al. (2018) developed a nonlinear optimization model to optimize
plant capacity for anaerobic digestion and maximize the net present worth. Their
analysis assessed the economic viability of using dairy cow manure for either (i)
on-farm production and the use of biogas to generate electricity and heat or (ii)
upgradation of biogas to biomethane, a natural gas substitute. In their research,
they used Idaho’s dairy farms as a case study. This analysis implies that 3,000
cows per farm are required for conducting an economical AD plant operation. For
farms with up to 3,600 animals, the highest net present value was achieved for the
on-farm use of biogas. Farms larger than that achieved their best economic results
via the production of biomethane. In total, about 45% of Idaho’s dairy manure
could be utilized by economically feasible biogas and biomethane plants. A higher
Biogas-to-Energy—The Combined Heat and Power (CHP) Systems 107

manure utilization rate could be achieved through joint, cooperative anaerobic


digestion plants and manure transportation.

BIOGAS IN THE CIRCULAR ECONOMY

Biogas is proven to fit in the circular economy system, where it is clearly explained
in the “flow of technical and biological materials.” Biogas has a role in “renewables
flow management” that includes the operation of biocycle and regeneration—
material farming and collection, extraction of biochemical feedstock, and AD/
composting and associated biogas management (WEF 2014).
Methane is a significant GHG of the worldwide air emissions and is 21 times
more potent than the (atmospheric) carbon dioxide. Of the total US methane
emission potential from wastewater, animal manure, landfill materials, and
industrial–institutional–commercial organic wastes, wastewater and animal
manure make up more than 50% of the approximately 7.86 million metric tons
of methane per year, as projected (NREL 2013). However, the global methane
emissions of wastewater and agricultural manure operations have been estimated
at more than 10% by the Year 2020 (Chitikela and Ritter 2018).
The EPA’s AGSTAR program tracks the AD projects of the livestock operations
in the United States. The realization of biogas to energy is slow in the country; for
example, the State of Oregon had accomplished only 8% of the opportunity by
the Year 2011; the projected biogas potential was more than 3.6 × 105 MJ/h (100
MW) (Weisberg and Roth 2011). However, the State of Oregon conducted a biogas
inventory in a comprehensive manner; this inventory provided the following
details that are worth noting (OR-DOE 2018).
A few key words are defined (not limited to) as follows:
Biogas production pathways: Different ways by which biogas is produced.
These include wastewater treatment plants, landfills, anaerobic digesters,
thermal gasifiers, and other methods that are defined by their primary
input. For example, dairy biogas is a production pathway while dairy
manure is a feedstock.
Carbon intensity: Amount of carbon dioxide released during the total
lifecycle, being production, transportation, distribution, consumption,
and disposal, of a product or service per unit of fuel (i.e., kilograms of
CO2 per Btu, or grams of CO2 per megajoule).
Contaminates: Elements or compounds that may harm machinery,
infrastructure, or air quality upon combustion of biogas and are
removed during cleaning, which includes sulfur compounds, moisture,
halogen compounds, silicon compounds, volatile organic compounds,
and particulate matter.
Diluent: Elements or compounds that dilute/lower the energy content
of biogas, such as carbon dioxide and nitrogen, and must be removed
during biogas upgrading prior to pipeline injection as RNG.
108 Renewable Energy Technologies and Water Infrastructure

Renewable natural gas (RNG): Biogas that has been processed to be


interchangeable with conventional natural gas for the purpose of meeting
pipeline quality standards or transportation fuel-grade requirements. It
is a resource which can be used and created in perpetuity from renewable
sources.
The gross annual methane potential in the State of Oregon has been estimated
at approximately 6.0 billion cubic feet of methane per year from agricultural
manure, wastewater, and food waste resources. The biogas production pathway
supply chains are reviewed; the RNG supply chain for wastewater treatment plants
has been presented, where the average carbon intensity for the pathway is calculated
at 19 gCO2e per MJ, as shown in Figure 5-9, and is as follows:
“Wastewater Treatment Plants—sewage at wastewater treatment plants
is processed on-site in an anaerobic digester to produce biogas, which

Figure 5-9. RNG supply chain for wastewater treatment plants.


Source: OR-DOE (2018).
Biogas-to-Energy—The Combined Heat and Power (CHP) Systems 109

is cleaned of some impurities prior to end uses. The biogas can be


used in three manners: combustion in a flare to destroy the methane,
combustion in an internal combustion engine to capture heat and/or
generate electricity, or cleaning and upgrading to pipeline-quality RNG
for use in vehicle or stationary applications. If used for the latter, the
RNG can be piped to an on-site facility or it can be trucked or injected
into the natural gas grid for transport to an offsite location.” Similar
and effective RNG supply chain details are available for greatly utilizing
the “agricultural manure,” where the average carbon intensity has been
estimated at −264 gCO2e per MJ (OR-DOE 2018).
This inventory reported on reductions in GHGs that “If the volume of RNG
that could be potentially captured and utilized in Oregon displaced fossil fuel
natural gas for stationary combustion, approximately (two) 2 million metric tons
(per year) of fossil fuel-based carbon dioxide would be prevented from entering
the atmosphere.”
The market potential exists in stationary fuels, transportation fuels, clean
fuels program(s), and carbon markets. However, the State of Oregon identified the
current barriers (that need to be overcome) for RNG development and successful
accomplishment in various markets, as follows: “The barriers identified fell into
the following categories: finance, information, transportation and stationary
fuel markets, existing statutes and rules governing the Oregon Public Utility
Commission, and a general category of ‘other,’ which includes an examination of
barriers around pipeline access, contract language, competition for feedstocks,
out-of-state producers, and incentives.”
The NBC (2009) project estimated a simple payback period of a biogas-to-
energy CHP project at a municipal wastewater plant—8.2 and 15.4 years for
reciprocating and microturbine systems, respectively, with power generation and
RECs (for more details of the project, please refer to the reference). Moreover,
several applicable funding opportunities were verified: Clean Renewable Energy
Bonds for USD 1.2MM, US Department of Economic Development (US-DED),
State of Rhode Island Economic Development Corporation for USD 750,000 [and,
the Federal American Recovery and Reinvestment Act (ARRA) funds, then].
With these potential availability of grants on the renewable energy projects, the
simple payback and net present values would certainly make RE projects highly
lucrative.
The author was involved in a biogas-to-energy CHP project, where the
municipal wastewater treatment plant’s AD process upgradation included the
following: FOG receiving; biogas conditioning to remove moisture, hydrogen
sulfide gas, and siloxanes; and two ICEs. The benefits have been realized on
power generation at 1,692 MMJ/h (>470 kW) and heat recovery and the design
included the facility to foresee “netzero-energy” operation(s) (Chitikela and
Simerl 2017).
There is potential for 12,000 new biogas systems to be developed in the
United States based on wastewater and livestock manure sources in the country
and that could be equivalent to the electricity production of approximately
110 Renewable Energy Technologies and Water Infrastructure

68,400 TJ (19 billion kWh) per year. Biogas systems can drive economic growth:
distributed energy system development and energy delivery; proven revenue
streams; job growth; production of high-quality and concentrated organic
fertilizer; avoidance of costs of fossil- or petroleum-based fuel transportation
and heat; the available system-payback paves the way for project financing; and
there is the availability of RECs and associated demand for this renewable energy.
The biogas-distributed energy systems support the local economy and are a win–
win to all stakeholders (USDA-EPA-DOE 2014). At a minimum, the biogas-to-
energy project(s) development to commissioning includes the following: the
preliminary study; RFQs (request for qualifications); selection of a qualified
provider; detailed design—conceptual verification, basis of the design, plans, and
specifications; environmental permits; cash flow—capital expenditure (CapEx),
operational expenditure (OpEx), savings, rebates, grants, and so on; and effective
construction or installation. Thereon, periodic measurement and verification of
the key performance indicators for the project duration (and beyond) should be
conducted (Chitikela and Ritter 2018).
Thus, the biogas-to-energy CHP projects are proven: class of renewable; in
the control of methane emissions; for local energy independency; and, to meet the
triple bottom line (TBL) including resiliency requirements.

SUMMARY

Biogas-to-energy is a proven renewable energy system and has successfully been


practiced worldwide. Livestock manure, sewage solids, or unstabilized biosolids
obtained as raw by-products from wastewater processing; FOG; vegetable and food
wastes; and other municipal and industrial highly biodegradable wastes are the
feedstocks to a biogas-to-energy CHP system. Biogas is obtained via anaerobic
digestion of said feedstocks and comprises a high concentration of CH4, which is
a GHG and, therefore, release of methane needs to be controlled and put to use in
recovering this renewable energy (and toward making other useful products, as
research is ongoing). Many anaerobic digester systems with various configurations
are used worldwide and the effective design and O&M of these systems are
critical. Biogas conditioning or cleanup is necessary since it contains other than
methane, such as moisture, hydrogen sulfide, siloxanes (based on feedstocks)
and carbon dioxide; and to suit the downstream use of combustion operations or
pipeline-quality supply, as applicable. The applicable local/state/federal regulatory
requirements should be followed, such as US 40CFR, Part 503—Standards for the
Use or Disposal of Sewage Sludge (and other energy standards, if applicable). The
economics of biogas-to-energy systems are sturdy or proven and generate a positive
cash flow (with a shorter payback period). Thus, the stabilization of biodegradable
and wasted feedstocks via anaerobic digestion and the generation of biogas is a class-
renewable since it also provides the feasibility to produce other environmentally
friendly byproducts and, fits within the “circular economy.”
Biogas-to-Energy—The Combined Heat and Power (CHP) Systems 111

References
Aui, A., and M. M. Wright. 2017. Life cycle cost analysis of the operation of anaerobic
digesters in Iowa. Ames, IA: Dept. of Mechanical Engineering, Iowa State Univ.
40CFR. 2019a. “Protection of environment.” Accessed July 25, 2019. https://www.ecfr.gov/
cgi-bin/textidx?SID=b13fefbf5f10e1884553b45260cff34d&mc=true&tpl=/ecfrbrowse/
Title40/40tab_02.tpl.
40CFR. 2019b. “Part 503 – Standards for the use or disposal of sewage sludge.” Accessed
July 25, 2019. https://www.ecfr.gov/cgi-bin/text-idx?SID=7aa0be8eb1c1f4b29d6052374
2c1c4e1&mc=true&node=pt40.32.503&rgn=div5.
40CFR. 2019c. “Part 60 – Standards of performance for new stationary sources.” Accessed
July 28, 2019. https://www.ecfr.gov/cgi-bin/text-idx?SID=b46b2c691bf7b181c96f61
7cad15fc85&mc=true&node=pt40.7.60&rgn=div5; https://www.ecfr.gov/cgi-bin/
text-idx?SID=369d2f86b3386315d03304c406239da1&mc=true&tpl=/ecfrbrowse/
Title40/40cfrv8_02.tpl#0.
40CFR. 2019d. “Part 63 – National emission standards for hazardous air pollutants
for source categories.” Accessed July 28, 2019. https://www.ecfr.gov/cgi-bin/text-
idx?​SID=​a 32b86079f09f0cc0​490efc2f7b14ff0&mc=true&node=pt40.15.63&rgn=
div5; https://www.epa.gov/stationary-sources-air-pollution/national-emission-stand​
ards-hazardous-air-pollutants-neshap-9.
Cherosky, P., Y. Li, and K. Mancl. 2011. Manure to energy through anaerobic digestion. Fact
Sheet AEX-653.1. Columbus, OH: Ohio State Univ.
Chitikela, S. R., and W. F. Ritter. 2018. “The wasted biodegradable organic materials – A
renewable energy resource(s) and the sustainability requirements.” In Presented at the
111th Annual Conf. & Exhibition, Air & Waste Management Association. Hartford, CT.
Chitikela, S. R., and J. Simerl. 2017. “Municipal Water and Wastewater Infrastructure
Management and the Sustainable Utility – A Performance Contracting (PC) Review.”
In Proc., World Environmental and Water Resources Congress. Sacramento, CA.
EPA (Environmental Protection Agency). 2018. Market opportunities for biogas recovery
systems at U.S. livestock facilities. Rep. No. EPA-430-R-18-005. Washington, DC: EPA.
EPA. 2019. AgSTAR data and trends. Washington, DC: EPA.
Hamilton, D. W. 2013. Anaerobic digestion of animal manures: Types of digestors. Fact
Sheet BAE-1750. Stillwater, OK: Oklahoma Cooperative Extension Service, Oklahoma
State Univ.
Katers, J. F., and R. Holzem. 2015. “4 Reasons anaerobic digesters fail.” Progressive Dairy,
Accessed July 14, 2020. https://www.progressivedairy.com/topics/manure/4-rea​
sons-why-anaerobic-digesters-fail.
Klavon, K. H., S. A. Lansing, W. Mulbry, A. R. Moss, and G. Felton. 2013. “Economic
analysis of small scale anaerobic digestion in the United States.” Biomass Bioenergy
54: 36–45.
Lauer, M., J. K. Hansen, P. Lamus, and D. Thran. 2018. “Making money from waste: The
economic viability of producing biogas and biomethane in the Idaho Dairy Industry.”
Appl. Energy 222: 621–636.
Lazarus, W. J. 2008. Farm based anaerobic digesters as an energy and odor control
technology — Background and policy issues. AER 843. Washington, DC. Office of Energy
Policy and New Uses, Office of the Chief Economist, USDA.
Lleleji, K. E., C. Martin, and D. Jones. 2008. Basics of energy production through anaerobic
digestion of livestock manure. Fact Sheet ID-406-W. West Lafayette, IN. Purdue
Cooperative Extension, Purdue Univ.
112 Renewable Energy Technologies and Water Infrastructure

Metcalf & Eddy|AECOM. 2014. Wastewater engineering, treatment, and resource recovery.
5th ed. New York: McGraw-Hill.
NBC (Narragansett Bay Commission). 2009. Bucklin point renewable biogas energy
feasibility study, and Rhode Island State Office of Energy Resources RI.
NREL. 2013. Energy analysis – Biogas potential in the United States. NREL/FS-6A20-
60178. Golden CO: NREL.
OR-DOE. 2018. Biogas and renewable natural gas inventory. SB334 (2017) – 2018 Report
to the Oregon Legislature. Salem, OR: OR-DOE.
Personal-Experience. 2019. Author’s experience working with various biogas-to-energy
CHP systems.
Sawyer, C. N., and P. McCarty. 1978. Chemistry for environmental engineering. 3rd ed. 2nd
Printing (Singapore). Singapore: McGraw-Hill.
Sharvelle, L., and L. Loetscher. 2011. Anaerobic digestion of animal waste in Colorado. Fact
Sheet 1227. Fort Collins, CO. Colorado State Extension, Colorado State Univ.
Ten-State-Standards. 2014. Recommended standards for wastewater facilities. Albany, NY:
Health Research, Health Education Services Division.
USDA-USEPA-DOE. 2014. “Biogas opportunities roadmap – Voluntary actions to reduce
methane emissions and increase energy independence.” Accessed July 25, 2019. https://
www.usda.gov/oce/reports/energy/Biogas_Opportunities_Roadmap_8-1-14.pdf.
WEF (World Economic Forum). 2014. Towards the circular economy: Accelerating the
scale-up across global supply chains. Geneva: WEF.
Weisberg, P., and T. Roth. 2011. Growing Oregon’s Biogas Industry; A Review of Oregon’s
Biogas Potential and Benefits. A White Paper by The Climate Trust and Energy Trust of
Oregon (February 2011). Energy Trust of Oregon. Portland, OR.
Wright, P. E., and C. A. Cooch. 2017. “Paper Number 1700626: Estimating the Economic
Value of the Greenhouse Gas Reductions Associated with Dairy Manure Anaerobic
Digestion Systems Located in New York State Treating Dairy Manure.” In Presented at
the ASABE Annual Int. Meeting, Spokane WA.
CHAPTER 6
Fuel Cells for Renewable
Wastewater Infrastructure
Bhuvan Vemuri, Govinda Chilkoor,
Jawahar Kalimuthu, Ammi Amarnath,
James E. Kilduff, Venkataramana Gadhamshetty

INTRODUCTION

Modern infrastructure entails a significant interdependence between water


security and energy security. Rapidly changing climate and population growth has
resulted in an increasing demand for energy and water (or vice versa), increasing
the stress on the water–energy nexus. Figure 6-1 shows the complex interplay of
energy production and water consumption, primarily in the context of energy and
water exchange between power plants, wastewater treatment plants [or the water
resource recovery facility (WRRF)], municipalities, and commercial entities. As
shown in Figure 6-1, action taken in one area (energy production) will impact the
other (water consumption) (DOE 2006).
For example, a 1,800 GJ/h (500 MW) thermoelectric power plant uses nearly
1.14 million m3 (300 million gal.) of freshwater on a daily basis, out of which 97% is
returned into the environment (Feeley et al. 2005). Noting that electricity demand
in the United States is expected to grow by 41% in the high-economic growth case
(annual GDP growth rate = 2.8%) and 20% in the low-economic growth case
(annual GDP growth rate = 1.9%) (EIA 2014), we can observe a corresponding
increase in water consumption (see EIA 2014 for definitions of the high and low-
economic growth cases). Water is also used for energy extraction; for example, the
horizontal drilling and hydraulic fracturing methods that have enabled gas and
crude oil production from previously inaccessible shales also entail significant use
of freshwater (Shrestha et al. 2017).
Energy is required to extract and deliver freshwater for human consumption.
Total water withdrawals in the United States have been reported to reach as high
as 1.2 trillion m3 [322 billion gallons (Bgal. or BG)] per day (USGS 2015). Nearly

113
114 Renewable Energy Technologies and Water Infrastructure

Figure 6-1. Interrelation between water and energy.


Source: DOE (2006).

41.4% of the water (0.5 trillion m3 or 133 Bgal. per day) is used by thermoelectric
power plants, 36.7% (0.45 trillion m3 or 118 Bgal. per day) for irrigation, 12.1%
(0.15 trillion m3 or 39 Bgal. per day) for public supply, 4.6% (56.0 million m3 or
14.8 Bgal. per day) by self-supplied industries, 2.3% (28.8 million m3 or 7.6 Bgal.
per day) for aquaculture, 1.2% (15.1 million m3 or 4 Bgal. per day) for mining, and
the remaining for commerce and self-supplied domestic and livestock operations
(USGS 2015). The use of potable water by municipalities and industries, in turn,
produces wastewater, which requires treatment by publicly owned treatment
works, industrial works, or a WRRF. This wastewater infrastructure is responsible
for nearly 3% to 4% of the net energy consumption in the United States and
accounts for greenhouse gas emissions of 45 million tons per year (Pirne and
Yonkin 2008).
In terms of the economic impact, the energy requirements of water and
wastewater infrastructure account for 35% of municipal budgets in the United
States (Pirne and Yonkin 2008). Among the energy costs, electricity accounts
for 25% to 40% of the operating costs for wastewater treatment and 80% for
drinking water treatment and distribution (Pirne and Yonkin 2008). Against
this background, there is a clear need to minimize both water and energy
consumption. On a positive note, the new technologies discussed in this study
can help use wastewater as a resource of energy. Anaerobic digestion (AD) is
one approach currently used to do this by generating biogas (or biomethane)
that can be used as a fuel to generate heat and electricity. However, impurities
in the form of carbon dioxide, carbon monoxide (CO), water vapor, nitrogen,
Fuel Cells for Renewable Wastewater Infrastructure 115

ammonia, siloxanes, and hydrogen sulfide in biogas reduce the overall efficiency
of internal combustion engines. It is also quite expensive to refine and compress
biogas (Ong et al. 2014). In addition, AD systems are more effective for treating
high-strength industrial wastewaters and sludge, as compared to municipal
wastewater. Bioelectrochemical processes can potentially treat low-strength
wastewater and simultaneously generate pure hydrogen and methane or directly
generate electricity.
This study focuses on the use of the emerging fermentation processes or
bioelectrochemical processes for rendering a WRRF, a “net-zero” consumer of
energy. The hydrogen that can be generated at the WRRF can then be used by
a range of established fuel-cell technologies; the power generated by the fuel
cells (FCs) can then meet the electricity demands of typical equipment and
process operations under the wastewater infrastructure. Figure 6-2 shows typical
equipment and power requirements for a 37,850 m3/day (10 MGD) activated
sludge municipal WRRF. The equipment include the following: pumps; motors
used in screens, grit chambers, skimmers, and sludge rakes; compressors or
aerators used in secondary treatment (activated sludge process); lamps used in
UV disinfection; motors used for mixing and filter presses in sludge treatment;
and heating, ventilation, and air-conditioning (HVAC) and lighting used in
buildings and grounds. Figure 6-2 also shows five different classes of FCs that
can be used to power various types of WRRF equipment with corresponding
power outputs. These include alkaline fuel cells (AFCs), molten carbonate fuel
cells (MCFCs), phosphoric acid fuel cells (PAFCs), proton exchange membrane
fuel cells (PEMFCs), and solid oxide fuel cells (SOFCs) rated from 0.11 MJ/h to
7.2 GJ/h (30 W to 2 MW) (Kirubakaran et al. 2009).
Section 6.2 provides extensive details for each type of fuel cell and its
potential uses in wastewater infrastructure. As shown schematically in Figure 6-3,
this study discusses a suite of biohydrogen production technologies and fuel-cell
designs that can replace energy-intensive secondary treatment processes, improve
the efficiency of wastewater treatment, and enable the use of wastewater as a
resource for generating energy carriers and electricity. We present an approach
that combines dark fermentation (DF) processes to convert the chemical oxygen

Figure 6-2. A hypothetical example of a 10 MGD WRRF that can use various types
of fuel cells to drive typical equipment and process operations.
Note: AFC = alkaline fuel cells; MCFC = molten carbonate fuel cell; PAFC = phosphoric acid fuel
cell; PEMFC = proton exchange membrane fuel cell; SOFC = solid oxide fuel cell.
116 Renewable Energy Technologies and Water Infrastructure

Pumping

Hydrogen from Fuel cell


Wastewater Biomass
Heating and cooling
Heat /
Hydrogen gas electricity

Lighting

Fermentation
effluent Microbial fuel cell
Transportation
Electricity

Microbial fuel cell effluent Treated water

Reuse
Ultra filtration

Figure 6-3. A conceptual schematic of a biomodule, UF, and a microbial fuel cell for
wastewater treatment and energy production.

demand (COD) in wastewater into hydrogen, which can be used in FCs to generate
heat and electricity [combined heat and power (CHP)]; the fermentation effluent
can be further treated in microbial fuel cells (MFCs) to generate electricity.
A major opportunity is to replace the activated sludge process with microbial
electrolysis cells (MECs) that treat wastewater under anaerobic conditions and
simultaneously produce biohydrogen.

HYDROGEN

Hydrogen is almost always bonded with other elements, forming compounds


such as water, hydrocarbons, and hydroxides in minerals. Unlike oxygen and
nitrogen, the gaseous form of hydrogen does not freely exist in nature. A range of
chemical, thermochemical, and biological processes generate hydrogen. Unlike
fossil fuels, the combustion of hydrogen does not generate greenhouse gases;
instead, it yields pure water. The classic methods of hydrogen production rely
on energy-intensive processes, including catalytic steam reforming of naphtha
or natural gas, gasification of coal, and electrolysis of water. This study discusses
the use of biological processes for generating hydrogen from wastewater under
ambient conditions.

Biological Processes for Hydrogen Production


Hydrogen produced via biological processes is known as biohydrogen.
Figure 6-4 provides an overview of four methods for producing biohydrogen.
Fuel Cells for Renewable Wastewater Infrastructure 117

Figure 6-4. Biohydrogen production methods.

Light-independent reactions include DF and MECs, whereas light-dependent


reactions include photofermentation and biophotolysis.

Dark Fermentation
Gram-positive bacteria, including Clostridia spp., can ferment both pure
substrates (glucose and sucrose) and complex substrates (corn stover and biomass)
to generate biohydrogen. The acetate fermentation pathway yields nearly four
moles of H2 per mole of glucose [Equation (6-1)], whereas the butyrate pathway
yields only two moles of H2 per mole of glucose [Equation (6-2)]. Operational
parameters, such as the hydraulic retention time, pH, and hydrogen partial
pressure, can be engineered to achieve the desired fermentation pathway

C 6H12O6 + 2H2O → 4H2 + 2CO2 + 2C 2H 4O2 (6-1)

C 6H12O6 → 2H2 + 2CO2 + C 4H8O2 (6-2)

DF offers high production rates and entails the use of a simple reactor design.
Major disadvantages include low yield and accumulation of volatile fatty acids
(VFAs). These disadvantages could be overcome by combining DF with downstream
processes that require volatile fatty acids as feedstock, including photofermentation,
microbial electrolysis, and microbial fuel-cell processes. Furthermore, the VFAs in
biohydrogen effluents can be used as carbon substrates in tertiary bioprocesses, such
as enhanced biological phosphorus removal (EBPR) processes.
DF has been demonstrated on a relatively large scale and for a variety of
applications. Wang et al. (2018) demonstrated the use of DF for generating
hydrogen from the space crew’s waste (3.999 mM H2/g waste). Van Ginkel et al.
(2005) used DF to treat the wastewater from apple and potato processing plants,
producing hydrogen at a rate of 0.7 to 0.9 liter H2 per liter and 2.1 to 2.8 liter H2
118 Renewable Energy Technologies and Water Infrastructure

per liter of wastewater, respectively. Han and Shin (2004) used a leaching-bed
reactor to generate hydrogen from food waste at a maximum production rate
of 04.21 m3 H2 per m3 of waste per day. An anaerobic sequencing batch reactor
was used by Mohan et al. (2007) to treat dairy wastewater with a COD removal
efficiency of 64% and a hydrogen yield of 1.105 mmol H2/m3/min; also, this team
has demonstrated a pilot of 10 m3 DF technology.

Microbial Electrolysis Cells


MECs for hydrogen production were first reported in 2005 (Liu et al. 2005). MECs
can use pure substrates as well as complex waste streams including DF effluents
containing high levels of VFA. These devices use the energy and protons produced
by microbial transformation of organic matter, coupled with an additional (small)
electric current (Figure 6-5). This current can be driven by a potential of only 200
to 600 mV to produce hydrogen gas in the cathode chamber. This is in contrast
to the electrolysis of water, where a potential of 1,210 mV is theoretically required
to produce hydrogen gas at a neutral pH, and a potential of 1,800 to 2,000 mV is
required under alkaline conditions owing to overpotential of electrodes (Cheng
et al. 2002). In an MEC, microbes consume organic matter and produce electrons
(e−) and protons (H+). Therefore, hydrogen production in an MEC requires less
energy than a traditional electrolysis to produce hydrogen gas. One of the main
advantages of MECs is their capacity to treat wastewater and produce hydrogen.
Significant progress in the MEC field is observed over the past decade.
Heidrich et al. (2013) demonstrated biohydrogen production (0.015 liter H2
liter−1 day−1) using domestic wastewater in a pilot-scale MEC (120 L). A 1,000 L
pilot-scale continuous flow MEC was used by Cusick et al. (2011) to treat winery
wastewater, in which they achieved a COD removal efficiency of 62% and biogas
(approximately 60% biomethane) of 0.19 liter/liter wastewater/day produced.
Call and Logan (2008) achieved hydrogen production (3.12 m3 H2/m3 reactor per
day) in a single-chambered MEC (no membrane), which reduces the cost of these
systems. However, compared with other biohydrogen production processes such
as DF, MECs require higher energy inputs; MECs also require external energy
inputs in the form of electric supply (approximately 800 mV). The hydrogen
conversion efficiency for MECs is also quite low. Another challenge is that MECs
require expensive electrode and membrane materials.

Figure 6-5. Schematic of a microbial electrolysis cell.


Fuel Cells for Renewable Wastewater Infrastructure 119

Photofermentation
As shown in Equation (6-3), photoheterotrophic bacteria can generate hydrogen
from organic substrates via photofermentation. This process requires an external
source of light (Figure 6-6)
C 6O6H12 (glucose) + 6H2O → 12H2 + 6CO2 (6-3)

Unlike DF, anaerobic fermentation of organic waste produces organic


acids such as acetic, butyric, and lactic acids, which are converted to hydrogen
and carbon dioxide by phototrophic bacteria in the presence of light. Because
of low light conversion efficiencies, the hydrogen production rates (in m3) are
small, but the yields (in mL/g substrate) are high in this process (Barbosa et al.
2001; Nath et al. 2008). Recently, the photosynthetic bacterium Rhodobacter
sphaeroides has been reported to produce biohydrogen from light, water, and
limited nitrogen condition using mixed and single carbon sources (Hakobyan
et al. 2019). Microalgae and cyanobacterial are potential renewable resources for
biohydrogen production from light and water. Algal biomass can produce H2 in
two different ways: (1) extracellular biohydrogen production using specialized
hydrogenase genes; and (2) DF using microalgal biomass/residues as substrates for
H2 production (Wang and Wan 2009; Wang and Yin 2018). A schematic diagram
of photofermentation is shown in Figure 6-6.
Although photofermentation offers a higher hydrogen yield than DF, hydrogen
production efficiency is about two orders of magnitude lower, and the simultaneous
accumulation of oxygen by algae inhibits the hydrogenase enzymatic activity and
further limits biohydrogen production (Wang and Yin 2018). Therefore, for the
same hydrogen production rate, the photofermentation process requires twice the
reactor size compared with DF. Although there are several feedstocks reported for
biohydrogen production through photofermentation, actual hydrogen production
yields remain much lower than the theoretical maximum value (Su et al. 2009).
Photofermentation requires a good reactor design that provides an external light

Figure 6-6. Schematic of the photofermentation process for hydrogen production.


120 Renewable Energy Technologies and Water Infrastructure

source, allows adequate light penetration, and minimizes light attenuation at


higher cell densities. Also, the challenges vary with the method of pretreatment,
raw material properties, light distribution and density, photobioreactor
configuration, and light-, heat-, and mass-transfer properties. Combining DF
with downstream photofermentation holds considerable promise and has helped
successfully improve overall hydrogen yield (Su et al. 2009).

Biophotolysis
Algae and cyanobacteria can generate H2 from water in the presence of sunlight
[Equation (6-4)], and this process is known as biophotolysis. The solar conversion
efficiencies by this process are quite low. In this process, microorganisms, such as
green microalgae or cyanobacteria, use sunlight. These microorganisms contain
photosynthetic (PS) pigments such as chlorophyll that can perform oxygenic
photosynthesis. Sunlight provides the energy to split oxygen and hydrogen ions
from water. The produced hydrogen ions can be combined through direct and
indirect routes to produce hydrogen gas. A schematic diagram of direct and
indirect biophotolysis is shown in Figure 6-7.

Figure 6-7. Schematic of biophotolysis: (a) direct biophotolysis, and (b) indirect
biophotolysis.
Fuel Cells for Renewable Wastewater Infrastructure 121

2H2O → 2H2 + O2 (6-4)

Microalgae and cyanobacteria are involved in biohydrogen production


through biophotolysis using two distinct pathways, namely, direct photolysis and
indirect photolysis. Microalgae, such as Chlamydomonas reinhardtii, use direct
biophotolysis and rely on photosystems PSI and PSII along with the hydrogenase
enzyme systems (Oh et al. 2013). Cyanobacteria use indirect biophotolysis, with
a simultaneous accumulation of electrons/reducing equivalents as carbohydrate
energy reserves, along with DF for H2 production.
The major advantage of biophotolysis is that it generates hydrogen using
water as the sole feedstock. Hydrogen gas produced in this process has a purity
of 98% (Kruse et al. 2005). However, in addition to hydrogen, microorganisms
produce oxygen, which reduces the overall hydrogen production efficiency.

FUEL CELLS

Table 6-1 provides an overview of different FCs that can be used to drive equipment
and process operations under water infrastructure. Table 6-1 also provides
information on the type of electrolyte, half-cell reactions, and performance
parameters including power density and the cost. FCs convert chemical energy
into electrical energy by reacting fuel and an oxidant. Hydrogen, methanol, and
hydrocarbons can be used as fuels (Wang and Jiang 2017), and ambient air can
be used as an oxidant. An FC consists of four different components: (i) an anode,
where the fuel such as hydrogen is oxidized, producing protons and electrons; (ii)
a cathode, where an electron acceptor such as oxygen is reduced to form water;
(iii) a conductive electrolyte that transports the protons from the anode to the
cathode; and (iv) an external circuit that electrically connects the anode with the
cathode.
Although FCs are commonly used to generate electricity, they can also be
used to operate in a CHP mode. The thermal energy produced in a fuel cell is on
account of electrochemical reactions at the electrodes, overpotential losses, and
Joule heating. AFCs are commonly used in spacecraft to generate drinking water.
The FCs are commonly classified based on the choice of the electrolyte used in
the cell. For example, when phosphoric acid is used as an electrolyte, they are
grouped as phosphoric acid fuel cells (PAFCs or PFCs). The type of electrolyte
further determines the choice of catalyst, and the operating temperature, which
can affect the fuel-cell application. Currently, there are several types of FCs that
have their own advantages, limitations, and potential applications. The following
sections give a brief overview of the six major fuel-cell designs, each classified
based on the electrolyte and fuel used.
122

Table 6-1. Comparison of Different Fuel Cells and Their Application in Wastewater Infrastructure.

Parameters
Fuel cell PEMFC AFC PAFC MCFC SOFC MFC
Electrolyte Solid polymer Liquid solution of Phosphoric acid Lithium and Stabilized solid Ion-exchange membrane
membrane KOH (H3PO4) potassium oxide electrolyte (optional)
(Nafion) carbonate (LiAlO2) (Y2O3, ZrO2)
Operating 50–100 50–200 ∼200 ∼650 800–1000 25
temperature
(°C)
Anode reaction H2 → 2H+ + 2e− H2 + 2(OH−) → 2H2O H2 → 2H+ + 2e− H2O + CO32− → H2O + H2 + O2 → H2O  C12H22O11 + 13H2O
+ 2e− CO2 + 2e− + 2e− → 12CO2+ 48H+ + 48e−
Cathode 1/2O2 + 2H+ + 2e− 1/2O2 + H2O + 2e− 1/2O2 + 2H+ + 1/2O2 + CO2 + 2e− 1/2O2 + 2e− 1/2O2 + 2H+ + 2e−
reaction → H2O → 2(OH)− 2e− → H2O → CO32− → O2− → H2O
Charge carrier H+ OH− H+ CO32− O2− H+
Fuel Pure H2 Pure H2 Pure H2 H2, CO, CH4, other H2, CO, CH4, other Organic carbon
hydrocarbons hydrocarbons
Oxidant Oxygen in air Oxygen in air Oxygen in air Oxygen in air Oxygen in air Oxygen in air, potassium
ferric cyanide
Efficiency 40%–50% ∼50% 40% >50% >50% >50%
Power density 3.8–6.5 ∼1 0.8–1.9 1.5–2.6 0.1–1.5 ∼1
(kW/m3)
Renewable Energy Technologies and Water Infrastructure

Installation <1,500 ∼1,800 2,100 ∼2,000–3,000 ∼3,000 ∼8,500


cost ($ US/
kW)
(Continued)
Table 6-1. Comparison of Different Fuel Cells and Their Application in Wastewater Infrastructure. (Continued)

Parameters
Fuel cell PEMFC AFC PAFC MCFC SOFC MFC
Capacity 30 W–250kW 10–100 kW 100 kW–1.3 MW 155 kW–2 MW 1 kW–1.7 MW –
Applications Pumping; aeration; Pumping; aeration; Pumping; aeration; Pumping; aeration; Pumping; aeration; Replace the aeration
lighting and lighting and heat; lighting and lighting and heat; lighting and process; Pumping;
heat; sludge- sludge-digestion; heat; sludge- sludge-digestion; heat; sludge- lighting
digestion; transportation; digestion; transportation; digestion;
transportation portable power transportation; transportation;
portable power portable power
Advantages High power High power density; Produce high- High efficiency; no Solid electrolyte; Producing energy while
density; quick quick start-up grade waste metal catalysts high efficiency; treating organic
start-up; solid heat; stable needed generate effluents, and decrease
noncorrosive electrolyte high-grade in sludge production
electrolyte characteristics waste heat
Drawbacks Expensive platinum Expensive platinum Corrosive liquid High cost; corrosive High cost; slow Nonreliable influent
catalyst; catalyst; sensitive electrolyte; liquid electrolyte; start-up; concentration; pH
sensitive to fuel to fuel impurities sensitive to fuel slow start-up; intolerance to imbalance; electrode
impurities (CO, (CO, CO2, CH4, impurities (CO, intolerance to sulfur material and
H2S) H2S) H2S) sulfur performance; membrane
fouling; low electricity
production, current
instability; high cost
Reference Kirubakaran et al. (2009) Do et al. (2018), Logan et al.
Fuel Cells for Renewable Wastewater Infrastructure

(2006)

Note: 1 kW = 3.6 MJ/h.


123
124 Renewable Energy Technologies and Water Infrastructure

Hydrogen Fuel Cells


We discuss three major types of hydrogen FCs that can be used to support
wastewater infrastructure. As discussed in Section 6.2, the hydrogen gas that is
required to drive these FCs can be sustainably generated by biological processes
or bioelectrochemical technologies.

Polymer Electrolyte Membrane Fuel Cell


Working Principle
A polymer electrolyte membrane fuel cell [also known as proton exchange
membrane fuel cell (PEMFC)] is used to produce protons and electrons at the
anode in the presence of a catalyst (e.g., platinum). The protons and electrons
generated at the anode surface flow through the electrolyte (the PEM) and the
external circuit, respectively (Figure 6-8), to produce electricity, water, and heat
at the cathode. However, PEMFCs require noble metals such as platinum as a
catalyst, which increases the initial and operating costs, in part because it is
susceptible to carbon monoxide (CO) poisoning if present in fuel (Baschuk and
Li 2001).

Figure 6-8. Schematic of a PEMFC.


Fuel Cells for Renewable Wastewater Infrastructure 125

Potential Applications in Wastewater Infrastructure


The PEMFC design offers major advantages, including low operating temperatures
(50°C to 100°C), lightweight and easy portability, high efficiency, high power
density, and low maintenance requirements. PEMFCs are suitable for commercial
applications, including use in water reclamation facilities to satisfy plant
electricity requirements. As shown in Figure 6-2, it can be seen that the PEMFC
has a wide range of power output, from 100 W to 250 kW; therefore, it can be used
for virtually every piece of equipment and process at a WRRF, from screening to
disinfection, and solids’ processing.

Alkaline Fuel Cell


Working Principle
AFCs were one of the first FCs developed and were used as the primary source
of electricity in the United States’s Apollo space program. As the name suggests,
AFCs use alkaline potassium hydroxide (KOH) as the electrolyte. Initially,
these AFCs used potassium hydroxide in water as the electrolyte, but because of
advancements in technologies in recent years, a polymer membrane (an alkaline
membrane) was developed as the electrolyte. Figure 6-9 shows a schematic of the
working principle of an AFC. These AFCs produce heat and electrical energy, in
which hydrogen and oxygen gas are fed at the anode and cathode, respectively.
The hydrogen supplied at the anode reacts with a hydroxyl ion (OH−) to generate
water and electrons. The electrons travel from the anode through an external
circuit to the cathode, in which electrons react with the supplied oxygen at
the cathode to produce water and hydroxyl ions. This hydroxyl ion migrates

Figure 6-9. Schematic of an AFC.


126 Renewable Energy Technologies and Water Infrastructure

through the electrolyte and back to the anode to sustain the anodic reaction
(Behling 2012). AFCs offer high performance and efficiencies greater than
60% when combined with heat. AFCs are operated in the temperature range
of 50°C to 200°C. However, the disadvantage with AFCs is the formation of
carbonate scaling on membrane surfaces due to the presence of dissolved
CO2. These carbonate impurities interfere with the electrochemical reactions
in AFCs, reducing the overall efficiency and power production. In the case of
liquid electrolyte AFCs, parameters including wettability, differential pressure,
and corrosion affect the performance.

Potential Applications in Wastewater Infrastructure


Because of their operating temperature range, these FCs are easy to start. These
FCs have a capacity rating of 10 to 100 kW, which makes these ideal for pumping,
aeration, lighting and HVAC, and solids’ processing at a WRRF. Because these
are easy to start, they can also be used for transportation requirements within
the facilities.

Phosphoric Acid Fuel Cell


Working Principle
PAFCs use phosphoric acid as an electrolyte. PAFCs are considered the “first
generation” of modern FCs and are the first ones commercially used. These FCs
use porous carbon electrodes containing platinum as a catalyst. The protons
generated at the anode surface flow through the electrolyte (phosphoric acid),
and simultaneously the electrons generated at the anode surface move through
an external circuit to produce electricity. The electrons and protons meet at the
cathode surface to sustain the oxygen reduction reaction (Figure 6-10).
Simultaneously, water and heat are produced at the cathode. A typical
PAFC is similar to a PEMFC, but instead of a solid polymer membrane,
phosphoric acid is used as the electrolyte. The coulombic efficiency of these
FCs is approximately 40% (as compared to combustion-based power plants
that have approximately 33%), but when combined for cogeneration of heat, the
efficiency of these cells reaches more than 85%. Because of the acidic nature of
the electrolyte, the components are susceptible to corrosion or oxidation over
time. Due to their lower efficiency as compared to other types of FCs, these
cells require a higher loading of catalysts to increase their efficiency, which, in
turn, increases the cost.

Potential Applications in Wastewater Infrastructure


PAFCs produce high-grade waste heat, which can be used for heating requirements
of biohydrogen production or digestion processes. These FCs have a capacity of
100 kW to 1.3 MW (range), which makes it ideal for pumping, aeration, lighting
and HVAC, and solids’ processing at a WRRF. The main disadvantages of these
FCs are corrosiveness of the electrolyte and sensitiveness to fuel impurities.
Fuel Cells for Renewable Wastewater Infrastructure 127

Figure 6-10. Schematic of a PAFC.

Hydrogen and Hydrocarbon Fuel Cells


In this section, we discuss two major types of FCs that run on fuels other than
hydrogen, and these FCs include carbon monoxide (CO), biomethane (which is
the major component of biogas produced in AD), and other hydrocarbons.

Molten Carbonate Fuel Cells


Working Principle
MCFCs are high-temperature FCs composed of a molten mixture of alkali metal
carbonate suspended in a lithium aluminum oxide matrix as the electrolyte, an
anode, a cathode, and an external circuit connecting the two electrodes. Figure
6-11 shows a schematic of the working principle of a typical MCFC. The high
operating temperature (600°C to 700°C) results in the formation of a highly
conductive molten salt carbonate, in which carbonate ( CO32− ) ions provide the
ionic conduction. In the anode compartment, hydrogen supplied as fuel reduces
the carbonate ions to form carbon dioxide and electrons to produce electricity. CO2
from the anode exhaust moves to the cathode compartment to react with oxygen
present in air and the electron supplied from the external circuit to form CO32−
128 Renewable Energy Technologies and Water Infrastructure

Figure 6-11. Schematic of an MCFC.

ions. The carbonate ions produced at the cathode are again converted back into
CO2 at the anode. As these are operated at a high temperature of approximately
650°C, salt becomes liquid, which enables the movement of carbonate ions and
the use of nonprecious metals as catalysts. Unlike other FCs (PEM, AFC, PAFC),
MCFCs can use carbon-based fuels, such as natural gas and biogas, which are
converted to hydrogen within the fuel cell by a process called “internal reforming,”
as shown in Equation (6-5). Because of their high efficiencies (that can reach up
to approximately 65%) and high operating temperatures, these FCs are being
currently developed for natural gas and coal-based power plants for electricity
generation and for use in industrial and military applications. Due to the presence
of carbonate ions and the high operating temperatures, the electrodes used in
these FCs are vulnerable to corrosion.
CH 4 + H2O = 3H2 + CO (6-5)

Potential Applications in Wastewater Infrastructure


MCFCs can be integrated with WRRFs as the biomethane produced during
the solids’ digestion process can be utilized as a fuel in the anode component.
Fuel Cells for Renewable Wastewater Infrastructure 129

Because of MCFCs’ capacity to generate electricity in the range of 414 MJ/h to


7.2 GJ/h (115 kW to 2 MW), they can be used to power the equipment and process
operations at a WRRF.

Solid Oxide Fuel Cells


Working Principle
SOFCs produce electricity directly from oxidizing fuels. The materials used for
all the components in an SOFC are either ceramic or metal. In an SOFC, oxygen,
which is continuously fed into the cathode compartment, undergoes the reduction
reaction to produce O2− (Figure 6-12). The electrolyte transports the produced
O2− from the cathode to the anode compartment. The fuel in the anode (usually
hydrogen or hydrocarbon) reacts with O2− to produce electrons, water, and
CO2. The electrons released at the anode pass through an external circuit to the
cathode to sustain the cathodic reaction (Huang and Goodenough 2009). SOFCs
use a nonporous ceramic compound, such as yttrium-stabilized zirconia, as the
electrolyte. These cells are operated at very high temperatures of 800°C to 1,000°C,
which eliminates the need for a precious metal catalyst. Similar to MCFCs, these

Figure 6-12. Schematic of an SOFC.


130 Renewable Energy Technologies and Water Infrastructure

SOFCs use carbon-based fuels, such as natural gas and biogas, which are converted
to hydrogen within the fuel cell by internal reforming, as shown in Equation (6-3),
which also reduces the cost. These cells have an efficiency of more than 50% in
converting fuel to electricity, but when the cogenerated waste heat is captured,
these can achieve an efficiency of approximately 85%, which is similar to MCFCs.
Figure 6-12 shows the schematic and process flow of a typical SOFC.

Potential Applications in Wastewater Infrastructure


Because of their high operating temperatures, SOFCs are limited to stationary
applications, such as residential facilities, utility power plants, commercial
cogeneration, and portable power, making them ideal for operation at a WRRF.
These FCs have a wide range of operating capacity, from 3.6 MJ/h to 6.1 GJ/h
(1 kW to 1.7 MW), which is ideal for running various WRRF/water reclamation
facility infrastructure.

Microbial Fuel Cell


The MFC is a bioelectrochemical device that converts chemical energy from
organic substrates, such as glucose, into electricity using microorganisms as the
catalyst. Generating electricity using microorganisms was first demonstrated by
Potter (1911) using Escherichia coli. Recently, the MFC has been getting good
traction in the scientific community for its potential as a sustainable wastewater
treatment method. Typical components of an MFC include an anode, a cathode,
and an ion-exchange membrane (Figure 6-13). The organic matter present in
wastewater is oxidized by electrogenic bacteria at the anode surface to generate
electrons and protons. The produced electrons pass through the external circuit to
reduce an electron acceptor in the cathode compartment, whereas the protons pass
via the ion-exchange membrane to the cathode chamber to participate in the redox
reaction. It is now well agreed that oxygen is the most suitable electron acceptor
as compared to chemicals, such as potassium ferric cyanide, which is toxic to the
environment, incurs high cost, and has scalability issues. Recently, air-breathing

Figure 6-13. Schematic of an MFC.


Fuel Cells for Renewable Wastewater Infrastructure 131

cathodes have been developed, which can utilize oxygen present in the air as the
terminal electron acceptor, eliminating the use of chemicals and reducing the
cost. Even though there are huge advancements in MFCs, their performance is
limited by nonreliable influent concentration, pH imbalance, electrode material
and performance, membrane fouling, low electricity production, current
instability, and high initial cost. Currently, there are few applications of MFCs for
commercial or large-scale use, but with increasing demand for an affordable and
efficient way to clean water with less sludge by-products, MFCs offer an attractive
alternative to existing technologies (Logan et al. 2006).

SUMMARY

Fermentation technologies provide an attractive option for generating hydrogen


from wastewater, which can drive FCs that support typical equipment and
processes operating under wastewater infrastructure. To develop such renewable
infrastructure, it is critical to identify and address technological barriers to fuel-cell
technology that is solely driven by biohydrogen. Some of the challenges will be those
associated with the development of biohydrogen infrastructure (i.e., production,
storage, and supply of biohydrogen), durability and reliability of biohydrogen
FCs, and end users’ acceptance of the biohydrogen/fuel-cell technology. A series
of technoeconomic studies are warranted to assess and compare the viability
of different fermentation technologies (DF, photofermentation, microbial
electrolysis, and biophotolysis). In the case of MFCs and MECs, technological
breakthroughs are required to address fouling challenges faced by materials
(catalysts and electrodes) and those pertaining to reactor design. Overall, the
biohydrogen/fuel-cell approach discussed in this study provides an exciting
route to minimize energy consumption by wastewater infrastructure and drive
wastewater processing operations toward a NET-ZERO-energy.

ACKNOWLEDGMENTS

The authors acknowledge the funding support from the Electric Power Research
Institute (No. 10003325), the National Science Foundation (No. 1736255), and
NASA-EPSCoR (No. NNX16A). The authors declare no conflict of interest.

References
Barbosa, M. J., J. M. S. Rocha, J. Tramper, and R. H. Wijffels. 2001. “Acetate as a carbon
source for hydrogen production by photosynthetic bacteria.” J. Biotechnol. 85 (1): 25–33.
Baschuk, J. J., and X. Li. 2001. “Carbon monoxide poisoning of proton exchange membrane
fuel cells.” Int. J. Energy Res. 25 (8): 695–713.
132 Renewable Energy Technologies and Water Infrastructure

Behling, N. 2012. Fuel cells: current technology challenges and future research needs.
Amsterdam: Elsevier.
Call, D., and B. E. Logan. 2008. “Hydrogen production in a single chamber microbial
electrolysis cell lacking a membrane.” Environ. Sci. Technol. 42 (9): 3401–3406.
Cheng, H., K. Scott, and C. Ramshaw. 2002. “Intensification of water electrolysis in a
centrifugal field.” J. Electrochem. Soc. 149 (11): 11–172.
Cusick, R. D., B. Bryan, D. S. Parker, M. D. Merrill, M. Mehanna, P. D. Kiely, et al. 2011.
“Performance of a pilot-scale continuous flow microbial electrolysis cell fed winery
wastewater.” Appl. Microbiol. Biotechnol. 89 (6): 2053–2063.
Do, M., H. Ngo, W. Guo, Y. Liu, S. W. Chang, D. D. Nguyen, et al. 2018. “Challenges in the
application of microbial fuel cells to wastewater treatment and energy production: A
mini review.” Sci. Total Environ. 639: 910–920.
DOE. 2006. Energy demands on water resources. Technical Rep. No. 201107. Washington,
DC: DOE.
EIA (Energy Information Administration). 2014. Annual energy outlook 2014: With
projections to 2040. Washington, DC: EIA.
Feeley, T. J., L. Green, J. T. Murphy, J. Hoffmann, and B. A. Carney. 2005. Department
of Energy/Office of Fossil Energy’s Power Plant Water Management R&D Program.
Pittsburgh, PA: National Energy Technology Laboratory.
Hakobyan, L., L. Gabrielyan, and A. Trchounian. 2019. “Biohydrogen by Rhodobacter
sphaeroides during photo-fermentation: Mixed vs. sole carbon sources enhance
bacterial growth and H2 production.” Int. J. Hydrogen Energy 44 (2): 674–679.
Han, S. K., and H. S. Shin. 2004. “Performance of an innovative two-stage process converting
food waste to hydrogen and methane.” J. Air Waste Manage. Assoc. 54 (2): 242–249.
Heidrich, E. S., J. Dolfing, K. Scott, S. R. Edwards, C. Jones, and T. P. Curtis. 2013.
“Production of hydrogen from domestic wastewater in a pilot-scale microbial electrolysis
cell.” Appl. Microbiol. Biotechnol. 97 (15): 6979–6989.
Huang, K., and J. Goodenough. 2009. Solid oxide fuel cell technology: Principles,
performance and operations. Amsterdam: Elsevier.
Kirubakaran, A., S. Jain, and R. K. Nema. 2009. “A review on fuel cell technologies and
power electronic interface.” Renewable Sustainable Energy Rev. 13 (9): 2430–2440.
Kruse, O., J. Rupprecht, K. P. Bader, S. Thomas-Hall, P. M. Schenk, G. Finazzi, et al. 2005.
“Improved photobiological H2 production in engineered green algal cells.” J. Biol. Chem.
280 (40): 34170–34177.
Liu, H., S. Grot, and B. E. Logan. 2005. “Electrochemically assisted microbial production
of hydrogen from acetate.” Environ. Sci. Technol. 39 (11): 4317–4320.
Logan, B., B. Hamelers, R. Rozendal, U. Schröder, J. Keller, and S. Freguia. 2006. “Critical
review microbial fuel cells: Methodology and technology.” Environ. Sci. Technol. 40 (17):
5181–5192.
Mohan, S. V., V. Lalit Babu, and P. N. Sarma. 2007. “Anaerobic biohydrogen production
from dairy wastewater treatment in sequencing batch reactor (AnSBR): Effect of organic
loading rate.” Enzyme Microbial Technol. 41 (4): 506–515.
Nath, K., M. Muthukumar, A. Kumar, and D. Das. 2008. “Kinetics of two-stage fermentation
process for the production of hydrogen.” Int. J. Hydrogen Energy 33 (4): 1195–1203.
Oh, Y., S. Raj, G. Jung, and S. Park. 2013. “Metabolic engineering of microorganisms for
biohydrogen production.” Biohydrogen, 45–65. New York: Elsevier.
Ong, M. D., R. B. Williams, and S. R. Kaffka. 2014. DRAFT comparative assessment of
technology options for biogas clean-up. Davis, CA: Univ. of California.
Fuel Cells for Renewable Wastewater Infrastructure 133

Pirne, M., and M. Yonkin. 2008. Statewide assessment of energy use by the municipal water
and wastewater sector. Rep. No 08-17. Albany, NY: New York State Energy Research and
Development Authority.
Potter, M. C. 1911. “Electrical effects accompanying the decomposition of organic
compounds.” Proc. R. Soc. Lond. Series B 84 (571): 260–276.
Shrestha, N., G. Chilkoor, J. Wilder, V. Gadhamshetty, and J. J. Stone. 2017. “Potential
water resource impacts of hydraulic fracturing from unconventional oil production in
the Bakken shale.” Water Res. 108: 1–24.
Su, H., J. Cheng, J. Zhou, W. Song, and K. Cen. 2009. “Combination of dark-and photo-
fermentation to enhance hydrogen production and energy conversion efficiency.” Int. J.
Hydrogen Energy 34 (21): 8846–8853.
USGS. 2015. “Water use data for the nation.” Accessed July 13, 2020. https://waterdata.usgs.
gov/nwis/water_use?format=html_table&rdb_compression=file&wu_year=2015&wu_
category=ALL.
Van Ginkel, S. W., S. E. Oh, and B. E. Logan. 2005. “Biohydrogen gas production from
food processing and domestic wastewaters.” Int. J. Hydrogen Energy 30 (15): 1535–1542.
Wang, J., M. Bibra, K. Venkateswaran, D. R. Salem, N. K. Rathinam, V. Gadhamshetty, et al.
2018. “Biohydrogen production from space crew’s waste simulants using thermophilic
consolidated bioprocessing.” Bioresour. Technol. 255: 349–353.
Wang, J., and W. Wan. 2009. “Factors influencing fermentative hydrogen production: A
review.” Int. J. Hydrogen Energy 34 (2): 799–811.
Wang, J., and Y. Yin. 2018. “Fermentative hydrogen production using pretreated microalgal
biomass as feedstock.” Microbial Cell Factories 17: 22.
Wang, S., and S. P. Jiang. 2017. “Prospects of fuel cell technologies.” Natl. Sci. Rev. 4 (2):
163–166.
CHAPTER 7
Sustainable Desalination
Using Renewable
Energy Sources
Veera Gnaneswar Gude

INTRODUCTION

Freshwater production through desalination processes can be achieved by two


major physical principles: (1) evaporation/condensation, and (2) separation/
filtration (Gude 2018). Evaporation/condensation processes are known as phase-
change operations, whereas separation processes are known as nonphase-change
operations. An evaporation process requires thermal energy to form pure water
vapor from a saline water source, which is being heated. This water vapor is then
condensed on a cooling surface to form freshwater. A small portion of electrical
energy is required for fluid transfer and to create pressure conditions suitable
for the evaporation process. A separation process involves a membrane, which
acts as a physical barrier to separate water molecules from saline water via
permeation or diffusion (Gude 2016a). Examples of evaporation/condensation or
phase-change processes include solar stills, multieffect evaporation/distillation
(MED), multistage flash distillation (MSF), thermal vapor compression (TVC),
and mechanical vapor compression (MVC). Separation processes may be based
on ion transfer or solvent transfer phenomena as in electrodialysis (ED) and
reverse osmosis (RO) processes, respectively. Other emerging hybrid (involving
both evaporation and separation principles) desalination processes are membrane
distillation (MD) and RO combined with MSF or MED processes (Gude et al.
2010). Phase-change desalination processes require both thermal and electrical
energy. Nonphase-change desalination processes require electrical energy only.
A comparison of energy requirements for major desalination processes is shown
in Figure 7-1.

135
136 Renewable Energy Technologies and Water Infrastructure

Figure 7-1. Comparison of energy requirements for different desalination processes.


Source: Modified after Gude (2018).

DESALINATION TECHNOLOGIES—PRINCIPLES OF OPERATION

Multistage Flash Desalination


MSF desalination process was the dominant technology in the desalination
industry until the year 2000. Its share was around 60% of the total world
desalination capacity. Currently, its share in the market is about 30% of the total
market and about 80% of the thermal desalination capacity (Gude 2016a). This
process consists of a number of flashing stages, a brine heater, pumps, a venting
system, and a cooling water control loop. Incoming seawater temperature is
increased by passing it through the final heat exchanger (El-Dessouky et al. 1999).
Then, it is passed through the brine heater where steam from an external source
supplies the energy for the process and heats the seawater in the heat exchanger to
the maximum process temperature (80°C to –90°C). Finally, it is released into the
first vacuum chamber where the water vapor condenses into a freshwater product
on the cooling water control loop, and this operation is repeated in a number of
stages to achieve energy recovery and recycling (Younos and Tulou 2005a).

Multieffect Distillation
The multieffect distillation (MED) process is different from the MSF process. The
operating pressure in the distillation effects is reduced in successive stages so
that heat recovery can be achieved in a number of stages (Gude 2015). The heat
removed in a previous stage functions as a heat source in the next stage. Feedwater
entering the first effect is heated to the boiling point. Both feedwater and heating
vapor to the evaporators flow in the same direction. The water that is remaining
is pumped to the second effect, where it is once more applied to a tube bundle.
Sustainable Desalination Using Renewable Energy Sources 137

This process continues for several stages called effects, and there are about 4 to 21
effects in a large desalination plant. In recently designed plants, the feedwater is
divided into several parts before entering the flash drums. The rest of the process
remains similar to the other design (Ghalavand et al. 2015).

Vapor Compression
In this process, the heat for evaporating the feedwater is generated by compressing
the vapor. Two methods are used to condense the water vapor and to produce an
amount of heat sufficient to evaporate the incoming feed seawater: a mechanical
compressor and a steam jet (thermal compressor). In the first method, seawater
is evaporated and the vapor is passed through a compressor (Ghalavand et al.
2015). The vapor is compressed, which leads to an increase in the vapor dew
point (in this condition, the compressed vapor dew point is higher than seawater
boiling point), so it can be condensed by indirect seawater contact, leading to the
evaporation of freshwater. The compression ratio should be maintained near unity
to increase energy efficiency. In this process, the seawater temperature is held
at 100°C (212°F). MVC units are built in a variety of configurations to promote
seawater evaporation. The compressor creates a vacuum in the evaporator and
then compresses the vapor taken from the evaporator and condenses it in a tube
bundle.

Energy Efficiency of Thermal Desalination


Thermal energy required for desalination can be provided from different heating
sources such as natural gas and steam generated from power plants. The energy
efficiency is reported either as gain output ratio (GOR) or as performance ratio
(PR). GOR is defined as the ratio of the mass of water produced through a
desalination process over a fixed quantity of energy consumed. PR is defined as
the mass, in pounds, of water produced by desalination per 1,054 kJ (1,000 Btu) of
heat provided to the process (Gude 2015; Ghalavand et al. 2015). This is equivalent
to the number of kilograms of freshwater produced per 2,326 kJ (approximately
2,206 Btu) of heat. As shown in Figure 7-1, the MED process has a high GOR
or PR compared with the MSF and VC processes. MED is also known to be
thermodynamically more efficient.

Reverse Osmosis
RO is the most commonly used technology in desalination. In this process,
the osmotic pressure is overcome by applying an external pressure higher than
the osmotic pressure on the feedwater (brackish/seawater); thus, water flows
in the reverse direction to the natural flow across the membrane, leaving the
dissolved salts behind the membrane with an increase in salt concentration
called brine. Most of the energy consumption for membrane desalination is used
for pressurizing the feedwater. A typical large seawater RO plant consists of four
major components: feedwater pretreatment, high pressure pumping, membrane
separation, and permeate post-treatment. Major design considerations for
138 Renewable Energy Technologies and Water Infrastructure

seawater RO plants are the flux, conversion, or recovery ratio, permeate salinity,
membrane life, energy consumption, and feedwater temperature. Energy
demand by membrane processes is reported as specific energy consumption in
kWh/m3 of freshwater produced (Gude 2011). The specific energy consumption
has been reduced by almost 10 times over the last two decades as a result of
developing energy recovery devices and energy-efficient and highly permeable
membranes.
Table 7-1 presents a list of desalination processes with process operating
conditions and renewable energy applicability. Hybrid technologies such as
MD and low-temperature desalination technologies need more pilot-scale
demonstrations before they can be considered for large-scale operations.

RENEWABLE ENERGY INTEGRATION WITH DESALINATION


PROCESSES

Conventionally, desalination processes are powered by energy derived from


combustion of fossil fuels, which contribute to acid rain and climate change by
releasing greenhouse gases and several other harmful emissions (Gude 2011).
Currently, the world’s established desalination capacity is around 10% of the
world’s total minimum freshwater demand [basis: a minimum of 100 L/cap/day
(26.4 gal./cap/day) for 7.8 billion of the current world population]. It requires
1.63 million tons of oil/day, which may release 179 metric tons of CO2/day. It
is clear that freshwater production through fossil fuel-powered desalination
technologies is not an environmentally sustainable alternative. It is necessary
to develop alternatives to replace conventional energy sources used in the
desalination process with renewable ones and reduce the energy requirements
for desalination by developing innovative low-cost, low-energy technologies and
process hybridizations (Gude 2015).
The large quantities of energy required for desalination processes can
be provided by various renewable energy sources, as shown in Figure 7-2. The
selection process is not simple. Some of the renewable energy technologies such
as photovoltaic (PV) and wind turbines have reached maturity, making their
application in desalination economical. It should be noted that the desalination
processes can function regardless of the source of energy. Heat energy and electricity
can be supplied from any renewable energy source. The energy production costs
can be improved through process integration and cogeneration schemes (Younos
and Tulou 2005b). Renewable energy applications are more favored in regions
where they are competitive with other energy production methods. As shown in
Figure 7-2, desalination technologies should be integrated with locally available
renewable energy sources considering factors such as plant capacity, location, and
availability of renewable energy sources. Technologies that require thermal energy
can be conveniently combined with locally available renewable energy sources such
as geothermal sources (Gude 2016b). However, a proper site and source evaluation
Table 7-1. A comparison of Various Desalination (Membrane and Thermal) Processes and Potential Applications.

Process Description Renewable energy applications Remarks


Solar still Operating temperature and pressure: Direct solar energy, solar Small and rural applications
40°C–80°C, atmospheric collectors, PVT collectors, possible, low capital and little
Source water: seawater/brackish water Solar ponds, and process maintenance costs, no energy
Recovery Rate: 35%–45% waste heat costs
Specific energy consumption: Not suitable for large-scale
5,040 kJ/kg applications because of lower
Multieffect solar Operating temperature and pressure: Direct solar energy, solar efficiency
still 40°C–80°C, 1 atm collectors, PVT collectors,
Source water: seawater/brackish water Solar ponds, waste heat
Specific energy consumption: source
5,040 kJ/kg;
MSF Operating temperature and pressure: Solar collectors and Large-scale applications, reliable
80°C–120°C, 1–2 atm geothermal source process, and experience in
Source water: seawater operations
Recovery rate: 35%–45% Can be combined with power
Specific energy consumption: generation (cogeneration)
200–350 kJ/kg; Cost and energy intensive, not
MED 50°C–90°C, 0.1–0.5 atm, Solar collectors, solar pond, suitable for small-scale
Source water: seawater geothermal source, and applications
Recovery rate: 35%–45% process waste heat
Sustainable Desalination Using Renewable Energy Sources

Specific energy consumption:


150–250 kJ/kg
(Continued)
139
Table 7-1. A comparison of Various Desalination (Membrane and Thermal) Processes and Potential Applications. (Continued)
140

Process Description Renewable energy applications Remarks


Low-temperature Operating temperature and pressure: Solar collectors, solar pond, High thermodynamic efficiency,
MED 40°C–70°C, 0.1–0.4 atm, PVT collectors, geothermal low cost, or free waste heat
Source water: seawater source, and process waste Less experience; large heat
Recovery rate: 35%–45% heat transfer areas
M/TVC Source water: seawater Solar collectors, solar pond, Low specific energy
Recovery rate: 25%–40% thermal collectors, consumption
Specific energy consumption: geothermal source, and Electrical and mechanical energy
150–240 kJ/kg process waste heat input
MD Operating temperature and pressure: Solar collectors, solar pond, High product recovery, low-
40°C–80°C, atmospheric PVT collectors, PV modules, temperature operation
Source water: seawater geothermal source, and Limited commercial applications;
Recovery rate: 35%–45% process waste heat require membranes
RO Operating temperature and pressure: PVT collectors, PV modules, Reliable and most widely used
<45°C, 20–60 atm wind energy, wave energy technology
Source water: seawater/brackish water High electricity requirements,
Recovery rate: 35%–50% (sea water), capital and O&M costs
50%–90% (brackish water)
Specific energy consumption:
120 kJ/kg
ED Operating temperature and pressure: PVT collectors, PV modules, Efficient for high-quality product
<45°C wind energy, wave energy from brackish water RO
Renewable Energy Technologies and Water Infrastructure

Source water: brackish water Highly sensitive to raw water


Recovery rate: 50%–90% (brackish water quality, pretreatment
Specific energy consumption:
144 kJ/kg
Source: Adapted from Gude (2015, 2018).
Sustainable Desalination Using Renewable Energy Sources 141

Figure 7-2. Criteria for selecting a suitable renewable energy source for various
desalination technologies.

should be conducted to prevent any environmental issues. If the region has a strong
solar insolation, concentrated solar panels (either parabolic troughs or tower type)
can be used to support even large-scale desalination operations. For membrane
technologies that primarily require electrical energy, solar PV, wind, or wave
energy sources can be considered depending on their availability (Figure 7-2).
The integration of renewable energy sources with desalination technologies
comes with some challenges. The integration of solar, wind, and wave energy
sources may require storage facilities. For example, solar energy is a self-renewable
resource that can be easily captured and harvested as thermal energy for many
beneficial uses. However, the concern with solar energy is that it is intermittent in
nature and its intensity depends on the hour of the day and local weather conditions.
One of the solutions to utilize the fluctuating solar energy on a continuous basis
is to incorporate the thermal energy storage (TES) system into it (Figure 7-2).
Energy accumulation, storage, and supply are the key elements of the TES concept
that result in better economics, resource management, and lower environmental
emissions of a variable energy source-powered desalination system (Dincer 2002).
TES helps manage the energy resource when supply and demand are mismatched.
Three types of TES systems are in commercial use: (1) sensible heat storage, (2)
latent heat storage, and (3) thermochemical storage systems (Gude et al. 2012).
The most widely used TES is the sensible heat storage system that stores energy by
142 Renewable Energy Technologies and Water Infrastructure

changing the temperature of the storage medium, that is, water, air, rock beds, or
sand. Water is cheap, chemically stable, and has a higher heat capacity compared
with many other fluids such as oils, air, or rock beds. The amount of energy stored
by a sensible heat storage system is proportional to the difference between the
storage input and the output temperatures, the mass of the storage medium, and the
medium’s heat capacity (Gude et al. 2012). The energy stored in the TES system can
be used for a variety of applications depending on the temperature of the medium;
for example, TES temperatures in the range of 60°C to 80°C (140°F to 176°F) are
suitable energy sources for low-temperature desalination and TES temperatures in
the range of 100°C to 400°C (212°F to 752°F) are suitable for power generation, as
well as cooling and other industrial process applications (Hamed et al. 2016). The
energy input can be provided by solar collectors, parabolic trough collectors, and
process waste heat releases. Similar to solar energy, wind and wave energy sources
also have high variability requiring battery energy storage. PV panels also require
battery energy storage to ensure round-the-clock operations.

SELECTION PROCESS OF THE DESALINATION PROCESS

The selection of an appropriate renewable energy-powered desalination process


requires a complete analysis of the plant from start to end. Irrespective of the
production volume, the following factors need consideration, when selecting a
desalination process (Gude et al. 2010):
1. Need for desalination:
a. capacity, and
b. quality.
2. Energy source, infrastructure, hardware:
a. Solar;
b. Wind;
c. Wave;
d. Geothermal; and
e. Nuclear.
3. Feedwater source:
a. Seawater;
b. Brackish water; and
c. Produced water (wastewater reuse).
4. Process:
a. Type of process (e.g., MED, MSF, RO, TVC);
b. Specific energy requirements; and
c. Characteristics (e.g., single stage, multistage, integrated).
Sustainable Desalination Using Renewable Energy Sources 143

5. Economic feasibility:
a. Financing;
b. Capital costs; and
c. Operating costs.
6. Environment:
a. Brine disposal;
b. Intake structures; and
c. Use of chemicals.
The selection of a desalination process depends on various factors such as the
water demand (desalination capacity), the intended use of water (product water
quality), the availability of energy sources, energy source-compatible desalination
technology and skills to operate and maintain the desalination process, feedwater
source and water quality characteristics, economic feasibility, and environmental
issues or considerations. The selection process is shown in Figure 7-3. If the water
source has a salinity of less than 2,000 mg/L, nanofiltration can be considered.
For source water with less than 5,000 mg/L salinity, ED or nanofiltration process
can be considered. If the water source salinity is higher than 5,000 mg/L, then
pretreatment and post-treatment schemes may be required.

Figure 7-3. Criteria for selecting a desalination technology.


144 Renewable Energy Technologies and Water Infrastructure

SUSTAINABILITY OF DESALINATION TECHNOLOGIES

The triple bottom line for assessing the sustainability of a desalination process
includes the critical components of environmental, economic, and social domains.
One of the critical needs of a desalination process, that is, energy, overlaps the
environmental and economic compartments. For example, the environmental
emissions from desalination processes can be minimized by using renewable
energy sources. Although renewable energy source integration may address
environmental issues partially, economic objectives may not be met completely.

Environmental Impacts of Desalination Processes


Desalination plants can have a significant impact (direct or indirect) on the
environment. Increased use of fossil fuels for desalination may accelerate air
pollution because of greenhouse gas emissions, which include carbon monoxide
(CO), nitric oxide (NO), nitrogen dioxide (NO2), and sulfur dioxide (SO2),
known to have a detrimental influence on local public health and environment
(Younos 2005). However, renewable energy integration can alleviate these issues.
For example, in a solar energy-driven desalination plant, the prime energy
consumption to produce a solar collector of 1 m2 area and the supporting frame
is about 7.0 GJ (Ardente et al. 2005). If the annual useful thermal energy harvested
by the total solar collector area is 5.5 GJ/m2 year, which is used for desalination,
then considering the pumping energy required in an active solar collector system
(between the solar collector system and the desalination system), the prime energy
payback period of the solar collector system is under 2 years (Gude et al. 2012). The
global warming potential (GWP) of a solar collector of 1 m2 area is estimated to be
around 721 kg eq. CO2. Annual CO2 eq. emission savings owing to thermal energy
generated by solar collectors is estimated as 407 kg eq. CO2 (basis: a specific global
warming factor of 0.066 kg eq. CO2 per MJ of useful thermal energy). Therefore,
the GWP of the solar collectors can be recovered in less than 2 years (Gude et al.
2012). These data support the clean idea of utilizing renewable energy harvested
by solar collectors for low-temperature desalination.

Brine Disposal
Desalination plants utilize significant amount of chemicals for pretreatment of
saline water and post-treatment of desalinated water. The concentrates generated
from brackish water via an RO process (with 60% to 85% recovery) would have
a concentration factor that is 2.5 to 7 times higher and the same for the seawater
RO (SWRO) (with 30% to 50% recovery) would be between 1.25 and 2.0 (NWC
2008). Thermal desalination technologies have a wide range of recovery ratios
between 15% and 50% with a concentration factor of 1.15 (owing to cooling water
mixing). Specific seawater concentrate properties, such as salinity, temperature,
and various chemicals used for coagulation, biocides for controlling biological
growth, antifoaming, anticorrosion, and cleaning chemicals, cause serious
environmental concerns for their proper disposal.
Sustainable Desalination Using Renewable Energy Sources 145

About 2 units of concentrate are generated for every unit of desalinated water
produced. The current practice of handling these concentrates is to discharge them
into the coastal waters, which could have detrimental effects on the aquatic life and
coastal environment (Gude 2016a). Brine/concentrate disposal technologies and
management options are not as robust and practical as desalination technologies,
and, in many cases, brine discharge poses a major threat for disapproval of new inland
desalination plants and some coastal desalination plants. Continuous recovery and
discharge of seawater at a specific location may also affect the quality of incoming
supply water for desalination in the long run. To mitigate major environmental
concerns related to brine/concentrate discharges, concentrates should be rediluted
with the seawater to minimize the effects related to high salt concentrations.
Brine disposal alternate selection requires a consideration of various factors,
as shown in Figure 7-4. These factors include the location of the plant, land
availability, and other treatment and beneficial applications. Surface discharge is
the most feasible option for coastal desalination plants. Many options are available
for inland communities, which include evaporation ponds, sewer disposal, spray
irrigation, deep well injection, and zero liquid discharge systems for valuable
resource recovery. The recovery of valuable chemicals, such as gypsum from
the concentrates, is an attractive option. The reuse of treated brackish water in

Figure 7-4. Criteria for selecting a suitable brine disposal and management alternate
for desalination technologies considering plant location and other beneficial uses.
146 Renewable Energy Technologies and Water Infrastructure

irrigation applications is another alternate to reduce the demands for freshwater


sources. Air pollution can be mitigated by using renewable energy sources. Finally,
improving the desalination process schemes would enhance the energetic and
environmental performance of the desalination processes. For instance, an SWRO
plant with a capacity of 100 million m3 (per year) would require an electrical load
of 180 GJ/h (50 MW) (Semiat 2008). A desalination and power plant cogeneration
scheme would operate at a higher efficiency than separate desalination and power
generation plants. The cogeneration scheme operation is based on the water
supply demands, which are subject to day–night and summer–winter fluctuations
resulting in an energy-efficient process scheme. Gas turbines can provide higher
efficiency with the use of high thermal energy released in off-gases, reducing the
actual energy utilization compared with other common uses.

Economic Considerations of Renewable Energy-Driven


Desalination Processes
The economics of renewable energy-driven desalination processes are not
straightforward. The economic feasibility of renewable energy application
depends on various factors such as financial packages and freshwater cost
affordability. A recent study focusing on economic analysis of various renewable
energy and desalination technologies considered solar PV, solar CSP, and nuclear
desalination technologies (Bitar and Ahmad 2017). Ten different combinations
were considered, which are as follows: Nuclear—RO, Nuclear—MED, Nuclear—
MSF, Solar—PV (RO), Solar CSP—PT (RO), Solar CSP—PT (MED), Solar
CSP—PT (MSF), Solar CSP—Tower (RO), Solar CSP—Tower (MED), and Solar
CSP—Tower (MSF). As shown in Figure 7-5, the solar PV-RO system provides the

Figure 7-5. Freshwater costs through different desalination and renewable energy
technology combinations.
Source: Data taken from Bitar and Ahmad (2017).
Note: CSP–PT = concentrated solar plant–parabolic trough.
Sustainable Desalination Using Renewable Energy Sources 147

cheapest option for desalination followed by a nuclear energy-driven RO process.


Nuclear energy-driven MED and MSF processes offer cheaper options than solar
CSP-driven MED and MSF desalination processes. However, it should be noted
that the application of nuclear energy in desalination and technological know-
how is still limited. Hybrid configurations such as PV–Wind–RO combination
are also found to be economically attractive (Gude and Fthenakis 2020). It should
be noted that the freshwater costs for various renewable energy and desalination
configurations including hybrid renewable energy schemes depend on various
factors as previously mentioned (Khan et al. 2018).

Regulatory Requirements
Desalination plant selection, design, and implementation is affected by many
regulations and permitting issues; these may vary in geographical requirements
(Younos and Lee 2019). Clean Water Act is the major act under which the National
Pollution Discharge Elimination System (NPDES) was established. In addition to
the NPDES permit requirements, other acts such as the Safe Drinking Water Act,
the Resource Conservation and Recovery Act, the Endangered Species Act, and
the water Desalination Act should be considered. More details about state-specific
regulations are discussed elsewhere (Younos and Lee 2019).

Social Aspects of Desalination Processes


Similar to the environmental impacts, the installation of desalination plants could
have significant social impacts during construction and operational phases. Social
issues for desalination technologies include a lack of public trust and confidence
in water providers, public response to dust, noise, and visual amenity impacts,
beach and ocean-based recreational activity impacts, disruption to properties and
activities along the pipeline, impact on future residential development, impact
on industrial development in the local area, impact on local economic activity,
risk to public safety, disruption of local businesses, and impact of construction
on road users (Gude 2016a). Another common issue is that communities, in
many cases, do not like to have a physical plant installed in nearby locations
or cities for various reasons, often termed as “Not In My Backyard (NIMBY).”
The construction of large desalination plants involves several inconveniences
to the local communities, including the closure of recreational facilities, local
transportation, noise and air pollution and aesthetical depreciation, and inflow
and outflow traffic transporting materials.

SUMMARY

The desalination industry has experienced tremendous growth in recent years


in response to increasing freshwater demands. Because of the continuous
improvements in desalination processes, desalination costs have declined
significantly for both thermal and membrane technologies over last few decades.
148 Renewable Energy Technologies and Water Infrastructure

Significant cost reductions can be achieved, especially for RO technology, favored


by growth rate, plant capacity, higher permeation rates, and improvements in
membrane materials. However, a further reduction in desalinated water costs
in the future is quite a challenging task despite the continuous improvements in
membrane and energy recovery technologies because the equipment, raw materials,
and energy costs are rapidly rising. Hence, a greater understanding of water scarcity
scenarios and discoveries targeting local supply issues will benefit from strategic
planning and management. Further, technological advances in novel membrane
development, process configurations, and process hybridization of membrane and
thermal processes are crucial for developments in the near future. More research
efforts should be dedicated to the development of desalination technologies such
as MD and adsorption desalination that can efficiently harvest solar energy and
waste process heat for water production. Process management schemes should
focus not only on minimizing specific energy consumption but also on enhancing
water recovery through various options. Studies focusing on technoeconomics
and environmental life-cycle impacts of renewable energy integrated desalination
technologies should be given priority to advance environmentally responsible
desalination schemes in the near future.

References
Ardente, F., G. Beccali, M. Cellura, and V. L. Brano. 2005. “Life cycle assessment of a solar
thermal collector.” Renewable Energy 30 (7): 1031–1054.
Bitar, R., and A. Ahmad. 2017. Solar vs nuclear: Which is cheaper for water desalination?
Policy Brief#2/2017. Beirut, Lebanon: AUB Policy Institute.
El-Dessouky, H. T., H. M. Ettouney, and Y. Al-Roumi. 1999. “Multi-stage flash desalination:
Present and future outlook.” Chem. Eng. J. 73 (2): 173–190.
Dincer, I. 2002. “On thermal energy storage systems and applications in buildings.” Energy
Build. 34 (4): 377–388.
Ghalavand, Y., M. S. Hatamipour, and A. Rahimi. 2015. “A review on energy consumption
of desalination processes.” Desalin. Water Treat. 54 (6): 1526–1541.
Gude, V. G. 2011. “Energy consumption and recovery in reverse osmosis.” Desalin. Water
Treat. 36 (1–3): 239–260.
Gude, V. G. 2015. “Energy storage for desalination processes powered by renewable energy
and waste heat sources.” Appl. Energy 137: 877–898.
Gude, V. G. 2016a. “Desalination and sustainability—An appraisal and current
perspective.” Water Res. 89: 87–106.
Gude, V. G. 2016b. “Geothermal source potential for water desalination—Current status
and future perspective.” Renewable Sustainable Energy Rev. 57: 1038–1065.
Gude, V. G. 2018. “Use of exergy tools in renewable energy driven desalination systems.”
Therm. Sci. Eng. Prog. 8: 154–170.
Gude, V. G., and V. Fthenakis. 2020. “Energy efficiency and renewable energy utilization
in desalination systems.” Prog. Energy 2: 022003.
Gude, V. G., N. Nirmalakhandan, and S. Deng. 2010. “Renewable and sustainable
approaches for desalination.” Renewable Sustainable Energy Rev. 14 (9): 2641–2654.
Gude, V. G., N. Nirmalakhandan, S. Deng, and A. Maganti. 2012. “Low temperature
desalination using solar collectors augmented by thermal energy storage.” Appl. Energy
91 (1): 466–474.
Sustainable Desalination Using Renewable Energy Sources 149

Hamed, O. A., H. Kosaka, K. H. Bamardouf, K. Al-Shail, and A. S. Al-Ghamdi. 2016.


“Concentrating solar power for seawater thermal desalination.” Desalination 396: 70–78.
Khan, M. A., S. Rehman, and F. A. Al-Sulaiman. 2018. “A hybrid renewable energy system
as a potential energy source for water desalination using reverse osmosis: A review.”
Renewable Sustainable Energy Rev. 97: 456–477.
NWC (National Water Commission). 2008. Emerging trends in desalination: A review.
Waterlines Rep. Canberra, Australia: NWC.
Semiat, R. 2008. “Energy issues in desalination processes.” Environ. Sci. Technol. 42 (22):
8193–8201.
Younos, T. 2005. “Environmental issues of desalination.” J. Contemp. Water Res. Educ.
132 (1): 11–18.
Younos, T., and J. Lee. 2019. “Desalination: opportunities and challenges.” Water Environ.
Technol. 22–27.
Younos, T., and K. E. Tulou. 2005a. “Overview of desalination techniques.” J. Contemp.
Water Res. Educ. 132 (1): 3–10.
Younos, T., and K. E. Tulou. 2005b. “Energy needs, consumption and sources.” J. Contemp.
Water Res. Educ. 132 (1): 27–38.
CHAPTER 8
Geothermal Energy
Audrey Angelos, Guangdong Zhu

GENERAL DESCRIPTION

Geothermal energy refers to the heat of the earth, which along with the decay of
radioactive isotopes has been in existence since the formation of the earth. This
heat is greatest at the earth’s molten core, creating a thermal gradient from the
center of the earth to the surface. Because of this gradient, high temperatures can
be obtained by drilling below the earth’s surface. Subsurface temperatures vary
depending on the location (Figure 8-1) (Glassley 2014). Geothermal resources
at different temperatures are used for different applications, including power
generation, direct heating and cooling, mineral recovery, and energy storage.
Today, the most prominent uses of geothermal energy are power generation and
direct heating and cooling.

Figure 8-1. Subsurface temperatures across the United States.


Source: DOE (2019).

151
152 Renewable Energy Technologies and Water Infrastructure

Figure 8-2. Cross section of a hydrothermal reservoir.


Source: DOE (2019).

GEOTHERMAL RESOURCES

A naturally occurring geothermal resource is called a hydrothermal reservoir.


An ideal site for a hydrothermal reservoir requires heat, permeability, and water.
A geothermal reservoir is formed when sufficient amounts of rainwater and
snowmelt enter into and are heated by underground thermal aquifers (Figure 8-2)
(Blodgett 2014). These types of reservoirs are the conventionally used methods
for harnessing geothermal power, because they are naturally occurring. The
heated liquid from the geothermal reservoir can be extracted and used either as
the working fluid or as the heat transfer fluid to a separate the working fluid, in
a variety of applications. In addition to power production, the geothermal brine
from hydrothermal reservoirs can be used for mineral recovery.

GEOTHERMAL RESOURCE APPLICATIONS

Geothermal energy is used for different applications depending on the resource


temperature. These applications can be split into two main categories: power
generation and direct use (EIA 2018). Power generation using a geothermal power
Geothermal Energy 153

plant takes on different forms depending on the temperature of the geothermal


resource used. Direct uses are also dependent on the resource temperature. Hot
geothermal resources (between about 170°C and 370°C) can be used for hydrogen
production, while cooler geothermal resources (between 50°C and 150°C) can be
used for a variety of applications such as building heating and cooling, cement
and aggregate drying, heat pumps, and green housing (Figure 8-3).

Figure 8-3. Various applications for geothermal energy based on the temperature
of the resource.
Source: DOE (2019).
154 Renewable Energy Technologies and Water Infrastructure

GEOTHERMAL POWER GENERATION

A geothermal power plant takes advantage of geothermal energy by converting


it into electricity through a thermodynamic cycle. Geothermal brine can be used
directly in this cycle, or a separate working fluid can be used to drive the turbine
and generator, converting thermal energy into electricity.
There are two aspects of geothermal electricity production that make it a
promising and unique form of energy production. First, it requires no fuel cycle
to generate heat, making geothermal a more renewable option than fossil-fueled
power plants, biomass reactors, and nuclear reactors. Second, it provides a reliable,
constant source of energy, making geothermal more accessible to the electricity
grid (and, therefore, more economically viable) than other renewable energies that
can only provide intermittent power, such as wind and solar photovoltaic sources.

POWER PLANT TYPES

Three main types of geothermal power technologies are being currently used: dry
steam, flash, and binary (see Figure 8-4 for schematic diagrams). The worldwide
locations of these different types of power technologies can be seen in Figure 8-5.
The system chosen is dependent on the temperature of the geothermal
reservoir used (see Table 8-1 for temperature ranges for each system type). Each
system will be explained in detail in the following sections.

Dry Steam
A dry steam (or supercritical) system requires a reservoir temperature above the
critical point of water (374°C). Because of the high temperatures, these geothermal

Figure 8-4. Simplified schematic drawings of the different technologies for


geothermal power plants.
Source: DOE (2019).
Geothermal Energy 155

Figure 8-5. Distribution of worldwide geothermal power plant technologies.


Source: EU Science Hub (2019).

Table 8-1. Temperature Ranges for Different Power Block Systems.

System type Dry steam Flash Binary

Resource temperature (°C) >374 180–374 107–182

Source: Renewable Energy World (2019).

reservoirs provide the greatest amount of energy per kilogram of fluid extracted. In
this system, the high-pressure steam from the reservoir decompresses as it moves
toward the surface and results in little separation of liquid from steam because of
the high temperature. This high-energy steam is sent through a turbine to produce
mechanical energy. Because there is little or no liquid to remove from the steam,
the design of the plant is simple, with only a separator necessary for removing any
particles (Glassley 2014). This simple and less expensive design may be offset by the
higher cost of drilling deeper wells to extract the supercritical geothermal brine.

Flash
Flash technology systems make use of wet steam geothermal resources (at slightly
lower temperatures). The geothermal fluid flashes to steam, either during its ascent
from the reservoir or above the surface. The steam is separated from the high-
pressure, high-temperature fluid and delivered to a high-pressure turbine. The
liquid is either flashed again at lower pressure and sent to a low-pressure turbine
(in a double-flash system) or it is condensed and injected back into the reservoir
(in a single-flash system). If sent to a low-pressure turbine, the steam exiting this
turbine is also condensed and reinjected. It is important to note for this system
that the vaporization of the geothermal brine (as it passes through the turbine)
results in some emissions, therefore depleting the geothermal reservoir over time.
156 Renewable Energy Technologies and Water Infrastructure

Organic Rankine Cycle/Binary


The final conventional technology used in geothermal power plants is the binary,
or organic Rankine, cycle. A binary cycle can make use of lower-temperature
geothermal resources, because the geothermal brine transfers its thermal energy to
a (often organic) working fluid with a lower boiling point than water. This working
fluid is vaporized and sent through a turbine, and the geothermal fluid is injected
back into the reservoir, leading to zero emissions of possible contaminants in the
geothermal fluid and less depletion of geothermal reservoir resources. Because the
working fluid has a lower boiling point, air cooling is often sufficient to accomplish
condensation at the turbine exit. Although this allows for more flexibility at the
location of binary-cycle plants (no water source needed), it also causes the plants
to be more sensitive to ambient air temperature changes (at higher ambient air
temperatures, the efficiency and power output is lower) because of the relatively
small temperature difference between the turbine inlet and the outlet.

CURRENT STATUS

Worldwide Capacity
According to the Global Status Report on Renewables conducted in 2019,
approximately 0.5 GW of new geothermal power capacity was added in 2018,
bringing the global total to about 13.3 GW. Turkey and Indonesia were the leaders
in new geothermal power installations. The countries with the largest geothermal
power capacity total are the United States, Indonesia, and the Philippines (the top
10 countries can be seen in Figure 8-6) (REN21 2019).
Technical potential for approximately 240 GW of electricity generation is
available from hydrothermal resources. The Geothermal Energy Association
predicts that by 2021, the global geothermal industry will reach 18.4 GW (GEA

Figure 8-6. Geothermal power capacity and additions by country.


Source: REN21 (2019).
Geothermal Energy 157

Figure 8-7. Operating capacity of geothermal power plants by system type.


Source: GEA (2016).

2016). This number does not consider the increase in the installed capacity that could
occur with an increase in enhanced geothermal system (EGS) implementation.

Technological Distribution
Based on data collected in 2016, dry steam systems make up a quarter of the installed
global geothermal capacity. Flash technologies (single, double, and triple) make up
a little more than half (58%) of this capacity, and binary plants make up about 16%
(Figure 8-7). Although binary systems make up the smallest percentage of installed
capacity, it is a relatively new technology that takes advantage of lower-temperature
geothermal resources and, therefore, is expected to grow in its installed capacity in
the future (GEA 2016). Turkey exemplifies the trend of increased binary-cycle plants.
It has the fourth largest installed capacity of geothermal power and the largest new
capacity under construction in 2018. All Turkey’s new geothermal plants under
construction use binary-cycle technology, and a majority of its installed geothermal
plants use binary-cycle technology as well (REN21 2019).

GEOTHERMAL DIRECT USE

Geothermal resources can be used directly as well as in a thermodynamic cycle.


Using geothermal energy for heating applications instead of traditional fossil fuel
sources can decrease costs and detrimental environmental impacts associated with
traditional heating systems. Geothermal energy is also being used in innovative
new ways beyond heating and electricity production, such as for mineral recovery
and energy storage.
158 Renewable Energy Technologies and Water Infrastructure

DIRECT HEATING APPLICATIONS

The most common direct use of geothermal resources is district and space heating.
District heating systems use multiple wells and a series of pipes to distribute
hot geothermal brine to a group of homes or buildings. Space heating typically
uses only one well per home or structure (DOE Office of Energy Efficiency
and Renewable Energy 2004). Space heating is usually accomplished with a
geothermal heat pump. A geothermal heat pump uses a constant temperature
ranging between 10°C and 16°C at about 3 m (10 ft) underground to heat and cool
buildings. A geothermal heat pump has three parts: the heat exchanger, the pump
unit, and the ductwork for air delivery.
In winter, the heat pump transfers heat from the ground to a building by
pumping heat from the heat exchanger into the ductwork. In summer, the reverse
process happens, and the heat pump removes heat from the indoor air and pumps
it back into the heat exchanger. This extra heat can be used to heat water as well.
Making use of the thermal energy in the ground allows geothermal heat pumps
to be much more energy efficient, environmentally conscious, and cost-effective
than conventional heating systems (NREL, n.d.).
Geothermal heat is also used in aquaculture (fish farming) and horticulture
(greenhouse farming). Using geothermal resources for heat rather than traditional
sources can save up to 80% of fuel costs (DOE Office of Energy Efficiency and
Renewable Energy 2004). In addition to agrobusiness applications, geothermal
brine can be piped under roads and sidewalks to melt snow. This natural and
renewable heat resource can be applied in many different settings to save money
and reduce the impact on the surrounding environment.

MINERAL RECOVERY

Geothermal brine can contain critical materials such as rare-earth elements and
lithium, which are materials necessary for many renewable energy technologies
(solar panels, wind turbines, and battery storage). Recovering these critical
materials from geothermal brine would reduce the United States’ dependence
on foreign nations for importing materials as well as potentially decreasing the
cost of renewable energy technologies (DOE Office of Energy Efficiency and
Renewable Energy 2015).
The chemical composition of geothermal brine varies widely, and the
elements of interest are often present only in trace levels. Many types of extraction
technologies exist that have already been implemented or piloted for commercial
use. These include packed column extraction, reverse osmosis, evaporation and
precipitation, inorganic substrate use, electrowinning, and oxidation. Mineral
recovery often involves multiple steps and multiple technologies to capture all
minerals of interest in a cost-effective way (Wall 2016). Although mineral recovery
of geothermal brines has been ongoing since geothermal energy use began, new
Geothermal Energy 159

technologies are in development to allow for the recovery of more valuable and
trace-level elements.

ENERGY STORAGE

With the increased energy demand worldwide and in the United States,
dispatchability is an important quality for power plants. Geothermal wells provide
a new opportunity for energy storage. Compressed-air energy storage (CAES)
technology is one example. In CAES, excess electricity is used to compress air
underground, which can then be released to turn a turbine during peak demand
times. With current CAES applications, a specific geological formation is needed,
limiting its use geographically. Using depleted shale gas wells to store compressed
air as well as waste heat from the compression process increases the geographical
opportunities for CAES (NREL 2018).
Besides CAES, the very nature of storing heat in a geothermal reservoir can
provide valuable dispatchability to the future grid with high nonflexible renewable
penetration.

FUTURE DIRECTION

Enhanced Geothermal Systems


EGS is a developing technology that allows geothermal reservoirs to be created at
sites that have sufficiently hot rocks but lack permeability or fluid saturation. For an
EGS, fluid (often water) is injected into the subsurface to cause pre-existing cracks
to reopen and increase permeability. With increased permeability, the injected
fluid can circulate through the hot rock, creating a human-made geothermal
reservoir (DOE Office of Energy Efficiency and Renewable Energy 2012).
EGS has the potential to greatly expand the use of geothermal energy, as
it requires only hot rock in the subsurface, making many more sites viable for
EGS than for hydrothermal systems. EGS can make use of binary cycles so that
the injected water and chemicals will be completely recycled and the working
fluid will require a lower thermal input. More than 100 GW of electrical capacity
is available that could be economically viable in the United States alone, which
would lead to a 40-fold increase in current geothermal power generating capacity
(DOE Office of Energy Efficiency and Renewable Energy). EGS could allow for the
construction of geothermal power plants in areas where there are not necessarily
already geothermal reservoirs, leading to increased flexibility in plant location
(IRENA 2017).
Geothermal direct heating and cooling using EGS could also be expanded to
geographical areas lacking traditional hydrothermal reservoirs.
160 Renewable Energy Technologies and Water Infrastructure

Figure 8-8. Correlating weaknesses and strengths between geothermal energy


and CSP.

Hybridization Opportunities with Concentrating Solar Power


Although geothermal power plants have many strengths when compared with
other renewable energy technologies, their weaknesses still prevent them from
being as competitive on the electrical grid as fossil fuel plants. The efficiency of
geothermal plants, in general, decreases with higher ambient temperatures; the
resources can deplete over time, and their efficiency is lower than other renewable
technologies such as concentrating solar power (CSP). To overcome some of
these weaknesses, geothermal energy can be coupled with CSP to create a more
economically competitive and a more efficient power plant.
Both geothermal power and CSP have operating strengths and weaknesses. The
operating strengths of geothermal power coincide well with CSP weaknesses, and
vice versa, creating an opportunity for synergistic hybridization. The weaknesses
and corresponding strengths of both systems are outlined in Figure 8-8.
Hybridization between geothermal and CSP is further encouraged by their
shared thermodynamic cycle, which allows for the use of the same power block
equipment. Sharing equipment means lower capital costs. In addition, there are
many locations throughout the world where geothermal and solar resources
coincide. Hybridizing these two renewable power sources creates the potential for
a base load, flexible, renewable electricity source that can meet current electricity
market demands at a competitive price.

PERSPECTIVE

Although geothermal energy use began in the early 1900s with the first geothermal
power plant, successfully harnessing geothermal energy is still a commercially
challenging field undergoing many improvements and changes (Blodgett 2014).
Power generation using geothermal energy is expanding and increasing because
Geothermal Energy 161

of new efforts at implementing EGS and binary power plants requiring lower
geothermal resource temperatures. Innovative direct-use applications for
geothermal are still being discovered and improved. With these new technological
advancements in power generation and direct use, geothermal energy has the
potential to solve many of the world’s energy problems.

References
Blodgett, L. 2014. “Geothermal 101: Basics of geothermal energy.” Geotherm. Energy
Assoc. Accessed September 19, 2021. https://geothermal.org/sites/default/files/2021-02/
Geothermal_101-Basics_of_Geothermal_Energy.pdf
DOE. 2019. Geovision: Harnessing the heat beneath our feet. Washington, DC: DOE.
DOE. 2004. Geothermal technologies program: Direct use. Washington, DC: DOE.
DOE. 2012. “What is an Enhanced Geothermal System (EGS)?” Fact Sheet. Accessed
September 19, 2021. https://www1.eere.energy.gov/geothermal/pdfs/egs_basics.pdf.
DOE. 2015. Mineral recovery creates revenue stream for geothermal energy development.
Washington, DC: DOE.
DOE. “How an Enhanced Geothermal System Works.” Accessed September 19, 2021.
https://www.energy.gov/eere/geothermal/how-enhanced-geothermal-system-works
EIA (US Energy Information Administration). 2018. Geothermal explained: Geothermal
power plants. Washington, DC: EIA.
EU Science Hub. 2019. “Earth’s geothermal hotspots: new dataset launched.” Accessed
September 21, 2021. https://ec.europa.eu/jrc/en/news/earths-geothermal-hot​spots-new-
data​set-launched.
GEA (Geothermal Energy Association). 2016. “2016 Annual U.S. & Global Geothermal
Power Production Report.” Accessed on September 22, 2021. https://www.eesi.org/
files/2016_Annual_US_Global_Geothermal_Power_Production.pdf.
Glassley, W. E. 2014. Geothermal energy: Renewable energy and the environment. 2nd ed.
Boca Raton, FL: CRC Press.
IRENA (International Renewable Energy Agency). 2017. “Geothermal power: Technology
brief.” Proc. IEEE 89: 111–113. Accessed September 19, 2021. https://www.irena.org/
publications/2017/Aug/Geothermal-power-Technology-brief.
NREL (National Renewable Energy Laboratory). n.d. Geothermal heat pump basics.
Golden, CO: NREL.
NREL. 2018. Geothermal technologies could push energy storage beyond batteries. Golden,
CO: NREL.
Renewable Energy World. 2019. “Geothermal electricity production.” Accessed September
19, 2021. https://www.renewableenergyworld.com/types-of-renewable-energy/tech-3/
geoelectricity/#gref.
REN21. 2019. “Renewables 2019 global status report.” Accessed September 19, 2021.
https://www.ren21.net/wp-content/uploads/2019/05/gsr_2019_full_report_en.pdf.
Wall, A. 2016. “Mineral recovery from geothermal brines: Resources, technologies, and
economics.” In Presented at California Geothermal Forum, Sacramento, CA. Golden,
CO: NREL.
CHAPTER 9
Wind Energy—Increasing
Resilience in Water
Infrastructure
Pamela A. Menges

INTRODUCTION

Resilience in water infrastructure is a fundamental expectation and requirement


because water is a critical resource.
The requirement to grow resiliency in critical resources and infrastructure
is fundamental as new technologies are integrated into their management
and transportation. The greater the degree of autonomy in the management
of resources, the greater the potential for resiliency. This is the point of ideal
convergence between water and wind energy.
Requirements to increase resiliency in water infrastructure have been
readily acknowledged, and because the relationship between energy and water
becomes better understood, new sources of reliable energy must be a part of the
solution. Loss of fuels owing to natural disasters, conflict, and infrastructure
failures makes renewable energy a critical component of any water resiliency
program. Furthermore, increasing issues around water safety, natural disasters,
and declining freshwater availability due to decaying infrastructure demonstrates
that, at any given time, nearly 2 billion people internationally have no reliable
access to freshwater. The National Academy of Engineering acknowledges that
the coupling of water management and treatment with wind energy sources will
advance resiliency and access to freshwater, even if most applications have been
limited to small water treatment plants and off-the-grid installation (NAE 2016).
Wind energy and hybrid wind–solar technologies create reliable, economical,
and, in some cases, portable sources of energy for water purification, desalination,
storage, and delivery. These technologies also offer greater resiliency and
security by creating microgrid support systems within critical water resource
management.

163
164 Renewable Energy Technologies and Water Infrastructure

Water security and safety is currently one of the greatest challenges


exhibited by engineered infrastructure. Growing populations, climate change,
and increasing urbanization necessitate requirements for augmenting freshwater
supplies and developing alternate water resources.
Energy, and most important “reliable energy,” is the underlying issue required
to solve water supply and management challenges. In the last 20 years, renewable
energy technologies have grown in use for increased resilience in engineered
systems, via improving reliability through the elimination of fuel as a primary
energy source.
These reliable and resilient renewable energy sources must be considered for
integration in the treatment and transport of water, as well as achieving water
security and safety.
Instances of power outrages owing to natural disasters, infrastructure
failures, and fuel losses are amplifying in frequency. The United States and other
technologically developed countries have experienced power outages that directly
affected the ability of not just rural but also urban areas to obtain fresh potable
water. The ability to respond to water emergencies associated with infrastructure
collapses is an emerging necessity in urban areas with aging water infrastructure
and growing populations.
This is occurring in developing countries, some with as many as 1 billion
people without safe, reliable drinking water. The lack of safe and reliable water
supplies will limit development, because water is the most universal component
for providing economic prosperity both industrially and technologically.
Urban areas, like Flint, Michigan, have become water-subsistent communities
because of poor decisions concerning existing infrastructure. This is a significant
health threat, and the cost of transporting potable water to communities by truck
is by some estimates hundred times more expensive than repairing damaged or
dangerous infrastructure (Lawler 2016).
This has been an epiphany for many public administrators. Hence, the drive
to increase the resilience not just in water safety and security but also in improving
access under less-than-ideal conditions, including mobile water treatment units
and basing water storage systems.
Because a greater understanding of the relationship between energy and
water is established, concerns for the types of energy that would create the
resilience required for the challenges ahead must be a central component to the
solution.
Since natural disasters, economic upheaval, and conflict all limit access to
fuels, renewables become a clear solution for mitigating adverse consequences.
Wind has been used for nearly two centuries to move water from wells and
holding basins and continues to possess extensive agricultural applications.
Hybrid systems integrating wind and solar not only offer solutions to augment
existing water infrastructure during periods of nominal performance but also act
as an effective solution to supporting water infrastructure during electrical grid
outages and additional emergencies.
Wind Energy—Increasing Resilience in Water Infrastructure 165

WIND TURBINE TECHNOLOGIES

To apply wind energy technologies to different water management requirements,


an understanding of the primary types of wind turbines is required.
Two primary types of wind turbines include, as seen in Figures 9-1 and 9-2,
horizontal axis wind turbines (HAWTs) and vertical axis wind turbines (VAWTs).
HAWTs are classified as such by the design of the axis of blade rotation with
respect to the ground and parallel to the wind, whereas in VAWTs, the axis is
vertical to the ground and perpendicular to the wind. Because wind has normally
three-axis components, these definitions are based on the prevailing or coupling
wind that imparts force on the rotors, horizontally or vertically.
Both HAWT and VAWT technologies offer advantages for different applications.
Where, HAWT technologies are, in general, more efficient, they require more
maintenance, more expensive installation owing to the height of the pole to meet
rotor dimensions, and safety considerations including a cautious turbine failure
projectile zone (TFPZ) in the case of a rotor failure, bird strikes, and noise or
vibration imparted to surrounding structures.
VAWT technologies are divided into spire or Darrieus type and drag or
Savonius type. Darrieus turbines have advantages over HAWT ones and some
Savonius types but are subject to main bearing failures, stability issues during
installations, and turbulence-induced damage as much as HAWT technologies.
The Savonius type was not considered the optimum choice for many years because
of induced drag issues and low efficiencies in the 12% to 19% range (Menges 2016).

Figure 9-1. Right HAWT source US Department of Energy, and Left VAWT Darrieus
wind turbine once used in the Magdalen Islands.
Source: Wiki Commons.
166 Renewable Energy Technologies and Water Infrastructure

Figure 9-2. Image of a High-Lift VAWT—Experimental Savonius used in plant IoT


wireless and virtual security applications. Advances in hybrid VAWT technologies,
including a new class of reduced drag Savonius, are creating more efficient
installations with highly reliable solar-wind hybrid uninterruptable power supply
(UPS) products. The new biomimetic derived high-lift or Hy-VAWT Savonius is the
most tolerant of not just turbulence, but also high-wind and extreme wind changes.
Its stability, reliability, and improved performance are indications that Hy-VAWT
may be the ideal technology for built and remote environments.
Source: Reproduced with permission of Star Sailor Wind, LLC.

However, advances in technologies have revived the Savonius, such as the built
environment wind turbine (BEWT), a micro grid that obtains remote wind
where reliability and safety issues are critical. As seen in Figure 9-2, these new
high-lift VAWT (Hy-VAWT) technologies employ air dams, mechanical flaps,
and biomimetic elements to increase efficiency, while maintaining rugged and
reliable mechanical systems. Some of the new Hy-VAWT technologies are now
approaching 25% to 29% efficiencies with ultra-high-wind tolerance and providing
enhanced safety (Menges 2016).
HAWT technology is the most commonly used in utility and high-
power applications. VAWT small wind turbines are growing in use in built
environments because of their simplicity and safety. New VAWT technology
(including purpose-built wind turbines) is leading the way in a new class of the
BEWT that focuses on safety, quiet operation, reliability, and appropriateness for
installation on built structures where low vibration, noise, and increased rotor
safety are critical.
HAWT technology requires a significant investment during installation,
especially for larger turbines, where often hundreds of meters of concrete are
used to support towers and sophisticated directional and pitch controls of the
rotors, hence requiring more maintenance cycles and planning for blade failures.
HAWT systems installed in built environments require significant structural
supports particularly when located on buildings. Other factors primarily
associated with HAWT or propeller wind turbines include noise, effects on birds,
and pitch controls often supported by tailstocks and other devices that increase
required clearances and safety.
Wind Energy—Increasing Resilience in Water Infrastructure 167

Safety procedures during the installation of HAWT or propeller wind turbines


include a study that determines the turbine failure projectile zone (TFPZ). This is
primarily an HAWT concern, particularly associated with increased turbulence
and icing conditions. However, gusting owing to compression of wind in congested
urban areas and variability of wind flow around inner city areas creates a unique
challenge to siting and risk analysis as well.
The advantage of HAWT technology is higher efficiency, which often
approaches 36% to 40% (Saidur et al. 2011). This, in turn, provides a higher power
coefficient based on the mechanical power output of the turbine divided by the
power of the free-stream air through the cross-sectional area of the rotor normal
to the free-stream direction.
VAWT technologies offer unique efficiencies when installed in configurations
that allow for higher-density installations. Whittlesey et al. (2010) found that
specific geometric spacing of VAWT arrays increases the individual turbine’s
power coefficient when placed in proximity to other VAWTs. These configurations
yield much higher power densities for a given area of land as compared to that of
HAWTs. A flow model of the interactions of wind flow between the VAWTs was
developed to investigate the effect of changes in the spatial arrangement of VAWTs
on the array performance coefficient. The array performance coefficient compared
the expected average power coefficient of turbines in an array to a spatially isolated
turbine. The geometric arrangement was “based on the configuration of shed
vortices in the wake of schooling fish is shown to significantly increase the array
performance coefficient based upon an array of 16 × 16 wind turbines. The results
suggest increases in power output of over one order of magnitude for any given
area of land as compared to HAWTs” (Whittlesey et al. 2010).
Other studies have demonstrated that biomimetic devices integrated within
VAWT rotor blades create increased lift with reduced incidence of drag on the
advancing blades. This innovation is one of several aiding the efficiency and value
of VAWTs for purpose-built applications in remote and resilient power. These
purposely designed Hy-VAWT technologies increase the average power generated
over the course of the year with improved reliability over HAWT technologies
(Menges 2016).
Another capability of the small 400 to 3,000 W VAWT and Hy-VAWT
technologies includes one preferred option that is mobility. They can be
transported where and when needed. This is particularly important in supporting
water security and establishing the concept of a water information network. This
wind information network, which includes environmental communications, has
been included in patents associated with Hy-VAWT wind turbines (Menges 2016).

WIND TECHNOLOGY POWERING MECHANICAL SYSTEMS

Two of the oldest wind turbine designs have been used for centuries to control
water infrastructure by mechanically powering pumps, gates, and weirs. The
168 Renewable Energy Technologies and Water Infrastructure

VAWT Savonius is still used, often in tandem with gears or flywheels to power
the movement of gates and weirs. This is still an area of application in certain
European countries.
The Aermotor, as seen in Figures 9-3 and 9-4, is a type of wind machine
specifically designed to function as a windmill producing mechanical power. This
type of windmill produces rotational energy through wind when the windmill
wheel (rotor disk) rotates. Directional control is supported by a tailstock. The
wheel assembly is attached to a hub, which drives gears that convert the rotary
motion to a pump rod that oscillates to move water from a well casing. Each
upstroke raises water from the well or cistern.
Where the Aermotor has not been used for generating electricity, some
customers have added electrical generators. Further, many Aermotor users have
created hybrid solar-mechanical solutions by employing solar to support electric
pumps when wind is less available. This is an integrated approach that many new
products developed as BEWT systems currently employ.
New applications of Savonius VAWTs have these high-torque machines
supporting mechanical power systems to move water and weir gates, screw
pumps, and other water control devices. One current design utilizes gears and

Figure 9-3. Image of an Aermotor windmill. Water is mechanically pumped from


wells, cisterns, and holding tanks. This practice has been in vogue since 1888 in
agricultural and small municipalities, with some reliability.
Source: Reproduced by permission of Aermotor Windmill Company.
Wind Energy—Increasing Resilience in Water Infrastructure 169

Figure 9-4. Diagram of an Aermotor pumping installation.

flywheels storing wind energy when it is most available. The high-torque nature of
most Savonius wind turbines creates a sustainable opportunity for moving higher
mass structures and gears.
In the application of renewable energy technologies in water infrastructure,
an ensemble of renewable technologies may be the most efficient and
successful application. So, moving beyond the concept of renewable energy as a
complementary or gap-filling generation technology, a new approach integrating
these technologies can establish hybrid wind and solar or integrated wind-solar
platforms for more than a secure backup. Hybrid technologies can augment
traditional utility grids by creating local and dedicated micro grids that offer
increased security and safety for critical infrastructure such as water. As a local
power system, the hybrid wind-solar micro grid can actually reduce costs by
reducing interconnected electrical transport over larger geographical areas.
Further, these hybrid systems are central to building wireless networks for water
infrastructure surveillance, increasing water security and safety.
In studying the use of wind power in the storage and distribution of water,
we can look to antiquity to see that wind was a reliable component for the control
of water.
Wind technology is millennia old, and extensive research has demonstrated
that wind power was used for much more than milling and water pumping in
170 Renewable Energy Technologies and Water Infrastructure

antiquity. In the Nile Valley during the Dynastic Period, wind-powered weir gates
appeared to have been employed along with some earth forming turning fluvial
flatlands into holding basins during extended seasonal wet periods for water
storage (Butzer 1975).
Wind-powered water transport has been utilized in agriculture for centuries.
New purpose-designed wind turbines offering high-torque and robust high-wind
design have also been emerging in recent years. These VAWT and Hy-VAWT
technologies are now implemented to control mechanical power systems in
demonstration projects and are, in general, applied to irrigation systems. Wind
energy stored from the mechanical movement of water from wells, holding tanks,
and irrigation ditches, in turn, becomes the potential energy in the storage water
to further control weir gates, doors, and valves. The hydrostatic force of the water
is utilized for energy to control the movement of the water itself.

EFFICIENCY AND THE BETZ LIMIT

Currently, there is no perfect wind technology. Objectively, HAWTs can achieve


36% efficiency but has limitations in the location and areas of installation,
maintenance, and concerns for safety in populated and urban locals.
In general, VAWTs offer 20% to 29% efficiency and have, depending on design,
lower maintenance and higher torque, which are important for mechanically
powered water systems. New purpose-designed Hy-VAWTs have efficiencies
approaching 28% with low maintenance and noise emissions, high torque, and
high wind tolerance and are safe for wildlife and humans (Menges 2016). One
caveat, the installation of a new Hy-VAWT requires elevations and clearances that
are unique to small VAWTs. They are currently limited by size, but they can be
stacked, generating up 36 kW per tower that may be installed in configurations
designed to optimize power production and output (Whittlesey et al. 2010).
Energy conversion is a central problem in wind energy. Any time energy
is converted from one form to another, losses occur. These losses vary based on
the type of energy conversion and the form of energy employed. Wind energy
has losses specifically associated with the efficiency by which a wind turbine
captures the energy of the wind. The losses associated with drag, turbulence,
aerodynamic flutter, and vibration are owing to the proximity of other structures,
and mechanical system losses.
Mechanical losses in HAWT systems are not unusual and are a by-product
of the advanced pitch and performance control systems. Many manufacturers of
large HAWT systems actually augment performance through hydraulics aiding
in start-up and control efficiencies.
The idea of efficiency of wind turbines being unique as aerodynamic machines
was initially defined by Albert Betz. Betz argued that wind turbines would be
limited by our ability to extract energy from the wind. This limit states that no
wind turbine can ever exceed 59.26% efficiency (Gipe 2009). The Betz limit is still
Wind Energy—Increasing Resilience in Water Infrastructure 171

used today to evaluate wind turbines for applications and for marketing, although
computational fluid dynamics and actual aerodynamic testing are more critical
tools. We still must accept that there is no perfect form of energy conversion where
combustion engines are at best 12% efficient and nuclear reactors are 4% efficient.
Currently, the best efficiency in commercial solar panels hovers around 20%.
Other important losses for us to recognize include energy losses, particularly
electricity losses owing to the type of electricity, direct current or alternating
current, and, of course, transmission losses of all generated electricity. Most
people are unaware that 80% of utility or grid transported electricity generated is
actually lost in transmission.

ASSESSING POWER FROM WIND

Wind turbines harvest energy from the wind. In general, the available power is
determined by a couple of ways. Beginning with the VAWT technologies, the
simplest power rating is associated with VAWT technologies whereby the actual
area of the rotors that is available to interact with the wind determines the power.
This applies to a majority of wind turbines with three or more rotors, because
the width of all the rotors during rotation or sweep is roughly the basis for rating
power. Efficiency and the equilibrium point of the Savonius type may be further
determined through wind tunnel testing.
New devices used to create an increased lift in the new Hy-VAWT technologies
do not change the power rating. They can only improve efficiency.
It is important to know the rated power from the manufacturer. This power
is, in general, based on a certain wind speed. Most manufacturers will use clever
means of marketing the most auspicious wind speed for their rated power.
Only two VAWT manufacturers rate their products at equilibrium. This means
basically, equal energy in from wind and equal generated power out. Both Star
Sailor Wind and Oy Windside rate power at equilibrium, which allows anyone
with a fundamental knowledge to ascertain the power rating for their installation.
So, if we have a 1 m-wide generic Savonius, which we also assume is 1 m high,
the total power is 50 W. Based on the square law, a 2 m wide, 1 m high would
have a power rating of 200 W, and a 6 m wide, 1 m high would be 1,600 W and so
forth. A fascinating concept regarding these high-wind-tolerant turbines is their
ability to be stacked in a 12 unit, 8 m turbine stack, offering 38,400 W or 38.4 kW
of power in a single installation. This is where the new Hy-VAWT technologies
will decrease overall expenditures (Menges 2016).
HAWT or propeller wind turbines power rating is based on the wind speed,
air density at a specific elevation, and the area of sweep of the rotor blades. Sea-
level density of air is approximately 1.225 kg/m3. It is the mass of air that acts to
turn the rotor blades. The power, in watts, available from the wind is found by
Power = 12 (ρ)×(swept area in square meters)×(wind velocity in m/s)3 = Watts
172 Renewable Energy Technologies and Water Infrastructure

Note, where the air density is ρ at the sea level is 1,225 kg/m3, this will change with
elevation and humidity. Determining the power of a wind turbine for a particular
location requires calculations based on average air density and humidity for
different seasons of the year.
Because the wind speed is cubed, if it were to double, the power will increase
by a factor of 8.
If we have a 10 m diameter HAWT, we determine the sweep area by determining
the radius, which is 5 m, and using the formula for a circle πR 2 we have

Sweep area = π52 = 78.54m2

For a wind speed of 4.47 m/sec or 10 mph, we can determine the power available
through

P = 12 (1.225) × 78.54 × (4.47)3 = 4, 300 W

If we increase the wind speed to 8.94 m/sec or 20 mph,

P = 12 (1.225) × 78.54 × (8.94)3 = 34, 400 W

So, by doubling the wind speed, we see that we obtain about eight times the power.

TURBINE TECHNOLOGY, APPLICATIONS, AND SITING

Evaluating sites for wind technologies is fundamental to achieving successful


applications in water management, treatment, storage, and transport.
Siting of wind turbines requires data on elevation, prevailing winds, and
surrounding structures and services, such as telephone and electrical wires,
roadways, and bridges, and environmental data that must also consider all wildlife
and human activity.
This said, wind offers options in energy resilience that little other renewable
energy can provide, if we understand how to use wind effectively. Siting is the
most important factor in the choice of wind technology, including the size and
type of the wind turbine. Site data will also determine what energy is expected
from the installation. Wind speed, direction, and distribution varies by season
and the greatest availability of wind is in the spring and fall where wind speeds
average greater than 2 m/sec. This is, in general, the cut in speed, when the wind
turbine will generate electricity.
Certain wind turbines, such as large HAWT systems, will actually require
electricity to start rotation. However, these large, utility-sized wind turbines are
unlikely candidates for our desalination, water treatment, and water infrastructure
security schemes.
Wind Energy—Increasing Resilience in Water Infrastructure 173

BUILT ENVIRONMENT WIND TURBINE

Wind turbines designed for built environments are becoming more popular and
can benefit water infrastructure power as their characteristics favor applications
in plant and infrastructure environments, as shown in Table 9-1.

Table 9-1. Comparison of Features and Characteristics of HAWT and VAWT


Technologies.

Horizontal axis wind turbine (HAWT) Vertical axis wind turbine (VAWT)
Wind resource
Wind speed frequency distribution Wind speed frequency distribution
Prevalent wind direction Prevalent wind direction
Turbulence (TI) Turbulence (TI)
Inflow angles Inflow angles (Darrieus type)
Building characteristics and geometry
Characteristics and geometry Characteristics and geometry
Roof shape (flat roof, pitched roof, Roof shape (flat roof, pitched roof,
parapets) parapets)
Orientation with respect to prevalent Orientation with respect to prevalent
winds winds
Significant structural considerations Few structural considerations as
owing to tower mounting is across parapets or on
flat roof structures
Tower height No tower, roof, and parapet mounting
Vibration emissions owing to flow- Low vibration emissions can be
induced vibration and blade flutter remedied by vibration-reducing
risk mounts
Vibration emissions owing to
mechanical turbine control
systems
Built environment turbine safety comparisons
Wind speed Balanced rotors with structural
advantage that experience
maximum torque only once during
revolution ((all VAWTs)
Turbulence (TI) risk High wind tolerance (Savonius type)
Extreme wind direction changes Turbulence (TI) reduced risk (all VAWTs)
Noise emissions may affect building Wind direction changes (Darrieus
occupants and have environmental type)
impact
Bird danger Low noise emissions
Turbine failure projectile zone (TFPZ) Low risk to birds/wildlife (Savonius
primarily a HAWT concern type)
particularly associated with
increased TI and icing conditions
174 Renewable Energy Technologies and Water Infrastructure

Siting issues in urban and suburban areas is complicated when employing


HAWT technology because of the sweep of the blades, installation requirements
including structural issues caused by the induced force on the building structure
by the mounted mast, turbulence (which increases maintenance), and delineating
a safe turbine failure projectile zone (TFPZ). HAWT technologies do not harvest
turbulence in the same way as the new Hy-VAWT technology, which can actually
benefit from controlling turbulence.
Wind turbulence, particularly in built urban areas, is a result of interactions
with a variety of building heights, prevailing wind, and geometries as well as
what has been called “wind compression” by architects creating increased air flow
around certain building structures. This increase wind speed in built areas can
be both beneficial and deleterious depending on the (particular) wind technology
integrated. Hy-VAWT Savonius technology performance can actually benefit
from wall and parapet installation harvesting the compression of inflow from
wall boundaries.

QUANTIFYING TURBULENCE AND TURBULENCE INTENSITY

Wind turbulence impacts the fatigue loads acting on wind turbines. HAWT
blades can be damaged by rapid and extreme changes in wind speeds, particularly
unstable fluctuations and turbulent winds. Therefore, accurate modeling of wind
velocities is critical in siting HAWTs in built environments.
Turbulence intensity (TI) is the standard deviation of the horizontal wind
velocity divided by the average wind velocity over time (t). Rapid fluctuations
increase TI, and steady winds decrease TI. TI typical range is 3% to 20% (Kumer
et al. 2016). Instrumentation used and the number and type of sensors significantly
influence the observed TI. Further, because TI measurements are measuring
horizontal wind components, vertical wind is not recorded. Turbulent kinetic
energy (TKE) measurements that evaluate horizontal and vertical components,
which are more commonly used in meteorological sectors, have been proposed
as a replacement for TI models. TKE measurements of ocean-based wind turbine
sites allow for a greater understanding of boundary layers and complex winds
common in ocean environments.
TI is the variable most likely to affect wind turbine performance at any given
site. It also has the greatest effect on the performance of HAWT technologies
and must be understood for any installation because of its effect on power curve
measurement (Kumer et al. 2016).
The International Electrotechnical Committee (IEC) definition of turbulence
for wind energy applications is TI, where TI = σU/U (wind-speed std.-deviation
to mean wind-speed).
The normal turbulence model (NTM) defines the turbulence observed and
analyzed under normal operating conditions. The parameters of NTM are based
on collected data from proposed or existing sites based on the International
Wind Energy—Increasing Resilience in Water Infrastructure 175

Standard IEC 61400-2, which limits the level of TI by wind speed as part of the
design process. HAWTs installed in locations that have observed TI higher than
the recommended NTM will reduce reliability and increase maintenance and
liability in built areas.

ENERGY AND WATER

Water treatment and desalination are two fundamental areas where renewable
energy resources like wind energy can be applied effectively. Diversity in energy
resources not only increases water security but can also provide support locally,
where reliable safe water resources are often unavailable.
To increase freshwater resources in a water-stressed region, one must
recognize the requirements to create fresh water from salty, brackish, or polluted
water supplies. Desalination and filtration processes require not just water
supplies but also reliable resilient energy.
The most common and important method of water treatment, offering safe
water, is desalination. In the 1970s, desalination was a process that removed salts
and minerals from ocean water. Today, desalination is a tool for the treatment of
brackish and wastewater and has been adopted almost globally as a method of
augmenting water supplies.
In regions such as the Middle East, desalination is responsible for 86% of
wastewater reuse and seawater for as much as 65% of drinking water. In other
regions including North America, only 8% of water is reclaimed using this process.
Desalination technologies rely on filtration, enhanced solar evaporation,
multieffect distillation, and membrane-based processes with the better-recognized
reverse osmosis (RO) and the lesser-known electrodialysis. RO systems make up
more than 60% of the desalination plants today.
RO is a hydraulic filtration process whereby the untreated water is pressurized
and forced through selective membranes that remove contaminants primarily
through size exclusion and charge repulsion. RO offers many advantages because
of the availability of small-packaged, pallet-mounted plants to municipal-
sized plants. Energy consumption is relatively low compared with other water
treatment methods. At present, the relative advantages of RO in capital cost,
energy consumption, and ease of operation make this option a realistic solution.
However, enhanced solar evaporation processes are becoming more affordable,
and even though they require larger land areas, may offer energy advantages in the
near future.
Many challenges still exist in RO technology that are also issues for wind
power. The siting of most large desalination plants will be close to seawater. These
concerns are primarily environmental, since winds are easily available and offer
nearly year around consistent wind speeds.
Desalination plants in coastal areas create unique issues environmentally
because they tend to interfere with and entrap marine microorganisms during
176 Renewable Energy Technologies and Water Infrastructure

the intake phase. Screens are useful in reducing microorganism capture but do
not protect the smallest organisms. Also, the concentration of brine in discharge
streams further deteriorates the marine environment. This creates a secondary
mixing requirement whereby brine concentrated effluents are mixed with either
captured storm water or untreated seawater to reduce salinity to a reasonable level.
Said challenges cost plant energy and must become part of the analysis, and
we must plan for fluctuations in power requirements during seasonal changes and
during high freshwater demand periods (Szeptycki et al. 2016).
Several issues exist in RO technology for inland water purification that
include poor resistance to chlorine by many types of polyamide membranes.
Unfortunately, sodium hypochlorite is common in many water treatment plants
(WTPs) because it limits the growth of biofilms. New membranes offer some
chlorine resistance, but it must be considered in terms of integration of RO
technologies for water purification inland.
Other filtration systems including activated carbon, diatomaceous earths,
and lime create fewer issues in technology selection and have lower power
requirements. If we consider the energy required for desalination of water based
on their total dissolved solids (TDS), as shown in Table 9-2, we get a reasonable
estimate for a variety of RO processes.
From Table 9-2, we know that the higher the level of TDS, the greater the
energy required to desalinate the water. This is where the choice from pretreatment
for RO plants is critical, not to prepare the water for the RO process, but for
determining the efficiency and cost of pretreatment in the energy analysis. As we
consider the integration of a RO plant for water desalination, we must factor in
the total-J (kWh) necessary to operate based on average water demand.
Pretreatment using activated carbon filters provides an efficient means to
reduce organic substances and the size of solid particulates. It also utilizes a high-
efficiency pump that saves energy as well. To complete an energy generation plan,
we must know about all projected energies utilized in a facility.
Assuming that an RO plant utilizes a state-of-the art energy recovery turbine
(ERT), whereby energy is recovered from the concentrate flow, the energy utilized

Table 9-2. Types of Water for Desalination and Their Total Dissolved Solids (TDS)
with Specific Energy Consumption for Unit Volume.

Minimum energy,
Water source TDS (mg/L) MM-J/m3 (kwh/m3)
Seawater 15,000–50,000 2.41 (0.67)
Brackish water 1,500–15,000 0.61 (0.17)
River water 500–3,000 0.14 (0.04)
Pure Water <500 <0.04 (<0.01)
Wastewater (untreated domestic) 250–1,000 0.04 (0.01)
Wastewater (treated domestic) 500–700 0.04 (0.01)
Source: NAE (2016, pp. 56).
Wind Energy—Increasing Resilience in Water Infrastructure 177

will be 15.84 MM J/m3 (4.4 kWh/m3). We must have a level of energy utilization
that can be met with a well-sited wind turbine:
Pretreatment water handling and filtration: 4.03 MM J/m3 (1.12 kWh/m3)
Reverse Osmosis process unit: 15.84 MM J/m3 (4.40 kWh/m3)
Post-treatment storage (tankage): 2.81 MM J/m3 (0.78 kWh/m3)
that is, a total of 22.68 MM J/m3 (6.30 kWh/m3).
Considering the following:
at 1 m3 = 264.172 gal., and the average 6 in. pipe carrying 2.649 m3/min
(700 gal./ min), the energy required would be 41.98 MM J (11.66 kWh)
for 2.52 GJ (699.6 kWh) or 60.44 GJ (16,790.4 kWh), and roughly 700,000
W or 700 kW from out power supplies.
Wind turbines would include the following options:
Installation of ten Polaris America 100 kW—P25-100 turbines for a total
power capacity of 1 MW at approximately USD 350,000 per turbine
covering 12 acres and producing approximately, 71.28 GJ (19,800 kWh).
The cut-out speed is 25 m/sec (55.9 mph). In addition, at the initial cost
of USD 3.5 million for installation.
Or
Installation of 24 Star Sailor Wind 38.4 kW towers for a total power
capacity of 921.6 kWh at approximately USD 129,000 per tower
covering 8 to 10 acres or equivalent to two (16 × 16) arrays producing
approximately 64.8 GJ (18,000 kWh). Star Sailor Wind towers have no
cut-out speed and will continue to produce power to 176 kmph (110
mph), when induced drag stops rpm. This option will be at a cost of
USD 3.1 million for initial installation.
Both options offer power sufficient to operate the plant at 2.649 m3/h (700
gph), producing as much as 2.536 m3/h (670 gph) of freshwater. The Polaris
HAWT offers greater power in the low end of the curve to 10 m/sec (22.4 mph).
The Star Sailor Wind tower arrays create further efficiencies by reducing land-use
requirements and producing improved performance based on array installation
geometries (Whittlesey et al. 2010).
Both wind turbine installations require battery banks and power management
systems that will use up to 8% of power generated. This conceptual system offers
a potential off-grid operation that will generate the greatest resiliency. However,
grid tie is recommended for all critical infrastructure facilities.
This exercise, of course, is offered as a beginning for a conceptual installation.
In an actual plan, an analysis of all energy consumption, distribution, and storage
associated with all predicted loads would be obtained as well as a 2 to 3 month site
survey of wind, potential obstacles associated with all elevation points.
A micro grid capable of running a 2.271 to 2.649 m3/h (600 to 700 gph) water
treatment plant requires diversity in integration and storage. The US Department
178 Renewable Energy Technologies and Water Infrastructure

of Energy Federal Energy Management Program has evaluated these methods in


the optimization of reverse osmosis operations (PNNL 2013).
Distillation offers a means of desalination and overall water purification and
includes multistage flash distillation, multiple-effect distillation, and mechanical
vapor compression processes. These thermally driven processes in which water
is evaporated through heating and recovered through condensation may be
effectively integrated with wind energy (NRC 2008).
The energy requirements to produce heat sufficient to support steam
distillation may be less than those of even the most efficient RO technology.
Further, solar thermal augmented plants, where the saltwater is preheated
through thermal solar panels and solar concentrators, increase energy efficiency,
as it would for any distillation process.
Any application of wind energy must be viable technologically and
economically. There are few out-of-the-box solutions at this point in time. Making
any use of wind energy in water treatment is a means of increasing resilience
in an energy augmentation program. As new and more energy-efficient RO and
filtration systems emerge, wind energy can become a primary energy source for
many types of water treatment and other unit operations.

ROLE OF WIND ENERGY IN WATER INFRASTRUCTURE SECURITY

The potential for wind energy application(s) to the resilience of water infrastructure
including the role in water security may not be obvious. Water security and safety
have been a primary issue in the national security community for over two decades.
Concerns over terrorism, industrial espionage, natural disasters, and energy are
key components. However, in recent years, concerns over local and regional water
management have become a growing concern in the national security sectors.
Growing issues around mismanagement and ineffective development have
led to water safety issues in areas of Michigan, New Jersey, and Ohio. Beyond
public safety, there is an underlying issue in the economics of freshwater. As
municipalities and states struggle with water safety, larger overriding concerns
for water insecurity that can drive civil disobedience and conflict exist.
Where many of these water-related instances have been self-inflicted, it does
not abrogate the responsibility of the government to guarantee safe and fresh water.
Other factors are that many technology-based industries as well as the food
and beverage industries depend on freshwater. Most of these organizations have
water treatment plants as part of their facilities and growing numbers have
wastewater treatment because of growing sewer issues in many older urban areas.
Water treatment plants, reservoirs, and industrial water resources all require
the ability to monitor not just water quality but also physical security. The advent
of virtual security systems has increased the use of networked cameras, sensors,
controllers, and other devices that are depended on for water treatment, storage,
and delivery. Wireless communications, edge computing, and ledger data
systems have further advanced the sophistication by which all infrastructure
can be monitored and controlled.
Wind Energy—Increasing Resilience in Water Infrastructure 179

The expansion of wireless as a ubiquitous technology has led to the growth of


the field of the Internet of Things (IoTs). IoT technologies offer both advantages
and disadvantages. The advantages are usually obvious, dedicated, networked
sensors and devices to monitor and control water. However, the disadvantages
are increasing the insecurity of IoT-networked systems.
An area rarely explored is the structure and security of cellular wireless
communications, certainly not beyond integrating devices. Unfortunately,
technologies like cellular wireless have been exploited successfully for many years.
This is fundamentally due to network insecurity associated with the Common
Channel Signaling System No. 7.
Common Channel Signaling System No. 7 (SS7) is a data communications
network standard that is intended to be used as a control and management network
for telecommunication networks. SS7 supports most functions and operations
including call management, data base query, and routing for an integrated services
digital network. This is the basis of the cellular telephone network (Gallager 2008).
The world was introduced to SS7 in 2014 with much fanfare as it became a
major news event. The underlying vulnerability inherent in SS7 allows almost
anyone with basic technical skills to listen to calls, track cellular phones, and read
and manipulate RMS feeds such as texts. This is also a fundamental vulnerability
in networking devices. Even though most devices use some form of encryption,
most data can be intercepted and control commands for valves, circuit controls,
and other critical components can be accessed (Goodin 2018).
The universe of IoT developed the new area of virtual security (VSEC). VSEC
is not just about the primary wire and wireless infrastructure to support things
like virtual fences, but the actual security of virtual environments encompassing
operations such as uninterruptable power supplies (UPS), remote sensing,
infrastructure and industrial monitoring, edge computing, and distributed ledgers.
The VSEC community’s primary challenge has become how to gain access to
energy for secure networking and communications. This has led to an explosion
in the development of wind- and solar-powered networks. One such technology
is based on purpose-designed and built Hy-Lift VAWT hybrid systems supporting
advanced remote sensing and edge computing that integrates distributed data
storage within an adaptive artificial intelligence (AI) network. The Self-Powered
Intelligent Networks (SPIN™)* offer not just the UPS systems required for network
security but also independent data analysis and storage through advanced edge
computing technology, thus removing the vulnerability in current cellular
wireless networks (Menges 2015).

EMERGING TECHNOLOGIES

Water treatment technologies have advanced over the last few decades by
increasing water quality and decreasing the cost of plant construction. Other
technologies have evolved, including more unusual concepts whereby ocean
environments are used to support energy production and water treatment. New
180 Renewable Energy Technologies and Water Infrastructure

technologies combined with new methods will optimize the integration of new
water–energy systems.
One such method exploits ocean water for energy conversion turning solar
energy into electricity based on the natural thermal gradients found in ocean
environments. Ocean thermal energy conversion (OTEC) exploits the temperature
difference between warmer surface waters and lower water temperatures at depth.
OTEC has several limitations, but the most specific is the requirement for the
region of the ocean where the seawater offers a thermal gradient in the upper
most 1000 m and the water has an average temperature more than 20°C. These
limitations aside, this process for generating electricity could power a substantial
desalination plant in the equatorial regions where these conditions are the most
prevalent (NRC 2008).
The integration of hybrid wind–solar technologies into the OTEC concept will
create further resiliency and energy efficiency by creating primary and secondary
energy supplies establishing expanded capabilities including uninterruptable
power supplies (UPS) and increased output of fresh water.
Other emerging technologies are based in new methods in remote sensing.
The implementation of new methods in remote sensing of water and water
resources will increase the stability of water availability and security and build
databases for planning and designing water infrastructure.
Simple and cost-effective methods utilizing remote sensing of groundwater
through surveillance wells integrating wind-powered technologies, such as self-
powered networks and micro-grid systems, provide real-time data on drought
development, flooding, and the potential impact of hurricanes and earthquakes.
Developing databases on groundwater availability and access will lead to an
increased understanding of how water resource planning will impact communities
and regions, thus increasing the chances of success of new engineering projects.
These methodologies can easily utilize renewable energy sources like wind and
wind-solar hybrid systems supporting sustainable solutions.
Expanding the use of eco services in engineering water infrastructure will
also lead to greater energy efficiency and infrastructure resilience and reduce
energy requirements. These eco services may include the integration of wetlands
into resiliency plans where renewable energy is implemented. Future innovation
in multifunctional integration methods and emerging technologies will expand
the use of the natural environment in increasing water resilience.
Further, the use of remote sensing in the monitoring of natural water resources
supports the use of eco services in water resiliency and resource engineering.
The expansion of eco services in engineering water infrastructure will also lead
to greater energy efficiency and infrastructure resilience and reduced energy
requirements. Combining with the renewable, wind energy resources, would
create a highly resilient component(s) on water infrastructure.
Future innovation in multifunctional integration methods and emerging
technologies will expand the use of the natural environment in increasing water
resilience.
Wind Energy—Increasing Resilience in Water Infrastructure 181

SUMMARY

Water infrastructure well supported by renewable energy systems, including


wind energy, is the big pathway into the future. Advanced wind energy systems
are available and are resilient. The capability of implementing resilience in the
built infrastructure to tolerate energy loss, as well as the outcome of natural and
human-made disasters, including the highly volatile aspects of climate change, is
key to establishing criteria for resilience in critical water infrastructure.
Current hybrid technologies, such as Star Sailor SPIN™* (trademark of Star
Sailor Wind) systems, can augment traditional utility grids by creating local and
dedicated micro grids providing not just increased security and safety for critical
infrastructure but also water information and remote sensing disseminating real-
time data on water security and availability. This will also increase the efficiency
and competency of local water management and treatment, reducing costs, and
increasing resiliency.

References
Butzer, K. W. 1975. Early hydraulic civilization in Egypt, 21–23. Chicago: Univ. of Chicago
Press.
Gallager, R. G. 2008. Principles of digital communications. MIT Open Course Ware.
Cambridge, MA: Massachusetts Institute of Technology.
Gipe, P. 2009. Wind energy basics. 2nd ed. Hartford, VT: Chelsea Green Publishing
Company.
Goodin, D. 2018. “SS7 routing-protocol breach of US cellular carrier exposed customer
data.” Ars Technica.
Kumer, V.-M., J. Reuder, M. Dorninger, R. Zauner, and V. Grubišić. 2016. “Turbulent
kinetic energy estimates from profiling wind LiDAR measurements and their potential
for wind energy applications.” J. Renewable Energy 99: 898–910.
Lawler, E. 2016. “Flint infrastructure fix could cost up to $1.5B, mayor Karen Weaver
says.” Accessed September 19, 2021. https://www.mlive.com/lansing-news/2016/01/
flint_infrastructure_fix_could.html.
Menges, P. A. 2015. Self-Powered Intelligent Networks (SPIN) for secure, robust wireless
communications and computing. Cincinnati: Star Sailor Wind, LLC.
Menges, P. 2016. A Hy-Lift VAWT data and specifications. Cincinnati: Star Sailor Wind,
LLC.
NAE (National Academy of Engineering). 2016. Frontiers of engineering: Reports on
leading-edge engineering from the 2016 symposium. Washington, DC: National
Academies Press.
NRC (National Research Council). 2008. Desalination: A national perspective. Washington,
DC: NOAA.
PNNL (Pacific Northwest National Laboratory). 2013. Reverse osmosis optimization.
Washington, DC: US Dept. of Energy, Federal Energy Management Program.
Saidur, R., M. Islam, and N. Rahim. 2011. Assessment of wind energy potentiality at Kudat
and Labuan, Malaysia using Weibull distribution function, 985–992. Amsterdam,
Netherlands: Elsevier.
182 Renewable Energy Technologies and Water Infrastructure

Szeptycki, L., E. Hartge, N. Ajami, A. Erickson, W. N. Heady, L. LaFeir, et al. 2016. “Marine
and coastal impacts of ocean desalination in California. A report of water in the west,
center for ocean solutions, Monterey Bay aquarium, and the nature conservancy.” http://
waterinthewest.stanford.edu/sites/default/ les/Desal_Whitepaper_FINAL.pdf.
Whittlesey, R. W., S. Liska, and J. O. Dabiri. 2010. “Fish schooling as a basis for vertical
axis wind turbine farm design.” Bioinspiration Biomimetics 5 (3): 035005.
CHAPTER 10
Solar Energy and Water/
Wastewater Infrastructure
Venkata Gullapalli

INTRODUCTION

The World Health Organization (WHO) estimated that a total of 2.1 billion people
around the world do not have access to safe drinking water, and approximately
1.1 billion people around the world do not have access to electricity. In the
United States, treatment of water and wastewater consumes approximately 3% of
the nation’s energy consumption. Conventional energy generation involves the
release of harmful greenhouse gases (GHGs). Using renewables not only reduces
these GHGs but also facilitates decentralization of energy generation by reducing
the grid load. Solar is one of the renewals that is gaining importance in the power
generation industry. The use of solar radiation for electricity generation has gained
traction since the 1980s. Solar plants can be either centralized or decentralized
and have shown great results when installed in regions where sunlight is abundant.
Incorporating renewables into water and wastewater treatment can help reduce
energy costs and GHG emission. This chapter will provide the following details:
electricity generation using solar radiation, solar disinfection (SODIS), and
incorporation of solar use into water and wastewater treatment.

SOLAR RADIATION

The electromagnetic radiation emitted by the sun is called solar radiation. The
availability of solar radiation depends on the season, time of the day, local weather,
and geographic location. Measurements of solar energy are typically expressed
as the total radiation on a horizontal surface or as total radiation on a surface
tracking the sun. Using a variety of technologies, solar radiation can be turned
into electricity and heat. However, the amount of energy that can be harvested
majorly depends on the solar radiation that is available. The two most commonly
used solar technologies are solar photovoltaics (PV) for electricity to power homes

183
184 Renewable Energy Technologies and Water Infrastructure

and businesses and concentrating solar power (CSP) to provide electricity for
large power stations (NREL 2020a). Solar PVs use light and CSPs use heat from
sunlight for generating electricity.

SOLAR PHOTOVOLTAICS

PV cells convert light (photons, P) directly into electricity (voltage, V). PV cells are
made of different semiconductor devices. When light (photons) strikes a PV cell,
photons of light can transfer their energy to electrons, allowing the electrons to
flow through the material as electrical current (NREL 2020a, b; EERE 2020a). This
is then harnessed by wires connected to the positive and negative sides of the cell.
A solar PV panel is a combination of many PV cells connected in chains. These
connected PV cells form larger units and are called modules or panels. Modules
can be used individually or connected to form arrays for larger production.
Because of the flexibility in making small or big arrays, PVs are suitable for
meeting small and large electricity demands.
Solar cells use a wide variety of semiconductors. The first-generation solar
cells are made of silicon. The second-generation ones include cadmium telluride
(CdTe) and copper indium gallium diselenide(CIGD); these are thin-film cells
and are made through a complex process of combining four elements. The third-
generation PV cells are organic–inorganic hybrid assemblies, nanostructured
semiconductors, and molecular assemblies (Ebhota and Jen 2018). Cells from
each generation have their own set of pros and cons, although there is a huge
advancement in solar cell production since the 1950s. There are no panels that
are available now that can convert all the sunlight they receive into energy. There
are several factors that affect the conversion efficiency of cells. Researchers are
constantly striving to innovate in the area of solar cell generation.
Solar PV modules are a part of the PV system. The PV system makes the
generated electricity in PV modules to be useful for domestic and commercial usage.
A PV system consists of mounting structures, inverters, and storage. Figure 10-1
represents a typical PV system: (I) energy source: the sun emits sunlight consisting
of photons; (II) energy conversion: a solar panel converts sunlight to electricity; (III)
energy conversion and conditioning: converting direct current (DC) into alternating
current (AC) power; (IV) energy storage: battery for storing electricity generated
during sunny times to use it nights or nonsunny times; and (V) Energy distribution:
distributing the generated/stored energy to (a) domestic or (b) commercial usage.
Another new development in PV systems is concentrated PV (CPV) systems. The PV
cells in a CPV system are built into concentrating collectors that use a lens or mirrors
to focus sunlight onto the cells (NREL 2020c). The primary advantages of CPV
systems are as follows: high efficiency, low system cost, and low-capital investment
to facilitate rapid scale-up; the systems use less expensive semiconducting PV
material to achieve a specified electrical output. Reliability, however, is an important
technical challenge for this emerging technological approach, and the systems, in
general, require highly sophisticated tracking devices.
Solar Energy and Water/Wastewater Infrastructure 185

Figure 10-1. Typical solar photovoltaic system.


Source: With permission from ucf.edu.

SOLAR THERMAL

Unlike solar PV, solar thermal uses thermal (heat) energy in sunlight. To date,
CSP is a technology that makes the best use of solar thermal to generate electricity.
CSP is the best fit for commercial generation of electricity. The basic mechanism
involved in CSP generation is, thermal energy from sunlight is used to generate
steam, and steam then drives a turbine connected to an electrical generator for
generating electricity (NREL 2020c; EERE 2020b). Salts are used to better thermal
energy storage. Salts are heated up to 500°C and the molted salts warm up water to
generate steam. Molten salts are stored in tanks and used during nonsunny times.
Usage of salts makes the system generate electricity round the clock. Figures 10-2
and 10-3 (DOE 2014) represent typical CSP systems. Figure 10-2 shows the system
that uses large mirrors to reflect and concentrate the thermal energy to a receiving
tower. The concentrated thermal energy boils water, and steam is sent to a storage
tank. Steam is then used to drive the turbine connecter to a generator for power
generation. The steam is then sent back to a condenser, and the water is sent
back to the receiver tower. Some power towers use molten salt in place of water
and steam. This hot molten salt can be used immediately to generate steam and
electricity, or it can be stored and used at later required frequencies.
Figure 10-3 shows the CSP system that uses linear concentrator systems.
Linear concentrators are long rectangular and curved mirrors. Concentrators
are tilted in a way to collect sunlight and focus it on receivers (tubes) that run
the length of the mirrors. Receivers are positioned along the focal line. Reflected
sunlight heats up the salts in a tube. The hot fluid then is used to boil water in a
conventional steam-turbine generator to produce electricity. Another CSP system
that is in use is the dish/engine system. This system uses mirrored dishes, which
are composed of small flat mirrors.
186 Renewable Energy Technologies and Water Infrastructure

Figure 10-2. Typical solar photovoltaic system.


Source: DOE (2014).

Figure 10-3. Typical concentrating solar power system.


Source: DOE (2014).

SOLAR PHOTOVOLTAICS AND CONCENTRATING SOLAR POWER


COMPARISON

Researchers are working to make both CSP and PV more viable and efficient. Each
system has its own pros and cons and is better in its own way. Both technologies
attract power generators for developing clean energy compared with conventional
systems that use coal, natural gas, or nuclear energy. CSP generates AC, and PV
Solar Energy and Water/Wastewater Infrastructure 187

Table 10-1. A Comparison between Solar PV and CSP.

Parameters Solar photovoltaic Concentrating solar power


Electric current Direct current Alternating current
production
Source of Light into electricity Solar thermal to electricity
energy
General PV modules, inverters in Solar field, thermal conversion
components some applications, (utilizing salts or others),
and associated power generation, and
electrical units associated electrical units
Application Community to Community to industrial scale
industrial scale
Technology Rapidly evolving, Ongoing advanced research
status efficient, and cost-
effective systems

generates DC, which is converted to AC for distribution. Because, unlike PVs,


CSPs use the thermal component of sunlight, this limits the power generation
by CSPs during low light conditions, whereas the light is diffused and can be
used by PVs to generate energy. Thermal energy storage technologies that are
available now are far advanced than electricity storage technologies. CSPs use
thermal energy storage technologies, which help generate electricity at any time
(night times, or on cloudy days). PVs use electricity storage units (batteries), which
are not so efficient when compared with thermal storage units. CSPs are better
for large-scale electricity generation with no irregularities in supply mostly. PVs
are good for small or decentralized electricity generation where much storage
is not required and less space is available. Constructing PV systems is easy and
requires less time and cost when compared with CSP plants. CSPs require large
space when compared with PVs. The other advantage of PVs is their portability;
they can be used in calculators, watches, cars, and aircrafts. As of now, CSPs are
better for industrial-scale adoption. In recent years, the price of PV panels has
come down substantially and is expected to drop even further. Research on CSP
technology is lagging behind PV technologies (Rashad et al. 2015). Table 10-1
shows a comparison between solar PV and CSP.

SOLAR IN WATER INDUSTRY

Solar radiation in water purification dates to prehistoric times. Many religious


practices involve leaving water under the sun before consuming, in the name of
offering it to the Sun God. Scientifically, this is called solar disinfection (SODIS),
where water is disinfected by solar radiation. Present scientific trends have taken
188 Renewable Energy Technologies and Water Infrastructure

it to the next level of using solar PV and CSP for operating water and wastewater
treatment systems or plants.
According to the EPA (2020), “A public water system provides water for
human consumption through pipes or other constructed conveyances to at least
15 service connections or serves an average of at least 25 people for at least 60
days a year”. A traditional public water system consists of three basic components
(NRC 1982): raw water source, treatment of water, and distribution of water to
users. The other component that adds to basic operations is storage, where treated
water is stored in tanks for supply during demand time.
There are approximately 155,693 public water systems in the United States
(CDC 2020). A total of 8% of US community water systems provide water to 82%
of the US population through large municipal water systems (CDC 2020). US
public drinking water systems use roughly 141.1 PJ (39.2 billion kWh) per year
(Bracken et al. 2013), where much of the electricity is used for pumping water. As
much as 40% of operating costs for drinking water systems can be earmarked for
energy (EPA 2020). Figure 10-4 (Bracken et al. 2013) shows the typical energy use
within a public surface water (SW) system.
Replacing or introducing solar pumps for pumping operations could result
in reducing energy costs for PWS. Solar pumps are powered by solar PV panels
and incur higher upfront investment costs but are exceptionally reliable and
incur less operating and maintenance (O&M) costs. Because of decentralized
installation and operation capabilities, solar pumps are more suitable for rural
and underdeveloped communities since grid electricity is unavailable. Raw
water source type and quality determines the required water treatment train
for producing potable water. Groundwater (GW), SW, and GW under direct
influence of surface water (GWUDISW) are three major sources of raw water
worldwide. The fourth and fifth sources that are gaining reputation in recent times
are saline (saltwater) and brackish water. As stated by EPA (2020), water with

Figure 10-4. Energy usage in a public surface water system.


Source: Bracken et al. (2013).
Solar Energy and Water/Wastewater Infrastructure 189

TDS concentration less than 3,000 mg/L is considered as fresh water, between
3,000 and 10,000 mg/L is considered as brackish, and in excess of 10,000 mg/L is
considered saline, although water with TDS concentration less than 3,000 mg/L is
considered fresh but is only considered when a source with TDS below 1,000 mg/L
is not available (Godsey 2020). Fresh source water does not require special
treatment techniques or methods in most cases because of its prevailing low salt
concentration. Brackish and saline water require special treatment processes to
reduce TDS concentration and to make it potable.

DESALINATION

The process of removing minerals and salts from saline and brackish water to
produce potable or irrigation water is called desalination. Desalination of sea and
brackish water is gaining reputation to keep up to or supplementing the demand
for fresh water. Desalination involves five steps of water processing: pumping,
pretreatment, micro- or ultrafiltration, desalination by advanced filtration, and
optional post treatment. Based on the mechanism, desalination processes are
divided into two categories: thermal and membrane technologies (Krishna 2004).

Thermal Technologies
Thermal desalination technologies involve heating of saline water and collecting
condensed vapor to produce pure water (Krishna 2004). Thermal desalination
technologies incur higher costs and involve an energy-intense process. Thermal
technologies include multistage flash distillation (MSF), multieffect distillation
(MED), and vapor compression distillation (NAS 2008).
MSF plants have been built since the late 1950s. MSF plants are typically for
producing fresh water up to 56,775 m3/day (15 million gal./day) (Tonner 2004). Each
of these processes can be structured as a “long-tube” or “cross-tube” design. In the
long-tube design (built at Freeport, The Bahamas, in 1961), tubing is parallel to the
concentrate flow, whereas in the cross-tube design, tubing is perpendicular to the
concentrate flow. Globally, distillation processes produce about 12.9 billion L/day
(3.4 billion gal./day), which is about 50% of the worldwide desalination capacity;
MSF plants represent about 84% of this capacity. Most of these plants have been
built overseas, primarily in the Middle East. Several MED plants have been built
in the United States and overseas. Three low-temperature MED plants with a
combined capacity of approximately 13,247 m3/day (3.5 million gal./day) have been
operating successfully in St. Thomas, US Virgin Islands, where desalinated water is
the principal water supply source (Kelsey et al. 2005); the MED units are operated
by the Virgin Islands Water and Power Authority. Steam from the source power
plant is directed to the evaporators in the desalination units, and product water is
obtained as a condensate of the vapor from each vessel. In general, several MED
plants are found overseas, in the Caribbean and Middle East regions in particular.
190 Renewable Energy Technologies and Water Infrastructure

Membrane Technologies
Membrane usage for separation of salts from water was started in the 1960s and
gained reputation in the 1980s (Shatat and Riffat 2012). The process includes
passing water through a synthetic membrane to move either water or salt to
induce two zones of differing concentrations to produce freshwater. Typical
separation processes are micro- and ultrafiltration. Three membrane technologies
used majorly in the water industry are reverse osmosis (RO), electrodialysis (ED),
and membrane distillation (MD).
RO involves separating dissolved salts from water by passing water through a
membrane under high pressure. RO systems started gaining importance since the
1970s and are relatively new when compared with other desalination technologies.
At present, RO is a better option for desalinating saline water when compared with
other desalination technologies. Energy is majorly used for pressurizing the feed/
raw water. It is required that the raw water be effectively prefiltered before entering
into an RO system. This helps in reducing incidents of membrane clogging and
increasing the membrane lifetime, although scaling in membranes does occur
is low when compared with that in thermal desalination techniques. As far as
the product output is concerned, one study states, “for every 5 gallons of usable
water, 40–90 gallons of water are sent to the wastewater system” (1 gal. is 3.785 L)
(Shatat and Riffat 2012). As researchers work to develop innovative membranes to
increase the capacity of RO systems, it is anticipated that there will be a reduction
in the reject or wastage in RO systems in the future.
ED is an electrochemical separation process that employs electrically charged
ion exchange membranes with an electrical potential difference as a driving force.
It depends on the fact that most salts dissolved in water are ionic, being either
positively (cationic) or negatively (anionic) charged, and they migrate toward
electrodes with an opposite electric charge. Membranes can be constructed to
permit the selective passage of either cations or anions (Shatat and Riffat 2012).
ED is not economically viable for saline water but is best suited for brackish
water. Comparitively, ED requires less pretreatment than RO. Energy usage is
proportional to salt removal.
MD is a thermally driven separation process, in which only vapor molecules
transfer through a microporous hydrophobic membrane. The driving force in the
MD process is the vapor pressure difference induced by the temperature difference
across the hydrophobic membrane (Alkhudhiri et al. 2012). MD requires that the
raw/feedwater be organic free, and, although energy consumption is similar to
that in MSF and MED plants, MD can be operated at low temperature ranges.

SOLAR IN DESALINATION

Desalination of water consumes a significant level of energy and membrane


technologies use electricity for pressurizing water. However, thermal technologies
use both thermal and electricity for operation. Ullah et al. (2018) reports that
Solar Energy and Water/Wastewater Infrastructure 191

the global desalination capacity sums up to 0.09 trillion L (0.024 trillion gal.)
of water per day. Sixty-five percent of the capacity is desalinated by RO, 21%
through MSF, 7% by MED, and the remaining 7% through other technologies. All
desalination techniques are energy intensive and require high energy to operate.
Al-Karaghouli and Kazmerski (2013) reported (in the year 2016) that around
1.04 PJ/day (290 GWh/day) electricity was consumed for water desalination; this
is a 10% increase compared with the consumption in 2010. A majority share of this
power is generated by using fossil fuels, which contribute to the release of GHGs.
The report developed by IRENA and ETSAP on renewable energy usage in water
desalination (in 2012) includes the average energy consumed by desalination per
gallon: 19.02 kJ/L (0.02 kWh/gal.) for MSF and 9.51 kJ/L (0.01 kWh/gal.) for MED
or RO or ED. Introducing renewable energy sources, such as solar, can help reduce
the release of GHGs. Solar-powered desalination systems are divided into two
categories, namely direct and indirect.
• Direct systems are similar to thermal desalination processes. These are called
solar stills, where saline water is heated and water vapor is condensed and
collected for potable uses. Solar stills are good for small-scale potable water
generation. The space occupied by stills is proportional to the amount of water
desalinated. Solar-powered humidification–dehumidification is an advanced
and convoluted form of solar still, where water is heated using collectors,
and involves different thermal processes. Solar-powered humidification–
dehumidification is good for applications in small communities in rural areas
where sunlight is abundant. To date, CSPs are known to be the best direct
solar desalination systems (Shahzad et al. 2019).
• Indirect systems are a coupled unit(s), a solar collector, and a desalination
part(s). One of the best indirect systems now available is a PV system powering
an RO or ED system, where solar panels are used to generate electricity
that is used for powering pumps in an RO or ED system. Studies are being
conducted to determine the best coupled systems. Both RO systems and PV
systems are still in the evolving stage. Making PV and RO systems more
efficient and less expensive can result in better solar-powered desalination
systems. Approximately, 45% of the renewable energy-powered desalination
plants use solar PV, 25% solar thermal, 20% wind, and 10% some type of
hybrid system that is derived by coupling different renewable energy sources
(Quteishat and Abu-Arabi 2004).

SOLAR WATER DISINFECTION

SODIS is the process of disinfecting water by exposing it to sunlight. Descriptions


of SODIS of water have existed in communities in the Indian subcontinent for
nearly 2000 years. In the distant past, drinking water was to be placed outside in
open containers so as to be “blessed” by the sun (Baker 1981). Downes and Blunt
(1877) proved that sunlight is effective in reducing or killing bacteria. Disinfection
192 Renewable Energy Technologies and Water Infrastructure

by sunlight happens because of the UV radiation and infrared radiation present


in solar radiation (Mbonimpa et al. 2012). The shorter the wavelength, the
stronger is its disinfection capacity and, based on this, UVC is considered as the
strongest disinfectant. Stratospheric oxygen absorbs all the UVC and 90% of
the UVB radiation and 5% to 10% of UVA reaching the earth’s surface; thus,
UVA and UVB are available for disinfection. Inactivation of bacteria under solar
radiation occurs in three ways: thermal inactivation, optical inactivation, and
combined thermal and optical inactivation.
Thermal inactivation is a process of inactivating bacteria by the application of
heat. This is one of the oldest techniques for water disinfection. This significantly
improves the microbiological quality of water but does not fully remove the
potential risk of waterborne pathogens if water temperatures do not reach boiling
point (Rosa et al. 2010). Infrared radiation is another aspect of solar radiation,
where heat is produced by radiation at wavelengths between 700 nm and 1
mm. The infrared radiation absorbed by water is responsible for increasing its
temperature. Microorganisms are sensitive to heat, and water can experience a
bacterial reduction of 99.9% prior to reaching boiling temperature. This can be
achieved by heating water up to 50°C to 60°C for 1 h (Meierhoer et al. 2002). In
SODIS, water is retained in airtight containers to increase the water temperature.
Several enhancement methods have been attempted by accelerating the rate of
thermal inactivation of organisms using absorptive materials or painting the
containers black (Sommer et al. 1997) to aid in the absorption of solar radiation.
Thermal enhancement has been achieved by (i) painting sections of bottles with
black paint (Martin-Dominguez et al. 2005), (ii) circulating water over a black
surface in an enclosed casing that is transparent to UVA light (Rijal et al. 2003),
and (iii) using a solar collector attached to a double glass envelope container
(Saitoh and El-Ghetany 2002). Solar reflectors can also increase the temperature of
water but not to the same extent as the use of absorptive materials or blackening/
coating of bottle surfaces (Mani 2006).
Optical inactivation of bacteria is a process of optical irradiation; solar
irradiance is optical irradiation. The incident solar irradiance on the outer Earth
atmosphere has an intensity of approximately 4,896 MJ/h/m2 (1,360 W/m2). This
value varies with geographical position within the elliptical sidereal orbit of
the Earth (the path that the Earth follows to revolve around the Sun during a
day or month or season). The irradiance intensity on a horizontal surface at the
ground level on the equator is reduced to approximately 1,120 kJ/s/m2 (1,120 W/
m 2) (averaged over the period of hours during which sunlight is available)
remaining after absorbed by atmospheric components including water vapor,
ozone, oxygen, and others. Thus, 1.12 MJ/s/m2 (1,120 W/m2) is the average solar
irradiance intensity reaching the Earth’s surface during sunny times and 1 s is the
unit exposure time of optical energy that is available in each second to inactivate
microbial matter. This value reduces in a cosine function as latitude increases
away from the equator (McGuigan et al. 2012). UVA and UVB reaching the
Earth’s surface play a major role in optical inactivation of the bacterial population.
This concept is good when applied to a zero turbid water that helps the sunlight
Solar Energy and Water/Wastewater Infrastructure 193

pass through or transmit deeper levels of water and can effectively reduce the
bacterial count.
Inactivation of bacteria by applying both thermal (heat) and optical irradiance
is called “combined thermal and optical inactivation.” The combined thermal and
optical inactivation is the mechanism involved in most SODIS systems. These
systems help the effective usage of both solar UV and IR for disinfection.
Compound parabolic collector (CPC) reflectors are widely used SODIS
systems. The walls of a CPC act as reflectors and the water-containing tube acts
as a receiver. This has been proven as a successful solar water disinfection system
to date (Mbonimpa et al. 2012). Research studies have proven that the use of
reflective materials gives better disinfection rates than the use of heat-absorbent
materials because of the effective use of sunlight by both direct (radiation from
the sun) and indirect (radiation from reflectors) means (Mani 2006). The material
used for reflecting radiation depends on the pricing of the material commercially
available. Research is ongoing for developing the most viable reflection material.
SODIS systems are effective in tropical regions, where sunlight is abundant.
This is also a good application for decentralized and low-capital-fund-available
regions.

WASTEWATER PROCESSING

Wastewater is generated from domestic, industrial, commercial, and agricultural


activities or by a combination of these. Pollutants include oxygen-demanding
substances, pathogens, nutrients, organic and inorganic chemicals, and water
temperature. EPA estimates that there are approximately a million miles of publicly
and privately owned sewer lines to collect and transfer wastewater from the source
to wastewater treatment plants. Treatment of wastewater before releasing it into
natural water bodies was started in the late 1800s (EPA 2004). The Clean Water Act,
which was passed in 1972, requires that all wastewater be treated before releasing it
into the watershed. Most wastewater treatment trains involve primary treatment,
secondary treatment, and tertiary treatment. Primary treatment involves the
removal of large floating objects by passing wastewater through screens. After
screening, water is passed into a grit chamber, where silt, sand, and small stones
settle to bottom and are removed. Secondary treatment removes about 85% of the
organic matter by making use of bacteria (EPA 2004); the treatment techniques
used are trickling filters and/or activated sludge processes. Effluent from primary
treatment is pumped into secondary treatment aeration tanks; aeration tanks are
where the primary treated sewage is mixed with air and bacteria-loaded sludge.
The mixture is then allowed to remain aerated for several hours during which
the bacteria breaks down the organic matter and then the liquid/solids mixture
is settled or separated. After the secondary treatment, the effluent is disinfected
and discharged to a water body as permitted. Chlorine is mostly used to disinfect
the water and then dechlorination is performed before releasing the effluent
194 Renewable Energy Technologies and Water Infrastructure

Figure 10-5. Energy consumption of a conventional activated sludge system.


Source: Guo et al. (2019).

into natural water bodies. Dechlorination is done for reducing the ill-effects of
chlorine on aquatic life (the flora and fauna). Heavy metals, nutrients, and toxic
compounds cannot be removed during primary and secondary treatments;
advanced wastewater treatment techniques include biological, chemical, and
physical techniques to remove/reduce said pollutants. A few of these techniques
are filtration, carbon adsorption, distillation, and RO. Effluents that undergo high-
level purification can be reused for industrial, recreational, and even for potable
purposes (based on the applicable regulatory requirements).
Similar to potable water treatment, the collection and treatment of wastewater
is also an energy-intense process operation(s). Collecting and moving of
wastewater to the treatment plant is mostly done via gravitation (sewer network),
but it requires pumping in some areas and results in high electricity consumption
to run pipe flows (the force mains). The pumping time is more during storm events
if sewers are combined. At the treatment plants, energy (in the form of electricity)
is consumed for the operation of aeration blowers, lifts, and return pumps. In
secondary treatment, 70% of the energy is consumed by lift pumps, biological
treatment, and disposal, with aeration being the high energy consumer summing
up to 60% (Qiao and Zhou 2018). Figure 10-5 shows the energy consumption of a
conventional activated sludge system (Guo et al. 2019).

SOLAR IN WASTEWATER PROCESSING

Solar energy has long been used for wastewater purification. Stabilization ponds
are where sewage flows through different aerobic, anaerobic, and facultative
Solar Energy and Water/Wastewater Infrastructure 195

pond systems as different stages or trains. Pollutant concentration is reduced by


natural processes. These natural processes include means of physical, chemical,
and biological and provide the required effluent quality; in general, various pond
trains are used at small wastewater systems worldwide. The major source of
energy for carrying out these processes is sunlight. Sunlight plays different roles
by supplying different forms of energy to these systems; thermal energy is used
for triggering chemical reaction(s), and the UV radiation from sunlight helps in
reducing or removing bacteria. Electricity generated from a solar PV can be used
for various pumping operations used to move wastewater or sewage. Sludge(s)
generated in wastewater treatment should be dewatered before disposal, and some
techniques use thermal energy to dry solids; solar thermal can be a best fit in this,
where WWTPs can utilize the thermal energy from the sun to reduce the water
content in sludge(s). A research study conducted by Mathioudakis et al. (2009)
on solar thermal drying of sludge showed convincing results in dewatering the
sludge and, thus, reducing the volume. The results were reported as the time taken
to reduce the moisture content to below 10% and volume reduction to 85%; in
summer, the target(s) was achieved in 7 to 12 days, whereas it took up to 32 days
in autumn. Solar drying is most effective in the summer months, but it can also be
made effective by making use of commercially available solar-absorbing materials
and optical catalysts such as TiO2 (DOE 2005). As some of the desalination
technologies can be used for wastewater treatment, incorporating solar energy
for power generation and utilizing the thermal energy from sunlight can reduce
the usage of fossil-fuel-generated electricity in wastewater treatment.

SUMMARY

Water and energy are two basic needs for humans and this nexus is very much
required. The demand for safe drinking water is rising because of the depletion
of conventional safe water resources (rivers, lakes, aquifers, and so on), and high
levels of environmental pollution are affecting natural water bodies. A highly
polluted source water requires a high standard treatment train and consumes
significant level(s) of energy. Electricity generation via conventional fossil-fuel
firing practices results in releasing significant amounts of GHGs. GHGs cause
climate change patterns, which, in turn, affects seasonal rainfall(s) and may
result in extended drought conditions in some parts of the world. Because of
the conversion of pervious to impervious surfaces in worldwide developmental
construction activities presently, aquifers are being significantly less recharged
and most of the rainwater flows into streams or rivers, causing havoc and then
loss to oceans. Solar renewable energy has proven to support water production
and traditional water or wastewater treatment. Incorporating solar as a renewable
energy resource into water and wastewater treatment not only assists in effective
water processing and production but also results in the reduced use of fossil-fuel-
generated electricity and, thus, reduction in GHG emissions to the atmosphere.
196 Renewable Energy Technologies and Water Infrastructure

References
Al-Karaghouli, A. A., and L. L. Kazmerski. 2013. “Energy consumption and water
production cost of conventional and renewable-energy-powered desalination processes.”
Accessed September 19, 2021. https://www.irena.org/-/media/Files/IRENA/Agency/
Publication/2012/IRENA-ETSAP-Tech-Brief-I12-Water-Desalination.pdf
Alkhudhiri, A., N. Darwish, and N. Hilal. 2012. “Membrane distillation: A comprehensive
review.” Desalination 287: 2–18.
Baker, K. H., J. P. Hegarty, B. Redmond, N. A. Reed, and D. S. Herson, 2002. “Effect
of Oxidizing Disinfectants (Chlorine, Monochloramine, and Ozone) on Helicobacter
pylori”, Applied and Environmental Microbiology, 68 (2): 981–4.
Bracken, N., J. Macknick, A. Tovar-Hastings, P. Komor, M. Gerritsen, and S. Mehta. 2013.
“The year of concentrating solar power electricity use and management in the municipal
water supply and wastewater industries.” WRF & EPRI.
Carns, K., and EPRI Solutions. 2005. “Bringing energy efficiency to the water & wastewater
industry: How do we get there?” In Presented at WEFTEC 2005, WEF, Alexandria VA.
CDC (Centers for Disease Control and Prevention). 2020. Accessed September 19, 2021.
https://www.cdc.gov/healthywater/drinking/public/index.html
DOE (US Department of Energy). 2014. “The year of concentrating solar power.” Accessed
September 24, 2020. https://www.energy.gov/sites/prod/files/2014/10/f18/CSP-report-
final-web.pdf.
Downes, A., and Blunt, T. P. 1877. “Researches on the effect of light upon bacteria and other
organisms.” In Proceedings of the Royal Society of London, 488–500.
Ebhota, W. S., and T.-C. Jen. 2018. “Efficient low-cost materials for solar energy applications:
Roles of nanotechnology.” In Recent developments in photovoltaic materials and devices,
edited by N. Prabaharan, M. A. Rosen, and P. E. Campana. London: IntechOpen.
EERE (Energy Efficiency & Renewable Energy). 2020a. “Solar photovoltaic cell
basics.” Accessed September 24, 2020. https://www.energy.gov/eere/solar/articles/
solar-photovoltaic-cell-basics.
EERE. 2020b. “Concentrating solar-thermal power basics.” Accessed September 24, 2020.
https://www.energy.gov/eere/solar/articles/concentrating-solar-power-basics.
EPA (US Environmental Protection Agency). 2020. “Information about public water systems.”
Accessed September 24, 2020. https://www.epa.gov/dwreginfo/infor​mation-about-pub​
lic-water-systems.
FSEC Energy Research Center. n.d. “How a PV system works.” Accessed September 24,
2020. http://www.fsec.ucf.edu/en/consumer/solar_electricity/basics/how_pv_system_
works.htm.
Godsey, W. 2020. “Fresh, brackish or saline water for hydraulic fracs: What are the
options?” Accessed September 22, 2021. https://www.epa.gov/sites/production/files/
documents/02_Godsey_-_Source_Options_508.pdf
Guo, Z., Y. Sun, S. Y. Pan, and P. C. Chiang. 2019. “Integration of green energy and
advanced energy-efficient technologies for municipal wastewater treatment plants.”
Int. J. Environ. Res. Public Health 16 (7): 1282.
IRENA (International Renewable Energy Agency). 2019. Renewable capacity statistics
2019. Abu Dhabi: IRENA.
Kelsey, R., H. Smith, D. Porter, G. Scott, and T. Siewicki, 2005. Bacterial Loading, BMP
Efficiency and Hydrology Modeling Turpentine Run. St. Thomas, USVI: St. Thomas,
Water Resources Research Institute Conference, University of the Virgin Islands.
Solar Energy and Water/Wastewater Infrastructure 197

Krishna, H. J. 2004. “Introduction to desalination technologies.” Accessed September


22, 2021. https://www.twdb.texas.gov/publications/reports/numbered_reports/doc/
R363/C1.pdf.
Mani, S. K. 2006. “Development and evaluation of small-scale systems for solar disinfection
of contaminated drinking water in India”, Ph.D. thesis, School of Applied Sciences,
Northumbria University.
Martin-Dominguez, A., M. T. Alarcón Herrera, I. R. Martin-Dominguez, and A. Gonzalez-
Herrera. 2005. “Efficiency in the disinfection of water for human consumption in rural
communities using solar radiation”, Solar Energy. 78 (1): 31–40.
Mathioudakis, V. L., A. G. Kapagiannidis, E. Athanasoulia, V. I. Diamantis, P. Melidis,
and A. Aivasidis. 2009. “Extended dewatering of sewage sludge in solar drying plants.”
Desalination 248: 733–739.
Mbonimpa, E. G., B. Vadheim, E. R. and Blatchley III. 2012. “Continuous-flow solar UVB
disinfection reactor for drinking water”, Water Research. 46 (7): 2344–2354.
McGuigan, K., R. Conroy, H. J. Mosler, M. du Preez, E. Ubomba-Jaswa, and P. Fernandez-
Ibañez. 2012. “Solar water disinfection (SODIS): A review from bench-top to roof-top”,
Journal of Hazardous Materials. 235–236: 29–46.
Meierhofer, Regula and M. Wegelin. 2002. Solar Water Disinfection – A guide for
the application of SODIS. Solar Water Disinfection: A Guide for the Application of
SODIS.
NAS (National Academy of Sciences). 2008. “Desalination: A national perspective.”
Accessed September 19, 2021. https://www.nap.edu/read/12184/chapter/1
NRC (National Research Council). 1982. Vol. 4 of Drinking water and health. Washington,
DC: National Academies Press.
NREL. 2020a. “Solar energy basics.” Accessed September 19, 2021. https://www.nrel.gov/
research/re-solar.html
NREL.2020b. “Solar research.” Accessed September 19, 2021. https://www.nrel.gov/solar/
NREL. 2020c. “Concentrating solar power basics.” Accessed September 19, 2021. https://
www.nrel.gov/research/re-csp.html.
Qiao, J. F., and H. B. Zhou. 2018. “Modeling of energy consumption and effluent quality
using density peaks-based adaptive fuzzy neural network.” IEEE-CAA J. Autom. Sin.
5: 968–976.
Quteishat, K., and M. Abu-Arabi. 2004. On promotion of solar desalination in the MENA
region. Muscat, Oman: Middle East Desalination Research Center.
Rashad, M., A. A. Elsamahy, M. Daowd, and A. Amin. 2015. “A comparative study on
photovoltaic and concentrated solar thermal power plants.” Recent Adv. Environ. Earth
Sci. Econ., 167–73.
Rijal, G. K., and R. S. Fujioka. 2003. “Use of reflectors to enhance the synergistic effects of
solar heating and solar wavelengths to disinfect drinking water sources”, Water Science
Technology. 48(11–12): 481–8.
Rosa G., Miller, L., and Clasen, T. 2010. “icrobiological effectiveness of disinfecting water
by boiling in uatemala”, The American Journal of Tropical Medicine and Hygiene. 82(3):
473–7.
Saitoh, T., and El-Ghetany, H. 2002. A pilot solar water disinfecting system: performance
analysis and testing. Solar Energy, 72: 261–269.
Shahzad, M. W., M. Burhan, D. Ybyraiymkul, and K. C. Ng. 2019. “Desalination with
renewable energy: A 24 hours operation solution.” In Water and wastewater treatment,
edited by M. Eyvaz. London: IntechOpen.
198 Renewable Energy Technologies and Water Infrastructure

Shatat, M., and S. B. Riffat. 2014. “Water desalination technologies utilizing conventional
and renewable energy sources.” Int. J. Low-Carbon Technol. 9 (1): 1–19.
Tonner, J. 2004. “Potential for thermal desalination in Texas.” Accessed September 22,
2021. http://www.twdb.texas.gov/publications/reports/numbered_reports/doc/r363/
c4.pdf.
Ullah, Ihsan, and Mohammad G. Rasul. 2019. “Recent Developments in Solar Thermal
Desalination Technologies: A Review.” Energies. (12)1: 119.
CHAPTER 11
Renewable Energy
Technologies for Water
Quality Monitoring
Varun K. Kasaraneni

INTRODUCTION

Freshwater is a limited resource and essential to keep the water bodies clean and
accessible because they are crucial for human survival, economic growth, and
environmental sustainability (Adu-Manu et al. 2017; Cosgrove and Loucks 2015).
At the heart of maintaining healthy water bodies and providing clean water to
citizens is water quality monitoring (WQM), which is a fundamental tool in the
management of freshwater resources. At its core, WQM serves several purposes,
such as the following:
• Determining the composition of water with reference to its physical, chemical,
biological, and radiological properties.
• Determining the levels of contaminants and identifying any existing
problems or any issues that could emerge in the future.
• WQM data help reveal the health of streams, rivers, and lakes and help
identify any changes or trends in the health of water bodies over a period.
• WQM data are extremely helpful in developing pollution prevention and
management strategies and emergency plans in case of events such as oil
spills and radiation leaks, and mass erosion.
• Regulatory compliance is a part of the requirement to meet a range of
water quality goals set by Acts such as the Clean Water Act (CWA) and Safe
Drinking Water Act (SDWA).
• Determining whether progress is being made in remediation projects and
conservation efforts.

199
200 Renewable Energy Technologies and Water Infrastructure

Traditionally, WQM has relied primarily on a manual approach, where a


user would collect one or more water samples and transport these samples to
a laboratory for subsequent analysis. This traditional approach to WQM had
several limitations including limited spatiotemporal variability and enormous
labor needs for maintaining a regional monitoring program. In addition, such
collection methods can introduce errors such as sample cross contamination and
misidentification (USGS 2006).
Over the last two decades, new technologies, such as sensors, laser-based
systems, fiber optics, and microelectronic mechanical systems, have addressed some
of these limitations. The arrival of automatic programmable samplers improved
on the sample collection using fixed sampling stations where water samples
are collected continuously or periodically. These technologies revolutionized
sample collection and enabled in situ measurement of several key water quality
parameters and supported long-term data acquisition and monitoring processes.
The recent addition of wireless and internet-enabled devices allows for remote
real-time visualization of the water quality data collected in lakes, streams, and
other water bodies (Adu-Manu et al. 2017; Pellerin et al. 2016). Continuous data
transmission of key water quality parameters ensures that water quality events are
not “missed or obscured” and rapidly responded to when necessary.
Collectively, the application of wireless technologies and remote sensing
to WQM are known as large-scale wireless monitoring networks (LSWMNs)
connected to an Internet of Things (IoT). An example of a common LSWMN setup
is shown in Adu-Manu et al. (2017). These advanced monitoring technologies may
be convenient but may have increased energy requirements because of the energy
demands of wireless communication (Patel and Patel 2016). This may result in the
use of bigger batteries and frequent replacement of batteries. Batteries can also be
unreliable because of untimely and irregular discharging that can result in a loss
of data, instrument calibration and so on. Further, replacing and maintaining
batteries for a large network of monitoring stations is time-consuming, labor
intensive, and challenging.
The use of in situ renewable energy sources such as solar panels and microbial
fuel cells (MFCs) that constantly recharge batteries can help avoid unnecessary
work, reduce or eliminate energy dependence on traditional carbon-based
electricity, and increase resiliency in the field. As energy resources become scarce,
the widespread deployment of large-scale networks will face trade-offs between
comprehensive “collect-it-all” data gathering and selective data-gathering
strategies.
Renewable energy systems equipped with energy-efficient modules can resolve
this dilemma because they have low carbon emission and power consumption.
They represent the next generation of smart water monitoring because they are
flexible, intelligent, and less likely to experience power interruptions due to low
system throughput leading to data gaps. This chapter will present the applications
of renewable energy technologies in WQM and discuss case studies in China,
Tanzania, and the United States involving the use of LSWMNs employing solar
panels and/or MFCs.
Renewable Energy Technologies for Water Quality Monitoring 201

SOLAR-POWERED MONITORING NETWORKS

LSWMNs are an essential tool for water quality managers. Automatic data streams
shared on wireless sensor nodes preclude the necessity of traveling to individual
collection points and collecting and preserving fragile samples for transport back
to laboratories for analysis. Some of the most common applications for these
systems include storm water sampling for nutrients, stream monitoring, and
tracking harmful algae blooms in lakes (Adu-Manu et al. 2017). LSWNs can be
used with any suitable sensor array that has Wi-Fi capabilities and, therefore, can
be customized to meet the specific needs of the community (Xie and Yuning 2018).

Stormwater Sampling and Monitoring


LSWNs are ideal for monitoring stormwater systems, particularly water systems
that still have combined sewer overflows. EPA (2016) updated the National Pollution
Discharge Elimination System (NPDES) to bring about more transparency in the
regulated water industry with a program called next-generation compliance. The
program strongly advocates for real-time monitoring with “less expensive mobile
technologies,” which are referred to as real-time control (RTC) (EPA 2018). The
EPA document highlights 11 case studies in the Appendix, including an RTC in
South Bend, Indiana, that was built to control their CSO overflow events. This
system was called CSOnet and worked on the basic framework of an energy-
efficient “dense wireless sensor actuator network” for monitoring and controlling
stormwater flow in the collection system during heavy rain events (Montestruque
and Lemmon 2015). The architecture is built around a Chasquimote, which is a
type of rugged sensor node installed in the sewer system, an electronic gated and
mounted above ground to collect signals from the sensor node, and a wide area
network (WAN) that retransmits data to a central database. Repeaters were used
to retransmit data when the spatial gaps in the system exceeded the reach of the
sensor (a sample WAN figure found in Montestruque and Lemmon 2015). The
system reduced CSO overflow events by 25% over an 8 year period starting in
2006. The power subsystem runs on a series of 13 V lithium batteries and switching
power supplies, but the entire system uses four D-sized batteries (6 V) that last
a year or more when sampling at 5 min intervals. The use of cellular service to
transmit data increased power consumption and reduced data transmission times
from 5 min intervals to once every 3 h. It was stated by the authors that the
network could be easily be powered by the electrical grid or solar cells.
This type of water monitoring network can also measure water quality
parameters and help manage, untreated stormwater during heavy precipitation
events. There are currently no operational storm sampling networks in the United
States or Canada that use a wireless solar network (WSN) set to monitor large
municipal water systems, but one can easily see how these systems can become
self-sufficient with renewable energy as the technology improves. However, there
is a functional prototype WQM system developed in Nanjing China (in 2015) that
collects traditional water quality parameters as opposed to just monitoring and
202 Renewable Energy Technologies and Water Infrastructure

Figure 11-1. Solar wireless sensor array used in the Tianbao Street eco-road project
in Nanjing, China.
Source: Xie and Yuning (2018).

controlling water levels (Xie and Yuning 2018). The Tianbao eco-road project is an
example of an intelligent monitoring system of stormwater implemented using the
“Internet of Things” and a 4G wireless sensor network. The system monitored site
hydrological conditions like rainfall, surface runoff, soil moisture content, and
total rainfall. Water quality parameters measured by the sensor array included
total nitrogen (TN), total phosphorus (TP), pH, turbidity, and suspended matter
(Xie and Yuning 2018) (Figures 11-1 and 11-2). This basic framework has the
potential for use in stormwater systems in the United States.
The research team concluded that an intelligent monitoring system powered
by solar panels can provide continuous accurate data for stormwater managers.
This project was published as a prototype, so the authors did not include
performance data on the solar array or sensor output, but, in general, it is well
known that solar arrays preform best with good continuous sun exposure. The
electric current draw will depend largely on the sensor’s electrical requirements.

Discharge and Nonpoint Source Monitoring


Monitoring of streams and lakes has become increasingly more important in
some areas of the United States where suburban development has accelerated and
urban runoff, mining runoff, fracking wastes, and industrial discharge threaten
the local watershed. The Susquehanna River Basin Project in Pennsylvania and
the Sunapee Lake Project in Vermont are both good examples of solar-powered
Renewable Energy Technologies for Water Quality Monitoring 203

Figure 11-2. Architecture of the wireless network used in the stormwater


management system in the ecological road of Tianbao Street, Nanjin, China.
Source: Xie and Yuning (2018).

LSWMNs. Traditionally, this type of sampling was performed by intensive grab


sampling and a heavy time commitment. Because the time commitment and lag
in sampling time could lead to the possibility of important events being missed,
which can lead to further degradation of the water. This is particularly problematic
in waterways that have been classified as high value or pristine.
The Susquehanna River Basin Project is a solar-powered wireless monitoring
network covering hundreds of stream miles in the Susquehanna River Valley
of Pennsylvania. The system was set up (in 2010) with the express purpose of
tracking changes in the water quality near fracked gas wells. Fracking can cause
an increase in ion content and suspended particles in water, so the network
sensors were designed to measure turbidity, dissolved oxygen (DO), pH, and
conductivity. There were 58 stations installed in priority watershed headwaters in
the Marcellus Shale Region that stretches from eastern New York and PA and then
down through Central PA. Each location consisted of a specially protected sonde
placed in a free-flowing stream. The sonde collects water quality parameters at 5
min intervals and is transmitted to a solar-powered data platform that transmits
data via cellular or satellite every 2–4 h (Figure 11-3). According to researchers,
the use of solar-powered panels eliminated the need for electrical hook-ups,
thereby enabling remote locations. The data generated from the sensors alerted
researchers of poor water quality and were made available to the public via web
204 Renewable Energy Technologies and Water Infrastructure

Figure 11-3. Wireless solar network station being installed by SRBC staff along
Hammond Creek, PA.
Source: SRBC (2014).

portal (SRBC 2014). At the end of 2015, the commission had collected enough data
to analyze the system’s usefulness to publish a non-peer-reviewed report.
The Sunapee Lake is a “Class A” surface waterbody in southern New
Hampshire covering 142 km2 with 35 tributaries feeding into it. The Class A rating
indicates that it is a pristine water body and, therefore, a source of clean drinking
water for people in its watershed. Increased turbidity, opacity, conductivity, and
cyanobacteria populations had become a commonly observed phenomenon by
community members living in the watershed. The Sunapee Lake Project was
formed (in 2012) to study the potential degrading impacts of land development
and increased percentage of impervious surfaces on the lake (Kelley 2016).
Water quality data on conductivity/chloride, total phosphorous, turbidity, and
pH were collected. One of the monitoring and sampling tools used in this study
was Teledyne’s ISCO brand portable sampler, which can collect samples based
on user-defined parameters to be picked up and analyzed later at the lab. This
ISCO sampler was powered by solar panels attached to an electric power box on
a pole. An onboard data logger captured preprogrammed intervals of turbidity
measurements. Because of a lack of funds to deploy a full wireless network at
key discharge locations, the researchers were unable to conclude the cause for
pollution. The efforts of the SRBC demonstrated that a wireless network is only as
good as the extent of the area in which it is employed, and the amount and quality
of the data they can collect.
Renewable Energy Technologies for Water Quality Monitoring 205

Water Quality Monitoring with Data Buoys


Long-term WQM on open water bodies on inland lakes is becoming increasingly
important, given the number of contaminated water bodies, the number of
sampling locations within a water body that need monitoring, and the sheer
number of contaminants. WQM data buoys provide a reliable and consistent
method to monitor multiple locations on large surface water bodies simultaneously
over long periods of time, get real-time data that are essential to make policy
decisions, and act on environmental and public health issues in a timely manner.
A buoy station, in general, has a floatation base and some type of solar
array arranged around a tower. Data buoys come in a variety of sizes and
shapes depending on the application but most include a base or buoy platform;
instruments for sensing, such as a sonde, anemometer, temperature probes, and
other sensors; a data logger for collecting information; a power supply to keep the
unit running and stable; and telemetry equipment for sending the collected data
to their destinations. A reliable power supply is required to run this equipment
properly, which can be achieved with the use of solar power to keep the system
running for a long period of time independent of an AC power source. This feature
allows the operator to deploy the buoy station in remote locations because it no
longer requires constant attention (Fawcett et al. 2006; Papoutsa et al. 2012). Solar
power also enables online systems and servers via various telemetry options for
sensor control and remote data access. This means the entire operation can run
with minimal demands for maintenance and without any wasteful and annoying
interruptions. The use of renewable energy technology, such as solar power in
buoys, is crucial in enabling real-time data capture and real-time access. The
practical effect of using solar power is that it enhances response times to an
unusual event or spots a contaminant in critical drinking water supplies or other
ecologically sensitive water sources where quality is high value.
These types of buoys are currently being used in monitoring programs across
the United States. At the Presque Isle Bay (PIB) on Lake Erie, the Regional Science
Consortium along with the Pennsylvania DEP has buoys powered by solar panels
that provide data about weather and water quality parameters [Figure 11-4(a)].
The water quality buoy in the PIB monitors for harmful algal blooms (HABs) and
records temperature, dissolved oxygen, pH, conductivity, turbidity, blue-green
algae, and total chlorophyll. The data from the buoy provide information about
ecological conditions and warn pets and beach goers from harmful exposure from
any potentially hazardous conditions. The data are live-streamed to a website and
are also available on a mobile app (RSC 2019).
In the Chesapeake Bay, 10 NOAA Fisheries buoys powered by solar panels
help marine scientists and fishermen to better manage, protect, and enjoy the
Bay’s marine resources [Figure 11-4(b)]. These 10 buoys use cell phone and
internet technology to share real-time data about the Bay’s weather and water
quality. Information on past and current water quality data including dissolved
oxygen, salinity, turbidity, and so on are made available through the portal. This
can help users plan for trips and make decisions while on water. The data obtained
206 Renewable Energy Technologies and Water Infrastructure

Figure 11-4. (a) Buoy system at Presque Isle Bay on Lake Erie, (b) NOAA fisheries buoy
in Chesapeake Bay, and (c) USGS monitoring buoy deployed on Lake Hopatcong, NJ.
Source: NOAA (2020), RSC (n.d.), USGS (2019a).

are also used by scientists to develop robust predictive models for water quality
in the Bay (NOAA 2020).
Another recent example is the USGS monitoring buoy deployed on Lake
Hopatcong, New Jersey [Figure 11-4(c)]. The buoy gathers near real-time data
on water quality parameters such as water temperature, specific conductance,
dissolved oxygen, organic matter, pH, turbidity, and pigments produced by
cyanobacteria (HABs). The buoy helps determine the severity of HABs and the
data gathered will help elucidate patterns of HAB development, duration, and
decline on the lake (USGS 2019a, b).
HABs are of increasing concern as they pose a threat to human and aquatic
ecosystem health and cause economic damage. Many unanswered questions
remain about the occurrence, environmental triggers for toxicity, and the ability
to predict the timing, duration, and toxicity of HABs (Graham et al. 2016). Data
gathered by solar-powered buoys can help answer these questions. The applications
of renewables in WQM are enabling predictions of water quality issues that could
prevent environmental and public health occurrences (Qin et al. 2015). Although
most of the aforementioned examples are near shore, new advanced buoys are
designed for offshore WQM, which comes with its own set of challenges. The
data buoys are equipped with cellular, spread spectrum radio, and satellite data
transmission to shore. These advanced communication features require more
power to provide uninterrupted service, which are in the form of large 10 W solar
panels mounted to the top of the buoy.

MICROBIAL FUEL CELLS FOR WATER QUALITY MONITORING

MFCs are recent technological innovation that operate on the simple idea of
collecting the chemical energy from the metabolic processes of microorganisms
and converting it into electricity. MFCs can run a large-scale wireless network
Renewable Energy Technologies for Water Quality Monitoring 207

Figure 11-5. A two-chamber microbial fuel cell diagram illustrating common


electrochemical and electro-microbiological processes occurring in the anaerobic
and aerobic compartments.
Source: Adapted from Du et al. (2007) as presented in Slate et al. (2019).

similar to their solar counterparts and are a potential solution to the problems
encountered in developing countries with densely packed urban areas lacking
both running water and electricity.
MFCs typically have an anode and a cathode connected by a resistance
module that measures the current generated by the biofilm. In most MFCs, the
anode and cathode are separated by an ionic exchange membrane (usually Nafion)
(Figure 11-5). The system relies on the dissimilative respiration of electroactive
microorganisms in an anaerobic environment transferring electrons via the
cathode to the anode. The electrons then travel through a circuit to generate
current and merge in the cathode electrode, which triggers the oxidation reduction
of electron acceptors like oxygen (Chen et al. 2017; Slate et al. 2019).
MFCs are self-sustaining because they employ a pump on the cathode side
to facilitate microorganism renewal to generate current. Any disturbances to this
system will change the amount of current generated. Simple MFCs are capable of
measuring biochemical oxygen demand (BOD), organic loads, and/or toxicants
by recording the resultant drop in voltage when those substances enter the
system. Complex MFCs use current generated by the microorganisms to power
temperature, pH, and dissolved oxygen sensors.
Most MFCs employ an ionic membrane that sits between the cathode and
the anode and functions as a BOD sensor. The membrane together with a pump
and required tubing can make this project costly in the long run. One way to
circumvent the problem related to the membrane is to design a membrane-less
system (Chouler et al. 2018). Chouler et al. proposed a screen-printed paper
microbial fuel cell (pMFC) that is based on the idea of “paper electronics” using
carbon-based electrodes for the anode and cathode (Adu-Manu et al. 2017;
208 Renewable Energy Technologies and Water Infrastructure

Cosgrove and Loucks 2015) (Figure 11-6). In the place of a membrane, the paper
acts a separator between the two electrodes. The capillary action of the paper
creates “autonomous microfluidics” that can be manipulated by merely changing
the structure of the paper. The paper structure acts as a pump that supplies
the cathode with a continuous source of fresh microbes to generate electricity.
Another advantage of this system is that no external potentiostat is necessary,
therefore obviating the need of an AC power supply.
This MFC design functions in the same way as a traditional MFC, in that
disturbances to the electroactive microorganism will cause a change in current,
signaling a pollution event. Its advantages lie in its low cost and low environmental
impact. This proof-of-concept model displayed better performance when the
paper was cross-linked with glyoxal acid to improve its tensile strength and

(a) (b)

(c) (d)

Figure 11-6. Membrane-less paper electronic MFC: (a) schematic of the pMFC and
electrical connection, (b) photograph of the pMFC, showing size, (c) principle of
operation of the pMFC, and (d) assembly of the fpMFC by folding two pMFCs back
to back.
Source: Chouler et al. (2018).
Renewable Energy Technologies for Water Quality Monitoring 209

increase the operational life of the device. Its performance improved when two
devices were stacked back to back (Chouler et al. 2018).
One of the challenges with an MFC-based sensor is that it produces power in
microwatts, which can be problematic when the system is designed to run a sensor
array instead of merely recording changes in voltage during a pollution event.
Chen et al. (2017) explored the use of a power management system in a traditional
pump and membrane MFC. They also employed a radio frequency activation
field in their design that remotely controlled the sensors switching them on when
needed (the figures can be found in Chen et al. 2017).
This MFC system employs a two-chamber system of manganese dioxide–
covered porous graphite on the cathode side and an aluminum alloy on the
anode side separated by an ionic membrane. The purpose of the two-chamber
system is to convert Mn(IV) oxide into Mn(II) oxidation. This reaction releases
two electrons that move from the anode to the cathode side generating a current
from the cathode to the anodic side. The system is self-sustaining only as long as
an oxygen reduction reaction can take place to keep generating electricity. The
best way to accomplish this is to keep the system fed with high levels of active
bacteria such as the Leptothrix discophora SP-6. The drawback of this system is
that high levels of bacteria must constantly be present for it to generate energy.
To compensate for the low power, they built a power management module that
employs the use of a super capacitor and a booster converter that helps boost the
power from 1.8 to 3.0 V. As proposed, the power management module can work
with as little 0.33 V input (the figures of power management modules can be
found in Chen et al. 2017).
This proof-of-concept MFC collected temperature and pH from one sensor
and temperature and dissolved oxygen from a second sensor. Conclusions from
field tests indicate that under optimum conditions it can reliably collect data at
a better rate than traditional manual collection regimens. Some drawbacks to
this system are that 20 W solar panels do not always collect enough energy to
power the gateway, humidity can cause hardware failures if not countered with
desiccant, the two probes cannot be read simultaneously because it will cause
a surge in voltage consumption, and uncovered components are susceptible to
wildlife grazing. Electroactive microorganisms require at least two hours to
accumulate enough energy to charge the power management module.
Another type of low-cost MFC is the bulk tank design that was explored
by Velasquez-Orta et al. (2017) in Tanzania, Africa, where unplanned urban
areas are growing rapidly in the absence of sewer and water systems. They tested
the concepts of a low-cost MFC placed in situ in wells dug to the level of the
groundwater to monitor fecal contamination (Figure 11-7). The concept is based
on the sediment microbial fuel cell (SMFC) used in ocean environments to
measure salinity. The team replicated the sediment/bulk liquid (SED/BL) design
and built three additional “galvanic” monitoring systems, namely, a sediment/
sediment liquid MFC (SED/SED), a bulk liquid/air MFC (BL/Air), and a bulk
liquid/ bulk liquid (BL/BL) (Velasquez-Orta et al. 2017). In these MFC models,
the cathode and anode were made of graphite felt and a stainless-steel mesh
210 Renewable Energy Technologies and Water Infrastructure

(a) (c)

(b) (d)

Figure 11-7. MFC bulk tanks: (a) sediment/sediment (SED/SED), (b) bulk liquid/air
(BL/air), (c) bulk liquid/bulk liquid (BL/BL), and (d) sediment/bulk liquid/SED/BL.
Source: Velasquez-Orta et al. (2017).

was employed as a current collector. The two electrodes were connected using a
resistor box. The electrodes were placed directly into the environmental media
(sediment and water) instead of using the traditional anaerobic reactor, pump, and
membrane setup. The SED/BL and BL/BL configurations produced the highest
electrical currents of the prototypes, which reinforce the idea of placing the
cathode in an environment where continuous bacteria growth if facilitated will
produce a continuous output of current. All the prototype models were responsive
to pollution events but failed to deliver reproducible results over time. One factor
effecting sensor output and overall system performance was temperature, which
is an environmental variable and is difficult to control in a low-cost system SMFC
(SED/BL).
The MFC-based sensor is an innovative solution to WQM that holds promise
as a proof-of-concept design. Its primary advantage is that it is a self-sufficient
system powered by abundant microorganisms that can be hooked to existing
large-scale monitoring networks and can be deployed in difficult conditions.
The MFC-based sensors with further refinement and mathematical optimization
of the current response could bring communities with no running water or
sanitation a first line of defense warning when fecal coliform bacteria render
Renewable Energy Technologies for Water Quality Monitoring 211

the water unfit for drinking. Some of the disadvantages of these systems are the
cost of materials, the ability to generate only micro voltages, and inconsistent
performances over test periods. These problems are being actively addressed today
as research teams work to develop low-cost systems that could become marketable
in the near future.

SUMMARY

Large wireless monitoring networks for WQM are becoming an integral part of our
water infrastructure. Many cities employ them to monitor and control stormwater
events to reduce and eliminate raw sewage entering open bodies of water in a bid to
protect public health. Wireless networks are also being used to monitor freshwater
resources, such as streams, rivers, and lakes, where industrial activity, mining,
fracking, and suburban development are becoming more prevalent. Solar-powered
wireless sensor networks make use of the Internet of Things (IoT) for data collection
through Wi-Fi, cellular, and satellite links. In countries where cost is a factor, MFCs
may become advanced enough over time to be deployed in communities where
housing development has outpaced sanitary sewer infrastructure. The drawbacks
of the renewable energy wireless networks include the following: security of the
system (cellular service is vulnerable to denial-of-service attacks, for example),
reliability and life of solar panels, costs of the systems, and interruptions in data
service owing to instrument failure. Despite these challenges, LSWMNs in WQM
are institutionalized in governmental WQM goals and are seen as playing an
integral role in delivering safe clean drinking water to communities around the
world and protecting public heath over the long term.

References
Adu-Manu, K. S., C. Tapparello, W. Heinzelman, F. A. Katsriku, and J.-D. Abdulai. 2017.
“Water quality monitoring using wireless sensor networks: current trends and future
research directions.” ACM Trans. Sens. Netw. 13 (1): 4:1–4:41.
Chen, Q., Y. Liu, G. Liu, Q. Yang, X. Shi, H. Gao, et al. 2017. “Harvest energy from the
water: A self-sustained wireless water quality sensing system.” ACM Trans. Embedded
Comput. Syst. 17 (1): 3:1–3:24.
Chouler, J., Á Cruz-Izquierdo, S. Rengaraj, J. L. Scott, and M. Di Lorenzo. 2018. “A screen-
printed paper microbial fuel cell biosensor for detection of toxic compounds in water.”
Biosens. Bioelectron. 102: 49–56.
Cosgrove, W. J., and D. P. Loucks. 2015. “Water management: Current and future
challenges and research directions.” Water Resour. Res. 51 (6): 4823–4839.
Du, Z., H. Li, and T. Gu. 2007. “A state of the art review on microbial fuel cells: A promising
technology for wastewater treatment and bioenergy.” Biotechnology Advances 25 (5):
464–482.
EPA. 2016. Compendia of next generation compliance examples in water, air, waste, and
cleanup programs. Overviews and Factsheets. Washington, DC: EPA.
212 Renewable Energy Technologies and Water Infrastructure

EPA. 2018. Smart data infrastructure for wet weather control and decision support. Overviews
and Factsheets. Washington, DC: EPA.
Fawcett, A., S. Bernard, G. C. Pitcher, T. A. Probyn, and A. du Randt. 2006. “Real-time
monitoring of harmful algal blooms in the southern Benguela.” Afr. J. Mar. Sci. 28 (2):
257–260.
Graham, J. L., N. M. Dubrovsky, and S. M. Eberts. 2016. Cyanobacterial harmful algal
blooms and U.S. Geological Survey science capabilities. Open-File Report, USGS
Numbered Series. Reston, VA: US Geological Survey.
Kelley, K. 2016. “Lake Sunapee Data Collection Guide.” In Vermont EPSCoR Symp.,
Burlington, VT.
Montestruque, L., and M. D. Lemmon. 2015. “Globally coordinated distributed storm
water management system.” In Proc., 1st ACM Int. Workshop on Cyber-Physical Systems
for Smart Water Networks, 1–6. New York: Association for Computing Machinery.
NOAA. 2020. “Locations | Chesapeake Bay Interpretive Buoy System.” Accessed September
19, 2021. https://buoybay.noaa.gov/locations
Papoutsa, C., A. Kounoudes, M. Milis, L. Toulios, A. Retalis, K. Kyrou, et al. 2012.
“Monitoring turbidity in Asprokremmos dam in Cyprus using earth observation and
smart buoy platform.” J. Eur. Water 38: 25–32.
Patel, K. K., and S. M. Patel. 2016. “Internet of things-IOT: Definition, characteristics,
architecture, enabling technologies, application & future challenges.” Int. J. Eng. Sci.
Comput. 6 (5): 6122–31
Pellerin, B. A., B. A. Stauffer, D. A. Young, D. J. Sullivan, S. B. Bricker, M. R. Walbridge,
et al. 2016. “Emerging tools for continuous nutrient monitoring networks: Sensors
advancing science and water resources protection.” JAWRA J. Am. Water Resour. Assoc.
52 (4): 993–1008.
Qin, B., W. Li, G. Zhu, Y. Zhang, T. Wu, and G. Gao. 2015. “Cyanobacterial bloom
management through integrated monitoring and forecasting in large shallow eutrophic
Lake Taihu (China).” J. Hazard. Mater. 287: 356–363.
RSC (Regional Science Consortium). 2019. “WQData LIVE.” Accessed November 29, 2019.
https://v2.wqdatalive.com/public/55.
RSC. n.d. “Nearshore Buoy System at Presque isle Bay.” Accessed September 6, 2020.
https://www.regsciconsort.com/research/current-projects/nearshore-buoy-system/.
Slate, A. J., K. A. Whitehead, D. A. C. Brownson, and C. E. Banks. 2019. “Microbial fuel
cells: An overview of current technology.” Renewable Sustainable Energy Rev. 101: 60–81.
SRBC (Susquehanna River Basin Commission). 2014. “SRBC remote water quality
monitoring network—Reports.” Accessed September 6, 2020. https://www.srbc.net/
our-work/fact-sheets/docs/remote-water-quality.pdf.
USGS. 2006. Chapter A4. Collection of water samples. Techniques of Water-Resources
Investigations, USGS Numbered Series. Reston, VA, 231: USGS.
USGS. 2019a. “USGS monitoring buoy deployed on Lake Hopatcong, New Jersey.”
Accessed September 6, 2020a. https://www.usgs.gov/media/images/monitor-buoy.
USGS. 2019b. “Water quality data viewer.” Accessed November 29, 2019. https://ny.water.
usgs.gov/maps/habs/.
Velasquez-Orta, S. B., D. Werner, J. C. Varia, and S. Mgana. 2017. “Microbial fuel cells for
inexpensive continuous in-situ monitoring of groundwater quality.” Water Res. 117:
9–17.
Xie, M.-K., and C. Yuning. 2018. “A framework for the intelligent monitoring system of
stormwater management based on the internet of things and wireless sensor networks.”
J. Digital Landscape Archit. 3: 310–318.
CHAPTER 12
Integrating Renewable Energy
in Water Infrastructure: Global
Trends and Future Outlook
Juneseok Lee, Tamim Younos

INTRODUCTION

Increased urbanization, population growth, aging built infrastructure, emerging


contaminants, competitive water uses, and measures to mitigate the effects of
climate change in urban areas all highlight the urgent need to expand and enhance
our water infrastructure, including water supply and distribution networks,
wastewater treatment, and stormwater management (Larsen et al. 2016; NRC 2008;
Younos et al. 2019; Ruberto et al. 2013). These centralized water infrastructure
elements are highly energy dependent and account for 4% to 10% of the nation’s
total energy use, most of which is generated from fossil-fuel–based sources, with
consequently significant economic, social, and environmental impacts (CRS 2013;
Sanders and Webber 2012). To effectively address the immediate and long-term
energy requirements of water infrastructure, a paradigm shift toward a more
sustainable management of water and energy issues is critical in built environments
(Lee et al. 2018). The advantages of water conservation strategies and energy-use
efficiency, although obvious, are limited due to the ever-increasing demand for
water that will inevitably result in higher energy consumption (Tanverakul and
Lee, 2016). Therefore, it is reasonable to consider whether integrating renewable
energy into water infrastructure could be a viable way to reduce dependency on
fossil-fuel–based electricity use in the water sector.
Many researchers (Lee et al. 2018; Dixon et al. 1999; Pidou et al. 2007;
Agudelo-Vera et al. 2011, 2012a, b; Walsh et al. 2014; Dallman et al. 2016;
Tavakol-Davani et al. 2016) have observed that both practical applications and
research developments in the domain of renewable energy applications for water
infrastructure are rapidly evolving areas that encompass both technology and
science. At present, however, there are only limited examples of renewable energy

213
214 Renewable Energy Technologies and Water Infrastructure

use to power water infrastructure or other industrial applications reported in the


literature. In this chapter, we consider how renewable energy is being utilized for
applications such as water infrastructure elements around the world, focusing
specifically on solar and wind energy resources (Lee and Younos 2018). We
conclude by addressing the outlook for renewable energy applications in terms of
decentralized water–energy systems and ongoing research efforts.

BACKGROUND

Water and Energy Nexus


The water and energy nexus refers to the interdependency of water and energy
production systems (CRS 2013; Sanders and Webber 2012; Lawson et al. 2014).
Although water use is a critical factor for energy production and power generation
(Younos 2012; Younos et al. 2016), water infrastructure, which includes potable
water production, supply/distribution, and wastewater treatment/discharge,
consumes significant amount of energy. According to EIA (2019), fossil fuels
(petroleum, natural gas, and coal) are sources for more than 80% of the energy
used in the United States every year. Conventional urban water infrastructure
systems typically involve extensive centralized networks serving large populations.
The major characteristics include, but not limited to, a dependency on imported
water supplies, an increasing demand for wastewater collection and treatment
because of population growth, an increased level of impervious cover leading
to greater stormwater runoff, and a complex and often chaotic system of pipe
networks delivering potable water to consumers and transporting wastewater and
stormwater runoff away from urban areas (Dallman et al. 2016; Lee et al., 2019).
At present, more than 52,000 conventional or centralized potable water
treatment and distribution systems operate in the United States, with most being
powered by conventional energy distribution systems (USEPA 2009). In 2009, the
electricity consumed by the nation’s public drinking water and wastewater utilities
for pumping, conveyance, treatment, distribution, and discharge amounted to
203.8 PJ (56.6 billion kWh); about 4% of the total national energy consumption
is attributed to water and wastewater services (CRS 2013). The energy used for
water treatment and delivery in the United States is reported to be in the range
of approximately 252 to 3,312 kJ/m3 (0.07 to 0.92 kWh/m3) (depending on the
source water quality and topography) with an estimated average of 1,368 kJ/m3
(0.38 kWh/m3) (AWWARF 2007, Kloss 2008). In California, 19% of the water-
related energy use is met by electricity and 30% by natural gas, with the remainder
being satisfied by burning 88 billion gal. (333 million m3) of diesel fuel every year
(Krebs 2007).
The energy demand for water infrastructure is projected to increase by
approximately 30% over the next few decades because of increased urban water
demand and the need to develop energy-intensive alternative water sources
such as saltwater and reclaimed wastewater for potable purposes (Stokes and
INTEGRATING RENEWABLE ENERGY IN WATER INFRASTRUCTURE 215

Horvath 2005). At present, our water infrastructure relies heavily on electricity


generated from fossil fuels, which is accompanied by constantly increasing levels
of carbon dioxide (CO2) emissions. Kloss (2008) estimated that an average of
about 300 kg of CO2 is created for every cubic meter of water delivered.

Water and Energy Conservation


Major consumers of potable water include commercial facilities, public facilities
such as government buildings and schools, and those living in apartment
buildings, dormitories, and private residences (Tanverakul and Lee 2016). For
each of these consumer groups, there are opportunities for water conservation
that could result in considerable energy savings, thus mitigating the high-
carbon footprint associated with water consumption (Younos and Parece 2012).
Consumer-related energy-use efficiency practices include water conservation
achieved by modifying consumer behaviors (Parece et al. 2013; Tanverakul and
Lee 2016) and the increased use of in-building water- and energy-saving fixtures
(e.g., EPA 2014).
A water- and energy-integrated decentralized infrastructure that reduces
energy use attributed to long-distance potable water transport is another option.
Ward et al. (2012) and Younos et al. (2016) discussed the potential energy savings
that could be achieved by lowering potable water consumption; the reduction
in CO2 levels attributed to greater use of rainwater harvesting systems has also
been reported (Dallman et al. 2016). Green-building designs can ameliorate much
of this adverse impact by incorporating practices that improve the efficient use
of water by implementing better design, operation, and maintenance across a
building’s entire life cycle (Cassidy 2003; USGBC 2006; Corbett and Muthulingam
2007). From an energy conservation perspective, however, the advantages of all
the water and energy conservation strategies noted previously may be limited
because of the ever-increasing demand for water that can only result in higher
energy consumption. Reducing the dependency of water infrastructure on fossil
fuels should, thus, be a critical objective in water infrastructure planning and
design (Lee and Younos 2018).

APPLICATIONS OF RENEWABLE ENERGY IN THE WATER


INDUSTRY

Solar Energy
At present, solar energy is a proven way to integrate renewable energy into large-
scale applications such as water and wastewater treatment and the desalination
of saltwater for potable purposes. Currently, the most promising solar energy
technology consists of photovoltaic (PV) arrays made of silicon chips that convert
solar energy into electricity through the transfer of electrons. In the United States,
several water utilities are already at least partially powered by solar energy. An
excellent example is the New Jersey American Water Canal Road Water Treatment
216 Renewable Energy Technologies and Water Infrastructure

Plant, which was installed in 2005. The system includes two 810 MJ/h (225 kW)
alternating current inverters and an internet-based data-acquisition system. The
original solar array consisted of 2,871 solar PV modules, each rated at 630 kJ/h
(175 W) for a total direct current (DC) output of 1.8 GJ/h (502 kW). In 2007,
the system was expanded by 313 MJ/h (87 kW) (a 17% increase) to achieve an
overall output of 2.1 GJ/h (590 kW). A further expansion of 392 MJ/h (109 kW)
DC was added atop the filter basins in 2008 to increase the overall capacity of the
site to 2.5 GJ/h (698 kW) DC. The solar array currently satisfies approximately
20% of the Canal Road Wastewater Treatment Plant’s peak usage (Lelby and
Burke 2011). Tables 12-1 and 12-2 show the estimated annual energy savings and
CO2 emission reductions attributed to solar energy, respectively, for water and
wastewater treatment plants for a similar project, the State of Massachusetts Pilot
Program (EPA 2009). Note that these are selected example facilities in the State
of Massachusetts.
Water scarcity in many parts of the world has increased the demand for
desalination of seawater and brackish water, particularly in high-population
coastal cities and island countries. Desalination technologies are highly energy
intensive (Younos and Tulou 2005), so particularly in sunnier parts of the world,

Table 12-1. Estimated Energy Savings and CO2 Reduction Attributed to Renewable
Energy Upgrades at Drinking Water Treatment Facilities, Massachusetts.

Estimated
Renewable energy Total annual annual CO2
Water treatment generation (up to) energy savings reduction
facility MJ/h (kW) GJ (kWh) (kg)
Ashland Howe Solar—162 (45) 700 (194,464) 233 × 103
Street Water
Treatment Plant
Easton Water Solar—180 (50) 216 (60,000) 47 × 103
Division
Falmouth Long Solar—54 (15) 1,002 (278,200) 216 × 103
Pond Water
Treatment Plant
Lee Water Solar and hydro­ 723 (200,940) 155 × 103
Treatment Plant electric—378 (105)
New Bedford— Solar—497 (138) 594 (165,000) 168 × 103
Quittacus Water
Treatment Plant
Townsend Water Solar—144 (40) 266 (73,844) 57 × 103
Treatment Plant
Worcester Water Solar and hydro­ 1,991 (553,152) 430 × 103
Treatment Plant electric—576 (160)
Source: Adapted from EPA (2009).
INTEGRATING RENEWABLE ENERGY IN WATER INFRASTRUCTURE 217

Table 12-2. Examples of Estimated Savings from Renewable Energy Upgrades at


Wastewater Treatment Facilities, Massachusetts.

Estimated
Renewable energy Total annual annual CO2
Wastewater treatment generation energy savings reduction
facility MJ (kW) TJ (kWh) (kg)
Charles River Pollution Solar—72 (20) 2.5 (705,300) 567 × 103
Control District
Great Lawrence Solar—1,476 (410) 17.7 (4,909,062) 5,420 × 103
Sanitary District
Pittsfield Wastewater Solar and 15.3 (4,255,737) 3,252
Treatment Plant biomass—6,372
(1,770)
Upper Blackstone Solar—1,440 (400) 3.0 (831,615) 636
Water Pollution
District
Source: Adapted from EPA (2009).

renewable energy technologies provide a significant opportunity to boost the


production of freshwater (Abou-Rayan and Djebedjian 2014). It is noted that
there are significant differences between direct-tied renewables and grid-tied
renewables in terms of intermittency challenges, not to mention storage capacities
and controls. In this chapter, we present all cases for the purpose of showing the
overall trend. Table 12-3 shows examples of desalination facilities powered by
solar and PV energy in various countries.

Wind Energy
Wind power rotates wind turbines, creating mechanical energy that can
be converted to electrical energy. Wind turbines come in both vertical axis
arrangements and multiple axis, horizontal arrangements. Turbines utilizing
wind energy for low power [36 to 360 MJ/h (10 to 100 kW)], medium power [360
to 1,800 MJ/h (100 kW to 0.5 MW)], and high power [>1,800 MJ/h (>0.5 MW)]
generation are mature technologies (Garcia-Rodriquez 2002). In the United States,
several water utilities are powered by wind energy. For example, the Washington
Suburban Sanitary Commission uses wind energy to power one-third of its
water and wastewater operations (Lelby and Burke 2011). Table 12-4 shows the
estimated annual energy savings and reduction in CO2 emissions attributed to
wind energy for selected wastewater treatment plants in the State of Massachusetts
Pilot Program (EPA 2009). Table 12-5 provides examples of desalination facilities
powered by wind energy around the world. Again, note that these are selected
example facilities.
218 Renewable Energy Technologies and Water Infrastructure

Table 12-3. Desalination Plants Utilizing Direct and Indirect Solar Energy.

Type of solar Desalination Plant capacity


Location energy technology (m3/day)
El Paso, Texas Solar pond MSF 16.19
Yanbu, Saudi Arabia Dish collectors FS 199.96
La Desired Island, French Solar-evacuated ME 40.01
Caribbean tube
Abu Dhabi, UAE Solar-evacuated ME 119.98
tube
Takami Island, Japan Solar-parabolic ME 15.99
trough
Almeria, Spain Solar-parabolic ME-Heat 71.99
trough Pump
Margarita de Savoya, Italy Solar pond MSF 49.99–59.99
Near Dead Sea Solar pond MED 2,999.61
Source: Garcia-Rodriquez (2002), Abou-Rayan and Djebedjian (2014).
Note: MSF = multiple effect evaporation; FS = freeze separation; ME = multiple effect; and
MED = multiple effect distillation.

Table 12-4. Examples of Estimated Savings from Renewable Energy Upgrades at


Wastewater Treatment Facilities in Massachusetts.

Wastewater Renewable energy Total annual energy Estimated annual


treatment generation saving CO2 reduction
facility GJ/h (kW) TJ (kWh) (kg)
Barnstable Wind and solar— 3.06 (850,000) 825 × 103
Wastewater 3.6 (1,000)
Treatment
Plant
Falmouth Wind—11.34 15.25 (4,235,000) 3,181 × 103
Wastewater (3,150)
Treatment
Plant
Source: EPA (2009).

OUTLOOK

Future applications of solar and wind energy technologies clearly include their
potential integration in green, decentralized water infrastructure systems. We
define a Decentralized Green Water-Infrastructure System (DGWIS) as one
that integrates locally available water sources (i.e., rainwater, greywater, and
groundwater) with renewable local energy sources (i.e., solar and wind) to support
INTEGRATING RENEWABLE ENERGY IN WATER INFRASTRUCTURE 219

Table 12-5. Examples of Desalination Plants Incorporating Wind Energy.

Power
generated Desalination Plant capacity
Location MJ (kW) technology (m3/day)
Shark Bay, Western 115 (32) RO—brackish water 129.98–167.98
Australia
Ruegen Island, Germany 720 (200) MVC 119.98–299.96
Source: Garcia-Rodriquez (2002), Abou-Rayan and Djebedjian (2014).
Note: RO = reverse osmosis; and MVC = mechanical vapor compression.

the treatment and distribution of both potable and nonpotable water for a wide
range of buildings, including residential, industrial, commercial, government,
and office premises (Lee et al. 2018; Lee and Younos 2018). Rainwater harvesting
and greywater recycling have long been identified as alternative water sources
for the sustainable management of water resources (Dixon et al. 1999; Pidou
et al. 2007; Agudelo-Vera et al. 2011, 2012a, b; Walsh et al. 2014; Dallman et al.
2016; Tavakol-Davani et al. 2016), and the general characteristics of graywater
and rainwater harvesting systems have been studied by many researchers (Pidou
et al. 2007; Eriksson et al. 2009; Li et al. 2009; Malinowski et al. 2015; NWRI 2015).
Younos (2014) noted that recent technological advances in prefiltration,
first-flush design, and the availability of small-scale water treatment units mean
that captured rainwater could be much more widely used as a potable water
source. Advances in small-scale and packaged water treatment technologies,
such as reverse osmosis, carbon filters, and UV disinfection devices, allow small-
scale decentralized water production systems to be installed as satellite systems
within buildings to treat and use locally available water sources, including
captured rainwater and reclaimed graywater. A typical small-scale packaged
water treatment system with a capacity of up to 50,000 L per day can easily be
configured as a water treatment unit at the individual building level in urban
areas (Younos 2014). It is noted that plumbing codes can limit some options in
the residential environment.
DGWIS will incorporate advanced small-scale water treatment technologies
based on patterns of anticipated water use. The production and consumption of
energy at the individual building level for harvesting and treating water on site
are expected to increase service reliability and technical efficiency and reduce
environmental impacts by decreasing the energy required. An added benefit would
be that decentralized systems should also improve levels of service by reducing
service interruptions in the transmission and distribution networks. It is our
contention that alternative technologies, specifically DGWIS, can provide at least
a partial solution to the water–/energy supply and water quality challenges facing
communities across the United States and around the world (Lee and Younos,
2018). The potential contributions of self-sufficiency and higher wastewater
recovery rates for sustainable water resource management have been highlighted
220 Renewable Energy Technologies and Water Infrastructure

in several recent studies (Rygaard et al. 2011; Agudelo-Vera et al. 2011, 2012a).
We, therefore, seek to chart a new paradigm shift in water–energy management
by evaluating and optimizing the capture and reuse of rainwater/greywater to
augment domestic water supplies, while at the same time reducing water-related
energy usage by adopting renewable energy technologies in buildings (Lee and
Younos 2018).
Recently, one of us had the opportunity to visit the ReNEWW house in West
Lafayette, Indiana, where real-time flow and temperature monitoring devices are
installed on every fixture (sampling the flow once every second) for both hot
and cold water as well as ambient air monitoring. Within the ReNEWW house,
it is possible to adjust the water heater storage capacity from 0.19 to 1.14 m3 (50
to 300 gal.) within the house simply by switching the drinking water source
from municipal to rainwater harvesting systems and implementing in-building
filtration/UV treatment for rainwater. The house’s water systems are powered
by solar panels installed on the rooftop. Even though, at present, the primary
function of the ReNEWW house is to perform research, it is clear that in the
future, the results of this research will be put to good use in support of our vision
of self-efficient/net-zero emissions/maximum recycling/deep-sustainability living
even at the residential level.
As noted previously, a paradigm shift in technology/science is also affecting
centralized supply systems. For example, more efficient energy use in water and
wastewater facilities/utilities represents a major step toward energy conservation
and reducing the amount of fossil-fuel–based electricity used in the water industry
(Lelby and Burke 2011; EPA 2010). Al-Smairan (2012) describes a case study that
uses PV solar energy to power a remote area groundwater pumping station in
Jordan, reporting that PV water pumping systems are more cost-effective than
diesel engines in powering the pumping systems at the case study site.
Although we strongly advocate for the use of renewable energy (in this
chapter), we must be cautious in selecting and utilizing the renewable energy
systems, with respect to (and not limited to): higher upfront cost, intermittency in
power generation, storage capacity, and geographic limitations. As the technology
is progressing rapidly, it is our hope that the said limitations will be insignificant in
using renewable energy in the years to come (Lee and Younos, 2018). In addition, it
is clear that further research, development, and applications of effective renewable
energy technologies (including the applicable socioeconomic aspects) will have
a significant potential to reduce the energy used by and carbon costs associated
with water infrastructure maintenance, operations, and updating.

SUMMARY

• In this chapter, we discussed renewable energy applications in water


infrastructure, focusing primarily on utilizing solar and wind energy
resources both within the United States and around the globe. It is clear
INTEGRATING RENEWABLE ENERGY IN WATER INFRASTRUCTURE 221

that solar and wind energy can provide significant opportunities for water
infrastructure engineering that could help develop and shape next-generation
technologies and urban water management strategies.
• Advantages of water conservation strategies and energy-use efficiency,
although obvious, are limited because of the constantly increasing water
demand that will result in higher energy consumption. Therefore, integrating
renewable energy into water infrastructure should be a priority objective for
reducing current water infrastructure dependency on the fossil-fuel–based
electricity.
• It is clear that both demand- and supply-side operations could benefit from
applying renewable energy within the water industry. As noted previously,
DGWIS could offer a way forward for effecting improvements in the energy
supply mechanisms for future water infrastructure developments. However,
it is only to be expected that critical issues, such as social acceptance and
broader applicability and the robustness of the new systems, will arise as
novel systems come into service. Both practical applications and the latest
research developments in the domain of integrating renewable energy into
existing water infrastructure are rapidly evolving and deserve far greater
attention from the research community and the water industry as a whole.

References
Abou-Rayan, M., and B. Djebedjian. 2014. “Advances in desalination technologies: Solar
desalination.” In Vol. 30 of Potable water—Emerging global problems and solutions. The
Handbook of Environmental Chemistry, edited by T. Younos and C. A. Grady, 181–211.
Heidelberg, Germany: Springer.
Agudelo-Vera, C. M., W. Leduc, A. R. Mels, and H. Rijnaarts. 2012a. “Harvesting urban
resources towards more resilient cities.” Resour. Conserv. Recycl. 64: 3–12.
Agudelo-Vera, C. M., A. Mels, K. Keesman, and H. Rijnaarts. 2012b. “The urban harvest
approach as an aid for sustainable urban resource planning.” J. Ind. Ecol. 16 (6): 839–850.
Agudelo-Vera, C. M., A. R. Mels, K. J. Keesman, and H. H. M. Rijnaarts. 2011. “Resource
management as a key factor for sustainable urban planning.” J. Environ. Manage. 92 (10):
2295–2303.
Al-Smairan, M. 2012. “Application of photovoltaic array for pumping water as an
alternative to diesel engines in Jordan Badia, Tall Hassan Station: Case study.” Renewable
Sustainable Energy Rev. 16 (7): 4500–4507.
AWWARF (American Water Works Association Research Foundation). 2007. Energy
index development for benchmarking water and wastewater utilities. Denver: AWWARF.
Cassidy, R., ed. 2003. White paper on sustainability: A report on the green building
movement. A supplement to Building Design & Construction, November issue. Accessed
June 24, 2020. http://archive.epa.gov/greenbuilding/web/pdf/bdcwhitepaperr2.pdf.
Corbett, C. J., and S. Muthulingam. 2007. Adoption of voluntary environmental standards:
The role of signaling and intrinsic benefits in the diffusion of the LEED green building
standards. Los Angeles: UCLA Institute of the Environment and Sustainability.
CRS (Congressional Research Service). 2013. Energy–water nexus: The water sector’s energy
use. 7-5700. Washington, DC: CRS.
222 Renewable Energy Technologies and Water Infrastructure

Dallman, S., A. M. Chaudhry, M. K. Muleta, and J. Lee. 2016. “The value of rain: Benefit
cost analysis of rainwater harvesting systems.” Water Resour. Manage. 30: 4415.
Dixon, A., D. Butler, and A. Fewkes. 1999. “Water saving potential of domestic water reuse
systems using greywater and rainwater in combination.” Water Sci. Technol. 39 (5):
25–32.
EIA (Energy Information Administration). 2019. “U.S. primary energy consumption
by energy source, 2019.” Energy Encyclopedia, Institute for Energy Research, EIA.
Accessed June 24, 2020. https://www.eia.gov/energyexplained/us-energy-facts/.
EPA (Environmental Protection Agency). 2009. Massachusetts energy management pilot
program for drinking water and wastewater case study. EPA-832-F-09-014. Washington,
DC: EPA.
EPA. 2010. Evaluation of energy conservation measures for wastewater treatment facilities.
EPA 832-R-10-005. Washington, DC: EPA.
EPA. 2014. WaterSense: An EPA partnership program. Washington, DC: EPA.
Eriksson, E., H. R. Andersen, T. S. Madsen, and A. Ledin. 2009. “Greywater pollution
variability and loadings.” Ecol. Eng. 35 (5): 661–669.
Garcia-Rodriquez, L. 2002. “Seawater desalination driven by renewable energies: A
review.” Desalination 143 (2): 103–113.
Kloss, C. 2008. Managing wet weather with green infrastructure—Municipal handbook—
Rainwater harvesting policies. EPA-833-F-08-010. Washington, DC: EPA.
Krebs, M. 2007. Water-related energy use in California. Assembly Committee on Water,
Parks and Wildlife, Public Interest Energy Research Program. Sacramento, CA: California
Energy Commission.
Larsen, T. V., S. Hoffmann, C. Luthi, B. Truffer, and M. Maurer. 2016. “Emerging solutions
to the water challenges of an urbanizing world.” Science 352 (6288): 928–933.
Lawson, S. E., Q. Zhang, M. Joshi, and P. Tzu-Han. 2014. “The effects of water–energy
nexus on potable water supplies.” In Vol. 30 of Potable water—Emerging global problems
and solutions. The Handbook of Environmental Chemistry, edited by T. Younos and C.
A. Grady, 125–151. Dordrecht, Netherlands: Springer.
Lee, J. and Younos, T. 2018. “Sustainability Strategies at the Water–Energy Nexus:
Renewable Energy and Decentralized Infrastructure.” Journal American Water Works
Association, February 2018. https://doi.org/10.1002/awwa.1001
Lee, J., K. Bae, and T. Younos. 2018. “Conceptual framework for decentralized green water-
infrastructure systems.” Water Environ. J. 32 (1): 112–117.
Lee, J., G. Güngör Demirci, and J. Keck. 2019. “Drinking water infrastructure.” In
Encyclopedia of water: Science, technology, and society. Hoboken, NJ: Wiley.
Lelby, V. M., and M. E. Burke. 2011. Energy efficiency best practices for North American
drinking water utilities. Denver: Water Research Foundation.
Li, F., K. Wichmann, and R. Otterpohl. 2009. “Review of the technological approaches for
greywater treatment and reuses.” Sci. Total Environ. 407 (11): 3439–3449.
Malinowski, P. A., A. S. Stillwell, J. S. Wu, and P. M. Schwarz. 2015. “Energy–water nexus:
Potential energy savings and implications for sustainable integrated water management
in urban areas from rainwater harvesting and gray-water reuse.” J. Water Resour. Plann.
Manage. 141 (12): A4015003.
NRC (National Research Council). 2008. Urban stormwater management in the United
States. Washington, DC: National Academies Press.
NWRI (National Water Research Institute). 2015. Framework for direct potable reuse. Los
Angeles: NWRI.
INTEGRATING RENEWABLE ENERGY IN WATER INFRASTRUCTURE 223

Parece, T. E., L. Grossman, and E. S. Geller. 2013. “Reducing carbon footprint of water
consumption: A case study of water conservation at a University Campus.” In Vol. 25
of Climate Change and water resources. The Handbook of Environmental Chemistry,
edited by T. Younos and C. A. Grady, 199–218. Heidelberg, Germany: Springer.
Pidou, M., F. A. Mamon, T. Stephenson, B. Jefferson, and P. Jeffrey. 2007. “Greywater
recycling: treatment options and applications.” Proc. Inst. Civil Eng. Eng. Sustainability
160 (3): 119–131.
Ruberto, A., J. Lee, and A. Bayer. 2013. “Water–energy nexus analysis of a public university
in California.” Water Effic. 8 (3): 36–41.
Rygaard, M., P. J. Binning, and H. J. Albrechtsen. 2011. “Increasing urban water self-
sufficiency: A new era, new challenges.” J. Environ. Manage. 92 (1): 185–194.
Sanders, K. T., and M. E. Webber. 2012. “Evaluating the energy consumed for water use in
the United States.” Environ. Res. Lett. 7 (3): 034034.
Stokes J. H., and A. Horvath. 2005. “Life cycle energy assessment of alternative water
supply systems.” Int. J. Life Cycle Assess. 11 (5): 335–343.
Tanverakul, S. A., and J. Lee. 2016. “Decadal review of residential water demand analysis
from a practical perspective.” Water Pract. Technol. 11 (2): 433–447.
Tavakol-Davani, H., S. Burian, J. Devkota, and D. Apul. 2016. “Performance and cost-based
comparison of green and gray infrastructure to control combined sewer overflows.” J.
Sustainable Water Built Environ. 2 (2): 04015009.
USGBC. 2006. Accessed September 19, 2021. https://www.usgbc.org/resources/
international-association-plumbing-and-mechanical-officials-publication-iapmo-ansi-
upc-1-2
Walsh, T., C. Pomeroy, and S. Burian. 2014. “Hydrologic modeling analysis of a passive,
residential rainwater harvesting program in an urbanized semiarid watershed.” J.
Hydrol. 508: 204–253.
Ward, S., D. Butler, and F. A. Memon. 2012. “Benchmarking energy consumption and CO2
emissions from rainwater harvesting systems: An improved method by proxy.” Water
Environ. J. 26 (2): 184–190.
Younos, T. 2012. “Water dependency of energy production and power generation systems.”
Water Resour. Impact 14 (1): 9–12.
Younos, T. 2014. “Bottled water: global impacts and potential.” In Vol. 30 of Potable water –
Emerging global problems and solutions. The Handbook of Environmental Chemistry,
edited by T. Younos and C. A. Grady, 213–227. Heidelberg, Germany: Springer.
Younos, T., J. Lee, and T. Parece. 2019. “Twenty-first century urban water management:
the imperative for holistic and cross-disciplinary approach.” J. Environ. Stud. Sci. 9 (1):
90–95.
Younos, T., K. O’Neill, and A. McAvoy. 2016. “Carbon footprint of water consumption in
urban environments: Mitigation strategies.” In Vol. 47 of Sustainable water management
in urban environments. The Handbook of Environmental Chemistry, edited by T.
Younos and T. E. Parece, 33–56. Heidelberg, Germany: Springer.
Younos, T., and T. E. Parece. 2012. “Water use and conservation.” In 21st century geography:
A reference handbook, edited by J. Stoltman, 446–456. Los Angeles: Sage.
Younos, T., and K. E. Tulou. 2005. “Overview of Desalination Techniques.” J. Contemp.
Water Res. Educ. 132 (1): 3–10.
Appendix

LIST OF ACRONYMS AND ABBREVIATIONS

°C Degrees Celsius
°F Degrees Fahrenheit
A Ampere
AC Alternating current
ACE Affordable clean energy
ACEE American Council for an Energy-Efficient Economy
AD Anaerobic digester
AEP American electric power
AFC Alkaline fuel cell
Amp hrs Ampere hours
BBD Biomass-based diesel
BCAP Biomass Crop Assistance Program
BEWT Built environment wind turbine
BOD Biochemical oxygen demand
BOEM Bureau of Ocean Energy Management
Btu British thermal unit
CAA Clean air act
CAES Compressed-air energy storage
CAFES Corporate average fuel economy standard
CapEx Capital expenditure
CBO Congressional Budget Office
CD Controlled drainage
CdTe Cadmium telluride
CFR Code of Federal Regulations
CH4 Methane
CHP Combined heat and power
CIGS Copper-indium-gallium-selenide
CO Carbon monoxide
CO2 Carbon dioxide
COD Chemical oxygen demand
CPP Clean power plan
CPV Concentrated photovoltaic
CREBs Clean renewable energy bonds
CRS Congressional Research Service
c-Si Crystalline silicon
CSOs Combined sewer overflows

225
226 Appendix

CSP Concentrated solar power


CT-PURA Connecticut Public Utilities Regulatory Authority
DAF Dissolved air flotation
DC Direct current
DEQ Department of Environmental Quality
DF Dark fermentation
DG Distributed generation
DLA Designated leasing areas
DO Dissolved oxygen
DWM Drainage water table management
EBPR Enhanced biological phosphorus removal
ED Electrodialysis
EERS Energy efficiency resource standards
EGS Enhanced geothermal system
EIS Environmental impact statement
EISA Energy Independence and Security Act
EPAct Energy Policy Act
EPCA Energy Policy and Conservation Act
EUEI European Union Energy Initiative
FERC Federal Energy Regulatory Commission
FLPMA Federal Land Policy and Management Act
FOG Fats, oil, and grease
FPA Federal Power Act
ft Foot/feet
FWPA Federal Water Power Act
gal. Gallon(s)
gal./year Gallon(s) per year
GDP Gross domestic product
GHGs Greenhouse gases
GJ Giga joules
GOR Gain output ratio
gpm Gallons per minute
GW Gigawatts
GWEC Global Wind Energy Council
GWP Global warming potential
H2S Hydrogen sulfide
HABs Harmful algal blooms
HAPs Hazardous air pollutants
HAWT Horizontal axis wind turbines
HDPE High-density polyethylene
HHV Higher heat value
hr Hour
HRT Hydraulic retention-time
ICE Internal combustion engine
IDALS Iowa Department of Agriculture and Land Stewardship
Appendix 227

IEC International Electrotechnical Committee


IoT Internet of Things
IOUs Investor-owned utilities
IOWA DNR Iowa Department of Natural Resources
IPP Independent power producer
IRENA International Renewable Energy Agency
ISOs Independent system operators
ISU Iowa State University
ITC Investment tax credit
ITDG International Technology Development Group
J Joules
J/hr. Joules per hour
JAXA Japan Aerospace Exploration Agency
kg Kilogram
KOH (Alkaline) potassium hydroxide
kPa kilopascal
KPIs Key performance indicators
kW kilowatt
kWh Kilowatt hour
KY-NREPC Kentucky Natural Resources and Environmental Protection
Cabinet
lb Pound
LFG Landfill gas
LOG Local Government Outreach Stakeholder Group
LSWMNs Large-scale wireless monitoring networks
LUC Land-use changes
MACRS Modified accelerated cost recovery system
MA-DER Massachusetts Department of Energy Resources
MCFC Molten carbonate fuel cell
MD Membrane distillation
MEC Microbial electrolysis cells
MED Multieffect evaporation/distillation
MFCs Microbial fuel cells
MPCA Minnesota Pollution Control Agency
MSDS or, SDS (Material) safety data sheets
MSF Multistage flash distillation
MSW Municipal solid waste
MTBE Methyl tertiary butyl ether
MTF Mississippi River/Gulf of Mexico Watershed Nutrient Task Force
MVC Mechanical vapor compression
MW Megawatt
MWh Megawatt hour
NBC Narragansett Bay Commission
NCSL National Conference of State Legislators
NEA National Energy Act
228 Appendix

NECPA National Energy Conservation Policy Act


NERC North American Electric Reliability Corporation
NF3 Nitrogen trifluoride
NGO Non-government organization
NIMBY Not in my backyard
NOAA National Oceanic and Atmospheric Administration
NOX Nitrogen oxides
NPDES National pollutant discharge elimination system
NPS Nonpoint source
NREL National Renewable Energy Laboratory
NSPS New source performance standards
NTM Normal turbulence model
O&M Operation and maintenance
OCS Outer continental shelf
OH− Hydroxyl ion
OpEx Operational expenditure
ORC Organic Rankine cycle
OTEC Ocean thermal energy conversion
PAFC Phosphoric acid fuel cell
PEMFC Proton exchange membrane fuel cell
PEMFC Polymer electrolyte membrane fuel cell
pMFC Paper microbial fuel cell
PNNL Pacific Northwest National Laboratory
PPA Power purchase agreement
ppm Parts per million
PR Performance ratio
psi Pounds per square inch
PSRP Processes to significantly reduce pathogens
PTC Production tax credit
PUC Public Utility Commission
PURPA Public Utility Regulatory Policies Act
PV Photovoltaic
PVC Polyvinyl chloride
REC Renewable energy certificate
RECs RE credits or certificates
RECs Renewable energy credits
RES Renewable electricity standard
RES Renewable energy standards
RF Radio frequency
RFS Renewable fuel standard
RIEDC Rhode Island Economic Development Corporation
RNG Renewable natural gas
RO Reverse osmosis
ROW Right-of-way
RPS Renewable portfolio standards
RTCs Real time controls
Appendix 229

RTOs Regional transmission operators


RTS Remote temperature sensor
RUE Right-of-use and easement
RVP Reid vapor pressure
s Second
SEIA Solar Energy Industries Association
SF6 Sulfur hexafluoride
SIPs State implementation plans
SO2 Sulfur dioxide
SOFC Solid oxide fuel cell
SPIN Self-powered intelligent networks
SRT Solids retention time
SS7 Signaling System No. 7
TBL Triple bottom line
TFPZ Turbine failure projectile zone
TI Turbulence intensity
TN Total nitrogen
TP Total phosphorus
TRI Toxics release inventory
TVC Thermal vapor compression
UNDP United Nations Development Programme
UN-IPCC United Nations Intergovernmental Panel on Climate Change
UPS Uninterruptable power supplies
USC US Code
USDA United States Department of Agriculture
USDA NASS USDA National Agricultural Statistics Service
USDOE United States Department of Energy
USEPA United States Environmental Protection Agency
USGS United States Geological Survey
UTIA University of Tennessee Institute of Agriculture
VAs Volatile acids
VAWT Vertical axis wind turbines
VC Vapor compression
VDC Volt direct current
VFAs Volatile fatty acids
VOCs Volatile organic compounds
VRE Variable renewable energy
VSEC Virtual security
VSred Volatile solids’ reduction
WAN Wide area network
WAS Waste activated sludge
WNA World Nuclear Association
WQM Water quality monitoring
WRRFs Water resource recovery facilities
WSN Wireless solar networks
ZLD Zero liquid discharge
Index

Note: Page numbers followed by f and t indicate figures and tables.

absorbed glass mat (AGM), 74 billion gallons (Bgal. or BG), 113


acronyms and abbreviations, 225–229 biochemical oxygen demand
activated sludge process, 115 (BOD), 207
advanced wastewater treatment biodiesel, 53; classes of, 55; diesel
techniques, 194 engine, 53; Pacific Biodiesel, 54;
Aermotor: pumping installation, 169f; processing, 54–55, 54f;
windmill, 168f. See also wind energy production from feedstocks, 56f;
Affordable Clean Energy (ACE), 4 transesterification reaction, 53, 55f;
Agricultural Act of 2014, 3. See also US biodiesel production, 55–57,
US renewable energy policy 56f. See also ethanol
AgSTAR program, 99 biofuels, 7–8, 39; advanced, 45, 57.
alkaline fuel cells (AFCs), 115, See also biodiesel; ethanol;
125–126, 125f. See also fuel cells renewable energy resources
alternating current (AC), 61, 184 biogas, 114; in circular
American Recovery and Reinvestment economy, 107–110; -distributed
Act (ARRA), 109 energy systems, 110; generation
anaerobic digesters (ADs), 94; systems, 94–96; methane makeup
complete-mix, 99f; in biogas system, 93; on-farm
configurations, 97f; designs for production, 96; production
farm-operated, 100f; in US Farms pathways, 107; -to-electric
from 2000 to 2018, 100f generation, 102–105. See also
anaerobic digestion (AD), 114; biogas-to-energy renewable systems;
advantages, 115; biogas from 2000 fuel cells; hydrogen
to 2018, 101f; covered anaerobic biogas-to-energy renewable
lagoon, 98f; livestock and poultry systems, 93, 110; AgSTAR
manure, 96–100, 105–107; plug-flow program, 99; anaerobic digestion
anaerobic digestor, 98f. See also configurations, 97f; carbon
biogas-to-energy renewable systems intensity, 107; contaminates, 107;
anaerobic digestion system diluents, 107; feedstocks, biogas
(ADS), 106 and energy generation, and end
Appalachian Electric Power (AEP), 76 use, 106f; livestock and poultry
applicant, 29 manure, 96–100, 105–107;
artificial intelligence (AI), 179 microturbine and reciprocating
engine comparison, 104t; natural
best management practices gas, 102; renewable natural gas, 108;
(BMPs), 49 RNG supply chain for wastewater

231
232 Index

biogas-to-energy renewable Clean Power Plan (CPP), 3; of


systems (cont.) 2015, 3–4. See also US renewable
treatment plants, 108f; sludges/ energy policy
biosolids and biogas-to-energy, Clean Water Act (CWA), 193, 199
101–102; Ten State Standards, 95; Coanda effect hydro-shear screen, 68
uses of biogas from 2000 to Code of Federal Regulations (CFR), 31
2018, 101f. See also anaerobic combined heat and power systems
digesters; biogas (CHP systems), 93, 103, 116
biohydrogen, 116; combined thermal and optical
infrastructure, 131; production inactivation, 193
methods, 117f. See also hydrogen Common Channel Signaling System
Biomass Crop Assistance Program No. 7, 179
(BCAP), 3 compound parabolic collector
Biomass Energy and Alcohol Fuels (CPC), 193
Act, 2. See also US renewable compressed-air energy storage
energy policy (CAES), 159
biomethane. See biogas concentrated photovoltaic (CPV), 32,
biophotolysis, 120–121, 120f. See also 184
hydrogen concentrated solar power (CSP), 32,
Bipartisan Budget Act of 2018, 6. 160, 184; and solar PV, 187t
See also US renewable energy Congressional Budget Office (CBO), 46
policy Consolidated Appropriations Act
brine, 137; disposal, 144. See also of 2016, 5. See also US renewable
desalination technologies energy policy
building block scenarios (BBS), 48 contaminants, 107. See also biogas-to-
built environment wind turbine energy renewable systems
(BEWT), 166 controlled drainage (CD), 52
buoy station, 205 copper–indium-gallium–selenide
Bureau of Land Management (CISG), 8, 184
(BLM), 31–33 corn-growing states, 49
Bureau of Ocean Energy Management Corporate Average Fuel Economy
(BOEM), 33–34; regulatory (CAFE), 3
roadmap, 34f corporate average fuel economy
business as usual (BAU), 47 standard (CAFES), 40
cyanobacteria, 121
capital expenditure (CapEx), 20, 110
carbon intensity, 107. See also biogas- dark fermentation (DF), 115, 117–118.
to-energy renewable systems See also hydrogen
cellulases, 44 Database of State Incentives for
cellulosic ethanol processing, 43–44, Renewables & Efficiency (DSIRE), 35
44f. See also ethanol data buoys, 205
Chasquimote, 201 Decentralized Green Water-
chemical oxygen demand (COD), 116 Infrastructure System (DGWIS), 218
Class A rating, 204. See also solar- Department of Energy (DOE), 8, 34
powered monitoring networks Department of Environmental
Clean Air Act (CAA), 40 Quality (DEQ), 29
Index 233

desalination processes, 135, 147, electrodialysis (ED), 135, 190


189; comparison and potential Emergency Economic Stabilization
applications, 139t–140t; criteria for Act of 2008, 6. See also US
selecting energy source, 141f; criteria renewable energy policy
for selecting technology, 143f; energy energy efficiency resources standards
requirement comparison, 136f; (EERS), 4. See also US renewable
membrane technologies, 190; energy policy
principles, 135; renewable energy Energy Independence and Security
integration with, 138, 141–142; Act (EISA), 40, 57; of 2007, 3;
selection process of, 142–143, volume renewable standards set
143f; solar-powered, 190–191; by, 41t. See also US renewable
technologies, 136–138; technology energy policy
sustainability, 144–147; thermal energy-intensive alternative water
technologies, 189. See also solar sources, 214
energy and water/wastewater Energy Policy Act (EPAct), 2, 40;
infrastructure 1992, 2; 2005, 2–3. See also US
desalination technologies, 136; renewable energy policy
brine disposal and management Energy Policy and Conservation Act
alternate, 144, 145f; criteria for (EPCA), 2
selecting, 143f; economics of energy recovery turbine (ERT), 176
renewable energy-driven, 146–147; enhanced biological phosphorus
energy efficiency of thermal removal (EBPR), 117
desalination, 137; environmental enhanced geothermal system
impacts of desalination (EGS), 157, 159. See also geothermal
processes, 144; freshwater energy
costs, 146f; multieffect environmental impact statement
distillation, 136–137; multistage (EIS), 33
flash desalination, 136; ethanol, 39; annual nitrate plus nitrite
regulatory requirements, 147; loads, 51f; annual total nitrogen
reverse osmosis, 137–138; social loads, 51f; annual total phosphorus
aspects of desalination, 147; loads, 52f; cellulosic ethanol
surface discharge, 145–146; processing, 43–44, 44f; classes
sustainability of, 144–147; vapor of, 41–42; corn processing, 42;
compression, 137 dry milling for corn ethanol, 43f;
designated leasing areas (DLAs), Energy Policy Act, 40; full lifecycle
31, 32 corn ethanol GHG emissions for
diesel engine, 53. See also biodiesel 2014, 47f; global ethanol production
diluents, 107. See also biogas-to- 2007–2017, 40f; greenhouse
energy renewable systems gases, 46–48; legislation, 40–41;
direct current (DC), 61, 184, 216 modern-day ethanol industry, 39;
dissolved oxygen (DO), 7, 203 phosphorus sources, 49; pros and
dry milling process, 43f. See also cons, 52; sources of phosphorus
ethanol and nitrogen in MRB, 49, 50f; US
ethanol plant capacity, 45f; US
Electric Consumers Protection Act of ethanol production and consumption
1986, 6 2000–2018, 44–46, 45f;
234 Index

ethanol (cont.) geothermal brine, 158; geothermal


volume renewable standards direct use, 157; geothermal
set by EISA, 41t; water quality power capacity and additions by
impacts, 48–52; wet milling for country, 156f; geothermal power
corn ethanol, 42f. See also biodiesel generation, 154; geothermal power
European Union Energy Initiative plant technology distribution, 155f;
(EUEI), 78 geothermal resources, 152;
hybridization opportunities, 160;
Federal Energy Regulatory hydrothermal reservoir, 152,
Commission (FERC), 1, 17, 34–35 152f; mineral recovery, 158–159;
federal investment tax credit, 5. operating capacity of geothermal
See also US renewable energy policy power plants by system type, 157f;
Federal Land Policy and Management organic Rankine cycle/binary, 156;
Act (FLPMA), 31 perspective, 160–161; power plant
Federal Power Act (FPA), 1 types, 154; subsurface temperatures
Federal Water Power Act (FWPA), 1 across the United States, 151f;
fermentation technologies, 131 technological distribution, 157;
Food, Conservation, and Energy Act technologies for geothermal power
of 2008, 3. See also US renewable plants, 154f; temperature ranges for
energy policy different power block systems, 155f;
freshwater, 199; resources, 211 use, 157; weaknesses and strengths
fuel cells (FCs), 113, 121, 131; of CSP and, 160f; worldwide
alkaline, 125–126, 125f; biomodule, capacity, 156–157
UF, and microbial, 116f; Geothermal Energy Act, 2. See also
classification, 121; components, 121; US renewable energy policy
hydrogen, 124–125; hydrogen geothermal power plant, 154; capacity
and hydrocarbon, 127; internal and additions by countries, 156f;
reforming, 128; 10MGD dry steam, 154–155; flash, 155;
WRRF, 115f; microbial, 130–131, operating capacity of, 157f;
130f; molten carbonate, 127–129, organic Rankine cycle/binary, 156;
128f; phosphoric acid, 126, 127f; technologies for, 154f; technology
polymer electrolyte membrane, 124– distribution, 155f; temperature
125, 124f; solid oxide, 129–130, 129f; ranges for different power block
in wastewater infrastructure, 122t– systems, 155f
123t. See also biogas; hydrogen global warming potential (GWP), 144
gram-positive bacteria, 117
gain output ratio (GOR), 137 greenhouse gases (GHGs), 4, 17,
geothermal brine, 158 46–48, 93
geothermal energy, 151; applications groundwater (GW), 188
based on resource temperature, 153f; Gulf’s hypoxic zone, 48
current status, 156; direct heating GW under direct influence of surface
applications, 158; dry steam, water (GWUDISW), 188
154–155; energy storage, 159;
enhanced geothermal systems, 159; harmful algal blooms (HABs), 205
flash, 155; future direction, 159; hazardous air pollutants (HAPs), 102
Index 235

heating, ventilation, and air- investor-owned utilities (IOUs), 18


conditioning (HVAC), 115 Iowa Department of Natural
high-density polyethylene (HDPE), 67 Resources (Iowa DNR), 49
higher heat value (HHV), 103, 104
high-lift VAWT (Hy-VAWT), 166 landfill gas (LFG), 17
horizontal axis wind turbines landowner, 30
(HAWTs), 165–167; source in the land-use changes (LUC), 46
US, 165f; vs. VAWT, 165–167, 173t. large-scale wireless monitoring
See also wind energy networks (LSWMNs), 200, 201
House bill, 5; HR 8444 1977–78. Leading by Example (LBE), 4
See National Energy Act of 1978. Local Government Outreach
See also US renewable energy policy Stakeholder Group (LOG), 29
HR 8444 1977–78. See National low-emission practices, 48
Energy Act of 1978
hydraulic ram, 61, 62f material safety data sheets (MSDSs), 8
hydraulic retention time (HRT), 94 mechanical vapor compression
hydrogen, 113, 116, 131; (MVC), 135
biohydrogen, 116; biohydrogen megawatt (MW), 31; capacity fee, 31;
production methods, 116–117, rate, 32
117f; biophotolysis, 120–121, 120f; membrane distillation (MD), 135, 190
dark fermentation, 117–118; fuel membrane fuel cells (MCFCs), 115.
cells, 124–125; and hydrocarbon See also fuel cells
fuel cells, 127; microbial methyl tertiary butyl ether (MTBE), 39
electrolysis cells, 118, 118f; microalgae, 121
photofermentation, 119–120, 119f; microbial electrolysis cells
production from algal biomass, 119. (MECs), 116, 118, 118f. See also
See also biogas; fuel cells hydrogen
hydropower, 6–7; electricity, 61; microbial fuel cells (MFCs), 12,
plant size and capacity, 62t. See also 17, 116, 116f, 130–131, 130f, 200,
renewable energy resources 206; -based sensor, 210–211; bulk
hydrothermal reservoir, 152, 152f. tanks, 210f; challenges with MFC-
See also geothermal energy based sensor, 209; low-cost,
209–210; membrane-less paper
impulse turbines, 69. See also electronic, 208f; paper microbial
micro-hydropower fuel cell, 207; two-chamber
independent power producer microbial fuel cell, 207f. See also
(IPP), 25 fuel cells; water quality monitoring
independent system operators micro-hydropower, 61, 80; basic
(ISOs), 18 concept of, 64f; battery-based micro-
internal reforming, 128. See also fuel hydropower system, 74f; bill of
cells materials, 81t–89t; case study project
International Electrotechnical at Glen Alton, 65f; case study project
Committee (IEC), 174 cost, 75; case study site, 64–65;
Internet of Things (IoTs), 179, 200, 211 components of hydraulic ram, 62f;
investment tax credit (ITC), 5 cost–benefit analysis, 75–76; cost
236 Index

micro-hydropower (cont.) Mississippi River/Gulf of Mexico


summary, 75t; decision matrix for Watershed Nutrient Task Force
turbine–generator interface, 71t; (MTF), 48
design components, 65; electrical Modified Accelerated Cost Recovery
component design, 71–75, 71f; System (MACRS), 6. See also US
for electricity generation, 61; renewable energy policy
emerging technologies, 79–80; flow molten carbonate fuel cells, 127–129,
control weir at case study site, 67f; 128f. See also fuel cells
generation, 62–63; HL-100 air multieffect distillation (MED), 136–
heater dump-load controller, 73f; 137, 189. See also desalination
hydraulic component design, 65– technologies
68; hydraulic head, 66f; hydropower multistage flash desalination (MSF
in India, 77; impulse turbines, 69; desalination), 135, 136, 189. See also
lucid energy pipe system, 79f; desalination technologies
LV1500 four-nozzle turbine, 70f; municipal solid waste (MSW),
magnet electric generator 26, 28
connected to Turgo turbine
wheel, 72f; mechanical component Narragansett Bay Commission
design, 68–71; MorningStar Tri-45 (NBC), 103
battery overcharge controller, 75f; National Academy of Engineering
operation of Coanda effect screen (NAE), 163
unit, 69t; outback GTFX3048 National Energy Act (NEA), 1
inverter, 73f; penstock, 63; projects National Energy Act of 1978 (HR 8444
in Nepal, 77–78; regulatory 1977–78), 2. See also US renewable
requirements, 76; selection matrix energy policy
for penstock pipe material, 68t; National Energy Conservation Policy
small-scale hydropower system, 63f; Act (NECPA), 2
small-stream, 78–79; surge National Pollution Discharge
protector, 73f; system design, 64, Elimination System (NPDES),
66t; turbine flow control, 70f; 147, 201
turbine-generator system, 70f; natural gas, 102. See also biogas-to-
Turgo impulse wheel, 69; universal energy renewable systems
UB4D 12 V, 200AH sealed AGM net capacity factor, 32
battery, 74f; as viable technology in New Stationary Sources (NSPS), 102
developing countries, 76–78 nongovernment organizations
micro turbines, 103 (NGOs), 99
million megawatt hours (MWh), 100 normal turbulence model (NTM), 174
Minnesota Pollution Control Agency North American Electric Reliability
(MPCA), 49 Corporation (NERC), 18
Mississippi River Basin (MRB), 48; “Not In My Backyard (NIMBY)”, 147
annual nitrate plus nitrite load, 51f;
annual total nitrogen load, 51f; ocean thermal energy conversion
annual total phosphorus load, 52f; (OTEC), 180
sources of phosphorus and on-farm biogas production, 96. See also
nitrogen, 49, 50f. See also ethanol biogas-to-energy renewable systems
Index 237

operational expenditure (OpEx), 20, real-time control (RTC), 201


110 reciprocating engines, 104
operation and maintenance RE credits or certificates (RECs),
(O&M), 20, 75, 188 20, 25
operator, 30 regional transmission operators
optical inactivation of bacteria, 192 (RTOs), 18
optical irradiation, 192 Reid vapor pressure (RVP), 41
outer continental shelf (OCS), 33 reliable energy, 164
owner, 30 remote temperature sensor
(RTS), 74
Pacific Biodiesel, 54. See also biodiesel renewable electricity generator, 25
Pacific Northwest National Renewable Electricity Standard
Laboratory (PNNL), 34 (RES), 4
packaged water treatment renewable energy (RE), 17, 19;
technologies, 219 Energy Policy Act of 2005, 20;
paper microbial fuel cell (pMFC), 207. Energy Policy requirements, 20;
See also microbial fuel cells procurement of, 20; production
penstock, 63. See also and consumption of the United
micro-hydropower States, 18f; programs, 33–34;
performance ratio (PR), 137 project technologies, 19;
phosphoric acid fuel cells (PAFCs or resources, 17; standards, 20–25;
PFCs), 115, 121, 126, 127f. See also systems, 17; types of, 18; US-code,
fuel cells law, and act, 19–20. See also US
photofermentation, 119–120, 119f. renewable energy generation
See also hydrogen technology mix
photosynthetic pigments (PS renewable energy certificate
pigments), 120 (REC), 20, 25
photovoltaic (PV), 5, 17, 32, 138, 215 Renewable Energy Initiatives, 2.
polymer electrolyte membrane fuel See also US renewable energy policy
cell, 124–125, 124f. See also fuel cells renewable energy in water
polyvinyl chloride (PVC), 67 infrastructure, 213, 214, 220–221;
power purchase agreement (PPA), 25 annual energy savings and CO2
Presque Isle Bay (PIB), 205 emission reductions, 216t, 217t;
production tax credit (PTC), 5 applications in water industry, 215;
project, 30 applications of solar and wind
proton exchange membrane fuel cell energy, 218; centralized water
(PEMFC), 124–125, 124f infrastructure elements, 213;
public utility commission (PUC), 25 energy demand for water
Public Utility Regulatory Policies Act infrastructure, 214–215; estimated
(PURPA), 2 savings from renewable energy
public water system, 188 in wastewater treatment, 218t;
pump as turbine (PAT), 80 ReNEWW house, 220; solar
energy, 215–217; solar energy in
rated capacity, 30 desalination plants, 218t; water and
rate of return, 32 energy conservation, 215; water
238 Index

renewable energy in water right-of-use and easement (RUE), 33


infrastructure (cont.) right-of-way (ROW), 31, 33
and energy nexus, 214–215; wind
energy, 217–218; wind energy in Safe Drinking Water Act (SDWA), 199
desalination plants, 219t sediment microbial fuel cell
renewable energy projects, 5, 28–31. (SMFC), 209
See also US renewable energy Self-Powered Intelligent Networks
generation technology mix; (SPIN), 179
US renewable energy policy Signaling System No. 7 (SS7), 179
renewable energy resources, 6; sludges/biosolids, 101–102. See also
biofuels, 7–8; Electric biogas-to-energy renewable systems
Consumers Protection Act of solar: cells, 184; energy, 215–217;
1986, 6; hydropower, 6–7; solar -powered desalination systems, 191;
photovoltaic, 8; wind energy, 8–9. -powered humidification–
See also US renewable energy policy dehumidification, 191; -powered
renewable energy standards (RES), 20 wireless sensor networks, 211;
renewable energy technologies, 80, 205 radiation, 183–184;
renewable fuel organizations, 42 stills, 191; thermal, 185; water
Renewable Fuel Standards (RFS), 3, 39 disinfection, 191–193; and wind
renewable natural gas (RNG), 108. energy rule, 31–33. See also
See also biogas-to-energy renewable renewable energy in water
systems infrastructure; US renewable energy
Renewable Portfolio Standards generation technology mix
(RPS), 4, 20; eligibility, 27–28; Solar Energy and Energy
and goals of United States, Conservation, 2. See also US
21t–24t; required, annual renewable renewable energy policy
energy percentages, 27t; State of solar energy and water/wastewater
Connecticut, 25–27; sustainable infrastructure, 183, 195; Clean
biomass fuel, 26. See also US Water Act, 193; concentrating solar
renewable energy generation power system, 186f; energy usage
technology mix in conventional activated sludge
renewable production tax credit, 5–6. system, 194f; energy usage in
See also US renewable energy policy public surface water system, 188f;
renewables, 34–35. See also US solar in desalination, 189–190,
renewable energy generation 190–191; solar in wastewater
technology mix processing, 194–195; solar in water
ReNEWW house, 220. See also industry, 187–189; solar PV and
renewable energy in water CSP, 187t; solar radiation, 183–184;
infrastructure solar thermal, 185; solar water
request for qualifications (RFQs), 110 disinfection, 191–193; wastewater
RES Act of 2019, 4. See also US processing, 193–194. See also solar
renewable energy policy photovoltaics
reverse osmosis (RO), 135, 137–138, solar photovoltaics, 184, 185f,
175, 190. See also desalination 186f; and CSP, 186–187, 187t;
technologies modules, 184
Index 239

solar-powered monitoring triple bottom line (TBL), 110


networks, 201; buoy system, 206f; turbine failure projectile zone
Class A rating, 204; discharge and (TFPZ), 165, 167, 174
nonpoint source monitoring, turbulence intensity (TI), 174
202–204; solar wireless sensor turbulent kinetic energy (TKE),
array, 202f; stormwater sampling 174
and monitoring, 201–202, 203f; Turgo impulse wheel, 69. See also
Sunapee Lake Project, 204; micro-hydropower
Susquehanna River Basin
Project, 203; water quality uninterruptable power supplies
monitoring with data buoys, (UPS), 179
205–206; wireless solar network United Nations Development
station, 204f. See also water quality Programme (UNDP), 78
monitoring United Nations Intergovernmental
solar radiation, solar disinfection Panel on Climate Change
(SODIS), 183, 187 (UN-IPCC), 10
solid oxide fuel cells (SOFCs), 115, US Department of Economic
129–130, 129f. See also fuel cells Development (US-DED), 109
solids retention time (SRT), 94 US ethanol production, 44–46.
Star Sailor SPIN•, 181 See also ethanol
state implementation plans (SIPs), 102 US renewable energy generation
Sunapee Lake, 204. See also solar- technology mix, 17, 35; energy
powered monitoring networks standards and goals, 21t–24t;
surface discharge, 145–146. See also local government on renewable
desalination technologies energy projects, 28–31; renewable
surface water (SW), 188 energy, 19–20; renewable energy
sustainable biomass fuel, 26 standards, 20–25; renewable
Synthetic Fuels Corporation Act, 2. portfolio standards, 25–28;
See also US renewable energy policy renewables, 34–35; RE
programs, 33–34; Right-of-
Ten State Standards, 95. See also Way, 31; solar and wind energy
biogas-to-energy renewable systems rule, 31–33; VA developed model
TETHYS, 34 ordinances, 29
thermal desalination energy US renewable energy policy, 1,
efficiency, 137. See also desalination 11–12; Agricultural Act of 2014, 3;
technologies Biomass Energy and Alcohol
thermal energy storage (TES), 141 Fuels Act, 2; Bipartisan Budget
thermal inactivation, 192 Act of 2018, 6; Clean Power
thermal vapor compression (TVC), 135 Plan of 2015, 3–4; Consolidated
total dissolved solids (TDS), 176 Appropriations Act of 2016, 5;
total nitrogen (TN), 49, 202 Emergency Economic Stabilization
total phosphorus (TP), 48, 202 Act of 2008, 6; energy efficiency
Toxics Release Inventory (TRI), 8 resources standards, 4; Energy
transesterification reaction, 53, 55f. Independence and Security Act of
See also biodiesel 2007, 3; Energy Policy Act of
240 Index

US renewable energy policy (cont.) water and energy conservation, 215


1992, 2; Energy Policy Act of water–energy nexus, 113, 114f,
2005, 2–3; federal investment tax 214–215. See also fuel cells;
credit, 5; Food, Conservation, and hydrogen
Energy Act of 2008, 3; future of, 9– renewable energy in water
11; Geothermal Energy Act, 2; House infrastructure
bill, 5; modified accelerated cost water infrastructure, 181;
recovery system, 6; National Energy centralized, 213; energy demand
Act of 1978 (HR 8444 1977–78), 2; for, 214–215; resilience in, 163.
Renewable Energy Initiatives, 2; See also renewable energy in water
renewable energy projects, 5; infrastructure
renewable energy resources, 6–9; water quality monitoring
renewable production tax credit, 5– (WQM), 199, 211; microbial fuel
6; RES Act of 2019, 4; Solar Energy cells for, 206–211; next-generation
and Energy Conservation, 2; compliance, 201; purposes, 199; in
Synthetic Fuels Corporation Act, 2; situ renewable energy sources, 200;
Total Energy Subsidies, 11t solar-powered monitoring
utility-interactive electricity networks, 201–206; traditional
generation plant, 72 approach, 200
water resource recovery facility
VA developed model ordinances, 29 (WRRF), 113, 115; 10MGD
valve regulated lead acid batteries WRRF, 115f
(VRLA), 74 water treatment plants (WTPs), 176
vapor compression, 137. See also wet milling process, 42f. See also
desalination technologies ethanol
variable renewable energy (VRE), 25 wide area network (WAN), 201
vertical axis wind turbines wind compression, 174
(VAWTs), 165; Darrieus wind wind energy, 8–9, 163, 181,
turbine, 165f; and HAWT 217–218; Aermotor pumping
technology, 165–167, 173t; high- installation, 169f; Aermotor
lift, 166f. See also wind energy windmill, 168f; assessing wind
virtual security (VSEC), 179 power, 171–172; built environment
volatile acids (VAs), 94 wind turbine, 173–174; desalination
volatile fatty acids (VFAs), 117 and energy consumption, 176t; in
volatile solids reduction (VSred), 103 desalination plants, 219t; efficiency
volt direct-current (VDC), 71 and Betz limit, 170–171; emerging
technologies, 179; energy and
waste activated sludge (WAS), 103 water, 175–178; HAWT and VAWT
wastewater processing, 193–194. technology, 165–167, 165f, 173t;
See also solar energy and water/ high-lift VAWT, 166f; project, 30;
wastewater infrastructure quantifying turbulence, 174–175;
wastewater processing or the resource turbine technology, applications,
recovery facilities (WRRFs), 94 and siting, 172; VAWT Darrieus
wastewater treatment techniques, wind turbine, 165f; wind energy
advanced, 194 in water infrastructure, 178–179;
Index 241

wind powered mechanical wind turbine, 30; technologies,


systems, 167–170; wind turbine 165–167; types, 165
technologies, 165–167; wind turbine wireless solar network (WSN), 201
types, 165. See also renewable World Health Organization
energy in water infrastructure; (WHO), 183
renewable energy resources

You might also like