Você está na página 1de 856

Renewable Energy Technology Guide

2012

2012 TECHNICAL REPORT

Renewable Energy Technology


Guide
2012
1023993
Final Report, December 2012

EPRI Project Manager


C. Lyons

ELECTRIC POWER RESEARCH INSTITUTE


3420 Hillview Avenue, Palo Alto, California 94304-1338 PO Box 10412, Palo Alto, California 94303-0813 USA
800.313.3774 650.855.2121 askepri@epri.com www.epri.com

DISCLAIMER OF WARRANTIES AND LIMITATION OF LIABILITIES


THIS DOCUMENT WAS PREPARED BY THE ORGANIZATION NAMED BELOW AS AN ACCOUNT
OF WORK SPONSORED OR COSPONSORED BY THE ELECTRIC POWER RESEARCH
INSTITUTE, INC. (EPRI). NEITHER EPRI, ANY MEMBER OF EPRI, ANY COSPONSOR, THE
ORGANIZATION(S) BELOW, NOR ANY PERSON ACTING ON BEHALF OF ANY OF THEM:
(A) MAKES ANY WARRANTY OR REPRESENTATION WHATSOEVER, EXPRESS OR IMPLIED, (I)
WITH RESPECT TO THE USE OF ANY INFORMATION, APPARATUS, METHOD, PROCESS, OR
SIMILAR ITEM DISCLOSED IN THIS DOCUMENT, INCLUDING MERCHANTABILITY AND FITNESS
FOR A PARTICULAR PURPOSE, OR (II) THAT SUCH USE DOES NOT INFRINGE ON OR
INTERFERE WITH PRIVATELY OWNED RIGHTS, INCLUDING ANY PARTY'S INTELLECTUAL
PROPERTY, OR (III) THAT THIS DOCUMENT IS SUITABLE TO ANY PARTICULAR USER'S
CIRCUMSTANCE; OR
(B) ASSUMES RESPONSIBILITY FOR ANY DAMAGES OR OTHER LIABILITY WHATSOEVER
(INCLUDING ANY CONSEQUENTIAL DAMAGES, EVEN IF EPRI OR ANY EPRI REPRESENTATIVE
HAS BEEN ADVISED OF THE POSSIBILITY OF SUCH DAMAGES) RESULTING FROM YOUR
SELECTION OR USE OF THIS DOCUMENT OR ANY INFORMATION, APPARATUS, METHOD,
PROCESS, OR SIMILAR ITEM DISCLOSED IN THIS DOCUMENT.
REFERENCE HEREIN TO ANY SPECIFIC COMMERCIAL PRODUCT, PROCESS, OR SERVICE BY
ITS TRADE NAME, TRADEMARK, MANUFACTURER, OR OTHERWISE, DOES NOT NECESSARILY
CONSTITUTE OR IMPLY ITS ENDORSEMENT, RECOMMENDATION, OR FAVORING BY EPRI.
THE FOLLOWING ORGANIZATION PREPARED THIS REPORT:
Electric Power Research Institute (EPRI)

NOTE
For further information about EPRI, call the EPRI Customer Assistance Center at 800.313.3774 or
e-mail askepri@epri.com.
Electric Power Research Institute, EPRI, and TOGETHERSHAPING THE FUTURE OF ELECTRICITY
are registered service marks of the Electric Power Research Institute, Inc.
Copyright 2013 Electric Power Research Institute, Inc. All rights reserved.

ACKNOWLEDGMENTS
The following organization prepared this report:
Electric Power Research Institute (EPRI)
1300 West W. T. Harris Blvd.
Charlotte, NC 28262
Principal Investigator
C. Lyons
This report describes research sponsored by EPRI.
The following EPRI personnel are acknowledged for their contributions to this report:
L. Cerezo
T. Coleman
P. Jacobson
C. Libby
D. OConnor
A. Tuohy

This publication is a corporate document that should be cited in the literature in the following
manner:
Renewable Energy Technology Guide: 2012. EPRI, Palo Alto, CA: 2011. 1023993.
iii

PRODUCT DESCRIPTION

First published in 2000 as the Renewable Energy Technical Assessment GuideTAG-RE, the
Electric Power Research Institutes (EPRIs) annual Renewable Energy Technology Guide
provides a consistent basis for evaluating the economic feasibility of renewable generation
technologies. These technologies include wind, solar photovoltaic (PV), solar thermal, biomass,
municipal solid waste, geothermal, and emerging ocean energy conversion technologies.
Results and Findings
Based on rated capacity, the leading worldwide sources of renewable energy at the end of 2011
were wind (238 GW), biomass (72 GW), solar PV (70 GW), geothermal (11.2 GW), and solar
thermal (1.8 GW). Renewables generated an estimated 20.3% of global electricity by the end of
2011. In 2010, renewable energy supplied approximately 16.7% of global energy consumption.
Challenges and Objectives
Because so many conflicting and overly optimistic claims are made about the performance and
economic potential of all energy sourcesconventional fossil, nuclear, hydro, alternative fuel,
and renewablesit is important to conduct careful, objective assessments of the status,
performance, cost, environmental impacts, and other aspects of the technologies. The primary
objective of this report is to provide timely and unbiased assessments of the status, performance,
and cost of the renewable generation technologies.
Applications, Value, and Use
The renewable generation technology data reported in the EPRI Renewable Energy Technology
Guide are valuable to system planners and others who make decisions about whether renewable
technology belongs in the generation portfolio and who develop long-term strategies for a
sustainable generation portfolio.
An energy producer can use renewable energy to strengthen ties to the community, attract new
customers, and diversify its portfolio. EPRI believes that companies that have not yet
incorporated renewables into their generation mix should consider these technologies to
determine which ones might benefit their generation strategy. The information summarized in
this report encompasses more than three decades of EPRI research on renewable technologies.
The Renewable Energy Technology Guide will continue to be updated on an annual basis to
ensure that EPRI members have access to the most current information on the technical and
economic status of renewable technologies.

Approach
Information and data from EPRI reports and a variety of outside sources, including U.S.
government reports and other publicly available publications, were collected and analyzed. The
informationwhich concerned the status, performance, cost, installed capacity, and markets for
renewable energy generation technologieswas used to update Renewable Energy Technology
Guide:2011 (EPRI report 1021795).
Keywords
Biomass
Geothermal
Municipal solid waste
Renewable energy
Solar energy
Wind energy

vi

ABSTRACT
The Renewable Energy Technology Guide: 2012 is a fundamental industry reference that informs
the technical and economic assessment of renewable energy generation technologies. Design,
cost, and performance information contained in this report will enable Electric Power Research
Institute members to perform preliminary capital investment evaluations in a systematic and
informed manner.
Renewable power technologies addressed in this report that are commercially available or on the
threshold of commercialization include wind, biomass, municipal solid waste, solar
photovoltaics, solar thermal, and geothermal. Other sections cover emerging renewable power
technologies, such as ocean tidal, ocean wave, and river in-stream energy conversion. In
addition, the report discusses the challenges of integrating renewable energy technologies with
the grid and the potential to reduce greenhouse gas emissions through renewables.

vii

CONTENTS

1 INTRODUCTION .................................................................................................................... 1-1


1.1 Background ..................................................................................................................... 1-1
1.2 EPRIs Role ..................................................................................................................... 1-4
1.3 Objective ......................................................................................................................... 1-4
1.4 Definitions and Units ....................................................................................................... 1-5
1.5 Scope .............................................................................................................................. 1-5
1.6 References ...................................................................................................................... 1-6
2 ECONOMIC METHODOLOGY AND ASSUMPTIONS........................................................... 2-1
2.1 Regulated Utility Power Projects ..................................................................................... 2-1
2.1.1 Fixed Charge Rates................................................................................................. 2-1
2.1.1.1 Annual Fixed Charge Rates ............................................................................. 2-1
2.1.1.2 Nominal Levelized Annual Charges and Nominal Levelized Fixed
Charge Rates ............................................................................................................... 2-3
2.1.1.3 Real Levelized Annual Charges and Real Levelized Fixed Charge Rates ...... 2-3
2.1.2 Example of Annual Fixed Charge Rates for a 200-MW Wind Power Plant ............. 2-5
2.1.3 Other Capital ExpensesCapital Additions ............................................................ 2-5
2.1.4 Calculating Costs per Kilowatt-Hour ........................................................................ 2-5
2.1.4.1 Levelized Costs per Kilowatt-Hour ................................................................... 2-5
2.2 Non-Utility Generator Power Projects ............................................................................. 2-8
2.2.1 Types of Non-Utility Generators .............................................................................. 2-9
2.2.1.1 PURPA Cogenerators and Small Power Producers ...................................... 2-10
2.2.1.2 Exempt Wholesale Generators ...................................................................... 2-11
2.2.1.3 Merchant Power Plants .................................................................................. 2-11
2.2.2 Development of an Economic Pro Forma for a Merchant Plant ............................ 2-12
2.2.2.1 Differences Between Regulated Utility and Non-Utility Power Projects ......... 2-12
2.2.3 Conceptualizing the Analysis ................................................................................. 2-13
2.2.3.1 Total Capital Requirement ............................................................................. 2-14

ix

2.2.3.2 EPRI Capital Cost Definitions ........................................................................ 2-17


2.2.3.3 Income Statement .......................................................................................... 2-17
2.2.3.4 Income Taxes ................................................................................................ 2-21
2.2.3.5 Cash Flow Statement .................................................................................... 2-22
2.2.3.6 Economic Measures ...................................................................................... 2-24
2.2.3.7 Sensitivity Analysis ........................................................................................ 2-27
2.2.3.8 Project Risks and Financing .......................................................................... 2-29
2.2.4 Limitations of Examples ......................................................................................... 2-31
2.3 Guidelines for Economic Evaluation of Renewable Energy Projects ............................ 2-31
2.3.1 Design/Cost Estimate ............................................................................................ 2-31
2.3.2 Accuracy Ranges .................................................................................................. 2-32
2.3.3 Definitions of Economic Terms .............................................................................. 2-32
2.4 References .................................................................................................................... 2-34
3 WIND POWER ........................................................................................................................ 3-1
3.1 Introduction ..................................................................................................................... 3-1
3.2 Installed Wind Capacity ................................................................................................... 3-2
3.3 Wind Energy Principles ................................................................................................... 3-9
3.4 Developments in Wind Turbine Technology.................................................................. 3-11
3.4.1 Generators and Power Electronics ........................................................................ 3-11
3.4.1.1 Squirrel Cage Induction Generator ................................................................ 3-12
3.4.1.2 Variable-Speed Turbine ................................................................................. 3-12
3.4.1.3 Power Electronics .......................................................................................... 3-13
3.4.1.4 Squirrel Cage Induction Generator with Full Power Conversion.................... 3-13
3.4.1.5 Doubly Fed Induction Generator .................................................................... 3-13
3.4.1.6 Synchronous Generator ................................................................................. 3-13
3.4.1.7 Direct-Drive Low-Speed Wound Rotor Generators........................................ 3-14
3.4.1.8 Direct-Drive Low-Speed Permanent Magnet Generators .............................. 3-14
3.4.1.9 Superconducting Low-Speed Generators ...................................................... 3-15
3.4.2 Blades and Rotor ................................................................................................... 3-16
3.4.2.1 Passive Aerodynamic Control ........................................................................ 3-19
3.4.2.2 Active Aerodynamic Control........................................................................... 3-19
3.4.2.3 Stealth Rotor Blade ..................................................................................... 3-20
3.4.2.4 Smart Blades ............................................................................................... 3-20
3.4.2.5 Flatback Airfoils.............................................................................................. 3-20

3.4.2.6 Blade Manufacturing Processes .................................................................... 3-21


3.4.2.7 Blade Transport and Shipping ....................................................................... 3-21
3.4.2.8 Blade Test Facilities ....................................................................................... 3-21
3.4.3 Yaw Systems ......................................................................................................... 3-21
3.4.4 Sensors ................................................................................................................. 3-22
3.4.4.1 Optical Strain Gages ...................................................................................... 3-22
3.4.4.2 Wind MeasurementsLIDAR ........................................................................ 3-23
3.4.4.3 Condition Monitoring ...................................................................................... 3-23
3.4.5 Controls ................................................................................................................. 3-24
3.4.6 SCADA Data Collection and Transmittal ............................................................... 3-25
3.4.7 Drive Train ............................................................................................................. 3-26
3.4.7.1 Drive Trains with Gearboxes.......................................................................... 3-26
3.4.7.2 Hydrodynamic Fluid Coupling ........................................................................ 3-27
3.4.7.3 Direct Drive Train without Gearbox ................................................................ 3-28
3.4.8 Foundation............................................................................................................. 3-29
3.4.9 Tower..................................................................................................................... 3-30
3.4.9.1 Conventional Steel Towers ............................................................................ 3-30
3.4.9.2 Concrete and Hybrid Towers ......................................................................... 3-30
3.4.9.3 Integrated Towers and Foundations .............................................................. 3-31
3.4.9.4 Tall Towers .................................................................................................... 3-31
3.4.9.5 Two-Bladed Rotors ........................................................................................ 3-32
3.4.9.6 Vertical Axis Rotors ....................................................................................... 3-34
3.4.10 Offshore Foundations .......................................................................................... 3-36
3.4.10.1 General Classifications ................................................................................ 3-36
3.4.10.2 Monopile Foundations.................................................................................. 3-40
3.4.10.3 Gravity-Based Foundations ......................................................................... 3-43
3.4.10.4 Tripod Foundations ...................................................................................... 3-44
3.4.10.5 Jacket Foundations ...................................................................................... 3-45
3.4.10.6 Suction Bucket Foundations ........................................................................ 3-46
3.4.10.7 Floating Foundations ................................................................................... 3-47
3.5 Trends in the Turbine Supply Market ............................................................................ 3-48
3.6 Trends in Wind Turbine and Plant Sizes ....................................................................... 3-53
3.7 Onshore Capital and O&M Cost Trends........................................................................ 3-54
3.8 Developments in Offshore Wind Technology ................................................................ 3-57

xi

3.8.1 Trends in the Offshore Turbine Supply Market ...................................................... 3-58


3.8.2 Offshore Transmission Technology Status ............................................................ 3-59
3.8.3 Offshore Capital and O&M Cost Trends ................................................................ 3-59
3.8.3.1 Capital Costs ................................................................................................. 3-59
3.8.3.2 Wind Turbine Foundations ............................................................................. 3-60
3.8.3.3 Wind Turbine Generators............................................................................... 3-60
3.8.3.4 Inner Array Grid ............................................................................................. 3-61
3.8.3.5 Offshore Substation (OSS) ............................................................................ 3-61
3.8.3.6 Export Cabling and Onshore Interconnection ................................................ 3-61
3.8.3.7 Installation Costs ........................................................................................... 3-62
3.8.3.8 Distribution of CAPEX by Category ............................................................... 3-63
3.8.3.9 O&M Costs..................................................................................................... 3-65
3.8.3.10 Preventive Maintenance .............................................................................. 3-67
3.8.3.11 Corrective Maintenance ............................................................................... 3-68
3.8.3.12 O&M Expenditures ....................................................................................... 3-70
3.8.4 Offshore Plant Performance .................................................................................. 3-73
3.8.5 Offshore Plant Cost Estimates .............................................................................. 3-74
3.8.5.1 Northeastern United States............................................................................ 3-74
3.8.5.2 Great Lakes ................................................................................................... 3-85
3.8.5.3 United Kingdom ............................................................................................. 3-96
3.9 On-Shore Technology Performance and Cost Tables ................................................ 3-109
3.9.1 Site Assumptions ................................................................................................. 3-109
3.9.2 Plant Performance ............................................................................................... 3-111
3.9.3 Total Capital Requirement ................................................................................... 3-112
3.9.4 Operation and Maintenance Cost ........................................................................ 3-113
3.9.5 Levelized Cost of Electricity ................................................................................. 3-117
3.10 Grid Integration.......................................................................................................... 3-133
3.10.1 Production Variability ......................................................................................... 3-133
3.10.2 Ancillary Service Costs ...................................................................................... 3-134
3.10.3 Wind Energy Forecasting and Scheduling ........................................................ 3-135
3.10.4 Power Quality .................................................................................................... 3-136
3.11 Project Development Process and Market ................................................................ 3-136
3.11.1 Buy or Build Considerations .............................................................................. 3-136
3.11.2 Project Development Process ........................................................................... 3-137

xii

3.11.2.1 Pre-Development Activities ........................................................................ 3-137


3.11.2.2 Engineering, Procurement, and Construction Activities ............................. 3-140
3.11.2.3 Operation and Maintenance Activities ....................................................... 3-141
3.12 Environmental Issues ................................................................................................ 3-142
3.12.1 Avian and Bat Issues ......................................................................................... 3-143
3.12.2 Noise ................................................................................................................. 3-144
3.12.3 Visual Impact ..................................................................................................... 3-145
3.12.4 Shadow Flicker .................................................................................................. 3-146
3.12.5 Radar Interference............................................................................................. 3-147
3.13 Scouting New Potential Game-Changer Technologies ............................................. 3-148
3.13.1 Multi-Copter Aerial Drones for Cost-Effective Wind Turbines, Blades, and
Tower Inspection .......................................................................................................... 3-148
3.13.2 Direct-Drive, Axial Gap, Air-Cooled Permanent Magnet Generator .................. 3-149
3.13.3 Oversize Large Turbine Component Heavy-Lift Transportation Using
Balloons ........................................................................................................................ 3-149
3.13.4 Airborne Wind Turbines ..................................................................................... 3-149
3.13.5 Ro-Birds of Prey to Scare Birds ......................................................................... 3-149
3.14 References ................................................................................................................ 3-149
3.14.1 DOE-EPRI Wind Turbine Verification Program Reports .................................... 3-152
3.14.2 Other Reports .................................................................................................... 3-153
4 BIOMASS ELECTRICITY GENERATION .............................................................................. 4-1
4.1 Introduction ..................................................................................................................... 4-1
4.1.1 Basic Issues Associated with Biomass Fuel Utilization ........................................... 4-4
4.1.2 Technology Considerations for Using Biomass Fuels ............................................. 4-5
4.1.2.2 Goal Technologies ........................................................................................... 4-6
4.1.2.3 Prospectus ....................................................................................................... 4-7
4.2 Biomass Fuel Resources ................................................................................................ 4-7
4.2.1 Biofuels vs. Biomass Electricity ............................................................................. 4-10
4.2.2 Basic Properties of Solid Biomass Fuels ............................................................... 4-10
4.2.2.1 Proximate and Ultimate Analyses of Biomass Fuels ..................................... 4-11
4.2.2.2 Structural Relationships and Heteroatoms .................................................... 4-13
4.2.2.3 Inorganic Constituents in Biomass................................................................. 4-14
4.2.3 Performance Characteristics of Biomass Fuels ..................................................... 4-18
4.2.3.1 Biomass Fuel Volatility ................................................................................... 4-18

xiii

4.2.3.2 Fuel Nitrogen Characteristics of Biomass Fuels ............................................ 4-19


4.2.3.3 Ash Reactivity, Slagging, and Fouling ........................................................... 4-27
4.3 Characteristics of Biomass Cofiring Technologies ........................................................ 4-30
4.3.1 Overview of Solid Biomass Fuel Cofiring Systems ................................................ 4-31
4.3.2 Cofiring in Cyclone Boilers .................................................................................... 4-31
4.3.2.1 Cofiring System Design for Cyclone Boilers .................................................. 4-33
4.3.2.2 Capital Costs of Cofiring Systems for Cyclone Boilers .................................. 4-37
4.3.2.3 Cofiring Impacts in Cyclone Boilers ............................................................... 4-39
4.3.3 Cofiring in Pulverized Coal (PC) Boilers ................................................................ 4-41
4.3.3.1 Cofiring System Designs for Pulverized Coal Boilers .................................... 4-41
4.3.3.2 Capital Costs of Separate Injection PC Cofiring Systems ............................. 4-43
4.3.3.3 Impacts of Cofiring Biomass in PC Boilers .................................................... 4-44
4.3.4 Cofiring Gaseous Biomass Fuels .......................................................................... 4-46
4.3.4.1 Biomass Pre-Treatments for Cofiring with Coal ............................................. 4-47
4.3.4.2 The Global Biomass Pellet Market................................................................. 4-49
4.3.5 Conclusions Regarding Biomass Cofiring ............................................................. 4-50
4.4 Stand-Alone Biomass-Fired Systems............................................................................ 4-50
4.4.1 Wood-Waste-Fired Boilers .................................................................................... 4-50
4.4.1.1 Typical Capacities of Wood-Waste-Fired Plants............................................ 4-51
4.4.1.2 Wood-Waste-Fired System Design................................................................ 4-51
4.4.1.3 Comparisons to Other Biomass-Fired Rankine-Cycle Designs ..................... 4-57
4.4.2 Repowering Existing Coal Units for Biomass Firing .............................................. 4-58
4.4.3 Ownership Models and Institutional Issues ........................................................... 4-59
4.4.4 Alternative Systems for Gaseous Biomass Fuels .................................................. 4-59
4.4.5 Near-Commercial Technologies ............................................................................ 4-60
4.4.5.1 Types of Gasifiers .......................................................................................... 4-60
4.4.5.2 Syngas Quality for Downstream Use ............................................................. 4-66
4.4.5.3 Wood Gasification for Power and District Heating in Gssing, Austria .......... 4-67
4.4.5.4 Wood Gasification at Harbore...................................................................... 4-68
4.4.5.5 Gasification of Wood Pellets at Skive ............................................................ 4-69
4.4.5.6 Cofiring Syngas at Ruien ............................................................................... 4-70
4.4.5.7 Wood Gasification at Kokemki ..................................................................... 4-71
4.4.6 Non-Commercial Technologies ............................................................................. 4-72
4.5 Cost and Performance Summary .................................................................................. 4-73

xiv

4.5.1 100% Biomass Repowering of a Pulverized Coal Boiler ....................................... 4-73


4.5.1.1 Plant Design Assumptions ............................................................................. 4-74
4.5.1.2 Plant Performance Estimate .......................................................................... 4-75
4.5.1.3 Total Performance and Cost Estimates ......................................................... 4-75
4.5.1.4 Levelized Cost of Electricity Estimates .......................................................... 4-78
4.5.2 Biomass Cofiring with Coal in a Pulverized Coal Boiler ........................................ 4-79
4.5.2.1 Plant Design and Fuel Assumptions .............................................................. 4-79
4.5.2.2 Plant Performance Estimates ........................................................................ 4-80
4.5.2.3 Total Capital Requirement and Operation and Maintenance Cost
Estimates ................................................................................................................... 4-82
4.5.3 Biomass-Fired Bubbling Fluidized Bed Combustion Boiler Power Plants ............. 4-84
4.5.3.1 Plant Design Assumptions ............................................................................. 4-84
4.5.3.2 Plant Performance Estimate .......................................................................... 4-85
4.5.3.3 Total Capital Requirement and Operation and Maintenance Cost
Estimates ................................................................................................................... 4-85
4.5.3.4 Levelized Cost of Electricity Estimates .......................................................... 4-87
4.5.4 Biomass-Fired Stoker Boiler Power Plants ............................................................ 4-89
4.5.4.1 Plant Design Assumptions ............................................................................. 4-89
4.5.4.2 Plant Performance Estimate .......................................................................... 4-89
4.5.4.3 Total Capital Requirement and Operation and Maintenance Cost
Estimates ................................................................................................................... 4-90
4.5.4.4 Levelized Cost of Electricity Estimate ............................................................ 4-91
4.5.5 Biomass Gasification Power Plants ....................................................................... 4-92
4.5.5.1 Plant Design Assumptions ............................................................................. 4-92
4.5.5.2 Plant Performance Estimate .......................................................................... 4-93
4.5.5.3 Total Capital Requirement and Operation and Maintenance Cost
Estimates ................................................................................................................... 4-94
4.5.3.4 Levelized Cost of Electricity Estimate ............................................................ 4-96
4.6 Biomass to Electricity RD&D Initiatives for the Future .................................................. 4-99
4.7 U.S. Technical Tax Code Modifications to Promote Further Use of Biomass ............. 4-100
4.8 Conclusions ................................................................................................................. 4-100
4.9 References .................................................................................................................. 4-101
5 MUNICIPAL SOLID WASTE .................................................................................................. 5-1
5.1 Introduction ..................................................................................................................... 5-1
5.2 Overview of the Regulatory and Legislative Environment ............................................... 5-3

xv

5.2.1 Permit Overview ...................................................................................................... 5-4


5.2.2 Air Emission Permits ............................................................................................... 5-4
5.2.3 Solid Waste Permits ................................................................................................ 5-7
5.3 Legislative Overview ....................................................................................................... 5-7
5.3.1 Renewable Fuel Standards (RFS) ........................................................................... 5-8
5.3.2 Renewable Portfolio Standard ............................................................................... 5-10
5.4 Technology Considerations for Using MSW.................................................................. 5-12
5.5 Commercially Available Technologies........................................................................... 5-14
5.5.1 Mass Burn Technologies ....................................................................................... 5-14
5.5.1.1 Typical Mass Burn Technical Parameters ..................................................... 5-16
5.5.1.2 Mass Burn Experience and Vendors in the United States ............................. 5-16
5.5.1.3 Modular Starved Air Combustion Systems .................................................... 5-18
5.5.2 RDF Co-Firing Technologies ................................................................................. 5-20
5.5.2.1 RDF Dedicated Boiler .................................................................................... 5-21
5.5.2.2 RDF Fluidized Bed ......................................................................................... 5-24
5.5.2.3 RDF Experience and Vendors in the United States ....................................... 5-27
5.5.3 LFG Utilization Technologies ................................................................................. 5-27
5.3.3.1 LFG to Electricity............................................................................................ 5-29
5.5.3.2 LFG as Direct Heat Source ............................................................................ 5-31
5.5.3.3 LFG to Pipeline .............................................................................................. 5-33
5.5.4 Syngas Co-Firing and Hybrid Cycles ..................................................................... 5-34
5.5.5 Advanced Thermal Conversion Technologies ....................................................... 5-37
5.5.5.1 Pyrolysis ........................................................................................................ 5-37
5.5.5.2 Gasification .................................................................................................... 5-39
5.5.5.3 Plasma Arc .................................................................................................... 5-42
5.5.6 Aerobic Composting .............................................................................................. 5-43
5.5.7 Chemical Decomposition ....................................................................................... 5-47
5.5.8 Technology Summary............................................................................................ 5-49
5.6 Evaluation of a 25-MW Mass Burn WTE Facility........................................................... 5-49
5.6.1 The MSW Stream .................................................................................................. 5-50
5.6.2 Advantages and Disadvantages of Waste-to-Energy Systems for MSW .............. 5-51
5.6.3 MSW Receipt and Storage .................................................................................... 5-54
5.6.4 Waste to Energy Technologies for Combustion .................................................... 5-55
5.6.5 Mass Burn of MSW................................................................................................ 5-56

xvi

5.6.6 Air Emissions Control ............................................................................................ 5-57


5.6.7 Energy Generation ................................................................................................ 5-58
5.6.8 Facility Sizing......................................................................................................... 5-58
5.6.9 System Maintenance and Reliability ...................................................................... 5-58
5.6.10 Mass Burn Process ............................................................................................. 5-59
5.6.11 Capital and O&M Related Scope and Equipment ............................................... 5-61
5.6.11.1 MSW Fuel Costs .......................................................................................... 5-63
5.6.11.2 Cost Estimate Summary .............................................................................. 5-63
5.6.11.3 Levelized Cost of Electricity ......................................................................... 5-65
5.6.11.4 Levelized Fixed Charge Rate....................................................................... 5-65
5.7 Evaluation of Cofiring of Coal with RDF ........................................................................ 5-66
5.7.1 RDF Processing and Features .............................................................................. 5-66
5.7.2 RDF Production Technologies............................................................................... 5-67
5.7.3 Heating Value of RDF............................................................................................ 5-68
5.7.4 SlaggingA Key Combustion Parameter ............................................................. 5-69
5.7.5 RDF Co-Fire Energy Generation and Feedstock .................................................. 5-70
5.7.6 RDF Co-Firing Advantages and Disadvantages .................................................... 5-71
5.7.7 RDF Co-Firing Strategy ......................................................................................... 5-72
5.7.8 Preliminary Design Parameters ............................................................................. 5-73
5.7.8.1 Material Handling ........................................................................................... 5-75
5.7.9 Performance Parameters ...................................................................................... 5-75
5.7.9.1 Impact of RDF Firing on Boiler Performance ................................................. 5-77
5.7.10 Coal and RDF Quality Summary ......................................................................... 5-78
5.7.11 Capital and O&M Related Scope and Equipment ............................................... 5-80
5.7.11.1 RDF Fuel Costs ........................................................................................... 5-84
5.7.11.2 Cost Estimate Summary .............................................................................. 5-84
5.7.11.3 Levelized Cost of Electricity ......................................................................... 5-85
5.7.11.4 Levelized Fixed Charge Rate....................................................................... 5-86
5.8 Landfill Gas Evaluation ................................................................................................. 5-87
5.8.1 LFG Characteristics ............................................................................................... 5-87
5.8.2 LFG Treatment ...................................................................................................... 5-88
5.8.3 LFG Basic Design.................................................................................................. 5-88
5.8.3.1 System Maintenance and Reliability .............................................................. 5-89
5.8.4 LFG Issues ............................................................................................................ 5-89

xvii

5.8.5 LFG Benefits.......................................................................................................... 5-89


5.8.6 LFG-to-Electricity ................................................................................................... 5-90
5.8.7 Landfill Gas to Direct Heat..................................................................................... 5-91
5.8.8 Landfill Gas to Pipeline .......................................................................................... 5-92
5.8.9 Capital and O&M Related Scope and Equipment ................................................. 5-94
5.8.9.1 LFG Fuel Costs .............................................................................................. 5-98
5.8.9.2 Cost Estimate Summary ................................................................................ 5-98
5.8.9.3 Levelized Cost of Electricity ......................................................................... 5-104
5.8.9.4 Levelized Fixed Charge Rate....................................................................... 5-106
5.8.10 Conclusions ....................................................................................................... 5-107
6 SOLAR PHOTOVOLTAICS ................................................................................................... 6-1
6.1 Introduction ..................................................................................................................... 6-1
6.2 Cost and Economic Issues .............................................................................................. 6-6
6.2.1 Japan ..................................................................................................................... 6-10
6.2.2 Germany ................................................................................................................ 6-11
6.2.3 Technology Performance and Cost Tables ........................................................... 6-12
6.2.3.1 Distributed PV Performance and Cost Estimates .......................................... 6-13
6.2.3.2 Utility-Scale PV Plant Performance and Cost Estimates ............................... 6-13
6.2.4 Performance and Cost History and Projections ..................................................... 6-24
6.2.5 PV Performance and Cost Estimate Model ........................................................... 6-25
6.2.5.1 Resource Estimates ....................................................................................... 6-25
6.2.5.2 Distributed PV Cost and Performance Estimates .......................................... 6-26
6.2.5.3 Hypothetical Case Studies ............................................................................. 6-28
6.2.5.4 Central Station PV Cost and Performance Methodology ............................... 6-32
6.3 Environmental Issues .................................................................................................... 6-37
6.4 Design and Deployment Issues..................................................................................... 6-39
6.4.1 Utility-Scale Issues ................................................................................................ 6-40
6.4.2 Operating and Maintenance Labor Requirements ................................................. 6-40
6.4.3 Grid Connection..................................................................................................... 6-42
6.5 Equipment Markets and Key Participants ..................................................................... 6-44
6.5.1 Equipment Markets................................................................................................ 6-44
6.5.2 Key Participants..................................................................................................... 6-46
6.5.3 Resources ............................................................................................................. 6-47
6.5.3.1 Internet Sites .................................................................................................. 6-47

xviii

6.5.3.2 PV Manufacturers .......................................................................................... 6-49


6.5.4 Applicable Codes and Standards .......................................................................... 6-49
6.6 References .................................................................................................................... 6-50
7 GEOTHERMAL ENERGY ...................................................................................................... 7-1
7.1 Introduction ..................................................................................................................... 7-1
7.2 Resources and Current Installed Capacity ...................................................................... 7-3
7.2.1 Hydrothermal Geothermal Resource Assessment .................................................. 7-3
7.2.1.1 Resource Temperature .................................................................................... 7-4
7.2.1.2 Depth ............................................................................................................... 7-7
7.2.1.3 Resource Permeability ..................................................................................... 7-7
7.2.2 Geo-Pressured Geothermal Resource Assessment ............................................... 7-8
7.2.3 Hot Dry Rock Resource Assessment ...................................................................... 7-8
7.2.4 Installed Geothermal Capacity .............................................................................. 7-11
7.3 Technology Description ................................................................................................. 7-13
7.3.1 Hydrothermal Power .............................................................................................. 7-14
7.3.1.1 Major Common System Components and Features ...................................... 7-15
7.3.1.2 Direct Steam .................................................................................................. 7-16
7.3.1.3 Flash Steam ................................................................................................... 7-17
7.3.1.4 Binary Power Plants....................................................................................... 7-19
7.3.1.5 Low Enthalpy/Reverse Air Conditioning Cycles ............................................. 7-21
7.3.2 Geo-Pressured ...................................................................................................... 7-21
7.3.3 Hot Dry Rock (Enhanced Geothermal Systems) ................................................... 7-21
7.3.3.1 Supercritical Cycles ....................................................................................... 7-23
7.3.4 Hot Sedimentary Aquifer (HSA) ............................................................................. 7-23
7.3.5 Down-Hole Closed-Loop Systems ......................................................................... 7-24
7.3.6 Geothermal Hybrid Plants ..................................................................................... 7-24
7.3.6.1 Geothermal-Solar Hybrid Cycles ................................................................... 7-25
7.3.6.2 Geothermal + Gas Peaker Cycles ................................................................. 7-26
7.3.6.3 Geothermal-Fossil Hybrid Cycles .................................................................. 7-27
7.3.6.4 Geothermal-Biomass Hybrid Cycles for Moderate-Low Enthalpy
Sources ...................................................................................................................... 7-27
7.3.7 Non-Electric (Direct Use) Geothermal ................................................................. 7-27
7.4 Technology Status......................................................................................................... 7-28
7.4.1 Recent U.S. Developments ................................................................................... 7-29

xix

7.4.1.1 Alaska ............................................................................................................ 7-29


7.4.1.2 Arizona ........................................................................................................... 7-30
7.4.1.3 California ........................................................................................................ 7-30
7.4.1.4 Colorado ........................................................................................................ 7-30
7.4.1.5 Hawaii ............................................................................................................ 7-30
7.4.1.6 Idaho .............................................................................................................. 7-31
7.4.1.7 Louisiana ....................................................................................................... 7-31
7.4.1.8 Mississippi...................................................................................................... 7-31
7.4.1.9 Nevada .......................................................................................................... 7-31
7.4.1.10 New Mexico ................................................................................................. 7-31
7.4.1.11 North Dakota ................................................................................................ 7-31
7.4.1.12 Oregon ......................................................................................................... 7-31
7.4.1.13 Texas ........................................................................................................... 7-32
7.4.1.14 Utah ............................................................................................................. 7-32
7.4.1.15 Washington .................................................................................................. 7-32
7.4.1.16 Wyoming ...................................................................................................... 7-32
7.4.2 Hot Dry Rock Technology Status .......................................................................... 7-32
7.4.3 Geothermal Technologies for Electricity Production Deployment Curve
(Grubb Curve)................................................................................................................. 7-34
7.4.4 Geophysical Methods in Geothermal Exploration ................................................. 7-34
7.4.5 Geothermal Drilling Technology ............................................................................ 7-36
7.5 Cost and Economic Issues ............................................................................................ 7-37
7.5.1 Example Geothermal Project Costs ...................................................................... 7-39
7.5.2 Factors that Influence Geothermal Power Plant Costs .......................................... 7-40
7.5.3 Engineering and Economic Evaluation .................................................................. 7-40
7.6 Environmental Issues .................................................................................................... 7-43
7.6.1 Land Use ............................................................................................................... 7-43
7.6.2 Subsidence ............................................................................................................ 7-44
7.6.3 Emissions .............................................................................................................. 7-45
7.6.4 Thermal Discharge ................................................................................................ 7-47
7.6.5 Induced Seismicity................................................................................................. 7-47
7.7 Design, Deployment, and O&M Issues ......................................................................... 7-48
7.7.1 Industry Trends...................................................................................................... 7-48
7.7.2 Research Needs and Recommendations .............................................................. 7-49
7.8 Key Research Participants ............................................................................................ 7-50

xx

7.9 References .................................................................................................................... 7-52


8 SOLAR THERMAL ................................................................................................................. 8-1
8.1 Introduction ..................................................................................................................... 8-1
8.1.1 Solar Thermal Technologies.................................................................................... 8-4
8.1.2 Hybridization ............................................................................................................ 8-6
8.2 Resources ....................................................................................................................... 8-7
8.3 Technology Description ................................................................................................. 8-10
8.3.1 Parabolic Trough ................................................................................................... 8-10
8.3.2 Central Receiver .................................................................................................... 8-12
8.3.3 Dish/Engine ........................................................................................................... 8-17
8.3.3.1 Concentrators ................................................................................................ 8-18
8.3.3.2 Engines .......................................................................................................... 8-19
8.3.4 Linear Fresnel Reflector ........................................................................................ 8-20
8.3.5 Solar Chimney ....................................................................................................... 8-21
8.4 Technology Status......................................................................................................... 8-22
8.4.1 Parabolic Trough ................................................................................................... 8-26
8.4.2 Central Receiver .................................................................................................... 8-28
8.4.3 Dish/Engine ........................................................................................................... 8-32
8.4.4 Linear Fresnel Reflector ........................................................................................ 8-34
8.5 Cost and Economic Issues ............................................................................................ 8-34
8.5.1 Engineering and Economic Evaluation .................................................................. 8-35
8.5.2 Parabolic Trough ................................................................................................... 8-37
8.5.3 Central Receiver .................................................................................................... 8-39
8.5.4 Dish/Engine ........................................................................................................... 8-40
8.6 Environmental Issues .................................................................................................... 8-40
8.6.1 Land Use ............................................................................................................... 8-41
8.6.2 Water Use.............................................................................................................. 8-42
8.6.3 Molten Salt for Central Receiver ............................................................................ 8-42
8.6.4 Oil for Heat Transfer in Trough Applications ......................................................... 8-42
8.6.5 Potential for Greenhouse Gas Reduction .............................................................. 8-42
8.7 Design, Deployment, and O&M Issues ......................................................................... 8-42
8.7.1 Parabolic Trough ................................................................................................... 8-43
8.7.2 Central Receiver .................................................................................................... 8-45
8.7.3 Dish/Engine ........................................................................................................... 8-48

xxi

8.7.4 Linear Fresnel Reflector ........................................................................................ 8-48


8.8 Equipment Markets and Key Participants ..................................................................... 8-48
8.8.1 Internet Resources ................................................................................................ 8-52
8.9 References .................................................................................................................... 8-53
9 OCEAN TIDAL ENERGY ....................................................................................................... 9-1
9.1 Introduction ..................................................................................................................... 9-1
9.2 U.S. Tidal In-Stream Energy Highlights: Late 2011 to Mid-2012..................................... 9-5
9.2.1 U.S. Federal Highlights............................................................................................ 9-5
9.2.1.1 Federal Appropriations..................................................................................... 9-5
9.2.1.2 Funding Awards ............................................................................................... 9-6
9.2.1.3 Federal Regulations ......................................................................................... 9-8
9.2.2 U.S. State Highlights ............................................................................................... 9-8
9.2.2.1 Projects in Alaska ............................................................................................ 9-8
9.2.2.2 Projects in Washington .................................................................................... 9-9
9.2.2.3 Projects in California ...................................................................................... 9-10
9.2.2.4 Projects in New York...................................................................................... 9-11
9.2.2.5 Projects in Massachusetts ............................................................................. 9-12
9.2.2.6 Projects in Maine ........................................................................................... 9-12
9.2.2.7 Projects in New Hampshire............................................................................ 9-14
9.2.2.8 Projects in New Jersey .................................................................................. 9-15
9.2.3 U.S. Developer and Project Deployment Highlights .............................................. 9-16
9.2.3.1 Verdant Power ............................................................................................... 9-16
9.2.3.2 Vortex Hydro Energy ..................................................................................... 9-17
9.2.3.3 Ocean Renewable Power Company .............................................................. 9-17
9.2.3.4 Free Flow Power ............................................................................................ 9-18
9.2.3.5 UEK Corporation ............................................................................................ 9-18
9.3 Worldwide Tidal In-Stream Energy Highlights: Late 2011 to Mid-2012 ......................... 9-18
9.3.1 Canada .................................................................................................................. 9-18
9.3.1.1 Fundy Ocean Research Centre for Energy Developments............................ 9-18
9.3.1.2 The Canoe Pass Tidal Energy Corporation (CPTEC).................................... 9-19
9.3.1.3 Ocean Renewable Power Company .............................................................. 9-19
9.3.2 United Kingdom (UK)............................................................................................. 9-19
9.3.2.1 Marine Current Turbines (MCT) ..................................................................... 9-21
9.3.2.2 Pulse Tidal ..................................................................................................... 9-22

xxii

9.3.2.3 Swanturbines ................................................................................................. 9-23


9.3.2.4 Tidal Generation Limited ................................................................................ 9-24
9.3.2.5 Atlantis Resources Corporation ..................................................................... 9-25
9.3.2.6 Hammerfest Strm......................................................................................... 9-25
9.3.2.7 Bluewater ....................................................................................................... 9-26
9.3.2.8 Kawasaki ....................................................................................................... 9-26
9.3.2.9 Voith Hydro .................................................................................................... 9-26
9.3.2.10 Flumill .......................................................................................................... 9-26
9.3.2.11 Scotrenewables ........................................................................................... 9-26
9.3.3 Ireland.................................................................................................................... 9-27
9.3.3.1 OpenHydro .................................................................................................... 9-27
9.3.4 France ................................................................................................................... 9-28
9.3.5 Australia................................................................................................................. 9-29
9.3.6 New Zealand ......................................................................................................... 9-29
9.3.7 India ....................................................................................................................... 9-29
9.3.8 Africa ..................................................................................................................... 9-30
9.3.9 Japan ..................................................................................................................... 9-30
9.4 Tidal Power and Energy Resources .............................................................................. 9-30
9.4.1 Data Sources ......................................................................................................... 9-30
9.4.2 Tidal Power Flux and Average Annual Energy ...................................................... 9-31
9.4.3 Extractable Tidal Power and Energy ..................................................................... 9-33
9.4.4 Tidal Power Forecasting ........................................................................................ 9-34
9.5 TISEC Research and Development .............................................................................. 9-34
9.5.1 Harnessing Tidal Energy ....................................................................................... 9-34
9.5.2 TISEC System Developers .................................................................................... 9-36
9.5.3 Survival in Storms and Hostile Marine Environments ............................................ 9-39
9.5.4 Effect of Tidal Power Plants on the Environment .................................................. 9-39
9.5.5 Permits for Tidal Power Plants .............................................................................. 9-40
9.5.6 Overview of Regulatory Status for Tidal Power Plants .......................................... 9-42
9.5.6.1 FERC Pilot Project License............................................................................ 9-42
9.6 Design, Performance, Cost, and Economic Feasibility ................................................. 9-42
9.6.1 TISEC Sites ........................................................................................................... 9-42
9.6.2 TISEC Device Assessments.................................................................................. 9-43
9.6.2.1 Assessment for Selected Tidal Power Plants in Maine and Washington....... 9-44

xxiii

9.6.2.2 Assessment for Typical TISEC Device Installed in Washington State........... 9-46
9.6.3 TISEC Economic Feasibility .................................................................................. 9-49
9.7 Environmental Impact Issues ........................................................................................ 9-51
9.8 Installed Capacity and Estimated Growth ..................................................................... 9-53
9.9 Research Focus ............................................................................................................ 9-53
9.10 Conclusion................................................................................................................... 9-55
9.11 Internet Resources ...................................................................................................... 9-55
9.12 References .................................................................................................................. 9-55
10 OCEAN WAVE ENERGY ................................................................................................... 10-1
10.1 Introduction.................................................................................................................. 10-1
10.2 U.S. Wave Energy Highlights: Late 2011 to Mid-2012 ................................................ 10-6
10.2.1 U.S. Federal Highlights ........................................................................................ 10-6
10.2.1.1 Federal Appropriations................................................................................. 10-6
10.2.1.2 Funding Awards ........................................................................................... 10-7
10.2.1.3 Federal Regulations ................................................................................... 10-10
10.2.2 State Highlights ................................................................................................. 10-10
10.2.2.1 Projects in Hawaii ...................................................................................... 10-10
10.2.2.2 Projects in Oregon ..................................................................................... 10-12
10.2.2.3 Projects in California .................................................................................. 10-15
10.2.3 U.S. Developer and Project Deployment Highlights .......................................... 10-18
10.2.3.1 Ocean Power Technologies ....................................................................... 10-18
10.3 Worldwide Wave Energy Highlights: Late 2011 to Mid-2012 .................................... 10-19
10.3.1 United Kingdom and Ireland .............................................................................. 10-19
10.3.1.1 Wave Energy Developers .......................................................................... 10-19
10.3.1.2 Wave Energy Test Centers ........................................................................ 10-21
10.3.2 Portugal and Spain ............................................................................................ 10-24
10.3.2.1 Developments in Portugal .......................................................................... 10-24
10.3.2.1 Developments in Spain .............................................................................. 10-25
10.3.3 Denmark, Norway and Sweden ......................................................................... 10-25
10.3.4 Australia, New Zealand, and Tasmania ............................................................. 10-26
10.3.4.1 BioPower Systems Pty Ltd......................................................................... 10-27
10.3.4.2 Oceanlinx Limited ...................................................................................... 10-27
10.3.4.3 Aotearoa Wave and Tidal Energy Association (AWATEA) of New
Zealand .................................................................................................................... 10-28

xxiv

10.4 Wave Power and Energy Resources ........................................................................ 10-28


10.4.1 Measurement Data Sources .............................................................................. 10-33
10.4.2 Wind-Wave Model Data Sources ...................................................................... 10-33
10.4.3 Wave Power Forecasting .................................................................................. 10-34
10.5 Wave Energy Conversion (WEC) Technology Description ....................................... 10-35
10.5.1 Harnessing Wave Energy .................................................................................. 10-35
10.5.2 WEC System Developers .................................................................................. 10-37
10.5.3 Survival in Storms and Hostile Marine Environments ........................................ 10-45
10.5.4 Effect of Wave Power Plants on the Environment ............................................. 10-45
10.5.4.1 Report to Congress: Potential Environmental Effects of Marine and
Hydrokinetic Energy Technologies .......................................................................... 10-45
10.5.5 Permits for Offshore Wave Power Plants .......................................................... 10-47
10.5.6 Overview of Regulatory Status for Offshore Wave Power Plants ...................... 10-47
10.5.7 WEC Power Plant Footprints ............................................................................. 10-48
10.6 Design, Performance, Cost, and Economic Feasibility Issues .................................. 10-49
10.6.1 WEC Sites ......................................................................................................... 10-49
10.6.2 WEC Design, Performance, and Cost ............................................................... 10-50
10.6.2.1 Assessment for Linear Absorber Devices in Massachusetts and
Oregon ..................................................................................................................... 10-51
10.6.2.2 Assessment for Typical Wave Power Device Installed in Northern
California .................................................................................................................. 10-53
10.6.3 WEC Economic Feasibility ................................................................................ 10-56
10.7 Installed Capacity and Estimated Growth ................................................................. 10-59
10.8 Research Focus ........................................................................................................ 10-59
10.9 Conclusions ............................................................................................................... 10-60
10.10 Internet Resources .................................................................................................. 10-61
10.11 References .............................................................................................................. 10-61
11 RIVER IN-STREAM ENERGY ............................................................................................ 11-1
11.1 Introduction.................................................................................................................. 11-1
11.1.1 Main Components of River Hydrokinetic Resource ............................................. 11-3
11.1.2 The Conversion of River In-Stream Hydrokinetic Energy to Electricity ............... 11-4
11.2 U.S. In-Stream River Energy Highlights: Late 2011 to Mid-2012 ................................ 11-6
11.2.1 U.S. Federal Highlights ........................................................................................ 11-6
11.2.1.1 Federal Appropriations................................................................................. 11-6
11.2.1.2 Funding Awards ........................................................................................... 11-7

xxv

11.2.1.3 Federal Regulations ..................................................................................... 11-8


11.2.2 State and Regional Highlights ............................................................................. 11-9
11.2.2.1 Developments in Alaska .............................................................................. 11-9
11.2.2.3 Mississippi River Projects .......................................................................... 11-11
11.2.2.4 Detroit River Projects ................................................................................. 11-14
11.3 Canadian River In-Stream Energy Highlights: Late 2011 toMid-2012....................... 11-15
11.4 River In-Stream Power and Energy Resources ........................................................ 11-17
11.5 River In-Stream Energy Conversion Development ................................................... 11-20
11.5.1 Harnessing In-Stream River Energy .................................................................. 11-21
11.5.2 RISEC Technology Developers ......................................................................... 11-22
11.5.2.1 Aero Hydro Research and Technology Associates ................................... 11-23
11.5.2.2 Free Flow Power ........................................................................................ 11-24
11.5.2.3 Hydro Green .............................................................................................. 11-25
11.5.2.4 Lucid Energy Technologies........................................................................ 11-26
11.5.2.5 New Energy Corp. ..................................................................................... 11-27
11.5.2.6 Ocean Renewable Power Corp. ................................................................ 11-28
11.5.2.7 UEK Corp. .................................................................................................. 11-29
11.5.2.8 Verdant Power ........................................................................................... 11-29
11.5.2.9 Vortex Hydro .............................................................................................. 11-31
11.5.2.10 Whitestone Power & Communications..................................................... 11-31
11.5.3 Survival in Hostile River Environments .............................................................. 11-32
11.5.4 Environmental Impacts of In-Stream River Power Plants .................................. 11-33
11.5.5 Permits for In-Stream River Power Plants ......................................................... 11-33
11.5.6 Overview of Regulatory Status for River In-Stream Energy Conversion
Projects......................................................................................................................... 11-33
11.6 Design, Performance, Cost, and Economic Feasibility ............................................. 11-34
11.6.1 RISEC Sites....................................................................................................... 11-34
11.6.2 RISEC Devices Studied..................................................................................... 11-34
11.6.3 RISEC Design, Performance, and Cost ............................................................ 11-37
11.6.3.1 EPRI Methodology ..................................................................................... 11-37
11.6.3.2 Kilowatt-Scale Assessment: Alaska ........................................................... 11-39
11.6.3.3 Megawatt-Scale Assessment: Mississippi River, Louisiana ...................... 11-41
11.6.4 RISEC Economic Feasibility .............................................................................. 11-43
11.7 Environmental Issues ................................................................................................ 11-45
11.8 Installed Capacity and Estimated Growth ................................................................. 11-46

xxvi

11.9 Research Focus ........................................................................................................ 11-47


11.10 Conclusions ............................................................................................................. 11-48
11.11 Internet Resources .................................................................................................. 11-48
11.12 References .............................................................................................................. 11-48
12 GRID INTEGRATION CHALLENGES AND TECHNOLOGIES ......................................... 12-1
12.1 Basics of Power System Operation and Planning ....................................................... 12-1
12.1.1 Functions of System Generation ......................................................................... 12-1
12.1.2 Expectations Based on Traditional Generator Performance ............................... 12-2
12.1.3 Typical Power System Operations and Planning ................................................ 12-4
12.1.3.1 Resource Adequacy and Transmission Planning ........................................ 12-5
12.1.3.2 Unit Commitment ......................................................................................... 12-6
12.1.3.3 Economic Dispatch ...................................................................................... 12-6
12.1.3.4 Role of Ancillary Services ............................................................................ 12-7
12.1.3.5 Voltage Control .......................................................................................... 12-11
12.1.3.6 Frequency Response ................................................................................. 12-11
12.2 Characteristics of Variable Generation and Experience to Date ............................... 12-12
12.2.1 Variability ........................................................................................................... 12-14
12.2.2 Uncertainty ........................................................................................................ 12-17
12.2.3 Non-Synchronous Nature of the Resource ........................................................ 12-17
12.2.4 Distributed Nature of the Generation ................................................................. 12-18
12.2.5 Remote Location of Some Renewable Resources ............................................ 12-18
12.2.6 Experience of Integration of Wind Power .......................................................... 12-18
12.2.7 Experience of Integration of Solar Power .......................................................... 12-19
12.3 Variable Generation Integration Impacts ................................................................... 12-21
12.3.1 Transmission Connected Generation Impacts .................................................. 12-21
12.3.2 Distribution Connected Generation Impacts ...................................................... 12-25
12.4 Integration Technologies and Strategies for Transmission-Connected
Renewable Generation...................................................................................................... 12-30
12.4.1 Variable Generator Interface to Maintain System Stability ................................ 12-30
12.4.2 Variable Energy Forecasting Tools ................................................................... 12-31
12.4.3 Increased Diversity through Increased Transmission and BA Cooperation ...... 12-33
12.4.4 Flexibility from Conventional (Non-Variable Generation) Plant ......................... 12-34
12.4.5 Flexibility from Storage and Demand Side Resources ...................................... 12-35
12.4.6 Grid Planning and Operating Tools ................................................................... 12-36

xxvii

12.5 Integration Technologies and Strategies for Distribution-Connected Renewable


Generation ........................................................................................................................ 12-37
12.6 References ................................................................................................................ 12-43
12.6.1 In-Text Citations ................................................................................................ 12-43
12.6.2 General .............................................................................................................. 12-45
12.6.3 Renewable Generation Grid Impacts and Integration ....................................... 12-45
13 GREENHOUSE GAS EMISSIONS CONTROL .................................................................. 13-1
13.1 Greenhouse Gases ..................................................................................................... 13-1
13.2 Role of Renewable Energy Technology ...................................................................... 13-1
13.3 Fossil Carbon Intensity of Fuels .................................................................................. 13-2
13.4 CO2 Emissions Offsets ................................................................................................ 13-3
13.4.1 Factors Affecting CO2 Emissions Offset .............................................................. 13-3
13.4.1.1 Wind ............................................................................................................. 13-4
13.4.1.2 Solar PV and Thermal.................................................................................. 13-4
13.4.1.3 Biomass, Geothermal, and Hydro ................................................................ 13-4
13.4.1.4 Planting Trees and Other Fast-Growing Crops............................................ 13-4
13.4.2 Levelized CO2 Emissions Offsets ........................................................................ 13-5
13.5 Factors Affecting Generation Cost of Renewable Technologies ................................. 13-7
13.5.1 Wind .................................................................................................................... 13-7
13.5.2 Biomass ............................................................................................................. 13-10
13.5.3 Solar Photovoltaics ............................................................................................ 13-11
13.5.4 Geothermal ........................................................................................................ 13-13
13.5.5 Solar Thermal .................................................................................................... 13-15
13.6 References ................................................................................................................ 13-17

xxviii

LIST OF FIGURES
Figure 1-1 Renewable energy as share of U.S. total primary energy consumption, 2011 ......... 1-2
Figure 1-2 U.S. renewable energy consumption by source, 2011 ............................................. 1-3
Figure 1-3 U.S. non-hydroelectric renewable power sources, 19892011 ................................ 1-3
Figure 2-1 Day-ahead prices for the California Power Exchange ............................................ 2-19
Figure 2-2 Example price duration curve ................................................................................. 2-20
Figure 2-3 Components of busbar cost for example 200-MW wind generation plant .............. 2-24
Figure 3-1 Historical (2011) and projected (20122016) installed wind generation
capacity in the United States, Europe, and remainder of the world ................................... 3-3
Figure 3-2 End-of-year operating wind capacity by country and region: 20082011 ................. 3-6
Figure 3-3 Wind capacity and forecast: 19902021 .................................................................. 3-7
Figure 3-4 Installed wind generation capacity in the United States by state as of 3Q 2012 ...... 3-8
Figure 3-5 Rayleigh probability density function for wind speed .............................................. 3-10
Figure 3-6 Coefficient of performance (Cp) vs. tip speed/wind speed ratio ............................. 3-10
Figure 3-7 Components of a typical wind turbine ..................................................................... 3-11
Figure 3-8 Schematic of Enercon direct drive turbine .............................................................. 3-14
Figure 3-9 Direct-drive 1.5-MW generator under testing at NRELs National Wind
Technology Center ........................................................................................................... 3-15
Figure 3-10 Example of an eight-pole, air-cooled superconducting generator ........................ 3-16
Figure 3-11 Cross section of wind turbine blade Source: DNV-GEC ....................................... 3-17
Figure 3-12 Bend-twist coupling achieved by fiber orientation ................................................ 3-18
Figure 3-13 Vortex generators (counter-rotating array) ........................................................... 3-19
Figure 3-14 Gearbox internal schematic showing one planetary and two parallel-shaft
stages ............................................................................................................................... 3-27
Figure 3-15 Cross section of hydrodynamic drive system for a wind turbine ........................... 3-28
Figure 3-16 Electrical diagram of a typical direct-drive turbine ................................................ 3-29
Figure 3-17 Example hybrid steel/concrete and concrete towers ............................................ 3-31
Figure 3-18 Artists rendering of the 2-B 6-MW wind turbine ................................................... 3-33
Figure 3-19 Artists rendering of the Vertiwind offshore VAWTs .............................................. 3-35
Figure 3-20 Guoneng 1-MW, 8-bladed VAWT with 36-m blades ............................................. 3-36
Figure 3-21 Offshore wind support structures .......................................................................... 3-38
Figure 3-22 Offshore wind monopile foundation showing grouted connection ........................ 3-39
Figure 3-23 Monopile foundation ............................................................................................. 3-41
Figure 3-24 Gravity-based foundations .................................................................................... 3-43

xxix

Figure 3-25 Tripod foundation .................................................................................................. 3-44


Figure 3-26 Jacket foundations ................................................................................................ 3-45
Figure 3-27 Tension leg floating foundation ............................................................................. 3-47
Figure 3-28 Manufacturing plants for turbine blades and other components ........................... 3-50
Figure 3-29 Top 10 suppliers market share and presence ..................................................... 3-51
Figure 3-30 Clipper Windpowers 2.5-MW Liberty Series wind turbine nacelle, with four
modular permanent magnet generators ........................................................................... 3-52
Figure 3-31 Breakdown of estimated capital costs for land-based wind plants ....................... 3-55
Figure 3-32 Offshore wind: CAPEX distribution by category ................................................... 3-64
Figure 3-33 Factors that influence wind farm availability ......................................................... 3-66
Figure 3-34 Distribution of turbine failures by components ...................................................... 3-69
Figure 3-35 Comparison of symmetrical and asymmetrical wind farm layouts ........................ 3-73
Figure 3-36 Northeastern project location ................................................................................ 3-76
Figure 3-37 Northeastern U.S. project: wind farm layout ......................................................... 3-77
Figure 3-38 Power curve of STW-3.6-120 ............................................................................... 3-82
Figure 3-39 Great Lakes project area ...................................................................................... 3-86
Figure 3-40 Wind farm layout ................................................................................................... 3-87
Figure 3-41 Power curve of STW-3.6.120 (8.2 m/s) ................................................................ 3-91
Figure 3-42 UK project locations ............................................................................................ 3-100
Figure 3-43 UK wind farm layout ........................................................................................... 3-101
Figure 3-44 Power curve of STW-3.6-120 (8.9 m/s) .............................................................. 3-104
Figure 3-45 Breakdown of estimated capital costs for selected installed offshore UK wind
plants .............................................................................................................................. 3-109
Figure 3-46 California plant-provided O&M costs over time compared with contracted
O&M costs in years 15 (inflation rate = 2.5%) .............................................................. 3-116
Figure 3-47 California plant-provided other expenses over time (inflation rate = 2.5%) ...... 3-117
Figure 3-48 Sensitivity of LCOE to varied parameters (California case)................................ 3-121
Figure 3-49 LCOE probabilistic analysis resultswind plant 1 (California) ........................... 3-122
Figure 3-50 LCOE probabilistic analysis resultswind plant 2 (Texas) ................................ 3-122
Figure 3-51 LCOE probabilistic analysis resultswind plant 3 (Michigan) ............................ 3-123
Figure 3-52 LCOE probabilistic analysis resultswind plant 4 (New York)........................... 3-123
Figure 3-53 LCOE probabilistic analysis resultswind plant 5 (Washington) ....................... 3-124
Figure 3-54 LCOE probabilistic analysis resultswind plant 6 (Georgia) ............................. 3-124
Figure 3-55 LCOE probabilistic analysis resultswind plant 7, tax incentive (Brazil) ........... 3-125
Figure 3-56 LCOE probabilistic analysis resultswind plant 8, CDM (Australia) .................. 3-125
Figure 3-57 Sensitivity of LCOE to varied parametersno PTC (California case) ................ 3-128
Figure 3-58 LCOE probabilistic analysis resultswind plant 1, no PTC (California) ............. 3-129
Figure 3-59 LCOE probabilistic analysis resultswind plant 2, no PTC (Texas) .................. 3-129
Figure 3-60 LCOE probabilistic analysis resultswind plant 3, no PTC (Michigan).............. 3-130

xxx

Figure 3-61 LCOE probabilistic analysis resultswind plant 4, no PTC (New York) ............ 3-130
Figure 3-62 LCOE probabilistic analysis resultswind plant 5, no PTC (Washington) ......... 3-131
Figure 3-63 LCOE probabilistic analysis resultswind plant 6, no PTC (Georgia) ............... 3-131
Figure 3-64 LCOE probabilistic analysis resultswind plant 7, no incentives (Brazil) .......... 3-132
Figure 3-65 LCOE probabilistic analysis resultswind plant 8, no incentives (Australia) ..... 3-132
Figure 3-66 Distribution of average wind power in the United States .................................... 3-139
Figure 3-67 Number of O&M personnel at a wind power plant versus the number of
turbines and project rating .............................................................................................. 3-142
Figure 3-68 Shadow flicker impact illustration (AWEA) ......................................................... 3-146
Figure 3-69 The importance of distance in shadow flicker prevention (CH2MHill) ................ 3-147
Figure 4-1 Renewable electricity projections, excluding hydropower (billion kWh/yr)................ 4-3
Figure 4-2 Breakdown of biomass feedstock ............................................................................. 4-9
Figure 4-3 Effect of biomass plant scale on feedstock requirements and plantation area
required ............................................................................................................................ 4-10
Figure 4-4 Maximum volatile yield of biomass fuels compared to reference coals .................. 4-19
Figure 4-5 Fuel nitrogen concentrations for the fuel samples analyzed in detail ..................... 4-20
Figure 4-6 Distribution of volatile and char fuel nitrogen in selected biomass and
reference coal samples .................................................................................................... 4-21
Figure 4-7 Maximum nitrogen volatile yields for selected biomass and coal samples ............. 4-22
Figure 4-8 Volatile nitrogen and carbon evolution from fresh sawdust .................................... 4-23
Figure 4-9 Volatile nitrogen and carbon evolution from urban wood waste ............................. 4-23
Figure 4-10 Volatile nitrogen and carbon evolution from fresh switchgrass ............................ 4-24
Figure 4-11 Volatile nitrogen and carbon evolution from weathered switchgrass .................... 4-24
Figure 4-12 Nitrogen and carbon volatilization normalized to total volatiles formed from
sawdust ............................................................................................................................ 4-25
Figure 4-13 Nitrogen and carbon volatile evolution from weathered switchgrass
normalized to total volatile evolution ................................................................................ 4-25
Figure 4-14 Nitrogen/carbon atomic ratios for biomass chars formed during DTR
pyrolysis at various temperatures .................................................................................... 4-26
Figure 4-15 Normalized nitrogen/carbon atomic ratios for biomass solid fuel and char,
compared to Pittsburgh #8 bituminous coal and Black Thunder Powder River Basin
sub-bituminous coal ......................................................................................................... 4-27
Figure 4-16 Schematic of biomass cofiring with coal in a utility boiler ..................................... 4-30
Figure 4-17 Plan view of the Willow Island biomass cofiring system ....................................... 4-35
Figure 4-18 Elevation view of the Willow Island biomass cofiring system ............................... 4-36
Figure 4-19 Receiving sawdust at the Willow Island Generating Station ................................. 4-37
Figure 4-20 Influence of cofiring in cyclone boilers on furnace exit gas temperature .............. 4-39
Figure 4-21 Impact of cofiring on slag viscosity ....................................................................... 4-40
Figure 4-22 Design of the sawdust injection system at Seward Generating Station................ 4-43
Figure 4-23 NOX reduction at the Albright Generating Station ................................................. 4-46

xxxi

Figure 4-24 Representative wood fuel handling system .......................................................... 4-52


Figure 4-25 Schematic of biomass-fired stoker boiler power plant .......................................... 4-53
Figure 4-26 Typical boiler and turbine-generator approach ..................................................... 4-54
Figure 4-27 Schematic of biomass-fired atmospheric fluidized boiler power plant .................. 4-55
Figure 4-28 Air pollution control system, cooling tower, and demineralizer for a typical
wood-waste-fired plant ..................................................................................................... 4-57
Figure 4-29 Gasifier types for biomass .................................................................................... 4-61
Figure 4-30 Updraft gasifier ..................................................................................................... 4-62
Figure 4-31 Downdraft gasifier ................................................................................................. 4-63
Figure 4-32 Fluid bed gasifier .................................................................................................. 4-64
Figure 4-33 Circulating pyrolysis .............................................................................................. 4-65
Figure 4-34 Schematic of biomass gasification application for combined heat and power ...... 4-68
Figure 4-35 Process flow diagram for the Vlund Gasifier at Harbore .................................. 4-69
Figure 4-36 Fjernvarme process flow diagram for coproduction of 6 MW electric and 10
MW thermal ...................................................................................................................... 4-70
Figure 4-37 Overall configuration of the biomass gasifier at Ruien ......................................... 4-71
Figure 4-38 Schematic of biomass gasification and power generation system at
Kokomki ......................................................................................................................... 4-72
Figure 4-39 Example bubbling bed retrofit to a pulverized coal boiler ..................................... 4-74
Figure 4-40 LCOE cost comparisonscomparison by region ................................................. 4-97
Figure 5-1 MSW management by country ................................................................................. 5-2
Figure 5-2 States with WTE facilites .......................................................................................... 5-3
Figure 5-3 Waterwall furnace section ...................................................................................... 5-14
Figure 5-4 Typical mass-burn waterwall system ...................................................................... 5-15
Figure 5-5 Typical modular combustion system ...................................................................... 5-19
Figure 5-6 Simplistic RDF processing facility ........................................................................... 5-22
Figure 5-7 Typical RDF combustion facility ............................................................................. 5-23
Figure 5-8 Fluidized bed .......................................................................................................... 5-25
Figure 5-9 Typical RDF fluid bed system ................................................................................. 5-25
Figure 5-10 RDF fluidized bed gasification system .................................................................. 5-26
Figure 5-11 Simplified landfill gas collection system ................................................................ 5-28
Figure 5-12 Landfill gas collection and electricity production layout ........................................ 5-29
Figure 5-13 Typical LFG engine/generator set ........................................................................ 5-30
Figure 5-14 LFG cleaning diagram with CO2 wash .................................................................. 5-32
Figure 5-15 Landfill gas compression and cleaning unit .......................................................... 5-32
Figure 5-16 Electrabel Plant, Ruien, Belgium .......................................................................... 5-36
Figure 5-17 Process diagram of a pyrolysis system ................................................................ 5-39
Figure 5-18 Typical gasification system ................................................................................... 5-40
Figure 5-19 EnTech process schematic .................................................................................. 5-41

xxxii

Figure 5-20 Cross-section of a plasma arc furnace ................................................................. 5-42


Figure 5-21 Process flow for anaerobic digestion system ....................................................... 5-46
Figure 5-22 ArrowBio facility in Sydney, Australia ................................................................... 5-47
Figure 5-23 Simplified mass burn waste-to-energy facility schematic ..................................... 5-55
Figure 5-24 Waste-to-energy system schematic ..................................................................... 5-60
Figure 5-25 Simplified system schematic ................................................................................ 5-72
Figure 5-26 RDF burner ........................................................................................................... 5-74
Figure 5-27 Dump grate ........................................................................................................... 5-74
Figure 5-28 Hydrogen chloride dew point ................................................................................ 5-77
Figure 5-29 RDF co-firing processing diagram ........................................................................ 5-81
Figure 5-30 MEDAL CO2 removal system ............................................................................... 5-93
Figure 5-31 Landfill gas utilization process diagram ................................................................ 5-95
Figure 6-1 Worldwide PV industry growth .................................................................................. 6-4
Figure 6-2 View down one row in 12-MW section of EdF Energies Nouvelles 115-MW
plant at Toul, France using First Solar modules ................................................................. 6-6
Figure 6-3 California Solar Initiative installed projects by month ............................................... 6-8
Figure 6-4 Annual U.S. installed grid-connected PV capacity by sector .................................... 6-9
Figure 6-5 Historic PV market growth by country .................................................................... 6-10
Figure 6-6 German PV installations between January 2009 and March 2010 showing
surge of projects, particularly larger than 10 kW, at years end in anticipation of
declining feed-in tariff ....................................................................................................... 6-12
Figure 6-7 LCOE probabilistic analysis: fixed a-Si modules, Las Vegas, NV (4th Quarter
2011$) .............................................................................................................................. 6-21
Figure 6-8 Worldwide average PV module selling price vs. cumulative sales ......................... 6-24
Figure 6-9 Trends in worldwide average PV module selling price vs. cumulative sales .......... 6-25
Figure 6-10 U.S. shipments of PV by market sector (top) and end use (bottom) in 2009........ 6-45
Figure 6-11 Distribution of installed PV applications in countries participating in the
International Energy Agency (IEA) Photovoltaic Power Systems Program in 2011 ......... 6-46
Figure 6-12 Annual solar cell production by country ................................................................ 6-47
Figure 7-1 Map illustrating the Ring of Fire ............................................................................. 7-2
Figure 7-2 Distribution curve of geothermal energy as a function of worldwide
temperature ........................................................................................................................ 7-7
Figure 7-3 U.S. geothermal resource potential map (estimated temperatures [C] at 6 km
depth) (top), 2004 SMU heat flow assessment (bottom) .................................................... 7-9
Figure 7-4 Estimated EGS resources in the western United States ........................................ 7-10
Figure 7-5 Total installed capacity in 2007 (GHC Bulletin, September 2007) .......................... 7-11
Figure 7-6 Total U.S. installed geothermal capacity as of March 2009.................................... 7-13
Figure 7-7 Hydrothermal power plant at the Geysers, California ............................................. 7-14
Figure 7-8 Dual-flash (two-flash) geothermal power plant ....................................................... 7-18
Figure 7-9 Binary cycle geothermal power plant (air-cooled design) ....................................... 7-19

xxxiii

Figure 7-10 HDR geothermal production process ................................................................... 7-22


Figure 7-11 Hot sedimentary aquifer production process ........................................................ 7-24
Figure 7-12 Geothermal-solar hybrid production process for a geothermal flash steam
plant ................................................................................................................................. 7-25
Figure 7-13 Geothermal-solar hybrid production process for a hot water/brine non-flash
resource ........................................................................................................................... 7-26
Figure 7-14 Geothermal-gas peaker production process ........................................................ 7-26
Figure 7-15 Geo-biomass hybrid cycle .................................................................................... 7-27
Figure 7-16 The Grubb curve ................................................................................................... 7-34
Figure 7-17 Steam pipelines for the Ohaaki, New Zealand plant run through land used
for grazing and agriforestry .............................................................................................. 7-45
Figure 8-1 Distribution of direct-normal insolation worldwide (top, kWh/m2/yr) and U.S.
(bottom, Wh/m2/day) .......................................................................................................... 8-9
Figure 8-2 Solar trough collector field at Kramer Junction, California ...................................... 8-11
Figure 8-3 Andasol 1 and 2 parabolic trough plants, Spain .................................................... 8-11
Figure 8-4 Solar Two central receiver in operation .................................................................. 8-14
Figure 8-5 25-kW SAIC dish/engine system at the DOE Mesa Top Thermal Test Facility ...... 8-17
Figure 8-6 LFR Array at Liddell Power Station, Australia ........................................................ 8-20
Figure 8-7 Cross-section diagram of solar chimney ................................................................ 8-21
Figure 8-8 Solar chimney sculpture ......................................................................................... 8-22
Figure 8-9 Martin Next-Generation Solar Energy Center ......................................................... 8-27
Figure 8-10 Gemasolar plant under construction, May 2010 ................................................... 8-29
Figure 8-11 Gemasolar molten salt storage tanks under construction, June 2010 .................. 8-29
Figure 8-12 Gemasolar receiver tower and heliostat ............................................................... 8-30
Figure 8-13 Sierra SunTower plant using modular tower approach ........................................ 8-31
Figure 8-14 Thirty Infinia PowerDish units demonstrated in Spain .......................................... 8-33
Figure 9-1 Earth, Moon, and Suns influence on tides ............................................................... 9-4
Figure 9-2 The Earths tidal bulge .............................................................................................. 9-4
Figure 9-3 Typical hydrokinetic water turbine ............................................................................ 9-5
Figure 9-4 Tidal power density ................................................................................................... 9-5
Figure 9-5 Vortex-induced vibrations oscillate objects in fluid currents ................................... 9-17
Figure 9-6 OpenHydro device at the EMEC Fall of Warness tidal test site.............................. 9-21
Figure 9-7 Marine Current Turbines SeaGen .......................................................................... 9-22
Figure 9-8 Pulse Tidal generator ............................................................................................. 9-23
Figure 9-9 Swanturbines Cygnet ............................................................................................. 9-24
Figure 9-10 Tidal Generation Limited DeltaStream ................................................................. 9-25
Figure 9-11 OpenHydro Gravity Base installation .................................................................... 9-27
Figure 9-12 OpenHydro Gravity Base photo ............................................................................ 9-27
Figure 9-13 OpenHydro deployment vessel ............................................................................ 9-28

xxxiv

Figure 9-14 United States tidal current reference stations ....................................................... 9-31
Figure 9-15 Annual average tidal current speed probability distributions for Dog Island
Transect, Western Passage, Maine ................................................................................. 9-32
Figure 9-16 Annual tidal energy resource for six U.S. tidal current sites ................................. 9-32
Figure 9-17 Tidal energy conversion system configurations .................................................... 9-35
Figure 9-18 Capital cost breakdown for a typical tidal power technology ................................ 9-48
Figure 9-19 Operational cost breakdown for a typical tidal power technology ......................... 9-48
Figure 9-20 Cost of electricity as a function of power density .................................................. 9-49
Figure 9-21 Cost projection as a function of development status ............................................ 9-50
Figure 10-1 Wind blowing over fetch of water, producing waves ............................................. 10-4
Figure 10-2 Particle motion in different water depths .............................................................. 10-4
Figure 10-3 Vector field for particle motion in waves ............................................................... 10-5
Figure 10-4 Wave power flux ................................................................................................... 10-5
Figure 10-5 Schematic of PG&E WaveConnect system ........................................................ 10-16
Figure 10-6 Aquamarine Power Oyster ................................................................................. 10-20
Figure 10-7 EMEC testing configuration in 2011 ................................................................... 10-23
Figure 10-8 Wave Star 1:10 machine with buoys raised ....................................................... 10-25
Figure 10-9 Wave Star 1:2 machine with buoys raised ......................................................... 10-25
Figure 10-10 Floating power plant AS Poseidon ................................................................... 10-26
Figure 10-11 Seabased AB linear generator ......................................................................... 10-26
Figure 10-12 BioPower BioWave ........................................................................................... 10-27
Figure 10-13 Oceanlinx oscillating water column .................................................................. 10-27
Figure 10-14 West Coast reference stations ......................................................................... 10-33
Figure 10-15 Hawaii reference stations ................................................................................. 10-33
Figure 10-16 NOAA Wave Watch III global coverage ............................................................ 10-34
Figure 10-17 Wave energy device principles ......................................................................... 10-36
Figure 10-18 Wave energy device concept ........................................................................... 10-37
Figure 10-19 Capital cost breakdown for a typical wave energy technology ......................... 10-55
Figure 10-20 Operational cost breakdown for a typical wave energy technology (in $/kWyear) ............................................................................................................................... 10-55
Figure 10-21 Cost of electricity as a function of the power density at the deployment site.... 10-56
Figure 10-22 Cost projection as a function of development status ........................................ 10-57
Figure 10-23 Levelized COE comparison to wind: Oregon example with federal and state
financial incentives ......................................................................................................... 10-58
Figure 11-1 Major North American rivers and their yearly discharges in km3/year .................. 11-3
Figure 11-2 Steps affecting hydrokinetic turbine efficiency ...................................................... 11-5
Figure 11-3 Example of a hydrokinetic turbine ........................................................................ 11-5
Figure 11-4 Water power density ............................................................................................. 11-5

xxxv

Figure 11-5 Hydrokinetic turbine deployment from barge at Ruby, Alaska, on the Yukon
River ................................................................................................................................. 11-9
Figure 11-6 Village of Eagle during spring breakup of 2009 .................................................. 11-10
Figure 11-7 Hydro Greens Hydrokinetic Turbine (left) and the deployment site at lock
and dam #2 on the Mississippi River at Hastings, Minnesota (right).............................. 11-12
Figure 11-8 Preliminary permit sites held by Free Flow Power on the Mississippi River;
Memphis District (upper left), Vicksburg District (upper right), and New Orleans
District (lower left) of the U.S. Army Corps of Engineers ............................................... 11-13
Figure 11-9 Pre-test inspection of Free Flow Power 3-meter turbine in flume at the
USGS Conte Anadromous Fish Laboratory ................................................................... 11-14
Figure 11-10 Location of Verdants Cornwall Ontario Renewable Energy (CORE) Project
in the St. Lawrence River ............................................................................................... 11-15
Figure 11-11 Site of the CORE Project on the St. Lawrence River (left), and Verdant
Turbine (right) ................................................................................................................. 11-16
Figure 11-12 Positioning the anchor for the EnCurrent turbine deployment at Fort
Simpson, NT, Canada .................................................................................................... 11-16
Figure 11-13 Stream velocity profile across the Yukon River at the USGS Gauging
Station at Eagle, Alaska for a discharge rate of 183,000 ft3/s ........................................ 11-18
Figure 11-14 Channel cross section at the USGS Gauging Station at Eagle, Alaska, at a
discharge rate of 183,000 ft3/s ....................................................................................... 11-18
Figure 11-15 Velocity versus discharge at the USGS Gauging Station at Eagle, Alaska ...... 11-19
Figure 11-16 Velocity distribution at the USGS Gauging Station at Eagle, Alaska ................ 11-19
Figure 11-17 Average discharge by month at the USGS Gauging Station at Eagle,
Alaska ............................................................................................................................ 11-20
Figure 11-18 Average velocity by month at the USGS Gauging Station at Eagle, Alaska..... 11-20
Figure 11-19 River in-stream energy conversion devices ...................................................... 11-21
Figure 11-20 Experimental configuration of the AHRTA oscillating turbine ........................... 11-24
Figure 11-21 Free Flow Power SmarTurbine generator ........................................................ 11-24
Figure 11-22 Hydro Green turbines ....................................................................................... 11-25
Figure 11-23 Hydro Green turbine array configuration .......................................................... 11-26
Figure 11-24 Lucid Energy Gorlov Helical Turbine ................................................................ 11-26
Figure 11-25 Lucid Energy array configuration ...................................................................... 11-27
Figure 11-26 New Energy system concept ............................................................................ 11-28
Figure 11-27 New Energy EnCurrent turbine ......................................................................... 11-28
Figure 11-28 Ocean Renewable Power Co. OCGen module ................................................ 11-28
Figure 11-29 UEK prototype demonstrated in 2000 .............................................................. 11-29
Figure 11-30 Verdant Power Free Flow turbine ..................................................................... 11-30
Figure 11-31 Verdant Power turbine lowered into the East River prior to mounting on a
monopile ......................................................................................................................... 11-30
Figure 11-32 Vortex-induced vibrations oscillating objects in fluid currents........................... 11-31
Figure 11-33 Whitestone Power & Communications Microturbine River-In-Stream
device ............................................................................................................................. 11-32

xxxvi

Figure 11-34 Ice jam during spring breakup: upstream view of the Yukon River at Eagle,
Alaska ............................................................................................................................ 11-33
Figure 11-35 Site location overview, with chosen sites indicated in yellow ........................... 11-35
Figure 11-36 Water depth profiles at three Alaska sites during typical river discharges........ 11-35
Figure 11-37 Average water velocities by month at three sites in Alaska .............................. 11-36
Figure 11-38 Average power densities by month at three sites in Alaska ............................. 11-36
Figure 11-39 Cost projection as a function of development status ........................................ 11-38
Figure 11-40 RISEC capital expense as a function of plant scale ......................................... 11-42
Figure 11-41 RISEC O&M expense as a function of plant scale ........................................... 11-42
Figure 11-42 Cost of electricity as a function of mean power flux ......................................... 11-43
Figure 12-1 System operation over minutes/hours (top) to days (bottom) ............................... 12-5
Figure 12-2 Ancillary services distinguished by their deployment times and durations ......... 12-11
Figure 12-3 Wind output varying diurnally, seasonally, and with weather changes ............... 12-14
Figure 12-4 Daily variation of solar PV system output over a month ..................................... 12-15
Figure 12-5 Hourly load shapes with and without wind generation ........................................ 12-16
Figure 12-6 Number of P observation changes in average power for 4 ramp rate
intervals: 10 seconds, 1 minute, 10 minutes, and 1 hour ............................................... 12-16
Figure 12-7 Example of wind power plant output correlation with distance ........................... 12-17
Figure 12-8 Example of wind plant output variation on area control error (ACE)................... 12-19
Figure 12-9 Relative cost of photovoltaic electricity due only to resource variability ............. 12-20
Figure 12-10 Monthly solar energy variation over one year ................................................... 12-20
Figure 12-11 Results from estimates for the increase in balancing and operating costs
due to wind power .......................................................................................................... 12-22
Figure 12-12 Large wind ramp in Texas showing need for increased situational
awareness ...................................................................................................................... 12-24
Figure 12-13 Power factor curves for example wind plant ..................................................... 12-25
Figure 12-14 Today's typical distribution feeder topology ...................................................... 12-26
Figure 12-15 Wind predictions adjusted using self-learning statistical methods .................... 12-32
Figure 12-16 Distributed controller results aggregated to manage area power and
system voltage profiles ................................................................................................... 12-38
Figure 12-17 Cascaded restoration of distributed generators ................................................ 12-38
Figure 12-18 Concept of distribution microgrids of various sizes and levels allowing
reliability islands and grid tie operation .......................................................................... 12-40
Figure 12-19 Distributed controller must be integrated with overall distribution control
system to maximize system value and reduce capacity requirements........................... 12-41
Figure 12-20 FRT limiting curves proposed in the new German grid codes for connecting
PV systems to the medium-voltage power grid .............................................................. 12-42
Figure 13-1 Levelized CO2 emissions offset vs. % fossil fuel generation ................................ 13-6
Figure 13-2 Levelized CO2 emissions control cost vs. cost premium and CO2 emissions
offset ................................................................................................................................ 13-6

xxxvii

Figure 13-3 Levelized CO2 emissions control cost vs. cost premium and fossil fuel mix at
50% generation from fossil fuels ...................................................................................... 13-7
Figure 13-4 Levelized CO2 emissions control cost at 100-MW California wind plant vs.
CO2 emissions offset and system generation cost (4th Quarter 2010$) .......................... 13-9
Figure 13-5 Levelized CO2 emissions control cost vs. location and system generation
cost (4th Quarter 2010$) .................................................................................................. 13-9
Figure 13-6 Levelized CO2 emissions control cost vs. system generation cost for 100%
biomass-repowered and bubbling fluidized bed plants and 10% biomass co-fired
coal plant (incremental cost vs. coal for co-fired plant, 3rd Quarter 2010$) ................... 13-11
Figure 13-7 Levelized CO2 emissions control cost vs. system generation cost for 50-MW
solar photovoltaic plants, Las Vegas, NV (top) and Columbus, OH (bottom)
(December 2010$) ......................................................................................................... 13-13
Figure 13-8 Levelized CO2 emissions control cost for 50-MW flash-steam and binarycycle power plants vs. base system generation cost (December 2010$) ...................... 13-15
Figure 13-9 Levelized CO2 emissions control cost for four solar thermal technologies
(December 2010$) ......................................................................................................... 13-17

xxxviii

LIST OF TABLES
Table 2-1 Calculation of annual fixed charge rates .................................................................... 2-2
Table 2-2 Calculating levelized annual charges ........................................................................ 2-4
Table 2-3 Wind power generationnominal dollar terms .......................................................... 2-5
Table 2-4 Wind power generationconstant dollar terms ......................................................... 2-7
Table 2-5 Sources and uses of construction funds .................................................................. 2-15
Table 2-6 Example income statement ($1000) ........................................................................ 2-18
Table 2-7 Example cash flow statement ($1000) ..................................................................... 2-23
Table 2-8 Example sensitivity analysis .................................................................................... 2-28
Table 2-9 Confidence rating based on cost and design estimate ............................................ 2-31
Table 2-10 Accuracy range estimates for RETG cost data (ranges in percent)1 ..................... 2-32
Table 2-11 TPC-TPI adjustment factors .................................................................................. 2-33
Table 3-1 Wind power overview ................................................................................................. 3-1
Table 3-2 Operating wind generation capacity: end of year 2005end of year 2011................. 3-5
Table 3-3 Forecast wind generation capacity additions: 20122016 ......................................... 3-5
Table 3-4 Estimated U.S. wind generation capcaity by depth ................................................. 3-37
Table 3-5 Top 10 suppliers 2011 ............................................................................................. 3-49
Table 3-6 Selected onshore wind turbine suppliers and turbine models over 2.3 MW ............ 3-53
Table 3-7 Areas of potential technology improvement ............................................................. 3-56
Table 3-8 Typical average foundation costs ............................................................................ 3-60
Table 3-9 Offshore wind: CAPEX costs from project data ....................................................... 3-62
Table 3-10 Offshore wind: total CAPEX and CAPEX share by component category .............. 3-63
Table 3-11 Northeastern U.S. project: design criteria and boundary conditions ...................... 3-78
Table 3-12 Northeastern U.S. project: estimated CAPEX ....................................................... 3-79
Table 3-13 Northeastern U.S. project: total plant investment .................................................. 3-80
Table 3-14 Annual energy yield, Northeastern U.S. project ..................................................... 3-82
Table 3-15 Estimate of full load hours, Northeastern U.S. project ........................................... 3-83
Table 3-16 Financial assumptions, Northeastern U.S. project ................................................. 3-84
Table 3-17 Levelized cost of electricity, Northeastern U.S. project.......................................... 3-85
Table 3-18 Design criteria and boundary conditions for Great Lakes project .......................... 3-88
Table 3-19 Estimated CAPEX for Great Lakes project ............................................................ 3-89
Table 3-20 Total plant investment, Great Lakes project .......................................................... 3-90
Table 3-21 Annual energy yield, Great Lakes project .............................................................. 3-92

xxxix

Table 3-22 Estimated full load hours, Great Lakes project ...................................................... 3-93
Table 3-23 Financial assumptions, Great Lakes project .......................................................... 3-94
Table 3-24 Levelized cost of electricity (LCOE), Great Lakes project ..................................... 3-95
Table 3-25 UK Round 2 offshore wind project status .............................................................. 3-97
Table 3-26 Design criteria and boundary conditions, UK project ............................................. 3-99
Table 3-27 Estimated CAPEX for UK project ......................................................................... 3-103
Table 3-28 Annual energy yield, UK project .......................................................................... 3-105
Table 3-29 Estimate of full load hours, UK project ................................................................. 3-106
Table 3-30 Financial assumptions for the UK project ............................................................ 3-107
Table 3-31 Levelized cost of electricity, UK project ............................................................... 3-108
Table 3-32 Project descriptions ............................................................................................. 3-110
Table 3-33 Conceptual layout design assumptions ............................................................... 3-111
Table 3-34 Total capital requirement estimates summary (2011 US$) .................................. 3-112
Table 3-35 Plant-provided O&M costs (2011$), average of years 15 .................................. 3-114
Table 3-36 Contracted O&M costs (2011$), average over extended warranty period
(years 15) ..................................................................................................................... 3-115
Table 3-37 Wind probabilistic analysisvaried parameters .................................................. 3-118
Table 3-38 Static parameters for financial analysis including incentives ............................... 3-120
Table 3-39 Initial, mean, and 95th percentile LCOE values, including incentives (2011$) .... 3-121
Table 3-40 Static parameters for financial analysis without incentives .................................. 3-127
Table 3-41 Initial, mean, and 95th percentile LCOE values with no incentives (2011$) ........ 3-128
Table 4-1 Biomass electricity generation overview .................................................................... 4-1
Table 4-2 Biomass energy consumption in the U.S. economy by sector and type, 2010
(Quads) .............................................................................................................................. 4-2
Table 4-3 Proximate and ultimate analyses for typical woody biomass fuels .......................... 4-11
Table 4-4 Proximate and ultimate analyses for typical herbaceous biomass fuels .................. 4-12
Table 4-5 Proximate and ultimate analyses for typical manures ............................................. 4-12
Table 4-6 Nitrogen and ash concentrations in biomass fuels (values in lb/106 Btu) ................ 4-14
Table 4-7 Ash analyses of various biomass fuels compared with Pittsburgh #8 coal .............. 4-15
Table 4-8 Ash analyses of switchgrass and other herbaceous crops ...................................... 4-15
Table 4-9 Slagging and fouling index for selected biomass fuels ............................................ 4-16
Table 4-10 Ranges in concentrations of trace metals in woody biomass (mg/kg in dry
wood) ............................................................................................................................... 4-17
Table 4-11 Ranges in concentrations of trace metals in woody biomass burned at a pulp
mill in the Pacific Northwest (mg/kg in dry wood) ............................................................. 4-17
Table 4-12 Volatility measures for representative biomass fuels ............................................. 4-18
Table 4-13 Chemical fractionation analyses of various biofuel ashes ..................................... 4-28
Table 4-14 Chemical fractionation of potassium in switchgrass for four samples ................... 4-29
Table 4-15 Complete chemical fractionation of a switchgrass sample .................................... 4-29

xl

Table 4-16 Representative biomass cofiring tests and demonstrations .................................. 4-32
Table 4-17 Capital cost of the Willow Island cofiring demonstration ........................................ 4-38
Table 4-18 Capital cost summary for the Bailly Generating Station cofiring system................ 4-38
Table 4-19 Estimated capital cost of the Albright cofiring demonstration if constructed as
a new facility ..................................................................................................................... 4-44
Table 4-20 Pre-treatment options for addressing cofiring constraints ..................................... 4-48
Table 4-21 Typical water rates for condensing turbines in wood waste-fired plants ................ 4-55
Table 4-22 Typical water rates for backpressure turbines in wood waste-fired plants............. 4-56
Table 4-23 Plant design assumptions for biomass repowering base case .............................. 4-74
Table 4-24 Raw and dried biomass fuel composition .............................................................. 4-75
Table 4-25 Plant performance summary for PC coal-fired and repowered biomass-fired
bubbling fluidized bed plants ............................................................................................ 4-76
Table 4-26 100% Biomass repowering performance and cost estimates (3Q 2010$) ............. 4-77
Table 4-27 LCOE estimates for 100% biomass repowering (3Q 2010 $) ................................ 4-78
Table 4-28 Summary of biomass cofiring base case ............................................................... 4-79
Table 4-29 Woody biomass fuel quality: heat content, proximate, and ultimate analyses ....... 4-80
Table 4-30 Performance estimates for cofiring coal and undried biomass .............................. 4-81
Table 4-31 Performance estimates for cofiring coal and dried biomass .................................. 4-81
Table 4-32 Performance estimates for cofiring co-milled coal and torrefied biomass.............. 4-82
Table 4-33 Biomass cofiring performance and cost summary (3Q 2010$) .............................. 4-83
Table 4-34 Design assumptions for bubbling BFB boiler plants .............................................. 4-84
Table 4-35 Plant performance estimates for BFB boiler plants ............................................... 4-85
Table 4-36 Performance and cost estimates for bubbling fluidized bed boiler plants (3Q
2010$) .............................................................................................................................. 4-86
Table 4-37 Levelized cost of electricity estimates for biomass-fired bubbling fluidized bed
plants (3Q 2010$) ............................................................................................................ 4-88
Table 4-38 Design assumptions for stoker boiler plant ............................................................ 4-89
Table 4-39 Plant performance estimates for stoker boiler plant .............................................. 4-89
Table 4-40 Performance and cost estimates for stoker boiler plant (3Q 2010$) ..................... 4-90
Table 4-41 Levelized cost of electricity estimates for 50-MW stoker boiler plant fired by
woody biomass (3rd-Quarter 2010$) ............................................................................... 4-92
Table 4-42 Design assumptions for biomass gasification plant ............................................... 4-93
Table 4-43 Plant performance estimates for gasification plant ................................................ 4-94
Table 4-44 Performance and cost estimates for biomass gasification (2012$) ....................... 4-94
Table 4-45 LCOE regional costs with component breakdown ................................................. 4-97
Table 4-46 LCOE without PTC ................................................................................................ 4-99
Table 5-1 MSW overview ........................................................................................................... 5-1
Table 5-2 Permits and review processes ................................................................................... 5-4
Table 5-3 Fuel categories specified by Renewable Fuel Standard 2 (RFS2) regulation ........... 5-8
Table 5-4 Cellulosic biofuel pathways for use in generating RINs ............................................. 5-9

xli

Table 5-5 States defining MSW as a renewable fuel eligible to meet renewable portfolio
standards ......................................................................................................................... 5-10
Table 5-6 States defining waste-to-energy as renewable in state law ..................................... 5-11
Table 5-7 Federal statutes and policies defining WTE as renewable (as of 10/1/10) .............. 5-12
Table 5-8 U.S. operating mass-burn/waterwall facilities and vendors ..................................... 5-17
Table 5-9 Recent WTE expansions and procurements in North America................................ 5-18
Table 5-10 Operating modular system facilities and vendors .................................................. 5-20
Table 5-11 Biomass power facilities co-firing with coal ............................................................ 5-21
Table 5-12 Comparative fuel properties of RDF ...................................................................... 5-24
Table 5-13 U.S. RDF facilities .................................................................................................. 5-27
Table 5-14 Summary of LMOP projects, by conversion technology ........................................ 5-28
Table 5-15 List of representative electricity-generating projects .............................................. 5-30
Table 5-16 List of representative direct heat projects .............................................................. 5-33
Table 5-17 List of representative high-Btu projects ................................................................. 5-33
Table 5-18 Primary constituents of biogas ............................................................................... 5-35
Table 5-19 Operating mixed waste compost facilities .............................................................. 5-45
Table 5-20 Technology maturity .............................................................................................. 5-49
Table 5-21 Typical waste fuel parameters compared with coal ............................................... 5-50
Table 5-22 Municipal solid waste generation and disposal ...................................................... 5-52
Table 5-23 Air emission criteria ............................................................................................... 5-57
Table 5-24 Total capital equipment .......................................................................................... 5-61
Table 5-25 Representative fixed and variable O&M cost categories ....................................... 5-62
Table 5-26 Mass burn capital operations and maintenance costs ........................................... 5-63
Table 5-27 Mass burn facility levelized cost of electricity ........................................................ 5-65
Table 5-28 Mass burn facility levelized fixed capital charge component ................................. 5-66
Table 5-29 Typical heat content of waste and other materials ................................................ 5-69
Table 5-30 Moisture and ash percentages of RDF and coal ................................................... 5-70
Table 5-31 Fuel quality characteristics .................................................................................... 5-79
Table 5-32 RDF capital equipment .......................................................................................... 5-82
Table 5-33 Fixed and variable cost categories ........................................................................ 5-82
Table 5-34 Marginal capital and operations and maintenance costs ....................................... 5-84
Table 5-35 Marginal RDF co-fire levelized cost of electricity ................................................... 5-86
Table 5-36 Marginal RDF co-fire levelized fixed capital charge component ............................ 5-86
Table 5-37 Landfill gas characteristics ..................................................................................... 5-88
Table 5-38 Summary of gas supply parameters for LFG engines ........................................... 5-90
Table 5-39 Landfill gas quality characteristics for injection into pipeline .................................. 5-92
Table 5-40 LFG capital equipment requirements1 ................................................................... 5-96
Table 5-41 Fixed and variable cost categories ........................................................................ 5-97
Table 5-42 LFG to electricity capital operations and maintenance costs ................................. 5-99

xlii

Table 5-43 LFG to direct heat capital operations and maintenance costs ............................. 5-100
Table 5-44 LFG to high-pressure pipeline capital operations and maintenance costs .......... 5-102
Table 5-45 LFG to electricity levelized cost of electricity ....................................................... 5-104
Table 5-46 LFG to direct heat levelized cost of energy ......................................................... 5-105
Table 5-47 LFG to high-pressure pipeline levelized cost of energy ....................................... 5-105
Table 5-48 LFG to electricity levelized fixed capital charge component ................................ 5-106
Table 5-49 LFG to direct heat levelized fixed capital charge component .............................. 5-106
Table 5-50 LFG to high-pressure pipeline levelized fixed capital charge component ............ 5-106
Table 6-1 Overview of solar photovoltaics ................................................................................. 6-2
Table 6-2 PV technology summary .......................................................................................... 6-14
Table 6-3 Photovoltaic site assumptions ................................................................................. 6-15
Table 6-4 First-year electricity generation and capacity factors ............................................... 6-16
Table 6-5 Utility-scale solar photovoltaic power plant total capital requirement estimates
(4th Quarter 2011$) .......................................................................................................... 6-18
Table 6-6 Utility-scale solar photovoltaic power plant operation and maintenance cost
estimates (4th Quarter 2011$) ......................................................................................... 6-19
Table 6-7 Levelized cost of electricity with and without 30% investment tax credit (4th
Quarter 2011$) ................................................................................................................. 6-20
Table 6-8 Utility-scale solar photovoltaic power plant performance and cost estimate
summary (4th Quarter 2011$) .......................................................................................... 6-22
Table 6-9 Site data for 27 locations in the continental United States ...................................... 6-26
Table 6-10 Information included in Clean Power Estimator databases ................................... 6-27
Table 6-11 Rooftop case-study electric-rate scenarios ............................................................ 6-28
Table 6-12 Rooftop case-study utility rates .............................................................................. 6-30
Table 6-13 Rooftop case-study system performance .............................................................. 6-30
Table 6-14 Rooftop case-study capital costs ........................................................................... 6-31
Table 6-15 Rooftop case-study economic-analysis results ...................................................... 6-32
Table 6-16 Solar PV power plant capital, O&M, and lease cost projections for fixed flatplate thin-film photovoltaic power plants (Dec 2009$)...................................................... 6-34
Table 6-17 Solar PV power plant capital, O&M, and lease cost projections for one-axis
tracking crystalline silicon flat-plate photovoltaic power plants (Dec 2009$).................... 6-34
Table 6-18 Solar PV power plant capital, O&M, and lease cost projections for two-axis
tracking high concentration photovoltaic power plants (Dec 2009$) ................................ 6-35
Table 6-19 Solar PV power plant capital and O&M cost scaling factors vs. rated plant
output relative to a 5-MWac plant ...................................................................................... 6-35
Table 6-20 Utility-scale PV power plant O&M cost estimates (2011$)..................................... 6-42
Table 7-1 Geothermal energy overview ..................................................................................... 7-1
Table 7-2 Proposed geothermal resource temperature classification ........................................ 7-5
Table 7-3 Installed geothermal power worldwide ..................................................................... 7-12
Table 7-4 U.S. installed capacity as of March 2009 ................................................................. 7-13

xliii

Table 7-5 Estimated costs for a 50-MW geothermal plant (Dec 2009 $) ................................. 7-39
Table 7-6 Estimated geothermal development costs (Dec 2009 $) ......................................... 7-40
Table 7-7 Design assumptions for geothermal plants .............................................................. 7-41
Table 7-8 Performance and total capital requirement estimates for geothermal power
plants (December 2010 $) ................................................................................................ 7-42
Table 7-9 Levelized cost of electricity estimates for geothermal power plants (30-year
project life, constant December 2010 $) .......................................................................... 7-43
Table 7-10 Land area requirements for current renewable technologies ................................ 7-44
Table 7-11 Averages of four significant pollutants, as emitted from geothermal and coal
facilities ............................................................................................................................ 7-46
Table 8-1 Central receiver technology ....................................................................................... 8-2
Table 8-2 Dish/engine technology ............................................................................................. 8-2
Table 8-3 Linear Fresnel reflector technology ........................................................................... 8-3
Table 8-4 Trough technology ..................................................................................................... 8-3
Table 8-5 Central receiver demonstration projects .................................................................. 8-13
Table 8-6 Technology monitoring guide for solar thermal power plants .................................. 8-23
Table 8-7 Technology process development map: solar thermal power plants ....................... 8-24
Table 8-8 Comparison of the Solar Two and Gemasolar projects ........................................... 8-28
Table 8-9 Performance and cost estimates for solar thermal technologies (December
2010$) .............................................................................................................................. 8-35
Table 8-10 Constant-dollar levelized cost of electricity for solar thermal power plants
(December 2010$) ........................................................................................................... 8-37
Table 8-11 Selected data for solar trough systems ................................................................. 8-38
Table 8-12 Peak central-receiver system efficiency components ............................................ 8-46
Table 8-13 Annual average central receiver system efficiency components ........................... 8-47
Table 9-1 Ocean tidal energy overview ..................................................................................... 9-1
Table 9-2 Offshore in-stream tidal energy conversion device developers ............................... 9-20
Table 9-3 Offshore in-stream tidal energy conversion device developers ............................... 9-37
Table 9-4 Pending FERC preliminary permits for tidal current projects (as of August 2,
2012) ................................................................................................................................ 9-40
Table 9-5 Issued FERC preliminary permits for tidal current projects (as of August 2,
2012) ................................................................................................................................ 9-41
Table 9-6 Cost and performance estimates for selected tidal power plants ............................ 9-45
Table 9-7 Typical cost, performance and economic profiles for tidal power plants (2011$) .... 9-47
Table 9-8 Cost estimates for selected U.S. feasibility evaluation sites (Dec 2009$) ............... 9-50
Table 9-9 Installed and planned U.S. tidal energy capacity ..................................................... 9-53
Table 10-1 Ocean wave energy overview ................................................................................ 10-1
Table 10-2 Alaska available wave energy resource breakdown (TWh per year) ................... 10-29
Table 10-3 West Coast available wave energy resources (TWh per year) ............................ 10-29
Table 10-4 Hawaii available wave energy resources by major island (TWh per year) .......... 10-29

xliv

Table 10-5 East Coast available wave energy resources by state (TWh per year) ............... 10-30
Table 10-6 Gulf of Mexico available wave energy resources by state (TWh per year) .......... 10-30
Table 10-7 Percent technically recoverable wave energy by region for capacity packing
density of 10 MW/km under assumptions described in text ........................................... 10-31
Table 10-8 Percent technically recoverable wave energy by region for capacity packing
density of 15 MW/km under assumptions described in text ........................................... 10-31
Table 10-9 Percent technically recoverable wave energy by region for capacity packing
density of 20 MW/km under assumptions described in text ........................................... 10-32
Table 10-10 Wave energy conversion device developers as of November 2011 .................. 10-38
Table 10-11 FERC active preliminary permits ....................................................................... 10-47
Table 10-12 FERC, MMS, and state lands permitting, licensing and leasing framework ...... 10-47
Table 10-13 WEC device areal footprints .............................................................................. 10-48
Table 10-14 Cost and performance estimates for linear absorber (Pelamis) wave power
plants (December 2009 dollars) ..................................................................................... 10-52
Table 10-15 Typical cost, performance and economic profiles for wave energy plants
(2011$) ........................................................................................................................... 10-54
Table 11-1 Overview of river in-stream energy ........................................................................ 11-1
Table 11-2 Manning coefficients for representative river substrates ....................................... 11-4
Table 11-3 In-stream river energy conversion device developers ......................................... 11-23
Table 11-4 Cost and performance estimates for Alaska in-stream river power plants........... 11-40
Table 11-5 RISEC energy cost, performance and economic profiles (2011$) ....................... 11-41
Table 11-6 Cost estimates for three feasibility evaluation sites ............................................. 11-44
Table 11-7 Major environmental issues and mitigation recommendations ............................ 11-45
Table 11-8 Installed and planned U.S. river in-stream power capacity (MW) ........................ 11-46
Table 12-1 Functions and services provided by generation .................................................... 12-3
Table 12-2 Comparison of output controllability for various generation technologies .............. 12-9
Table 12-3 Comparison of non-thermal renewable generation technologies ........................ 12-12
Table 12-4 Distributed power system performance expectations at various connection
points in the electric system ........................................................................................... 12-21
Table 12-5 Grid penetration scenarios and changing role of distribution generation ............. 12-29
Table 12-6 Energy storage characteristics by application (kilowatt-scale)............................. 12-39
Table 13-1 Lifetimes in the atmosphere and relative infrared absorption strengths of the
greenhouse gases ............................................................................................................ 13-1
Table 13-2 Fossil carbon intensity of coal, oil, natural gas, and wood fuels ............................ 13-2
Table 13-3 Carbon intensity of generation technologies .......................................................... 13-3
Table 13-4 Wind plant performance and cost vs. location (4th Quarter 2010$)....................... 13-8
Table 13-5 Levelized cost of electricity for 50-MW biomass-fired stoker and fluidized bed
boiler power plants ($/MWh, 3rd Quarter 2010$) ........................................................... 13-10
Table 13-6 Performance and cost estimates for 10-MW solar photovoltaic power plants
(4th Quarter 2010$) ........................................................................................................ 13-12

xlv

Table 13-7 Levelized cost of electricity for 50-MW flash-steam and binary-cycle
geothermal power plants (December 2010$) ................................................................. 13-14
Table 13-8 Constant-dollar levelized cost of electricity for solar thermal technologies
(December 2010$) ......................................................................................................... 13-16

xlvi

INTRODUCTION

1.1 Background
Renewable energy technologies are those that utilize inexhaustible or naturally replenished
resources to generate electricity with minimal environmental impact. Energy sources addressed
in this update of the Renewable Energy Technology Guide (RETG) include wind, biomass,
municipal solid waste (MSW), solar photovoltaic (PV), geothermal, solar thermal, ocean tidal,
wave energy, and river in-stream energy conversion (RISEC).
Passage of the Public Utilities Regulatory Policies Act (PURPA) in 1978 was a turning point for
renewable technologies in the United States. PURPA was drafted in response to the oil crisis of
the early 1970s and guaranteed a market at some price for any energy an alternative provider
could produce. Solar and wind power, in particular, made great strides during this period.
However, to some extent, lower oil prices during the 1980s and 1990s lessened interest in,
development of, and government support for renewable technologies.
In the past decade, concerns about the long-term impacts of carbon dioxide and other greenhouse
gases have reinvigorated the industry. The 1992 United Nations Framework Convention of
Climate Change (UNFCCC) set an overall framework for intergovernmental efforts to tackle the
challenges posed by global climate change and was ratified by 192 countries. The 1997 Kyoto
Protocol was a critical driver of subsequent funding for renewable energy projects in the
European Union and, to a lesser degree, elsewhere. As a result, Germany, Japan, the United
Kingdom, and other European Union (EU) countries have set aggressive goals for meeting
increasingly large percentages of their energy needs with renewable energy technologies. In
January 2008, the EU adopted country-specific goals for achieving overall 20% renewable
energy usage by the year 2020. New agreements negotiated in Durban, South Africa in
December 2011, known as the Durban Platform, appear poised to continue and extend efforts
to curb greenhouse gas emissions among both developed and developing nations. The 2012
United Nations Climate Change Conference was held at the Qatar National Convention Center in
Doha. During that meeting, the Kyoto Protocol, which was due to expire at the end of 2012, was
extended to 2020. In addition, the engaged countries agreed to continue the work established at
the Durban Platform and begin developing an agreement to continue to carry on the work of the
Kyoto Protocol beyond 2020. The Durban Platform will be developed by 2015 and executed in
2020, following the expiration of the Kyoto Protocol.
Meanwhile, many individual utilities, municipalities, states, nations and other policymaking
bodies have established renewable portfolio standards (RPSs) that set renewable energy goals,
generally requiring a utility to obtain a certain percentage of its generating capacity from
renewable resources by a certain date. As of November 2012, 38 U.S. states plus the District of
Columbia had RPS mandates or voluntary goals. Other tools employed to encourage and support
the deployment of renewables include the investment tax credit (ITC), the production tax credit
1-1

Introduction

(PTC), and feed-in tariffs (FITs). As renewable energy technologies evolve and become more
cost-competitive, and their environmental benefits become more valuable, they are being
deployed with increasing frequency and public support.
Figures 1-1 and 1-2 show the relative contribution of renewable energy to total primary energy
consumption in the United States in 2011, based on data reported by the DOE Energy
Information Administration (EIA) [1]. In 2011, renewable energy contributed just 9% of the total
energy consumption. Figure 1-3 shows the long-term trends for non-hydro renewable generation,
which in particular highlights the dramatic increase in wind power deployments in the past
decade, thanks in part to subsidies and decreasing capital costs.

Figure 1-1
Renewable energy as share of U.S. total primary energy consumption, 2011
Source: EIA, September 2012

1-2

Introduction

Figure 1-2
U.S. renewable energy consumption by source, 2011
Source: EIA

Figure 1-3
U.S. non-hydroelectric renewable power sources, 19892011
Source: EIA

Renewable energy resources are broadly accessible throughout the world. Some resources are
almost universally available; others are limited to particular areas. Locations with high annual
insolation are obviously best suited to solar power. Solar energy can still be usefully and
economically applied in locations with lower quality resources including, in some cases, extreme
northern or southern latitudes. Wind energy is generally deployed along coastlines or mountain
passes with reliable sustained winds, although the central plains of North America host some of
1-3

Introduction

the worlds best wind resources and many of the newest and largest wind power projects. In
addition, the Midwest is an excellent source of biomass, which can be grown through managed
agricultural programs to provide a continuous supply of biomass fuels. Geothermal resources are
concentrated in geologically active areas such as along the Pacific Ring of Fire, which includes
the Philippines, Indonesia, and California. Hydroelectric resources exist in most countries.
The benefits of using renewable energy are many and extend beyond issues of abundance and
environmental impact. Diversifying energy resources provides security against economic or
political events that may impact one particular resource. They offer some protection against risks
associated with fluctuating fossil-fuel prices and supplies. Investment in renewable energy can
contribute to local economic growth and employment. Many renewable energy plants can be
built in a modular fashion proportionate to load growth patterns and local needs.
The value of renewable energy is determined by a number of factors, including its ability to meet
mandated renewable energy capacity requirements, respond to customer demands for renewable
energy, and differentiate a power provider from its competitors. Renewables can also offer
significant environmental benefits such as offsetting carbon dioxide, NOx, and SO2 emissions.
Some renewable energy technologies such as biomass and geothermal (and, to a lesser extent,
some solar thermal technologies) are dispatchable, making them valuable generation assets. The
output of non-dispatchable technologies such as solar PV can correspond well with summer
afternoon load peaks. The Electric Power Research Institutes (EPRIs) ongoing efforts to
develop accurate forecasting models will improve the value of wind energy in a companys
generation portfolio.

1.2 EPRIs Role


Since the 1970s, EPRI has published periodic updates to its Technical Assessment Guide (TAG),1
which provides the most up-to-date technical and economic data available on a wide range of
generation technologies. The TAG has become the recognized industry standard for information
essential in preliminary planning for generation technology evaluation and application.
In its earliest editions, the TAG also included renewable energy technologies. However, in 1999
EPRI concluded that the field of renewable energy had grown sufficiently, and was developing
and changing so rapidly, that it would be worthwhile to prepare a dedicated Renewable Energy
Technical Assessment Guide (TAG-RE) and update it annually. Beginning with the 2009 edition,
the TAG-RE was renamed the Renewable Energy Technology Guide (RETG).

1.3 Objective
This RETG, updated for 2012, fulfills two roles. First, it serves as a one-stop information source
for a wide variety of renewable energy technologies evaluated on a consistent basis. EPRIs
experience and expertise, along with its reputation for objectivity and credibility, make the
RETG a uniquely valuable reference. The various renewable technologies have been evaluated
with sufficient technical depth to enable planners to conduct an accurate assessment of the
renewable energy options that are available to them.

TAG is a registered trademark of the Electric Power Research Institute.

1-4

Introduction

Second, the RETG provides information to educate all interested parties in the key technology
concepts, language, and issues related to renewables. It offers not only technical depth but also
breadth of coverage across many disciplines including project planning, resource availability and
management, regulatory processes, operating and maintenance requirements, market potential,
and future developments.
The information summarized in the RETG encompasses nearly four decades of EPRI research on
renewable technologies and draws upon a large foundation of knowledge built by the DOE and
other national and international experts. This information has been distilled and synthesized to
provide the readers of RETG the most up-to-date, accurate, practical intelligence available that
they can use to investigate, plan, and deploy the latest renewable energy technologies.

1.4 Definitions and Units


Unless otherwise stated, all dollar figures cited in the RETG are assumed to be U.S. dollars
corrected for inflation to end-of-year 2012 values.
Throughout the RETG, most notably the Biomass and Solar Thermal chapters, the subscripts e
(e.g., kWe or MWe) and t (e.g., kWt or MWt) are used to differentiate between electrical and
thermal power, respectively. In the Solar Photovoltaic chapter, the terms kWp or MWp may be
used to denote peak PV cell or module capacity.

1.5 Scope
The RETG consists of 13 chapters. Following this Chapter 1: Introduction, they are:
Chapter 2: Economic Methodology and Assumptions, which provides a detailed discussion of
the basis upon which data were gathered and presented.
Chapter 3: Wind Power, which covers one of the most widespread and economically
competitive renewable energy technologies. Wind power systems progressed substantially
through the past decades as a result of government incentives, with a steady trend of cost
reduction. Wind power is particularly popular in Europe, due partly to aggressive government
goals and public support, and is considered to be a commercially established and competitive
grid-power technology.
Chapter 4: Biomass Electricity Generation, which focuses on biomass fuel and combustion
properties, direct combustion, gasification, and cofiring with coal in utility boilers. This chapter
addresses resource availability, technology status, performance and cost history and projections,
operating and maintenance labor requirements, environmental emissions, installed capacity, and
cost and economic issues.
Chapter 5: Municipal Solid Waste, outlines some of the current and future technologies in the
waste-to-energy (WTE) sector including performance, installation history, emissions and in some
cases and cost and economic issues.
Chapter 6: Solar Photovoltaics (PV), which addresses power systems that convert sunlight
directly into electricity. These solid-state electronic devices have no moving parts, no fluids, no
noise, and no emissions of any kind. These benefits have positioned PV to be the preferred power
technology for many remote applications. In addition, developments on the horizon promise to
significantly reduce the costs of PV and open up new markets for distributed applications, such as
innovative building-integrated technologies, and increasingly, MW-scale projects.
1-5

Introduction

Chapter 7: Geothermal Energy, which provides the latest information on systems that generate
electricity by tapping underground steam reservoirs. Although available in relatively few areas,
geothermal power has been a practical reality in California, Italy, and Asia for decades. This
chapter also addresses the potential in both developed and developing countries for the more
common geothermal hot water and liquid-dominated hydrothermal resources.
Chapter 8: Solar Thermal, which addresses systems that use concentrated sunlight to heat a
working fluid and generate electricity in a thermodynamic cycle. Such systems range in scale
from 25-kW reflector dish-generator systems to 85-MW solar thermal trough plants that
concentrate solar energy using parabolic trough reflectors to heat oil, which is then used to
generate steam and electricity. Another technology option is central receiver technology, which
uses hundreds of mirrors to focus sunlight onto a central receiver. The 10-MW Solar Two
demonstration incorporated thermal storage through the use of a molten salt heat transfer fluid
and a large storage tanks.
Chapter 9: Ocean Tidal Energy, which addresses low-impact and emerging hydro systems that
capture ocean tidal current energy to generate electricity. These systems are distinct from
traditional hydroelectric installations, which typically generate electricity via a turbine powered
by water impounded behind a dam. Many ocean energy generation technologies are derived from
analogous technologies in the wind turbine industry.
Chapter 10: Ocean Wave Energy, which addresses low-impact and emerging hydro systems
that convert the kinetic energy of ocean waves into electricity.
Chapter 11: River In-Stream Energy, which reviews the potential for generating electrical
power from river energy. Technology and developer companies are working to modify tidal
energy turbines for use in river environments.
Chapter 12: Grid Integration Challenges and Technologies, which addresses power
electronics, energy storage, and other technologies need to integrate large wind plants and other
intermittent generation technologies into the electricity grid. Energy storage technologies such as
batteries, pumped hydro, compressed air energy storage (CAES), flywheels, and superconducting
magnetic energy storage (SMES) can be selected and designed to absorb short, intermediate, and
longer-term fluctuations of output from seconds to days. Other integration technologies, such as
line compensation, power electronics, integration with hydro, and solar and wind energy
forecasting, address short-term fluctuations.
Chapter 13: Greenhouse Gas Emissions Control, which discusses the role that renewable
energy power technologies can play in greenhouse gas emissions reduction and includes example
calculations showing how that role can be quantified.

1.6 References
1. Annual Energy Review 2011 U.S. Energy Information Administration, Washington, D.C.:
September 2012: DOE/EIA 0384 (2011).
2. Renewable Energy Technology Guide: 2011. EPRI, Palo Alto, CA: 2011. 1021795.
3. Renewable Energy Technology Guide: 2010. EPRI, Palo Alto, CA: 2010. 1019760.
4. Renewable Energy Technology GuideRETG 2009. EPRI, Palo Alto, CA: 2010. 1021379.

1-6

Introduction

5. Renewable Energy Technology GuideRETG: April 2009 Update. EPRI, Palo Alto, CA:
2009. 1019300.
6. Renewable Energy Technical Assessment Guide TAG-RE 2008. EPRI, Palo Alto, CA:
2008. 1015801.
7. Renewable Energy Technical Assessment Guide TAG-RE 2007. EPRI, Palo Alto, CA:
2008. 1014182.
8. Renewable Energy Technical Assessment Guide TAG-RE 2006. EPRI, Palo Alto, CA:
2007. 1012722.
9. Renewable Energy Technical Assessment Guide TAG-RE 2005. EPRI, Palo Alto, CA:
2005. 1010407.
10. Renewable Energy Technical Assessment Guide TAG-RE 2004. EPRI, Palo Alto, CA:
2004. 1008366.
11. Renewable Energy Technical Assessment Guide TAG-RE 2003. EPRI, Palo Alto, CA:
2003. 1004938.
12. Renewable Energy Technical Assessment Guide TAG-RE 2002. EPRI, Palo Alto, CA:
2002. 1004196.
13. Renewable Energy Technical Assessment Guide TAG-RE 2001. EPRI, Palo Alto, CA:
2001. 1004034.
14. Renewable Energy Technical Assessment GuideTAG-RE 2000. EPRI, Palo Alto, CA: 2001.
1000574.
15. Renewable Energy Technology Characterizations. EPRI, Palo Alto, CA, and the Office of
Utility Technologies, DOE, Washington, D.C.: 1997. TR-109496.
16. Renewables Global Status Report. REN21. Paris, France.

1-7

ECONOMIC METHODOLOGY AND ASSUMPTIONS


The economic methodology and assumptions presented here provide a uniform framework to
evaluate the economic feasibility of alternative renewable energy projects on a consistent basis.
Because both regulated utility and unregulated non-utility generators are developing renewable
energy projects, this section presents separate methodologies to apply depending on the
developer type.
The methodology descriptions presented below are excerpted from the 1999 Technical
Assessment Guide, Volume 3, Revision 8, Fundamentals and MethodsElectricity Supply [1].
The methodology is illustrated by a 200-MW wind project example using cost and performance
data from a 2009 EPRI report [2].

2.1 Regulated Utility Power Projects


The total annual cost of a regulated utility power project consists of the annual fixed charge on
the capital investment plus the operation and maintenance (O&M) costs. The latter costs include
operation and maintenance labor and material, fuel, raw materials and utilities, waste disposal
fees, and other costs.
2.1.1 Fixed Charge Rates
A very useful concept is the capital fixed charge rate or fixed charge rate. This rate represents
the annual charges customers would have to pay each year so that the utility recovers its capitalrelated revenue requirements, and is expressed as a percentage of the booked cost of a plant.
Fixed charge rates are typically measured in one of the following three ways:

Annual fixed charge rate

Nominal levelized fixed charge rate

Real levelized fixed charge rate

Subsections 2.1.2 through 2.1.4 address these concepts and their derivation as well as the
components and calculation of annual fixed charge rates [1].
2.1.1.1 Annual Fixed Charge Rates
Annual fixed charge rates express the annual capital revenue requirements as a percentage of the
booked cost. In Table 2-1, for example, the total present value of the annual capital charge or
booked cost is $604.27 million, and the annual capital revenue requirements based on that
booked cost are shown in column 1. The annual fixed charge rates are the ratio of the annual
capital revenue requirements to the booked cost of $604.27 million, and are shown in column 2.
The annual fixed charge rates decline over time as the annual capital revenue requirements
decline. So, for example, in the first year, the annual fixed charge rate is 19.09% and declines to
5.32% by the end of the book life. The assumed debt/equity ratio in Table 2-1 is 1.5.
2-1

Economic Methodology and Assumptions


Table 2-1
Calculation of annual fixed charge rates

End of Year

Nominal Annual
Fixed Charge Rates
(Percentage of Booked Cost)

(1)

(2)

2011

$115.34

19.09%

2012

112.38

18.60%

2013

108.34

17.93%

2014

104.46

17.29%

2015

100.74

16.67%

2016

97.16

16.08%

2017

93.72

15.51%

2018

90.41

14.96%

2019

87.20

14.43%

2020

84.02

13.90%

2021

80.83

13.38%

2022

77.65

12.85%

2023

74.46

12.32%

2024

71.28

11.80%

2025

68.09

11.27%

2026

64.91

10.74%

2027

61.72

10.21%

2028

58.54

9.69%

2029

55.35

9.16%

2030

52.17

8.63%

2031

48.98

8.11%

2032

46.49

7.69%

2033

44.70

7.40%

2034

42.90

7.10%

2035

41.11

6.80%

2036

39.32

6.51%

2037

37.52

6.21%

2038

35.73

5.91%

2039

33.94

5.62%

2040

32.14

5.32%

Booked Cost
($ millions)

2-2

Capital Revenue
Requirements
($ Million, Nominal)

$604.27

Economic Methodology and Assumptions

2.1.1.2 Nominal Levelized Annual Charges and Nominal Levelized Fixed Charge Rates
The second type of fixed charge is the nominal levelized fixed charge rate, which translates
booked costs into a constant annual nominal dollar charge with the same present value as the
actual annual capital revenue requirements. Calculating the nominal levelized annual charge can
be thought of as a two-step process (see Example 2-1).
Example 2-1
The first step is to calculate the present value of the annual capital revenue requirements.
Column 1 of Table 2-2 shows the annual capital revenue requirements in Table 2-1. Because the
annual revenue requirements are in nominal terms, they must be discounted using the nominal
discount rate. Column 2 shows the annual present value factors used to bring the costs from endof-year to present value. Column 3 lists the real or constant-dollar revenue requirements that
have inflation factored out of the values (e.g., $115.34/(1.03) = $111.98 for the year 2011).
Column 4 lists the annual present value factors based on the real discount rate.
Once the present value has been calculated, the levelized annual charges and the fixed charge
rates can be derived as noted in the table.
It is important to note that the levelized fixed charge rate depends on the capital structure and
cost of money, tax depreciation, construction length, book life, and taxes.
2.1.1.3 Real Levelized Annual Charges and Real Levelized Fixed Charge Rates
The third type of fixed charge rate is real levelized fixed charge. It is the fraction of the booked
cost, expressed in constant dollar terms, that customers would have to pay annually over the
book life of a plant to cover capital revenue requirements. The real levelized annual charge and
fixed charge rate are calculated the same way as their nominal counterparts.
However, real levelized fixed charges have an important advantage over their nominal
counterparts: they do not reflect general inflation. The nominal levelized charge reflects
assumptions about future inflation. If a power contract is specified in nominal terms, there are
risks associated with expected inflation. If inflation is less than expected, the buyer will pay too
much and the seller will earn greater-than-anticipated profits. The opposite would be true if
inflation were greater than anticipated. On the other hand, if a contract is specified in real terms,
the risks to the buyer and seller can be avoided. The real levelized charge can be adjusted for
actual inflation over the term of the contract.
How are levelized fixed charge rates used? Generation alternatives are often compared on the
basis of levelized costs. With the appropriate levelized fixed charge rate, the booked cost of
alternatives can be converted to levelized annual cost terms. The EPRI financial model computes
levelized fixed charge rates, which can be used to evaluate generating and other utility
investment options.

2-3

Economic Methodology and Assumptions


Table 2-2
Calculating levelized annual charges
Nominal Revenue Requirements Real Revenue Requirements
Capital Revenue
Capital Revenue
End of
Present Value
Present Value
Requirements
Requirements
Year
Factors
Factors
($ Million)
($ Million)
(1)
(2)
(3)
(4)
2011
$115.34
0.9193
$111.98
0.9469
2012
112.38
0.8451
105.93
0.8966
2013
108.34
0.7769
99.14
0.8490
2014
104.46
0.7142
92.81
0.8039
2015
100.74
0.6566
86.90
0.7612
2016
97.16
0.6036
81.37
0.7208
2017
93.72
0.5549
76.21
0.6825
2018
90.41
0.5101
71.37
0.6462
2019
87.20
0.4690
66.83
0.6119
2020
84.02
0.4311
62.52
0.5794
2021
80.83
0.3964
58.39
0.5486
2022
77.65
0.3644
54.46
0.5195
2023
74.46
0.3350
50.70
0.4919
2024
71.28
0.3079
47.12
0.4658
2025
68.09
0.2831
43.71
0.4411
2026
64.91
0.2603
40.45
0.4176
2027
61.72
0.2393
37.34
0.3954
2028
58.54
0.2199
34.38
0.3744
2029
55.35
0.2022
31.57
0.3546
2030
52.17
0.1859
28.88
0.3357
2031
48.98
0.1709
26.33
0.3179
2032
46.49
0.1571
24.26
0.3010
2033
44.70
0.1444
22.65
0.2850
2034
42.90
0.1328
21.11
0.2699
2035
41.11
0.1221
19.63
0.2556
2036
39.32
0.1122
18.23
0.2420
2037
37.52
0.1032
16.89
0.2291
2038
35.73
0.0948
15.62
0.2170
2039
33.94
0.0872
14.40
0.2054
2040
32.14
0.0801
13.24
0.1945
($ million)
$901.81
$901.81

Present value of annual


revenue requirements1
Escalation rate
Discount rate2
Capital recovery factor notation
Capital recovery factor value
Levelized annual charges3
($ million)
Booked cost
($ million)
Levelized annual fixed
charge rates4, 5

3.00%
8.78%
(A/P, 8.78%, 30)
0.0954
86.05
$604.27
14.24%

3.00%
5.61%
(A/P, 5.61%, 30)
0.0696
62.80
$604.27
10.39%

Notes:
1. The present value is at the beginning of 2010 and results from the sum of the products of the annual present value
factors times the annual requirements.
2. The discount rate for real-dollar analysis has inflation factored out: 1.0878/1.03-1.
3. The levelized annual charges (end-of-year) result from the present value times the capital recovery (A/P, i, n) factor.
4. The levelized annual fixed charge rate results from the levelized annual charges divided by the booked cost.
5. The real levelized annual charge is an end-of-year payment expressed in beginning-of-year 2008 real dollars.

2-4

Economic Methodology and Assumptions

2.1.2 Example of Annual Fixed Charge Rates for a 200-MW Wind Power Plant
To illustrate how plant characteristics affect annual fixed charges, the 1999 TAG report presents
several examples, including transmission; distribution; and renewable, nuclear, combined-cycle
combustion turbine, simple-cycle combustion turbine, and fossil fuel generation. The examples
differ with respect to construction period, tax life, and book life.
Tables 2-3 and 2-4 address the example of a 200-MW wind power plant. The tables present the
capital cost of the plant, the construction period, book life, tax life, and, based on these data, the
annual, cumulative, and levelized fixed charges. The fixed charges in Table 2-3 are expressed in
nominal terms, whereas those in Table 2-4 are expressed in constant-dollar terms.
2.1.3 Other Capital ExpensesCapital Additions
Capital additions are investments in a plant after it is in service. Because they are treated as
capital investments, the above discussion of capital-related revenue requirements applies to
capital additions as well. The annual revenue requirements associated with specific capital
additions depend on relevant book and tax depreciation schedules. For many analyses, it is
sufficient to treat them as expenses, in which case customers are assumed to pay for them on an
as-you-go basis. For the same level of capital additions, expensing them yields a smaller present
value of revenue requirements than capitalizing them; capitalizing would include a tax
component whereas expensing would not.
2.1.4 Calculating Costs per Kilowatt-Hour
Calculating costs on a per-kilowatt-hour basis provides a relatively easy way to compare new
generating alternatives and their costs with those of one or more existing resources.
2.1.4.1 Levelized Costs per Kilowatt-Hour
Comparing costs on a levelized or a present-value-of-revenue-requirements basis is appropriate
under the following conditions:

When the benefits or revenues are the same between alternatives. For example, if the
revenues to a utility will be the same for alternatives that are all designed to produce
300 MWnet. Therefore, the most economical plant can be selected by minimizing the
present value of revenue requirements.

When alternatives have different benefits but the benefits remain the same for each year. For
example, a 300-MWnet plant can be compared with a 200-MWnet plant by comparing ratios of
levelized cost per kilowatt-hour (often referred to as levelized busbar costs).
Table 2-3
Wind power generationnominal dollar terms
Instantaneous Construction Cost ($/kW at midyear of first year of construction)
Total plant cost1
2,045
Allowance for funds used during construction
60
(AFUDC;interest during construction)
Capital Cost ($/kW at commercial operation date)
Total plant investment
2,105
Due diligence, permitting, legal, development
150
Total booked cost (total capital requirement)
2,255

2-5

Economic Methodology and Assumptions


Table 2-3 (continued)
Wind power generationnominal dollar terms
Time Frames and Economic Factors
Plant construction period (years)
Tax recovery period (years)
Book life (years)

1
5
30

Apparent escalation rate during construction2


(%/year)
AFUDC rate, nominal (%/year)

6.00

Discount rates (%/year)


Nominal
Real

8.78
5.61

3.00

Fixed Charges
End of Year

Annual

Cumulative
Present Value

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30

0.1920
0.1788
0.1595
0.1467
0.1378
0.1289
0.1230
0.1201
0.1171
0.1142
0.1113
0.1083
0.1054
0.1024
0.0995
0.0965
0.0936
0.0906
0.0877
0.0847
0.0818
0.0788
0.0759
0.0729
0.0700
0.0671
0.0641
0.0612
0.0582
0.0553

0.1765
0.3277
0.4515
0.5563
0.6468
0.7246
0.7929
0.8541
0.9091
0.9583
1.0024
1.0419
1.0772
1.1087
1.1369
1.1620
1.1844
1.2043
1.2220
1.2378
1.2517
1.2641
1.2751
1.2848
1.2933
1.3008
1.3075
1.3133
1.3183
1.3228

Levelized3
0.1920
0.1857
0.1777
0.1709
0.1653
0.1605
0.1564
0.1530
0.1503
0.1479
0.1458
0.1439
0.1422
0.1406
0.1392
0.1379
0.1366
0.1355
0.1344
0.1334
0.1325
0.1316
0.1308
0.1300
0.1293
0.1286
0.1280
0.1273
0.1268
0.1262

Notes:
1. The total plant cost figure used in this illustrative example is likely low for end-of-2008 wind energy
installation costs.
2. The apparent escalation rate represents the total annual rate of change of capital costs during construction.
For this example, the inflation rate equals 3% and the real escalation rate equals 0%.
3. Levelized over the period from the first year through the current year. For example, the levelized fixed
charge rate for a 25-year book life is 0.0974; for a 30-year book life, it is 0.0921. Values where the book
life is less than the tax life should be considered as approximations only.

2-6

Economic Methodology and Assumptions


Table 2-4
Wind power generationconstant dollar terms
Instantaneous Construction Cost ($/kW at mid-year of first year of construction)
Total plant cost1
2,045
AFUDC (interest during construction)
60
Capital Cost ($/kW at commercial operation date)
Total plant investment
2,105
Due diligence, permitting, legal, development
150
Total booked cost (total capital requirement)
2,255
Time Frames and Economic Factors
Plant construction period (years)
1
Tax recovery period (years)
5
Book life (years)
30
Apparent escalation rate during construction2
3.00
(%/year)
AFUDC rate, nominal (%/year)
6.00
Discount Rates (%/year)
Nominal
8.78
Real
5.61
Fixed Charges
End of Year
Annual
Cumulative Present Value

Levelized3

0.1920

0.1765

0.1864

0.1736

0.3277

0.1777

0.1503

0.4515

0.1677

0.1342

0.5563

0.1591

0.1224

0.6468

0.1519

0.1112

0.7246

0.1455

0.1030

0.7929

0.1401

0.0976

0.8541

0.1354

0.0925

0.9091

0.1314

10

0.0875

0.9583

0.1278

11

0.0828

1.0024

0.1246

12

0.0782

1.0419

0.1216

13

0.0739

1.0772

0.1189

14

0.0697

1.1087

0.1164

15

0.0658

1.1369

0.1141

16

0.0620

1.1620

0.1119

17

0.0583

1.1844

0.1099

18

0.0548

1.2043

0.1080

19

0.0515

1.2220

0.1062

20

0.0483

1.2378

0.1045

21

0.0453

1.2517

0.1029

22

0.0424

1.2641

0.1014

23

0.0396

1.2751

0.1000

24

0.0370

1.2848

0.0987

25

0.0344

1.2933

0.0974

26

0.0320

1.3008

0.0963

27

0.0297

1.3075

0.0951

28

0.0275

1.3133

0.0941

29

0.0254

1.3183

0.0931

30

0.0235

1.3228

0.0921

Notes:
1. The total plant cost figure used in above illustrative example is likely low for end-of-2008 wind energy
installation costs.

2-7

Economic Methodology and Assumptions


2.
3.

The apparent escalation rate represents the total annual rate of change of capital costs during construction.
For this example, the inflation rate equals 3% and the real escalation rate equals 0%.
Levelized over the period from the first year through the current year. For example, the levelized fixed
charge rate for a 25-year book life is 0.0974; for a 30-year book life, it is 0.0921. Values where the book
life is less than the tax life should be considered as approximations only.

Levelized busbar costs are calculated by the following steps:


1. Calculate the present value of revenue requirements.
2. Calculate the levelized annual revenue requirements by multiplying the present value by the
capital recovery or (A/P,i,n) factor.
3. Calculate the levelized busbar cost by dividing the levelized annual cost by the uniform
annual generation in kilowatt-hours.
Frequently, alternatives need to be compared in which the amount of generation changes each
year. Sometimes analysts calculate alternative measures that may present analytical problems,
such as the following:

Levelizing the annual cost of generation. The annual cost of generation can be calculated by
dividing the annual revenue requirement by the annual generation. Levelizing these rates
(calculating a present value of the ratios and multiplying the result by an A/P factor) may
lead to mathematical problems because it does not properly account for the amount of
generation in each year.

Dividing the present value of revenue requirements by the present value of kilowatt-hours.
This can result in erroneous results under certain mathematical conditions.

For the case in which annual generation varies, other analytical methods must be used. These
methods include purchasing replacement power so that the generation remains constant over the
years, applying more sophisticated approaches included in production cost models, or
performing a net present value (or profitability) analysis that discretely evaluates the variations
in revenue.

2.2 Non-Utility Generator Power Projects


Electricity from non-utility generators (NUGs) has played an increasingly important role in the
electric utility industry since the passage of the Public Utility Regulatory Policies Act of 1978
(PURPA). As electric utility industry restructuring continues, the role of non-utility generation is
likely to increase. The Energy Policy Act of 2005 modified some of PURPAs mandatory purchase
requirements that obligate utilities to purchase power from qualifying small power facilities.
One of the key goals of restructuring was to shift risk from electricity customers to both utility
and non-utility power producers. Under the current regulatory structure, customers bear the
risks associated with and benefit from prudent utility investment decisions. Even though they are
giving up the potential for higher-than-expected profits as well as the potential for losses, electric
utilities are guaranteed a return on their prudent investments. With restructuring, the risk
associated with the generation portion of electricity production has shifted somewhat to power
producers. The extent to which risks are shifted depends on the specifics of industry
restructuring, which varies among states and regulatory jurisdictions.
2-8

Economic Methodology and Assumptions

Merchant plants are generally characterized as those that have some commodity risk for
electricity sales (i.e., the sale of electricity is not fully committed to long-term power sales
agreements). However, most renewable energy power plants being built remain either utilityowned or, more commonly, privately-owned, with their output pre-sold to utilities at
contractually agreed-upon rates.
The following are key differences between utility and non-utility power producers:

Obligation to serve: Electric utilities have traditionally had an obligation to serve and to
provide reliable electricity service. However, the obligation to serve is becoming less clear as
the industry restructures the regulation of generating and distribution companies. Non-utility
power producers develop projects for their potential economic rewards and have the option to
sell their power on a wholesale basis to a utility, on a retail basis to a customer, or directly to
a power pool.

Rates/prices: Rates for utilities are usually set using the revenue requirements approach.
Non-utility power producers typically attempt to set prices as high as the market will allow.

Risks and benefits: Customers of electric utilities bear the risks associated with prudent
investments. Because customers, not utilities, bear the risk, utilities earn a lower return on
investments associated with a monopoly. Non-utility power producers, or utilities providing
generation in competitive power markets, bear the risks associated with their investments but
can mitigate them to the extent they can negotiate contracts for fuel purchases and energy
sales.

The following subsections focus on the economic evaluation of non-utility or independent power
projects. Subsection 2.2.1 provides a brief discussion of the various types of non-utility power
producer projects.
Subsection 2.2.2 describes the development of an economic pro forma for an example
merchant plant.
Subsection 2.2.3 summarizes a sensitivity analysis for the example merchant plant pro forma and
discusses the major risks associated with financing merchant plants.
Subsection 2.2.4 provides a statement of caution on the limitations of the examples that are
discussed.
2.2.1 Types of Non-Utility Generators
An understanding of the economic evaluation methods for non-utility projects requires an
understanding of the potential economic rewards. Non-utility projects can be classified into the
following types:

PURPA cogenerators and small power producers

Wholesale generators exempt from the Energy Policy Act of 1992

Merchant power plants

2-9

Economic Methodology and Assumptions

2.2.1.1 PURPA Cogenerators and Small Power Producers


Title II of PURPA created a class of power producers called qualifying facilities (QFs) for which
utility ownership is limited to 50%. Qualifying facilitieswhich can include cogenerators and
small power producersare exempt from certain state regulations, Federal Energy Regulatory
Commission (FERC) regulations, and the recently repealed Public Utility Holding Company Act
(PUHCA). However, the Energy Policy Act of 2005 (EPAct05) amended provisions within
PURPA that affect cogeneration and small power production facilities, including eliminating the
ownership limitation on utilities.
Cogenerators produce two forms of energy: electricity and steam (or heat). Two types of
cogenerators have emerged. The first type comprises facilities designed primarily to meet
the thermal requirements of an industrial host, with electricity generation being a secondary
consideration. The second type has become known as PURPA machines, which are designed
primarily for the purpose of selling electricity to electric utilities while satisfying the thermal
requirements under Title II of PURPA.
In 2006, FERC, in compliance with EPAct05, issued a decision that did not significantly change
the definition of a qualifying facility. FERC decided not to dramatically tighten standards,
impose new and higher efficiency rules, or restrict access to QF status. Instead, the Commission
stated that combined heat and power (CHP) cogeneration continues to be intrinsically more
efficient than separate heat and power production, that current efficiency standards are
appropriate, that case-by-case evaluation is called for rather than any one-size-fits-all numeric
rule, that smaller cogenerators in particular present no risk of being sham projects to foist
overpriced power onto utilities, and that cogenerators should be allowed to continue selfcertifying as QFs. Small power producers continue to be qualifying facilities that use biomass,
waste, geothermal resources or renewable resources such as wind, solar energy, or water. There
are other non-utility generating units in addition to QFs, but they are regulated under the Federal
Power Act.
Power sales for qualifying facilities were typically based on a utilitys avoided cost rates.
Avoided cost rates were generally specified as the cost of a utilitys next plant to be built. By
purchasing power from a non-utility generator, the utility typically paid the non-utility generator
the avoided cost for not building this plant. Consequently, ratepayers would be indifferent as to
who supplied the power. Financing a qualifying facility was relatively straightforward because
the credit worthiness of the project was tied closely with that of the utility (or power purchaser).
Power sales usually consisted of a capacity charge and an energy charge. The capacity charge
frequently comprised a debt service portion that was not subject to escalation and a fixed
operating and maintenance cost that may have been subject to inflation over time. The energy
charge was typically based on the cost of fuel and variable operating costs and was subject to a
negotiated inflation or escalation rate over time.
The economic incentive for qualifying facilities was reduced in the late 1980s and early 1990s
when rates for some qualifying facilities were reduced below the local utilitys avoided cost. In
some states a qualifying facilitys rates had to be low enough to win a competitive bid for a
power solicitation.

2-10

Economic Methodology and Assumptions

The economic incentive for QFs was further reduced by competition from exempt wholesale
generators, which resulted from the Energy Policy Act of 1992. Many long-term power purchase
agreements for qualifying facilities have turned out to be significantly more expensive than the
market price of power. There are efforts in some states to restructure or buy out these contracts.
2.2.1.2 Exempt Wholesale Generators
The Energy Policy Act of 1992 (EPAct92) created another group of non-utility facilities called
exempt wholesale generators (EWGs), which was unaffected by EPAct05. The major
characteristics of EWGs include the following:

Ownership and operation is for wholesale delivery of electricity.

Ownership by utilities is allowed to be 100%.

There is no requirement for a thermal host.

Wholesale rate approval is required from FERC.

Exemptions from the Securities and Exchange Commission (SEC) regulations under PUHCA
are allowed.

Other major provisions of the EPAct92 include the following:

FERC is empowered to order transmission access to third parties (such as EWGs).

FERC is prohibited from ordering retail wheeling. However, states may choose to order retail
wheeling, which could further encourage the construction of non-utility generators.

EPAct92 also allowed EWGs to develop large central power stations that have economy-of-scale
advantages over smaller cogeneration plants. In addition, EPAct92 opened the way for states to
consider retail wheeling and the establishment of merchant power plants.
2.2.1.3 Merchant Power Plants
Merchant plants, which have arisen in the United States as electric wholesale competition has
taken hold in many regions, are generally characterized as those that have some commodity risk
for electricity sales (i.e., the sale of electricity is not fully committed to long-term power sales
agreements). The power will be sold either on a spot market basis to the power pool or under
contracts with varying terms to utilities. Pure merchant power more precisely refers to power
sales that are not covered by conventional long-term agreements. The term hybrid merchant
power refers to plants selling power under a combination of conventional power sales agreements
and spot market sales where the power is at risk for prevailing market conditions. For the purposes
of this report, the term merchant power plants will generally refer to plants in which a
substantial portion of the power sales (sometimes referred to as the offtake) is at risk.
Merchant power plants include existing, repowered, and greenfield units. Most currently
operating merchant plants are existing plants that have recently been sold by electric utilities.
These existing plants are either currently economical or represent a site that could be repowered
to be economical (e.g., by adding a new combustion turbine at an existing site).

2-11

Economic Methodology and Assumptions

Greenfield power plants are those that are built at a new site. A large number of greenfield
merchant power plants are being developed in areas where they will be able to produce power
less expensively than existing older generation or in areas where additional power will be
required to meet growing demand.
The economic analysis in this report centers on greenfield merchant power plants, but is broadly
applicable to repowering projects and to other non-utility power plants.
2.2.2 Development of an Economic Pro Forma for a Merchant Plant
The revenue requirements method (as discussed in Section 2.1) is generally not applicable to the
evaluation of non-utility projects. Although there is a variety of methods to evaluate non-utility
power projects, all methods depend on calculating cash flows. The cash flows represent all of the
revenues from the sale of electricity less the sum of all expenses, debt service, and income taxes.
The net cash flows represent cash available to equity holders. The major differences between the
two types of analyses are summarized below and followed by an example pro forma economic
analysis for a non-utility plant.
2.2.2.1 Differences Between Regulated Utility and Non-Utility Power Projects
2.2.2.1.1 Pricing

Regulated utilities are restricted in how they can set prices to recover the costs of investments.
Recovery of the total capital-related costs is determined using the revenue requirements approach
described in Section 5 of the 1999 EPRI TAG report [1]. In particular, the recovery of capitalrelated costs depends on return on equity, interest on debt, book depreciation, income taxes, and
property taxes and insurance. But taxes, in turn, depend on capital-related costs, interest on debt,
tax depreciation schedules, and income tax rates. Thus, capital-related revenues are determined
by the underlying costs.
Non-utility power producers have more flexibility. They are not bound to the revenue
requirements approach for determining how their costs will be recovered. Such flexibility is
necessary for non-utilities to respond to changes in market power prices. Whereas under the
revenue requirements approach, capital-related revenues decline over the life of a project, this
need not be the case for non-utility power projects. The prices of power for non-utility projects
can rise, fall, or remain constant over the contract period depending on the needs of the purchaser
and the internal requirements of the non-utility power producer.
2.2.2.1.2 Income Taxes and Depreciation

There are two types of depreciation for regulated utilities: book and tax (see Section 2.1). Under
the revenue requirements method, book depreciation is used to determine revenues, whereas tax
depreciation is used to calculate income taxes. For non-utility power producers, there is no such
distinction. Tax depreciation is calculated to reduce income taxes, which directly affect cash
flow. However, book depreciation (which may be needed for accounting purposes) does not
affect cash flow. Cash flow drives the economic viability of non-utility projects.

2-12

Economic Methodology and Assumptions

2.2.2.1.3 Capital Expenditures

Revenue requirements reflect the cost of constructing a plant as well as the allowance for funds
used during construction (referred to as interest during construction for non-utility projects).
For non-utility power producers, only the debt portion of construction expenditures is typically
capitalized. The equity portion of the interest during construction is recovered in the net cash
flows and is not capitalized.
2.2.2.1.4 Debt Financing

Debt is a major source of financing for construction projects. Regulated utilities tend to borrow
against their corporate balance sheet. Corporate stability and borrowing capability become
important issues for utility financing.
Non-utilities finance projects through both corporate debt and project finance (as discussed later
in this section). Corporate debt is becoming more common as the industry consolidates to fewer
and larger companies through mergers and acquisitions. Nevertheless, there is a significant role
for limited recourse project finance.
2.2.2.1.5 Economic Methodology

Regulated utilities typically evaluate potential projects using the present worth of revenue
requirements methodology. Financial structure (debt and equity proportions) is implicitly
included in the analysis by using an after-tax discount rate.
Non-utilities typically evaluate potential projects using a discounted net cash flow analysis. For
project finance opportunities, specific financial parameters are included in the costs of the project.
The after-tax net cash flows are then discounted with the owners minimum acceptable rate of return.
2.2.3 Conceptualizing the Analysis
This section illustrates cash flow calculations using the example of a 200-MW merchant wind
plant. General parameters for the example are described, followed by a discussion of income and
cash flow statements. Economic results for evaluating the cash flows are also discussed.
In this 200-MW wind project example, the total plant cost is assumed to be $2,120/kW. This
represents a significant increase from the $1,000/kW used in the 2004 and earlier editions of the
RETG, and accounts for the recent cost increases for wind turbines and other components caused
by high demand and higher steel prices.
A merchant plant needs to sell its power at a price sufficient to cover fuel costs, O&M costs,
income and property taxes, debt service, working capital, and return on and recovery of the
equity investment. This price is usually calculated at the busbar of the plant (essentially at the
plant boundary on the high-voltage side of the transformer) and may be separated into capacity
and energy components. Some regions of the country have a shortfall of generating capacity and,
consequently, have both the capacity and energy components. Other markets have only an
energy component. These payments can be further complicated by various risk-hedging business
deals (from options to tolling agreements, which are not considered in this report) with power
marketers, gas companies, or utilities.

2-13

Economic Methodology and Assumptions

For this 200-MW wind plant example, we have assumed the following parameters and costs:

Time frames

The durations of project development and permitting can vary widely and are not
explicitly included here. Reasonable cost allowances for these activities are discussed
below.

Duration of engineering, procurement, and construction: 12 months.

Commercial operation date: January 1, 2011.

Service life of plant: 20 years.

Technical and operating parameters

Capability: 200 net MW at rated wind speed.

Operation: intermittent.

200-MW wind plant composed of 100 2.0-MW wind turbines.

2.2.3.1 Total Capital Requirement


The major categories of total capital requirement include construction, development, and
financing costs. Table 2-5 displays general categories for the uses of the capital funds, as well as
their sources, as described later. Capital costs can vary significantly as a function of variations in
many items such as equipment type and cost, interconnection, site difficulties, and financing
structure. This example is designed to illustrate a reasonable magnitude of all-in costs.
2.2.3.1.1 Construction Cost

The largest component of capital cost is usually engineering, procurement, and construction
(EPC), which are included in the term total plant cost (TPC). TPC includes detailed design,
equipment and materials procurement, and installation and construction costs. Typically, an EPC
contractor supplies these services. TPC costs generally range from 60% to 90% of the total
capital requirement, depending on development difficulty and financing costs.
Other costs that are usually included with the total capital requirement include the following:

Land (excluded for wind plant as land is leased).

Operating and maintenance mobilization (or startup costs): This example assumes one month
of operating labor and 0.5 months of full-capacity fuel consumption as approximate cost
allowances (no fuel is used by wind plant).

Spare parts. Spare parts costs can range widely depending on the maintenance philosophy
and remoteness for the plant. This example assumes approximately 2% of the EPC cost for
spare parts.

2-14

Economic Methodology and Assumptions


Table 2-5
Sources and uses of construction funds
Commercial Operation:
January 1, 2011

(Example assumes a 100% construction loan and 60% debt/40% equity


financing at plant startup)
2008

2009

2010

Total Capital
$1,000
$/kW

Proportion
of Total

Description

Units

SOURCES OF FUNDS
Equity
Debt

$1,000
$1,000

0
0

194,112
291,168

194,112
291,168

40.0%
60.0%

Total Project

$1,000

485,280

485,280

100.0%

USES OF FUNDS
Land
Total plant cost
Operating and maintenance mobilization
Spare parts
Subtotal Construction Cost

$1,000
$1,000
$1,000
$1,000
$1,000

0
0
0
0
0

0
0
0
0
0

0
409,000
137
8,180
417,317

0
409,000
137
8,180
417,317

0
2,045
1
41
2,087

0.0%
84.3%
0.0%
1.7%
86.0%

Owner's Cost
Financial fees
Interest during construction
Reserves
Subtotal Owner's and Financing

$1,000
$1,000
$1,000
$1,000
$1,000

0
0
0
0
0

0
0
0
0
0

2,045
5,823
12,520
20,450
40,838

2,045
5,823
12,520
20,450
40,838

10
29
63
102
204

0.4%
1.2%
2.6%
4.2%
8.4%

Working capital
Contingency

$1,000
$1,000

0
0

0
0

6,260
20,866

6,260
20,866

31
104

1.3%
4.3%

Total Capital Requirement

$1,000

485,280

485,280

2,426

100.0%

Notes:
Construction escalation =
Construction interest =

0.00% per month,


0.50% per month,

0.00% per year


6.00% per year

2-15

Economic Methodology and Assumptions

2.2.3.1.2 Financing Cost

Merchant plants can be financed either as a limited recourse project or through general corporate
debt. Limited recourse project financing refers to projects that are dependent on the specific cash
flows of the project as defined in the project agreements. Qualifying facilities and exempt
wholesale generators were commonly financed on a limited recourse project finance basis. Some
merchant plant projects are also being financed on a limited recourse project basis. However,
there is a trend for financing merchant plant projects with corporate debt (as discussed later in
this section of the report). For this example, debt is assumed to be financed with limited recourse
project financing. Assumed financing costs and parameters for a project-financed transaction
include the following:

Construction loan:

100% debt: Frequently, construction loans may require equity on a pro rata basis with the
construction costs. However, a 100% debt construction loan is assumed to keep this
example conceptually simple.

Interest rate during construction of 8.0%.

Financial fees of approximately 2% of the loan amount.

Debt service reserve fund of approximately six months of debt service. The lending
institution may require that a debt reserve fund be included in the capital cost of a project.
The fund is set up to pay the debt service for the term loan if unexpected conditions
adversely affect the project revenues.

Term loan:

60% debt/40% equity

Commercial bank loan rate of 8.0% for a 10-year term

2.2.3.1.3 Development Cost

Development costs include a variety of costs that a non-utility generating company incurs
to develop a project. Examples include security deposits, permitting (including construction
permits and environmental permits), owners engineering and general and administrative costs,
development fees, legal fees, and easements and rights-of-way. These costs can vary widely
depending on the complexity of the site, the financing, and the terms of contracts with the EPC
contractor and equipment suppliers. For this example (Table 2-5), an approximate cost allowance
of 5% of the TPC price was assumed for total development costs.
2.2.3.1.4 Other Costs

A variety of other costs are frequently included in the total capital requirement, including the
following:

Working capital, which includes requirements for cash and inventories, and for this example
is assumed to equal one month of account receivables

Contingency, assumed for this example to be 5% of the total cost

2-16

Economic Methodology and Assumptions

2.2.3.2 EPRI Capital Cost Definitions


Section 4 of the 1999 EPRI TAG report [1] provides capital cost definitions for regulated
utilities. The three measures of capital cost are total plant cost, total plant investment, and total
capital requirement.
The total plant cost (TPC) is the sum of process facilities capital, general facilities capital,
engineering and home office overhead, and project and process contingencies. It is usually
expressed in constant dollars and assumes overnight construction,that is, the construction of
the facility occurs instantaneously. When expressed in constant dollars, the total plant cost is also
referred to by several terms: real direct capital cost, overnight capital costs, or overnight
construction costs.
The total plant investment (TPI) adjusts the TPC to account for the fact that construction occurs
over a period of years, and hence the costs are distributed over the same period and tend to
escalate over time. In addition, the project allows for interest accrued on each expense from the
date of the expense until the completion and commissioning of the facility. The accrued interest
is called allowance for funds used during construction (AFUDC) and is also referred to as
interest during construction. TPI is the sum of the TPC, an adjustment for the escalation of
capital costs during construction, and AFUDC. The 1999 EPRI TAG Report [1] provides
formulas for calculating escalation and interest during construction as a function of the escalation
rate, cost of money, and construction expenditure profile.
The total capital requirement (TCR) is the sum of the TPI and owners costs such as land and
property tax, insurance, preproduction, startup, and inventory costs.
These terms are maintained for consistency in this report. However, a project that is project
financed typically organizes its capital costs into construction costs (sometimes referred to as
hard costs) and financing and development costs (sometimes referred to as soft costs). EPRI
values for project and process contingencies are typically included within the TPC numbers that
are guaranteed by the EPC contractor. An overall project contingency (based on development,
financing, permitting, and the potential for EPC contractor overruns) is normally included by the
banks and is calculated right before the total capital requirement is calculated.
Since the industry convention for summarizing capital costs varies somewhat from a regulated
utility to a market-based plant, the cost components are identified in Table 2-5 for comparison
with other cost breakdowns. For example, EPRIs value for total plant investment usually
includes the sum of the total plant costs and the interest during construction. TPI is not shown in
the table, but can be easily calculated.
2.2.3.3 Income Statement
The income statement summarizes the revenues and expenses for each year of the project.
Table 2-6 presents an example income statement, which is discussed in the text that follows.

2-17

Economic Methodology and Assumptions


Table 2-6
Example income statement ($1000)
B

10

11

12

13

14

15

16

17

18

19

20

2010

2011

2012

2013

2014

2015

2016

2017

2018

2019

2020

2021

2022

2023

2024

2025

2026

2027

2028

2029

2030

3 Revenues
4

Capacity Payments

Energy Payments

53,425

54,761

56,130

57,533

58,971

60,446

61,957

63,506

65,093

66,721

68,389

70,098

71,851

73,647

75,488

77,375

79,310

81,293

83,325

85,408

Federal Production Tax Credit

13,199

13,529

13,867

14,214

14,569

14,934

15,307

15,690

16,082

16,484

66,624

68,290

69,997

71,747

73,541

75,379

77,264

79,195

81,175

83,204

68,389

70,098

71,851

73,647

75,488

77,375

79,310

81,293

83,325

85,408

8.7

8.9

9.2

9.4

9.6

9.9

10.1

10.4

10.6

10.9

11.2

11.4

11.7

12.0

12.3

12.6

12.9

13.3

13.6

13.9

Total Revenues

Avg. Electricity Rev. (cents/kWh)

9
10 Fuel Costs
11

Variable Fuel Costs

12

Fixed Fuel Costs

490

13

Total

14
15 Variable Operating Expenses
16

Supplies and Consumables

307

314

322

330

338

347

356

364

374

383

392

402

412

423

433

444

455

467

478

17

Water & Water Treatment

18

Other

19

Land Laease

1,603

1,643

1,684

1,726

1,769

1,813

1,859

1,905

1,953

2,002

2,052

2,103

2,156

2,209

2,265

2,321

2,379

2,439

2,500

2,562

1,909

1,957

2,006

2,056

2,108

2,160

2,214

2,270

2,326

2,385

2,444

2,505

2,568

2,632

2,698

2,765

2,834

2,905

2,978

3,052

1,164

20

Total

21
22 Fixed Operating Expenses
23

Operating labor

728

746

765

784

804

824

844

865

887

909

932

955

979

1,004

1,029

1,054

1,081

1,108

1,135

24

General and administrative

146

149

153

157

161

165

169

173

177

182

186

191

196

201

206

211

216

222

227

233

25

Maintenance labor and matl

3,400

3,485

2,858

3,002

3,154

3,314

3,482

3,658

3,843

4,038

4,242

4,457

4,683

4,920

5,169

5,430

5,705

5,994

6,298

6,616

26

Insurance

1,456

1,492

1,530

1,568

1,607

1,647

1,688

1,731

1,774

1,818

1,864

1,910

1,958

2,007

2,057

2,108

2,161

2,215

2,271

2,327

27

Property Taxes

971

995

1,020

1,045

1,071

1,098

1,126

1,154

1,183

1,212

1,242

1,273

1,305

1,338

1,371

1,406

1,441

1,477

1,514

1,552

28

Other

6,700

6,868

6,325

6,556

6,797

7,048

7,309

7,581

7,864

8,159

8,467

8,787

9,121

9,469

9,832

10,210

10,604

11,016

11,445

11,892

8,609

8,825

8,331

8,612

8,905

9,208

9,523

9,850

10,190

10,544

10,911

11,292

11,689

12,101

12,529

12,975

13,439

13,921

14,422

14,944

58,015

59,465

61,666

63,135

64,636

66,171

67,741

69,345

70,985

72,661

57,478

58,806

60,162

61,546

62,959

64,400

65,871

67,372

68,902

70,464

29

Total

30
31 Total Operating Expense
32 Earnings Before Interest, Taxes,
33

Deprec. and Amortization (EBITDA)

34
35 Pro Forma Income Tax Calculations
36

Tax Depreciation

37 Earnings before Interest + Taxes


38

Interest Paid

39

Interest Received (at 5.00% per year)

81,800

130,880

78,528

47,117

47,117

23,558

(23,785)

(71,415)

(16,862)

16,018

17,519

42,613

67,741

69,345

70,985

72,661

57,478

58,806

60,162

61,546

62,959

64,400

65,871

67,372

68,902

70,464

23,293

21,686

19,949

18,073

16,048

13,860

11,498

8,946

6,190

3,214

250

250

250

250

250

250

250

250

250

250

40
41 Net Operating Loss
42 Taxable Earnings
43

State/Local Tax

44

Federal Tax

45 Total Tax Obligation

0
(46,829)

0
(92,850)

0
(36,561)

0
(1,806)

1,721

29,002

56,493

60,649

65,044

69,697

57,478

58,806

60,162

61,546

62,959

64,400

65,871

67,372

68,902

70,464

(2,341)

(4,643)

(1,828)

(90)

86

1,450

2,825

3,032

3,252

3,485

2,874

2,940

3,008

3,077

3,148

3,220

3,294

3,369

3,445

3,523

(15,571)

(30,873)

(12,156)

(600)

572

9,643

18,784

20,166

21,627

23,174

19,111

19,553

20,004

20,464

20,934

21,413

21,902

22,401

22,910

23,429

(17,912)

(35,515)

(13,984)

(691)

658

11,093

21,608

23,198

24,879

26,659

21,985

22,493

23,012

23,541

24,082

24,633

25,196

25,770

26,355

26,952

(28,917)

(57,335)

(22,576)

(1,115)

1,063

17,909

34,884

37,451

40,165

43,038

35,493

36,313

37,150

38,005

38,877

39,767

40,675

41,602

42,547

43,511

46
47 Net Earnings after Taxes

2-18

Economic Methodology and Assumptions

2.2.3.3.1 Revenues

Forecasting revenues over the service life of a merchant power plant is one of the most critical
aspects of the economic analysis. Revenues for a utility generator are set by the revenue
requirements method and approved by a state commission. However, there is no guarantee of
revenues with a merchant plant because the price of power is set by the electricity markets.
Section 2 of the 1999 EPRI TAG report [1] provides additional information.
A forecast of market prices for a deregulated power market varies with time of day and time of
year, as shown for a historical time period in Figure 2-1. Detailed forecasting of the market price
for power typically requires the following steps:

Project the cost of fuel for the generating units in the region.

Evaluate the cost of generation from existing units.

Project the cost of generation from planned units

Project the dispatch order of units that will comprise the market.

Include variations in electricity consumption as a function of time-of-day, time-of-year, and


weather.

Develop a price duration curve for each year of the service life.

Figure 2-1
Day-ahead prices for the California Power Exchange

2-19

Economic Methodology and Assumptions

In an annual market price duration curve, data are arranged to show the fraction of a year that the
market price is at or above a given level. An example of a price duration curve for 1998 is shown
in Figure 2-2. For the California Power Exchange, the price was greater than or equal to
approximately $29/MWh for 20% of the year. The integrated price duration curve shows the
cumulative average for any given portion of a year. For example, a market utilization of 20% of
the year would have resulted in an average price of $45/MWh. The average for the entire year
was $24.4/MWh.
This information (appropriately adjusted for transmission costs) can then be used in the
calculation of revenues for the merchant plant.

Figure 2-2
Example price duration curve

For the example (not related to the California Power Exchange data), the revenue assumptions
include the following:

For simplicity, only an energy component is assumed for the price of electricity (see line 8 of
Table 2-6). The energy component is based on average power sales and assumed to be
approximately 8.5 cents/kWh in January 2010, escalating at 2.5% per year. For some
projects, there would also be a capacity component (line 7 of Table 2-6) of revenues.

In addition, it is assumed that the 200-MW wind plant qualifies for the 10-year Federal
Production Tax Credit, which was increased to 2.1 cents/kWh by the Emergency Economic
Stabilization Act of 2008 and is assumed to escalate at 2.5%/yr for 10 years.

Average annual capacity factor: 35%.

2-20

Economic Methodology and Assumptions

2.2.3.3.2 Operating and Maintenance Expenses

Operating and maintenance costs need to be projected for each year of the service life. For this
example, the operating cost assumptions (included in lines 15 through 31 of Table 2-6) include
the following:

Average annual capacity factor: 35%.

Average annual net plant heat rate (lower heating value): N/A for wind plant.

Delivered price of natural gas: N/A for wind plant.

Variable operating costs: 0.05 mills/kWh (at Jan. 1, 2011).

Annual land lease cost: 3% of electricity sale revenue.

Labor: five employees at an average fully burdened wage rate of $35/hour (at Jan. 1, 2011).

General and administrative expenses: 20% of operating labor.

Maintenance labor and material: $17/kW-yr (at Jan. 1, 2011).

Insurance: 0.5% of the total capital requirement.

Property taxes: 1.25% of the total capital requirement.

All costs escalate at 2.5% per year (for simplicity).

Income statements often include a book depreciation component for the calculation of book
income. However, this income statement calculates earnings before interest, taxes,
depreciation, and amortization (EBITDA) to focus on the calculation of net cash flows (see
the next subsection).

2.2.3.4 Income Taxes


Income taxes are included on a conceptual basis. For the example, key assumptions include:

All of the depreciable capital investment qualifies for 20-year MACRS tax depreciation
(line 36, Table 2-6) (for simplicity)

Interest paid on the debt service is tax deductible (Line 38, Table 2-6)

Interest received from the debt service reserve increases the taxable earnings (line 39,
Table 2-6)

A state income tax rate of 5% (line 43, Table 2-6)

Marginal federal income tax rate of 35% (line 44, Table 2-6)

This example assumes that the project structure includes a single-purpose corporation. It is
important to note that some projects are structured as partnerships. Although partnerships are
not taxed directly under federal and most state laws, a tax computation is needed so that
investors can compute their individual cash return from the project and also to compare it
with other opportunities. In this case, the tax information needed by the partners is computed,
but tax payments are excluded from the projects cash flow forecast.

2-21

Economic Methodology and Assumptions

2.2.3.5 Cash Flow Statement


The cash flow statement calculates the after-tax net cash flow for the project. Table 2-7 presents
an example cash flow statement.
The cash flow statement begins with the EBITDA (line 5 in Table 2-7) as brought forward from
line 33 of Table 2-6 and includes the following adjustments (lines 925, Table 2-7):

Less income taxes

Less debt service (principal and interest payments for the loan)

Plus interest received from the debt reserve fund

Less any new contributions to reserves (sometimes additional reserves may be required)

Plus return of the reserves at the end of the debt service term (the reserve fund is returned if it
has never been drawn down)

Less any adjustments to working capital (growth in working capital requirements)

Less equity investment during construction

The net cash flows represent the cash flows associated with the equity investor (owner of the
project). During the construction period, the cash flows are negative. Throughout the service life
of the plant, the net cash flows are initially negative and become positive in the out years.
The net present value (line 31, Table 2-7) and the internal rate of return (IRR on line 32,
Table 2-7) are economic measures of the project that reflect the present worth of profit over the
service life and the profitability of the project, respectively. The cumulative IRR reflects the IRR
for the relevant service life. For example, the IRR for a 20-year project life is projected to be
12.8%, whereas the IRR for the first 19 years of net cash flows is projected to be 12.2%. The
meaning of these measures is discussed in the following section.
Figure 2-3 shows the variation of the components of busbar electricity sales price versus year.
The fixed costs represent income taxes, debt service, and working capital. The fixed costs
decline when the debt service for the loan ends (a 10-year loan in this example). The O&M
component represents the fixed and variable operating and maintenance component. The net cash
flow represents the capital recovery for the equity investor.

2-22

Economic Methodology and Assumptions


Table 2-7
Example cash flow statement ($1000)
A

Commercial Operation: January 1, 2010

1
2

-2

-1

10

11

12

13

14

15

16

17

18

19

20

2008

2009

2010

2011

2012

2013

2014

2015

2016

2017

2018

2019

2020

2021

2022

2023

2024

2025

2026

2027

2028

2029

2030

58,015

59,465

61,666

63,135

64,636

66,171

67,741

69,345

70,985

72,661

57,478

58,806

60,162

61,546

62,959

64,400

65,871

67,372

68,902

70,464

658

11,093

21,608

23,198

24,879

26,659

21,985

22,493

23,012

23,541

24,082

24,633

25,196

25,770

26,355

26,952

4
5 EBITDA

6
7

Taxes Paid

(17,912) (35,515) (13,984)

(691)

9 Cash Flow From Operations

75,927

94,980

75,651

63,825

63,978

55,078

46,132

46,147

46,105

46,002

35,493

36,313

37,150

38,005

38,877

39,767

40,675

41,602

42,547

43,511

43,393

43,393

43,393

43,393

43,393

43,393

43,393

43,393

43,393

43,393

1,023

1,023

1,023

1,023

1,023

1,023

1,023

1,023

1,023

1,023

10
11

Debt Service

12

Interest Received

13

Contribution to Reseserves

14

Disbursement (return) of Reserves

15
16 Additions to Working Capital
17

Accounts Receivable

182

187

191

196

201

206

211

216

222

200

205

210

216

221

227

232

238

18

Fuel Stocks

(1,082)
0

19

Spare Parts

205

210

215

220

226

231

237

243

249

255

262

268

275

282

289

296

304

311

15,645

16,709

18,900

20,358

21,850

23,374

24,933

26,526

(9,763)
0
(12,758)

20
21 Capitalized Refurbishments
22
23 Contributed Capital

194,112

0 (194,112)

28,155

29,820

58,304

58,344

59,688

61,061

62,461

63,890

65,348

66,836

68,353

92,985

0 (194,112) (178,467) (161,759) (142,859) (122,501) (100,651) (77,276) (52,343) (25,817)

2,338

32,158

90,462

148,806

208,494

269,555

332,016

395,906

461,255

528,090

596,443

689,428

0 (194,112)

3,276

0 (194,112) (160,555) (108,332) (75,447) (54,398) (33,207) (20,926) (17,601) (14,273) (10,998)

24
25 Net Cash Flow Before Tax
26

Cumulative

27
28 Net Cash Flow After Tax
29

Cumulative

33,557

52,224

32,884

21,049

21,191

12,281

3,325

3,328

3,161

36,319

35,851

36,676

37,519

38,379

39,257

40,153

41,066

41,998

66,032

(7,837)

28,482

64,333

101,009

138,528

176,908

216,165

256,317

297,384

339,381

405,414

-1.2%

3.2%

5.7%

7.5%

8.8%

9.8%

10.6%

11.3%

11.8%

12.2%

12.8%

30

66,512

31 Net Present Value (at time = 0)

-38.8%

32 Cumulative IRR on Net Cash Flow After Tax

-21.2%

-13.0%

-6.7%

-3.8%

-3.1%

-2.4%

-1.8%

33
34 Coverage Ratios (before tax)

Minimum Average

35

Total Debt Service

1.34

1.51

1.34

1.37

1.42

1.45

1.49

1.52

1.56

1.60

1.64

1.67

na

na

na

na

na

na

na

na

na

na

36

Debt Service Including Reserves

1.67

1.93

1.81

1.84

1.89

1.93

1.96

2.00

2.03

2.07

2.11

1.67

na

na

na

na

na

na

na

na

na

na

37
38 Reserves
39

Contribution to Reserves

20,450

40

Reserve Balance

20,450

20,450

20,450

20,450

20,450

20,450

20,450

20,450

20,450

41

Disbursement of Reserves

42

Operating Income with Reserves

20,450

78,465

79,915

82,116

83,585

85,086

86,621

88,191

89,795

91,435

72,661

7,282

7,464

7,651

7,842

8,038

8,239

8,445

8,656

8,873

9,095

8,013

8,213

8,419

8,629

8,845

9,066

9,293

9,525

9,763

10,007

8,385

8,594

8,809

9,029

9,255

9,486

9,723

9,967

10,216

10,471

10,733

11,001

11,276

11,558

11,847

12,143

12,447

12,758

13,077

43
44 Working Capital Balances
45

Accounts Receivable

46

Fuel Stocks

47

Spare Parts

0
8,180

2-23

Economic Methodology and Assumptions

200

Busbar Cost ($/MWh, end-of-year $)

150

100

Federal PTC
Net Cash Flow
Fixed Costs

50

O&M
Land Lease
Fuel

(50)

2030

2029

2028

2027

2026

2025

2024

2023

2022

2021

2020

2019

2018

2017

2016

2015

2014

2013

2012

2011

(100)

Figure 2-3
Components of busbar cost for example 200-MW wind generation plant

2.2.3.6 Economic Measures


Several common measures address the economic viability of a project. The busbar cost includes
capital, O&M, fuel, and other costs of generating electricity. The profitability of the net cash
flows is typically addressed by calculating the net present value (NPV) and the IRR. Other
common measures that address the riskiness of a project include the payback period and debt
coverage ratios.
2.2.3.6.1 Busbar Cost

The busbar cost is the unit cost of generating one unit of electricity (e.g., 1 MWh), and is equal
to the total annual costs minus credits received, divided by the total annual electricity generation.
For the 200-MW wind plant example, the cost components include fixed and variable O&M cost,
other fixed costs including income taxes and debt service, and other net cash flow. The credit is
the 10-year federal production tax credit (PTC), which is assumed to be 2.1 cents/kWh in 2009,
escalates at 2.5%/yr for 10 years, and is zero after 10 years. Figure 2-3 illustrates how the various
cost components and the PTC vary over the 20-year operating period of the wind plant. The 10year PTC is shown as a negative contribution to the busbar cost during the first 10 years of
operation, after which it is eliminated. The fixed cost component decreases after 10 years
because the 10-year loan has been retired and the debt service component of fixed cost is
eliminated. The average annual busbar cost increases from 8.7 cents/kWh in 2009 to
13.9 cents/kWh in 2028.

2-24

Economic Methodology and Assumptions

2.2.3.6.2 Net Present Value

The NPV represents the present value (or present worth) of profit using the time value of money
principles described in Section 3 of the 1999 EPRI TAG report [1]. This calculation results from
discounting the net cash flows at the minimum acceptable rate of return for an equity investor.
The method is also referred to as the discounted cash flow method.
The net present value must be defined at a certain point in time. Frequently, the NPV is
calculated at the commercial operation date. In this case, the total capital requirement (at the
commercial operation date) is subtracted from the net cash flows that are discounted or brought
back to the same date. For the example provided, the total capital requirement is considered at
the commercial operation date, since the construction loan is based on 100% of the capital
investment and the interest during construction essentially includes the future worth of the
construction expenditures at the commercial operation date.
The NPV may also be calculated at the beginning of the construction time period if equity is
required to fund construction (as opposed to a 100% construction loan). In any event, the NPV
can be calculated at the point in time that makes the most sense for the evaluation.
In addition to determining the present value of cash flows for potential projects, the discounted
cash flow method can be used in other ways as well. Up to this point, revenues from the project
were assumed to be known. With an assumed rate of return, the discounted cash flow approach
can then be used to compute the annual revenues, or rate structure, necessary to make the project
a viable one. Essentially, this is what is done in the revenue requirements method.
Another application would be to use the discounted cash flow method to negotiate with a
potential project constructor. If all the other variables are known (or estimated), the discounted
cash flow approach could be used to compute the initial cash outlay that would make the project
feasible. Alternatively, the discounted cash flow method can be used to project values for
liquidated damages if a contractor fails to meet specific targets on plant performance or capacity.
Other types of studies can include an assessment of capital/energy tradeoffs, such as an
evaluation of the potential benefits from spending capital dollars to reduce heat rate. In many of
the previously discussed cases, solutions are found using trial-and-error algorithms, and
spreadsheets can find solutions for even complex problems very quickly.
Finally, a contract may cover a period that does not extend over the entire economic life of a
project. There is, however, value to the project for the period beyond the life of the contract,
which is referred to as its terminal value. Additionally, there may be a salvage value or a
dismantling and decommissioning cost at the end of the service life. Analysts often omit these
values or costs because they are subject to uncertainty since they occur far in the future, have
relatively little impact on the results, are hard to estimate, and make the analysis slightly more
conservative. However, there may be some cases in which it is appropriate to include the
terminal value in calculating discounted cash flows.
2.2.3.6.3 Internal Rate of Return

The IRR addresses the profitability of a project. Mathematically, the IRR is defined as the
discount rate that sets the present worth of the net cash flows over the service life equal to the
equity investment at the commercial operating date.
2-25

Economic Methodology and Assumptions

An IRR of 20% does not necessarily mean that net cash flows will represent 20% of the equity
investment for each and every year of service life. However, an IRR of 20% does mean that the
equity investor will earn the equivalent of 20% of the outstanding equity balance each year. That
balance will be reduced in some fashion over the life of the plant.
Many companies have a minimum acceptable rate of return that must be met before a potential
project is seriously considered. This minimum acceptable return is known as the hurdle rate,
which can be used to screen potential projects based on their internal rates of return. In the
example above, the project would be viable if its 20% internal rate of return is greater than the
hurdle rate.
There are several caveats to be aware of when calculating the IRR:

The IRR solution is a trial-and-error solution that is typically solved by a convergence


routine available in spreadsheet software.

The solution is based on solving an nth-degree polynomial that may have multiple real,
positive roots. More than one change in the sign of the coefficients of the net cash flows is an
indication of possible multiple roots. The net cash flow line in the example in Table 2-7 has
only one real positive root due to only one change in sign of the coefficients (negative net
cash flow at the commercial operating date changing to positive net cash flows for the
service life). A standard engineering economics textbook should be consulted for situations
where multiple roots are suspected.

Changes in the IRR are not scalar, and a small change in the cash flows can have a large
effect on the IRR.

Comparing IRR values may be misleading. Although the internal rate of return allows
investors to rank options based on their potential rate of return, it does not take into account a
projects size. For example, it does not allow an analyst to compare a $1 million project with
a 25% internal rate of return with a $10 million alternative having a 20% internal rate of
return. An incremental analysis may be required. Again, a standard engineering economics
textbook should be consulted for these situations.

2.2.3.6.4 Payback Period

The payback period represents how long it will take the net cash flows to recover the capital
investment (i.e., how many years for the cumulative net cash flows to become positive). There
are two versions of the payback period measure:

The simple payback period typically represents the number of years for the net cash flows to
pay back the capital investment. The calculation requires summing the net cash flows for the
preceding years. The time value of money is not considered in the calculation.

The discounted payback period represents the number of years for the present worth of net
cash flows to recover the capital investment. The time value of money is considered in the
calculation.

The discounted payback period is the preferred measure since it provides a proper economic
signal for the importance of the net cash flows. The simple payback period can be misleading
since cash flows in future years have the same apparent value as cash flows in the current year.
2-26

Economic Methodology and Assumptions

The payback period is a measure of secondary economic importance since it does not consider
all of the net cash flows over the life cycle of a project. Significant cash flows after the payback
period are not included in the measure number. Nevertheless, the payback period remains a
popular indicator of the risk level of the project.
2.2.3.6.5 Debt Coverage Ratios

For projects that are debt financed, debt coverage ratios address the level of risk for the project
from the perspective of the lending institution. The debt coverage ratio is calculated as the
EBITDA divided by the annual debt service. EBITDA represents the revenues minus expenses
(not including book depreciation) on a before-income-tax basis. The debt service represents the
principal and interest payments to service the loan. The ratio is designed to address the ability of
the projects cash flows to service the debt; the higher the ratio, the lower the level of risk for the
project.
The minimum debt coverage ratio should always be greater than 1.0. Depending on the risk of
the project, banks may require that the minimum never go below a certain threshold, such as 1.3
or 1.5.
Lending institutions also calculate the average debt coverage ratio over the life of the loan. This
measure reflects the ability to service the loan over its term. Depending on the level of risk, the
lending institution may require that the average over the life of the loan never drop below a
certain threshold, such as 1.7 or 2.0.
2.2.3.7 Sensitivity Analysis
Sensitivity analysis is important for understanding the key drivers for the economic pro forma.
An economic analysis typically begins with specifying a base case as defined in the previous
example. Sensitivity analysis is based on re-evaluating the economic analysis by changing input
parameters by reasonable amounts. The analyst can then study the results to determine the key
drivers in the pro forma model. An analyst can also determine what reasonable variations in
input parameters may cause the project to miss its financial goals.
Table 2-8 presents a sensitivity analysis for the previous example. The effects of reasonable
variations in key input parameters on key economic results are shown. For example, a 20%
decrease in the TPC would cause the IRR to increase from 12.8% for the base case to 18.8% (the
pro forma assumes that certain development costs are based on the TPC costthese would
decrease accordingly). Debt coverage ratios and payback periods would improve accordingly.

2-27

Economic Methodology and Assumptions


Table 2-8
Example sensitivity analysis

Input Data
Parameter
CAPITAL COST
Total Plant Cost

% Change

After Tax

Value

Case

IRR1

Net Present Value2


$ million (Dec 2008)
Result

Senior Debt
Coverage Ratios3
% vs. Base Minimum Average

Simple
Payback
Period

Discounted
Payback
Period

(years)

(years)

$ thousand, December 2009


20%

490,800

8.6%

1,917

4%

1.10

1.25

13

20

Base

409,000

12.8%

66,512

122%

1.34

1.51

11

14

-20%

327,200

18.8%

131,108

240%

1.69

1.90

40%

11.1%

53,097

97%

2.01

2.27

15

Base

60%

12.8%

66,512

122%

1.34

1.51

11

14

80%

16.8%

80,037

146%

1.00

1.13

12

13

20

FINANCING
Debt Proportion

Loan Term

Years
Base

Interest Rate

9.3%

12,690

23%

1.15

1.29

13

10

12.8%

66,512

122%

1.34

1.51

11

14

12

15.4%

103,401

189%

1.50

1.69

11

6.0%

13.8%

81,166

148%

1.47

1.65

13

8.0%

12.8%

66,512

122%

1.34

1.51

11

14

10.0%

11.7%

51,172

94%

1.22

1.38

11

16

11

Rate
Base

REVENUES
cents/kWh, January 1, 2010
Power Revenue

20%

10.2

10

17.1%

138,989

254%

1.58

1.77

Base

8.5

12.8%

66,512

122%

1.34

1.51

11

14

-20%

6.8

11

8.1%

(5,964)

-11%

1.10

1.24

13

>20

40%

12

16.5%

126,394

231%

1.55

1.75

11

Base

35%

12.8%

66,512

122%

1.34

1.51

11

14

30%

13

8.9%

6,631

12%

1.12

1.27

13

20

12

Capacity Factor

Federal Production

cents/kWh, January 1, 2010

Tax Credit
Base

3.0

14

14.6%

91,769

168%

1.47

1.65

2.1

11.9%

54,667

100%

1.29

1.45

11

15

0.0

15

8.9%

7,580

14%

1.03

1.17

13

20

Notes
1. Internal rate of return on discounted net cash flow after tax.
2. Discount rate is 8.5%.
3. Senior debt coverage ratios are based on operating income (before tax, not including reserves) divided by debt service.

As can be seen from the table, the key drivers in the analysis are the power sales revenue,
capacity factor, federal PTC, and total plant cost. A 20% decrease in electricity revenues would
cause the IRR to decrease from 12.8% to 8.1%, and the debt coverage ratios to be low and
possibly unacceptable depending on the goals of the project.
The sensitivity analysis provides a basis for beginning the probabilistic risk analysis that is
described in Sections 8 and 9 of the 1999 EPRI TAG report [1].

2-28

Economic Methodology and Assumptions

2.2.3.8 Project Risks and Financing


Non-utility generating plants can be financed as limited recourse projects or through general
corporate debt.
Limited recourse project financing refers to those projects that are dependent on the specific cash
flows of the project as defined in specific project agreements. Limited recourse financing usually
requires the formation of a special-purpose company to protect the exposure of the parent
company. The financial exposure of the owner is typically limited to the equity invested in the
physical and intangible assets of the project. However, the interest of the owner in the assets is
junior to that of the lenders. This type of financing requires acceptable risk allocation through
proper contract structuring with financially strong contractual parties, performance guarantees,
and liquidated damages (intended to backstop guarantees provided by the parties to the various
project contracts and agreements). The backstop for the guarantees can be assigned to the
appropriate original equipment manufacturers.
Qualifying facilities (under PURPA) were commonly financed on a limited recourse project
finance basis because these projects had a revenue stream that was secured by long-term power
sales agreements. The ultimate risk of power sales became a function of the creditworthiness of
the utility that had contracted to buy the power. Most utilities have been creditworthy because of
their regulated monopoly status.
The regulatory restructuring of the industry has allowed the development of merchant power plants
that sell power on a commodity basis (as previously shown in Figure 2-1). Limited recourse
financing for a merchant plant requires more stringent assurances that the project can meet its
goals. The amount and type of limited recourse debt financing will be a function of the following:

Properly structured project ownership. An appropriate legal structure and experienced and
well-managed sponsors are fundamental requirements. Experienced contractors are required
for engineering, procurement and construction.

The proportion of the plants output that will be under contract. It is expected that this
percentage will vary widely between developers with different risk strategies. A sound
security package in terms of the project agreements will secure the appropriate cash flows
with a low degree of volatility.

The competitiveness of a project in its region. The project will have to be one of the lowestcost producers. However, being one of the lowest-cost producers in a commodity-driven
market has several implications:

Maximizing the capacity factor will become relatively less important than maximizing
the profitability of the plant production.

The economics of a project may be driven by capacity payments in some areas of the
country, while in other regions they will be driven by energy payments.

Different plant types will occupy a niche based on the time of day or season of the year.
For example, some natural gas combined-cycle units may be most profitable in the
shoulder or intermediate peaking times of day. Simple-cycle natural-gas-fired plants
and pumped storage plants are typically most economical for the time of day and season
of the year when power is at peak demand. The economic viability of run-of-river hydro
plants will depend on the hydrologically wet seasons.
2-29

Economic Methodology and Assumptions

The capabilities for understanding and forecasting market pricing.

The ability of the project cash flows to cover the risks of power sales. Many projects will
have structural enhancements that take the following forms:

Methods to shift risk, such as subordinating fuel expenses to debt service payments, or to
utilize a tolling agreement for fuel supply and power output

Methods to share risk, such as higher equity investments, limited parent guarantees, or
subordinated debt provided by vendors and/or EPC contractors

Methods to account for the timing of risks, such as cash sweeps, reserve accounts and
commodity swaps, and remarketing of unused fuel.

Achieving an investment-grade rating for limited recourse project financing is difficult when
power sales are subject to commodity risk. The volatility of commodity prices leads to risk that
the plant may not be able to meet its financial goals. Consequently, lenders will require that
projects have a greater equity proportion of financing to absorb the volatility risk (i.e., the risk
that the cash flows will be insufficient to pay the debt service).
A typical merchant plant project that is project financed can be characterized as follows:

Equity proportion of the capital structure of 30% to 50% depending on the risk of the project.

Average annual debt coverage ratios of at least 1.75 to 2.0 (probably closer to 2.0).

Debt terms of 10 to 20 years with cash sweep provisions.

Relatively higher debt reserves than were required for qualifying facilities because of the
seasonality of electricity pricing. (Debt reserves of 6 to 12 months of debt service payments
may be required, depending on the level of risk. Debt service reserves can take the form of
cash that is financed, letters of credit, or other financial commitments.)

Other financial structural enhancements such as subordinated debt, limited parent guarantees
and commodity swaps.

A well-defined legal structure.

An internal rate of return on after tax equity that is typically at least 15%.

As more equity is required for limited recourse project financing, the financing complexity and
costs increase. Essentially, a premium is paid for limited recourse debt even though the higher
levels of equity are designed to absorb the risk of volatility in the cash flows. Consequently,
large firms with healthy balance sheets and diversified assets may have an advantage in
financing merchant power plants through general corporate debt. Corporate debt allows a
company to avoid burdensome contractual arrangements and project security arrangements that
are required for project financing and, in some cases, allows the company to tolerate more
commodity-based risk.

2-30

Economic Methodology and Assumptions

The advantage of financing with corporate debt is one of the factors driving the industry to
consolidate into larger companies. Large balance sheets (sometimes formed as a result of
mergers and acquisitions) allow for financing through corporate debt. Large companies with
diversified portfolios of power plant assets can also balance downside risks that occur in one
regional market with upside risks that occur in another. Financing with corporate debt may
provide funding that is more flexible, has shorter lead-time, and often has reduced costs.
In conclusion, the financing of merchant power plants is evolving along several parallel paths. A
framework for limited recourse project financing is developing in the financial community that
requires significant proportions of equity investment. Additionally, there is an increasing trend
for financing with corporate debt as the industry consolidates to larger firms.
2.2.4 Limitations of Examples
This section has outlined methods for evaluating power projects developed by non-utility
producers and provided comparisons to evaluations for utility projects. Many simplifying
assumptions have been applied that may not be appropriate for an actual project. The appropriate
method for any entity to use in making economic decisions depends on many factors including
its strengths and weaknesses, its strategic aim, its competitors, and the economic and legal
environment. In addition, as competition increases, industry participants will assess their
missions and their risk profiles, and the distinctions between regulated and non-utility entities
can be expected to blur.

2.3 Guidelines for Economic Evaluation of Renewable Energy Projects


2.3.1 Design/Cost Estimate
The rating system shown in Table 2-9 indicates the level of effort involved in the design and
cost estimate.
Table 2-9
Confidence rating based on cost and design estimate
Letter Rating1

Key Word

Description

Actual

Data on detailed process and mechanical


designs; historical data from existing units

Detailed

Detailed process design


(Class III design and cost estimate)

Preliminary

Preliminary process design


(Class II design and cost estimate)

Simplified

Simplified process design


(Class I design and cost estimate)

Goal

Technical design/cost goal


(or value developed from literature data)

Note:
1.

If complete design and cost-estimating methods and assumptions are not available, the rating is reduced by
one unit.

2-31

Economic Methodology and Assumptions

2.3.2 Accuracy Ranges


The accuracy of cost estimates has been discussed in detail in many texts and papers on the
subject. Table 2-10 presents estimates of the range of accuracy for the cost data presented in this
report.
Table 2-10
Accuracy range estimates for RETG cost data (ranges in percent)1
Design and
Cost Estimate
Rating

A
Mature

B
Commercial

C
Demo

D
Pilot

E&F
Lab & Idea

5 to + 5

10 to + 10

15 to + 20

C. Preliminary

10 to + 10

15 to +15

20 to +20

25 to +30

30 to +60

D. Simplified

15 to + 15

20 to +20

25 to +30

30 to +30

30 to +80

30 to + 70

30 to +80

30 to +100

30 to +200

A. Actual
B. Detailed

E. Goal
Note:
1.

This table indicates the overall accuracy of cost estimates. Accuracy is a function of the level of costestimating effort and the degree of technical development of the technology. The same ranges apply to
O&M costs.

2.3.3 Definitions of Economic Terms


The following briefly defines various economic terms used throughout this RETG report:
Plant size: Net electrical capacity of the unit that forms the basis for the cost estimates.
Available for commercial orders: Estimate of the earliest date that a commercial-size unit could
be ordered with normal performance guarantees, if the technology is not already commercially
available. This date assumes an orderly development program, with success at each step.
Because most technology development programs have unexpected delays and unsuccessful steps
that must be repeated, the actual commercial order date is likely to be later than the one shown.
Furthermore, the very first commercial units will be somewhat more expensive (in constant
dollars) than the mature unit designs that form the basis for the cost estimate.
First commercial service: Estimate of the earliest date that a commercial-size unit could be put
into operation. The first commercial service date estimate is usually based on adding the idealized
construction time plus an allowance for design time to the available for commercial orders date.
Hypothetical in-service year: Year in which the plant is assumed to be placed into service. In
most cases this is the same as the first commercial service.
Plant capital cost: Basis for the cost estimate in terms of the number of units included in the plant.
General facilities and engineering fee: General facilities capital is the total construction cost of
the general facilities, including roads, office buildings, shops, laboratories, etc. The engineering
fee is the engineering and home office overhead and fee paid to the architect/engineering
company, and is considered representative of the type of generation or storage unit.
2-32

Economic Methodology and Assumptions

Project and process contingency: Project contingency is a capital cost contingency factor
covering the cost of additional equipment or other costs that are not apparent in the preliminary
design and that would result from a more detailed design of a definitive project. Process
contingency is a capital cost contingency factor applied to a new technology in an effort to
quantify the uncertainty in the technical performance and cost of the commercial-scale
equipment.
Total plant cost (TPC): Total of all direct and indirect construction costs at a single point in time,
excluding the time value of money (interest, etc.) during the construction period. Capital cost
estimates expressed at a single point in time are often referred to as instantaneous or overnight
construction costs.
Total cash expended (TCE): Amount of money that would have been spent for all labor,
materials, and indirect costs at the time the unit goes into service. TCE is expressed in mixedyear dollars accumulated over the construction period.
Allowance for funds used during construction (AFUDC): Interest paid on the cash expended
during the construction period. This interest is accumulated and compounded until the plant goes
into service.
Total plant investment (TPI): Sum of TCE and AFUDC. TPI is expressed in mixed-year dollars
accumulated at the same point in time as total plant cost. For cases where the expenditure rate is
uniform, TPI, TCE, and AFUDC can be calculated by multiplying total plant cost (TPC) by the
appropriate adjustment factor (AF) shown in Table 2-11. That is:
TPI = TPC AF (TPI)
TCE = TPC AF (TCE)
AFUDC = TPC AF (AFUDC)
Table 2-11
TPC-TPI adjustment factors
Adjustment Factors (AF)1

Idealized
Construction
Time (Years)

AF (TPI)

AF (TCE)

AF (AFUDC)

1.0

1.0

1.025

0.980

0.045

1.050

0.961

0.089

1.076

0.942

0.134

1.103

0.924

0.179

1.131

0.906

0.225

1.160

0.889

0.271

1.189

0.872

0.317

Note:
1. Based on 4.1% cost escalation and 9.2% AFUDC interest rate.

2-33

Economic Methodology and Assumptions

Owner costs: This includes prepaid royalties, startup costs, inventory capital, catalysts and
chemicals, and land.
Total capital requirement: All capital costs that would go into the rate base on a hypothetical inservice date. Since the technology may not be available until a much later point in time, the total
capital requirement should be multiplied by the appropriate cost escalation factor for a future inservice date.
Total capital replacement: Capital cost that would be incurred by employing a particular
technology; the total capital cost over the lifetime of the unit. This represents major replacement
or refurbishing of equipment such as replacement of blades in a combustion turbine.
Operation and maintenance (O&M) costs: Costs are expressed at the same point in time as the
total capital requirement and represent the costs for a normal year of operation. The extra costs
associated with plant startup in the first year are included in the total capital requirement.
Incremental costs are broken down into variable and consumable components. Consumables are
computed directly from the plant material balance. The variable component primarily represents
variable maintenance and is estimated from an algorithm incorporating the units expected
capacity factor.
Unit availability: Fraction of time a generating unit is able to supply power at various capacity
levels. Estimates are based on mature plants and are likely to be significantly more optimistic
than the initial experience with first-of-a-kind plants.
Minimum load: Estimate of the minimum load from which the unit could be ramped up to full
load in a reasonable time (one hour or less).
Pre-construction, licensing, and design time: Time periods for pre-construction activities depend
on local regulations and conditions. The time period is intended to be a reasonable
approximation, but actual times may range from one-half to two times the estimated time.
Idealized plant construction time: The idealized construction time is intended to represent the
minimum construction time required to build a plant under ideal conditions. Actual construction
times will likely be somewhat longer due to weather, regulatory problems, labor strikes, etc.
Unit life: Estimate of the book life of the plant. Maintenance costs include sufficient funds to
replace minor equipment that wears out before the unit life shown.
Technology development rating and design/cost estimate rating: Rating systems designed to
indicate the quality of the information used to develop the data.

2.4 References
1. TAG Technical Assessment Guide, Volume 3, Rev. 8: Fundamentals and Methods
Electricity Supply. EPRI, Palo Alto, CA: 1999. TR-100281-V3R8.
2. Wind Power Technology Status and Performance and Cost Estimates2009. EPRI, Palo
Alto, CA: 2009. 1020362.
3. Technical Assessment Guide, Electricity Supply1993, Volume 1, Rev. 7. EPRI, Palo Alto,
CA: 1993. TR-102275-V1R7.

2-34

Economic Methodology and Assumptions

4. U.S. DOE Energy Information Administration, Cost and Quality of Fuels for Electric Utility
Plants 1994, July 1995, pp. 15, 27.
5. U.S. DOE Energy Information Administration, Annual Energy Outlook 1996, January 1996,
p. 79.
6. U.S. DOE Energy Information Administration website, Status of State Electric Industry
Restructuring Activity, http://www.eia.doe.gov/cneaf/electricity/chg_str/regmap.html, July
2003.

2-35

WIND POWER

3.1 Introduction
Wind is generated by small regional differences in atmospheric pressure caused by solar heating
of the Earths surface, radiation cooling at night, the passage of air over warm or cold ocean
water, the passage of fronts and storms, and other complex meteorological phenomena. Since the
dawn of civilization, people have relied on the wind to propel sailing vessels and power graingrinding mills, saw mills, water pumps, electric generators, and other devices.
In recent years, the modern wind turbine has been developed and commercialized for electricity
generation, principally in Europe and the United States. Wind is now one of the fastest-growing
forms of electricity generation in the world. During 2011 alone, more than 41,000 MW of new
wind capacity was installed worldwide. As of August 2012, the installed wind generation
capacity was 50,000 MW in the United States and approximately 250,000 MW worldwide
(Table 3-1).
This section of the RETG addresses worldwide historical and projected installed wind generation
capacity, wind turbine technology, likely development pathways, wind turbine performance and
project cost estimates, O&M labor requirements, grid integration, onshore and offshore project
development process and market, environmental issues, potential for greenhouse gas reduction
and evolving markets for equipment and services.
Table 3-1
Wind power overview
Installed Capacity
(August 2012)

245,000 MW estimated worldwide

50,000 MW in the United States

World leaders:

- European Union: 100,000 MW; China: 63,000 MW


Technology Readiness

Environmental Impact

Mature, commercial

High confidence in cost estimates and projections

New technologies /enhancements being commercialized

Minimal during manufacturing and operation.

New technologies minimize avian/bat mortality.

Sight, shadow flickering and sound impacts must be considered in


planning and operation.

Radar interference issues in many sites to be addressed.

3-1

Wind Power
Table 3-1 (continued)
Wind power overview
Economic Status

Policy Status

Trends to Watch

Competitive in many markets.

Tax and cash incentives still desirable to facilitate most projects.

US production tax credit expiration in 12/2012, will have an impact on


new power to be developed in 2013.

Encouraged by RPS and energy policies in most jurisdictions.

Encounters occasional resistance from the public, environmental


groups, and others concerned primarily with visual impact of turbines.

Near-term global capacity growth of approximately 23% per year.

Larger individual turbines (taller towers, longer blades).

Larger aggregate wind farms.

Advanced materials and controls.

Rapid expansion into high wind resource areas (U.S. plains, North
Atlantic, Asia) as costs becomes competitive in more markets.

Wind energy storage/ stricter grid integration requirements

Entrance into low-wind-speed environments.

Offshore wind capacity growth in Europe and China

Substantial cost reduction of wind turbines in the near term due to


oversupply and market competition.

3.2 Installed Wind Capacity


The worldwide installed wind generating capacity continues to grow at a rapid rate. Figure 3-1
shows the historical installed wind generating capacity in 2011 and the projected capacity from
2012 through 2016 in the United States, Europe, and the rest of the world. During 2011, installed
capacity increased by about 21%, from 199,520 MW in 2010 to 241,029 MW, and is forecast to
more than double by 2016 to 488,424 MW.

3-2

Wind Power

Historical

Figure 3-1
Historical (2011) and projected (20122016) installed wind generation capacity in the
United States, Europe, and remainder of the world
Source: BTM Consult ApS, March 2012

Table 3-2 presents the installed capacity by country at the end of each year during the period
2005 through 2011. As of the end of 2011, the Peoples Republic of China led the world in
installed wind capacity (62,412 MW), followed by the United States (47,804 MW), Germany
(29,248 MW), and Spain (21,350 MW). China installed the most wind capacity during 2011 with
17,631 MW, approximately a 4% decrease over the countrys 2010 installed capacity. China was
followed by the United States (6,810 MW), Spain (1,050 MW), Germany (2,007 MW), and India
(3,300 MW)
The U.S. market has traditionally fluctuated from year to year depending on the status of the
federal production tax credit (PTC). The federal PTC expired on December 31, 2003 and was
extended to December 2005 by tax legislation passed by Congress in September 2004. Since
then, Congress has extended the PTC four times, to December 2007, 2008, 2009, and 2012. As a
result of the PTC extensions, wind capacity additions have soared in the United States.
Furthermore, the enactment of the American Recovery and Reinvestment Act (ARRA) in
February 2009 introduced a greater degree of project finance flexibility that is expected to spur
wind development. In addition to extending the PTC to 2012, ARRA also allows all PTC-eligible
projects to qualify for the investment tax credit (ITC). The stipulation essentially grants owners
of non-solar renewable energy facilities the option to irrevocably choose the ITC in lieu of the
PTC and, in turn, earn an up-front tax credit equal to 30% of a projects capital costs. The option
remains in effect for the current period of the PTCthrough 2012 for wind energy facilities and
3-3

Wind Power

through 2013 for other qualified renewable energy systems. Wind developers, along with wind
energy advocates, have been extremely busy this year encouraging lawmakers to extend the PTC
credit before it expires on December 31, 2012. They are pushing Congress to extend the PTC out
for at least four years to match the solar energy programs credit. To date (November 2012), they
have been unsuccessful in getting the PTC extended. Not having the PTC in place will greatly
affect the continued wind installation in the United States. In addition to the uncertainty of the
extension of the PTC, natural gas prices are severely impacting the competitiveness of wind
generation. Natural gas generation provides a more predictable and grid-friendly generation
technology for utilities.
The Chinese market continues to experience great growth in the wind market. The growth is due
in large part to governmental policies aimed at increasing renewable energy in the nations
portfolio. The National Development and Reform Commission (NDRC) set renewable energy
target of 11.4% of total energy consumption by 2015 and 15% by 2020. The Chinese
government foresees 90 GW total wind power installation by the end of 2015. Despite the
extensive installation growth, much work has yet to be completed with the electricity
infrastructure necessary to transport the generation to the nations loads. Many of the locations in
China with the greatest wind resources are located far from the industrial load centers. Major
investment in the countrys transmission system will need to take place. At the end of 2011,
approximately 24% of the installed wind capacity was still awaiting grid connection; that
amounts to over 15 GW waiting to be connected. Recently announced government policies in
China to manage these issues have slowed down development of wind power, but in the long
term should assist in allowing future deployment of wind technology.
As with China, positive governmental policies continue to support the growth of wind energy in
Europe as well. The general trend in Germany has been the gravitation toward projects being
further inland. These projects will serve as proving ground for installations that are designed for
lower wind resource areas, meaning larger rotor diameters and taller towers. Repowering of
older turbines with new turbines, in many cases larger capacity turbines will also contribute to
Germanys increase in wind capacity. Repowering of older turbine units contributed 238 MW.
This trend will continue as a reaction to the best sites for wind power being already occupied. In
addition, Germanys offshore market is expected to continue to grow thanks to a commitment for
the grid company to provide the electrical connection, including the undersea cable, from a wind
farm to the coast and an improved tariff. Germany as a whole reached a wind power penetration
level of ~11% of electricity consumption. France saw a 26% reduction in the level of wind
installations in 2011. In 2010, 1,186 MW of wind energy was installed. In 2011, France installed
875 MW of wind energy. Recent modifications to the permitting process have made obtaining a
wind permit a longer and more costly line item. In addition, regulatory uncertainty and
complexity has led to limited grid access in some cases. Despite the setbacks, France is still
expected to improve its wind installations going forward. In 2011, Spains new installations
reach a capacity of 1,050 MW. That is approximately a 30% reduction compared with 2010. The
main driver behind the slower market has been the financial crisis impacting the entire country.
Near-term installations will undoubtedly continue to be affected by the countrys financial
troubles.

3-4

Wind Power

Table 3-3 presents the forecast annual capacity additions for 2012 through 2016 and the
projected installed capacity at the end of 2016 by country. As discussed later, much of the new
wind capacity in Europe will be installed at offshore sites. From 2012 through 2016, China is
expected to install the most new wind generation capacity (82,000 MW), followed by the United
States (46,500 MW), India (23,000 MW), Germany (13,700 MW), France (6,800 MW), and
Spain (2,850 MW). By the end of 2016, the countries with the largest projected installed wind
generation capacity are expected to be China (144,412 MW), the United States (93,834 MW),
Germany (42,948 MW), India (39,266 MW), and Spain (24,200 MW). Figure 3-2 compares the
operating wind capacity in different nations and regions of the world at the end of 2008, 2009,
2010, and 2011 indicating both overall capacity growth and shifting shares of the total.
Table 3-2
Operating wind generation capacity: end of year 2005end of year 2011
Source: BTM Consult ApS, March 2012
Operating Wind Capacity by Country (MW)
Region/Country
Europe

2005

2006

2007

2008

2009

2010

2011

41,044

48,627

56,824

65,971

76,580

87,671

97,588

5,617

8,963

13,973

22,174

37,147

58,277

79,282

189

198

493

667

1,871

2,705

1,496

1,653

2,132

2,628

3,220

4,890

5,368

6,062

379

488

569

797

1,116

1,214

1,426

59,234

74,283

94,005

122,158

160,084

199,520

238,216

North America

Asia
South & Central America
Pacific Region
Total Middle East & Africa
Total World

Table 3-3
Forecast wind generation capacity additions: 20122016
Source: BTM Consult ApS, March 2012
Cum MW

Forecast Wind Capacity Additions (MW)

Cum
MW

End 2011

2012

2013

2014

2015

2016

Total
Added

End
2015

Europe

97,588

11,100

13,075

13,625

16,500

18,750

73,050

170,638

United States

47,084

8,250

7,500

9,000

10,000

12,000

46,750

93,834

Asia

21,005

19,150

20,400

21,500

21,900

26,500

109,450

188,732

Total World
(Including Unlisted Countries)

241,029

43,195

47,805

52,560

58,180

68,105

269,845

510,874

Region/Country

3-5

Wind Power

Figure 3-2
End-of-year operating wind capacity by country and region: 20082011
Source: BTM Consult ApS, March 2012

Figure 3-3 presents the BTM Consult forecasts of wind power development for 2012 to 2021.
The forecast is based on an analysis of a number of technical, economic, and political factors in
each country. The factors include national energy plans, government support for renewable
energy, present market for wind turbines, new project announcements, wind resource
assessment, and expected technology development and cost reduction.

3-6

Wind Power

1,100,000
1,000,000
900,000
800,000

MW

700,000
600,000
500,000
400,000
300,000
200,000
100,000
0
1990

2011

2016

2021

Figure 3-3
Wind capacity and forecast: 19902021
Source: BTM Consult ApS, March 2012

Historically, the U.S. market fluctuates from year to year depending on the status of the federal
PTC. The federal PTC expired on December 31, 2003, and was extended to December 2005 by
tax legislation passed by Congress in September 2004, and then to December 2007 by the 2005
Energy Policy Act. It was further extended to December 2008 by Congress when it passed a tax
and trade policy bill in December 2006. It was then extended until December 2009 when
Congress passed an energy improvement act in October 2008. And it was more recently
extended to 2012 by ARRA. As a result of the PTC extensions in 2005, 2006, 2007, and 2008,
wind capacity additions in those years exploded in the United States, totaling about 2,430; 2,454;
5,244; and 8,358 MW, respectively. Stipulations embedded in ARRAthe extension of the PTC
to the end of 2012, the option of project owners to choose the ITC in lieu of the PTC, and the
cash grant programhelped to support wind energy capacity additions in 2010 and beyond. As
the sunset of the PTC draws near, the developers have scurried to complete projects to ensure
participation in the program. he American Wind Energy Association (AWEA) reports that 1,832
MW of wind capacity were installed during third quarter 2012, and 4,728 MW through the third
quarter, which is approximately 41% greater than the first three quarters of 2011. Over 8,430
MW are currently under construction across 29 states and Puerto Rico. A major factor behind the
busy year is the possible expiration of wind PTC in 2012.

3-7

Wind Power

Developers are not confident that the PTC will be extended so they are attempting to install as
much wind energy as possible now to take advantage of the production tax credit. Figure 3-4
shows the distribution of U.S. wind generation capacity among the states as of the third quarter
of 2012. At the end of the third quarter of 2012, all but 11 states have wind power installations.
In 2011, Nevada did not have any wind capacity; in 2012, Nevada added 152 MW of wind
capacity. However, that does not mean that a utility within a state does not purchase wind
energy. For instance, in October 2012, Alabama Power signed a 20-year purchase power
agreement (PPA) for the electricity generated by TradeWind Energys 202-MW Buffalo Dunes
wind project in Kansas.
Forecasts of future U.S. wind capacity vary depending on different assessments of how planned
projects will proceed. As shown in Table 3-3, BTM Consult ApS forecasts that cumulative U.S.
wind capacity will reach approximately 93,384 MW in 2016.

Wind Power Capacity State by State3Q 2012, Total: 51,630 MW


Figure 3-4
Installed wind generation capacity in the United States by state as of 3Q 2012
Source: American Wind Energy AssociationAWEA 2012

3-8

Wind Power

3.3 Wind Energy Principles


Capturing the winds kinetic energy depends on the efficiency of the conversion process as well
as the power of the wind, which can be expressed by the equation:
Pw = 1/2 A V3
where = density of the air, A = swept area of the rotor, and V = velocity of the wind.
It can be seen that the power in the wind varies proportionally with the cube of the wind speed,
which has important bearing on the design and siting of wind turbines. As a result, even a small
increase in wind speed can substantially boost the power available in the wind. For example, a
25% increase in wind speed approximately corresponds to a doubling in the power contained in
the wind, which illustrates the importance of accurate resource assessment to a projects success.
Because of the strong link between wind speed and power, and because of the variable nature of
the resource, the direct use of average wind speeds to estimate energy generation can lead to
gross inaccuracies. Therefore, it makes sense to talk about the wind resource in statistical terms.
For sites where resource measurements have not been taken over a significant period of time and
actual distributions are unavailable, meteorologists have found that the Weibull probability
function provides a good approximation of the distribution of wind speeds over time. The
Weibull function is:
f(v) = (k/Vc)*(V/Vc)(k-1)*exp((V/Vc)k)
where k is a shape parameter between 1 and 3, and Vc is the characteristic velocity defined as the
weighted average wind speed at hub height. The Rayleigh distribution, shown in Figure 3-5, is a
special case of the Weibull function in which k = 2. It requires only the characteristic wind speed
to define an approximate distribution.
Although the greatest power production occurs at higher wind speeds, most sites do not
experience those high wind speeds with great frequency. Due to laws of conservation of mass
and momentum, there is a theoretical limit to the fraction of energy in the wind that can be
extracted by the rotor, called the Betz limit. The theoretical limit of rotor energy collection
efficiency is 59%. Because every other system in the turbine contributes losses as well, only a
fraction of the power in the wind can be extracted by a practical wind turbine. The actual
mechanical power output is:
Pm = Cp (1/2 A V3) = Cp Pw
where Cp is the coefficient of performance and represents overall system efficiency.

3-9

Wind Power

Figure 3-5
Rayleigh probability density function for wind speed

The coefficient of performance, Cp, varies with wind speed and blade tip speed, and is often
plotted as a function of the ratio of the blade tip speed to the wind speed because it clearly shows
the optimum operating point and is applicable to both constant-speed and variable-speed wind
turbines. Figure 3-6 illustrates a typical curve for the coefficient of performance versus the tip
speed/wind speed ratio.

Figure 3-6
Coefficient of performance (Cp) vs. tip speed/wind speed ratio

3-10

Wind Power

3.4 Developments in Wind Turbine Technology


Most utility-scale wind turbines currently available from established turbine manufacturers
utilize the Danish concept turbine configuration. This configuration uses a three-bladed rotor,
an upwind orientation (blades positioned upwind of the tower), and an active yaw system to keep
the rotor oriented into the wind. Figure 3-7 illustrates the major components of a typical utilityscale wind turbine design. The nacelle contains the drive train, which usually consists of a lowspeed shaft connecting the rotor to the gearbox, a two-or three-stage speed increasing gearbox,
and a high-speed shaft connecting the gearbox to the generator. Each turbine is equipped with a
transformer to step up the generator voltage to the on-site collection system voltage. Sometimes
this transformer is mounted within the nacelle, and sometimes pad-mounted transformers are
used near the base of the tower. The current status and recent developments of each of the main
wind turbine components are described in the following sections. The main sources of the
information are the 2010 EPRI report, Advanced Wind Turbine Technology Assessment2010
[67] and the 2009 report, Wind Power Technology Status and Performance and Cost Estimates
2009 [68].
Wind Speed &
Direction Sensors

Controller
Main Shaft

Blades

Hub
Cooling
System
(Radiator)

Nacelle

Tower
Generator
Access Door
Mechanical
Brake

Gearbox

Yaw Motor

Figure 3-7
Components of a typical wind turbine
Source: Danish Wind Industry Association, www.windpower.org

3.4.1 Generators and Power Electronics


In the early days of wind energy, wind plants were small relative to conventional sources of
generation. Consequently, wind plants were not required to help stabilize the power grid and
were even required to disconnect from the grid during transient events like grid faults. In recent
years, the power capacity of wind plants has grown significantly such that annual capacity
additions are comparable to those of conventional power plants. The penetration of wind energy,
which is a ratio of generation from wind to the total electrical load of the grid, has also increased.
This growth has caused a dramatic shift in the grid interconnection requirements imposed on
wind energy. Wind plants are increasingly being required to behave more like conventional
generation sources.

3-11

Wind Power

Grid integration requirements vary from region to region, but most follow the same basic
construct. Wind plants are required to contribute to the stability of the grid voltage by controlling
the flow of reactive power. Providing voltage support in this way is required during normal grid
operation, as well as during grid faults. There are limits to how quickly the power output of wind
plants can change over the course of fractions of seconds to one hour. Meeting these new, more
stringent grid-integration requirements, in addition to reducing structural loading and improving
energy capture, affects the design of the wind turbines, especially the generator topology.
3.4.1.1 Squirrel Cage Induction Generator
In the early 1990s, the standard wind turbine topology consisted of a constant-speed
asynchronous squirrel cage induction generator (SCIG) connected directly to the grid through a
transformer. Constant-speed wind turbines with SCIGs are relatively simple. The rotational
speed of the generator is essentially controlled by the frequency of the grid, and the power and
torque of the turbine are controlled by pitching the turbine blades. However, the reaction time of
the blade pitch system limits the wind turbines ability to limit torque and power, which results
in significant torque and power spikes during wind gusts. These spikes increase fatigue loading
and adversely affect the grid.
The stability of the grid voltage is maintained by controlling the flow of reactive power. SCIGs
always consume reactive power, and their reactive power consumption is not controllable. Power
factor correction capacitors are typically used to compensate for the reactive power consumed by
SCIGs, but SCIGs are unable to contribute to the stability of the grid. Most manufacturers have
traded the simplicity of the fixed-speed SCIG for the controllability available in more complex
generator configurations.
3.4.1.2 Variable-Speed Turbine
The rotational speed of many contemporary wind turbines is decoupled from the grid frequency
in order to achieve variable-speed operation. Variable-speed operation has two primary
advantages over fixed-speed operation. Variable-speed turbines are able to limit power and
torque by letting the rotational speed of the turbine rotor vary during wind gusts. The improved
control of torque reduces mechanical loading, and the improved control of power makes the
output electricity more palatable to the grid. The second advantage of variable-speed operation is
improved aerodynamic efficiency, which leads to increased energy capture in the lower wind
speed range. However, as discussed below, sophisticated power electronics are required to
enable variable-speed operation. Due to the parasitic electrical load associated with power
electronics, the net improvement in energy capture is reduced.
A patent held by General Electric (GE) has historically limited the use of variable-speed
technology in North America by other manufacturers. Recently GE has licensed the technology
to several other manufacturers; however, it is unclear if the license includes the capability to
provide reactive power at each turbine, which is held under a separate patent owned by GE. With
the imminent expiration of GEs variable-speed patent, manufacturers who already utilize
variable-speed technology in Europe are likely to implement the technology in the United States.

3-12

Wind Power

3.4.1.3 Power Electronics


Power electronics are solid-state devices (semiconductors) that convert the form of electric
power by changing the frequency and/or voltage and current. Power electronics play a significant
role in the generators of many wind turbines by allowing variable-speed operation and improving
controllability. One of the most common solid-state devices utilized in wind turbine generators is
the insulated-gate bipolar transistor (IGBT). IGBTs function as fast switches that can withstand
several thousand volts and conduct current on the order of 1000 amps. In wind turbine
applications, power electronics are primarily used to rectify (convert ac to dc) and/or invert
(convert dc to ac) currents and voltages.
3.4.1.4 Squirrel Cage Induction Generator with Full Power Conversion
The controllability and operating range of the SCIG can be appreciably improved by connecting
the SCIG to the grid through a full-load frequency converter. A full-load frequency converter is a
power electronics package that rectifies the variable-frequency ac output of the generator and
then inverts it at the frequency of the grid. The addition of the full-load frequency converter
allows high bandwidth control of both the active and reactive power of the turbine. This enables
the wind turbine to contribute to the stability of the grid during normal grid operation and
extreme grid events such as faults. The full-load frequency converter also makes variable-speed
operation possible, which has benefits in reducing mechanical loading.
3.4.1.5 Doubly Fed Induction Generator
The doubly fed induction generator (DFIG) is an asynchronous machine that is currently the
most common wind turbine generator type. The stator of the DFIG is connected directly to the
grid through a transformer, similar to a SCIG. However, the generator rotor of the DFIG is
connected to a power electronics package, called a partial-load frequency converter, through slip
rings. In this configuration, a fraction of the power exchanged between the wind turbine and the
grid flows through the power electronics frequency converter, and the remaining power flows
through the stator. Similar to the full-load frequency converter previously mentioned, the partialload frequency converter allows active and reactive power to be controlled and permits variablespeed operation. The dynamic range of the DFIG is slightly less than generator topologies with
full-load frequency converters, but the cost of the power electronics is less for partial-load
conversion than for full-load conversion.
3.4.1.6 Synchronous Generator
There are several types of synchronous generators used in wind turbines; however, all of them
have an electrical or mechanical means of decoupling the speed of the turbine rotor from the
frequency of the grid. As opposed to induction generators, synchronous machines appear very
stiff to the mechanical components of the drive train. Wind turbines with synchronous generators
must have variable-speed turbine rotor operation in order to prevent severe mechanical loading
from wind gusts. The most common means of providing this decoupling is to connect a full-load
frequency converter between the generator and the grid. The operation and benefits of the fullload frequency converter previously discussed for the SCIG with full power conversion are true
for synchronous generators. Another, less common, means for decoupling the turbine rotor speed
from the grid frequency is through the use of a variable-speed gearbox. In this configuration, the
synchronous generator is connected directly to the grid, and a variable-speed gearbox is
connected between the generator and the turbine rotor.
3-13

Wind Power

The synchronous generator can either be a permanent magnet generator (PMG) or a wound rotor
synchronous generator (WRSG).
3.4.1.7 Direct-Drive Low-Speed Wound Rotor Generators
Some turbine manufacturers have chosen to avoid the reliability issues that persistently plague
gearboxes by eliminating them from the design. These direct-drive designs connect the rotor
shaft directly to a low-speed generator. Since the generator speed is inversely proportional to
pole count, direct-drive generators have multiple poles and are of large diameter. Usually the
generator housing is integrated into the nacelle frame structure.
Figure 3-8 is a schematic of an Enercon direct-drive turbine. This first-generation low-speed
generator is exemplified by the doubly fed, asynchronous induction generator with a wound
rotor. The large diameter of the generator influences the design of the nacelle.

Figure 3-8
Schematic of Enercon direct drive turbine
Source: Enercon website, http://www.enercon.dk/en/_home.htm. Retrieved September 10, 2010

3.4.1.8 Direct-Drive Low-Speed Permanent Magnet Generators


What are sometimes termed second-generation direct-drive turbines often employ fullconversion permanent magnet (PM) generators. Using permanent magnets to produce the rotor
field simplifies the mechanical design by, for example, eliminating the need for slip rings. In
addition, PM generators are generally more efficient, particularly at partial load. Other
advantages include the higher magnetic field strength and lighter weight compared to a wound
rotor [70]. Figure 3-9 shows a 1.5-MW prototype PM generator tested at NRELs National Wind
Technology Center.

3-14

Wind Power

Figure 3-9
Direct-drive 1.5-MW generator under testing at NRELs National Wind Technology Center
Source: Lee, Fingersh. NREL Photographic Information eXchange via
http://www.nrel.gov/data/pix/Jpegs/14689.jpg

One disadvantage of PM generators, however, is the higher material cost for the permanent
magnets, which are made from alloys of the rare earth elements neodymium or samarium-cobalt.
In addition, because 95% of the worlds PM material is currently supplied by China, the rapid
and widespread deployment of low-speed PM generators could lead to supply problems that
might drive up prices [71]. On the other hand, 45% of the known sources of rare earth elements
are found outside China, and some projections show that annual demand of rare earth magnets
will amount to less than 0.2% of known global supply by 2015. Other projections suggest the
global supply may be as much as 200 million metric tons.
3.4.1.9 Superconducting Low-Speed Generators
For Europe to meet its goal of installing 300 GW of wind power by 2030, the high population
density onshore means that 120 GW will need to be installed offshore. For the United States to
reach its goal of 20% wind energy by 2030, 50 GW of offshore wind power will be needed. Asia
is also developing an offshore market. To come anywhere close to meeting these combined
offshore market demands, offshore turbines rated up to 10 MW and possibly larger will most
likely be necessary. Turbines of that size can take advantage of technologies that are not
economically feasible at a smaller scale.

3-15

Wind Power

One of the technologies emerging for direct-drive turbines larger than 5 MW is the application of
high-temperature superconductors (HTSs) in the generator [72, 73]. High-temperature
superconducting generators may have the potential to produce comparatively smaller, lighter and
more efficient machines for the 5- to 10-MW turbines currently being designed for the offshore
market. NREL is sponsoring American Superconductor Corp. (ASC) in conducting research into
the economic impact and costs of a direct-drive, superconducting 10-MW generator that uses
high-temperature superconducting tape instead of copper wire for the generators rotor [74]. The
use of superconducting tape for windings results in a significantly smaller generator for the same
power rating. A smaller generator will be lighter and, because it is superconducting, its losses
will be lower than those of a conventional turbine. According to ASC, the weight of the 10-MW
superconducting generator they are building is approximately 120 metric tons. In contrast,
conventional designs of that power rating weigh 300 metric tons. Figure 3-10 shows an exploded
view of a superconducting generator.

HTS race-track coils


Rotor support
Rotor iron

Back iron
Cu stator
Figure 3-10
Example of an eight-pole, air-cooled superconducting generator
Source: Abrahamsen, A. B., et. al. [73], with permission

3.4.2 Blades and Rotor


The purpose of the blades and rotor is to convert the kinetic energy of the wind to mechanical
rotational energy that can then be converted into electricity. Airfoils along the blade are oriented
so that a component of the lift vector is directed into the rotor plane of rotation, producing torque
about the low-speed shaft at the rotor center. The remaining component of lift is perpendicular to
the plane of rotation, and produces bending loads on the blade as well as thrust loading, which
must be accommodated by the structural load path (shafts, bearings, tower-top connection, tower,
and foundation). General design goals for the blade and rotor include maximizing aerodynamic
efficiency (energy production) while minimizing loads on the blades and turbine structure. For
example, the use of thicker airfoils, particularly at the blade root, has resulted in improved
structural integrity while maintaining aerodynamic performance.

3-16

Wind Power

Figure 3-11 shows the cross-section of a typical blade construction. Aerodynamic loading is
primarily carried by the spar caps on the upper and lower airfoil locations. A sandwich-style
shell provides the aerodynamic shape and resists panel buckling. Shear webs span the upper and
lower shells. In the case of Figure 3-11, shear webs are at the front and rear of the spar caps,
effectively forming a box-beam; however, several variants on the shear web layout are in
widespread use.

Figure 3-11
Cross section of wind turbine blade
Source: DNV-GEC

Blade technology is advancing in the areas of design methods, load-mitigating technology,


materials, and manufacturing. A few examples are addressed in the following paragraphs. The
overarching objectives of these advancements are to enable increased energy capture with
reduced or mitigated loads, weight, and cost.
Design practices are continually being refined to take advantage of developments in computing
speed and cost. Three-dimensional modeling of blade shapes and finite element analysis are now
the industry standard. Computational fluid dynamics is increasingly being used, although the
majority of design work is still performed using more simplified modeling approaches.
Several technologies are being developed to reduce aerodynamic loading in blades associated with
turbulence or other transient changes in wind. Advanced control methods are being developed to
independently pitch the blades. Several manufacturers and research institutions are working on
small fast-acting devices to control the aerodynamic condition of the blade via active control
technology. Passive load control is also being investigated. One of the major areas of interest is in
bend-twist coupling, where an increase in aerodynamic loads will cause the blade to twist in such
a way as to reduce the loading. This is shown schematically in Figure 3-12, where the orientation of
the fibers is used to achieve the desired twist coupling. At least one manufacturer has developed a
prototype rotor that uses geometry to realize bend-twist coupling. Measures must be taken to ensure
sufficient clearance between the blade tip and tower when the blade is under aerodynamic loading.
These measures may include a combination of increased rotor overhang distance, an upward tilting
of the nacelle, or a coning angle of the hub. LM Glasfiber has a patented approach of adding precurve to the blade so that the tip clearance is increased prior to loading.
3-17

Wind Power

Figure 3-12
Bend-twist coupling achieved by fiber orientation
Source: Laird, D. Sandia National Laboratories 2006 Wind Turbine Blade Workshop

The growth in blade weight with increased length is nonlinear. Generally, the weight increases
faster than the blade length. This means that, without material innovations, the average rating of
a utility-scale wind turbine will reach an upper limit unless the growth of blade weight can be
controlled. Advancements in materials are ongoing to reduce blade weight and cost. Per unit
weight, carbon fibers are significantly stronger and stiffer than glass fibers; however, carbon
fibers are much more expensive. As a result, the majority of blade manufacturers use fiberglass.
Notable exceptions are Vestas and Gamesa, which use carbon fiber in the spar caps of their
largest blades.
Blades are typically constructed from fiber-reinforced plastic (FRP) materials. Strength-perweight and stiffness-per-weight are both improved when the amount of load-carrying fibers is
increased relative to the plastic matrix. Newer manufacturing processes have been introduced
to increase the fiber content of the blade laminate. Nearly all new production is done with either
vacuum assisted resin transfer molding (VARTM) or the use of preimpregnated fabrics
(prepreg), rather than the traditional wet layup process. One manufacturer has developed an
innovative process in which a complete blade is produced in a single VARTM process, rather
than separate manufacture of skins and shear webs followed by a secondary assembly step. The
approach eliminates the need for bond lines, reducing weight and increasing structural integrity.
Several approaches have been considered to facilitate the transport and erection of
multimegawatt-scale blades. Several manufacturers are evaluating or developing a mid-span
joint in the blade so it can be manufactured and shipped in parts prior to assembly at the site.

3-18

Wind Power

3.4.2.1 Passive Aerodynamic Control


Passive aerodynamic devices have been used by manufacturers for many years to fine tune the
performance of rotor blades. They include vortex generators, shown in Figure 3-13, leading edge
slats tall strips, spoilers, and winglets [75]. Vortex generators and leading-edge slats affect what
is known as the boundary layer over an airfoil section, which in turn affects lift and drag
characteristics. The general effect of this control method is to delay the onset of trailing-edge
separation and subsequent stall, and thereby increase lift.

Figure 3-13
Vortex generators (counter-rotating array)

Fences and winglets would be used to mitigate the effects of three-dimensional flow along the
blade, thus improving predictability and performance. Fences would be deployed on inboard
sections where three-dimensional flow is strongest and can potentially interfere with the section
further outboard. Winglets mitigate tip loss effects by blocking the flow around the tip from the
high-pressure surface of the blade to the low-pressure side. Winglets may have an impact on
acoustic emissions as well.
Applying passive aerodynamic devices on the inboard sections of the blade could yield increases
in annual energy output on the order of 1% to 3% without adding much to the cost of the blade.
3.4.2.2 Active Aerodynamic Control
As with passive aerodynamic devices, active devices also fall into two major categories:
boundary layer control and camber modification. The technologies that fall under this heading
include leading edge slats, active vortex generators, plasma actuators, and boundary layer
suction/blowing. Circulation control is a technology that does not strictly fit in the category of
boundary layer control, but has a similar effect on lift characteristics. The feasibility of using
active boundary layer modification of local airfoil sections to control turbine loads appears less
advantageous.

3-19

Wind Power

3.4.2.3 Stealth Rotor Blade


Aviation radar interference is an increasingly common issue that developers face when planning
and developing new wind projects. Wind turbines can reflect radar signals, creating false returns
and blind spots that can sometimes appear on the operators screens. Aviation, defense, and
weather radar systemsparticularly older systemscan be affected. This issue has the potential
to hold up the installation of hundreds of megawatts of wind development, both in the United
States and in Europe. Manufacturers are investigating ways to reduce the interference footprint
caused by wind turbines and wind farms. Vestas and technology services company QinetiQ are
working to develop a stealth rotor blade that will be less visible to radar by applying a 5-mm
coating to the blades that absorbs incoming radar signals, similar to the coatings used on stealth
aircraft. [76, 77] However, because the coating adds 1,200 kg to the weight of a large turbine
blade, Vestas is working to incorporate the coating into the fiberglass-reinforced epoxy and
plastic foam of the rotor blades composite structure.
Alternative approaches to reducing the impacts of wind turbines on radar involve developments
in the signal-producing algorithm used by the radar system, as well as transitioning from analog
radar stations to more modern digital equipment. Such upgrades will occur at some point in the
future, but it is unclear when or who will ultimately bear the upgrade costs.
3.4.2.4 Smart Blades
Smart blades, also known as intelligent rotors, are wind turbine blades that are instrumented
for active load measurement, control, and damage detection. The idea is to continuously monitor
and control the performance and loads of a wind turbine at the source where the wind meets the
rotor. This is a significant area of basic research at organizations such as Sandia National
Laboratories.
In general, the use of sensors for active measurement and control is increasing. Potential benefits
of smart blade technology include the ability to provide real-time information to the control
system to prevent catastrophic wind turbine damage. Smart blade systems are complementary
to active aerodynamic control systems, providing real-time sensor data about the state of the
rotor needed to properly adjust control surfaces such as flaps or microtabs.
Merging a smart monitoring and control system with active aerodynamic devices such as flaps
or ailerons leads to a smart turbine that can adapt and respond in real time to fluctuations in the
wind speed at the rotor blades.
3.4.2.5 Flatback Airfoils
Using a systems approach to simultaneously optimize manufacturing and the structural and
aerodynamic performance of rotor blades, researchers at Sandia National Laboratories have
developed a design that employs flatback airfoils along the inboard portion of the blade. Flatback
airfoils are characterized by a blunt trailing edge in place of the conventional trailing edge that
tapers to a point.
Compared with a blade with the same thickness but a sharp trailing edge, blades that incorporate
flatback airfoils are likely to be stronger and lighter, generate more lift along the inboard section,
and have less sensitivity to blade soiling. The flatback geometry simplifies blade construction
and the truncated blade width (also known as chord length) makes them easier to transport.
3-20

Wind Power

3.4.2.6 Blade Manufacturing Processes


Demand for an increased number of wind turbine blades that are manufactured more quickly has
spurred innovation in the materials and process sector. Current blade manufacturing strategies
are both labor and capital intensive. As a result, many manufacturers are looking for automated
solutions.
Industrial automation companies have been applying their knowledge of aerospace automation to
the wind industry in recent years. Robots are used to lay dry tows or strips of woven fiberglass
cloth in preparation for them to be sprayed, vacuum-bagged, resin infused and cured. Using this
process, the Spanish company MTorres has a goal of laying down 2,000 lb (907 kg) of material
per hour in order to complete one blade every eight hours. MTorres uses proprietary software to
provide instructions to the lamination machine controller, which generates and executes machine
codes for carrying out the lamination process. The automation technology is still in development,
with plans to begin production as early as 2011, possibly in the United States.
3.4.2.7 Blade Transport and Shipping
Several approaches have been considered to facilitate the transport and erection of multimegawatt-scale blades. Several manufacturers are evaluating or developing a mid-span joint in
the blade so it can be manufactured and shipped in parts prior to assembly at the site.
3.4.2.8 Blade Test Facilities
With the continuing advancements in blade technology, there is increased interest in blade
research and test facilities in the United States. In collaboration with NREL, two new structural
test faculties are planned: one in Boston, Massachusetts, and one on the Texas Gulf Coast.
3.4.3 Yaw Systems
All horizontal-axis turbines use some type of system to rotate (yaw) the rotor into the wind.
Older turbine designs with downwind rotor configurations, and some small turbines with tailfins,
rely on the balance of wind forces to passively align the rotor to the wind direction, but
all upwind turbines use an active system. The most common arrangement uses a large-diameter
rolling-element turntable bearing (slewing ring) mounted to the top of the tower. The nacelle
frame is mounted to the rotating ring of this bearing.
The nacelle is rotated about the tower top with multiple yaw drives, which are mounted in the
nacelle frame and engage gear teeth in the fixed bearing ring. Most yaw drives are made up of
multistage planetary or worm-gear units driven by electric motors, although some turbines utilize
hydraulic motors instead. The number of yaw drives required varies with the size of the turbine,
from two for a typical sub-megawatt-class turbine to as many as eight for the largest
multimegawatt-class turbines.
The speed of yaw rotation is very slow, on the order of tens of minutes for one full rotation.
Yaw alignment is detected using the wind vane mounted to the top of the nacelle; when the rotor
is not aligned directly into the wind, the vane will be offset from the rotor axis of rotation,
indicating a yaw error. The control system then starts and stops the yaw motors in the correct
direction until the yaw error is below a specified threshold. To keep the turbine from constantly
yawing in response to short-term variations in the wind speed, algorithms are used to filter the
yaw error measurement.
3-21

Wind Power

The aerodynamic forces on the rotor can cause very large yaw torques. Early turbine designs
used brakes on the yaw motors to counteract this torque and keep the rotor in position.
Experience has shown that these forces change rapidly, and over time can damage the yaw and
cause wear on the yaw gear teeth. Modern designs incorporate friction brakes on the nacelle
frame itself. These are often hydraulically actuated so that the friction can be at least partially
released when the yaw drives are active.
An alternative arrangement for the yaw system uses a sliding bearing instead of a rolling-element
bearing. The sliding elements are low-friction polymer pads that are arranged about a machined
steel flange at the tower top. This arrangement provides both support for the nacelle frame and
friction to hold the rotor in position, and the yaw drives have to be over-sized to overcome this
friction.
Because electric yaw drive motors are started and stopped with contactors, the yaw movement
is sudden and can subject the yaw drives to very high forces. At least one manufacturer is
supplying variable-speed yaw drives to allow a smooth start and controlled forces. Another
common challenge with the conventional arrangement is wear of the teeth on the stationary yaw
gear. Some manufacturers are installing automatic greasing systems to ensure that the gear teeth
are adequately coated with lubricant. One manufacturer has sidestepped this problem entirely,
and has designed a unique hydraulically driven yaw system that uses a series of cylinders and
linkages.
3.4.4 Sensors
The most critical sensors on a wind turbine are the speed sensors on the rotor and generator
shafts. These sensors are the primary defense against excessive turbine over-speed, which can
be extremely destructive. Other critical sensors include the wind speed and direction sensors,
vibration sensors, sensors that detect when the electrical cables from the nacelle in the tower
are wrapped too tightly, temperature sensors on components that are susceptible to temperature
extremes, and pressure and flow sensors for the cooling and lubrication fluid systems. Voltage,
current, and power sensors monitor the electrical system and utility grid.
The evolution of wind turbine sensors has followed that of standard industrial controls. Many
modern sensors provide digital communications on standard computer network protocols such as
CANBUS. Other than the evolution from analog to digital sensors, there are only a few significant
advances in the sensors used on wind turbines. One of these advances is the use of ultrasonic wind
measurement sensors. These sensors provide wind speed, direction and temperature from a single
unit with no moving parts and a single serial cable. Although more expensive than conventional
sensors, ultrasonic sensors offer improved reliability and simplicity of installation. Ultrasonic
sensors can be heated more effectively than standard sensors; however, the data sensors may
experience additional fouling in conditions with a high degree of dust or precipitation.
3.4.4.1 Optical Strain Gages
A sensor that is being introduced into some of the larger, more complex turbines is the optical
strain gage. These sensors are built into the blades for condition monitoring and in conjunction
with blade pitch mechanisms for structural load alleviation. The long-term reliability of these
sensors has not yet been proven, but many major turbine vendors are experimenting with them
and it is possible that some have included them in production turbines.
3-22

Wind Power

3.4.4.2 Wind MeasurementsLIDAR


Active research projects involving nacelle-mounted wind speed sensors use light detection and
ranging (LIDAR) as one of the primary technologies under consideration. These optical remote
sensing units are designed to analyze the oncoming wind, enabling more informed control
decisions that will optimize power output and minimize structural loads. Typically, the wind
speed and direction are measured using sensors mounted on top of the nacelle, downwind of the
rotor. As a result, the wind speed and direction measurements slightly lag the actual wind
conditions at the rotor. Wind turbine rotor operating efficiency decreases by about 0.5% to 1%
for every degree of misalignment with the oncoming wind. Field testing is under way on LIDAR
systems that use lasers to measure the wind speed and direction up to 300 m in front of the rotor,
providing a 20-second advanced notice of the instantaneous wind speed and direction at hubheight. This advanced notice provides sufficient time to orient the nacelle and adjust blade angles
to improve energy capture and also reduce peak loads. One LIDAR system manufacturer claims
that addition of the LIDAR system could increase energy capture by 10%. In addition, the
nacelle-mounted LIDAR systems use fiber optics instead mirrors and are less expensive, more
robust, and less susceptible to misalignment.
Although the LIDAR forecasting systems do provide valuable wind information to the control
system, they add expense and complexity to the turbine. Nacelle-mounted LIDAR systems are
estimated to cost from $100,000 to $200,000, which would be 7% to 10% of the turbine capital
cost, depending on turbine size. In addition, because measuring the inflow specific to one
machine would not be useful for improving rotor performance at a neighboring machine, each
wind turbine requires its own LIDAR system. Also, the addition of the nacelle-mounted LIDAR
hardware adds complexity and could negatively impact both reliability and maintenance costs.
3.4.4.3 Condition Monitoring
In an effort to properly maintain wind turbine equipment, minimize plant downtime and maximize
energy generation, modern wind turbines increasingly use condition monitoring (CM) systems that
provide diagnostic data on the physical state of the equipment to identify and prevent potentially
serious failures. Some current CM techniques include bore-scopes, blade monitoring, oil analysis,
vibration monitoring, and manual field observation.
Different CM systems can be run while the wind turbine is on-line or off-line. By using on-line
CM systems, operators can monitor real-time operational data. Collected data can be used to
identify the environmental conditions that exist prior to CM events, as well as the resulting actual
loading impacts and specific gear failures. Some on-line CM methods include vibration analysis,
oil particle counting, and integrated gear and bearing analysis tied into a central monitoring
service. Off-line CM techniques require plant downtime, so they can be costly and labor
intensive. Off-line CM activities include strain and structural testing, recording torque and shear
measurements, and sampling oil and other turbine fluids. To keep costs down, when the turbine
off-line non-destructive CM testing techniques such as ultrasound, thermography and infrared
testing are used.
Improvements to CM systems are continuously being developed. Many of the new technologies
include remotely monitoring and responding to alarms, developing new algorithms to identify
flaws, and automating even more of the O&M activities. Additional information is provided in the
2010 EPRI report, Wind Turbine Blade Structural Health Monitoring, Methods and Benefits [69].
3-23

Wind Power

3.4.5 Controls
The following are the primary objectives of a wind turbine controller:

Keep the turbine safe and within its design load envelope.

Maximize the turbine energy production.

Make supervisory decisions, such as when to start, stop, or yaw the turbine.

Detect abnormal conditions such as high component temperatures and take the appropriate
action.

The controller achieves these objectives via a combination of sensors, software, hardware logic,
and actuators. The controller takes inputs from sensors and issues commands to actuators based
on a predetermined set of parameters. Controls may be divided into several subsystems as
follows:

Main turbine controller

Blade pitch system and controller(s)

Generator/power electronics controller

Safety system

The main controller is responsible for most supervisory decisions, fault detection, and providing
input into the closed-loop controls of essential subsystems. The main controller hardware
consists of a central processor, input/output modules, and communication interfaces. These are
typically mounted in one or more racks in the nacelle and/or a cabinet in the tower base and are
linked with a communications cable in the tower. There is often a user interface panel at the
tower base for maintenance personnel to manually control the turbine.
Blade pitch controllers on pitch-regulated turbines typically accept pitch commands from the
main controller and run their own closed-loop control algorithms to rapidly meet the commanded
pitch positions. In addition to blade position commands, the pitch controllers report other data
back to the main controller such as status, motor voltage, current, and temperature. This
information is typically communicated digitally over slip rings. Some turbine manufacturers are
implementing independent blade pitch systems, with controllers for each blade. The primary
advantage of independent blade pitch is to ensure that any single-point failure within the turbine
will only disable one blade pitch system, allowing the other blade pitch systems to remain
functional and capable of stopping the turbine. Independent blade pitch also allows reductions
in the size and cost of mechanical braking systems, the benefits of which have an effect on other
turbine components such as the gearbox.
The generator and power electronics controls for variable-speed turbines are a subsystem similar
to the pitch controller. Their primary function is to accept generator torque commands from the
main controller and rapidly meet them. This controller also communicates status information to
the main controller over a digital communication link.
The turbine safety system is not part of the control system and is actually super-ordinate to
the main controller as a design guideline, which allows the safety system to override the main
controller if necessary to protect the turbine. The safety system is often a series of relays that are
3-24

Wind Power

connected to critical sensors such as rotor and generator speed, vibrations, utility voltage, and
the yaw system. When any of these sensors measures an alarm-level signal, the relays open
(or close) and trigger an emergency stop using blade pitch and/or mechanical brakes.
Wind turbine controls have followed the trends for controls in other industries, namely the increase
in performance and decrease in cost of electronics and computing power. In the early days of wind
turbine development, many of the controls used hard-wired logic with sensors and relays and a
minimum of actual computing power. Modern turbines use controllers based on high-speed CPUs
with a considerable amount of memory and high-speed communications from well known vendors
such as Intel and Motorola. Turbine vendors use a variety of hardware depending on their
preference. Many companies have vertically integrated with controller vendors, others use off-theshelf control hardware from independent vendors, and still others use custom-developed controller
hardware. Generally, each turbine manufacturer provides some degree of proprietary software.
As turbines have grown larger in size, the relative per-unit cost of controllers has decreased;
however, their absolute cost has increased because of the greater complexity of most modern
turbines. The development time and costs for control hardware and software have also increased
considerably to support turbine control sophistication, plus the supervisory control and data
acquisition (SCADA) systems and other remote debugging and operations features that function
over the Internet.
The future development of wind turbine controls is focused primarily on the software and
algorithms used to operate, diagnose, and maintain the turbines. In particular, condition
monitoring features are becoming more common but require considerable sensors, algorithms,
bandwidth, and data storage.
3.4.6 SCADA Data Collection and Transmittal
SCADA systems are used on all commercial wind farms. A SCADA system provides
information on turbine operating status, alarms and warnings, and performance parameters such
as rotational speed, power output, and line current and voltage. All systems allow the ability to
remotely acknowledge and reset noncritical faults and to change some parameters in the control
system, such as maximum output power.
Most turbine manufacturers provide a SCADA interface that collects the information from each
turbine controller and displays the data in tabular and graphical format. Available displays
normally include an overall layout of the farm with indicators for the status of each turbine,
meteorological conditions, indicators for aggregate power output, total delivered megawatthours, and event and alarm logs. Often the user can access information at the individual turbine
level such as gearbox temperature or pitch position.
Although wind turbine controllers update at a rate of around 20 Hz, SCADA systems usually poll
the data at a much slower rate, on the order of seconds. Historical data, such as wind speed and
power output, are stored as the average, minimum, maximum, and standard deviation over a 10minute period. Some systems allow storage of selected parameters at a higher rate. In order to
increase visibility of project performance, some owners are directly importing higher-resolution
data from the turbine controllers into a time-based enterprise system that allows more options for
data analysis. Several third-party SCADAs are also available that can interface directly with the
turbine controller.
3-25

Wind Power

3.4.7 Drive Train


The wind turbine drive train performs two functions. First, it supports the weight of the rotor and
the thrust and yawing loads introduced by the blades. Second, it transmits the power-producing
torque loads from the rotor to the generator.
3.4.7.1 Drive Trains with Gearboxes
The vast majority of wind turbines are designed around a common drive train configuration,
which consists of a long main shaft and main bearing at the front of the nacelle, and a generator
at the rear of the nacelle. A gearbox is situated in the middle to increase the low rotation speed of
the rotor to match that of the generator. All of these components are laid out on the nacelle
frame, which is made from cast iron or fabricated steel or, often, a combination of both. In this
configuration, the high-speed shaft of the gearbox is connected to the generator with an
articulating coupling to compensate for inevitable misalignment. The low-speed section of the
gearbox is connected to the main shaft with a fixed compression-style shaft coupling.
There are two methods of mounting the gearbox to the nacelle frame. The most common is the
three-point design, where the main shaft is supported by one spherical rolling-element bearing at
the very front of the machine to carry the thrust and weight of the rotor, and the gearbox itself is
pinned to the frame on both sides to counteract the rotor torque and the overturning load from
the rotor. The second method for supporting the gearbox is referred to as the four-point
configuration. The main shaft is supported by two main bearings, one at the front and one at the
other end of the shaft, in front of the gearbox. The gearbox is still pinned to the nacelle frame on
both sides, but only to prevent the gearbox case from rotating; the overturning loads from the
overhang of the rotor are carried by the second main shaft bearing.
Because wind turbine generators are traditionally asynchronous (induction) machines connected
to the grid, they must operate at high speeds (e.g., 1200 or 1800 rpm) compared to the rotor.
This is true for variable-speed turbines also, although the speed range may vary up to 50% from
nominal. Because the rotor speed is usually between 10 and 20 rpm for a megawatt-class turbine,
the gearbox must then increase the rotor speed by anywhere from 60 to 110 times. Almost all
gearboxes are based on a three-stage design in order to achieve this high ratio and carry the
torque loads efficiently. Figure 3-14 shows an example of a typical three-stage gearbox with a
planetary low-speed stage and two parallel-shaft stages that all reside within a common gear
housing.

3-26

Wind Power

Figure 3-14
Gearbox internal schematic showing one planetary and two parallel-shaft stages

Wind turbine gearboxes have in general been a weak point of the turbine drive train. Much
research has gone into understanding why gearboxes fail and how gearbox life can be extended.
Improvements have been realized in design tools to analyze how the gears, bearings, shafts, and
cases respond under load, and best practices for design manufacture and maintenance have
recently been standardized. New techniques for improving the carrying capacity of gearboxes
include flexible planet gear supports, integrated planet gear bearings, improved lubrication
delivery and super-finishing of gear teeth.
3.4.7.2 Hydrodynamic Fluid Coupling
One exception to the common gearbox configuration that has recently been introduced at a large
scale is the hydrodynamic fluid coupling, placed in the drive train between the rotor and the
generator. Figure 3-15 is a schematic of the WinDrive hydrodynamic coupling developed and
commercialized by Voith and in service on some DeWind turbine models.

3-27

Wind Power

Figure 3-15
Cross section of hydrodynamic drive system for a wind turbine
Source: Voith Brochure, The WinDriveAn Innovative Drive Train Concept for Wind Turbines,
retrieved September 2010 via www.voithturbo.com/windrive_publications.php3

The hydrodynamic coupling combines the characteristics of a mechanically geared system with a
fluid/hydraulic system. It converts the continuously varying speed and torque of the rotor to a
constant-speed output to drive the generator. The fluid subsystem is continuously variable and
control is maintained through an adjustable guide vane within the hydrodynamic unit. In effect,
this system operates with a continuously variable gearbox ratio. While the addition of a
hydrodynamic coupling increases the complexity of the drive train relative to a conventional
drive train, this system allows the generator to operate at synchronous speed while the rotor turns
at variable speed without sophisticated power electronics. In addition, in the United States, the
technology enables variable-speed operation without the need to pay licensing fees associated
with patents on variable-speed technology.
3.4.7.3 Direct Drive Train without Gearbox
Some turbine manufacturers have chosen to avoid gearboxes entirely and connect the rotor shaft
directly to a low-speed generator. Figure 3-16 is an electrical drawing of a direct-drive turbine.
As discussed in an earlier section, the generator speed is inversely proportional to pole count,
and direct-drive generators have multiple poles and are of large diameter. For a megawatt-class
turbine, manufacture and shipping of these direct-drive generators is expensive and a logistical
challenge. Several manufacturers have made a compromise with the medium-speed design,
using a single-stage gearbox unit with a generator configuration that is somewhere between
conventional and the direct-drive design.

3-28

Wind Power

SG

Figure 3-16
Electrical diagram of a typical direct-drive turbine

3.4.8 Foundation
The turbine foundation must support the weight of the turbine, and also prevent the turbine from
overturning due to the high thrust loads from the rotor. Foundation design is always specific to
the geological conditions at the site. The turbine manufacturer will provide the design peak and
fatigue loads to the foundation designer who will consider the appropriate soil conditions,
climate, and earthquake potential. The approval of a licensed civil engineer is normally needed.
The common types of foundations for wind turbines with tubular steel towers are reinforced
concrete structures. The turbine tower is usually bolted to the foundation with a circular pattern
of fasteners on the inside and outside of the tower. The spread-foot design is shaped like an
inverted T and uses a large-diameter pad that distributes the turbine over-turning loads over
a large area.
Another design that is common in North America consists of a hollow cylinder of pre-stressed
concrete that is sunk into an excavation and then back-filled with a slurry of soil, sand, and
cement. With this design, the foundation walls resist the over-turning loads by bearing against
the sides of the excavation. A much smaller quantity of concrete is required than for the spreadfoot design, but the excavation can be problematic in some geological conditions.
A foundation design used in Australia is based on multiple prestressed concrete pilings that
are arranged in a circle and joined to a foundation ring at the surface. Similar designs are used in
arctic regions with permafrost ground conditions. The steel pilings are installed by drilling a hole
deep into the ground, placing the piling, then backfilling the hole with a slurry of soil and water,
which then freezes and holds the piling in place. The pilings include thermal siphons that
circulate refrigerant solution to ensure the permafrost temperature is maintained. The base of the
tower is also elevated approximately 1.5 m above ground level to facilitate the dissipation of
heat.

3-29

Wind Power

3.4.9 Tower
3.4.9.1 Conventional Steel Towers
All wind turbine towers installed in North America today are fabricated towers. The tower is
fabricated from rolled steel plate formed into welded tubular sections with flanged ends that are
bolted together on site. The towers are painted inside and out. Tubular towers are preferred over
steel lattice towers for aesthetic and practical reasonsaccess to the nacelle is safer and more
convenient with a ladder mounted inside an enclosed tower, and the bottom of a tubular tower
serves as an enclosed location for switchgear and controls. However, the maximum height of
a tubular tower is limited by the diameter of the base section. Most modern wind turbines are
mounted on 60- to 80-m towers.
Currently, all tubular towers are fabricated in a manufacturing facility and shipped to the site by
rail or truck. This limits the base diameter to approximately 4.1 m. To overcome this limitation,
some very tall prototype turbines have been installed on lattice towers that can be assembled on
site. One company is introducing a unique version of a lattice tower that is wrapped in a fabric
skin, offering some of the protection and visual appeal of a tubular tower. Another company is
reintroducing a technology that was used for kilowatt-class turbines in years past, in which the
tubular steel tower is fabricated in longitudinal sections and assembled into tube sections on-site,
reducing the shipping size of the parts.
In Europe, another approach has been developed to deal with the need for taller towers and
the high cost of manufacturing and transporting large-diameter steel tower sections. Several
manufacturers are making at least the base section of the tower out of reinforced concrete. The
concrete section itself is made of short cylindrical sections that are cast in moulds and stacked
on-site. Given that concrete is also required for the foundation, this approach may prove to be
economically feasible in North America.
3.4.9.2 Concrete and Hybrid Towers
Some of the most innovative developments in wind turbine tower design over the past 10 years
feature the use of reinforced concrete. The main drivers for these innovations are the limitations
on base diameter of the traditional tubular steel towers (due to bridge clearances and other
transportation restrictions) and dramatically increasing steel prices worldwide in recent years.
Figure 3-17 shows several concepts of hybrid and concrete towers. There are many advantages to
using concrete for wind turbine towers. The cost of the cement and rebar needed to construct a
concrete tower is considerably less than the cost of steel for an equivalent structure. In addition,
unlike steel towers that must be shipped in tubular segments, concrete tower sections can be
either cast in situ (e.g., concepts H2 or H3) or precast in arc segments that are manageably sized
to meet transportation requirements (e.g., concept H1).

3-30

Wind Power

Steel
tower

Pre-str.
concrete
section
H1

H2
Hybrid tower concepts

H3

Pre-stressed
concrete

Figure 3-17
Example hybrid steel/concrete and concrete towers

The arc segments can then be combined to form circular sections that are stacked on each other
at the site. Concrete towers constructed in this way have no restriction on base diameter imposed
by external circumstances, so design optimization is possible. This has implications both for the
foundation design as well as tower height. Concrete towers can be not only wider at the base but
also taller than steel towers, potentially increasing the power output of a turbine. Also, the
maintenance on concrete towers is less than that for steel, which may need to be repainted
periodically to prevent corrosion. Finally, designing a concrete tower for longer life adds little to
the cost of construction.
3.4.9.3 Integrated Towers and Foundations
Because concrete towers or hybrids tend to have a wider base than steel towers, foundation
pressure is significantly lower than for a steel tower. This simplifies the foundation technology
and reduces the amount of required concrete by 70%. For example, a basic annular ring 1 m
thick may be sufficient for the foundation, depending on the soil conditions. The foundation for a
comparable steel tubular tower would need to be several meters thick, requiring a significantly
more complicated pour during construction.
3.4.9.4 Tall Towers
Tall towers (100 m and taller) reach potentially higher speed winds and increase annual wind
energy generation relative to shorter towers. They are geared primarily for the onshore wind
market, where the influence of the boundary layer is generally more pronounced than offshore.
The boundary layer is the layer of air between ground level, where the wind speed is near zero,
to the elevation above which the wind speed is no longer influenced by surface friction or
3-31

Wind Power

obstacles such as trees, uneven terrain and buildings. The height of the boundary layer varies but,
generally speaking, wind turbines operate within this layer. The variation in wind speed with
height above ground is often referred to as wind shear. When wind shear is positive, the taller a
wind turbine is, the higher the average wind speed. This effect is diminished over water, where
the surface friction is considerably lower than on land, and where there are few or no obstacles to
disturb the air flow.
The ability to erect taller towers for land-based installations can lead to significant increases in
annual energy production (AEP), given the appropriate wind conditions. A study by NREL
suggests that increasing tower height by 20 m can increase AEP by 20% for sites with positive
wind shear. Because of this, taller towers can potentially open up vast new areas to development
that were previously uneconomic. Also, as turbine blade length increases, taller towers are
needed to keep the entire rotor sufficiently elevated for good performance.
3.4.9.5 Two-Bladed Rotors
Two-bladed wind turbines have been of particular interest in the United States due to the
perception that they could result in a lighter weight, lower-cost turbine than more traditional
three-bladed designs. Over the past three decades significant effort has been made to develop and
bring to market two-bladed rotors. In the 1980s, the U.S. Department of Energy (DOE)
sponsored research and development of multi-megawatt two-bladed turbines. Several companies
manufactured two-bladed turbines in the 1990s and early 2000s, and many configuration
variations were tried, including upwind, downwind, stall regulated, variable blade pitch, fixed
speed, variable speed, independently hinged flapping blades, and power output ranging from 200
kW to 4 MW. However, in todays U.S. onshore wind turbine market, only Nordic Windpower is
building a utility-scale, two-bladed turbine. Nordics machine is a 1-MW upwind, stall-regulated,
teetered turbine. To date Nordic has experienced very limited success in the market. Two-bladed
designs may decrease rotor cost, but they can have higher equipment loading and operational
noise as well as other disadvantages, as discussed later.
Two-bladed turbines offer some advantages as well as a few drawbacks. To achieve the same
aerodynamic performance, a two-bladed rotor will have approximately the same total solidity as
an equivalent three-bladed rotor (in a simple analysis). In that case the blade chord and thickness
are roughly 1.5 times that of the three-bladed rotor, and total rotor mass can be reduced by more
than one third. Greater structural efficiency and lighter weight as compared with three-bladed
rotors is often assumed to be a primary advantage of two-bladed rotor technology. However, for
onshore turbines, which are already constrained by blade transportation limits, the wider blade
chord is a disadvantage.
Equivalent performance of a two-bladed versus three-bladed rotors is achieved by increasing the
rotor speed of the two-bladed machine. Increased speed improves the dynamics of the twobladed rotor and is an advantage for low- and medium-speed generators. However, higher tip
speed also means increased noise, which can be a disadvantage for onshore turbines.

3-32

Wind Power

Many two-bladed rotors employ teeter, whereby the hub is allowed to seesaw in and out of the
rotor plane. Ideally, teetering means that unsteady blade bending loads are reduced or eliminated,
reducing loads throughout the turbine. In theory, two-bladed turbines can be lighter and less
costly than three-bladed turbines. In practice, however, because teeter motion is necessarily
limited (typically to several degrees), during operation there are times when blades reach the
teeter limit and the benefits of teeter are essentially erased.
Dynamically, the two-bladed, teetered rotor is more challenging than the three-bladed rotor. The
symmetry of the three-bladed rotor means its operating dynamics are easier to predict than those
of the two-bladed rotor, which lacks inertial symmetry. The application of advanced controls,
which have been in development for three-bladed rotors for many years, could help two-bladed
rotors achieve greater weight and cost savings.
Another perceived disadvantage of two-bladed rotors is the visual asymmetry, as compared to
the balanced view of a three-bladed rotor. However, the slower rotational speed of modern wind
turbines is thought to mitigate that objection to some degree.
While the current market for onshore two-bladed turbines is quite limited, some of the perceived
drawbacks do not apply in the offshore market. Visual and noise objections, for example, are not
likely to be issues offshore. There is significantly more development of offshore two-bladed
concepts rather than onshore.
Several companies are currently developing two-bladed turbines for the offshore market. One is
the Dutch company 2-B, which is developing a 6-MW, two-bladed, 140-m turbine with a lattice
tower. The company claims they can reduce COE up to 45% compared to todays three-bladed
offshore technology. One advantage they cite is the ability to fully assemble and pre-test the
nacelle and rotor onshore, stack the assembled rotors on a ships deck, and ship them to the site
ready for installation. Figure 3-18 shows an artists rendering of the 2-B 6-MW turbine.

Figure 3-18
Artists rendering of the 2-B 6-MW wind turbine
Source: Windpower Monthly, November 2011

3-33

Wind Power

Danish organization Global Innovation Center has teamed with Chinas Envision Energy to
design a two-bladed, 3.6-MW, 128-m rotor, direct drive offshore wind turbine. This turbine
employs partial pitch control on the outboard half of the blades. They cite a reduction in extreme
loads of 30% compared to an equivalent full-span pitch, three-bladed rotor. It is unclear whether
the loads reduction takes into account the teeter stop limits, however. A prototype of this turbine
is currently being installed in Denmark for testing.
Chinese company Mingyang has a commercially available 3-MW offshore turbine incorporating
a 2-bladed rotor and drive train configuration designated Super Compact Drive (SCD). The
turbine was developed jointly with Aerodyn, and first produced in 2010. In 2011, Mingyang
announced completion of testing for the 6-MW design, reported to have a rotor diameter of
140 m and a tower-head mass of 205 tons. Recently however, Mingyang has reported concerns
about the reliability and suitability of their technology for offshore applications.
The United Kingdom company Condor Wind Energy is developing a 5-MW offshore turbine
with a 2-bladed rotor. The 5-MW turbine arose after extensive testing of a 1.5-MW research
turbine. The rotor speed of 20.5 rpm for this 140-m diameter rotor means that the blade tip speed
is 127 m/s, which is significantly higher than the typical three-bladed turbine, which has a blade
tip speed of 70 to 80 m/s. While the high tip speed reduces torque loads, it also increases fatigue
cycles and could wear parts more quickly. One feature of the 5-MW version is that each unit is
equipped with a helicopter pad on top of the nacelle for delivering crew and equipment to the
site. During servicing, the blades are parked horizontally, enabling safer delivery than for threebladed rotors with helipads. Condor Wind Energy purports that its two-bladed turbine
technology will yield a 45% reduction in COE. While the combination of reduced rotor cost,
reduced torque loading, and easier construction will all contribute to reducing overall COE,
based on our experience we believe a more realistic reduction is 5% to 10%.
3.4.9.6 Vertical Axis Rotors
Although utility-scale, vertical axis wind turbines (VAWTs) have all but disappeared from the
landscape over the last decade, interest in the technology has resurged in recent years. In
particular, VAWTs are being developed for use in offshore wind projects. Conceptually, VAWTs
have some distinct advantages over horizontal axis wind turbines (HAWTs). Many fewer
components and systems are requiredthere is no massive tower or nacelle, no yaw system, no
pitch system, no complex blade geometry. VAWTs are omnidirectional, and can be more easily
maintained since the drive train is closer to the ground. It is claimed that vertical axis turbines can
be fabricated more cheaply than an equivalent horizontal axis turbine. However, horizontal axis
turbines have higher aerodynamic efficiency than vertical axis machines, so the overall impact on
the cost of electricity (COE) is less clear. Historically, the disadvantages of vertical axis turbines
include high cyclic (fatigue) loading due to the blade rotating in and out of the wind twice during
each revolution, non-optimal energy capture because the lower half of the rotor is near the ground,
and high loading of the main bearing, which carries the weight of the entire rotor structure.

3-34

Wind Power

Last year, Vertiwind of France began the development and testing of an offshore prototype
VAWT, pictured in Figure 3-19. The Vertiwind concept is a 2-MW, direct-drive, permanent
magnet turbine approximately 90 m in height mounted to a triangular structure tethered to the
seabed. The low center of gravity should aid the stability of the floating structure. The company
says that the entire unit, including floating foundation, can be assembled at a port and towed out
to the site with no need for special cranes. One concern with the design is the potential for high
fatigue loading at the two points where the blades attach to the supports.

Figure 3-19
Artists rendering of the Vertiwind offshore VAWTs

Guoneng Wind Power Generation, a Chinese manufacturer, is also exploring VAWT technology
(Figure 3-20). The 1-MW prototype it installed in 2011 is designed to operate in winds from 3 to
35 m/s. This is considerably higher than the cut-out wind speed for HAWTs, which is generally
25 m/s. The turbine houses two 500-kW, permanent magnet generators. A picture of the
prototype turbine is shown in Figure 3-20. One feature that stands out is the considerable number
of supports attached to each blade, which degrades aerodynamic performance. This turbine is
mounted on a tower, bringing the rotor up to where wind speeds are higher.

3-35

Wind Power

Figure 3-20
Guoneng 1-MW, 8-bladed VAWT with 36-m blades

A third VAWT concept under development takes a different approach, eliminating the tower or
any central support altogether. The 2.6-MW Lux turbine, for which a 40-kW prototype has been
built and tested, consists of six blades connected by cables and secured with guy wires. The Lux
VAWT is targeted for either onshore or offshore applications.
3.4.10 Offshore Foundations
One of the major differences between an onshore and an offshore wind energy installation is the
more complex foundation structures required offshore. Offshore, the foundations (or substructures) are much larger because they have to extend up above the highest wave crest in
order to support the tower and the access platform at a safe distance from the impact of the
waves. Further, in addition to the wind loading, the offshore foundation/substructure is subject to
hydrodynamic loads and a harsher environment than onshore structures (with respect to
corrosion, for example).
It is therefore not surprising that the foundation/substructure is one of the elements of offshore
wind development where a lot of innovation is taking place.
3.4.10.1 General Classifications
Foundation types are often grouped according to the depth of water for which they are most
suited. Table 3-4 presents the industrys current classification for offshore wind turbine
foundations based on water depth and the total offshore wind project potential capacity in the
United States for each of those depth classifications.

3-36

Wind Power
Table 3-4
Estimated U.S. wind generation capcaity by depth
Category

Depth

Estimated Wind Capacity at Depth

Shallow water

030 meters

430 GW

Transition depth

3060 meters

541 GW

Deep water (floating foundations)

6090 meters

1,533 GW

Shallow water foundations typically consist of monopiles and gravity-based foundations.


Transition depth foundations include some monopoles, but jacketed or tripod foundation types
are primarily used. Deep-water foundations are just entering the prototype stage at this time and
consist of floating foundations anchored to the seabed by tethers.
Support structures for offshore wind turbines consist of three parts, namely the tower (usually a
steel tube with one or two segments), mounted onto the substructure (jacket, tripod, gravity etc.),
which is fixed to the sea bottom by means of the foundation (piles or lower part of gravity base
foundation). For the purpose of discussion in this report, the foundation and substructure parts of
the offshore wind tower will be included and treated as one component in the foundation
discussed. Please refer to Figure 3-21 for details. The design and utilization of floating wind
turbines is described at the end of this section.
The tower is usually designed and delivered by the supplier of the wind turbine.2 It is usually
connected to the substructure by means of a bolted flange connection. The design and delivery of
the substructure and foundation are usually the responsibility of the project owner. However, as
all components must work and fit together, close cooperation is essential between the designers
of the tower, the substructure and the foundation.

The wind turbine consists of the tower and the rotor-nacelle-assembly (RNA).

3-37

Wind Power

Figure 3-21
Offshore wind support structures

Pile foundations bear the loads by the skin friction and point pressure of piles into the ground.
Compared to their weights, piles are able to carry relatively high loads. The piles are installed by
either ramming or drilling, or a combination of the two, into the seabed. Ramming is the most
widely used installation method, in which the pile is rammed through support structures
connected to the foundation structure (pile sleeves).
The second step involves installing the supporting structure onto the piles. The connection
between the piles and the support structure of the foundation is usually a grout connection
(Figure 3-22). Some problems have been detected on grout connections used in recent offshore
wind farms since the beginning of the early 1990s, such as those in Denmark, the UK, and the
Netherlands.

3-38

Wind Power

Figure 3-22
Offshore wind monopile foundation showing grouted connection

The problems with the grout connection have included vertical slippages of the plain steel tubeto-tube connections, which has occurred over time on many wind turbine foundations.3 To avoid
this slippage, different concepts have been developed to better secure the connections between
the interconnecting tubes. One concept uses conical tubes instead of cylindrical tubesthis leads
to a positive locking connection between the higher and lower tubes.
A second concept uses so-called shear keys (horizontal ribs on the inner side of one and the outer
side of the other tube) to increase the friction between the tubes. Combinations of both shear
keys and conical tube approaches are also possible.
Another way to secure the connection between the piles and the foundation support structure is
the so-called swaging technique, an alternative to grout connections. The foundation pile inside
the jacket sleeve is deformed by means of high pressure. This creates a connection inside the pile
sleeve, providing a better mechanism to fix the jacket into the seabed.
Piles foundations for deep-water use are made of steel, and generally have pile diameters of
1 to 2 m. Monopiles can have diameters up to 6.5 m, which is the upper limit of the piling
devices available on the market. One downside to the use of pile foundations is the sound made
during construction, as the piles are rammed into the seafloor. This processpilingleads to
very high noise emissions during the erection phase of an offshore wind farm. This noise can, in
turn, pose a permitting problem, as the noise is harmful, especially to sea mammals. In some
countries, such as Germany, an upper limit for the noise level during the erection phase is
defined in the criteria of the permitting authority. Different methods for noise reduction during
piling (e.g., bubbling around the piles) have been tested, but no such method is commercially
available at this time nor has any been permitted thus far.
3

Peter Schaumann et al: Durchrutschende Grout-Verbindungen in OWEA Tragverhalten, Instandsetzung und


Optimierung. Stahlbau 79 (2010).

3-39

Wind Power

Suction bucket is an alternative solution without the noise drawbacks of the deep piles. This
foundation type uses a construction shaped like an inverted bucket underneath the legs of the
foundation structure. After the bucket touches down on the seabed, it is pressed into the seabed
through the creation of a vacuum inside the bucket. The suction has the effect of drawing the
base construction down into the seabed. The diameter of the base construction is approximately 3
to 20 m; its height is of similar magnitude.
The load capacity of this foundation is measured in the same way as for piled foundations,
namely by the skin friction of the wall area and by the point pressure. The advantage of this
foundation is that the secondary steel components (boat landing, platforms, etc.) are already
mounted on the structure and need not be fitted with additional equipment. As there are no piles
to be rammed into the seabed, the noise emissions for this foundation type are significantly
reduced. However, the steady lowering of several suction points in a 3- or 4-point foundation can
pose a challenge during offshore installation. Additionally, in soils with high bulk density or
rocky structure, it can be difficult to create sufficient suction pressure. Experience with this type
of foundation structure is very limited to date; only one wind turbine has been installed using this
foundation type, and that was in shallow water in Denmark.
As gravity-based foundations can be erected without the need for piling or drilling devices, their
installation generates very low noise. On the other hand, the transport and handling of these
heavy foundations can be challenging. If the load capacity of the seabed is low, the ground has to
be prepared before the gravity-based foundations can be deposited. This preparation requires
removing the top layer of sediment and then inserting load-bearing sand or gravel layers. In
contrast to pile foundations, gravity based foundations have an additional advantage in that they
can be dismantled completely.
The following sections examine the primary foundation types in more specific detail. For each
category, a summary of advantages and disadvantages is provided.
3.4.10.2 Monopile Foundations
In simplest terms, a monopile is like an extended wind turbine tower. The monopile foundation
structure usually consists of two partsthe actual monopile and the upper transition piece. The
pile (blue portion in Figure 3-23) is anchored to the seabed as a deep foundation pile, which ends
several meters above the mud line or shortly above the mean sea level (MSL). The upper
transition piece (yellow portion in Figure 3-23) forms the rising structure, which extends both
below and above the water level up to the tower base.

3-40

Wind Power

Figure 3-23
Monopile foundation

In a pile, a single steel pipe (or, alternatively, a reinforced concrete pipe) of large diameter
(46.5 m) is piled (or, rarely, drilled) into the ground. The transition piece is aligned with and
joined to the monopile, to extend above sea level to the desired height. Potential misalignments
between the two sections can be compensated by a grouted joint, in which the two interlocking
pipes are bonded with a high strength grout. The potential for misalignment is the first reason to
use a transition piece between the pile and the tower. The second reason is the potential damage
of the secondary steel during pile driving. This latter reason is why components such as ladders,
platforms etc. are installed on the transition piece.
Monopiles have been used in water depths as high as 20 to 25 meters, with projects currently
planning to use monopiles in depth of up to 35 to 40 m, such as that in the German Baltic Sea.
However, in greater depths and with the installation equipment available today, monopiles are
feasible only for smaller wind turbines of up to 3.6 MW. Monopile diameter increases
significantly with depth and turbine size; a manageable upper limit today is considered to be 6.5
to 7 m. Monopiles can also be used for larger turbines, but today only in shallow waters. The use
of monopiles for 5-MW+ turbines in deeper water would be possible in principle, but would
require monopile diameters that are not manageable today: the upper limit for the installation
equipment (ramming etc.) is around 7 m. Another limit is the wall thickness that would be
required for larger monopiles to guarantee adequate stiffness. At these sizes, however, other
constraints may begin to impact project size, including pile length and mass, as transport vessels
and roadway/ access limitations begin to impinge. The stiffness of the foundation is an important
dimension of the foundation design, as it must be evaluated in comparison to the bending motion
and frequency of the wind turbine itself, in order to avoid resonance effects. The thickness of the
steel plate (i.e., the wall thickness of the monopile) required for larger and stiffer foundations
may also become a constraint during the manufacturing process, as the monopile tubes are
formed from steel plates.4 The ability to shape steel of the required thickness may not be within
the capability of the monopile supplier/source.

Kimon Argyriadis (GL Renewables Certification): Floating Offshore Wind Turbines. Hamburg Offshore Wind, 6
May 2010.

3-41

Wind Power

Main Advantages and Disadvantages

Given their low complexity, monopile structures are simple to produce and less costly than other
types of offshore wind turbine foundations. At the same time, the significant wall thickness of
the pile material imposes handling challenges, for appropriate forming and welding.
Monopiles are used in more offshore wind projects than all other offshore foundation types. Due
to the large base of experience in the use of monopiles, and the availability of production from
several manufacturers in different countries, market risk is low. Unlike solutions that utilize
several piles, the soil bearing capacity is of particular importance for the monopile. This reflects
the requirement that the surrounding soils have sufficient absorption ability to accommodate the
monopiles lateral bending moment.
The movement of monopile foundations into location requires transport barges and/or
construction ships of sufficient size. Noise emissions during the construction process are very
high due to considerable noise made during the piling process. On the plus side, the smooth
surface of the monopile represents a lower risk of corrosion (from rust formed on the metal
structure caused by the aggressive salt-water conditions) than for foundation types with more
complex and bumpy constructions, such as tripods and especially jackets. Because of their shape
and large surface area, however, monopile foundations are sensitive to wave loading (i.e., the
load on the foundation caused by the action of waves). Monopiles are also sensitive to soil
conditions at the site and to scour, the removal of sediment around the foundation by the flow of
water around the base.
Promising Developments

Besides monopiles that are piled or hammered into the soil in the traditional fashion, some
research has been done on monopiles that are fabricated of concrete and drilled into the seabed.
For a research project, Ballast Nedam (the Netherlands) developed a new foundation concept,
which uses a drilling method based on the horizontal tunnel-drilling methods that are already
used onshore by Ballast Nedam.5 Advantages of this foundation type are comparable to those of
gravity-based concrete foundations, in that this foundation type is expected to be comparatively
cheap and will be less vulnerable to (steel) price fluctuations. Noise emissions during installation
as well as sensitivity against soil conditions are very low. At a water depth of ~30 m, the weight
of the foundation will be 1,450 t for a 3.6-MW wind turbine generator (WTG) and 2,200 t for a
5.0-MW WTG. Costs are estimated be in the range of ~500,000 /MW for a
3.6-MW WTG and ~400,000 /MW for a 5.0-MW WTG.
In general, this concept appears interesting for several site conditions. As practical experience is
missing so far, however, additional research is required before this option can be evaluated
further.6

http://www.offshore-energy.nl/page_10352.asp (last checked 07/25/11).


See also Gran Loman, Vattenfall (2009): Conceptual Foundation Study for Kriegers Flak Offshore Wind Farm.

3-42

Wind Power

3.4.10.3 Gravity-Based Foundations


Gravity foundations use their weight to resist wind and wave loading (Figure 3-24). They are
typically designed as flat solid foundations, but may also be designed as boxes made of steel or
reinforced concrete that, after being sunk to the seabed, are ballasted with water or sand, as
shown in Figure 3-24.

Figure 3-24
Gravity-based foundations

Gravity-based foundations are comparatively cheap (hollow concrete bodies can be filled with
water or sand), and no drilling or piling is required. Depending on the load-carrying capacity of the
seabed, extensive preparations may be required before setting the gravity-based foundation on the
ground. Studies on gravity foundations for the first offshore wind farms in Denmark have shown
that these foundation types become uneconomical compared with other foundation types like
tripods and jackets, primarily due to the exponential increases in weight and therefore material and
especially installation costs with increasing water depth, due to the technical effort and increasing
mass and costs.7 For larger turbines, a water depth of 25 m seems to be a reasonable limit, although
some experts assume that 30 or even up to 40 m may also be workable.8
Main Advantages and Disadvantages

The engineering design basis and installation experience of gravity-based foundations is well
established. Fabrication of concrete gravity-based foundations for similar structure is wellknown civil engineering. The market risk9 for this category is also considered to be medium, on
the basis that there are fewer manufacturers with experience in these designs than there are
manufacturers for monopiles. However, by virtue of their most common construction materials
(concrete), the steel price risk for gravity-based foundations is low. With no piling required,
7

Erich Hau: Windkraftanlagen. Springer, Berlin (2003).


Schaumann, P. et al: Fatigue Design bei Offshore-Windenergieanlagen. Stahlbau 73 (2004).
9
We have used the term market risk in this study to refer to the risk embedded in the need to find one or more
manufacturers able to produce the foundations just when needed, for the specific project, in an adequate time frame
and at an acceptable price.
8

3-43

Wind Power

noise emissions during the project construction are low. Overall, however, environmental impact
in general is high due to the big footprint of most gravity-based foundation concepts. Corrosion
risk is low and foundation robustness is high. Because of the heavy loads involved in
transporting and placing the gravity-based foundations, big installation devicesbarges, cranes
and other handling equipmentare required. Like monopiles, gravity-based foundations are
sensitive to prolonged wave loading, scour and also show medium to high sensitivity against soil
conditions.
3.4.10.4 Tripod Foundations
The tripod foundation is a steel structure composed of a central shaft supported by three legs and
horizontal stiffeners, as shown in Figure 3-25. This foundation type makes its connection to the
seabed through each leg of the tripod, using either driven piles or suction caissons. These piles
pass through and are then grouted to tubes at the end of the tripod legs. Particularly when
compared to the monopile, this more complex structure provides stability and stiffness in the
foundation. This type of foundation is useful in greater water depths, up to 50 m.

Figure 3-25
Tripod foundation

3-44

Wind Power

Main Advantages and Disadvantages

The track record of tripods is limited, and the fabrication is relatively complexconsisting of
few but very thick welds and noncylindrical tube structures. Based on the number of
manufacturers experienced with this type of design, the market risk is considered medium to
high. Due to the steel construction, steel price risk is also considered very high. Noise emissions
can be high, due to the piling required during the construction phase. Suction caissons have the
potential to reduce noise during construction. Given the steel construction material, corrosion
risk is medium to high, and overall robustness is medium. The size of the transport and
installation devices required for tripod foundations deserves consideration, and is less
constraining than larger foundations like the gravity-based type. Also compared with monopiles
and gravity-based foundations, tripod foundations are less sensitive to wave loading and scour,
and sensitivity to soil conditions is low.
3.4.10.5 Jacket Foundations
The jacket foundation is a kind of latticework structure made of tubular hollow sections. It will
have a square base area of roughly 25 25 m on the seabed in the case of a four-legged jacket
structure, the example shown on the right in Figure 3-26. The foundation structure narrows from
its seabed base to the wind tower base. A three-legged structure is an alternative jacket
foundation design: in this case, the base is formed by an equilateral triangle (the left-side
example in Figure 3-26). Like a tripod foundation, jacket foundations offer higher stability and
stiffness than basic monopile foundations.

Figure 3-26
Jacket foundations

3-45

Wind Power

The installation process for the jacket foundation is similar to that for the tripod foundation in
that the foundation is anchored to the seabed by piles. In general, jacket foundations have high
stiffness and can be used in greater water depths (up to at least 60 m) than that of tripod
foundations.
Promising Directions

A new type of jacket foundations, the so-called twisted jacket, has been developed by Keystone
Engineering in the United States. This design will be tested on a met mast at the 4-GW Hornsea
wind farm zone off the UK.10 The twisted jacket foundation is a variation on a monopile, in that
it has a central pile driven into the seabed and three supporting legs angled around the central
pile to provide additional stability. This design is supposed to reduce costs significantly, but no
practical experience exists so far to support that contention.
Main Advantages and Disadvantages

Although jacket foundations are well known in the oil and gas industry, their track record is low
to medium for offshore wind applications. The fabrication is of medium complexitythe twisted
jacket is fabricated with steel tubes that are typically straightforward with relatively simple but
numerous welds. The installation requires the driving of a single pile into the seabed, which may
reduce costs over those of tripod and other jacket foundation designs. The main risks of this
foundation type arise from steel prices, lack of proven design, and lack of actual experience.
Noise emissions are likely to be very high and similar to monopile foundations during the
construction phase. The overall robustness of this foundation type is considered moderate.
Installation devices for transport and construction are again considered medium. Because of their
relatively small surface areas compared to larger foundation types, the sensitivity of jacket
foundations to wave loading, scour and soil conditions is considered very low.
3.4.10.6 Suction Bucket Foundations
Suction bucket foundations have not been used in commercial offshore projects so far. The suction
bucket foundation is a bucket-like steel foundation with its opening facing downwards, which is
sucked down to the soil by vacuum. The entire soil mass enclosed by the bucket contributes to the
stability of the foundation. This design can be used in water depths up to 40 (55) m.11,12
Main Advantages and Disadvantages

The track record of suction bucket foundations is very low, with only prototypes existing so far.
The complexity of fabrication is low to medium, as the structure employs mainly cylindrical
components with no corners or extended elements. Market risk is also medium because, although
the application to offshore wind is recent, the technology itself is relatively simple and expected
to be within the capability of many manufacturers. Because of the construction material,
however, exposure to steel price risk is considered high. As no piling is required, noise emissions

10

http://www.rechargenews.com/business_area/innovation/article253198.ece (last checked 07/25/11).


Ibsen, L.B. et al. Development of the bucket foundation for offshore wind turbines, a novel principle. In:
Proceedings of the US Wind Energy Conference, 26-28 August 2003.
12
www.lorc.dk.
11

3-46

Wind Power

would be low during the construction phase. Corrosion risk is presumed to be medium, with
again the caveat that there is little actual experience with this design.
Suction bucket designs are expected to be less robust compared with piled foundations. This
reflects the possibility that air gets sucked into the bucket, potentially reducing the stability of the
foundation. The size of the installation equipment required for the suction bucket foundations is
also considered to be medium, based on the relative size of the foundation elements. A big
advantage of this foundation design is presumed to be the ease of dismantling the facility during
decommissioning. When air is pumped into the bucket, the fixed foundation is freed from its
anchor position in the ground. Like monopiles, suction bucket foundations are sensitive to wave
loading and scour. The suction bucket design is also particularly sensitive to soil conditions,
requiring homogeneous soil structures.13
3.4.10.7 Floating Foundations
So far, floating foundations exist only as prototypes. Nonetheless, it is expected that they will
play a major role in future projects where water depths are greater than 50 to 60 m. One of those
conceptsthe tension leg systemis shown in Figure 3-27. This design utilizes submerged
floating bodies to support the rising structure. These floating bodies have excess buoyancy and
are anchored to foundation points on the seabed by means of tendons.

Figure 3-27
Tension leg floating foundation

13

Other issues of potential operational concern over timesuch as long-term stability after continued wave and
ocean action, potential misalignment between structural components, etc.are only now starting to be the focus of
research efforts in Europe.

3-47

Wind Power

Main Advantages and Disadvantages

As noted, the track record of floating foundations is very limited, as only prototypes exist so far.
Fabrication complexity is high for both the floating structure and the anchors. Because
experience is so limited, market risk must be considered high. Recent design concepts show that
floating foundations can become very heavy, which also leads to a high steel price risk. As no
piling is required, noise emissions are very low during the construction phase. Corrosion risk is
deemed medium to high, based on likely construction materials. Floating foundations are
expected to be less robust compared to piled foundations. The size of installation, transport, and
handling devices required for floating foundations is also lower than for other designsa major
advantage of this approach is the simple ability to ship the construction to an offshore location.
Another big advantage is the relatively easy dismantling ability. Floating foundations are
somewhat sensitive to wave loading; at the same time, sensitivity against scour and soil
conditions are deemed to be very low.

3.5 Trends in the Turbine Supply Market


This section summarizes trends in the current turbine supply market. The primary source of
information is the wind energy World Market Update 2010 published by BTM Consult ApS in
March 2011.
A handful of manufacturers supply the majority of the worlds wind turbine demand. In terms of
market share, the leading wind turbine supplier is Vestas, based in Denmark, which captured
approximately 13% of the global market share in 2011. Goldwind moved from the fourth
position to the second position, with 9.4% market share. GE Wind holds its same place as last
year with 8.8% market share. Surprisingly, Sinovel (based in Beijing) fell from second to
seventh in the supplier listing. Just last year, Sinovel had moved from the seventh position in
2008 to second, edging out General Electric (GE). Vestas and Sinovel delivered 10,228 MW in
2010. Sinovel and Goldwind, both Chinese suppliers, took advantage of the growing Chinese
market. GEs growth was hindered by the U.S. economic downturn. Refer to Table 3-5 for the
top-10 list of suppliers for 2010.

3-48

Wind Power
Table 3-5
Top 10 suppliers 2011
Source: BTM Consult ApS, March 2012

Accu. Supplied
MW
MW
2010
2011
VESTAS (DK)
45,547
5,213
GOLDWIND (PRC)
9,055
3,789
GE WIND (US)
26,871
3,542
GAMESA (S)
21,812
3,309
ENERCON (GE)
22,644
3,188
SUZLON GROUP (IND) 17,301
3,104
SINOVEL (PRC)
10,044
2,945
UNITED POWER (PRC) 2,435
2,859
SIEMENS (DK)
13,538
2,540
MINGYANG (PRC)
1,799
1,178
Others
34,882
8,693
Total
205,927 40,358

Share
2011
%
12.9%
9.4%
8.8%
8.2%
7.9%
7.7%
7.3%
7.1%
6.3%
2.9%
21.5%
100%

Accu.
MW
2011
50,760
12,844
30,412
25,120
25,832
20,405
12,989
5,294
16,078
2,976
43,575
246,284

Despite the challenges from the tough economic climate and regulatory uncertainty in some
markets, the wind industry achieved a 5.9% year over year (YOY) growth rate in 2011, a 2.5%
increase compared with 2010. The growth can be attributed to the continued expansion of the
leading manufacturers in emerging markets such as India, Brazil, Canada, and China.
The general trends in the market for wind turbine manufactures include scaling up turbine sizes,
utilizing larger and lighter turbine blades to provide viable options for low to medium wind
speed sites. In addition to globalization, consolidation within the wind energy sector continues.
Following the acquisition of Scanwind, Darwind and DeWind by GE Energy, Hara XEMC, and
Daewoo, respectively, in 2009, AREVA has completed the purchase of Multibrid in the summer
of 2010. At the end of 2010, United Technologies Corp (UTC) completed its purchase of
Clipper, although it announced in March 2012 that it intends to sell Clipper. Although not many
mergers and acquisitions have taken place this year, under the current market conditions, it is
expected that a number of important mergers and acquisiitios will occur in Europe or the United
States as Asian companies seek to enter the expoert market in these countries.
In 2011, manufacturers extended facilities and several major wind turbine companies expanded
their research and development capabilities in the United States (Figure 3-28). However,
weakened prospects after 2012 has led to industry-wide layoffs and increased pressure on lower
profit margins for manufacturers.

3-49

Wind Power

Figure 3-28
Manufacturing plants for turbine blades and other components
Source: NREL, 2011

The offshore sector will continue to drive innovation in the industry. Further progress with
offshore wind generation is expected within the next five years, mainly in Northern Europe and
Europe. In 2011, about 470 MW of offshore capacity was installed, which represents a 67.5%
deduction from 2010. Six new wind farms (362 MW) were installed in Europe with over 90%
being installed in the United Kingdom and two projects in China (108 MW). Although the 2011
installations were modest, there are nearly 4,000 MW currently under construction in Belgium,
Denmark, the United Kingdom, and China.
An interesting feature noticed among the turbine manufacturers is their presence in international
markets. In spite of an increasing global market, some manufacturers are mainly active in only
their respective domestic markets. This is evident in all of the Chinese suppliers, whereas the
other suppliers are more global (Figure 3-29).

3-50

Wind Power
18%

Global
Supplier

Domestic
Supplier

16%

14%

Global Market Share

12%

VESTAS (DK)

10%

GE WIND (US)

GOLDWIND (PRC)

GAMESA (S)
UNITED POWER (PRC)

8%

ENERCON (GE)
SUZLON GROUP (IND)

SINOVEL (PRC)

6%

SIEMENS (DK)

4%
MINGYANG (PRC)
2%

0%
0

Source: BTM Consult - A Part of Navigant - March


2012

10

15

20

25

No. of markets (Total supply >50MW)

Figure 3-29
Top 10 suppliers market share and presence
Source: BTM Consult ApS, March 2012

In terms of technological advancements, more effort is being applied to direct drive concepts. In
total, direct drive turbines account for approximately 21% of the global supplya 1.5% increase
from 2010. Several of the smaller turbine manufacturers are gaining market share with a number
of noteworthy developments, though the recent economic downturn has somewhat dampened
expectations. Suzlon purchased a majority stake (85%) in the German turbine manufacturer
REpower in 2007, which increased Suzlons global market share from 10.5% to 13.8% (although
its market share shrank to 9% in 2008 due to poor order inflow). The company purchased a
further stake in REpower in June 2009. Suzlon also purchased the worlds second largest
gearbox supplier, Hansen Transmissions, in 2006, giving the company greater control over its
sub-component supply chain. Ecotcnia, which operates five production facilities in Spain, was
acquired by the major power generation system supplier Alstom in October 2007, representing
Alstoms first move into the wind energy sector.
In 2007, U.S.-based manufacturer Clipper Windpower began serial production of the 2.5-MW
Liberty turbine, shown in Figure 3-30. The turbine drive train consists of a forged steel main shaft
integrated with the gear housing and a bull gear that drives four PMGs arranged around the
housing. This configuration eliminates the conventional three-stage gearbox and spreads the torque
loads across the four pinion gears. The drive train is matched with advanced power electronics
allowing for variable-speed operation and providing full power quality management capability.
3-51

Wind Power

China hosts a number of new entrants into the turbine supply market. However, recent
regulations will moderate overcapacity as seen earlier this year and last year. In June 2011, The
Chinese government stopped giving preferential treatment to original equipment manufacturers
unable to produce turbines with a capacity of 2.5 MW or higher, or with less than 1 GW of
annual assembly capacity.

Figure 3-30
Clipper Windpowers 2.5-MW Liberty Series wind turbine nacelle, with four modular
permanent magnet generators
Source: Clipper Windpower Inc.

The industry has seen a trend in the globalization of wind turbine technology via licensing of
manufacturing rights. Notably, American Superconductor Corp. (ASC), known for its hightemperature superconducting products, entered the wind turbine supply market with the
acquisition in 2007 of Austria-based Windtec, a designer and licenser of complete wind turbine
systems or subcomponents. ASC sells Windtecs proprietary designs to third parties for an initial
fee plus royalty payments for each installation. Purchase of the ASC design allows new
companies to quickly enter the turbine supply market with a tested and certified product. ASC
has sold turbine design licenses to several manufacturers including Sinovel (China), AAER
(Canada), Wikov (Czech Republic), and Fuhrlnder (Germany). See section 3.13 for further
information on technical advancements that have the potential to impact the wind industry
significantly.

3-52

Wind Power

3.6 Trends in Wind Turbine and Plant Sizes


The average size of individual wind turbines installed in 2010 was 1,655 kW, an increase of 95
kW over the average size in 2009. There was only slight growth in the average turbine size in
2010,14 when 83% of individual turbines installed were in the 1,500 to 2,500 kW range, and only
8.3% were in the 750- to 1,499-kW range. Less than 10% of installations used turbines outside of
these capacity ranges. The growth in turbine sizing was in part due to the large Chinese market
that installed the mainstream 1.5-MW turbine. In California, many of the projects that utilize
kilowatt-scale turbines installed in the 1970s and 1980s are being repowered with modern
megawatt-scale turbines, with each modern turbine replacing up to 10 of the original turbines in
some cases. Many of Germanys older fleet of turbines will be repowered in similar fashion
within the next five years.
The average wind turbine size is expected to continue to increase in the short term. However,
owing to the logistical constraints of transporting and lifting large turbine components, it is
expected that land-based turbine designs will be limited to rotor diameters of about 100 m to 110
m, which corresponds to generating capacities of approximately 3 to 5 MW, although the
offshore demand continues to encourage manufacturers to develop larger machines. In addition,
developers are able to take advantage of economies of scale with larger machines to help reduce
the $/kW of installed cost.
Table 3-6 summarizes selected turbine models over 3.5 MW that were delivered to market in
2010. This is not a comprehensive list of commercially available turbines in the market but rather
an indication of the next generation of large wind turbines expected to be widely deployed.
Table 3-6
Selected onshore wind turbine suppliers and turbine models over 2.3 MW
Country

Manufacturer

Model

Rating
(kW)

Rotor
Diameter
(Meters)

Control
Scheme

Status

France

Areva WInd

M5000

5000

116

Pitch (V)

Multibrid; offshore

Germany

Bard

5.0

5000

122

Pitch (V)

Offshore

Germany

Enercon

E126 7.5 MW

7500

127

Pitch (V)

DD annular generator

Spain

Gamesa

G10X-4.5

4500

128

Pitch (V)

IGBT-inverter

Germany

REPOWER

5M

5000

126

Pitch (V)

Onshore/offshore

Germany

REPOWER

6M

6000

126

Pitch (V)

Onshore/offshore

China

Sewind

W3600

3600

116/122

Pitch (V)

Onshore/offshore

Denmark

Siemens

SWT-3.6-107

3600

107

Pitch (V)

Onshore/offshore

China

Sinovel

SL 5000

5000

128

Pitch (V)

Onshore/offshore

IGBT = Insulated gate bipolar transistor DD = Direct drive

14

BTM Consult ApS. International Wind Energy Development World Market Update 2010, March 2011.

3-53

Wind Power

In addition to an increase in individual turbine rated capacities, the land-based turbine market has
experienced an increase in the overall size of a wind power plant. The average size of wind
projects installed in 2009 was 91 MW, an approximate 10 MW over 2008 levels. In the United
States, there are more community wind turbine projects being developed, which drives down the
average wind project size. The nations second largest project, Caithness Shepherds Flat, an
845-MW project in Oregon, was completed and commissioned in September 2012. The largest
project is the 1020-MW Alta Wind Energy Center wind farm in California.

3.7 Onshore Capital and O&M Cost Trends


In 2010, the U.S. wind industry has seen a 25% to 30% decrease in turbine pricing. Because the
wind turbine constitutes 70% to 75% of the total project overnight cost, the cost reduction is
evident in the installed cost. Wind turbine price quotes have been in the range of $900 to
$1,400/kW. The primary driver for the cost decreases is the overcapacity in the market, due in
large part to increased Chinese manufacturing. Prices in the near future are expected to stay
competitive, particularly as manufacturers manage through the possible decline in the US market
if the PTC is not extended beyond 2012. One concern that could cause prices to increase of
prices in the rare earth minerals market, which may become strained. Rare earth minerals are
used in generator production. China produces 95% of the worlds rare earth and recently took a
series of measures to tighten control over the industry, citing environmental concerns and
domestic demand. Export quotas have been reduced significantly; the Chinese industries have
been consolidated and China has announced plans to build national reserves.
As turbine rotors, blade lengths, and tower heights increase, construction continues to be a
challenge with limited high-capacity crane availability and transportation logistics
Looking forward, supply and demand will remain a key driver in the rising costs of turbines and
balance-of-plant construction.15
Figure 3-31 presents a breakdown of project costs by category, based on documentation of costs
from actual projects constructed in 2006 and 2007. The turbine, tower, and shipping costs are by
far the highest-cost components, contributing over 70% of the total project costs.

15

Andrew Fowler, June 2, 2008. AWEA Windpower 2008 Presentation, Trends in Wind Power Prices B.O.P. and
Turbine Costs, Renewable Energy Systems Americas (RES).

3-54

Wind Power

Figure 3-31
Breakdown of estimated capital costs for land-based wind plants
Source: Renewable and Sustainable Energy Reviews, pp. 13721382

The technology advances of turbine technology discussed in the previous section are expected to
make an impact on the cost of wind turbines as the new technologies are implemented. Table 3-7
summarizes the expected technological advances in turbine technology and the range of
associated increases in capital cost and energy production. This information is based on an
analysis completed by the DOE in order to determine the feasibility of achieving 20% of the U.S.
electric supply mix from wind power by the year 2030 and the technological advances that
would be required. The analysis includes certain assumptions about future market conditions,
such as the long-term extension of the federal production tax credit, which would create a stable
market environment in which companies would be willing to invest in the required research and
development activities. It is expected that most of the listed improvements would not be attained
in the short term. Enlarged rotors are likely to provide the most improvements between now and
2020, whereas the advanced tower concepts may be developed later than the 2030 time frame.
O&M costs are highly variable and depend on a number of factors such as the O&M strategy
employed, the reliability of the equipment, the operating environment, and the roles and
responsibilities of the equipment manufacturer in providing service and warranty repairs. In the
traditional O&M model, the turbine supplier provides warranted service and repairs for a fixed
fee during the first few years of the project. Five to eight years ago, five-year warranties were
available in the market; however, the trend in recent years among turbine suppliers is to reduce
the term to two years and to eliminate operations tasks from the contract. Options are still
available to extend the warranty period beyond the initial two-year term, but owners of large
fleets tend to take over all turbine service, repairs and operations at the close of the warranty
period. Owners that do not have large fleets have also been turning to third-party operators to
service and repair their projects.

3-55

Wind Power
Table 3-7
Areas of potential technology improvement
Cost Increments
(Best/Expected/Least, Percent)
Technical Area

Potential Advances
Annual Energy
Production

Turbine
Capital Cost

Advanced tower
concepts

* Taller towers in difficult locations


* New materials and/or processes
* Advanced structures/foundations
* Self-erecting, initial, or for service

+11/+11/+11

+8/+12/+20

Advanced (enlarged)
rotors

* Advanced materials
* Improved structural-aero design
* Active controls
* Passive controls
* Higher tip speed/lower acoustics

+35/+25/+10

6/-3/+3

Reduced energy losses


and improved availability

* Reduced blade soiling losses


* Damage-tolerant sensors
* Robust control systems
* Prognostic maintenance

+7/+5/0

0/0/0

Drive train
(gearboxes, generators,
and power electronics)

* Fewer gear stages or direct drive


* Medium/low speed generators
* Distributed gearbox topologies
* Permanent-magnet generators
* Medium voltage equipment
* Advanced gear tooth profiles
* New circuit topologies
* New semiconductor devices
* New materials

+8/+4/0

11/6/+1

Manufacturing and
learning curve

* Sustained, incremental design


and process improvements
* Large-scale manufacturing
* Reduced design loads

0/0/0

27/13/3

+61/+45/+21

36/10/+21

Totals

Source: 20% Wind Energy by 2030: Increasing Wind Energys Contribution to U.S. Electricity Supply, DOE,
May 2008. DOE/GO-102008-2578

3-56

Wind Power

3.8 Developments in Offshore Wind Technology


As of the end of 2009, the worldwide installed offshore wind power capacity was 2,112 MW,
located primarily in waters off the coast of Denmark, the United Kingdom, Ireland, Sweden, and
the Netherlands.16 An additional 1,500 MW is currently under construction, primarily in
Germany and the UK, and BTM Consult ApS estimates there will be just under 13,500 MW of
offshore wind capacity installed between 20102014.17
There are currently no offshore wind farms installed in the United States. The vast onshore wind
capacity that has yet to be developed and the high cost of offshore development have prevented
significant interest and investment in offshore wind. Other challenges to offshore wind
development in the United States include the lack of specialized equipment for transporting and
erecting offshore turbines and a traditional lack of a defined permitting process.
Some of the major challenges to offshore wind development are beginning to be addressed,
however, and commercial interest appears to be on the rise. For example, in April 2009, the
Department of the Interior finalized a long-awaited framework for renewable energy production
on the U.S. Outer Continental Shelf (OCS). The framework establishes a program to grant leases,
easements, and rights-of-way for orderly, safe, and environmentally responsible renewable
energy development activities, such as the siting and construction of offshore wind farms, on the
OCS. In addition to establishing a process for granting leases, easements, and rights-of-way for
offshore renewable energy development, the new program also establishes methods for sharing
revenues generated from OCS renewable energy projects with adjacent coastal states.
The Interior Departments Minerals Management Service (MMS) and the Federal Energy
Regulatory Commission (FERC) cleared the way for the publication of these final rules by
signing an agreement in April 2009 that clarifies their agencies jurisdictional responsibilities
for leasing and licensing renewable energy projects on the OCS. The MMS has exclusive
jurisdiction with regard to the production, transportation, or transmission of energy from nonhydrokinetic renewable energy projects, including wind and solar. FERC will have exclusive
jurisdiction to issue licenses for the construction and operation of hydrokinetic projects,
including wave and current, but companies will be required to first obtain a lease through MMS.
Regulatory activity and other developments are helping to fuel project work. There were more
than 2,000 MW of wind plants under consideration in the United States as of September 2009,
and DOE estimates 54,000 MW of electricity could come from offshore wind farms in the
United Sates by 2030.18 Proposed offshore projects in the United States are primarily
concentrated in the northeast and mid-Atlantic regions, where land-based wind development is
limited due to the high population density and limited wind resource. Fewer constraints on noise
and visual impact, in addition to the higher quality wind resource and relatively close proximity
to significant electric demand along the coast, contribute to the appeal of offshore wind in the
northeast.
16

Offshore Wind Technology presented by Walt Musial, National Renewable Energy Laboratory, at AWEA
Offshore Wind Workshop, September 9, 2008.

17

BTM Consult ApS. International Wind Energy Development World Market Update 2009, March 2010, p. 64.

18

Offshore Wind Technology presented by Walt Musial, National Renewable Energy Laboratory, at AWEA
Offshore Wind Workshop, September 9, 2008.

3-57

Wind Power

3.8.1 Trends in the Offshore Turbine Supply Market


The primary differences between offshore and onshore wind turbines are size and foundation
requirements. Because of the high cost of offshore wind turbine foundations and undersea
electric cables, offshore wind turbines are typically larger than their onshore counterparts to take
advantage of economies of scale. In addition to the difference in size, offshore wind turbines
have been modified in a number of ways to withstand the corrosive marine environment, such as
implementation of a fully sealed or positive-pressure nacelle to prevent corrosive saline air from
coming in contact with critical electrical components, structural upgrades to the tower to
withstand wave loading, and enhanced condition monitoring and controls to minimize service
trips.
Vestas and Siemens have historically supplied the bulk of the offshore turbine market, with
Vestas holding a cumulative operational offshore market share of about 60%. Vestas introduced
the V90 3.0-MW to the commercial market in 2005, and well over 500 units are now deployed
both onshore and offshore. Siemens installed the first 25 of its 3.6-MW flagship machines at the
Burbo Bank wind farm in the UK in 2007. These turbines, for both onshore and offshore
application, were also installed at two offshore projects in the UK: one at Inner Dowsing and one
at Lynn. General Electric also entered the commercial offshore market with the installation of
several 3.6-MW turbines off the coast of Ireland in 2004, but has since discontinued production
of offshore turbines.19
There are several new entrants into the commercial offshore market. German-based
manufacturers REpower, Bard Engineering, and Multibrid have installed pilot projects using 5MW offshore turbines, and each company expects to manufacture up to 100 turbines per year by
the end of the decade.20 In 2008, Enercon reported installation of two prototype 6-MW turbines
with 126-m rotor diameters in Emden, Germany.21 WinWinD of Finland, a near-shore wind
entrant, installed five of its 3-MW WWD-3 Multibrid-type wind turbines at the Kemi Ajos
Harbour wind farm in Finland.22
Other companies report research and development activities of multimegawatt-class offshore
turbines. In 2007, Clipper Windpower announced plans to develop the worlds largest wind
turbine at 7.5 MW to be tested in Blyth, UK. Gamesa has announced a 4.5-MW turbine that will
feature the worlds largest rotor diameter at 128 m. Sinovel Wind Co. of China is undertaking
research and development on 2-, 3-, and 5-MW wind turbines, for both onshore and offshore
application. Germany-based Nordex has commenced a concept study for a new
3- to 5-MW offshore wind turbine, with a prototype envisaged for 2010 and commercial
production planned to begin by 2011 or 2012. However, Nordex abandoned efforts to develop
offshore turbines after negotiations with a partner to form a joint venture failed. The offshore
business unit was dissolved and further attention given to the companys onshore division.

19

40,000 MW by 2020: Building offshore wind in Europe. Renewable Energy World, January 3, 2008.

20

Renewable Energy World, July 2008, p. 104.

21

Enercon website: http://www.enercon.de/www/en/nachrichten.nsf.

22

http://www.abb.com/cawp/seitp202/261428146168c203c12573bf002f3466.aspx.

3-58

Wind Power

3.8.2 Offshore Transmission Technology Status


Currently, offshore wind farms are installed at distances from shore ranging from 0.8 km to
20 km. Undersea cables connect the wind turbines within a project to an offshore substation
and from the substation to the mainland. Most offshore wind farms utilize high-voltage ac
transmission lines to transmit power from the offshore substation to the mainland. High-voltage
dc (HVDC) transmission is a new technology that experiences lower electrical line losses than
high-voltage ac; however, rectifier and inverter losses are introduced when converting from ac to
dc at the offshore substation and from dc back to ac at the onshore grid-connection point. The
lower line losses are expected to outweigh the additional electrical conversion losses and cost
differential only for projects located a significant distance from shore. In Germany, one HVDC
cable connection to an offshore wind project is near completion and contracts have been awarded
for three more HVDC-wind projects.
3.8.3 Offshore Capital and O&M Cost Trends
This section contains a general discussion of key offshore wind project component costs,
focusing on two example U.S. projects: one in the Northeast, and the other in the Great Lakes
region. The basic facility design assumptions for each facility are given, along with a discussion
of the principal considerations underpinning the selection of the major components and the cost
implications of those choices. Following the discussion of these two projects, a similar
discussion is undertaken for an installation in the United Kingdom.
Wind farm components examined in this section include:

Wind turbine foundations

Wind turbine generators (WTGs)

Offshore substation (OSS)

Inner array grid (IAG)

Export cabling including onshore interconnection

Installation costs

3.8.3.1 Capital Costs


The largest factor in the levelized cost of electricity (LCOE) for an offshore wind project is the
project capital costs. This section discusses the major factors and considerations embedded in the
analysis of the capital costs (CAPEX) for the facilities considered in this economic analysis. In
this analysis, capital costs were developed in the followingtwo ways:

Average actual costsavailable reports and specific project data were examined to establish
actual component costs. This yielded a range of costs within which a developer should expect
each category of costs to fall.

Total capital cost percentagespecific component categories are often estimated as a


fraction of that total capital cost.

Both these methods were used in developing the capital costs for the projects discussed in this
report.
3-59

Wind Power

3.8.3.2 Wind Turbine Foundations


Section 3.4.10 provides an extensive discussion and analysis of the different wind turbine
foundation options. As discussed earlier, water depth is the single critical factor in the selection of
offshore wind project foundations. Building on that analysis, monopile foundations were chosen
for the Northeast (NE) project and tripod foundations for the Great Lakes (GL) project. The UK
offshore wind project discussed later in this report selected monopoles for the foundation type.
Once the type of foundation is chosen, foundation capital costs are dependent on two factors;
WTG type and hub height. That choice then determines the weight-bearing requirements of the
foundation, for the tower head mass, rotor loads, and other inherent weights. The projects
discussed in this report are assumed to use the Siemens 3.6 MW offshore wind turbine with an
80-meter hub height. Other factors influencing the choice of foundation type include the water
depth, the soil conditions, ice conditions, wave heights, and currents.
Foundation costs for offshore wind projects can range from $840/ kW to $1,700/ kW. Table 3-8
shows typical costs for different foundation types for WTG in the size range of three MW and
above. Although a fourth category of foundations, gravity-based, was described in Section
3.4.10, they have not been included in this cost comparison due to insufficient data on gravitybased foundations.23
Table 3-8
Typical average foundation costs
Monopile

Jacket

Tripod

Typical cost range, /kW


(average)

600720
(650)

9401,215
(1,100)

725980
(850)

Typical cost range, $/kW


(average)

8401,008
(910)

1,3161,700
(1,540)

1,0151,372
(1,190)

3.8.3.3 Wind Turbine Generators


The WTG consists of several components, including the tower, the rotor and the nacelle
(together, the rotor-nacelle assembly [RNA]), where the tower is installed on the foundation by
means of a flange. The RNA price is usually in the same order of magnitude for all wind farms,
but the tower price is site specific. This is because the length, the wall thickness, and the
diameter of the tower will vary, depending on the height of the foundation. The level of the
foundation/tower interface depends on the type of foundation used and, more importantly, on the
wave conditions specific to that site (usually the design driver is the 50-year wave). The Siemens
3.6 MW offshore wind turbine was chosen for this study because of its frequent use in existing
offshore wind projects and the availability of cost and operations data for the unit. Experience
with different projects and different offshore WTGs sized between 3.6 MW and 6 MW shows
that costs for the WTG assembly are in the range24 of 1,275 to 1,687 /kW (1,7852,362 $/kW),
with an average value of 1,460 /kW (2,044 $/kW).
23

To date, only two examples of gravity-based foundations have been identified in place, for which data were
available. This includes one small WTG (maximum 2 MW in size) in shallow water, and a second project of 5 MW
in deep water.
24
These estimates are for the WTG purchase only and do not include delivery, installation and/or commissioning.

3-60

Wind Power

3.8.3.4 Inner Array Grid


The inner array grid (IAG) is composed of the cables placed between the separate WTG (or the
WTG strings) and the offshore substation. The IAG typically uses 33-kV AC cables with built in
communication fibers (glass fiber cables). The voltage level of 33 kV is currently the industry
standard for European offshore wind farms. A lower voltage (20 kV) would produce higher cable
losses; higher voltages are currently not possible, owing to the following technical design
criteria:

Most offshore WTGs make use of medium voltage switchgears with a rated voltage of 36
kV, running at an operating voltage of 33 kV.

The IAG cannot run at an operating voltage of 36 kV, as the grid operators in Europe usually
request a maximum overvoltage of 20% for a short period of time. This level would create
too high a voltage for 36 kV rated switchgears to survive.

IAG cabling of higher voltages would require switchgears also rated to those higher voltages.
Switchgears with this capability are quite expensive and large, too large to fit in the base or
transition pieces of current WTGs, where the typical tower/transition piece measures 4.5 m to
6 m in diameter. In addition, use of higher voltage IAG would require higher safety/
shielding distances.

Costs for the IAG are typically estimated including installation of the cabling. Typical IAG costs
are in the range of 100 to 173 /kW (140 to 242 $/kW) with an average value of 130 /kW
(182 $/kW).
3.8.3.5 Offshore Substation (OSS)
The OSS is the offshore platform that is connected to the lower voltage strings of the IAG and to
the higher voltage export cable(s) to the coast (or to another platform). The OSS is mainly
equipped with busbars, switchgears, and step-up transformers to increase the voltage from the
usual 33 kV of the IAG to the higher voltage of the export cable(s). Typical costs for offshore
substations are in the range of 155 to 215 /kW (217 to 301 $/kW) with an average value of 181
/kW (253 $/kW).
3.8.3.6 Export Cabling and Onshore Interconnection
The export cable (usually one or two, depending on installed power and redundancy philosophy)
connects the OSS to the onshore grid or substation. Export cables are usually 120 kV to 160 kV
and operate at ac for wind farms near the coast, where the cable length is typically less than 100
km. With larger distances to the coast (100 km), a dc cable connection may be required as ac
cables are subject to high losses over such distances. Because both of the hypothetical wind
farms under discussion in this report are well within 80 km from shore (20 and 8 km,
respectively, for the NE and GL locations), we have assumed an AC export cable for both
facilities. In Europe, the export cable and the onshore interconnection are usually not considered
as part of the CAPEX for offshore wind farms, as it is often built by the transmission system
operator (TSO) or special purpose companies. In eight EU countries active in offshore wind

3-61

Wind Power

development, five25 require the project developer to build and pay for the offshore export
cabling. In the United States, where no general policy has yet been enacted, we assume that
responsibility for export cabling lays 100% with the wind farm developer. Typical costs for
the export cabling and onshore interconnection are in the range of 250 to 1,300 /kW (350 to
1,820 $/kW) with an average value of 820 /kW (1,148 $/kW).
3.8.3.7 Installation Costs
The installation costs for offshore wind farms are estimated to range between 7% and 17% of the
total CAPEX. These costs depend heavily on project specific parameters, such as water depth,
coastal distance (or distance to base port), the number and size of both the foundations and the
WTG and the distances across the entire wind farm. Review of project data and available studies
has shown that installation costs for the IAG and the OSS are covered within the CAPEX for
those two components. As a result, in this section we report installation costs for two categories
only, the foundations and the WTG for the 3-MW+ sized projects. Typical costs for the
installation of foundation and WTG range between 290 and 500 /kW (406 to 700 $/kW) with an
average value of 385 /kW (539 $/kW). Table 3-9 summarizes the ranges and average costs by
major wind farm component, as discussed in this section.
Table 3-9
Offshore wind: CAPEX costs from project data
Cost Ranges
($/KW)

Cost Average
($/KW)

Foundation costs for WTG 3 MW+

8401,680

1,260

Wind turbine generators3 MW+

1,7852,362

2,044

Inner array grid

140242

182

Offshore substation

217301

253

Export cabling and onshore interconnection

3501820

1148

Installation costs (WTG and foundations only)

406700

539

3,7387,105

5,426

CAPEX Component

Total

25

The five that require the developer to build the offshore transmission line(s) are the UK, Sweden, the Netherlands,
Ireland, and France. Germany and Denmark consider this cost to be the responsibility of the TSO, whereas Belgium
has a cost-sharing formula: 1/3 TSO; 2/3 developer.

3-62

Wind Power

3.8.3.8 Distribution of CAPEX by Category


A number of studies from European and U.S. sources give an indication of typical CAPEX for
offshore wind farms. Data from these studies, as well as experience from recent projects (such as
due diligence) will be used as the basis for the cost estimates for both the NE project and the GL
project. Similar base data for both projects will be used, as there is only one offshore wind farm
in the world built in fresh water (or brackwater) and there is no study available dealing with the
costs for such fresh water wind farms. The cost estimate data has been adapted to the differing
site specific data, such as wind speed, water depth, and the like.
We recognize, however, that the analysis across the sites summarized in these studies must still
be informed by knowledge of the local conditions in which our three hypothetical projects are
located. Specifically, two important design drivers for the Great Lakes wind farm will influence
its costs: (a) the freezing of Lake Michigan during winter periods (which affects foundations and
O&M) and (b) lower average ambient temperatures. The latter have an impact on the whole
generation system, especially the longevity of steel and adapted cooling/heating systems.
Table 3-10 summarizes numerous studies of the actual costs seen in both U.S. and European
settings, for both total CAPEX and the share of that expenditure represented by different cost
components. This analysis leads to an average for the total estimated CAPEX of $4,840/KW. This
result is reasonably consistent with the earlier analysis of CAPEX by component category (Table
3-9). This CAPEX will be assumed in each of the U.S. cases presented in this report, with
adjustments for project specific requirements such as wind speed, water depth, and so on. This
table is followed by Figure 3-32, which shows the distribution of CAPEX by equipment category.
Table 3-10
Offshore wind: total CAPEX and CAPEX share by component category
Total CAPEX
(Estimated)

CAPEX by Category (% of total)

Source
WTG

Foundation

Grid

Installation

Other

/kW

$/kW26

The Offshore Report,


MAKE Consulting,
2009

44

22

14

15

3,300

4,620

Windenergie Report
Deutschland
Offshore, Fraunhofer
IWES, 2009

45

25

21

3,600

5,040

not

not

not

not

estimated

estimated

estimated

estimated

3,700

5,180

not

not

not

not

not

estimated

estimated

estimated

estimated

estimated

3,000

4,200

Offshore Wind in
Europe, Market Study,
KPMG 2010
Offshore wind power:
big challenge, big
opportunity, Carbon
Trust, 2008

26

15

An exchange rate of 1.4 $ / has been used in the conversion of currencies.

3-63

Wind Power
Table 3-10 (continued)
Offshore wind: total CAPEX and CAPEX share by component category
Total CAPEX
(Estimated)

CAPEX by Category (% of total)


Source
WTG

Foundation

Grid

not

not

not

estimated

estimated

estimated

KEMA project
data, 2011

36

22

28

Average values

41

23

19

13

Offshore Wind
Energy
Installation and
Decommissioning
Cost Estimation
in the US Outer
Continental Shelf,
Energy Research
Group, 2010

Installation

20

Figure 3-32
Offshore wind: CAPEX distribution by category
27

An exchange rate of 1.4 $ / has been used in the conversion of currencies.

3-64

Other

not

/kW

$/kW
27

3,070

4,300

4,070

5,698

3,457

4,840

estimated

Wind Power

It is important to note that this analysis is based on historical data, largely European in origin.
Although the data themselves cluster sufficiently well to merit confidence in the result, it should
be noted that these figures might change significantly as U.S. offshore wind infrastructure grows
in depth and capability in all five categories of CAPEX, but particularly U.S.-based foundation
manufacturing and installation.
3.8.3.9 O&M Costs
This section explores the cost of operation and maintenance (O&M) for offshore wind farms.
Estimating O&M costs depends on many variables including turbine type, distance to shore,
location of O&M base, vessel types, and many other related factors. This section begins with an
overview of each O&M component, including a rough cost estimate. These generic factors have
been used in the project-specific estimates included in all three projects discussed in this report.
O&M costs for an offshore wind farm are significantly higher than for onshore wind farms and
therefore contribute substantially to the energy generation costs. These higher costs are the direct
result of the projects offshore location, which inherently includes difficult and costly access and
a generally higher cost to affect needed maintenance and repairs.
Maintenance is essential to guarantee the availability of a wind farm. Wind farm availability is
defined as the percentage of time the facility is able to produce electricity in proportion to 100%
of the time available in a certain period (usually one year). Availability is a function of the
reliability, maintainability, serviceability, and operation of the wind farm and each component of
that facility.
For offshore wind farms, site accessibility and the maintenance strategy adopted by facility
operators are of equal importance in maximizing the total availability of the offshore generation.
Figure 3-33 provides a diagram of the factors that influence the availability of an offshore wind
farm. Each of these is defined as follows:

Reliability of a system is the probability that the system will perform its tasks. This
probability is usually determined as a percentage of time or the mean time between failures
(MTBF). For a wind turbine, the average MTBF is between 1 and 4 months, depending on
the turbine type or the development stage of a wind turbine.

Availability is the probability that the system is operating satisfactorily. The major difference
between reliability and availability is the O&M strategy of the system. A system can be very
reliablethat is, its failure frequency is extremely low, but when no maintenance or repair
action is taken after a failure, its availability becomes very poor.

Maintainability is a more qualitative issue that addresses the ease of repair issue. It can be
expressed in terms of hours needed to complete a repair action.

Serviceability regards, in a similar way, the ease of regular (scheduled) maintenance.

Accessibility is associated with the ability to get to the site to perform necessary maintenance
and repairs. This can be affected by weather conditions, vessel availability, distance from
port, and project design.

3-65

Wind Power

Figure 3-33
Factors that influence wind farm availability

Turbine reliability, maintainability, and serviceability are influenced by the turbine chosen. Each
type of turbine has specific failure rates for the components used, which influence its reliability.
Maintainability and serviceability are influenced in part by the type of components that are used
in the wind turbine and partly by the design of that turbine. During turbine design, designers can
take into account the nature and frequency of maintenancefor example, placing a filter that
needs frequent replacement in an easily accessible location.
The reliability, maintainability, and serviceability of a wind turbine and the influence of each
parameter on the units overall availability generally apply to both onshore and offshore turbines.
For onshore turbines, these three parameters are good indicators to predict overall availability.
Additional factors that influence offshore availability include both the accessibility of the wind
farm and the maintenance strategy adopted by operators. The reduced accessibility of offshore
facilities, and the logistical arrangements required (ships, jack-up barges, or helicopters), are the
main factors that differentiate offshore O&M from onshore O&M.
Generally, the costs for maintaining an offshore wind farm will be determined by both corrective
and preventive maintenance. The maintenance costs over the lifetime of the WTG can be split up
into three major phases. Each of these phases is characterized by different WTG operating
patterns and therefore by different O&M strategies and activities.
Phase 1: Startupduring the commissioning period, the burn-in problems usually require
additional maintenance effort (and thus costs). Time should be spent on finding the right settings
of software, changing minor production errors, etc. Turbine manufacturers usually provide a
fixed price, five-year contract that covers commissioning, preventive and corrective
maintenance, warranties, and machine damage.

3-66

Wind Power

Phase 2: Normal operating lifeduring this period, preventive maintenance will be completed
and random failures might still occur. After about 10 years of operation, it is very likely that
some of the main systems of the turbines will need attention, including pitch motors, hydraulic
pumps, lubrication systems, rotor blades, gearbox, generator etc. Experience is so far quite
limited to indicate how often a major overhaul should be carried out. The exact point in time at
which the overhaul(s) should take place is presently not known, perhaps after seven years, 15
years, or not at all. The major overhaul in fact is to be considered as condition-based
maintenance.
Phase 3: Wear-out phaseduring this period, the failure rate can be expected to increase due to
aging of the components. As the end of the WTGs lifetime approaches, the rate of corrective
maintenance is likely to increase. At present, it is not possible to predict how much additional
corrective maintenance will be required, as most available failure statistics rely on data from
WTG that have less than 15 years of operation.
The chosen maintenance strategy describes the equipment needed to perform maintenance. It
also lays out the framework and plans to accomplish the two types of maintenance required by an
offshore wind farm: preventive maintenance and corrective maintenance. These two types of
maintenance are discussed in more detail in the next sections.
3.8.3.10 Preventive Maintenance
Preventive maintenance is performed to avoid component failures and is an essential part of
ensuring the maximum availability of the wind turbines. Most wind turbine types require
between one (state of the art) and two preventive maintenance visits per year. Work is performed
by a team of at least three service engineers;28 for large units, two to three teams (six to nine
technicians) are generally sent. Depending on the type of service, the number of service
technicians and the size of the wind turbine, a preventive overhaul can take between two and
nine visits (48216 hours of standstill). Currently, best practice for offshore preventive
overhauls requires between two and four days (4896 hours of standstill).
Below we describe three different categories of preventive maintenance: planned, Condition
based, and reliability based:

Planned preventive maintenancethe objective of planned (preventive) maintenance is to


minimize unscheduled downtime. By performing maintenance in fixed intervals,
maintenance can be planned in periods of good access and relatively low energy yield,
thereby minimizing energy losses from the standstill period.

Condition-based preventive maintenancecondition-based preventive maintenance aims to


optimize maintenance costs and availability. Condition-based maintenance defines thresholds
that trigger maintenance action. While the condition of components can be determined by
sensors, taking samples or visual inspection, most turbines can be equipped with a conditionmonitoring system (CMS). The CMS will continually monitor one or more critical
components. CMS will detect early wear of components, thereby triggering replacement
before any unplanned breakdown can occur.

28

In Europe it has become an industry standard to always send at least three technicians to an offshore WTG for
safety reasons (it is easier to rescue an injured technician with two people and offshore wind farms are built in
remote areas).

3-67

Wind Power

Reliability-based preventive maintenancereliability-based preventive maintenance aims to


optimize the time plan of planned maintenance. By analyzing SCADA data, corrective
maintenance logs and the results of condition-based maintenance, operators can evaluate the
reoccurrence of planned maintenance. The analysis may indicate that failures occur more often
than anticipated, resulting in additional corrective maintenance. By adjusting the activities
included in planned maintenance of a component, for example, failures can be prevented.

To the extent possible, timing of preventive maintenance can also improve project availability by
reducing the amount of power production loss due to O&M activities. Wind conditions at an
offshore wind facility are not uniform throughout the year, and there are typically periods of low
wind conditions. Timing preventive maintenance activities that require WTG shutdown during
low wind periods can improve project overall availability.
3.8.3.11 Corrective Maintenance
Corrective maintenance is defined as maintenance required to keep WTGs in operation that are
not part of the normal plant maintenance cycles. Often corrective maintenance is associated with
component failure, but can include corrective actions conducted on a non-routine basis to
mitigate a potential failure or investigate a site condition. The wind farm will also be equipped
with a SCADA monitoring system. The SCADA system enables monitoring and control of the
wind from a companys control center (onshore or offshore). Although safety reasons limit the
control capabilities of the SCADA system, most manufacturers allow remote starts, stops,
adjusting of maximum power, setting and changing parameters and yawing of the turbines.
Besides providing this level of remote control, the SCADA system is used for data acquisition.
Data logged by the SCADA system includes the following:

Historical and real-time data from a variety of control and safety sensors

Meteorological data

A log with warnings, events and alarms

Production data

Availability of the wind turbines

The SCADA system can indicate when a turbine is not functioning properly, thereby triggering
corrective maintenance. The SCADA system is not the only indicator for corrective maintenance.
During preventive maintenance or any extra service visit, inspections by service technicians may
reveal equipment conditions that cannot be corrected during routine maintenance visits and may
require further corrective maintenance.
Based on generalized data, on average, two to six failures requiring corrective maintenance occur
per year in state-of-the-art onshore WTG.29 Several types of failures can be solved by remote
means; others require on-site action by a maintenance team. Although these figures are valid for
onshore WTG, it is expected that state-of-the-art offshore WTG will achieve at least these

29

Durstewitz, Ensslin. +15 years of practical experiences with wind power in Germany.

3-68

Wind Power

frequencies, if not a lower frequency of events. The current experience of offshore wind farms
shows lower overall project availability than that of onshore wind projects. It is also expected
that the new larger WTG models will have higher numbers of failures in their initial years than
current machines because of a lack of historical operating data.30
Turbine suppliers are understandably reluctant to provide data on the failure rates and frequency
of corrective maintenance of their turbines. Data gathered on different turbines and turbine
components shows that the average frequency of corrective maintenance for turbines is between
two to five occurrences per year, but with a large scatter (between 1 and 10) depending on the
units size, type, and age. Corrective maintenance rates are higher during the first three to five
years of operation when compared to the following five years.
Turbine corrective maintenance can be caused by many different components. Figure 3-34
depicts the distribution of failures by different categories of wind turbine components. The
majority of failures occur in electrical and control systems, including power electronics and
sensors.

Figure 3-34
Distribution of turbine failures by components

30

IWES/WMEP. Windenergie Report Deutschland 2009 Offshore.

3-69

Wind Power

When normal turbine operation is impeded or influenced by equipment problems, the severity of
the problem and the influence on operation determines the corrective maintenance action
required.

(Remote) resetTurbines are controlled and monitored by a variety of software. What could
seem to be a hardware problem could also be caused by software. A reset of the turbine could
solve the situation when this is caused by software. Resetting the turbine in many cases can
be done remotely. For safety reasons, however, it is not possible to perform a remote reset for
a large variety of equipment problems, especially if the root cause cannot be identified.

Small failureFailures can occur that have little or no impact on the operation of the wind
farm or wind turbine. These failures require corrective maintenance. When the failure doesn't
influence normal operation, the decision may be made to postpone corrective action until
preventive maintenance is planned.

Large failureFailures that have impact on the normal operation of the wind farm or wind
turbine require immediate corrective maintenance. Depending on the component and nature
of the failure, corrective maintenance can either be performed immediately or scheduled as
needed to accommodate mobilization of required vessels and/or the lead time needed to
secure spare parts.

3.8.3.12 O&M Expenditures


The limited data available for offshore wind projects can be expected to have the following
characteristics:31

The O&M costs are approximately 40 to 70 $/kW installed capacity per year.

Approximately two-thirds of O&M costs are caused by corrective maintenance.

Experience shows that O&M expenditures for offshore facilities can contribute to approximately
25 to 30% of the LCOE. By comparison, the equivalent O&M contribution for an onshore
facility is between 10% and 15%. The higher O&M burden for offshore facilities is attributable
to several factors, including (a) equipment costs for offshore O&M activities, such as crane ships
and other vessels; (b) higher wages for offshore technicians;and (c) accessibility (as offshore
locations may be impeded by weather conditions, high waves, and strong winds). Furthermore,
the capacity factors are generally higher for offshore wind farms. Other factors that come into
play include the following:

Turbine maintenanceturbine failure rate, indicating the demand for corrective maintenance

Transportation and vessel availabilitythe vessels and equipment required to access the
units, for lifting and repair, and the fluctuating prices for these elements

Weather conditionsthe influence of the wave height and wind speed on the operational
windows for accessing and repairing the turbine

Distance to the shore

Logistical considerationsfor example, personnel availability, the crew size, stock control,
contracts with equipment suppliers and offshore companies

31

Engels, Obdam. Current developments in wind2009.

3-70

Wind Power

Turbine maintenance. If offshore turbines fail, maintenance technicians need to access the
turbines and carry out maintenance. Especially in case of failures of large components, offshore
turbines are being modified to make replacements of large components easy, such as by making
modular designs, or by building in an internal crane to hoist larger components. The ease by
which repairs can be made influences the O&M expenditures.
Transportation and vessel availability. For present-day offshore wind farms, small boats like
catamarans (e.g., Windcat), mono hulls (e.g., Fob Lady), or SWATH ( Small Waterplane-Area
Twin Hull) boats are being used to transfer personnel from the harbor to the turbines. In case of
bad weather, helicopters also are being used. Rigid inflatable boats (RIBs) are used only for short
distances and during very good weather. For intermediate sized components like a yaw drive,
main bearing, or pitch motor, it is often necessary to use a larger vessel for transportation of
equipment, such as a supply vessel.
For both construction and maintenance of an offshore wind farm, several purpose-built vessels
are used, such as cable laying vessels and heavy lifting vessels. These vessels can be deployed all
over the world. The Merchant Marine Act of 1920, commonly known as the Jones Act, requires
vessels engaged in the transport of passengers or cargo between U.S. destinations to be built and
flagged in the United States, and owned and crewed by U.S. citizens. The Merchant Marine Act
could influence the number, type, and cost of vessels available for offshore wind work on U.S.
projects, thereby creating possible capacity constraints.
Each of the vessel types listed above can be specified by several characteristics. For offshore
operation and maintenance, the following characteristics are of importance:

Weather windowthe nature and extremes of weather that the vessel can reliably and
consistently handle, especially wave height and wind speed

Logistics timelead time to arrange for and prepare the vessel for its intended use

Traveling speedwhich determines traveling time between turbines and locations

Functional capabilitiesthese may take many forms, e.g., the maximum number of
maintenance technicians or personnel that can be transported, maximum hoisting capacity
and height, maximum load that can be transferred

Vessel costsof the system per hour or per day and the costs for mobilization and
demobilization

The cost of vessels can vary considerably depending on size and capabilities. Day rates may vary
between 2,500 for a small supply vessel to 100,000 or more for jack-up vessels. The
availability of vessels can also be an issue. Large lifting vessels are scarce and therefore not
always available when needed, which can result in long downtimes.
Weather conditions. Offshore weather conditions, mainly wind speeds and wave heights, do
have a large influence on the O&M procedures of offshore wind farms. The maintenance
activities and replacement of large components can only be carried out if the wind speed and
wave heights are sufficiently low. Preventive maintenance actions are therefore usually planned
in the summer period with lower wind speeds and wave heights and fewer storms. If failures
occur in the winter season, it does happen that technicians cannot access the turbines for repair
actions due to bad weather, and this may result in long downtimes and thus revenue losses.
3-71

Wind Power

Logistical considerations. The location and capabilities of available harbor locations also have
an influence on the O&M strategy. All workers and goods have to be transferred between the
offshore wind farm and a selected harbor. The harbor capabilities determine which type of
vessels can enter the harbor, and under what weather conditions vessels can enter. The following
items related to harbor facilities influence O&M strategy and expenditures:

Proximity to wind farmthe location of the harbor determines the travel time to the wind
farm. Long traveling times result in longer down times, longer shifts for maintenance
personal and increased probability that weather will change before the wind farm is reached.

24/7 quayside access requiredalthough most O&M facilities currently operate within a 12
hour daily working window, it is essential that 24 hour access is available to facilitate 24
hour operation as required. This could, for example, be necessary during a major overhaul.

Speed restrictionsthe location of an O&M facility within a port environment and travel
distance through the port are key considerations. An O&M facility located deep within a
large port may prove less attractive when compared to a smaller port further away and not
subject to the same speed restrictions.

Conflicting traffica busy port can impact dramatically on the ability of an O&M operator
to respond to an emergency, thus a quieter port with less conflicting traffic is preferred.

Tidal constraintssome ports are not accessible during a certain period due to tidal
constrains. This can impact maintenance planning, travel time and quayside access.

Local, skilled workforcealthough a developer and the associated turbine manufacturer will
provide the specific training required to operate and maintain a wind farm, a local, skilled
workforce is essential.

Turbine manufacturer requirementsmanufacturers will often have their own preferred


specification list for an O&M facility, which needs to be considered.

Provision of a helicopter servicefor larger, more remote wind farms, transport to and from the
site may be supported by a dedicated helicopter service. Transportation by sea is also greatly
reliant on stable weather conditions, which can restrict access to the wind farm. A helicopter
service will greatly increase the ability to access the wind farm in poor weather conditions. It is
likely that most developers will require a dedicated helicopter facility on the same site as the
O&M building to minimize transfer times for maintenance crew and equipment.

Large-scale repairs will require an alternative approach to day-to-day O&M. Although it is


possible that each of the wind projects discussed in this report will require a dedicated day-today O&M land based hub, large-scale repairs could be contained within one or two key
locations, with the potential to service a number of wind farms. The lay-down space
requirements are likely to be significant, extending to numerous acres and will need direct access
to deep-water ports with an ability to maneuver massive pieces of equipment onto barges and
jack-up vessels.
To develop an estimate of the O&M expenditures for Great Lakes and Northeastern projects, the
cost are based on models developed by ECN. The ECN cost model allows the modeling of
different O&M strategies (incorporating different choices and uses of vessel types, crew size,
turbine failure rates, etc.) to calculate the corresponding costs. Because the cost model requires
detailed project-specific information for these calculations, and because these two projects are
3-72

Wind Power

hypothetical, assumptions were made regarding the project-specific information. Because many
of the models assumptions reflect predefined data incorporated into the model, the relationships
and results are indicative of an offshore wind farm in Europes North Sea.
3.8.4 Offshore Plant Performance
In general, the wind farms performance or overall efficiency is higher in smaller wind farms and
higher with larger distances between the WTG. This is due to the reduced wake effects, defined
as the reduction of the wind speed behind the rotors in a wind farm. The wake effect is also
responsible for an increase in the intensity of the turbulence inside a wind farm, which in turn
can lead to higher maintenance costs for the WTG (as higher turbulence causes a higher rate of
wear and tear on the WTG components). Optimal wind farm cost efficiency, therefore, represents
a tradeoff between the spacing and distances between the turbines, and the construction costs for
the offshore wind project.
Another variable that affects both cost and wind farm efficiency is the arrangement and
placement of turbines within the farm area. The layout can utilize either symmetrical or
asymmetrical placement of the WTG. Figure 3-35 shows two layouts of wind projects: an
optimized symmetrical layout on the left and an optimized stochastic layout on the right.
Experience and literature32 suggest that a symmetrical wind farm layout typically has lower
overall efficiency than an asymmetrical layout. For this reason, an asymmetrical layout was
chosen for purposes of this analysis. According to Neubert et al.,33 even an optimized
symmetrical layout will have a net energy output 1% below the wind farm efficiency of an
optimized asymmetrical layout.

Figure 3-35
Comparison of symmetrical and asymmetrical wind farm layouts

32

A. Neubert, A. Shah, W. Schlez (2010): Maximum Yield from Symmetrical Wind Farm Layouts. DEWEK
2010.
33
Ibid.

3-73

Wind Power

Depending on the wind turbine type, the minimum distances between turbines as required by the
manufacturers will be five rotor diameters (in mean wind direction) and four rotor diameters
(perpendicular to the mean wind direction).
The shorter distances between turbines lead to higher wear and tear on the units due to higher
turbulences and wake effects inside the wind farm, which can in turn lead to a reduced unit
lifetime. For this reason, greater distances between the WTGs were used in both the Northeastern
and Great Lakes projects.
3.8.5 Offshore Plant Cost Estimates
3.8.5.1 Northeastern United States
For the purposes of this assessment, four northeastern coastal states were evaluated: Maine,
Massachusetts, Rhode Island, and New Jersey. The following factors were key to this analysis:

Wind resourcethe availability of data showing wind speeds in excess of 9 m/s

Ocean depth and floor conditionsthe availability of reliable and detailed information on
which to base design of the foundation structure for the hypothetical facility

Proximity to major populations/electricity markets

Economic supportsthe availability and nature of any incentives for the developers of
offshore wind facilities, as these factors weigh into the economic calculation of later tasks

Proximity to transmission grid

Based on this analysis, sites off the coast of New Jersey were selected for further consideration
(Figure 3-36). Our selection of New Jersey reflects several key factors:

The states proactive work to identify offshore tracks for wind industry leasing interest.34

The innovative program of economic supports created under the Offshore Wind Economic
Development Act. This program created a renewable energy credit (OREC) specifically for
offshore wind, and an aggressive state purchasing mandate is part of that program.

New Jerseys proximity to the major electricity markets of New Jersey and New York City.

Existing transmission in close proximity to the potential site.

The tract is made up of eight 5km-square blocks from the New Jersey Ocean Baseline Study.
These blocks are marked as blocks 7O, P, Q, and R, and 8O, P, Q, and R. Together, these blocks
constitute a site of 10 km wide by 20 km across, for 200 square km. The advantages of this
location include:

Proximity to major northeastern markets

Proximity to major harbor and shipping facilities

Inclusion within the study area of the New Jersey ocean baseline study area; therefore
environmental impacts have already been established/recorded

In April 2011, NJ DEP issued a call for nominations from firms interested in leasing tracts offshore for potential
wind generation. See the press release at http://www.state.nj.us/dep/newsrel/2011/11_0053.htm.
34

3-74

Wind Power

Inclusion within New Jerseys state/federal leasing area

Location significantly offshore yet in shallow depths

Diminished visibility onshore, in heavily traveled tourist/vacation area

Excellent wind resource

Potential disadvantages of this location include the following:

Land access to nearest shore location may be problematic, given narrow roads on barrier
beaches and small vacation towns.

Closest shore area is not industrial, so staging is not as close as ideal. Barging from Port
Authority facilities may be required.

The selected tract begins roughly nine miles off the coast and extends eastward 10 km, making
the eastward edge of the potential tract 22 miles off the New Jersey coast. Depths increase
slightly over the 200 square km, from 22 m at the western edge to 31 m at the eastern boundary.
The tract is characterized by a gently sloping sandy bottom at 17 to 27 m in the area selected for
the wind farm, and has net mean tides of five feet. Average wind speeds at this location are
assumed to be 8.8 m/s, and the probability of winter ice formation is low.
Assessment of site environmental conditions relies on an extensive study completed by the NJ
Department of Environmental Protection in their Ocean/ Wind Power Ecological Baseline
Studies, released in late 2010.35 The Environmental Sensitivity Index (ESI) developed for the
New Jersey coast cataloged in detail the marine resources in this location. Sensitivities found in
this area include:

The identified tract is one of the largest studied that shows generally low sensitivity on most
parameters throughout the majority of the surveyed area.

Two shipwrecks are present within this tractcreating a single point as a non-development
area, because of the species that utilize the wreck for habitat.

One block (R7) contains a shoal, with relatively higher ESI in recognition of the essential
fisheries habitat and commercial and recreational fishing areas in that portion of the block.

Avian density is less than 50 birds per km2, on par with virtually the entire NJ coast.

The identified tractall eight surveyed blocksis considered generally representative of the
vast stretch of NJ coast, with between 13 and 24 (of the 40 species recorded in the area)
dependent on the habitat in these blocks. Only one small section of R7 was deemed to have
greater essential fish habitat (EFH), likely associated with the shoals in that block.

Five of the eight blocks had expanses where sea turtle sightings might occur, although at a
frequency less likely than many other coastal areas.

Marine mammalsboth threatened and endangered (T+E) and all other marine mammals
were noted in this location. The eight-block target location shows three blocks with the
lowest frequency of marine mammal sightings, two blocks with the highest frequency.

35

New Jersey Department of Environmental Protections four-volume series of Ocean/ Wind Power Ecological
Baseline Studies is available at http://www.nj.gov/dep/dsr/ocean-wind/report.htm.

3-75

Wind Power

Figure 3-36
Northeastern project location

Project Description

The turbines selected for this project are the Siemens STW3.6 MW120. To reach the 450 MW
projected output, a total of 125 WTGs will be mounted on monopile foundations and 80 meter
towers. In the absence of detailed information on the turbulence or other parameters of the wind
conditions, a base area for the wind farm of 10 6 km, with the long side oriented toward the
main wind direction will be used. This produces a wind farm layout of 8 rows of 16 WTGs each
with one row with only 13 WTGs.
The wind farm layout is shown in Figure 3-37. The distances between the WTGs are six rotor
diameters or 720 m (6 120 m = 720 m) in mean wind direction and five rotor diameters or
600 m (5 120 m = 600 m) perpendicular to the mean wind direction. We also assumed a 500 m
safety distance around the wind farm. This safety buffer draws on European experience, is
usually considered part of the wind farm and represents a safety distance to shipping lines, cable
routes, pipelines, and the like. As an additional safety measure, offshore wind farms are usually a
restricted area for any public or recreation ship traffic.

3-76

Wind Power

Figure 3-37
Northeastern U.S. project: wind farm layout

Given the assumed base area, orientation and wind farm layout, the efficiency of the
Northeastern U.S. Project is estimated to be 90%.
Cost Analysis
Estimated CAPEX

The Northeastern U.S. Project is defined according to the design criteria and boundary
conditions summarized in Table 3-11. These are the basis for the estimated cost elements in
Table 3-12, which are used in the economic modeling in the following sections.

3-77

Wind Power
Table 3-11
Northeastern U.S. project: design criteria and boundary conditions
Parameter

Description / Comment

Rated capacity

450 MW

No. and type of WTG

125 SWT-3.6-120

Rotor diameter

120 m

Hub height

80 m

Defined as meters above MSL

Location

Atlantic off New Jersey

East of Egg Harbor

Distance from coast

20 km

Measured from center of wind farm

Water depth

921 m
(max. 27 m)

Tidal range

< 1.5 m

Wave heights

13 m

Ambient temperature
range

min. ca. 10C


max. ca. 40C

KEMA estimate for 50year wave


height
Location could require special
adaptation of cooling systems (hot
climate version) as temp. are >30C
from May through October

Soil conditions

Mostly sand with areas


of mud and silt deposits

It is assumed that hammering piles is


possible

Wind conditions

8.8 m/s average annual


mean wind speed

At hub height (80 m)

Ice formation

Very low

Assumed from temperature range and


ocean salt content

Foundations

Monopiles

IAG design

3-78

Value

15 half rings of 8 WTG


each, with
1 half rings of 5 WTG
each

Rated voltage IAG

33 kV

Offshore substation
(OSS)

With two transformers


33/150 kV

Rings are normally open and can be


closed for redundancy / all WTG are
equipped with 3 feeder switch gears
In center of wind farm

Wind Power
Table 3-12
Northeastern U.S. project: estimated CAPEX
Component

Sum
[M]

Sum
[M$]

[/kW]

[$/kW]
2,240

Comments
Higher value due to hot climate
version

Wind turbine generators

720

1,008

48

1,600

Foundations

293

410

21

650

Average value due to


910 comparable conditions and
water depth

Inner array grid

59

83

130

Average value due to


182 comparable conditions and
water depth

Offshore substation

90

126

200

280

Export cabling and


onshore interconnection

112

157

250

Lower value due to small


350 distance to coast and low water
depth

Installation

173

242

12

385

Average value due to


539 comparable conditions and
water depth

70

98

155

217 Average value

1,517

2,124

100

3,371

4,720

152

212

10

337

472

1,669

2,336

110

3,708

5,192

Other (project
development, insurance,
project management, etc.)
Subtotal:
Contingencies
Total

Higher value due to hot climate


adaptation

Conservative assumption
some sources use 5%

On this basis, the estimated total CAPEX for the Northeastern Offshore Wind Project is 1,669
M ($2,336 M) or 3,709/kW ($5,193/kW).
In addition to CAPEX, the TPI was estimated for the Northeastern Project, as shown in Table 313. TPI assumes the TPC with the addition of construction loan and other financing costs. Other
assumptions of the TPI calculation include a nine-month construction cycle and 40% of the cost
paid in month 1 with the balance of costs levelized over remaining 8 months (at 50/50
debt/equity ratio and debt interest rate).

3-79

Wind Power
Table 3-13
Northeastern U.S. project: total plant investment

End of Year

Total Cash
Expended TPC

Before Tax
Construction
Loan Cost at
Debt Financing
Rate

2012 Value of
Construction
Loan Payments

TOTAL PLANT
INVESTMENT

($M)

($M)

($M)

($M)

2012

$2,336

$118.53

$2,454.71

$2,454,71

2013

$0

$0

$0

$0

TOTAL (TPI)

$2,454.71

Estimated O&M Costs

As described above, the Northeastern Project is assumed to have 125 turbines of 3.6 MW,
located off the New Jersey coast.
To facilitate quick response to any failures occurring in the wind farm, qualified harbor facilities
must be located as close to the wind farm as possible. One harbor will be selected when there is a
harbor facility in close proximity that meets all the requirements. This would be the most ideal
situation since all operations can be coordinated from one location. A detailed assessment of
local options could also yield the result that no single nearby harbor meets all requirements. At
that point, operators must determine whether to select (a) a single harbor further away that meets
all requirements but will result in longer travel times, or (b) two harborsone harbor nearby for
day-to-day O&M activities, for which the harbor requirements are more easily met, as only
smaller vessels are used. A second harbor situated further away would be used for coordinating
O&M activities that require larger or special vessels.
After a quick assessment, two harbors were selected for the Northeastern project. The harbor that
will be used for the day-to-day O&M is the harbor of Atlantic City, which lies 30 km from the
wind farm. Because this harbor is not equipped with the appropriate freight handling equipment
for large components, however, large spare parts will have to be transported from Port Newark,
which is located 100 km from the wind farm.
Estimating project-specific O&M expenditures depends on many different factors. No detailed
information is available on wind and wave conditions for this hypothetical site. Also constructing
a specific O&M strategy for this location is out of the scope of this project; therefore, only a
general range of O&M expenditures can be given. O&M costs per year are approximately $50
per installed kW.

3-80

Wind Power

The Northeastern project will have specific properties that influence the overall costs of the
O&M operation. The following section details some of these cost drivers, including:

Distance to harborfor each maintenance action a vessel has to travel to the wind farm. The
distance from the harbor to the wind farm has significant influence on total O&M
expenditures. Longer traveling times result in higher fuel consumption and longer down
times. Maintenance teams have less time to complete repairs, which can result in that some
maintenance actions require two shifts to complete instead of one. As the wind farm is 30 km
from the harbor, it will take a crew vessel traveling at 25 knots 45 minutes on average to
reach the wind farm. This is an average travel time for offshore wind farms. Larger
components are shipped from Newark, which will take a longer time due to the longer
distance and the lower traveling speed of the larger, specialized vessels.

Two harbor facilitiesusing two harbor facilities has the advantage that day-to-day
operations can be done from a harbor nearby the wind farm. The disadvantage is that two
separate locations have to be rented and equipped with appropriate equipment and facilities
to facilitate the O&M requirements.

Accessibilityaccessibility of the wind turbines is influenced by wind speed and wave


conditions. Wave conditions, in particular, can strongly reduce the accessibility of a location.
Personal transfer and hoisting operations are limited by the significant wave height a vessel
can handle. When in an early stage of the wind farm development, vessels are selected for the
O&M activities. An assessment has to be made of the wave conditions at the wind farm
location to determine what kind of vessel will be used for O&M activities. Selecting a vessel
that cannot handle the wave conditions at the location will lead to additional downtime due to
low accessibility. Selecting a vessel or system that can easily deal with the wave conditions
may result in unnecessarily high charter costs for the vessel.

Estimated Energy Yields

To estimate the energy yield for the both the NE and GL wind farms, we refer to the gross
energy yields for the Siemens SWT-3.6-120 wind turbine that is assumed in both cases. For the
NE project, the assumed average wind speed of 8.8 m/s is shown in Figure 3-38.

3-81

Wind Power

Figure 3-38
Power curve of STW-3.6-120

As no detailed wind conditions are known from these hypothetical sites, normal offshore
conditions have been assumed, following those seen in the North Sea (although these may be
very conservative for the less rugged NJ coast). Table 3-14 summarizes the annual energy yield
calculated from the average wind speed and other conditions shown.
Table 3-14
Annual energy yield, Northeastern U.S. project
Number of WTG

125

Installed capacity (MW)

450

Farm efficiency (%)

90

Mean wind speed (m/s)

8.8

k-factor

2.3

Energy yield single WTG (MWh/yr)

16,856

Energy yield wind farm (MWh/yr)

2,107,000

Gross energy yield (MWh/yr) (includes losses)

1,896,300

We have used the values of the gross energy output in Table 3-14 for calculating the net energy
output, along with a typical guaranteed availability and other losses. The results are depicted in
Table 3-15.

3-82

Wind Power

The technical availability estimate for the wind farm, 95%, is realistic and is a quite common
industry value for the guaranteed technical availability. Electrical efficiency is expected to be
98% for the Northeastern project, assuming that the point of common coupling (PCC) and thus
the kWh counter is inside the wind farm. Losses due to neighboring wind farms are not assumed,
as no other developments are known to be planned in the immediate vicinity. Other losses, which
are not covered by the availability guarantee of the WTG supplier, have also been considered.
These additional unavoidable losses may arise from (a) the 95/5% manufacturer operational
guaranteethat is, 5% of the operating hours might be unavailable because of issues with the
WTG; (b) other unavoidable and/or allowable standstills, such as normal maintenance, customer
and authority visits, ice formation on rotor blades, no accessibility due to bad weather, grid faults
and so forth.
In recognition of these factors, the following worst-case calculation is presented, based on
common O&M contracts and common allowable standstills per WTG:

Normal maintenance

47 days/year (100168 hr/year)

Waiting time for heavy lift vessel

35 days/year

Ice formation on rotor blades

13 day/year

Inaccessibility due to bad weather

510 days/year

Automatic cable unwinding

0.5 day/year

Grid faults

0.5 day/year

Total

1426 days/year (~47% further losses)

A realistic estimate for the best case would be some 7 to 10 days of allowable standstills per
year, which equals to losses of some 2% to 3%. Based on these assumptions, estimated annual
energy yield for the NE wind farm to be as shown in Table 3-15.
Table 3-15
Estimate of full load hours, Northeastern U.S. project
Estimated Availability
Gross Energy Yield [MWh/yr]

Base Case

Worst Case

1,896,300

1,896,300

- Availability

96.0 %

95.0 %

- Electrical Efficiency

99.0 %

98.0 %

2.0 %

7.0 %

1,766,200

1,648,900

6.9

13.4

3,924

3,649

- Other losses not covered by availability guarantee


Net Energy Yield [MWh/yr]
Total Losses (%)
Full Load Hours

Estimates for the realistic annual energy yield can be assumed to be roughly 1,766 GWh/yr for
the Northeastern project.

3-83

Wind Power

Economic Analysis

As described earlier, the study methodology included development of the Offshore Wind
Economic Analysis Model. This tool was combined with the costs and output discussed in the
previous sections to calculate the LCOE on the Northeastern project. This calculation uses the
financial assumptions summarized in Table 3-16.
Table 3-16
Financial assumptions, Northeastern U.S. project
Rated plant capacity
Annual energy production at busbar (EPB)
Therefore, capacity factor

450 MW
1,766,200 MWh/yr
44,80%

Year constant dollars

2012

Federal tax rate

35%

State
State tax rate
Composite tax rate (T)
Book life

New Jersey
9.00%
41%
20 years

Construction financing rate

8.0%

Common equity financing share

50%

Preferred equity financing share

0%

Debt financing share

50%

Common equity financing rate

10.5%

Preferred equity financing rate

0%

Debt financing rate

5.6%

Nominal discount rate before tax

8.1%

Nominal discount rate after tax

7.4%

Inflation rate

1.76%

Real discount rate before tax

6.3%

Real discount rate after tax

5.6%

Federal investment tax credit (ITC)

0%

Federal production tax credit (PTC)

$0.022/kWh

Annual state investment tax credit

$10 Million

Total state investment tax credit limit


Renewable ENERGY CERtificate (REC)
Annual REC limit

$ 100 Million
$0/kWh
$0

Based on these assumptions, the LCOE for the Northeastern U.S. Project was calculated. EPRIs
methodology was used to yield the results summarized in Table 3-17.

3-84

Wind Power
Table 3-17
Levelized cost of electricity, Northeastern U.S. project
(TPI * FCR) +
(O&M+LOR)/EPB

Cost of Energy (COE) =


Nominal Rates
Total Plant Investment (TPI)

$2,454,709,160

Fixed Charge Rate (FCR)

10.37%

Annual O&M Cost (O&M)

$57,546,300

Levelized Overhaul & Replacement Cost (LO&R)


Energy Production (EPM)

$0
1,766,200 MWh/yr

Cost of EnergyTPI

14.4100 /kWh

Cost of EnergyO&M

3.2582 /kWh

Cost of EnergyLO&R

0 /kWh

COE Nominal

$0.176682/kWh

COE Nominal

17.6682 /kWh
Real Rates

Total Plant Investment (TPI)


Fixed Charge Rate
Annual O&M Cost (O&M)
Levelized Overhaul & Replacement Cost (LO&R)
Energy Production (EPM)
Cost of EnergyTPI
Cost of EnergyO&M
Cost of EnergyLO&R

$2,454,709,160
10.02%
$57,546,300
$0
1,766,200 MWh/yr
13.9286 /kWh
3.2582 /kWh
0 /kWh

COE Real

$0.171868/kWh

COE Real

17.1868 /kWh

3.8.5.2 Great Lakes


The Great Lakes Wind Council (GLWC) issued a report in October 2010 that identified and mapped
potential leasing areas for offshore wind energy development. This report provides guidance on
model legislation governing development, as well as informing and engaging the public on offshore
wind energy development issues. The GLWC report identified the following five areas as having the
best potential for development of offshore wind projects in the Great Lakes region:

Southern Lake Michigan near Berrien County

Northern Lake Michigan near Delta County

Central Lake Superior near Alger County

Central Lake Huron (out from Saginaw Bay)

Southern Lake Huron near Sanilac County

3-85

Wind Power

For the purposes of this study, the chosen location for the Great Lakes offshore wind project is in
Southern Lake Michigan near Berrien County as a representative site for the development of
Great Lakes offshore wind projects, as shown in Figure 3-39. The conditions of this site are
typical of those to be expected for the development of any most feasible offshore wind projects
in the Great Lakes. The following parameters used in determining the levelized electric cost for a
Great Lakes offshore wind project are based upon the site conditions to be found in this area.
Wind conditions are NREL Class 5, producing average wind speeds in the range of 8 m/s. Based
upon wind maps produced for the state of Michigan, wind speeds of 8.2 m/s will be used as the
average condition for this facility. Weather conditions at this location will impose significant
operational and performance considerations. In recognition of these constraints, the operational
assumptions for this site include:

Temperature range of 50C to 35C and water temperature of 0C

Ice formation on the lake surface and on WTG beginning December 15 and lasting through
March 15; ice thickness assumed to be 8 in. (20 cm)

Blown ice speed of 6 mph; maximum wind speed of 120 mph

Disruption of turbine maintenance cycle, which depends on the availability of access vessels,
from January through March due to closure of the Great Lakes shipping season

Wind turbine reliability and production correspondingly decreased from January through
March, due to the inability to perform maintenance during this period.

Project Description

At 200 MW, the Great Lakes project is smaller than the Northeastern project. This theoretical
offshore wind project would consist of 55 Siemens SWT 3.6-MW120 turbines, placed on a
tripod foundation and 80 meter towers. Based on wind speed information available, the layout
designed is shown in Figure 3-40. This plot assumes a base area of ~ 6.4 3.64 km, with the
long side oriented toward the main wind direction. The layout features 11 rows of 5 WTG each
to accommodate the 55 machines.

Figure 3-39
Great Lakes project area

3-86

Wind Power

The Siemens 3.6-MW turbines have a rotor diameter of 120 meters. To accommodate these
appropriately, we have used a distance between the WTG of 660 m, equivalent to 5.5 rotor
diameters (5.5 120 m = 660 m) in mean wind direction, and 540 m (4.5 rotor diameters
120 m) perpendicular to the mean wind direction. We also assumed a 500 m safety distance
around the wind farm; such a safety zone is analogous to the 23 km2 required in some European
offshore wind farms. This 500-m strip is usually considered part of the wind farm and represents
a safety buffer between the facility and shipping lines, cable routes, pipelines etc. Offshore wind
farms are usually a restricted area for any public or recreational boating traffic.

Figure 3-40
Wind farm layout

Using this base area and layout, WTG orientation and distances, we expect the wind farm
efficiency to be 90%.

3-87

Wind Power

Cost Analysis
Estimated CAPEX

The Offshore Wind Economic Analysis Model utilized the design parameters and the cost
parameters to estimate CAPEX for the Great Lakes project. In this section we summarize the
design and boundary conditions of the wind farm and finish with a CAPEX calculation including
the major facility components.
The Great Lakes project is defined according to the design criteria and boundary conditions
summarized in Table 3-18.
Table 3-18
Design criteria and boundary conditions for Great Lakes project
Parameter

Value

Description/Comment

Rated elec. capacity

200 MW (198 MW)

No. and type of WTG

55 SWT-3.6-120

Rotor diameter

120 m

Hub height

80 m

Defined as meters above MSL

Location

Lake Michigan

Near Berrien County, MI

Distance from coast

20 km

Measured from center of wind farm

Water depth

3045 m

Tidal Range

N/A

Fresh-water lakevery low tides assumed

Wave heights

8m

KEMA estimate for 50-year wave height

min. 50C

Ambient temperature
range

max. ca. 35C

Location definitely requires special adaptation


of material types and heating systems (cold
climate version) as temps. are extremely low

Soil conditions

Mostly sand with areas of


mud and silt deposits

It is assumed that hammering piles is possible

Wind conditions

8.2 m/s average annual


mean wind speed

At hub height (80 m)

Ice conditions

Up to 0.2 m

During mid-December through mid-March

Foundations

Tripods

IAG design

6 half rings of 8 WTG


each
1 half ring of 7 WTG each

Rated voltage IAG

33 kV

OSS

With two transformers


33/150 kV

3-88

Rings are normally open and can be closed for


redundancy/all WTG are equipped with 3
feeder switch gears

In center of wind farm

Wind Power

Based on the cost descriptions and the preceding design criteria, the CAPEX for the Great Lakes
project is estimated in Table 3-19.
Table 3-19
Estimated CAPEX for Great Lakes project
Component

Sum
[M]

Sum
[M$]

[/kW]

[$/kW]

Comments

Wind turbine generators

330

462

43

1,650

2,310

Higher value due to hot climate


version

Foundations

180

252

24

900

1,260

Higher cost due to foundation


design

Inner array grid

28

39

140

196

Average value do to comparable


conditions fresh water

Offshore substation

40

56

200

280

Cost low due to location and


proximity to transmission grid

Export cabling and


onshore interconnection

48

67.2

240

Lower value due to small


336 distance to coast and low water
depth

100

140

13

500

Higher costs due to foundation


700 design, water depth and vessel
availability issues

42

59

210

294 Average value

768

1,075

100

3,840

5,376

77

108

10

384

536

845

1,183

110

4,224

5,914

Installation
Other (project
development, insurance,
project management, etc.)
Subtotal:
Contingencies
Total

Conservative assumption
some sources use 5%

On this basis, estimates for the total CAPEX of the Great Lakes Offshore Wind Project is
850 M ($1,190 M) or 4,252/kW ($5,952/kW).
In addition to CAPEX, TPI for the Great Lakes project was also estimated, as shown in Table 320. TPI assumes the TPC with the addition of construction loan and other financing costs. Other
assumptions of the TPI calculation include a nine-month construction cycle; 40% of the cost paid
in month 1 with the balance of costs levelized over the remaining eight months (at 50/50
debt/equity ratio and debt interest rate).

3-89

Wind Power
Table 3-20
Total plant investment, Great Lakes project
End of Year

Total Cash
Expended
TPC
(M$)

Before Tax
Construction
Loan Cost at
Debt Financing
Rate

2011 Value of
Construction Loan
Payments

Total Plant
Investment
(M$)

(M$)

(M$)
2011

$1,182.72

$60.01

$1,242.73

$1,242,73

2012

$0-

$0

$0

$0

TOTAL (TPI)

$1,242.73

Estimated O&M Costs

As described earlier, the Great Lakes project is an offshore wind farm of 55 turbines of 3.6 MW.
The project is positioned in the southwest of Lake Michigan. To enable quick response to any
failures, the facilities at Burns Harbor, Indiana, were found to meet all requirements at a distance
of 16 km from the wind farm.
As described earlier, estimating project-specific O&M expenditures is very dependent on many
different factors. Therefore, for this project the yearly O&M costs are in the range between $40
and $70 per installed kW. The following section details several site-specific cost drivers that
have been identified for the Great Lakes Project. These include the following:

Lake accessLarge parts of Lake Michigan are frozen during the winter months (especially
from January until March), which makes accessing the turbines by boat impossible during
this period. Turbines can be reached by helicopter, but this is also restricted by weather (e.g.,
icing conditions). This reduced accessibility may therefore result in long downtimes when a
failure occurs. As a result, revenue losses due to any failure will be higher than in other
months.

Low temperaturesLow temperatures contribute to reduced accessibility and may also


influence the functioning of the wind turbine. Turbines are designed to operate in a certain
range of weather conditions. When operating toward the limits of the design range, more
failures may occur. Lubrication may not function due to increased viscosity, or sensors may
trip more often.

When a turbine is in operation, components generate heat, which increases the temperature in
the turbine. When a failure occurs and a turbine shuts down, less heat is produced. This has
to be compensated by electrical heaters in the turbine. When an electrical failure results in
loss of power, this could lead to additional failures as the temperatures could drop below
design parameters.

Vessel availabilityOperators often secure availability contracts for vessels used regularly
for O&M activities. Vessels required for more specific operationsfor example, cablelaying vessels or jack-up vessels for exchanging large componentsare often purpose-built
and service many clients. The costs of these vessels are influenced by their availability and
the distance they must travel for each project. Until the Great Lakes region sees development

3-90

Wind Power

of significant offshore wind activity, vessels with the right capabilities may be required to
travel from the Atlantic coast or other distant harbors, resulting in long mobilization times
and high costs. In addition, because access to the Great Lakes through the Saint Lawrence
River is restricted by ice formation, large lifting vessels will not be able to reach the wind
farm during winter months, with corresponding impact on the availability of the wind farm.
To compensate for the high mobilization cost, clustering of maintenance activities should be part
of the O&M strategy. When clustering maintenance activities a heavy lifting vessel will be
chartered when, for example, a minimum of three turbines require the vessel for maintenance.
Clustering allows maintenance activities to be divided over the three turbines. A downside is the
increased downtime for the turbines because they will not be repaired until a certain threshold of
turbines require repair. For purposes of this analysis, the estimates O&M costs for the Great
Lakes project are $60/kW.
Estimated Energy Yield

To estimate the energy yield for the Great Lakes wind farm, the gross energy yield for the WTG
has been calculated, based on the turbines power curve. Please refer to Figure 3-41 for the
power curve of the Siemens STW-3.6-120 at 8.2 m/s average wind speed for this site. Table 3-21
summarizes the gross energy yield for the Great Lakes project.

Figure 3-41
Power curve of STW-3.6.120 (8.2 m/s)

As no detailed wind conditions are available from these hypothetical sites, we have assumed
normal offshore conditions, again based on experience in the North Sea. These have been
adapted to some extent for the Great Lakes project.

3-91

Wind Power
Table 3-21
Annual energy yield, Great Lakes project
55

Number of WTGs

200 (198)

Installed capacity (MW)


Farm efficiency (%]

90

Mean wind speed (m/s)

8.2

k-factor

2.0

Energy yield single WTG (MWh/yr)

14,921

Energy yield wind farm (MWh/yr)

820,655

Gross energy yield (MWh/yr) (incl. park losses)

738,590

We have used the values of the gross energy output in Table 3-21 to calculate the net energy
output, given a typical guaranteed availability and other losses. The results are summarized in
Table 3-22.
The technical availability of 95% for the wind farm is a realistic assumption and is quite
common industry value for the guaranteed technical availability. Electrical efficiency is expected
to be 99% for the Great Lakes project, with the assumption that the PCC, and thus the kWh
counter, is inside the wind farm. Losses due to neighboring wind farms have not been assumed.
Other losses, which are not covered by the common availability guarantee of the WTG supplier
(service provider), have also been considered. As discussed with the Northeastern US project,
these additional unavoidable losses include the 5% of operating hours not covered by the
manufacturers guarantee and therefore possible unavailability due to issues with the WTG, as
well as other forms of allowable or unavoidable standstills.
The following worst-case calculation based on typical O&M contracts shows common allowable
standstills per WTG. While these assumptions are similar to those for the Northeastern U.S.
project, the rationale is different. In the case of the Great Lakes project, estimates of downtime
take into account the colder temperatures and deeper water depth of the Great Lakes location.
This contrasts with the Northeastern project, where the downtime calculation reflects the larger
project size and greater distances.

Normal maintenance:

47 days/year (100168 h/a)

Waiting time for heavy lift vessel

35 days/year

Ice formation on rotor blades

13 day/year

Inaccessibility due to bad weather

510 days/year

Automatic cable unwinding

0.5 day/year

Grid faults

0.5 day/year

Total

3-92

1426 days/year (~47 % further losses)

Wind Power

A realistic estimate for the best case would be some 7 to 10 days of allowable standstills per
year, which equals to losses of some 2% to 3%. Based on these assumptions, estimated annual
energy yields of the Great Lakes wind farm are according to the values in Table 3-21.
Table 3-22
Estimated full load hours, Great Lakes project
Estimated Availability
Gross energy yield (MWh/yr)

Base Case

Worst Case

738,590

738,590

- Availability

96.0 %

95.0 %

- Electrical efficiency

99.0 %

98.0 %

2.0 %

7.0 %

687,900

639,500

6.9

13.4

3,440

3,200

- Other losses not covered by availability guarantee


Net energy yield (MWh/yr)
Total losses (%)
Full load hours

The above calculations show that realistic annual energy yields can be assumed to be in the
neighborhood of 688 GWh/yr for the Great Lakes project.
Economic Analysis

Given the costs and output discussed in the previous sections, the LCOE for the Great Lakes
project has been calculated. This calculation made use of the financial assumptions summarized
in Table 3-23.

3-93

Wind Power
Table 3-23
Financial assumptions, Great Lakes project
Rated plant capacity
Annual energy production at busbar (EPB)

450 MW
1,766,200MWh/yr

Therefore, capacity factor

44,80%

Year constant dollars

2012

Federal tax rate

35%

State

Michigan

State tax rate

9.00%

Composite tax rate (T)

41%

Book life

20 Years

Construction financing rate

8.0%

Common equity financing share

50%

Preferred equity financing share

0%

Debt financing share

50%

Common equity financing rate

10.5%

Preferred equity financing rate

0%

Debt financing rate

5.6%

Nominal discount rate before tax

8.1%

Nominal discount rate after tax

7.4%

Inflation rate
Real discount rate before tax

6.3%

Real discount rate after tax

5.6%

Federal investment tax credit (ITC)

0%

Federal production tax credit (PTC)

$0.022/kWh

Annual state investment tax credit

$ 10 Million

Total state investment tax credit limit


Renewable energy certificate (REC)
Annual REC limit

3-94

1.76%

$ 100 Million
$

0/kWh
$

Wind Power

Based on these assumptions, the LCOE for the Great Lakes project was calculated. EPRIs
methodology was used to yield the results summarized in Table 3-24.
Table 3-24
Levelized cost of electricity (LCOE), Great Lakes project
(TPI * FCR) +
(O&M+LOR)/EPB

Cost of Energy (COE) =


Nominal Rates
Total Plant Investment (TPI)

$1,242,726,852

Fixed Charge Rate (FCR)

10.55%

Annual O&M Cost (O&M)

$29,780,800

Levelized Overhaul & Replacement Cost (LO&R)


Energy Production (EPM)

$0
687,916 MWh/yr

Cost of EnergyTPI

19.0655 /kWh

Cost of EnergyO&M

4.3291/kWh

Cost of EnergyLO&R

0 /kWh

COE Nominal

$0.233947/kWh

COE Nominal

23.3947 /kWh
Real Rates

Total Plant Investment (TPI)


Fixed Charge Rate
Annual O&M Cost (O&M)
Levelized Overhaul & Replacement Cost (LO&R)
Energy Production (EPM)
Cost of EnergyTPI
Cost of EnergyO&M
Cost of EnergyLO&R

$1,242,726,852
10.02%
$29,780,800
$0
687,916 MWh/yr
19.4094 /kWh
4.3291 /kWh
0 /kWh

COE Real

$0.227386/kWh

COE Real

22.7386 /kWh

3-95

Wind Power

The higher costs for the Great Lakes project are due to following three factors:

Smaller project sizeThe lower economies of scale will increase the average cost per kW
for the project

Vessel availability and locationCurrently, the types of vessels used for offshore wind
project construction and operation do not exist on the Great Lakes. Movement of such
vessels from ocean ports may not be viable due to the limits of the Saint Lawrence Seaway.
The relocation of such vessels to the Great Lakes also increases project costs.

Deeper waterThe Great Lakes project is on average 25 m deeper than the Northeast
Project. This necessitated the use of tripod foundations, resulting in a significant increase in
foundation costs. The estimated increase in foundation costs is over $600/kW. This cost
differential can be expected to decrease as increased use of tripod foundations results in more
experience with installation.

Lake iceIce formation on Lake Michigan can affect design, construction, and operation of
the project. All foundations and the offshore substation need to be designed for the potential
ice loads that can occur on the Great Lakes. The analysis assumes the overall construction
period of the smaller Great Lakes project will be the same as the larger Northeastern project,
because it may not be possible to perform construction while ice is on the Lake Michigan.
Access for operations, maintenance, and repairs can be impacted by lake ice during winter
months.

3.8.5.3 United Kingdom


The United Kingdom installed its first offshore wind turbine at Blyth Harbor in 2000. Since then,
the UK has placed in operation approximately 1.6 GW of offshore wind power and has another
1.9 GW in construction and approximately 7 GW in some level of development. It is clear that
the UK is the world leader in offshore wind power development.
The UK has a national target of providing 15% of all of its energy consumption, including
transportation and heating fuels, from renewable energy sources by 2020. It is estimated that to
meet the 15% target, the UK may need to obtain 40% of its electricity from renewable energy
sources. If offshore wind power were to provide most of this renewable electric production, then
it is estimated the UK will need to build between 15 GW and 20 GW of offshore wind power.
The costs of offshore wind power have risen in the UK from 1.4 million/MW in 2000 to 3.6
million/MW in 2010. This increase has additional importance in light of the nations renewable
energy goals. Recent studies indicate that offshore wind may be the most costly of the UKs
options for renewable electric production.36 These studies attribute much of the major cost
increases for offshore wind to component cost increases. The offshore wind industry is still on a
very steep learning curve and suffers from a lack of certain critical resources. It is expected that,
as the manufacturing base for offshore wind components increases, the design and engineering
matures, and maritime vessel fleets increase and personnel are trained in offshore construction,
costs of offshore wind will decrease.

36

The Energy Challenge Energy Review Report 2006, Department of Trade and Industry, London, 2006.

3-96

Wind Power

As a result of concerns regarding the cost of offshore wind and the UKs renewable energy
targets, several studies have examined the economics of offshore wind. The following are the
reports on UK wind project costs used in the development of this section:

Great Expectations: The costs of offshore wind in UK waters understanding the past and
projecting the future, Technology and Policy Assessment Function of the UK Energy
Research Centre, September 2010

UK Electricity Generation Costs Update, Mott MacDonald, Brighton, UK, June 2010

UK Offshore Wind: Building an Industry Analysis and scenarios for industrial development,
Douglas Woods, Renewable UK, London, UK, June 2010

The Energy Challenge Energy Review Report 2006, Department of Trade and Industry,
Norwich, UK July 2006

These studies provide a clear view of the challenges facing the UK in developing offshore wind
projects and meeting its renewable energy goals. The balance of this section draws on these
reports and other data sources to present cost and economic data on offshore wind projects in the
UK.
The UK is developing its offshore wind projects in a series of rounds. Each round was created
when the UK auctioned leases for developing offshore wind projects. For the most part, the
projects with leases in Round 1 are either in operation or under construction. The first of the
Round 2 projects began construction in October 2007 (300 MW Thanet wind farm). There is
approximately 7 to 8 GW of offshore wind projects in Round 2. Approximately 656 MW of
Round 2 projects are in operation and another 1.9 GW under construction, as can be seen in
Table 3-25.
Table 3-25
UK Round 2 offshore wind project status
Location

Status

Capacity

Developer/Turbines

Triton Knoll

Pre-application

1,200

RWE npower renewable

Docking Shoal

Submitted (Dec 2008)

540

Centrica

Westernmost Rough

Submitted (Dec 2009)

240

DONG Energy

Race Bank

Submitted (Jan 2009)

620

Centrica

Dudgeon

Submitted (Jun 2009)

560

Warwick energy

London Array II

Approved (Dec 2006)

370

DONG Energy / E.ON UK


Renewables / Masdar

Gwynt y Mor

Approved (Dec 2008)

576

RWE Innogy / SWM /


Siemens

Humber Gateway

Approved (Feb 2011)

300

E.ON UK Renewables

West of Duddon Sands

Approved (Sep 2008)

389

ScottishPower / DONG
Energy

3-97

Wind Power
Table 3-25 (continued)
UK Round 2 offshore wind project status
Location

Status

Capacity

Developer/Turbines

Lincs

Under construction (Apr 2010)

270

Centrica / DONG /
Siemens Project Ventures

London Array 1

Under construction (Feb 2011)

630

DONG Energy /E.ON UK


Renewables /Masdar

Walney II

Under construction (Feb 2011)

184

DONG Energy

Sheringham Shoal

Under construction (Mar 2010)

317

Scira Offshore Energy


Limited

Greater Gabbard

Under construction (May 2009)

504

SSE Renewables & RWE


npower renewable

Gunfleet Sands I&II

Operational (Apr 2010)

173

DONG Energy (Siemens


3.6 MW turbines)

Walney I

Operational (Jul 2011)

184

DONG Energy

Thanet

Operational (Sep 2010)

300

Vattenfall (Vestas V90


3MW)

TOTAL MW

7,356

In Operation

657

Under Construction

1,904

The average distance from shore of the non-operating Round 2 projects is about 23 km, with the
closest being 9 km and the furthest 40 km. For the remainder of this section, the engineering and
economic evaluation of offshore wind in the UK will use the London Array 1 project as a basis
for its discussion. This is a 630 MW project located just off the Thames Estuary in the North Sea.
The project was chosen because it represents a project of considerable size and has development
characteristics that are average for those projects being developed in Round 2.
Physical Characteristics

For the UK offshore wind project assessed, the London Array Project Phase I was chosen,
described in the companys website (http://www.londonarray.com/the-project). Upon
completion, the project will be approximately 1 GW, with roughly 341 Siemens 3.6 offshore
wind turbines, four offshore substations, over 800 km of cabling and cost in excess of 4.5
billion ($6.3 billion). The discussion in this section will address the first phase of the project,
currently under construction and expected to be in full operation in 2013. Reasons for choosing
the project for this study include the following:

It is a new project currently under construction.

It is in medium water depth of only 212 meters.

It is larger than the two U.S. projects, which provides some indications of economy of scale
impacts.

3-98

Wind Power

It uses the Siemens 3.6 MW offshore wind turbine.

A significant amount of data is available on the project.

The UK project is defined according to the design criteria and boundary conditions
summarized in Table 3-26. These are the basis for the estimated cost elements in Table 3-27,
which are used in the economic modeling that follows.
Table 3-26
Design criteria and boundary conditions, UK project
Parameter
Rated capacity
No. and type of WTG

Value
630 MW
175 SWT-3.6-120

Rotor diameter

120 m

Hub height

80 m

Location
Distance from coast

North Sea
20 km

Water depth

1523 m

Tidal range

4.6 m

Wave heights
Ambient temperature range
Soil conditions
Wind conditions

9m

Thames Estuary
Measured from centre of wind farm

KEMA estimate for 50 year wave


height

max. ca. 40C


Mostly sand with areas of
mud and silt deposits
8.9 m/s average annual
mean wind speed
Very low

Foundations

Monopiles

IAG design

12 rings of between 7 and


18 WTGs

Offshore Substation (OSS)

Defined as meters above MSL

min. ca. 10C

Ice formation

Rated voltage IAG

Description / Comment

33 kV

Monopiles are hammered in


At hub height (80 m)
Assumed from temperature range and
ocean salt content

In center of wind farm

Two stations with two


transformers 33/150 kV

The following is a description of the London Array Phase I project.

3-99

Wind Power

Location

The London Array project is located in the North Sea off the Thames Estuary, about 20 km from
shore (see Figure 3-42). The total project area for both Phase I and II is estimated at 230 km2.
The expected average wind speed for the site is in the range of 8.9 m/s. Water depth for the site
ranges from 15 m to 25 m.

Figure 3-42
UK project locations

Project Ownership

The estimated total cost for London Array 1 is approximately 2.2 billion. The financial risk of
such a large project required the creation of a consortium so the risk could be spread across
several parties. The consortium for the London Array project has three players: Dong Energy is
the controlling partner with 50% ownership; E.ON Group has a 30% stake in the project; and
Masdar has the remaining 20% share of the project. Capsule descriptions of the three companies
follow.

DONG EnergyDenmark-based DONG Energy is a leading European energy group.


DONG Energy was founded in 2006 when six Danish energy companies merged: DONG,
Elsam, ENERGI E2, Nesa, Copenhagen Energy, and Frederiksberg Forsyning. They procure,
produce, distribute, and trade in energy and related products across northern Europe. DONG
Energy is a market leader in offshore wind technology and has built around half of the
offshore wind farms operating today. They are heavily involved in the production and
expansion of renewable energy in the UK. The company is involved in building three new
major UK offshore wind farms and currently operates the offshore wind farms Gunfleet
Sands (172 MW), Burbo Bank (90 MW) and Barrows (90 MW).

E.ON GroupGerman based E.ON Group is one of the worlds largest power and gas
companies. At facilities across Europe, Russia, and North America, more than 85,000
employees generated just under 93 billion in sales in 2010. They are a leading energy
supplier in the UK, with around 8 million customers. E.ON has been involved in renewable
energy since 1991, when they invested in their first onshore wind farm. They now own and
operate 22 wind farms in the UK including the 60-MW Scroby Sands offshore wind farm off
the coast of Great Yarmouth and the 60-turbine Robin Rigg Wind Farm in the Solway Firth.
Many more projects are in the pipeline.

3-100

Wind Power

MasdarMasdar, a renewable energy company based in Abu Dhabi, is a subsidiary of the


Mubadala Development Company. Masdar operates through five integrated units, including
an independent, research-driven graduate university, and seeks to become a leader in making
renewable energy a real, viable business and Abu Dhabi a global center of excellence in the
renewable energy and clean technology category. Masdar specializes in developing and
commercializing renewable and sustainable energy technologies. The London Array is just
one of many projects that the company is working on, both in Abu Dhabi and internationally.
Masdar is also investing in a number of innovative clean tech companies to further develop
their technologies. Their biggest project is Masdar City, a carbon-neutral, zero-waste
cleantech hub that is fully powered by renewable energy.

Wind Turbines

The London Array 1 will contain 175 Siemens 3.6 MW wind turbines (SWT-3.6-120), as shown
in Figure 3-43. This turbine has a rotor diameter of 120 meters. Hub height for the project will be
at 87 m. The Siemens contract for the wind turbines was placed at 1.0 billion. Installation of the
wind turbines and tower will be by a German and Danish group comprising of Per Aarsleff A/S
and Bilfinger Berger Ingeniuerbeau GmBH.

Figure 3-43
UK wind farm layout

Wind Turbine Foundations

Monopile foundations were chosen for this project because the water depth is 25 m or less. Each
monopole weighs approximately 268 tons. These will be driven 20 m to 50 m into the seabed and
capped with a transition piece to the turbine towers. The monopiles were supplied by a German
and Danish group (Per Aarsleff A/S and Bilfinger Berger Ingeniuerbeau GmBH). Estimated
value of the monopile supply and installation contract is 250 million.

3-101

Wind Power

Offshore and Onshore Substations

The London Array 1 will have two offshore and one onshore substations. The offshore substation
will step up the power from the wind farm to 150 kVA. The onshore substation is located 54 km
from the wind farm at Cleve Hill in Graveney, Essex. The Cleve Hill substation will step up the
power from the wind farm to 400 kV for interconnection with National Grids 400 kV
transmission system. Equipment supply and construction of all three substations was awarded to
Future Energy, a joint venture among Fabricom, Iemants, and Geosea. Estimated cost for the
offshore substations is 85 million each. The onshore substation is expected to cost 30 million.
Inner Array Cabling

The London Array 1 wind farm is expected to require 210 km of array cable to link in all 175
turbines to the offshore substations. JDR Cable Systems will supply the 210 km of 33 kV array
cables that will link the turbines to each other and to the offshore substations. The expected cost
of the array cabling is approximately 85 million.
Onshore Cabling

There will be four cables connecting the offshore substations to the onshore substations. The
cabling to be used will be 150 kV XLPE submarine power cables. Nexans Norway AS has been
awarded the contract to supply the 220 km of 150 kV subsea export cable connecting the
offshore substations to the shore.
Cost Analysis
Estimated CAPEX

The estimated CAPEX for this project (Table 3-27) is based on information obtained from
various sources. A press release in December 2009 discusses the execution of 2 billion in
contracts. The European Investment Bank, a finance bank for the project, estimated the total
CAPEX for this phase of the project at 2.2 billion GBP, 2.568 billion.37 This figure is used in
the analysis for this report.

37

London Array Offshore Windfarm, European Investment Bank, 08/0602010,


http://www.eib.org/projects/pipeline/2009/20090108.htm.

3-102

Wind Power
Table 3-27
Estimated CAPEX for UK project
Sum
[M]

Sum
[M$]

[/kW]

[$/kW]

1,167

1,634

50

1,853

2,594

Foundations

420

588

18

667

934

Higher cost due to foundation


design

Inner array grid

70

98

111

156

Average value do to comparable


conditions fresh water

Offshore substation

140

196

222

311

Cost low due to location and


proximity to transmission grid

Export cabling and


onshore interconnection

163

229

259

363

Lower value due to small


distance to coast and low water
depth

Installation

257

360

11

408

571

Higher costs due to foundation


design, water depth and vessel
availability issues

Other (project
development, insurance,
project management, etc.)

117

163

185

259

Average value

2,335

3,268

100

3,706

5,188

233

327

10

371

519

2,568

3,595

110

4,076

5,707

Component
Wind turbine generators

Subtotal
Contingencies
Total

Comments
Higher value due to hot climate
version

Conservative assumption
some sources use 5%

Because this project is structured financially differently from a U.S. project, the total CAPEX is
assumed to include construction finance costs. Thus, TPI equals CAPEX for this project.
Estimated O&M Costs

As described above, the London Array 1 Project is assumed to have 125 turbines of 3.6 MW,
located in the North Sea in the Thames Estuary. To facilitate quick response to any failures
occurring in the wind farm, qualified harbor facilities must be located as close to the wind farm
as possible. One harbor will be selected when there is a harbor facility in close proximity that
meets all the requirements. This would be the ideal situation since all operations can be
coordinated from one location. A detailed assessment of local options could also result that no
single nearby harbor meets all requirements. At that point, operators must determine whether to
select (a) a single harbor further away that meets all requirements but will result in longer travel
times, or (b) two harborsone harbor nearby for day-to-day O&M activities, for which the
harbor requirements are more easily met because only smaller vessels are used. A second harbor
situated further away would be used for coordinating O&M activities that require larger or
special vessels.

3-103

Wind Power

After a quick assessment, two harbors were selected for the London Array 1 Project. The harbor
that will be used for the day to day O&M are the ports of Ramsgate or Sheerness, both of which
lie approximately 50 km from the wind farm.
Estimating project-specific O&M expenditures depends on many different factors. O&M
operations are well understood in the North Sea. O&M costs per year range between $40 and $70
per installed kW. A report by Ernst & Young indicates that expected O&M costs should be in the
range of 44/kW to 70/kW.38 The Energy Research Center of the Netherlands report, Current
developments in wind2009: Going to great lengths to improve wind energy, gives an estimate
of $1.07/kWh (0.00764/kWh). For the purposes of this study, O&M costs will be set at 70/kW.
Estimated Energy Yield

To estimate the energy yield for the UK wind farm (Table 3-28), we refer to the gross energy
yields for the Siemens SWT-3.6-120 wind turbine that is assumed in all three cases. For the UK
Project, the assumed average wind speed of 8.9 m/s is shown in Figure 3-44.

Figure 3-44
Power curve of STW-3.6-120 (8.9 m/s)

38

Ernst & Young, Cost of and financial support for offshore wind: A report for the Department of Energy and
Climate Change, 27 April 2009, p. 9.

3-104

Wind Power
Table 3-28
Annual energy yield, UK project
Number of WTG

175

Installed capacity (MW)

630

Farm efficiency (%]

90

Mean wind speed (m/s)

8.9

k-factor

2.3

Energy yield single WTG (MWh/yr]

17,002

Energy yield wind farm (MWh/yr)

2,975,350

Gross energy yield (MWh/yr)


(incl. park losses)

2,677,815

We have used the values of the gross energy output in Table 3-28 for calculating the net energy
output, along with a typical guaranteed availability and other losses.
The assumed technical availability of the wind farm of 95% is realistic and is common industry
value for the guaranteed technical availability. Electrical efficiency is expected to be 98% for the
UK Project, assuming that the PCC, and thus the kWh counter, is inside the wind farm. Losses
due to neighboring wind farms are not assumed, as no other developments are known to be
planned in the immediate vicinity. Other losses, which are not covered by the availability
guarantee of the WTG supplier, have also been considered. These additional unavoidable losses
may arise from (a) the 95/5% manufacturer operational guaranteethat is, 5% of the operating
hours might be unavailable due to issues with the WTG; (b) other unavoidable and/or allowable
standstills, such as normal maintenance, customer and authority visits, ice formation on rotor
blades, no accessibility due to bad weather, grid faults, and so on.
In recognition of these factors, the following worst-case calculation was developed, based on
common O&M contracts and common allowable standstills per WTG:

Normal maintenance:

47 days/year (100168 hr/year)

Waiting time for heavy lift vessel

35 days/year

Ice formation on rotor blades

13 day/year

Inaccessibility due to bad weather

510 days/year

Automatic cable unwinding

0.5 day/year

Grid faults

0.5 day/year

Total

1426 days year (~47 % further losses)

A realistic estimate for the best case would be some 7 to 10 days of allowable standstills per
year, which equals to losses of some 2% to 3%. Based on these assumptions, the estimated
annual energy yield for the UK wind farm is as shown in Table 3-29.

3-105

Wind Power
Table 3-29
Estimate of full load hours, UK project
Estimated Availability
Gross energy yield (MWh/yr)

Base case
2,677,815

- Availability

96.0%

- Electrical efficiency

99.0%

- Other losses not covered by availability guarantee

2.0%

Net energy yield (MWh/yr)


Total losses (%)
Full load hours

2,494,095
6.9
3,924

The realistic annual energy yield for the London Array Phase 1 offshore wind farm can be
assumed to be roughly 2,495 GWh/yr for the UK Project.
Economic Analysis

Given the costs and output discussed in the previous sections, the Offshore Wind Economic
Analysis Model was used to calculate LCOE for the UK Project. This calculation used the
financial assumptions summarized in Table 3-30.

3-106

Wind Power
Table 3-30
Financial assumptions for the UK project
Rated plant capacity
Annual energy production at busbar (EPB)
Therefore, capacity factor

630 MW
2,494,095 MWh/yr
45.2 %

Year constant dollars

2012

Federal tax rate

35%

State tax rate


Composite tax rate (T)
Book life

0%
35%
20 Years

Construction financing rate

0.0%

Common equity financing share

50%

Preferred equity financing share

0%

Debt financing share

50%

Common equity financing rate

10.5%

Preferred equity financing rate

0%

Debt financing rate

5.6%

Nominal discount rate before tax

8.1%

Nominal discount rate after tax

6.9%

Inflation rate

1.76%

Real discount rate before tax

6.3%

Real discount rate after tax

5.2%

Federal investment tax credit (ITC)

0%

Federal production tax credit (PTC)

$0.0/kWh

Renewable energy certificate (REC)

$0/kWh

Annual REC limit

$0

Annual REC limit

$0

Based on these assumptions, the calculated the LCOE for the London Array Phase 1 Project is
summarized in Table 3-31. A breakdown of the estimated capital costs for selected UK
installations is given in Figure 3-45.

3-107

Wind Power
Table 3-31
Levelized cost of electricity, UK project
(TPI * FCR) +
(O&M+LOR)/EPB

Cost of Energy (COE) =


Nominal Rates
Total Plant Investment (TPI)

$3,594,752,000

Fixed Charge Rate (FCR)

11.33%

Annual O&M Cost (O&M)

$82,271,280

Levelized Overhaul & Replacement Cost (LO&R)


Energy Production (EPM)

$0
2,494,095 MWh/yr

Cost of EnergyTPI

16.3244 /kWh

Cost of EnergyO&M

3.2986 /kWh

Cost of EnergyLO&R

0 /kWh

COE Nominal

$0.196231/kWh

COE Nominal

19.6231 /kWh
Real Rates

Total Plant Investment (TPI)


Fixed Charge Rate
Annual O&M Cost (O&M)
Levelized Overhaul & Replacement Cost (LO&R)
Energy Production (EPM)
Cost of EnergyTPI
Cost of EnergyO&M
Cost of EnergyLO&R

3-108

$3,594,752,000
10.90%
$82,271,280
$0
2,494,095 MWh/yr
15.7139 /kWh
3.2986 /kWh
0 /kWh

COE Real

$0.190125/kWh

COE Real

19.0125 /kWh

Wind Power

Figure 3-45
Breakdown of estimated capital costs for selected installed offshore UK wind plants
Source: Study of the Costs of Offshore Wind Generation-A Report to the Renewables Advisory
Board and DTI, Offshore Design Engineering (ODE) Limited, Sept 14, 2006

3.9 On-Shore Technology Performance and Cost Tables


This section addresses the performance and capital, O&M, and levelized cost of electricity
estimates for five representative utility-scale onshore wind power plants. The estimates are based
on data presented in the 2012 EPRI report, Engineering and Economic Evaluation of UtilityScale Wind Power Plants [36].
3.9.1 Site Assumptions
Table 3-32 summarizes the rated capacities and locations of the eight onshore wind project cases.
Four U.S. sites were chosen in the western (California), south-central (Texas), northeast (New
York) and northwest (Washington) where extensive wind development is already occurring; one
site is in the north central region (Michigan), where initial wind power development is under
way; and one is in the southeast, where there is currently no wind development. For the southeast
location, it was decided to assume a conceptual site in Georgia where there may be potential for
onshore wind energy development. Both the Brazil and Australia sites were selected because
wind development is currently active in these countries.

3-109

Wind Power
Table 3-32
Project descriptions

The onshore project rated capacities were chosen to be consistent with typical current installed
capacities in each state within the last four years. For the Georgia site, the rated capacity is
assumed to be approximately 50 MW, which is similar to that of the northeast, which has similar
land constraints.
Table 3-33 presents the conceptual layout design assumptions for each site. The projects are
based on a generic 2-MW wind turbine, which represents the average onshore turbine power
rating for recent installations. Generic turbine models were chosen to avoid favoring a specific
turbine manufacturer. The wind turbine spacing and wind rose for each site were based on NREL
and private party wind data and consider the typical site terrain, infrastructure and other
obstacles allowed. The wind turbine crosswind and downwind separations inTable 3-33 are
given in rotor diameters.

3-110

Wind Power
Table 3-33
Conceptual layout design assumptions

Note that none of these assumptions is representative of conditions for the entire state or region.
The costs and energy capture values used in this report are meant to be examples; they are not
meant to represent average costs and performance for every potential project within the state.
For all cases, a shallow-depth concrete gravity foundation (spread footing) is assumed to be used.
At present, this foundation type is the most commonly used design for onshore wind energy
applications throughout the United States and the world. These foundations typically have
octagonal bases bearing on either native soil or engineered fill and are normally backfilled with
compacted native soil with the pedestals protruding above the finished grade.
3.9.2 Plant Performance
Plant performance is measured by the annual capacity factor, which is the annual kilowatt-hour
energy production as a fraction of the energy that would be produced if plant operated at its rated
capacity 100% of the time. For wind plants, it is a function of average of wind speed, wind speed
frequency distribution, energy losses due to planned and unplanned outages and electrical losses,
soiling and icing of wind turbine blades, wake turbulence, and other losses. The annual capacity
factors presented in this report were developed by analyzing performance data from modern
wind farms that were installed no earlier than 2004. The assessment also excluded wind farms
that had less than six months of available operating data. Ventyx Velocity Suite was used to
gather the data.
The resulting annual capacity factor estimates for the five conceptual onshore and one offshore
wind plants in the United States, as well as those in Brazil and Australia, are as follows:

California: 33%

Texas: 41%
3-111

Wind Power

Michigan: 33%

New York: 30%

Washington: 30%

Georgia: 29%

Brazil: 42%

Australia: 38%

3.9.3 Total Capital Requirement


Table 3-34 summarizes the total capital requirement estimates for the main cost categories listed
above for each of the six U.S. wind farm locations, as well as Brazil and Australia. The costs do
not include financing and legal fees, but do include interest during construction (AFUDC), and
owners costs during construction.
The total capital requirement consists of the TPC, escalation and interest during construction,
and owners costs. The TPC includes wind turbine procurement and transportation, balance-ofplant direct labor and material, and indirect and EPC construction costs. The TPI is the sum of
TPC and escalation and interest during construction (also known as allowance for funds used
during construction, or AFUDC). Total capital requirement is the sum of TPI and owners costs
(due diligence, permitting, legal, development, taxes, and fees). Larger projects generally benefit
from economies of scale; this is reflected in the cost estimates.
Table 3-34
Total capital requirement estimates summary (2011 US$)

Source: DNV Project Cost Database (2011 USD$)


Note: Assumes Brazilian Real (R$) to U.S. dollar (US$) exchange rate of 2.08; assumes Australian dollar (AU$) to
U.S. dollar (US$) exchange rate of 0.98.

The total capital requirement estimates for the onshore wind projects range from a little less than
$2,000/kW for the 100-MW Brazilian wind plant to about $2,500/kW for the 100-MW
Australian plant. The estimate for the most expensive domestic wind farm, the 50-MW Georgia
3-112

Wind Power

site, can be attributed to the turbine cost for the low-wind speed region. Both the hub height and
rotor diameter are significantly increased for the Southeast. In addition, it does not have the
benefit of economies of scale, as it is only 50-MW because land constraints and it reflective of
the region being in the early stages of wind technology adopotion.
3.9.4 Operation and Maintenance Cost
Both plant-provided and contracted O&M costs were evaluated for this study. Plant-provided
O&M (plant O&M) is an O&M strategy whereby the owner of the wind farm provides personnel
and equipment for upkeep of the facility. All costs and liability associated with the failure of
plant equipment, other than equipment covered by a manufacturers warranty, are borne by the
owner. Contracted O&M strategy assumes that O&M costs are outsourced by the owner to a
third-party firm (often the original equipment manufacturer [OEM]). The third-party contractor
provides all necessary technical personnel for operation and maintenance of the facility.
For this study it was assumed that O&M costs associated with the WTGs are covered by the
WTG manufacturer for the first two years of service under the wind turbine purchase agreement.
We also assumed that an extended WTG warranty is purchased by the owner for an additional
three years of coverage (contract O&M). During this five-year contract O&M period we
assumed that the owner is only responsible for maintenance of the balance-of-plant including a
site manager and minimal technical support for the contract O&M provider. The purchase of the
extended WTG warranty is assumed to cost nominally $40,000 per WTG per year and adjust
both for inflation and project size.
O&M costs considered include the following:

Fixed labor costs for the operation and maintenance of the facility, including administrative
expenses
Maintenance costs, consisting of spare parts, consumables, cranes, and the like for the WTGs
and balance of plant (BOP) equipment
Other expenses, including insurance, property tax, land lease fees, and management
Fixed labor costs include staffing the required maintenance personnel. Only minimal costs are
considered for the contracted O&M scenario, as only a site manager and minimal technical staff
are required to support the contracted O&M team.
Maintenance costs include scheduled and unscheduled WTG maintenance and BOP
maintenance. Scheduled WTG maintenance is the routine, planned work on the turbines, based
on the OEM manuals (commonly performed every six months per WTG). Unscheduled WTG
maintenance is a budgetary value for covering the unexpected repairs of WTG equipment when
equipment breaks. This budgetary value includes costs for labor, materials and crane rental. BOP
maintenance includes upkeep in the project collection substation, meteorological towers, pad
mounted transformers and all other BOP equipment.
Land lease agreement costs are site specific and may be dependent on individual negotiations
with land owners. A land lease agreement with a private land owner may require a fixed payment
plus a percentage of gross revenue. Other types of land lease payment agreements may only
entitle the land owner to a percentage of gross revenues of all WTGs located on the property.
Some agreements may also include provisions for land use that do not include WTGs, such as
3-113

Wind Power

access roads, transmission lines, and the like. Other types of land agreements might involve
leases from a federal agency such as the Bureau of Land Management (BLM). For this estimate,
annual land lease costs are assumed to be 2.5% of gross energy revenue.
Insurance costs are assumed to be 0.2% of total plant cost.
Property taxes are assumed to be 0.2% of total plant cost and are included in the O&M costs.
Tables 3-35 and Table 3-36, respectively, summarize plant-provided and contracted O&M
averaged over the first five years of the project life. Costs are thousands of dollars per year
unless otherwise noted. Comparing the two tables, one can see that, on average, the total contract
O&M cost ($/kW/yr) exceeds the cost of plant O&M at each site. This reflects the fact that the
contract O&M provider (often the manufacturer) assumes all risk both for repair and
maintenance as well as for any performance or availability guarantees that are in place. Also
evident in Table 3-35 is the influence of project size on O&M costs. Although several factors
influence the differences in O&M cost among the different regions, one can see that, in general,
larger projects (e.g., Texas and Washington) have lower O&M costs per kW than smaller
projects (e.g., New York and Georgia).
For a wind plant in Georgia, O&M costs are also affected by the taller tower. The O&M cost
estimator assumes that crane costs increase by approximately 50% when tower height increases
from 80 m to 100 m. Over the 20-year project life, that translates into a 10% increase in WTG
maintenance costs (not including BOP or other O&M costs). However, low labor rates in that
region of the country mask, to some extent, the impact of a larger crane. In addition, the full
impact of increased crane costs tends to occur later in the project life as parts failure rates
increase. The net result for Georgia as that even with a taller tower, during the first five years of
operation the O&M costs for the example project in that region were slightly lower than for a
similar-sized project in New York. For contract O&M, however, the total costs for the Georgia
wind plant exceed those of the New York wind plant.
Table 3-35
Plant-provided O&M costs (2011$), average of years 15

3-114

Wind Power
Table 3-36
Contracted O&M costs (2011$), average over extended warranty period (years 15)

*Includes only site manager and minimal support labor for contract O&M.
**Includes warranty/extended service contract fee, plus BOP.

Figures 3-46 and Figure 3-47 show a breakdown of the total O&M costs for a representative
project in California. Figure 3-46 includes O&M costs related to labor and maintenance, and
Figure 3-47 shows a breakdown of other O&M costs. The first column in each figure shows
costs for contract O&M averaged over the five-year extended warranty period. The remaining
columns in each figure show average plant O&M in five-year increments over the life of the
project. Notable in Figure 3-46 is that plant-provided labor and maintenance O&M cost for years
1 through 5 is approximately 18% lower than the cost of the equivalent level of O&M provided
by the OEM. When the OEM provides O&M service under a warranty and/or maintenance
agreement they take on more risk than the plant operator would. The plant owner is typically
willing to pay extra for contract O&M for the first year or two of a project to avoid risk
associated with infant mortality failures and the potential for reduced WTG availability in the
first six to 12 months of operation. Despite the increased cost of contract O&M compared with
plant O&M, there are several reasons that the plant-owner may extend the contract O&M beyond
the typical two-year warranty period. These include the following:

Capability for remote monitoring from a central location. This capability has the potential to
anticipate problems and thereby increase WTG availability.

Direct and timely WTG upgrades and service letters.

Ready access to spares potentially leading to increased availability.

Over time the cost of plant-provided O&M increases, as shown in Figure 3-46, due partly to
inflation but also due to increased parts replacement cost. Parts replacement cost (including
hardware, crane and additional labor) increases faster than inflation because failure rates
increase as hardware ages. Figure 3-46 shows that whereas in the early years of a project
labor costs compose nearly half of all O&M costs (excluding other O&M costs), parts
replacement cost dominates O&M cost in the later years.
3-115

Wind Power

Figure 3-46
California plant-provided O&M costs over time compared with contracted O&M costs in
years 15 (inflation rate = 2.5%)

The other O&M expenses shown in Figure 3-47 include administrative, financial and legal costs,
insurance costs, property taxes and lease payments to landowners. We assumed these costs,
which grow with inflation, to be similar for contract O&M or plant O&M over the same time
period. Plots showing other O&M costs for the remaining regions look very similar to Figure
3-47 and therefore have not been included.

3-116

Wind Power

Figure 3-47
California plant-provided other expenses over time (inflation rate = 2.5%)

3.9.5 Levelized Cost of Electricity


A probabilistic analysis of the LCOE was performed for each of the wind sites evaluated in the
previous sections of this report. This analysis was performed using the risk analysis software
@Risk, which uses Monte Carlo simulation to estimate the range of possible outcomes and the
probabilities of these outcomes based on varying selected input parameters to the model. The
LCOE for each site was evaluated by performing 10,000 iterations to generate the probability
distribution of the probable LCOE values. The probability distribution of the LCOE values is
then used to estimate the probability that the LCOE is below a given value.
The assumptions made for the varied parameters were based on the most likely values for capital
cost, plant performance (capacity factor), operations and maintenance costs, and financing
assumptions made for each site. Note that debt and equity rates were selected taking into account
the maturity of the wind market as well as the typical values for a given region or country. A
maximum and minimum range, or standard deviation, was then assumed for each variable,
applicable to the type of statistical distribution assumed. Table 3-37 shows the assumed ranges,
their probability distributions and related assumptions.

3-117

Wind Power
Table 3-37
Wind probabilistic analysisvaried parameters
Initial
Value

Varied Parameter

Statistical Distribution and


Distribution Assumptions

(Most Likely)
Capacity Factor (%)

Normal

Wind Plant 1 (CA)

33%

Bounded at 0% and 100%

Wind Plant 2 (TX)

41%

= 0.1 * initial most likely


value

Wind Plant 3 (MI)

33%

Wind Plant 4 (NY)

30%

Wind Plant 5 (WA)

30%

Wind Plant 6 (GA)

29%

Wind Plant 7 (Brazil)

42%

Wind Plant 8 (Australia)

38%

Total Plant Cost ($/kW)

Triangular

10% / +15% of most likely

Wind Plant 1 (CA)

$2,022

Wind Plant 2 (TX)

$1,842

Wind Plant 3 (MI)

$2,154

Wind Plant 4 (NY)

$2,161

Wind Plant 5 (WA)

$2,100

Wind Plant 6 (GA)

$2,408

Wind Plant 7 (Brazil)

$1,922

Wind Plant 8 (Australia)

$2,581

value

Fixed O&M ($/kW/yr)


(annual average cost over 20 year project life,

Triangular

including 2.5% inflation, except Brazil)

10% / +15% of most likely

Wind Plant 1 (CA)

$59

Wind Plant 2 (TX)

$50

Wind Plant 3 (MI)

$58

Wind Plant 4 (NY)

$63

Wind Plant 5 (WA)

$57

Wind Plant 6 (GA)

$65

Wind Plant 7 (Brazil) (5.0%


inflation)

$75

Wind Plant 8 (Australia)

$69

3-118

value

Wind Power
Table 3-37 (continued)
Wind probabilistic analysisvaried parameters
Varied Parameter

Initial Value
(Most Likely)

Debt Rate*

Statistical Distribution and


Distribution Assumptions
Triangular

Wind Plant 1 (CA)

5.5%

Wind Plant 2 (TX)

5.5%

Wind Plant 3 (MI)

6.0%

Wind Plant 4 (NY)

5.5%

Wind Plant 5 (WA)

5.5%

Wind Plant 6 (GA)

7.0%

Wind Plant 7 (Brazil)

7.9%

Wind Plant 8 (Australia)

7.0%

Equity Rate*

2 percentage points

Triangular

Wind Plant 1 (CA)

6.5%

Wind Plant 2 (TX)

6.5%

Wind Plant 3 (MI)

7.0%

Wind Plant 4 (NY)

6.5%

Wind Plant 5 (WA)

6.5%

Wind Plant 6 (GA)

8.0%

Wind Plant 7 (Brazil)

10.0%

Wind Plant 8 (Australia)

10.0%

1.5 percentage points

*Values provided by EPRI

Financial Assumptions, Including Incentives


The financial assumptions presented in Table 3-38 include those relevant to the calculation of the
levelized cost of energy but were not varied (static) in the probabilistic analysis. We assumed
that the evaluated U.S. plants will receive the production tax credit (PTC) and five-year
accelerated depreciation (Modified Accelerated Cost Recovery System [MACRS]), that Brazil
will receive a reduction in tax rate, and that Australia will receive Clean Development
Mechanism (CDM) carbon credits worth $13 per MWh.

3-119

Wind Power
Table 3-38
Static parameters for financial analysis including incentives
Plant Lifespan and Financial Timing Assumptions

Reference Year for Cost and Economic Assumptions

2011

Units

Operating Period (project lifespan)

20

years

Fixed O&M Escalation


U.S. and Australia
Brazil

2.5
5.0

%/yr

Variable O&M Escalation


U.S. and Australia
Brazil

2.5
5.0

%/yr

Debt-to-Equity Ratio

60/40

%/%

Debt Financing Term

10

years

Depreciation method (U.S. only)

5-year

MACRS

Tax Rate
U.S. and Australia

35

Operations and Maintenance (O&M)

Capital Cost

Tax-Related Assumptions

Brazil

20

Production Tax Credit (U.S. only)


CDM carbon offset credit (Australia only)

22
13

$/MWh
$/MWh

General Inflation
U.S. and Australia
Brazil

2.5
5.0

Sensitivity Analysis, Including Incentives


Figure 3-48 is a tornado chart showing the sensitivity of the LCOE for the California case. The
parameters are ordered according to the Spearman rank correlation coefficient, which is a
measure of the statistical dependence between LCOE and each of the five varying parameters.
The tornado charts for the other sites are similar. The chart shows that the LCOE is most
sensitive to the net capacity factor, followed by the capital cost, return on equity, interest on debt,
and fixed O&M cost.

3-120

Wind Power

Figure 3-48
Sensitivity of LCOE to varied parameters (California case)

Probability Distributions
Table 3-39 summarizes the initial, mean (P50), and 95th percentile (P95) values of the LCOE,
and Figures 3-49 through Figure 3-56 present the probability distributions of the LCOE for each
of the sites. The figures show the minimum, maximum, and 5%, 50%, and 95% values of the
LCOE at which the LCOE is less than values shown in the figures. LCOE values are presented in
2011 dollars.
Table 3-39
Initial, mean, and 95th percentile LCOE values, including incentives (2011$)
Results

LCOE initial
value ($/MWh)

Mean
($/MWh)

95th
Percentile

Wind Plant 1 (CA)

$47.55

$49.57

$59.01

Wind Plant 2 (TX)

$28.71

$30.65

$37.63

Wind Plant 3 (MI)

$52.23

$54.67

$65.08

Wind Plant 4 (NY)

$60.09

$62.07

$73.21

Wind Plant 5 (WA)

$55.85

$58.33

$68.99

Wind Plant 6 (GA)

$77.63

$80.46

$94.40

Wind Plant 7 (Brazil)

$60.30

$62.66

$73.07

Wind Plant 8 (Australia)

$88.97

$91.29

$105.72

3-121

Wind Power

Figure 3-49
LCOE probabilistic analysis resultswind plant 1 (California)

Figure 3-50
LCOE probabilistic analysis resultswind plant 2 (Texas)

3-122

Wind Power

Figure 3-51
LCOE probabilistic analysis resultswind plant 3 (Michigan)

Figure 3-52
LCOE probabilistic analysis resultswind plant 4 (New York)

3-123

Wind Power

Figure 3-53
LCOE probabilistic analysis resultswind plant 5 (Washington)

Figure 3-54
LCOE probabilistic analysis resultswind plant 6 (Georgia)

3-124

Wind Power

Figure 3-55
LCOE probabilistic analysis resultswind plant 7, tax incentive (Brazil)

Figure 3-56
LCOE probabilistic analysis resultswind plant 8, CDM (Australia)

3-125

Wind Power

3.9.6 Levelized Cost of Electricity with No Incentives


In a similar manner to the previous section, a probabilistic analysis of the LCOE was performed
using @Risk for each of the wind sites, where it was assumed that no incentives (e.g., no PTC,
CDM carbon credit, or reduction in taxes) would be available. The probability distribution of the
LCOE values was then used to estimate the probability that the LCOE is below a given value.
The assumptions made for the varied parameters were based on the most likely values for capital
cost, plant performance (capacity factor), O&M costs, and financing assumptions made for each
site. A maximum and minimum range, or standard deviation, was then assumed for each
variable, as applicable to the type of statistical distribution assumed. Table 3-40 shows the
assumed ranges, their probability distributions and related assumptions. In this case, all
incentives (PTC, carbon credit, tax incentives) assumed to be zero.
Financial Assumptions with No Incentives
The financial assumptions presented in Table 3-40 include those relevant to the calculation of the
LCOE but were not varied (static) in the probabilistic analysis. For this sensitivity, it is assumed
that all of the evaluated plants for the U.S. market will not receive the $22 per MWh production
tax credit (PTC), but will receive five-year accelerated depreciation (MACRS). For this scenario,
Brazil and Australia do not receive the MACRS incentive or any other incentive.

3-126

Wind Power
Table 3-40
Static parameters for financial analysis without incentives
Plant Lifespan and Financial Timing Assumptions

Reference Year for Cost and Economic Assumptions

2011

Units

Operating Period (project lifespan)

20

years

U.S. and Australia

2.5

%/yr

Brazil

5.0

Operations and Maintenance (O&M)


Fixed O&M Escalation

Variable O&M Escalation


U.S. and Australia

2.5

Brazil

5.0

%/yr

Capital Cost
Debt-to-Equity Ratio

60/40

%/%

Debt Financing Term

10

years

Depreciation method, U.S. only

5-year

MACRS

Tax Rate

35

Production Tax Credit (U.S. only)

$/MWh

CDM carbon offset credit

$/MWh

U.S. and Australia

2.5

Brazil

5.0

Tax-Related Assumptions

General Inflation

Sensitivity Analysis with No Incentives


Figure 3-57 is a tornado chart showing the sensitivity of the LCOE for the California case with
no incentives. The tornado charts for the other sites are similar. The chart shows that the LCOE
is most sensitive to the capacity factor, followed by the capital cost, return on equity, interest on
debt, and fixed O&M cost.
Probability Distributions
Table 3-41 summarizes the initial, mean (P50), and 95th percentile (P95) values of the LCOE
with no incentives, and Figures 3-58 to Figure 3-65 present the probability distributions of the
LCOE for each of the sites. The charts show the minimum, maximum, and 5%, 50%, and 95%
values of the LCOE values at which the LCOE is less than values shown in the figures. LCOE
values are presented in constant, study year dollars.

3-127

Wind Power

Figure 3-57
Sensitivity of LCOE to varied parametersno PTC (California case)
Table 3-41
Initial, mean, and 95th percentile LCOE values with no incentives (2011$)
Results

LCOE,
value

initial

Mean
($/MWh)

95th
Percentile

($/MWh)
Wind Plant 1 (CA)

$68.78

$70.79

$80.31

Wind Plant 2 (TX)

$49.79

$51.76

$58.80

Wind Plant 3 (MI)

$73.83

$76.26

$86.59

Wind Plant 4 (NY)

$81.31

$83.31

$94.63

Wind Plant 5 (WA)

$77.08

$79.55

$90.56

Wind Plant 6 (GA)

$99.96

$102.79

$116.80

Wind Plant 7 (Brazil)

$63.87

$66.21

$75.21

Wind Plant 8 (Australia)

$102.98

$105.30

$120.35

3-128

Wind Power

Figure 3-58
LCOE probabilistic analysis resultswind plant 1, no PTC (California)

Figure 3-59
LCOE probabilistic analysis resultswind plant 2, no PTC (Texas)

3-129

Wind Power

Figure 3-60
LCOE probabilistic analysis resultswind plant 3, no PTC (Michigan)

Figure 3-61
LCOE probabilistic analysis resultswind plant 4, no PTC (New York)

3-130

Wind Power

Figure 3-62
LCOE probabilistic analysis resultswind plant 5, no PTC (Washington)

Figure 3-63
LCOE probabilistic analysis resultswind plant 6, no PTC (Georgia)

3-131

Wind Power

Figure 3-64
LCOE probabilistic analysis resultswind plant 7, no incentives (Brazil)

Figure 3-65
LCOE probabilistic analysis resultswind plant 8, no incentives (Australia)

3-132

Wind Power

3.10 Grid Integration


3.10.1 Production Variability
Even though wind is an intermittent generation resource, generation output from wind farms is
more stable than the term intermittent may suggest. Grid-connected wind farms typically have
capacity factors ranging from 25% to 40% or more over the course of a year. A common
misconception about wind power is that turbines are either on or off and sit dormant for much of
the year. In actuality, wind turbines are capable of partial output, and most wind farms generate
at some level during 70% to 90% of the year. Further, there are a number of factors that smooth
the output of a wind farm relative to variations in actual wind speed.
First, wind turbines do not begin to generate power until winds reach cut-in speed, usually 4 to
5 m/s, and will generate at the rated capacity at or above the rated wind speedtypically 12 to
14 m/s. This creates a floor and a ceiling over the range of wind speeds at which wind
turbine output will vary. In addition, the turbine rotor of variable-speed machines acts like a
flywheel, capturing the energy of sudden gusts and converting it to power more smoothly.
Second, the wind at modern turbine tower heights is steadier than the winds closer to the ground.
Third, geographic dispersion of wind turbines has a powerful mitigating effect on wind plant
output variation. Several studies have demonstrated this fact with wind plants spread among
multiple sites, but the smoothing effect of geographic dispersion also occurs within a single wind
farm. Individual wind gusts are typically very local phenomena, with a breadth of 30 m or less.
Because even a medium-sized project would be spread over several miles of land, individual
turbines do not experience the same wind speed simultaneously and ramp up and down at
differing rates.
Although variability over time due to wind speed changes can pose integration challenges, wind
power generally varies less rapidly than customer demand. Experience to date has shown that at
penetration levels of up to 10% to 20% of delivered total system load measured on an energy
basis and for large interconnected electricity grids with no significant transmission bottlenecks,
wind variability has little impact relative to the larger, shifting variability of utility system
demand. In addition to systems with transmission constraints, island and other systems that are
not interconnected may be affected at relatively low levels of wind penetrationespecially those
with low nighttime loads and combined heat and power generation. That being said, the more
accurately these output variations can be anticipated and planned for, the more reliably and
economically wind can be integrated into a large utility system.
Alternately, energy storage and backup power generation resources at strategic locations in the
electricity grid can help alleviate some of the integration issues related to wind power. In fact,
some turbine manufacturers, such as GE, are reportedly working to embed inertial storage
capability directly into their turbine models to alleviate grid ramping issues. Cost of production
issues are stymieing commercial rollout. Application of energy storage and other integration
technologies to wind and other renewable technologies is more comprehensively addressed in
Chapter 12.

3-133

Wind Power

3.10.2 Ancillary Service Costs


In the last few years, several empirical studies have been performed in the United States
analyzing the impacts and ancillary services costs of wind power plant integration into specific
utility systems. Work is ongoing to establish methodologies for identifying integration costs
more accurately. Good utility practice for integrating wind will evolve for some time.
However, some initial conclusions can be drawn from work to date.
Recent studiescarried out in cooperation among NREL, other researchers, and utility operators
and plannersanalyze the impacts and costs to integrate wind power into several different utility
systems, including Xcel Energy, PacifiCorp, and BPA. The results of the Xcel Energy evaluation
are presented in the Utility Wind Interest Group report, Characterizing the Impacts of Significant
Wind Generation Facilities on Bulk Power System Operations Planning, Excel Energy-North
Case Study (May 2003), also available from EPRI (EPRI 1004807, December 2003). The studies
estimate ancillary service costs for a range of utility operations time frames including unit
commitment and reserves, load following, and regulation. The methodologies vary to some
extent between the studies given differences in the wind regimes modeled, the degree of wind
penetration, the mix of utility generation assets, and the cost structures. However, a common
finding among all the studies is that, overall, impacts and costs are relatively low at low wind
penetration levels. Total ancillary cost estimates ranged from negligible costs in the case of
BPAs system, to $1/MWh for Xcels Minnesota system, to a high of $5$6/MWh for
Pacificorps system to integrate 2,000 MW of wind generationequal to nearly 20% of system
capacity. A paper released in 2003 summarizes these studies and provides references to the
specific analyses [4].
Building on these studies, the California Energy Commission (CEC) sponsored a similar detailed
study of the cost to integrate renewables into the states larger power system in support of the
California Public Utility Commissions (CPUC) RPS implementation efforts. The study is
motivated by the RPSs least-cost, best-fit bid selection criterion, which requires that indirect
costs be considered in addition to the energy bid price when selecting eligible renewable
projects. Accordingly, the three-phase study quantifies the costs of integrating various renewable
technologies, including wind power.
Phase I analyzes the costs to integrate existing renewables [5]. Phase II of the study identifies
and analyzes the key attributes of renewable generators that affect integration cost. Phase III
[http://www.energy.ca.gov/reports/500-04-054.PDF] provides a final methodology to model the
characteristics and costs of renewables in California. Given the transparent and open approach of
the study, broad participation, extensive data availability from the Cal-ISO, and depth of the
analysis, the California study will likely serve as a benchmark for future efforts that attempt to
accurately assess the integration costs for renewables, including wind power, into larger utility
systems.
During 2004, two additional wind integration studies were completed in the United States. One
sponsored by the Minnesota Department of Commerce addressed the impacts of expected higher
wind penetration in the future. The other, sponsored by the New York State Energy Research and
Development Authority (NYSERDA) addressed the impacts of wind development on the New
York State electricity system. Recent wind integration studies are summarized in the 2006 EPRI
report, Survey of Wind Power Integration Studies (1011883, 2006).
3-134

Wind Power

3.10.3 Wind Energy Forecasting and Scheduling


In recent years, the art of forecasting wind plant generation over the next zero to 48 or
72 hours has advanced rapidly. Current forecasting systems can provide system operators with
information necessary to integrate intermittent wind generation in the most reliable and
economic manner. Wind production forecasts can allow a system operator to schedule a wind
plants capacity and energy with some accuracy, effectively avoiding some capacity and fuel
costs while maintaining reliability standards. These systems also help wind plant operators
schedule turbine maintenance. Several companies and organizations in the United States, Europe,
and elsewhere in the world offer wind energy forecasting services and are providing next-hour,
next-day, and extended forecasts to wind plant operators, utility system operators, and electric
utilities.
Wind forecast systems integrate regional weather and project production models, utilizing
real-time inputs from on-site and off-site weather stations and wind plant SCADA systems. They
produce rolling output forecasts for periods of up to a month or more in advance, which can be
updated as frequently as every hour. The accuracy of these forecasts increases dramatically with
the proximity of the hour forecast. The week-ahead average forecast error is typically in the
range of 20% of nameplate. The average error for a forecast provided 60 minutes ahead for the
next hour is in the 8% range. Expected error for a forecast provided 30 to 40 minutes ahead of
the next hour should be closer to 4%. These error ranges will vary depending on the specific
wind regime and project layout.
During 2003 and 2004, EPRI published several reports on wind energy forecasting system
development and testing at two wind plants in California and one in Texas [2731]. These
reports assessed the capability of current forecasting technology to predict hourly wind speed
and energy generation over the next 48 hours. The mean absolute errors of the wind speed and
wind energy generation were 30% to 40% and 50% to 60% of the annual average wind speed
and energy generation, respectively.
During 2005, the California Energy Commission and EPRI completed development and testing
of a regional wind energy forecasting system, designed to generate five-minute forecasts over a
three-hour period, updated every five minutes. In addition, the project evaluated improved
technology for intermediate-term forecasting of hourly wind generation over a 48-hour period.
The results are described in a four-volume final report published in 2006 [3235].
To date, most wind projects simply deliver energy into a utilitys system on an unscheduled basis
without the use of forecasting. However, wind production forecasting is increasingly being used
for power scheduling purposes. For example, wind forecasting is now being used by the
California ISO to forecast approximately 500 MW of wind generation across the state.
Forecasting is also being used extensively in the Pacific Northwest to schedule the output of
individual projects, including the 300-MW Stateline project.
For utilities that are required to schedule generation with a control area operator, electronically
delivered next-hour wind production forecasts can be automatically integrated into next-hour
schedule modifications to minimize imbalances. As noted earlier, forecast error in these nexthour time frames is expected to be small and may be smaller than a utilitys average control area
load forecast error. Because these two errors are not correlated, the relatively small wind forecast
error should make little net contribution to a utilitys existing level of scheduling imbalances.
3-135

Wind Power

3.10.4 Power Quality


In recent years, wind turbine manufacturers have placed increasing emphasis on improving the grid
integration and power quality of wind turbine generators. This push is being driven largely by two
factors. First, as wind economics have improved and wind power has reached higher penetration
levels, expectations have increased for state-of-the-art performance and better
power quality. Second, many growth opportunities for wind around the world are on weaker
transmission systems where power quality is key for expansion. Slowing growth in the European
onshore wind market is focusing European manufacturers into the large North American market.
Unlike the European power grid, which is characterized by greater density of population, power
generation, and transmission lines, the U.S. market, particularly in the West, is characterized by
expansive geography and long distances between population centers and generation. Much of the
wind resource is located in remote regions with access to only smaller, less robust transmission lines.
General Electric, the only major U.S. turbine manufacturer, has led the push for power quality
in the U.S. market with its patented dynamic volt-ampere-reactive (VAR) system. This system
regulates voltage in real time, supplying reactive power to the grid through each turbines power
electronics instead of through slower-acting capacitors or static VAR compensators. More
recently, European manufacturers have developed similar systems to compete effectively in
North America.
In 2004, Danish manufacturer Vestas developed the V90, a 3.0-MW full variable-speed pitchregulated turbine that provides similar VAR support and ride-through benefits as the GE turbine.
In 2005, the company received an order for 36 units for an offshore project near the Dutch coast.
In addition to OEM-provided power quality systems, several after-market dynamic VAR support
systems are available, specifically designed for use with wind generation.

3.11 Project Development Process and Market


3.11.1 Buy or Build Considerations
As wind energy costs continue to drop and utility experience with purchasing wind energy
increases, more utilities are likely to consider developing, owning, and operating their own wind
projects. Two important initial considerations are project scale and capital intensity. Wind
project costs decline rapidly with project scale. In Class 5 and 6 wind regimes, projects of 100
MW and larger can typically deliver energy at costs below competing new thermal generation.
However, at smaller scales, this advantage can quickly disappear. Smaller project costs are
driven up by pre-development, financial transaction, and construction mobilization expenses that
typically remain relatively flat regardless of project size. Also, wind projects are capital
intensive, typically coming in near $1,150/kW before 2005 (now $1,700 to $2,200/kW) installed
for a 100-MW project. For utilities that must carefully allocate capital across many competing
uses, entering a power purchase agreement with a wind developer may provide more strategic
benefit than committing to wind plant ownership.
For utilities that are interested in investing in a wind project, the question of balancing
cost management and risk must be addressed at each project development stage. The predevelopment stage is high risk. Pre-development companies typically invest up to a half million
dollars for site pre-screening, resource assessment, land and easement acquisition, environmental
review, and permitting on a specific site before viability can be determined. Generally, about one
3-136

Wind Power

in four sites clears all required hurdles and demonstrates resource quality sufficient to produce
energy in the target cost range. The high levels of in-house expertise maintained by experienced
pre-development companies and consultants helps to reduce pre-development risk and improve
the probability of identifying and successfully developing sites with the optimal resource in any
given area. Utilities choosing to take on the pre-development task can avoid the significant fees
charged by pre-development specialists for packaged sites, but may expose themselves to a
high probability of project failure or disappointing results. Utilities may also hire experienced
consultants to guide the pre-development process and manage sub-tasks in disciplines unfamiliar
to utility personnel, thereby reducing risk.
3.11.2 Project Development Process
The vast majority of wind capacity on-line in the United States has been developed, financed,
built, and operated by wind independent power producer companies, with the energy being
sold under long-term power purchase agreements with utilities. This arrangement stems from
the complexity of wind project development and project operation. Specialized activities require
expertise across a range of disciplines, a number of which do not fall within the areas of
expertise of typical electric utilities. In addition, equipment procurement and operations are
subject to substantial economies of scale. A small but growing portion of U.S. wind capacity is
owned and operated by utilities. Much of the utility-owned facilities have traditionally comprised
small wind installations with fewer than five turbines and been operated by municipal and public
utilities. However, large-scale, multimegawatt utility wind ownership is on the upswing.
MidAmerican Energy developed a 310-MW project in its Iowa territory, intends to embark on
future projects totaling 545 MW, and is proposing to add another 1,000 MW of owned wind
generation, thereafter. Other utilitiesincluding Alliant Energy, Kansas City Power & Light,
Minnesota Power, Nebraska Public Power District Oklahoma Gas & Electric, Puget Sound
Energy, Sacramento Municipal Utility District, and WE Energies, among othershave either
already invested in or have announced plans to own and develop wind projects.
3.11.2.1 Pre-Development Activities
The project development process falls into three distinct stages: (1) pre-development;
(2) engineering, procurement, and construction (EPC); and (3) operations. Primary predevelopment tasks include resource assessment, transmission access evaluation, land acquisition,
permitting, pre-engineering/costing, marketing, and, sometimes, finance.
Wind resource quality can vary significantly from site to site. Obviously some locations are
windier than others, but even within a known wind resource area, the wind resource can vary
substantially between adjoining land parcels, depending on topography. This is further
complicated by the fact that, for a given site, wind resources generally exhibit seasonal, diurnal,
and hourly variations. Accordingly, optimizing site selection for resource quality is perhaps the
most important and difficult pre-development task. Wind resource quality is characterized by
wind speed and directiontypically expressed as frequency distributions, the wind shear or
variation of wind speed with elevation, and the turbulence intensity.
Prior to final site selection, the wind resource is measured for an extended period of timeat
least one year, and usually two to three yearsto statistically quantify the resource. If the data
collected show a high correlation to a nearby meteorological tower with a longer operating
historyat an airport, for examplea shorter period of data collection may be acceptable.
3-137

Wind Power

A meteorological tower or mast is erected at one or more locations to continuously measure wind
speed, direction, temperature, and sometimes other weather parameters. The measurements are
made at multiple elevations above the ground (typically 10, 30, and 50 meters) to allow the wind
shear to be estimated. As a rule of thumb, meteorological tower height should be at least
two-thirds the planned turbine hub height.
The recent trend toward turbine towers 80 m and taller will likely engender broader use of
new 60-m meteorological towers and SODAR, sonic detection, and ranging technology that uses
sound to passively profile the vertical shear from the ground up to several hundred meters. The
resulting data are stored on site by a data logger and periodically downloaded on site or remotely
by modem. Data are analyzed to resolve erroneous values and calculate average wind speeds,
directions, and temperatures over annual, seasonal, monthly, and hourly time intervals. The
information is often expressed in wind speed frequency distributions and wind roses, which
graphically show the relative frequency of wind speed and direction, and wind energy.
Most regions of the United States have specific locations where significant wind resources exist,
especially on mountain passes, ridges, and coastal areas near oceans and large lakes. Figure 3-66
shows the distribution of wind power in the United States.

3-138

Wind Power

Figure 3-66
Distribution of average wind power in the United States
Source: Renewable Resource Data Center

3-139

Wind Power

Wind speed class characterizes the wind resource, with the darkest areas representing high-speed
Class 7 winds and the lightest representing low-speed Class 1 winds. In the last few years, more
detailed wind maps have been generated for most of the country that show the wind resource
distribution in finer detail than was available in previous maps. A listing and web links to these
updated wind maps is available from the DOE at
www.eere.energy.gov/windandhydro/windpoweringamerica/.
For cost-effective wind power generation, a wind power plant must be located at a site with high
wind speeds and must use reliable, efficient wind turbines. Generally, Class 3 is representative of
a low-to-moderate wind resource and Classes 4 through 6 are considered to be the most desirable
for commercial wind projects.
Site selection is typically an iterative screening process in which the pros and cons for a large
number of potential sites are evaluated with increasing levels of scrutiny until the highest
viability site or sites are selected for further pursuit. Pre-engineering, costing, and output
projection also include many variables and components that are unique to wind generation
and require specialized knowledge.
In todays market, the wind project pre-development stage is often conducted by small,
regionally focused, independent wind pre-development companies that actively seek out sites
and, in turn, sell a completed pre-development package to large energy companies that selffinance, build, own, and operate multiple wind projects. A completed pre-development package
typically includes: fully-documented wind resource studies, land leases, land use entitlements
(permits), an interconnection agreement, and a power purchase agreement. The build, own,
operate (BOO) companies typically also maintain in-house pre-development groups that actively
seek and pre-develop sites and market output to utilities or other buyers.
3.11.2.2 Engineering, Procurement, and Construction Activities
At the EPC stage, there is also a range of cost and risk considerations that should be evaluated.
Wind generation technology and associated power management equipment is fully commercial
but still evolving rapidly. Advances are made annually in improving project engineering, layout,
equipment, and material choices. Accordingly, to achieve optimal project performance, engineers
with substantial and current wind generation experience should be utilized.
During the 1980s and 1990s, most wind projects were developed by independent predevelopment companies that also managed EPC tasks and project finance, tapping financial
markets for debt and equity. However, over the past several years, large utility or oil and gas
company affiliates have consolidated dominant wind development market share using a balance
sheet finance approach. The independent project finance model may re-emerge in the later half
of this decade as the broader financial markets recover their confidence in electricity generation
investments.
At the procurement level, economies of scale in purchasing can be important. Large wind buildown-operate companies often maintain multiyear bulk purchasing agreements with major
equipment vendors, which can allow them to gain the pricing benefit of very large purchase
orders and spread them across numerous projects, both large and small. Utilities and developers
purchasing equipment and services for single projects do not have this advantage. Nevertheless,
the wind turbine market in the United States is extremely competitive and has the benefit of a
3-140

Wind Power

diversity of high-quality manufacturers all competing to increase market share. Turbine


equipment, which typically accounts for approximately 70% to 75% of total project cost, can
often be purchased in smaller volumes at competitive prices, particularly if the purchaser can
allow for flexible lead times. Purchase order timing is important because the supply/demand
balance in the U.S. turbine market can be subject to discontinuities related to tax incentive
deadlines. Also, long lead-time orders allow turbine manufacturers to order major turbine
components for just-in-time delivery, saving inventory costs.
Regarding construction, wind projects are highly modularized, and the major tasks of pouring
foundations and erecting turbines can be accomplished very efficiently in a production line
manner as long as the work is properly staged, equipment deliveries are timely, and weather is
cooperative. Typically, crews can erect one turbine per day. Large projects may employ multiple
crews. This allows even very large wind projects in the 200-MW range to be completed within
six months from first groundbreaking to commercial startup. However, because high crane work
and regular crane movement are major components of the job, excessive wind, rain, snow, or
other force majeure events can cause substantial delays and cost overruns. Accordingly,
allocation of construction risk should be a key consideration when evaluating a buy or build
decision.
3.11.2.3 Operation and Maintenance Activities
Wind project operations and maintenance are following the trends in other generating
technologies and are becoming increasingly systematized. More sensors are being used to gather
statistical performance data on turbines and subcomponents, and around-the-clock remote
monitoring is becoming the norm. Wind turbine manufactures typically offer full-wrap warranty
service contracts for up to five years, and ongoing support and service relationships are available
from both turbine manufacturers and wind O&M companies. Generally, current wind turbine
technology is highly reliable, with the historical fleet availability for leading manufacturers in
excess of 97%. Well-established turbine models typically show availability in excess of 98%.
However, because wind technology is still evolving rapidly and new turbine models are being
introduced annually, technology risk can be a factor.
With advanced instrumentation and SCADA systems, modern wind power plants are capable of
operating unattended, and on-site labor is required only for routine scheduled maintenance and
periodic troubleshooting. Figure 3-67 includes two graphs: one of O&M personnel versus
number of turbines, and one of O&M personnel versus size of facility. Both graphs include data
from the Turbine Verification Program (TVP) projects as well as a number of other commercial
projects. Some of the commercial projects are based on proposals rather than actual experience;
however, because the data for these planned projects did not appear to distort the results, they are
included in the graphs.
The graphs show a strong relationship between the number of O&M personnel at a wind power
plant and the number of turbines (r2 = 0.92). Although there also appears to be a relationship
between the number of people and the size of the project (r2 = 0.70), the size of the maintenance
crew is more dependent on the number of turbines than on the rated capacity of the project. This
is due to the fact that, although some of the maintenance activities on larger wind turbines may
require more time or different equipment to complete repairs, many maintenance activities
require the same level of effort regardless of turbine size.
3-141

Wind Power

O&M Personnel versus Number of Turbines


14
y = 0.0533x + 0.6185

O&M Personnel

12

R 2 = 0.9196

10
8
6
4
2
0
0

25

50

75

100

125

150

175

Number of Turbines

O&M Personnel versus Project Rating


14
y = 0.0477x + 1.2231

O&M Personnel

12

R 2 = 0.7002

10
8
6
4
2
0
0

50

100
Project Rating (MW)

150

200

Figure 3-67
Number of O&M personnel at a wind power plant versus the number of turbines and
project rating

3.12 Environmental Issues


Environmental and social issues associated with wind energy include public acceptance,
permitting, land use, soil erosion, visual and noise impacts, and impacts on resident and
migratory bird and animal populations. These issues are addressed in the 1999 National Wind
Coordinating Committee report, Permitting of Wind Energy Facilities [18].
Experience in the DOE-EPRI Wind Turbine Verification Program indicates that, through
vigorous public outreach and involvement programs and a proactive approach to assessing
environmental impacts, it is possible to address community concerns and, ultimately, receive
support and foster a sense of ownership by the community.

3-142

Wind Power

3.12.1 Avian and Bat Issues


Avian interaction with wind facilities became a central issue for the wind industry in the late
1980s, when bird carcasses were first reported in the Altamont Pass wind resource area in
northern California. More recently, concerns over endangered bat species have been raised by
observations of increased bat mortality at the Mountaineer wind project in West Virginia
and other wind projects.
Because numerous birds are killed each year by man-made structures such as buildings,
communication towers, and bridges, not to mention cars, trucks, and airplanes, the idea that local
bird populations might be affected was not new to the wind industry. However, the presence of
raptor carcasses drew attention to the issue in the Altamont Pass, and the fact that the states
golden eagle population could potentially be affected raised the stakes.
In 1989 the California Energy Commission, along with the counties of Solano, Alameda, and
Contra Costa, contracted with BioSystems to conduct the first extensive study of bird fatalities
in the Altamont Pass. The results from that initial research were published in 1992. Although it
left many questions unanswered, the BioSystems study did indicate that wind turbines were
indeed killing birds, particularly raptors, and in numbers significant enough to cause concern. It
also marked the beginning of the debate over what has become the most contentious
environmental issue to face the wind industry.
Since 1992, much more research has been done at the Altamont Pass, as well as numerous other
wind resource areas, to identify the species found in each area and define their use and mortality
rates; understand why raptors seem particularly susceptible to collisions with wind turbines;
and identify the factors that lead to bird deaths (i.e., turbine designs, placement, geography,
vegetation, and prey availability at the site; habitat encroachment in surrounding areas; and
interaction behaviors such as flying, perching, and hunting). To date, this body of research
suggests that biological resource and risk factors vary significantly from site to site, and that
results from the Altamont Pass are not necessarily representative of other sites. In fact, all
research to date on sites in the United States indicates that significant bird mortality rates are
only correlated with one commercial wind energy facilityAltamont Pass. However, this does
not mean that the issue can be dismissed.
In 1999, the National Wind Coordinating Committee published a definitive report, Studying
Wind Energy/Bird Interactions A Guidance Document [17]. The report provides extensive
information on designing and conducting field avian studies to better understand the risk that
wind facilities create to local avian populations. If widely adopted and implemented, this report
will lead to a body of research that will produce credible and comparable results. The report
stresses that, while a general protocol for conducting research should be adopted, each site must
be evaluated on the basis of its unique set of parameters.
The current wind industry trend toward larger wind turbines is also considered to be beneficial
to reducing the avian mortality risk at wind facilities. Compared to the small 50- to 200-kW wind
turbines installed during the 1980s at Altamont Pass and other wind areas, the large wind
turbines being installed today typically use taller towers and operate at much lower rotor speeds,
typically 20 to 30 rpm. This may elevate the turbine rotor above the elevation of much of the
avian habitat, while slower rotor speed should reduce the probability of striking birds. In addition
to new wind projects, repowering projects in California are replacing large numbers of small
turbines with fewer large machines, which should reduce the avian risk at existing wind projects.
3-143

Wind Power

The industry must continue to take this issue seriously and strive to minimize avian and bat risk at
all wind facilities through ongoing research and proper siting. The National Wind Coordinating
Committee continues to take a lead role in coordinating and disseminating research on biological
impacts of wind farms and is an excellent resource for information on this subject [9].
3.12.2 Noise
Although wind generation facilities generate noise during both construction and operation, it is
the long-term operating noise that receives the most attention during permitting. In August 2002,
the National Wind Coordinating Committee issued an update to its report, Permitting of Wind
Energy Facilities, which addresses noise characteristics, impacts on receptors, prediction and
measurement, and mitigation strategies [18].
The noise produced by operating wind generation facilities is much different in both level and
character from the noise generated by large power plants and other industrial facilities. It is
generally considered to be low-level noise and consists of both mechanical and aerodynamic
components.
Mechanical noise is typically only tonal noise at discrete frequencies, such as noise caused by
wind turbine mechanical components, vortex shedding from blades, and unstable flows over
holes or slits. Mechanical components that generate noise include gearboxes, yaw drives, brakes,
and cooling fans for generators. Mechanical noise can increase as components move out of
alignment and gears and bearings wear over time.
Aerodynamic noise is caused by airflow over and past the turbine blades and tends to increase
with rotor speed and wind speed. It consists of three components: broadband, impulsive, and
low-frequency noise. Broadband noise is characterized by a continuous distribution of sound
pressure with frequencies greater than 100 Hertz (Hz) and is often caused by the interaction of
turbine blades and atmospheric turbulence, producing the familiar swishing or whooshing
sound. Impulsive noise typically appears as a thumping noise as the wind turbine blades of
downwind turbines pass through the wind shadow of the tower. Low-frequency noise, in the
range 20 to 100 Hz, is also associated with older downwind turbines and results from interactions
between the blades and wake turbulence caused by towers and nearby wind turbines.
In general, the more the noise from a new source exceeds the background level or generates a
different tonal characteristic than background noise, the more it will be unacceptable to the local
community. The perception of wind generator noise by a receptor depends on the noise intensity,
frequency, frequency distribution, and pattern; background noise level; proximity of the receptor
to the wind turbine; terrain and vegetation between the wind generator and the receptor; and the
nature of the receptor. Background noise tends to increase with wind speed and thus mask the
wind generator noise. Therefore wind generator noise is generally most noticeable at lower wind
speeds. If the receptor is at a location where the wind speed is lower than at the wind turbine,
e.g., in a hollow or depression in the terrain, then the background noise will also be lower there.
In general, a wind turbine is likely to be heard at twice the distance in hilly terrain as in flat
terrain.
Although no federal and few state noise standards exist, the U.S. Environmental Protection
Agency has promulgated noise guidelines. Many local governments have enacted local noise
ordinances that must be considered when siting wind facilities.
3-144

Wind Power

The NWCC Permitting Handbook cites two proposed noise measurement techniques for wind
energy systems, one prepared by the Solar Energy Research Institute (now NREL) in 1987,
A Proposed Metric for Assessing the Potential of Community Annoyance; the other prepared
by the American Wind Energy Association (AWEA) in 1989, Procedure for Measurement of
Acoustic Emissions from Wind Turbine Generator Systems, Tier 12.1. In addition, the
International Electrotechnical Commission (IEC) proposed a standard in 1988 that was
rejected in the balloting process.
Turbine noise studies should include separate measurements of low-frequency and A-weighted
noise levels across a range of wind speeds, turbulence conditions, distances from the turbine,
and locations of the receptor relative to wind direction. Preferably before the wind project is
installed, background noise measurements should be conducted at representative dwellings
up to one-quarter mile from the site for flat terrain, and up to one-half mile for uneven terrain.
Receptors at less windy locations should be emphasized.
Several software packages are available that predict noise level profiles for the area around
a wind facility. However, prediction of wind generator noise levels at a given site is difficult
because conditions vary between sites. For example, the variations of wind turbulence and
noise with wind speed depend on wind direction, terrain, vegetation, and other site variables
and are rarely the same at two sites. Thus, the measured noise levels at one site are not
necessarily representative of the noise levels that would occur at another site. Noise mitigation
measures include requirements for noise setbacks of 400 to 1,000 meters from the edge of the
property line; installation of sound insulation and baffles in the turbine nacelle; use of low-speed
cooling fans, special finishing of gear teeth, vibration isolators and soft mounts for major
components; and design of wind turbine rotors for low noise. If local residences are shielded
from the wind, a larger setback may be required than in an exposed location.
3.12.3 Visual Impact
Wind projects are usually located in rural and remote areas, and their visual impacts are somewhat
different from those of conventional power plants. Visual impact considerations during permitting
typically focus on the impact of the wind project on the viewshed, or visual appearance of the
project setting from different locations. They include the impacts on the natural terrain and
landscape; the form, line, color, and texture of the viewshed; and the visibility and appearance of
landmarks in the area. In addition, the compliance of the project with public preferences, local
goals, policies, and guidelines regarding visual quality is crucial toward gaining acceptance.
The elements of wind projects that affect visual impact include the relative elevations of the site and
the surrounding area; the presence of dense woodland or other vegetation cover at the site; the
number of wind turbines; wind turbine spacing and uniformity; height and color of the wind turbine
nacelle, tower, and rotor; the type of tower used (lattice or tubular steel); markings and lighting
required for FAA approval; roads built on slopes; and the locations and sizes of service buildings,
substations, electrical switchgear, transmission lines, and other on-site facilities. In general,
deploying fewer and wider-spaced turbines of uniform type, color, tower design, and rotational
direction enhances a projects visual appearance and makes it more likely to gain public acceptance.
In a March 2003 study, the Renewable Energy Policy Project (REPP) found that property values
within the viewshed of recently constructed commercial-scale wind farms have not been
negatively affected [10]. The study assembled a database covering every wind development that
3-145

Wind Power

had come online in the United States after 1998 with 10-MW installed capacity or greater. REPP
then compared actual property sales records for all land within the viewshed and for a
comparable community not within the viewshed. The report found that for the great majority of
projects, property values actually rose more quickly in the viewshed than they did in the
comparable community. Moreover, values increased faster in the viewshed after the projects
came on-line than they did before. In all, 30 cases were analyzed; in 26 of them, property values
in the affected viewshed increased more than those in other areas without wind turbines.
3.12.4 Shadow Flicker
Shadow flicker is a result of the alternating changes in light intensity that can occur at times
when the rotating blades of wind turbines cast moving shadows on the ground or on structures
(see Figure 3-68). As the turbine blades pass in front of the sun, a shadow moves across the
surrounding area, causing a flicker effect as the turbine rotates. Because of the necessary location
of the sun in the sky, shadow flicker is limited in time and location. The location of the turbine
shadow varies by time of day and season.

Figure 3-68
Shadow flicker impact illustration (AWEA)

Flickering occurs in a butterfly-like pattern around each turbine. AWEAs standard of 1,000 m
from a wind turbine is often cited as the distance necessary to prevent significant impact from
shadow flicker. Based on observed modeling and actual project analysis, 1,400 m is the distance
to have very little impact throughout the year. Figure 3-69 illustrates the important role of
distance in reducing the impact of shadow flicker.

3-146

Wind Power

Figure 3-69
The importance of distance in shadow flicker prevention (CH2MHill)

These shadows can be disruptive and has caused concerned in Northern Europe and certain parts
of the United States. Many opponents of new wind farms cite the concern that the shadow flicker
causes headaches, nausea, dizziness, and disorientation and can trigger seizures in people who
suffer from epileptic seizure. However, typical epileptic seizures are precipitated by light flashes
in the range from 5 to 30 Hz. Shadow flickering caused by todays wind turbines rotating at a
much slower rate is in the range of 0.6 to 1.0 Hertz. Wind power advocates argue that empirical
evidence does not support the claims of negative health impacts from wind farm installations.
Shadow flicker can be addressed in a variety of ways, including instituting a significant buffer
zone from the wind farms, landscaping to block the shadows, or stopping the turbines during the
most vulnerable times. During the project design, developers are often requested or required to
analyze shadow clicker to determine where the shadows would fall and for how long over the
course of a year.
3.12.5 Radar Interference
Wind turbines, like other tall, metallic structures, can interfere with communication or radar
signals when these signals are interrupted by the turbine structure or the rotor plane. Wind
turbines can sometimes cause electromagnetic interference affecting TV and radio reception.
Electromagnetic interference can be caused by near-field effects, diffraction, or reflection and
scattering. Although instances of TV or radio interference are infrequent and typically
straightforward to mitigate, the interaction of wind turbines and navigational or defense radar
signals is the subject of considerable recent attention.
3-147

Wind Power

The increase of the number of wind turbines around military installations and airports has
increased risks to radar systems that track aircraft and other vehicles. The concern is that the
rotating blades will scatter waves or be mistaken for aircraft or other moving vehicles on radar
screens, ultimately negatively impacting training exercises and operational readiness. In the
majority of cases, interference is either not present, is not deemed significant, or can be readily
mitigated. Understanding the extent of a wind installations radar interference potential and
developing mitigation techniques can be more complicated than for other forms of potential
interference, as it depends on turbine height, rotor sweep area, blade rotation speed, and the
landscape surrounding a wind energy project.
There are essentially two types of interferences: direct interference and Doppler interference.
Direct interference happens with high reflectivity and reduces radar sensitivity, sometimes
producing false images (ghosting) or shadow areas (dead zones). Doppler interference
creates false targets and affects both airborne and fixed radar.
The following tools and practices are being studied to manage or mitigate the potential impact of
wind turbine interference:

Conducting studies and doing extensive modeling in the early stages of project development
to ensure that the wind farm location is not in an area of high radar activity. Wind farm
layout optimization, terrain masking/shielding, or reduction of the radar cross-section area
may be sufficient to address identified interference problems.

Using an absorbent or reflective material as a coating for equipment to minimize the


turbines radar signature.

Installing additional radars to cover the shadow spots, relocating radar installations to
accommodate the new wind farms, or altering air traffic routes around new wind farms.

Using long-range radars that are wind far friendly (e.g., Lockheed Martins AN/TPS-77).

Developing signal processing algorithms that reduce and eliminate interferences.

3.13 Scouting New Potential Game-Changer Technologies


As part of EPRIs Technology Innovation there is a Wind Scouting Program to identify
innovative concepts that may be game changers and contribute to address key challenges and
reduce cost of the production of electricity based on wind.
The following sections contain a summary description of selected set of breakthrough concepts,
not yet fully commercial, with potential to become wind industry game changers.
3.13.1 Multi-Copter Aerial Drones for Cost-Effective Wind Turbines, Blades, and
Tower Inspection
A multi-copter drone, with high definition and thermal cameras, can be used for blades and
tower in-service inspection. The images are sent life to the engineer on the ground with the
down-link system. The system is able to provide clear high resolution images. Advantages are
that the turbine is stopped less than 1 hour for inspection compared with several hours when
climbing to inspect. Video inspections are recorded and there no need to climb or descend the
turbines.
3-148

Wind Power

3.13.2 Direct-Drive, Axial Gap, Air-Cooled Permanent Magnet Generator


This technology can produce the same torque with less than half the mass (Nm/kg) of a
comparably rated iron core direct drive generator. Advantages include lower losses, higher
partial load efficiency, and modularity that contribute to low cost of maintenance, assembly, and
transportation without a large crane. Application to onshore and offshore large wind turbines will
allow substantial reduction of CAPEX and maintenance cost.
3.13.3 Oversize Large Turbine Component Heavy-Lift Transportation Using
Balloons
The Aeroscraft Rigid Variable Buoyancy Vehicle, designed to control lift in all stages of air or
ground operations, including to offload cargo without re-ballasting, may be used for economic
transportation of large-scale 10-Mw+ wind turbine blades, towers, and other heavy components
up to 500 tons. It can operate without ground infrastructure and minimize the cost of roads,
reduce environmental impact, and allow easier access to remote sites. The system allows vertical
takeoff and landing and can lower down slung maximum payload while in aerial hover mode.
3.13.4 Airborne Wind Turbines
An airborne wind turbine is a design concept for a wind turbine that is supported in the air
without a tower. Airborne wind turbines may operate in low or high altitudes.
Airborne turbine systems have the advantage of tapping of tapping an almost constant wind
without requirements for slip rings or yaw mechanisms and without the expense of tower
construction.
There are two varieties of airborne turbines: aerodynamic and aerostat variety. An aerodynamic
airborne wind power system relies on the wind for the support.An aerostat-type wind power
relies at least in part on buoyancy to support the wind-collecting elements.
Benefits common to both types include high operational performance and capacity factors,
removal of all placement limitations, fast deployment without towers or heavy cranes, reduced
environmental visual impact, and bats/bird-friendly applications to onshore/offshore/distributed
generation. Typical range of heights is 200 to 600 m, and there is a wide range of wind speeds.
This may contribute to reduce LCOE ~50% compared with conventional wind turbines.
3.13.5 Ro-Birds of Prey to Scare Birds
Mechanical birds of prey may be used for bird control, using the inbuilt natural reaction of birds
to the natural behavior of a bird of prey. The flapping of its wings is just as important as the
silhouette. There are different robot birds designed to scare small to medium-sized birds, and
others for large birds. The target is to reduce the environmental impact of wind turbines on birds
and reduce the requirement to curtail operations.

3.14 References
1. American Wind Energy Association (AWEA) news release, August 20, 2003, available at
http://www.awea.org/news/news030820nmw.html.
2. Wind Turbine Verification Project Experience: 1999. EPRI, Palo Alto, CA: 2000. 1000961.
3-149

Wind Power

3. Long Islands Offshore Wind Energy Development Potential: A Preliminary Assessment,


Long Island Power Authority and the New York State Energy Research and Development
Authority, April 2002.
4. Parsons, B., M. Milligan, B. Zavadil, D. Brooks, B. Kirby, K. Dragoon, and J. Caldwell,
Grid Impacts of Wind Power: A Summary of Recent Studies in the United States, June
2003, presented at the EWEA Conference in Madrid, Spain.
5. California Renewables Portfolio Standard Renewable Generation Integration Cost Analysis,
Phase I: One-Year Analysis of Existing Resources, Results and Recommendations, Final
Report, October 9, 2003. See
http://cwec.ucdavis.edu/rpsintegration/RPS_Int_Cost_PhaseI_pubrev1.pdf.
6. Milligan, M., Modeling Utility-Scale Wind Power Plants, Part 2: Capacity Credit.
NREL/TP-500-29701. Golden, CO: National Renewable Energy Laboratory, 2002.
7. MAPP Reliability Handbook, Section 3, Subsection 4.7.2.7, Procedures for setting Monthly
Net Capability for Variable Capacity Generation.
8. NEPOOL Manual for Installed Capacity, Manual M-20, Supplement D.
9. Updated research and information on current activities is available at NWCCs website at
http://www.nationalwind.org/.
10. The Effect of Wind Development on Local Property Values, REPP, Washington, DC,
May 2003. See www.repp.org/articles/static/1/binaries/wind_online_final.pdf.
11. Planning Your First Wind Power Project: A Primer for Utilities. EPRI, Palo Alto, CA: 1994.
TR-104398.
12. Building Community Support for Local Renewable and Green-Pricing Projects. EPRI,
Palo Alto, CA: 1999. TR-114203.
13. Renewable Energy Technology Characterizations. U.S. Department of Energy and
EPRI, Palo Alto, CA: 1997. TR-109496.
14. Power-Electronic, Variable-Speed Wind Turbine Development: 1988-1993. EPRI,
Palo Alto, CA: 1995. TR-104738.
15. Gipe, P., Wind Energy Comes of Age. John Wiley & Sons, Inc., New York, 1995,
pp. 216219.
16. Blanco, M.I., The Economics of Wind Energy, Renewable and Sustainable Energy
Reviews, Elsevier, 2009, Vol. 13, Issues 67, pp. 13721382.
17. Studying Wind Energy/Bird Interactions: A Guidance Document, National Wind
Coordinating Committee, Washington, D.C., December 1999.
18. Permitting of Wind Energy Facilities, National Wind Coordinating Committee, Washington,
D.C., August 1999.
19. Windicator, Wind Power Monthly, Knebel, Denmark: October 2005.
20. International Wind Energy Development, World Market Update 2004, BTM Consult ApS,
Ringkobing, Denmark, March 2005.

3-150

Wind Power

21. Stretching the BoundariesWind Energy Technology Review 20042005, Renewable


Energy World Magazine, February, 2005.
22. www.re-focus.net, March 2002.
23. Offshore Wind Could Quadruple UK Wind Energy, Renewable Energy World Magazine,
May-June 2001.
24. Wind Turbine Productivity Improvement and Procurement Guidelines. EPRI, Palo Alto,
CA: 2002. TR-104738.
25. Wind Power For Pennies, Technology Review, July/August 2002.
26. Characterizing the Impacts of Significant Wind Generation Facilities on Bulk Power System
Operations Planning, Xcel Energy-North Case Study, Final Report, Utility Wind Interest
Group, 2003: 1004807.
27. California Wind Energy Forecasting System Development and Testing Phase 1: 12-Month
Testing. EPRI, Palo Alto, CA, California Energy Commission, Sacramento, CA: 2003.
1007338.
28. California Wind Energy Forecasting System Development and Testing Phase 2: 12-Month
Testing. EPRI, Palo Alto, CA, California Energy Commission, Sacramento, CA: 2003.
1007339.
29. Texas Wind Energy Forecasting System Development and Testing, Phase 1: Initial Testing.
EPRI, Palo Alto, CA, U.S. Department of Energy, Washington, DC: 2003. 1008032.
30. Wind Energy Forecasting Applications in Texas and California: EPRI-California Energy
Commission-U.S. Department of Energy Wind Energy Forecasting Program: EPRICalifornia Energy Commission-U.S. Department of Energy Wind Energy Forecasting
Program. EPRI, Palo Alto, CA: 2004. 1004038.
31. Texas Wind Energy Forecasting System Development and Testing, Phase 1: 12-Month
Testing. EPRI, Palo Alto, CA, U.S. Department of Energy, Washington, DC: 2004. 1008033.
32. Wind Energy Forecasting Technology Update: 2004, Palo Alto, CA: 2005. 1008389.
33. California Regional Wind Energy Forecasting System Development, Volume 1: Executive
Summary. EPRI, Palo Alto, CA, California Energy Commission, Sacramento, CA: 2006.
1013262.
34. California Regional Wind Energy Forecasting System Development, Volume 2: Wind Energy
Forecasting System Development and Testing and Numerical Modeling of Wind Flow over
Complex Terrain. EPRI, Palo Alto, CA, California Energy Commission, Sacramento, CA:
2006. 1013263.
35. California Regional Wind Energy Forecasting System Development, Volume 3: Wind Tunnel
Modeling of Wind Flow over Complex Terrain. EPRI, Palo Alto, CA, California Energy
Commission, Sacramento, CA: 2006. 1013264.
36. California Regional Wind Energy Forecasting System Development, Volume 4: California
Wind Generation Research Dataset (CARD). EPRI, Palo Alto, CA, California Energy
Commission, Sacramento, CA: 2006. 1013265.
3-151

Wind Power

37. Engineering and Economic Evaluation of Utility-Scale Wind Power Plants. EPRI, Palo Alto,
CA: 2009. 1017599.
3.14.1 DOE-EPRI Wind Turbine Verification Program Reports
38. Tennessee Valley Authority Buffalo Mountain Wind Power Project First-and Second-Year
Operating Experience: 2001-2003. EPRI, Palo Alto, CA: 2003. 1008031.
39. Tennessee Valley Authority Buffalo Mountain Wind Power Project Development. EPRI,
Palo Alto, CA: 2003. 1004207.
40. Big Spring Wind Power Project Third-Through Fifth-Year Operating Experience:
2001-2004. EPRI, Palo Alto, CA: 2004. 1008384.
41. Big Spring Wind Power Plant Second-Year Operating Experience: 2000-2001. EPRI,
Palo Alto, CA: 2001. 1004042.
42. Big Spring Wind Power Plant First-Year Operating Experience: 1999-2000. EPRI,
Palo Alto, CA: 2000. 1000958.
43. Project Development Experience at the Big Spring Wind Power Plant. EPRI, Palo Alto,
CA: 1999. TR-113919.
44. Lessons Learned at the Iowa and Nebraska Public Power Wind Projects. EPRI, Palo Alto,
CA: 2000. 1000962.
45. Iowa/Nebraska Distributed Wind Generation Projects First-and Second-Year Operating
Experience: 1999-2001. EPRI, Palo Alto, CA: 2001. 1004039.
46. Project Development Experience at the Iowa and Nebraska Distributed Wind Generation
Projects. EPRI, Palo Alto, CA: 1999. TR-112835.
47. Kotzebue Electric Association Wind Power Project Third-Year Operating Experience:
2001-2002. EPRI, Palo Alto, CA: 2002. 1004206.
48. Kotzebue Electric Association Wind Power Project Second-Year Operating Experience:
2000-2001. EPRI, Palo Alto, CA: 2001. 1004040.
49. Kotzebue Wind Power Project Initial Operating Experience: 1998-2000. EPRI, Palo Alto,
CA: 2000. 1000957.
50. Project Development Experience at the Kotzebue Wind Power Project. EPRI, Palo Alto,
CA: 1999. TR-113918.
51. Wisconsin Low Wind Speed Turbine Project Third-Year Operating Experience: 2000-2001.
EPRI, Palo Alto, CA: 2001. 1004041.
52. Wisconsin Low Wind Speed Turbine First and Second Year Operating Experience:
1998-2000. EPRI, Palo Alto, CA: 2000. 1000959.
53. Wisconsin Low Wind Speed Turbine Project Development. EPRI, Palo Alto, CA: 1998.
TR-111438.
54. Green Mountain Power Wind Power Project Third Year Operating Experience: 1999-2000.
EPRI, Palo Alto, CA: 2000. 1000960.
3-152

Wind Power

55. Green Mountain Power Wind Power Project Second Year Operating Experience: 1998-1999.
EPRI, Palo Alto, CA: 1999. TR-113917.
56. Green Mountain Power Wind Power Project First Year Operating Experience: 1997-1998.
EPRI, Palo Alto, CA: 1998. TR-111437.
57. Green Mountain Power Wind Power Project Development. EPRI, Palo Alto, CA: 1997.
TR-109061.
58. Central & South West Wind Power Project Third Year Operating Experience: 1998-1999.
EPRI, Palo Alto, CA: 1999. TR-113916.
59. Central & South West Wind Power Project Second Year Operating Experience: 1997-1998.
EPRI, Palo Alto, CA: 1998. TR-111436.
60. Central & South West Wind Power Project First Year Operating Experience: 1996-1997.
EPRI, Palo Alto, CA: 1997. TR-109062.
61. Central and South West Wind Power Project Development. EPRI, Palo Alto, CA: 1996.
TR-107300.
62. DOE-EPRI Wind Turbine Verification Program TVP MI-112231 Status Report, 1998.
63. DOE-EPRI Wind Turbine Verification Program TVP MI-112231 Status Report, 2001.
64. Review of Operation and Maintenance Experience in the TVP Program, Report No.
LP-500-28620, 2000.
3.14.2 Other Reports
65. Musial, W. and S. Butterfield, Future for Offshore Wind Energy in the United States,
National Renewable Energy Laboratory NREL/CP-500-36313, presented at EnergyOcean
2004, Palm Beach Florida, June 2830, 2004.
66. Frandsen, S., Design of Offshore Wind Turbines, Professional Course, Developing
Offshore Wind Energy, European Renewable Energy Centers Agency (EUREC), June 2003.
67. Engineering and Economic Evaluation of Renewable Energy Technology. EPRI, Palo Alto,
CA: 2007. 1012726.
68. Advanced Wind Turbine Technology Assessment2010. EPRI, Palo Alto, CA: 2010.
1019772.
69. Wind Power Technology Status and Performance and Cost Estimates2009. EPRI, Palo
Alto, CA: 2010. 1020362.
70. Wind Turbine Blade Structural Health Monitoring: Methods and Benefits. EPRI, Palo Alto,
CA: 2010. 1021655.
71. Qu, R., Workshop on Next Generation Wind Power, Rensselaer Polytechnic Institute. May
2010. Development and Challenges of Permanent Magnet Wind Generators. Published
May 12, 2010. Retrieved September 2010 via http://www.rpi.edu/cfes/news-andevents/Wind%20Workshop/Development%20Challenges%20of%20PM_Generator_RPI_Qu
_v8.pdf.

3-153

Wind Power

72. de Vries, E., Permanent Solution? Renewable Energy World, April 8, 2010. Retrieved
September 2010 via
http://www.renewableenergyworld.com/rea/news/article/2010/04/permanent-solution.
73. Tan, A., A Direct Drive to Sustainable Wind Energy. Windsystems Magazine, March 2010.
Retrieved September, 2010 via
http://windsystemsmag.com/media/pdfs/Articles/2010_March/Zenergy_0310.pdf.
74. Abrahamsen, A.B., N. Mijatovic, E. Seiler, et al., Superconducting Wind Turbine
Generators. Superconducting Science and Technology, Vol. 23, 2010. Retrieved September,
2010 via http://dx.doi.org/doi:10.1088/0953-2048/23/3/034019.
75. Tyler, D., Clean Technica. American Superconductor, DOE Study Large Wind Turbine
Design. Published February 11, 2009. Retrieved August 2010 via
http://cleantechnica.com/2009/02/11/american-superconductor-doe-to-study-large-windturbine-design/.
76. Chow, R., and C.P. van Dam., A Focus on the Flow in the Inboard Part of the Blade.
Presented at the Sandia National Laboratories Blade Workshop, July 2022, 2010. Retrieved
August 2010 via http://windpower.sandia.gov/2010BladeWorkshop/PDFs/2-2-A2%20VanDam.pdf.
77. Fairley, P., Stealth-Mode Wind Turbines. Technology Review, November 2, 2009.
Retrieved August 2010 via http://www.technologyreview.com/energy/23837/.
78. Renewable Energy Focus. Updated: QinetiQ and Vestas Test Stealth Technology for Wind
Turbines. Published October 26, 2009. Retrieved September 9, 2010 via
http://www.renewableenergyfocus.com/view/4715/updated-qinetiq-and-vestas-test-stealthtechnology-for-wind-turbines/.
79. Karal, P.and R. Stancich, Whats New about Gravity Base Foundations for Offshore
Wind? Wind Energy Update, September 3, 2010. Retrived September 7, 2010 via
http://social.windenergyupdate.com/qa/what%E2%80%99s-new-about-gravity-basefoundations-offshore-wind.
80. Jonkman, J. and D. Matha. National Renewable Energy Laboratory, A Quantitative
Comparison of the Responses of Three Floating Platforms. Report NREL/CP-500-46726,
March 2010. Retrieved August 2010 via
http://permanent.access.gpo.gov/lps123972/46726.pdf.
81. Lombardi, C., Norway Oil Giant Floats Idea for Bobbing Windmills. CNET News, Green
Tech, August 19, 2010. Retrieved September 8, 2010 via http://news.cnet.com/830111128_3-20014140-54.html.

3-154

BIOMASS ELECTRICITY GENERATION

4.1 Introduction
Biomass fuels produced by living plant and animal matter provide electric utilities and other
electricity generators with dispatchable renewable power (see Table 4-1 for an overview). For
the most part, biomass fuels can be produced, concentrated, and stored for use when it is
economical to do so. The U.S. Energy Information Agency (EIA) estimates that biomass
supplied 4.3 1015 Btu (Quads) of energy to the U.S. economy in 2010, a 30% increase above
2007 levels. Altogether, biomass supplies some 21 EJ to the worlds economylargely in
North America, Scandinavia, China, and numerous developing economies. Over the past
decade, annual biomass consumption in the United States has ranged from 2.77 EJ to 3.8 EJ
(2.627 to 3.615 Quads), depending on economic activity among the wood products industries,
agribusinesses, electric utilities, and other industries. Biomass consumption in the U.S. economy
has been generally quite stable, though it did climb more than 7% in 2007 owing to increased
demand for biofuels in the transportation sector.
Table 4-1
Biomass electricity generation overview
Installed Capacity (est.)
(Jan. 1, 2011 est.)

62,000 MW worldwide (2010)

10,400 MW in United States (2010)

World leaders:

- Brazil: 7,800 MW
- Germany: 4,900 MW
- China: 3,700MW
Technology Readiness

Environmental Impact

Combustion systems are commercial, widespread; gasification/pyrolysis


in initial deployments.

High confidence in cost estimates/projections for combustion systems.

Moderate to low confidence in cost estimates and projections for


gasification and pyrolysis systems.

Atmospheric and solid emissions from combustion and gasification


technologies controllable using commercial technology.

CO2 net emissions can either decrease or increase, depending on the


feedstock and its alternate fate.For the most part, there is a very slight
net increase in emissions.

Can help meet local solid waste disposal challenges.

4-1

Biomass Electricity Generation


Table 4-1 (continued)
Biomass electricity generation overview
Economic Status

Policy Status

Trends to Watch

Competitive or near-competitive in many markets, particularly if CO2


fossil emissions costs are externalized.

Particularly attractive in cogeneration applications.

Biomass co-firing is a low-cost option that may be a good option for


older coal units that are being considered for retirement.

Cost of securing and transporting fuel remains a challenge on case-bycase basis.

Not consistently addressed by state RPS or energy policies


(except digester gas and ethanol production). Highly inconsistent
treatment of carbon accounting.

Cap-and-trade/tax legislation for greenhouse gases (GHG) uncertain.

Combustion technologies can encounter resistance from the public,


policymakers, and environmental groups; education is needed.

Policies and government R&D investment tend to favor biomass to


liquid fuels over other biomass energy.

European Union policies and regulations have produced a robust and


growing biomass-to-power industry.

Development of cellulosic ethanol and related impact on fuel availability


and the export of biomass and biomass products abroad

Repowering older coal units for biomass firing.

Progress toward more efficient and economical gasification.

Dedicated high-yield fuel crops for biofuels and power.

Resolution on carbon neutrality of biopower

Biomass pre-treatments such as cleaning, torrefaction, steam treating,


and others.

Biomass certification standards.

Biomass fuels are largely industrial fuels. Table 4-2 shows the distribution of biomass fuels
among the economic sectors, based on EIA data. These data do not include the more than 0.66 EJ
of biomass used in transportation in the form of ethanol added to gasoline and biodiesel added to
diesel fuel.
Table 4-2
Biomass energy consumption in the U.S. economy by sector and type, 2010 (Quads)
Biomass Type

Residential and Commercial

Industrial

Electricity

Total

Wood and Wood Waste

0.490

1.307

0.189

1.986

MSW/Landfill Gas

0.029

0.122

0.227

0.378

Other Biomass

0.005

0.046

0.025

0.076

Total

0.524

1.475

0.441

2.44

Source: EIA

4-2

Economic Sector

Biomass Electricity Generation

In the electricity sector, use of biomass for power generation is expected to double every 10
years through 2030. Biopower uses bioenergy systems to generate electricity through the
following methods:

Direct firing in dedicated 100% biomass fueled boilers

Cofiring where biomass substitutes a portion of coal (or oil/gas) in existing power plants

Repowering existing coal boiler to 100% biomass

Gasification, which converts biofuels or biomass into carbon monoxide and hydrogen by
reacting the raw materials at high temperatures with a controlled amount of oxygen and/or
steam

Pyrolysis, which converts biomass into easily stored and transported liquid by subjecting the
feedstock to high temperatures in the absence of oxygen

Combined heat and power facilities, which generate steam/hot water for process or district
heating and electricity

Anaerobic digestion processes, which capture methane from decomposing bio-matter

Figure 4-1 illustrates the percentage share of renewable electricity (in billions kWh) that biomass
is expected to contribute through 2035.

Figure 4-1
Renewable electricity projections, excluding hydropower (billion kWh/yr)
Source: EIA, Energy Outlook 2011

Historically, biomass consumption for energy use has remained at low levels. However, as
Figure 4-1 illustrates, biomass is expected to be one of the most important sources of U.S.
renewable electricity generation through 2035, trailing only hydropower, based on projected
kilowatt-hours delivered. In fact, biomass is predicted to grow to roughly 39% in 2035.

4-3

Biomass Electricity Generation

Furthermore, in scenarios that reflect the impact of a 20% federal renewable portfolio standard
(RPS) and in scenarios that assume carbon dioxide reduction requirements based on international
agreements, electricity generation from biomass is projected to increase more substantially. In
these scenarios, environmental benefit is the primary reason driving increased biomass
utilization.
Compared with coal, biomass feedstocks have lower levels of sulfur, sulfur compounds, and
mercury, and demonstrations have shown that biomass cofiring with coal can also lead to lower
combustion-based nitrogen oxide emissions. Perhaps the most significant environmental benefit
of biomass, however, is potential reduction in carbon dioxide emissions. If analyzed as a closedloop process harnessed to cultivate biomass, in which power is generated using feedstocks that
are grown specifically for the purpose of energy production, the CO2 emissions resulting from
biomass firing are virtually offset. Many varieties of energy crops are being considered
including hybrid willow, switchgrass, arundo donax, sorghum, and hybrid poplarand may
become commercially available in the United States in the short term.
Feedstock cost and feedstock availability are the main barriers to widespread use of biomass for
power generation. In the long term, bio-power potential will depend on technology advances as
well as on the level of competition for feed stocks with food and fiber production for arable land
use. Land-based competition may not be an issue until 2020, however, as enough land that has
been idled by farm programs appears to exist to satisfy biomass crop growing needs. Once
biomass crop acreage requirements exceed 30 million acres, which is likely by 2020, according
to Oak Ridge National Laboratory (ORNL) estimates, biomass crops will begin to compete with
traditional agricultural crops.
In addition, societal risks associated with widespread use of biomass include intensive farming,
overuse of fertilizers, chemicals use, and biodiversity conservation. They pose a challenge to
increased biomass usage. Certifications attesting the sustainable production of biomass feedstock
are needed to, for example, improve upon land management techniques. While overexploitation
of biomass resources in developed and developing countries should be avoided, sustainable
biomass can offer an important, productive use for marginal lands and bring socioeconomic
benefits to many rural regions of the world.
Separately, innovative densification pre-treatments, such as pelletization, torrefaction/upgrading,
and conversion into bio-oil (e.g. by pyrolysis) may help overcome the economically and
environmentally challenging logistics of long-distance transportation. The commercial
emergence of these pre-treatment methods could help facilitate a global international trade by
reducing logistic costs.
4.1.1 Basic Issues Associated with Biomass Fuel Utilization
Biomass fuels exhibit certain fundamental differences from other fossil fuels. Typically, biomass
fuels are either gathered or harvested from diffuse sources and concentrated at a given location.
Consequently, there are practical limits on the quantities that can be obtained at any location
without experiencing significant cost pressures. This phenomenon limited the capacities of the
early iron furnaces fueled by charcoal in Pennsylvania and other states in the late 18th century
and early 19th century [1], and remains an issue today, though at a significantly expanded scale.
This is in distinct contrast to fossil fuels such as coal that are produced in centralized locations
and distributed to users such as power plants.
4-4

Biomass Electricity Generation

Biomass fuels currently used as fuels today are, almost exclusively, residues from other
processes. They may be wood processing residues such as slash, hog fuel, bark, sawdust, or spent
pulping liquor. They may be agricultural and agribusiness residues such as bagasse. They may be
wastewater treatment gas or landfill gas. These are commodities that are presently outside the
commercial mainstream. In many cases, these commodities have both material and energy value.
Wood waste markets, for example, can include mulch for urban areas, bedding for livestock and
poultry, feedstocks for materials such as particleboard, and feedstocks for niche chemical and
related products. As a result, fuel pricing is highly sensitive to locale and the competitive
pressures of local and regional economies.
Significant efforts have also been made to develop short-rotation biomass crops to be used
exclusively as fuel or energy feedstocks. To date, these have not produced competitively priced
fuels, though development continues.
4.1.2 Technology Considerations for Using Biomass Fuels
Biomass-to-power technologies can be divided into two basic categories: (1) existing
technologies that are either commercialized or nearly ready for commercial implementation, and
(2) goal technologies that have been proposed and are of long-term interest. Most of this chapter
will focus on the first category, as its costs, technologies, and issues are better understood.
4.1.2.1 Commercially Available Technologies
Commercially available technologies include stand-alone condensing power plants, such as
those owned by electric utilities and independent power producers (IPPs). Stand-alone power
plants fired by wood waste and certain agricultural wastes were developed extensively during the
early years of the Public Utilities Regulatory Policies Act (PURPA). Stand-alone power plants
fired with municipal solid waste (MSW) and landfill gas have also been developed. For example,
landfill gas is used to fuel a 50-MWe steam boiler installation at the Puente Hills, Los Angeles
landfill. The energy installation at Puente Hills also includes a combustion turbine and internal
combustion engines.
Equally significantly, commercially available technologies include Rankine-cycle cogeneration
facilities such as those owned and operated by companies in the forest products industry, the
sugar industry, and other manufacturing interests. Such installations couple a medium-pressure
boiler with main steam conditions, typically 600 psig/750F, 850 psig/825F, 1250 psig/950F,
and 1450 psig/1005F to either backpressure turbines or automatic extraction turbines. Process
steam is typically generated at 50 psig, 150 psig, or 450 psig depending on the application. The
energy released when steam is expanded from throttle conditions to exhaust conditions is used
to generate electricity. In selected cases, such as the Snohomish County Public Utility District in
Washington, utilities team with forest products industry firms to jointly develop such
cogeneration capabilities.
Biomass cofiring technologies have historically been employed at numerous locations in the
United States and continously in Europe. For example, wood waste has been blended with coal
on the main belt leading to the Willow Island Generating Station #2 boiler, a 188-MWe cyclone
boiler owned by First Energy. Sawdust and other biofuels have been ground and separately
injected into the 105-MWe boiler at Greenidge Station in a project developed in the mid-1990s
by New York State Electric and Gas (NYSEG) and continued by AES, the current plant owner.
4-5

Biomass Electricity Generation

Southern Company has pioneered numerous cofiring approaches at Plant Hammond, Plant Yates,
Plant Gadsden, and other installations. Previously, Northern States Power implemented biomass
cofiring at its Allen S. King Generating Station. The New York Power Authority (NYPA)
cofired biodiesel with oil in its 890-MW Poletti station. Numerous vendors now offer
commercial systems, but because of unclear regulations and poor economics, biomass cofiring is
not commonly practiced in the United States.
An example of coupling atmospheric gasification of biomass to electricity-generating Rankinecycle installations is the Kymijarvi power plant in Lahti, Finland, owned and operated by Lahden
Lampovoima Oy. The Lahti gasification project, an EU Thermie demonstration, involved
numerous partners, including Lahden Lampovoima Oy, Foster Wheeler Energia Oy, VTT (the
Technical Research Center of Finland), Elkraft Power Co., Ltd. of Denmark, and Plibrico Ab
from Sweden. The project involved installing a Foster Wheeler atmospheric circulating fluidized
bed gasifier capable of converting sawdust, bark, wood chips, plywood trim, particleboard trim,
and recycled fuel (REF) into a producer gas for firing in the coal-fired boiler used for generating
electricity and district heat. Based on the success of that project and other gasification projects,
atmospheric circulating fluidized bed boilers can now be procured commercially for such
cofiring projects used to generate electricity.
For each of these commercially available technologies, the issues of fuel cost and the availability
of fuel in significant quantities remain driving issues. The process of gathering and concentrating
the fuel in a single location is a primary consideration. All these commercially available
technologies exhibit technical advantages and disadvantages associated with using biomass fuels.
Such technical considerations, along with economics, are discussed in subsequent sections of this
chapter.
4.1.2.2 Goal Technologies
The technologies that have long-term promise, but remain beyond current technical and
economic feasibility on a proven basis, are many and varied.
The practice of raising silvicultural or agricultural crops as a fuel supply is one such technology.
As mentioned previously, this technology was employed by the early iron industry in the
fledgling United States. It has appeared significantly in the literature since the early 1970s (see,
for example, Szego and Kemp [2], Bethel et al. [3] and Henry [4]). Government research in this
area has been significant over a long period of time. To date, however, experience has shown
that biomass cannot be grown and harvested at costs close to those associated with alternative
combustible resources. Further, there are significant economic issues associated with competition
for the resource, once grown, that make this a goal technology rather than a near-term
commercial reality. In the United Kingdom, however, government policies are in place that may
offer a significant role for dedicated biomass. The success of this program will be watched
carefully in the coming years.
Portland General Electric has identified arundo donax as a species that could be cultivated in
western Oregon and/or Washington exclusively for the purpose of power generation in a PC unit.
As of late 2011, a test plot has been developed, and an early crop harvested. Although arundo is
considered an invasive species, this might represent the first commercial dedicated energy crop
effort in the United States.
4-6

Biomass Electricity Generation

Pressurized gasification with hot or warm gas cleanup, coupled with combustion turbines in
integrated gasification combined-cycle (IGCC) applications, also remains a long-term
possibility. Scale limitations severely constrain the economics of this technology. Furthermore,
hot or warm gas cleanup remains an elusive target. A demonstration of low-pressure gasification
firing reciprocating engines is under way in Skive, Denmark. Pyrolysis to produce liquid fuels is
yet another goal technology, but with serious practical problems.
Atmospheric (or near-atmospheric) gasification with warm gas cleanup in a configuration that
provides fuel to reciprocating engines and district heating is also considered a goal technology,
though is much nearer to commercial readiness than pressurized biomass gasification used in
combined cycle mode. Demonstrations at Harbore, Denmark, featuring the Vlund gasifier,
provide evidence of approaching commercial deployment. This technology category is addressed
in more detail in Section 4.4.5.
4.1.2.3 Prospectus
Within this expanding framework, the generation of electricity from biomass can be evaluated.
In order to make such an evaluation, this chapter considers the physical and chemical
characteristics of various biomass fuels, and the technical characteristics and costs of various
biomass technologies.

4.2 Biomass Fuel Resources


Biomass is commonly defined as material derived from living organic matter (e.g., trees, grasses,
animal manure). Although there are alternate political/regulatory definitions of biomass, from the
physical and chemical perspectives (including the biochemical perspective) this definition
appears to be most accurate. The consequence is that fractions of municipal solid wastepaper,
wood waste, food waste, yard wasteare forms of biomass fuel. This definition also supports
the inclusion of both residues and potential energy crops as biomass fuel. Consequently, biomass
includes wood and wood waste, herbaceous crops and crop wastes, agribusiness wastes (e.g.,
food processing wastes such as bagasse), animal manures and methane-rich gas from anaerobic
digestion of such fecal mater, methane-rich gases generated from wastewater treatment plants
and landfills, and miscellaneous related materials.
Numerous national studies of biomass availability have been conducted (see, for example,
Walsh et al. [5]). Such studies estimate the availability of biomass as a function of price. Oak
Ridge National Laboratory estimates that over 5 EJ (or 5 Quads) of biomass as residue could
be available at a price of $50/dry ton (approximately $3/GJ or $3/106 Btu) nationally. This study
focused on forestry and wood residues, crop residues, and energy crops. It did not address animal
manures, landfill gas, wastewater treatment gas, or biomass components of municipal solid
waste. If those wastes are added in, the estimated quantity of residue-based biomass fuel
available at some price could well be on the order of 10 EJ (10 Quads).
A separate study conducted by the EIA [42] meanwhile estimates that the United States will have
an upper limit of 7.1 quadrillion Btu of biomass available at prices of $5 per million Btu or less,
by 2020. The available low-cost feedstock ($1 per million Btu) is almost exclusively composed
of urban wood waste and mill residues. At about $2 per million Btu, agricultural residues
become viable as a second source of biomass. And energy crops and forestry residues begin to
make significant contribution at prices around $2.40 per million Btu or higher.
4-7

Biomass Electricity Generation

As a practical matter, national estimates have policy implications but do not replace on-the-ground
field surveys for utilities and generating stations that require local availability estimates when
considering firing biomass singly or in combination with fossil fuels. Local estimates are based
upon forest industry or agricultural industry activity, plus local markets for forest industry
residues.
Residues include bark, sawdust, shavings, chips, hog fuel, logging residues, silvicultural residues
(e.g., pre-commercial thinnings), dead and diseased timber, and fire-damaged timber that cannot be
salvaged. Local residues may also include an array of agricultural materials such as the following:

Corn stover

Corn cobs

Out-of-date seed corn

Wheat straw

Rice straw

Rice hulls

Vineyard prunings

Orchard prunings

Oat hulls

Bagasse

Determination of locally available biomass depends upon local forest industry or agricultural
activity, local competing markets for the material, and transportation systems available. Local
conditions vary substantially. For example, in the Chicago area, bark commands a premium price
as mulch. In some areas with high concentrations of poultry farms, sawdust and rice hulls both
command good prices as bedding material. Chips command prices greater than fuel prices in
areas that are within economical haul distances of pulp mills or manufacturers of oriented strand
board (OSB), medium-density fiberboard (MDF), or related products. Other competing uses that
have been encountered include manufacturing charcoal, particleboard, and specialty products.
When considering agricultural residues, seasonality becomes a concern. For wastewater treatment
gas, the population served by the wastewater treatment plant is the primary factor and can include
industries as well as residential populations. Similarly, landfill gas utilization depends on landfills
serving large populations. As a consequence, California has more than 200 MWe of generation from
landfill gas [6]. In all cases, however, local resource availability is the governing factor.
Local resource availability is critical because biomass fuels have limited transportability. Green
wood wastes typically have bulk densities of 0.21 to 0.26 kg/m3 (16 to 20 lb/ft3). This compares
to coals with bulk densities of 0.68 to 0.78 kg/m3 (52 to 60 lb/ft3). Crop wastes can have bulk
densities of 0.11 to 0.13 kg/m3 (5 to 8 lb/ft3). Landfill and wastewater treatment gases are
typically saturated with water, and have heating values that are about half of that associated with
natural gas (on a dry gas basis). Further, landfill and wastewater treatment gases are not
compressed, while natural gas is transported under significant pressure. The typical estimated
maximum haul distances for woody biomass are on the order of 50 to 75 miles (80 to 120 km);
economical transport distances for other biomass fuels are significantly lower.
4-8

Biomass Electricity Generation

Energy crops are just beginning to be commercially grown in the United States, and RD&D is
under way to explore optimal energy crop species. The main potential energy crop species being
considered are hybrid poplar, hybrid willow, switchgrass, and as noted earlier, canes/reeds such
as arundo. POLYSYS, an agricultural policy simulation model of the U.S. agricultural sector,
assumes that energy crops production will be limited to areas that are climatically suited for their
production, thus excluding all states in the Rocky Mountain and Western Plains regions. Future
genetic improvements in energy crops may extend this range. POLYSYS excludes irrigation as a
viable management practice for energy crops production.
Crop, forest, and primary mills provide roughly 70% of the current biomass resource. The single
largest source of biomass is crop residue. Perennial crops have the largest growing potential for
energy production. Dedicated energy crops (switchgrass, willow, hybrid poplar) can often be
economically grown on land that is not suitable for conventional crops and can provide erosion
protection for agricultural set-aside or Conservation Reserve Program (CRP) lands. Figure 4-2
breaks out the various sources used for biomass generation.

Methane from
Domestic
Wastewater
0%

Switchgrass on
CRP Lands
20%
Crops
Residues
37%

Methane from
Landfills
3%

Methane from
Manure
Management
1%

Urban Wood
7%

Secondary Mill
1%

Forest Residues
13%
Primary Mill
18%

Figure 4-2
Breakdown of biomass feedstock

The key to producing low-cost plantation feedstocks is land availability and quality, which to a
significant extent determine the degree of site preparation necessary, the choice of species and
rotations (cutting cycles), required cultural management and soil amendments (fertilization, weed
control, etc.), and fuel transport and logistics. Land and site quality also define biomass
productivity, the major determinant of total feedstock cost. In addition, productivity is a key
factor in determining the total size of the plantation, annual feedstock production, and the size of
the conversion facility that can be supported. Figure 4-3 shows the effect of biomass plant scale
on feedstock requirements and transport distances.
4-9

Biomass Electricity Generation

Figure 4-3
Effect of biomass plant scale on feedstock requirements and plantation area required [43]

Greater plantation productivity directly translates into reduced land requirements. Higher
conversion efficiencies translates into more installed plant capacity for a given feedstock
requirement or less required plantation area for a given installed capacity.
4.2.1 Biofuels vs. Biomass Electricity
A 2009 study performed by researchers at Stanford University, the University of California
Merced, and others [44] concludes that, on average, using biomass to produce electricity is 80%
more efficient than transforming the biomass into biofuel. In addition, the electricity option was
deemed to be twice as effective at reducing greenhouse gas emissions. The analysis covered a
range of harvested crops, including corn and switchgrass, and a number of different energy
conversion technologies. Data collected were applied to electric and combustion-engine versions
of four vehicle types, and their operating efficiencies were estimated during city and highway
driving. The study accounted for the energy required to convert the biomass into ethanol and
electricity, as well as for the energy intensiveness of manufacturing and disposing of each
vehicle type. Bioelectricity far outperformed ethanol in most scenarios. (It some cases it was
found to be 80% more efficient.)
4.2.2 Basic Properties of Solid Biomass Fuels
Fuel properties can be characterized in terms of proximate and ultimate analyses, reactivity
measurements, ash elemental analyses, ash reactivity, and trace metal composition.

4-10

Biomass Electricity Generation

4.2.2.1 Proximate and Ultimate Analyses of Biomass Fuels


Research at The Energy Institute of Pennsylvania State University (PSU) has led to
characterization of a broad array of biomass fuels, including the proximate and ultimate analyses
shown in Tables 4-3 through 4-5.
Note that the sawdust in Table 4-3 is mixed hardwood-softwood sawdust. As a practical matter
the difference between softwood and hardwood is insignificant from a fuels perspective, unless
a fuel loaded with extractives is used, such as slash pine. The urban wood waste is clean and
does not include pentachlorophenol or copper chromium arsenate (CCA) treated wood. The
switchgrass shown in Table 4-4 is from the Plant Gadsden tests of Southern Company.
The data in the tables document the fact that all biomass fuels have similar heating values on a
moisture- and ash-free (MAF) basis; however, the various biomass fuels differ with regard to
their relative percentages of holocellulose (cellulose plus the hemicelluloses), lignin, and
extractives.
Table 4-3
Proximate and ultimate analyses for typical woody biomass fuels
Parameter

Moisture %

Fuel
Pine Chips

Pine
Shavings

Red Oak
Shavings

Fresh Mixed
Sawdust

Urban Wood
Waste*

45.0

45.0

28.8

40.0

30.8

Proximate Analysis (wt % dry basis)


Volatiles

84.7

84.7

79.5

80.0

76.0

Fixed carbon

15.2

15.2

19.0

19.0

18.1

Ash

0.1

0.1

1.5

1.0

5.9

Ultimate Analysis (wt % dry basis)


Carbon

49.1

49.1

51.6

49.2

48.0

Hydrogen

6.4

6.4

5.8

6.0

5.5

Nitrogen

0.2

0.2

0.5

0.4

1.4

Sulfur

0.2

0.2

0.0

<0.1

0.1

Oxygen

44.0

44.0

40.6

43.0

39.1

Ash

0.1

0.1

1.5

1.0

5.9

8502

8502

8069

8400

8364

HHV (Btu/lb)

Sources: Miller et al., 2000; Johnson et al., 2001; Tillman, 2001.


Note: *Urban wood waste shown is from the Bailly Generating Station Tests.

4-11

Biomass Electricity Generation


Table 4-4
Proximate and ultimate analyses for typical herbaceous biomass fuels
Parameter

Fuel
Fresh
Switchgrass

Weathered
Switchgrass

Reed Canary
Grass

Mulch Hay

15

15

65.2

19.5

Moisture %

Proximate Analysis (wt % dry basis)


Volatile matter

76.18

81.8

76.1

77.6

Fixed carbon

16.08

14.8

4.1

17.1

Ash

7.74

3.4

19.8

5.3

Ultimate Analysis (wt % dry basis)


Carbon

46.73

49.4

45.8

46.5

Hydrogen

5.88

5.9

6.1

5.7

Nitrogen

0.54

0.4

1.0

1.7

Sulfur

0.13

0.3

0.1

0.2

Oxygen

38.99

40.6

42.9

40.6

Ash

7.74

3.4

4.1

5.3

HHV (Btu/lb)

7750

8150

7103

8058

Sources: Miller et al., 2000; Johnson et al., 2001.

Table 4-5
Proximate and ultimate analyses for typical manures
Parameter

Moisture %

Fuel
Dairy TieStall Manure

Dairy FreeStall Manure

Poultry Litter

Sheep
Manure

Swine Waste

64.7

69.8

20.0

47.8

97.8*

Proximate Analysis (wt %, dry basis)


Volatiles

76.0

30.1

55.3

65.2

59.6

Fixed carbon

18.1

7.4

7.7

14.0

7.3

Ash

6.0

62.5

17.0

20.9

33.1

Ultimate Analysis (wt %, dry basis)


Carbon

48.6

22.6

38.1

40.6

38.0

Hydrogen

5.8

2.9

5.6

5.1

5.5

Nitrogen

1.4

1.1

3.5

2.1

3.2

Sulfur

0.1

0.1

0.6

0.6

0.6

Oxygen

38.1

10.8

30.9

30.7

19.6

Ash

6.0

62.5

17.0

20.9

33.0

8203

3644

6399

6895

7328

HHV (Btu/lb)

Source: Miller et al., 2000; Miller et al., 2002.


Note: *Swine waste is liquefied for transportation and disposal.

4-12

Biomass Electricity Generation

4.2.2.2 Structural Relationships and Heteroatoms


The proximate and ultimate analyses relate directly to summative analysisthe distribution of
cellulose, hemicelluloses (e.g., galactoglucomannans, arabinogalactan), lignin, and extractives
(e.g., pinoresinol). Typical distributions of these compounds in softwoods on an extractive-free
basis are cellulose, 45% to 50%; hemicelluloses, 25% to 35%; and lignin, 25% to 35%.
Extractives comprise 1% to 5% of softwoods. Hardwoods are characterized by somewhat lower
concentrations of lignin. Herbaceous materials also are characterized by lower lignin
concentrations. The resulting fuels are highly reactive, with little aromaticity and a significant
concentration of functional groups such as hydroxyl and methoxyl functionalities. Aromatic
structures, where they occur, exist as single aromatic rings rather than fused aromatic clusters.
All of this contributes to the high volatile/fixed carbon ratio of the biomass fuels.
Structurally, carbon 13 nuclear magnetic resonance (13C NMR) research has shown that woody
and herbaceous biomass has an aromaticity of about 8% in sawdust and 10% in weathered
switchgrass (Johnson et al. [7], Tillman [8]). This contrasts with 57% aromaticity in Black
Thunder PRB coal and 73% aromaticity in Pittsburgh #8 coal. Biofuels typically contain six
aromatic carbons per aromatic cluster, whereas the PRB coals contain approximately 10 aromatic
carbons per cluster and the eastern bituminous coals contain 13 to 15 aromatic carbons per
cluster [7]. Further, structural research shows that the concentration of methoxy functionalities
(OCH3) represents 8.67% of the carbon atoms in sawdust and 7.87% of the carbon atoms in
weathered switchgrass. Methoxy functional groups provide a useful measure of fuel reactivity.
As a comparison, 4.21% of the carbon atoms in Black Thunder PRB coal are contained in
methoxy functional groups and 2.27% of the carbon atoms in Beulah lignite are contained in the
OCH3 groups; this percentage drops to 0.79% for western bituminous coals and to 0% for
eastern and Midwestern bituminous coals [7].
The structural analysis is consistent with the coalification results reported by Jenkins et al. [9]
and numerous other authors. Biomass is at the extreme upper right-hand corner on the typical
coalification diagrams plotting hydrogen/carbon atomic ratios on the X-axis and oxygen/carbon
ratios on the Y-axis.
Heteroatom analysis shows significant variability. Nitrogen concentrations, for example, can
vary dramatically. Generally, the woody biofuels have lower fuel nitrogen contents; however,
urban wood waste with high concentrations of glues is a notable exception. Herbaceous materials
tend to have higher nitrogen concentrations, whereas manures have extremely high nitrogen
concentrations, commonly in the form of free ammonia. Concentrations of inorganic matter
(ash) are typically lowest in woody biomass and highest in manures as well. Sulfur contents can
vary significantly, with very low concentrations in woody and herbaceous materials and higher
concentrations in some manures, though even these materials contain less than 1% sulfur.
The high concentrations of nitrogen in urban wood waste are a function of glues such as ureaformaldehyde. Such glues are used in the manufacture of various grades of plywood, OSB, and
MDF, as well as particleboard. The higher concentrations of nitrogen in herbaceous crops result
more from fertilization practices. Common fertilizers such as 30-30-30 have high concentrations
of nitrogen, along with potassium and phosphorus.

4-13

Biomass Electricity Generation

Note the extremely high concentrations of inorganic ash in some of the dairy manure and swine
wastes. Table 4-6 shows the concentrations of fuel nitrogen and ash for selected biomass
materials, expressed in lb/106 Btu.
Table 4-6
Nitrogen and ash concentrations in biomass fuels (values in lb/106 Btu)
Nitrogen (lb/106 Btu)

Ash (lb/106 Btu)

Pine chips

0.24

0.12

Pine shavings

0.24

0.12

Red oak shavings

0.61

1.86

Fresh mixed hardwood-softwood sawdust

0.47

1.19

Urban wood waste

1.67

7.05

Fresh switchgrass

0.70

9.99

Weathered switchgrass

0.49

4.17

Reed canary grass

1.41

5.77

Mulch hay

2.11

6.58

Dairy tie-stall manure

1.71

7.31

Dairy free-stall manure

3.02

171

Poultry litter

5.47

26.6

Sheep manure

3.04

30.31

Swine manure

4.37

45.0

Fuel

Source: Miller et al., 2000.

4.2.2.3 Inorganic Constituents in Biomass


Ash composition is of considerable significance in addressing concerns associated with slagging
and fouling. Traditional ash analyses are presented in Tables 4-7 and 4-8 for representative
biomass fuels and Pittsburgh #8 coal.
Note that the potassium oxide content is always approximately 10%or greaterfor herbaceous
biomass and manures. Woody biomass, on the other hand, typically has lower potassium content
unless it is a short-rotation woody crop grown with an aggressive regime for fuel or fiber
purposes. The higher concentrations of potassium and phosphorus in the herbaceous crop
residues and in the animal manures again result from fertilization practices.
The consequence of the high alkali metal concentrations in the herbaceous materials and
manures is the potential for formation of slagging and fouling deposits. Miles et al. [10] propose
a useful measure of slagging and fouling potential for biomass fuels as follows:

<0.4 lb alkali (K2O + Na2O)/106 Btu, low slagging and fouling potential

0.4 to 0.8 lb alkali/106 Btu, probable slagging and fouling

>0.8 lb alkali/106 Btu, certain slagging and fouling

4-14

Biomass Electricity Generation


Table 4-7
Ash analyses of various biomass fuels compared with Pittsburgh #8 coal
Parameter

Fuel
Pittsburgh
#8 Coal

Pine
Shavings

Reed Canary
Grass

Dairy TieStall Manure

Poultry Litter

Al2O3

25.34

13.4

1.66

2.26

9.14

BaO

0.15

0.05

0.02

0.05

CaO

2.28

8.75

9.57

23.3

12.7

Fe2O3

18.34

5.94

1.47

1.37

4.04

K2O

2.22

4.94

18.1

10.7

9.94

MgO

0.82

3.35

5.29

8.91

4.01

MnO

0.49

0.11

0.14

0.36

Na2O

0.25

1.38

2.34

7.04

3.60

P2O5

0.4

1.44

13.8

14.7

14.0

SiO2

48.2

57.2

43.0

26.0

39.4

SO3

0.67

0.05

0.02

0.14

2.58

SrO

0.80

0.11

0.11

0.03

TiO2

1.16

4.99

5.08

0.51

0.325

0.339

0.741

1.539

0.699

Base/acid ratio

Source: Miller et al., 2002; Miller and Miller, 2002.

Table 4-8
Ash analyses of switchgrass and other herbaceous crops
Parameter

Fuel
Switchgrass
#1

Switchgrass
#2

Alfalfa Stems

Wheat
Straw

Rice Straw

SiO2

65.18

65.42

1.44

55.70

73.00

Al2O3

4.51

6.98

0.60

1.80

1.40

TiO2

0.24

0.34

0.05

0.00

0.00

Fe2O3

2.03

3.56

0.25

0.70

0.60

CaO

5.60

7.14

12.90

2.60

1.90

MgO

3.00

3.17

4.24

2.40

1.80

Na2O

0.58

1.03

0.61

0.90

0.40

K2O

11.60

7.00

40.53

22.80

13.50

P2O5

4.50

2.80

7.67

1.20

1.40

SO3

0.44

2.00

1.60

1.70

0.70

CO2

0.00

0.00

17.44

0.00

0.00

Base/Acid Ratio

0.33

0.30

28.00

0.51

0.24

Sources: Prinzing, 1996; Baxter et al., 1996a; Bryers, 1993.

4-15

Biomass Electricity Generation

Table 4-9 presents the Miles et al. [10] slagging and fouling index for selected biomass fuels.
Note the comparison of the manurestie-stall dairy manure and chicken litterto woody
biomass and herbaceous biomass.
Table 4-9
Slagging and fouling index for selected biomass fuels
Fuel

Index (lb/106 Btu of K2O and Na2O)

Pine shavings and chips

0.010.02

Fresh mixed hardwood/softwood sawdust

0.080.12

Clean urban wood waste

0.45

Alfalfa stems

3.38

Fresh switchgrass

0.801.22

Weathered switchgrass

0.330.51

Reed canary grass

1.18

Tie-stall manure

1.30

Chicken litter

3.60

Sources: Miller et al. (2002); Prinzing (1996); Baxter et al. (1996a); Miles et al. (1993).

Trace metal concentrations in biomass are typically lower than in coal. This has been shown
with repeated testing (see Tillman [8] [11], Payette et al. [12], and Hus and Tillman [13]). In
particular, mercury concentrations, expressed as lb/106 Btu or mg/kJare significantly lower for
biomass than for coal [12].
Trace metals in wood waste, agricultural materials, and other biomass fuels are a reflection
of where they came from. Trees grown in areas where there are high concentrations of trace
metals will exhibit somewhat elevated concentrations of same. Herbaceous crops will exhibit
metal concentrations associated with fertilization practices. Urban wood waste will exhibit
metal concentrations associated with glues, laminates, paints, coatings, and related materials
(see Tillman [14]). Recent studies have shown that mercury concentrations in sawdust are
typically 0.001 to 0.002 mg/kg in the as-received wood, or 0.002 to 0.004 mg/kg in the wood
waste on an oven dry basis [15]. Studies of urban wood waste have shown the following trace
metal concentrations in mg/kg of fresh material [13]:

Arsenic: 2.143 mg/kg

Chromium: 6.570 mg/kg

Lead: 2.923 mg/kg

Mercury: 0.012 mg/kg

Nickel: 2.643 mg/kg

Vanadium: 3.06 mg/kg

4-16

Biomass Electricity Generation

Previous studies [16] have shown wide ranges in trace metal concentrations, as shown in
Table 4-10. These ranges reflect differences in the originating environment of the wood and
wood waste. Table 4-11 shows ranges measured for wood waste burned at a pulp mill in the
Pacific Northwest.
Table 4-10
Ranges in concentrations of trace metals in woody biomass (mg/kg in dry wood)
Metal

Mean

Range

Arsenic

6.5

310

Barium

0.5910

Cadmium

10.4

326

Chromium

31.4

9.192

Lead

72

38127

Nickel

36.6

11.650

Vanadium

53

2779

Zinc

500

200794

Source: Campbell, 1990.

Table 4-11
Ranges in concentrations of trace metals in woody biomass burned at a pulp mill in the
Pacific Northwest (mg/kg in dry wood)
Metal

Mean

Range

Arsenic

0.475

BDL1.8

Barium

51.5

4366

Beryllium

BDL

BDL

Cadmium

BDL

BDL

Chromium

128.4

16524

Chromium +6

0.063

BDL0.3

Lead

2.71

1.16.1

Nickel

137.3

BDL556

99

35207

Zinc
Source: Tillman, 1994.

Trace metals in herbaceous crops and animal manures exhibit the same wide variability. Again,
the originating and processing environments govern trace metal concentrations.

4-17

Biomass Electricity Generation

4.2.3 Performance Characteristics of Biomass Fuels


Performance characteristics of biomass fuels are critical to understanding their behavior in
energy production systems. Fuel volatility, nitrogen evolution patterns, and ash characteristics
are among the parameters of most significance.
4.2.3.1 Biomass Fuel Volatility
Biomass fuels exhibit significant volatility measured through conventional analyses by the
volatile/fixed carbon (V/FC) ratio of the proximate analysis and the hydrogen/carbon (H/C) and
oxygen/carbon (O/C) atomic ratios derived from the ultimate analysis. Representative biomass
fuels exhibit a highly consistent, very volatile pattern. Typical volatility measures for the fuels
shown in Tables 4-3 through 4-5 are presented in Table 4-12.
Table 4-12
Volatility measures for representative biomass fuels
Volatility Measure

Statistical Measure
Average

Standard Deviation

95% Confidence
Interval

Volatile/fixed carbon ratio (V/FC)

5.14

1.27

2.657.63

Hydrogen/carbon atomic ratio (H/C)

1.51

0.12

1.271.75

Oxygen/carbon atomic ratio (O/C)

0.59

0.10

0.390.79

There is a reasonably consistent pattern of volatility measured through conventional techniques.


Further, these measures are significantly higher than the volatility measures for the range of
coals burned in the United States and throughout the world. Eastern bituminous coals from the
Pittsburgh Seam typically have V/FC ratios of 0.40, H/C ratios of 0.69, and O/C ratios of 0.03.
Illinois basin coals have V/FC ratios of 0.70, H/C ratios of 0.81, and O/C ratios of 0.06. PRB
coals from Wyoming can have V/FC ratios of 0.84, H/C ratios of 0.85, and O/C ratios of 0.19
[17].
Alternative fuel analyses provide significantly more information concerning the volatility
of biomass fuels in relation to coal. These analyses are directly related either to the chemical
structure of the fuel or to their performance in combustion environments. Such alternative
measures include maximum volatile yield in drop tube reactor (DTR) experiments, in an inert
environment, including both the percentage yield and the temperature at which that yield was
obtained.
DTR experiments conducted at The Energy Institute of Pennsylvania State University [7] [18]
and reported by Tillman [8] [19] measured the maximum volatile yield of four biomass fuel
samples. Drop tube reactor temperatures varied from 752F (400C) to 3092F (1700C).
All DTR experiments were conducted in an argon atmosphere. In addition to the biomass fuels,
Pittsburgh Seam #8 coal and Black Thunder PRB coal were tested as representative fossil fuels.
Figure 4-4 shows the results of these tests.

4-18

Biomass Electricity Generation

1 0 0 .0 0 %
9 0 .0 0 %
8 0 .0 0 %
7 0 .0 0 %
6 0 .0 0 %

M a x im u m
V o la tile Y ie ld (% )

5 0 .0 0 %
4 0 .0 0 %
3 0 .0 0 %
2 0 .0 0 %
1 0 .0 0 %
0 .0 0 %
S aw dust

U rb a n w o o d

F re s h

W e a th e re d

B la c k

P itts b u rg h

W a s te

S w itc h g ra s s

S w itc h g ra s s

Thunder

#8

F u e l T yp e

Figure 4-4
Maximum volatile yield of biomass fuels compared to reference coals
Sources: Johnson et al., 2001; Johnson et al., 2002

All biomass fuels exhibit maximum volatile yields of about 90+%. This compares with the
lower maximum volatile yields for the two representative coals. It is also significant to note that
the fresh sawdust and fresh switchgrass achieved maximum volatile yields at 1832F (1000C)
while the urban wood waste and weathered switchgrass required temperatures of 2732F
(1500C) to achieve maximum volatile yield.
4.2.3.2 Fuel Nitrogen Characteristics of Biomass Fuels
The overall reactivity of biomass fuels is clearly established. The reactivity of nitrogen from the
various biomass fuels significantly impacts its ability to influence NOX formationor NOX
control. Key measures include fuel nitrogen in lb/106 Btu as well as the following:

Maximum fuel nitrogen volatility, paralleling the maximum fuel volatility

Distribution of volatile fuel nitrogen and char-bound fuel nitrogen

Pattern of volatile fuel nitrogen evolution

These parameters are important in that the volatile fuel nitrogen is more readily reduced to N2
through staged fuel and staged air mechanisms. However, it is important to recognize that not all
nitrogen volatiles are identical, and they evolve at different time histories. Nitrogen volatiles that
evolve more rapidly are more likely to be released from the fuel particle in a fuel-rich region of
the furnace; nitrogen volatiles that are slow to evolve are more likely to be released in a fuel-lean
environment, where surplus oxygen exists.

4-19

Biomass Electricity Generation

Figure 4-5 shows fuel nitrogen concentrations for four biomass fuels extensively analyzed, along
with reference coals. Note the high concentration of fuel nitrogen in the urban wood waste. Total
fuel nitrogen concentration in the sample analyzed is consistent with that measured in the urban
wood waste cofired at the Bailly Generating Station [11]. It results largely from glues in the
plywood and particleboard, OSB, and MDF components of the urban wood waste. It also results
from coatings and coverings, including paints and laminates. The nitrogen contents in the fresh
and weathered switchgrass samples vary significantly, with the weathered switchgrass showing
less fuel nitrogen.

1.8
1.6
1.4
1.2

Fuel Nitrogen,
Lb/MMBtu

1
0.8
0.6
0.4
0.2
0
Sawdust

Urban wood

Fresh

Weathered

Black

Pittsburgh

Waste

Switchgrass

Switchgrass

Thunder

#8

Fuel Type

Figure 4-5
Fuel nitrogen concentrations for the fuel samples analyzed in detail
Source: Johnson et al., 2001; Johnson et al., 2002

Figure 4-6 shows the distribution of nitrogen in volatile or char form for these fuels. Note that
the sawdust contains virtually no char nitrogen. Nitrogen in sawdust has been shown to comprise
largely amine functional groups in the living portion of the wood. The char nitrogen is virtually
identical for the two switchgrass samples and for the urban wood waste. Clearly some of the
nitrogen in the glues does not evolve in volatile form, but remains in the char form. Note also
that the char nitrogen for biomass is significantly lower than that for the reference fuels. The
Black Thunder PRB coal shows increased char nitrogen, and the Pittsburgh Seam coal shows the
highest concentration of fuel nitrogen in char form. These data for coals are well correlated with
the degree of difficulty in controlling NOX emissions during pulverized coal combustion.

4-20

Biomass Electricity Generation

1.8
1.6

Volatile Fuel Nitrogen (light


green)

1.4

Char Fuel Nitrogen (dark green)

1.2

Distribution of
Fuel Nitrogen by
Type (lb/MMBtu)

1
0.8
0.6
0.4
0.2
0
Sawdust

Urban wood

Fresh

Weathered

Black

Pittsburgh

Waste

Switchgrass

Switchgrass

Thunder

#8

Fuel Type

Figure 4-6
Distribution of volatile and char fuel nitrogen in selected biomass and reference coal
samples
Source: Johnson et al., 2001; Johnson et al., 2002

Maximum volatile nitrogen yield significantly affects NOX management. It ties structural
considerations to fuel behavior in combustion settings. Figure 4-7 shows maximum nitrogen
volatile yield as a function of fuel sample. Note that sawdust shows the highest maximum
nitrogen volatile yield, whereas Pittsburgh Seam coal shows the lowest nitrogen volatile yield.
Weathered switchgrass shows the lowest nitrogen volatile yield among the biomass fuels. These
data are consistent with those previously presented in terms of overall fuel volatility and
distribution of fuel nitrogen.
Given these data, it is critical to evaluate the behavior of nitrogen volatiles in biomass and
compare that behavior to the nitrogen volatiles of reference coals. Baxter et al. [23, 24] initiated
investigations into the rate of nitrogen volatile release from coal particles during the pyrolysis or
devolatilization phases of combustion. These investigations recognize that fuel nitrogen in all
coal can be in pyrrole, substituted pyrrole, and pyridine forms, and can be in amine and aniline
forms in the lower rank coals (e.g., lignites) to a very limited extent. These various forms of
nitrogen in fuel volatilize at different temperatures and at different time histories, as the particle
is heated and devolatilized. Volatiles that evolve most rapidly are also most easily controlled.
They tend to evolve in fuel-rich or oxygen-deficient regions of the combustion system. Nitrogen
volatiles retained for longer times in the fuel particle are more likely to be released in fuel-lean
regions of the combustion zone, where excess oxygen is available to oxidize the volatile
nitrogen.

4-21

Biomass Electricity Generation

100.00%
90.00%
80.00%
70.00%

Maximum
Nitrogen Volatile
Yield (%)

60.00%
50.00%
40.00%
30.00%
20.00%
10.00%
0.00%
Urban wood
Sawdust

Waste

Fresh

Weathered

Switchgrass Switchgrass

Black

Pittsburgh

Thunder

#8

Fuel Type

Figure 4-7
Maximum nitrogen volatile yields for selected biomass and coal samples
Source: Johnson et al., 2001; Johnson et al., 2002

Drop-tube reactor experiments by Johnson et al. [7, 18] document volatile nitrogen evolution
patterns for four biomass fuel samples evaluated: sawdust, urban wood waste, fresh switchgrass,
and weathered switchgrass. Nitrogen volatilization patterns are shown as a function of
temperature, recognizing that the temperature history of the particle is a reasonable substitute for
the time history of the fuel particle during the devolatilization phase. In these experiments, the
volatile nitrogen evolution rate was compared to the volatile carbon evolution rate and to the
total volatile matter evolution rate. These volatilization patterns were developed over the entire
temperature range of 400C to 1700C.
Figures 4-8 through 4-11 present the volatile nitrogen and carbon evolution rates for
representative biomass fuels. Among these samples, nitrogen volatiles evolve most rapidly for
sawdust, where nitrogen volatiles evolved more rapidly than the carbon volatile matter,
particularly during the initial stages of pyrolysis. For urban wood waste, the nitrogen evolution
rate essentially parallels the carbon volatile evolution rate. The nitrogen volatile evolution rate
lags just slightly behind the volatile carbon evolution rate for fresh switchgrass. This is quite
similar to the result for urban wood waste. The nitrogen volatile evolution lags significantly
behind the carbon volatile evolution for the weathered switchgrass, indicating the potential for
more difficulty in using this fuel to reduce NOx emissions from biomass combustion.

4-22

Biomass Electricity Generation

Percent Carbon or Nitrogen in Volatile Matter

100.00
90.00
Nitrogen
80.00
70.00
60.00
Carbon
50.00
40.00
30.00
20.00
10.00
0.00
0

200

400

600

800

1000

1200

1400

1600

1800

1200

1400

1600

1800

Temperature (C)

Figure 4-8
Volatile nitrogen and carbon evolution from fresh sawdust
100.00

Volatile Yield of Carbon and Nitrogen

90.00
Carbon
80.00
70.00
60.00
Nitrogen
50.00
40.00
30.00
20.00
10.00
0.00
0

200

400

600

800

1000

Temperature (C)

Figure 4-9
Volatile nitrogen and carbon evolution from urban wood waste

4-23

Biomass Electricity Generation

Percent Carbon or Nitrogen Volatilized

100.00
90.00
80.00
Carbon

70.00
60.00
50.00

Nitrogen

40.00
30.00
20.00
10.00
0.00
0

200

400

600

800

1000

1200

1400

1600

1800

Temperature (C)

Figure 4-10
Volatile nitrogen and carbon evolution from fresh switchgrass
100.00%

Volatile Yield of Carbon or Nitrogen

90.00%
80.00%
Carbon
70.00%
60.00%
50.00%
40.00%
30.00%
Nitrogen
20.00%
10.00%
0.00%
0

200

400

600

800

1000

1200

Temperature (C)

Figure 4-11
Volatile nitrogen and carbon evolution from weathered switchgrass

4-24

1400

1600

1800

Biomass Electricity Generation

Percent Nitrogen or Carbon


Evolved as Volatiles

The data presented here can also be normalized to total volatile evolution, as is shown in Figures
4-12 and 4-13 for the sawdust and weathered switchgrass. Normalizing the data to proximate
analysis, consistent with the Baxter analysis [23, 24], provides a means for relating volatile
nitrogen evolution to traditional fuel analyses. Further, it provides a means for relating the
nitrogen evolution data to typical information used in burner design and operation.
100.00
90.00
80.00
70.00

Nitrogen Volatiiles Formed

60.00
50.00
40.00

Carbon Volatiles Formed

30.00
20.00
10.00
0.00
0.0%

10.0%

20.0%

30.0%

40.0%

50.0%

60.0%

70.0%

80.0%

90.0%

100.0%

Percent Volatile Matter Evolved


Figure 4-12
Nitrogen and carbon volatilization normalized to total volatiles formed from sawdust
100.00%

Percent Evolving as Volatile Matter

90.00%
80.00%
70.00%
60.00%
50.00%
40.00%

Carbon Volatile Yield


30.00%
20.00%
10.00%
0.00%
0.0%

Nitrogen Volatile Yield


10.0%

20.0%

30.0%

40.0%

50.0%

60.0%

70.0%

80.0%

90.0%

100.0%

Percent Total Volatile Yield From Fuel

Figure 4-13
Nitrogen and carbon volatile evolution from weathered switchgrass normalized to total
volatile evolution

4-25

Biomass Electricity Generation

The normalized data demonstrate that the carbon volatile evolution and the total volatile
evolution are virtually identical, with only a slight lag in the carbon volatile evolution caused by
the hydrogen in the fuel.
An alternative approach to evaluating these data is to calculate the nitrogen/carbon (N/C) atomic
ratios of the char remaining from pyrolysis or devolatilization at given temperatures. If the N/C
ratio is declining, then the nitrogen volatiles are evolving more rapidly than the carbon volatiles.
Alternatively, if the N/C ratio is increasing, then the nitrogen is preferentially being retained in
the char. These data can be presented as calculated N/C ratios. These data can also be presented
on a normalized basis, where the initial N/C ratio of the fuel is considered 1.0. N/C ratios in the
char can then be related to the N/C ratio in the fuel on a proportional basis. Both approaches are
presented in Figures 4-14 and 4-15. In Figure 4-14, where normalized data are presented, Black
Thunder and Pittsburgh #8 coals are also presented for comparative purposes. The sawdust
shows the only initial decline in N/C ratios, indicating the preferential evolution of nitrogen
volatiles. The urban wood waste and fresh switchgrass data show that the nitrogen volatiles
evolve at a rate just slightly slower than the carbon volatiles. The weathered switchgrass data
show that the nitrogen volatile evolution lags significantly behind carbon volatile evolution at the
early stages of combustion. The coal data show that the more reactive Black Thunder coal
nitrogen volatile evolution pattern essentially mirrors the carbon volatile evolution pattern, while
there is a significant lag in the nitrogen volatile evolution relative to the carbon volatile evolution
for the Pittsburgh #8 coal.
Nitrogen/Carbon Atomic Ratio in Solid Fuel or
Char

0.035
Urban Wood Waste
0.030

0.025

0.020

Fresh Switchgrass

0.015
Weathered Switchgrass

0.010
Sawdust
0.005

0.000
0

200

400

600

800

1000

1200

1400

1600

1800

Temperature (C)

Figure 4-14
Nitrogen/carbon atomic ratios for biomass chars formed during DTR pyrolysis at various
temperatures

4-26

Normalized N/C Atomic Ratio in Solid Fuel and


Char

Biomass Electricity Generation

2.5

Weathered Switchgrass
2

Fresh Switchgrass
1.5

Pittsburgh #8

Black Thunder
0.5

Fresh Sawdust

Urban Wood Waste

0
0

200

400

600

800

1000

1200

1400

1600

1800

Temperature (C)

Figure 4-15
Normalized nitrogen/carbon atomic ratios for biomass solid fuel and char, compared to
Pittsburgh #8 bituminous coal and Black Thunder Powder River Basin sub-bituminous
coal

The result of the nitrogen evolution analysis is an assessment of the relative desirability of the
four biomass fuels tested. Sawdust clearly has the lowest potential for generating NOx emissions
and weathered switchgrass has the most potential. Fresh switchgrass and urban wood waste are
about on equal footing with respect to nitrogen evolution; however, fresh switchgrass has the
advantage of a lower fuel nitrogen content. Actual impact on NOx emissions will vary
significantly, though, depending on the fraction of biomass burned, pre-existing combustion
conditions, and so on.
4.2.3.3 Ash Reactivity, Slagging, and Fouling
Ash reactivity, slagging, and fouling also provide significant concerns for biomass fuel
utilization. Ash elemental analyses have been previously presented. More significant data can be
obtained from chemical fractionation of the ash, however. In this procedure (see Miller et al. [25,
26] and Miller and Miller [27]), the fuel is successively leached in water, ammonium acetate, and
hydrochloric acid (HCl). The successive leaching provides insights into the reactivity of a given
inorganic constituent by revealing mineral associations of the cations. Matter leached with water
and ammonium acetate is highly reactive; matter leached in HCl is less reactive, whereas the
residue from this procedure is considered to be relatively unreactive. Examining selective
leaching results has shown to be an effective method of predicting deposition in coals as well.
Baxter et al. [24] and Bryers [28] have used this procedure to evaluate a wide array of biomass
fuels, and a very similar technique is used to assess the deposition probabilities of coals and
other mineral effects.
4-27

Biomass Electricity Generation

Table 4-13 presents chemical fractionation of an array of biomass fuels developed by Miller et
al. [25]. This includes wood, herbaceous material (Reed Canary grass), and manure samples.
Table 4-13
Chemical fractionation analyses of various biofuel ashes
Parameter

Fuel
Pine
Shavings

Reed
Canary
Grass

Sheep
Manure

Dairy TieStall Manure

Poultry Litter

Water soluble and


ion exchangeable

65

97

97

96

41

Acid soluble

Insoluble

35

20

Water soluble and


ion exchangeable

77

98

96

99

90

Acid soluble

Insoluble

23

10

Water soluble and


ion exchangeable

92

92

58

68

41

Acid soluble

41

31

58

Insoluble

Water soluble and


ion exchangeable

72

95

77

90

69

Acid soluble

10

21

19

Insoluble

18

12

Water soluble and


ion exchangeable

30

25

32

12

Acid soluble

13

Insoluble

70

72

68

75

100

Water soluble and


ion exchangeable

29

55

Acid soluble

Insoluble

71

45

100

100

100

Potassium

Sodium

Calcium

Magnesium

Aluminum

Silicon

Source: Miller et al., 2002b.

4-28

Biomass Electricity Generation

Table 4-14, from Baxter et al. [24], presents chemical fractionation of switchgrass.
Table 4-14
Chemical fractionation of potassium in switchgrass for four samples
Parameter

Switchgrass Sample
A

C1

C2

Water soluble

52

68

70

72

Ion exchangeable

35

24

20

16

Acid soluble

Residual

10

10

Source: Baxter et al., 1996b.

Baxter, as reported by Miles et al. [10], performed one complete analysis of switchgrass as
shown in Table 4-15. These results show the high reactivity of the alkali, and alkali earth
minerals along with the low reactivity of the silica, alumina, and titanium dioxide.
Table 4-15
Complete chemical fractionation of a switchgrass sample
Ash Component

Chemical Fraction (wt %)


Water Soluble

Ion Exchangeable

Acid Soluble

Residual

SiO2

100

Al2O3

100

TiO2

100

Fe2O3

40

60

CaO

80

10

MgO

32

48

13

Na2O

52

48

K2O

70

20

10

P2O5

70

15

SO3

26

69

Source: Miles et al., 1993.

Chemical fractionation provides key insights into the reactivity of ash and its potential to cause
slagging and fouling. It provides insights into potential combustion problems of various biomass
fuels when used in conjunction with total ash concentration. Reactivity places a premium on
low-ash biomass fuels and demonstrates some of the difficulties with the herbaceous materials. It
also demonstrates that trees grown for fuel with high doses of fertilizer containing potassium
could generate significant slagging and fouling deposits.

4-29

Biomass Electricity Generation

4.3 Characteristics of Biomass Cofiring Technologies


Currently the technologies that are commercially offered include various forms of cofiring along
with stand-alone (100% biomass-fired) Rankine-cycle generating systems. Cofiring refers to the
practice of firing biomass fuels as a supplement to coal (or very rarely, gas). Stand-alone
biomass plants may either supply superheated steam to condensing turbines where the only
product is electricity, or they may supply steam to backpressure or automatic extraction turbines
in cogeneration (also known as combined-heat-and-power or CHP) applications.
Cofiring systems are most readily adapted to electric generating stations. Biomass can be
integrated with the fuel supply to existing boilers designed to utility standards. The biomass can
be used in large reheat boilers where steam is used most efficiently. Such systems are less
expensive than stand-alone power plants on a $/kW supported by biomass basis. If the biomass is
unavailable temporarily, the operation of the unit is not compromised. However, integrating
biomass fuel into the coal stream still involves complex issues of materials handling and control.
Further, cofiring does not contribute additional capacity; instead, it displaces coal fuel at the unit.
When gaseous biomass fuels (e.g., landfill gas, wastewater treatment gas, producer gas) are fired,
additional considerations and benefits are available. Figure 4-16 represents a typical
configuration of biomass cofiring in a coal-fired utility boiler.

Mechanical
draft cooling
tower
Cooling
water
Turbine/
generator

Electricity

Steam (2400 psi, 1000 deg F,


reheat to 1000 deg F)

Moisture losses
Biomass

710 F
269 F

Alternate
Fuel
Receiving

Utility boiler

Air heater
610 F

Coal

Bottom ash

ESP
80 F
Combustion
air

Ferrous

Figure 4-16
Schematic of biomass cofiring with coal in a utility boiler

4-30

Stack
Fly ash

Biomass Electricity Generation

4.3.1 Overview of Solid Biomass Fuel Cofiring Systems


Biomass cofiring refers to a family of technologies designed to introduce these renewable fuels
into fossil-fuel-fired systems. The common approach is to fire solid or gaseous biomass fuels in
coal-fired boilers. Differences in applicable technology are a function of the following issues:

The biomass fuel being fired (woody biomass, herbaceous biomass, gaseous fuel from a
landfill, a wastewater treatment facility, or an atmospheric gasifier)

The boiler type (cyclone, tangentially fired pulverized coal, wall-fired pulverized coal,
fluidized bed)

The percentage cofiring sought

The design and operating parameters of the unit (e.g., the presence of spare pulverizer
capacity, the use of low-NOX burners).

There are certain combinations that are highly useful, but others simply do not work. Technology
selection is based upon putting the most appropriate combination in place.
Numerous facilities have demonstrated cofiring (Table 4-16). Of these, cofiring has continued at
the Greenidge Station, the Big Stone Plant, Plant Gadsden, and the Willow Island Generating
Station, at a minimum. It was previously practiced for many years at the King Generating Station
of Northern States Power and several other locations. In Europe, Electrabels Gelderland station
cofires commercially, as do many other sites in Europe, including the Drax power station, Dong
Energys Studstrup station, and EDF Polskas Krakow station.
4.3.2 Cofiring in Cyclone Boilers
Cyclone firing demonstrates that the key issue of cofiring in either cyclone boilers or pulverized
coal (PC) boilers is fuel handling. Combustion considerations dictate many of the environmental
and technical consequences of cofiring. Materials handling concerns have the major impact on
economics.
Cyclone boilers are commonly considered most appropriate for cofiring dissimilar solid fuels
biomass fuels, tire-derived fuel (TDF), petroleum coke, and other opportunity fuels. In cyclone
firing, fuel crushed typically to 0.375-inch 0-inch (9.5 mm 0 mm) is stored in bunkers. The
outfeed from the bunkers falls by gravity through the downcomers to the feeders, typically Stock
gravimetric feeders, and is then fed by gravity to the cyclone burners. Rather than being burned
in suspension, the fuel is burned in a cyclone, forming a slag layer on the cyclone barrel.

4-31

Biomass Electricity Generation


Table 4-16
Representative biomass cofiring tests and demonstrations
Owner at Time
of Demo

Boiler Type

Biomass
Fired

Responsible
Organization

Allen Fossil Plant

TVA

Cyclone

Sawdust

TVA/EPRI/DOENETL

Kingston Fossil Plant

TVA

T-fired PC

Sawdust

TVA

Colbert Fossil Plant

TVA

Wall-fired PC

Sawdust

TVA

Plant Hammond

Southern Co

Wall-fired PC

Sawdust

Southern Co.

Plant Kraft

Southern Co

T-fired PC

Sawdust

Southern Co.

Alabama Power

T-fired PC

Switchgrass,
sawdust

Southern Co.,
DOE, EPRI

Shawville Generating
Station

GPU Genco

Wall-fired PC
T-fired PC

Sawdust

GPU/EPRI/DOENETL(2)

Seward Generating
Station

GPU Genco

Wall-fired PC

Sawdust

GPU/EPRI/DOENETL(2)

Greenidge Station

NYSEG/AES(1)

T-fired PC

General
wood waste

NYSEG/
NYSERDA(3)

Blount St. Station

MG&E

Wall-fired PC

Switchgrass

DOE/EPRI

Michigan City
Generating Station

NIPSCO

Cyclone

Urban wood
waste

EPRI/DOE-NETL(2)

Bailly Generating Station

NIPSCO

Cyclone

Urban wood
waste

EPRI/DOE-NETL(2)

Albright Generating
Station

Allegheny Energy

T-fired PC

Sawdust

Allegheny
Energy/DOENETL(2)/EPRI

Willow Island Generating


Station

Allegheny Energy

Cyclone

Sawdust

Allegheny
Energy/DOENETL(2)/EPRI

Dunkirk Generating
Station

Niagara Mohawk

T-fired PC

Sawdust/
SRWC

DOE-GFO(4)

Alliant Energy

T-fired PC

Switchgrass

DOE-GFO(4)

TECO

Coal gasifier
IGCC

SRWC

DOE-NETL(2)

Southern Co

T-fired PC

Switchgrass

DOE-GFO(4)

Midkraft

Wall-fired PC

Straw

Elsamprojekt

Lahden
Lampovoima Oy

Wall-fired PC

General
wood waste

Plant

Plant Gadsden

Ottumwa Generating
Station
Polk County Generating
Station
Plant Gadsden
Studstrup(5)
Lahti (gasification)

FWOy/EU Thermie
Program

Notes:
1. Many of these units have changed hands since the demonstration (e.g., AES has purchased the Greenidge Station formerly owned by
NYSEG).
2. NETL is the National Energy Technology Laboratory of DOE.
3. NYSERDA is the New York State Energy Research and Development Agency.
4. GFO is the Golden Field Office of DOE.
5. One example of many European cofiring demonstrations.

4-32

Biomass Electricity Generation

The flow of fuel to the bunkers, coal feeders, and cyclones is appropriate for granular and
regularly shaped particles. Wood waste, such as sawdust at 18 lb/ft3, works well in cyclone
boilers. However, cyclone technology is not suitable for switchgrass and other herbaceous
materials, as reported by Bush et al. [29]. Jenike and Johanson of Westford, Massachusetts, state
that, in evaluating co-feeding and co-milling switchgrass and coal at Plant Gadsden for Southern
Co., they analyzed blends of switchgrass and Pratt Seam coal. Their research documented that a
blend of 5% switchgrass and 95% coal on a mass basis would not flow from the bunker. The
needle-like characteristics of the switchgrass particles, combined with the low bulk densities of
switchgrass (e.g., 5 to 7 lb/ft3), resulted in continuous bridging. As a consequence of the handling
characteristics of herbaceous materials, all cyclone demonstrations have been on woody biomass.
Cyclone firing is unique in that the coal is not pulverized, combustion occurs at very high peak
flame temperatures (e.g., 3300F to 3550F), and inorganic matter is largely converted into a
liquid slag, which is then tapped from the furnace bottom. Cyclone firing typically results in
about 70% of the mineral matter in the coal or fuel mass being removed as bottom ash or slag,
leaving only 30% of the mineral matter to exit the boiler as fly ash. Slag or bottom ash is
typically sold for sandblasting grit, roofing granules, and similar products.
Cyclone firing can generate high concentrations of NOX. Uncontrolled combustion NOX
emissions of 1.2 to 2.0 lb/106 Btu have been generated, depending on boiler size and
configuration (front wall vs. opposed wall configurations of the cyclone barrels). More recently,
most cyclone boilers have been equipped with overfire air systems to reduce NOX emissions to
the 0.3 to 0.5 lb/106 Btu range. No new cyclone boilers have been built since about 1975, due in
large part to the NOX emissions associated with these units.
Cyclones, originally designed for slagging coals of the Midwest, have had success in firing a
range of coals. Consequently, cyclone boilers were adapted to cofire petroleum coke [19], tirederived fuel (TDF) [30], and other opportunity fuels with coal. In the 1970s, some experiments
cofired refuse-derived fuel (RDF) and coal in cyclone boilers. In the 1990s, Northern Indiana
Public Service Co. (NIPSCO) cofired tar-laden soils from manufactured gas plants with coal in
one cyclone boiler. In the early 1990s, Northern States Power also began cofiring dry wood
waste from Andersen Windows with PRB coal at its 600-MWe King Generating Station.
4.3.2.1 Cofiring System Design for Cyclone Boilers
With the exception of the King Station, where dry sawdust from a neighboring window
manufacturing facility was blown into the secondary air system of three of 12 cyclone barrels, all
demonstrations have blended woody biomass with coal and/or other opportunity fuels on the coal
pile or on the conveyor leading to the bunker system. Consequently, the cofiring system has to
have sufficient transport capacity to load the entire requirement for biomass within a two-to fourhour window depending upon loading schedules at the plant.
Critical variables associated with cyclone cofiring systems include the following:

Project design life

Degree of automation desired

Space available, including both location and acreage

Soil conditions where the cofiring system is intended to go

Plant standards and preferences


4-33

Biomass Electricity Generation

Project design life can be between three and 10 years, as the cofiring system addresses existing
plants built before 1975. This impacts the robustness of the design. Degree of automation is
critical. If the plant accepts a low degree of automation, the capital cost can be minimized;
however, labor requirements increase.
As a minimum, cofiring requires a pile and a reclaim conveyor connected to the main belt
conveyor feeding the bunkers or silos. This requires a truck unload location and a conveyor with
a belt scale for control. One to two workers are necessary to operate the system. It also requires
accepting the risk of tramp material in the biomass. Absent accepting this risk, a magnet must be
installed somewhere on the conveyor. This system can be enhanced by a pole barn or similar
structure to contain biomass dust. At the other extreme, the system can be built to accept any
type of truck and automatically screen biomass for oversized particles, with a grinder for
oversized particles, a stock-out bin, a reclaim system, and controls to manage the blend
automatically. This may require one-quarter to one-half of a worker. Space available dictates
whether a hydraulic truck unloader can be successfully installed. It also dictates conveyor
lengths, conveyor selection, and ease of operation. Soil conditions can be critical in determining
whether spread footings are acceptable or whether pile construction is required. Plant standards
dictate whether or not the electrical installation must be explosion-proof, the extent to which fire
suppression systems are required, the method and degree of system controls, and the extent to
which redundancy is built into the system. All of these are issues with significant economic
consequences.
An example of a cyclone cofiring system is the Willow Island Generating Station boiler #2
demonstration hosted by Allegheny Energy Supply Co., LLC. The Willow Island Generating
Station boiler #2, located in Willow Island, West Virginia, is a 188-MWe (net) unit equipped
with five cyclone barrels, all located in two rows or elevations on the front wall of the boiler. The
boiler operates in a positive pressure mode. The cyclones are equipped with scroll feeders.
Willow Island #2 boiler burns a washed Pittsburgh Seam coal. Limestone is periodically added
to the coal to enhance the slag-forming properties of the fuel. The plant is equipped with a hot
side electrostatic precipitator (ESP) for particulate control. Overfire air (OFA) was subsequently
added to the #2 boiler to reduce NOX emissions.
The design criteria for the Willow Island cofiring system included the following:

The capability to blend 15% sawdust with coal (mass basis) on the main coal belt,
recognizing that normal operation would involve blending 10% sawdust with coal on the
main belt.

The capability to receive and process at least 150 to 200 tons (six to eight truckloads) of
sawdust daily in walking floor vans.

The capacity to store more than three days of sawdust on-site to supply the boiler with
biomass during weekends, including holiday weekends.

The construction of a facility with a design life of more than three years.

The capability to screen the entire mass of biomass fuel to protect the crushers and cyclone
barrels from undesirable material.

4-34

Biomass Electricity Generation

The capability to grind oversized particles to a fuel specification of less than 0.25-inch
particles, consistent with maximizing the volatile matter release in the cyclone barrels.

The capability to meter the flow of biomass to the cyclone barrels in order to measure the
blends being put up in the bunkers.

In addition, the design was robust in order to minimize maintenance commitments.

Figures 4-17 and 4-18 show the system design. In this system, walking floor vans discharge their
load to a live-bottom hopper. The bottom of the hopper is a conveyor belt carrying the sawdust to
a 50-ton/hr disc screen capable of screening 0.25-inch 0-inch sawdust particles. Oversized
material is discharged from the screen to a slow speed/high torque grinder that reduces the wood
particles to the desired size. Acceptable particles are then carried by mechanical conveyor to a
408-ton live bottom storage bin. This bin accepts the material through three openings. It has the
capability to level the storage pile of sawdust using three leveling screws at the top of the bin.
The floor of the bin is a walking floor capable of moving the entire mass of sawdust in the bin.

Figure 4-17
Plan view of the Willow Island biomass cofiring system

4-35

Biomass Electricity Generation

Figure 4-18
Elevation view of the Willow Island biomass cofiring system

Sawdust is discharged from the bin by twin counter-rotating augers that convey the material to a
weigh belt conveyor. The augers operate at variable speed and meter the sawdust to the weigh
belt conveyor. The weigh belt conveyor carries sawdust to the main coal conveyor from the
crusher house to the power plant bunkers. The system is designed to convey up to 75 ton/hr of
sawdust to the main coal conveyor. Sawdust is added to the conveyor belt before the automatic
belt samplers. TDF is added to the conveyor belt separately, after the belt samplers.
The design was based on walking floor van truck transportation. Value engineering dictated that
walking floor vans are used due to the costs of hydraulic chip van dumpers. Further, dust
management issues made use of trailer dump trucks less than desirable. Finally, there was
insufficient space at the Willow Island site to house a truck dump other than a walking floor
receiving hopper.
The design then focused on screen selection. The disc screen provided an acceptable particle size
for combustion without generating the levels of dust associated with trommel screens. Further,
disc screens experience little blinding. Covered mechanical conveyors, selected both for
maintenance and dust management reasons, were used to load the bunker, rather than pneumatic
conveyors previously used to feed silos.
The walking floor bin was chosen over a silo because of storage requirements and height
considerations, in addition to consequent maintenance concerns when dealing with the head
pulley of the conveyor feeding the bin. There were ample reasons for using either type of storage
system, and customer preference ultimately influenced the choice of the storage facility. Covered
contained storage of the prepared fuel is always employed at power plants due to dust
management considerations.
Some additional features were incorporated into the design. A small permanent vacuum system
was added to facilitate general site cleanup. Dock seals were added around the truck unloading
hopper to minimize sawdust spillage at that location (Figure 4-19 shows a truck unloading
sawdust). An inspection door was added to the twin auger conveyor moving sawdust from the
walking floor bin to the weigh belt metering conveyor. These features improve the overall
performance of the system.
4-36

Biomass Electricity Generation

The system was installed on a site requiring extensive piling, which added considerably to the
cost. The system also met plant requirements for an explosion-proof electrical system.

Figure 4-19
Receiving sawdust at the Willow Island Generating Station

4.3.2.2 Capital Costs of Cofiring Systems for Cyclone Boilers


The capital costs associated with the Willow Island Generating Station are shown in Table 4-17.
Because this unit was built with the potential to supply 75 ton/hr to the main coal beltequal to
15% of the coal feed on a mass basis, or 7% of the coal feed on a thermal input basisit had the
capacity to support 13 MWe (net). Normal operation was between 5% and 10% of the total coal
input on a mass basis, or 2.2% to 4.5% on a thermal input basis. The resulting capital cost of the
Willow Island system was approximately $180/kW based on the kilowatt cofiring capacity
provided by the biomass fuel. Assuming a normal operating capacity of 50 ton/hr feed to the coal
bunkers, or 10% of the total coal flow on a mass basis, then the capital cost of this system is
$275/kW supported by biomass.
The capital cost of the Willow Island cofiring system can be compared with that required to
construct the Bailly Generating Station cofiring system, as shown in Table 4-18. The Bailly
Generating Station cofiring system consists of a pole barn, a trommel screen, a pad and
aboveground reclaimer, and a metering conveyor belt. The cost was $1.18 million. It had the
capacity to supply 10% of the mass feed, or 5% of the thermal input for the Bailly Generating
Station #7 boiler, a 165-MWe cyclone unit. On that basis, the capital cost of the Bailly cofiring
system was $140/kW supported by biomass.

4-37

Biomass Electricity Generation


Table 4-17
Capital cost of the Willow Island cofiring demonstration
Number
1

Description

Capital Cost

Process Equipment

$920,200

Walking floor bin


Outfeed conveyor from bin
Cross feed weigh belt conveyor
Scale
Sawdust receiving hopper
Conveyor to disc screen
Disc screen
Overs grinder
Outfeed conveyor
Vacuum system
Motor control center
2

Engineering Studies

$194,400

Soils studies
Engineering
3

Construction

$1,114,400

Home Office Support and Other

$86,800
TOTAL

$2,315,800

Source: Tillman and Payette, 2004.

Table 4-18
Capital cost summary for the Bailly Generating Station cofiring system
Number
1

Description

Capital Cost

Process Equipment

$587,700

Trommel screen
Aboveground reclaim
Air slide metering conveyor
2

Engineering and Installation

$327,400

Home Office Support and Other

$266,500
TOTAL

$1,181,600

Source: Tillman, 1999.

The cost range, $140 to $180/kW supported by biomass, is a reasonable estimate of the capital
costs associated with cofiring in cyclone boilers, particularly at the 8- to 12-MWe range. Because
these systems are largely materials handling systems, the exponential scale factor for estimating
larger systems would be 0.65 to 0.8. The $1.18 million cost of the Bailly system could well
represent a minimum cost for cofiring at smaller scales.

4-38

Biomass Electricity Generation

The differential operating costs associated with automation is due to additional labor
requirements. Annual maintenance costs can be estimated at about 4% to 5% of the initial capital
investment.
4.3.2.3 Cofiring Impacts in Cyclone Boilers
Cofiring in cyclone boilers has been tested and demonstrated in a significant number of plants
including the Allen Fossil Plant of TVA, the Bailly Generating Station of NiSource, the
Michigan City Generating Station of NiSource, and the Willow Island Generating Station of
Allegheny Energy.
Efficiency impacts have been modest. At the Willow Island cofiring demonstration, 10% cofiring
(mass basis) resulted in an increase in the net station heat rate (NSHR) of 30 Btu/kWh.
Efficiency penalties of up to 0.5% have been measured when cofiring at 10% mass basis (5%
thermal basis) at the Allen Fossil Plant, the Bailly Generating Station, and the Michigan City
Generating Station. This is about equal to 50 Btu/kWh [11].
Operationally, cofiring increased the duty on the coal feeders to the cyclones, as would be
expected. In no case did cofiring limit unit capacity. Flame temperatures were largely unaffected.
Further, flame intensity of sawdust and finely divided biomass equals that of coal. Furnace exit
gas temperatures (FEGT) decreased in all cofiring tests at Willow Island and at the other cyclone
demonstrations. Figure 4-20 depicts this phenomenon at Willow Island.
3000

Furnace Exit Gas Temperature (F)

FEGT when firing coal alone


2500

2000

1500
FEGT when cofiring coal and sawdust
1000

500

0
0.0

1.0

2.0

3.0

4.0

5.0

6.0

7.0

8.0

9.0

Percent Sawdust in Fuel (mass basis)

Figure 4-20
Influence of cofiring in cyclone boilers on furnace exit gas temperature
Source: Tillman and Payette, 2004

4-39

Biomass Electricity Generation

Cofiring can also reduce the viscosity of slag formed in the cyclone barrel, increasing the
flexibility of coal chosen. Figure 4-21 shows the impact of cofiring on slag viscosity as
determined by FACTSAGE modeling performed by Miller et al. [25]. Note the significant
change in viscosity from 100% bituminous coal to various cofiring approaches.
10000

1000

Viscosity, poise

Fuel 1 (100% coal)

100

Fuel 1
Fuel 2

10

Fuel 3
Fuel 4
Fuel 5
Fuel 9

1
1200

1600

2000

2400

2800

3200

3600

4000

Temperature, F

Figure 4-21
Impact of cofiring on slag viscosity (Fuels 2 to 9 are various cofiring scenarios using woody
biomass, herbaceous biomass, and animal manures with Eastern bituminous coals.)
Source: Miller et al., 2002

Typically, cofiring has reduced uncontrolled NOX emissions; however, this has not always
occurred. At Bailly Generating Station, NOX reductions of about 0.1 lb/106 Btu occurred when
cofiring at 10% wood (mass basis), and were far more dramatic when blending petroleum coke
with wood and coal. NOX emissions reductions at Allen Fossil Plant were up to 0.2 lb/106 Btu
when cofiring at 10% (mass basis), depending on the base coal employed. At Willow Island
Generating Station, NOX emissions did not decrease or increase. SO2 emissions decreased as a
function of wood substitution for coal in the boiler.
Testing at Bailly Generating Station [11] demonstrated that carbon monoxide and hydrocarbon
emissions did not increase or decrease as a result of biomass cofiring. The influence on SO3
emissions was inconsequential (approximately 2 ppmvd @ 3% O2).
In all cases, mercury (Hg) emissions decreased as a function of the concentration of Hg in the
feed fuels.

4-40

Biomass Electricity Generation

4.3.3 Cofiring in Pulverized Coal (PC) Boilers


Cofiring in PC boilers can be accomplished either by blending biomass with coal on the main
belt feeding the coal bunkers or by separately injecting biomass directly into the furnace
equipped with either burner modifications or specially designed biomass burners. Blending on
the belt is limited to woody biomass. Herbaceous biomass such as switchgrass causes significant
problems in this application, as previously discussed [29]. With woody biomass, initial research
by TVA showed that the percentage of biomass in the feed to the bunkers must be less than 5%
(mass basis) in order to minimize problems with the mills. Testing at Shawville Generating
Station of GPU Genco (now Reliant Energy) demonstrated that pulverizers were negatively
impacted even at 3% biomass (mass basis). Ongoing work at the Gadsden Station is determining
the maximum allowable particle size for wood input to the pulverizer. Southern Company has
reported that impacts and limitations on pulverizer performance are quite variable, depending on
the fuel, mill type, recent maintenance history and other factors. If boilers are equipped with
Riley Atritta mills, the maximum biomass percentage may be about 8%.
Most demonstrations of PC cofiring in the United States in both tangentially fired boilers and
wall-fired boilers have focused on separately preparing and injecting biomass into the primary
furnace. Internationally, there has been much more interest in co-milling of biomass and coal.
Examples of domestic tests and demonstrations include the following:

Greenidge Station

Blount St. Station

Seward Generating Station

Gadsden Station

Ottumwa Generating Station

Albright Generating Station

Dunkirk Generating Station

Buck Station

Lett Station

4.3.3.1 Cofiring System Designs for Pulverized Coal Boilers


There have been several cofiring demonstrations using separate injection. Of these, the Albright
Generating Station of FirstEnergy provides key insights into design and cost considerations.
The process design for the Albright Generating System cofiring demonstration somewhat
paralleled the design for Willow Island in concept. However, numerous differences occurred in
equipment selection, largely due to the following factors:

Moving to an existing demonstration at Seward Generating Station, where equipment had


already been selected based on different project criteria

Favoring pneumatic conveyance of sawdust, rather than mechanical conveyance of sawdust

Injecting the sawdust directly into the boiler, requiring changes in the final handling and
control approaches
4-41

Biomass Electricity Generation

The Albright design involved accepting sawdust through a walking floor receiver. This receiver
is slightly wider than a walking floor van, and can hold the entire load of a single truck. This
provides for surge capacity in the unloading process. The discharge of the walking floor unloader
is a series of screw conveyors transporting the sawdust to a 30-ton/hr disc screen.
A disc screen was chosen for Albright Generating Station rather than a trommel screen or deck
screen. Dust management was the critical issue, and a disc screen with covers generates less dust
in the work area than a trommel screen. Disc screens have been successfully deployed in the
forest products industry as well as the waste-to-energy industry. This screen produces a product
for suspension firing at less than 6.35 mm (<0.25-inch). The size chosen was based upon
extensive research conducted from 1992 to 1995, and the success of both TVA and Southern
Company cofiring <6.35 mm (<0.25-inch) sawdust in PC boilers.
The <6.35 mm (<0.25-inch) sawdust from the screen was deposited in a bin and then
pneumatically transported to the top of the agricultural silo. Material rejected as being oversized
was originally ground in a research grinder. When that proved unsuccessful in a commercial
application, it was returned to the vendor and the oversized materialless than 5% of the
incoming feedstockwas disposed of. Some employees used a portion of this as mulch on their
own property, while some was landfilled. Eventually a two-stage grinder and associated dust
management system was purchased to reduce the size of oversized product.
From the silo, sawdust was reclaimed on demand, fed through a surge hopper and onto a weigh
belt conveyor feeding a live bottom bin. The live bottom bin discharged sawdust into two rotary
airlocks, each supported by blowers. This arrangement transported sawdust to opposite corners
of the T-fired boiler. Processing equipment was housed in a metal building in order to manage
the entire process. The building also housed the motor control center and all other electrical
systems.
The Albright system delivered up to 6 tons/hr to the boiler, supporting the generation of
approximately 6.0 MWe or about 4.3% of the net generation of the #3 boiler. Injection into the
boiler was accomplished with two specially designed injectors capable of following the burner
tilts. The boiler has four rows of burners, with the injectors installed between the B and C rows.
The burners and sawdust delivery system were designed with a tip velocity of 25.5 m/s
(5,000 ft/min) and with an air/fuel mass ratio of 2 kg air/kg fuel. This achieved a substoichiometric air/fuel ratio and a fuel speed sufficient to overcome the flame speed of biomass.
The first consideration was essential for NOX control; the second consideration was a safety
issue.
The system used two independent control systems: one to manage the flow of fuel to the silo and
a second to manage the flow of sawdust to the boiler. The second set of controls allows the
system operator to set the flow of sawdust up to 5.5 tonnes/hr (6 tons/hr) to the boiler at his
discretion. It also permits the operator to shut down the flow of sawdust with an emergency stop
if, for any reason, the operator believes that such action is in the best interest of the boiler and the
power plant.
The design for Seward Generating Station, also using sawdust, was virtually identical to Albright
with one exception: instead of having a separate biomass injector, the Seward wall-fired burners
were modified to fire sawdust into the center of the coal burner. The sawdust and the coal
interacted in the flame envelope (Figure 4-22).
4-42

Biomass Electricity Generation

Both Albright and Seward used separate blowers and separate transport lines for each biomass
injection point. An alternative used at both Plant Gadsden and Dunkirk Generating Station
acquired biomass from the bin using an eductor and a single transport line. This transport line
then employed a splitter to convey the biomass to individual burners.

Figure 4-22
Design of the sawdust injection system at Seward Generating Station

4.3.3.2 Capital Costs of Separate Injection PC Cofiring Systems


As demonstrated at Colbert Fossil Plant and Kingston Fossil Plant of TVA, Shawville
Generating Station of GPU Genco, and Picway Station of American Electric Power, the capital
costs associated with cofiring in PC boilers via blending on the main coal belt are comparable to
those of cofiring in cyclone boilers. The capital costs associated with separate injection are
somewhat higher.
Table 4-19 summarizes the capital costs for the Albright project, assuming that it would be
constructed from completely new equipment. Given that the system could support 6.0 MWe, the
installed capital cost was $300/kW output supported by biomass. It may be surprising that the
cost of reasonably automated cofiring systems for cyclone and separate injection PC boilers are
separated by only about $120/kW. However, it must be remembered that cyclone systems must
deliver 24 hours of fuel in a two- to three-hour span (while bunkers are filling). Separate
injection systems deliver significantly lower quantities of biomass to the boiler (or bunkers) on a
per-hour basis.

4-43

Biomass Electricity Generation


Table 4-19
Estimated capital cost of the Albright cofiring demonstration if constructed as a
new facility
Number
1

Description
Process Equipment

Capital Cost
$882,700

Burners
Disc screen
Overs 2-stage grinder
Dust control for grinder
Screw conveyor
Sawdust receiving bin, blower, and cyclone
Storage silo and unloader
Paddle conveyor
Surge hopper and live bottom bin
Weigh belt feeder
Rotary airlocks and blowers
Walking floor unloader
Motor starters, motor vontrol venters, and related
System controls
2

Engineering Studies

Construction

$710,000

Home Office Support and Other

$135,600

TOTAL

$82,100

$1,810,400

Source: Tillman and Payette, 2004.

Capital costs associated with herbaceous crop cofiring systems are likely to be substantially
higher than those associated with biomass. Experience at the Ottumwa Generating Station and
Greenidge stations, as well as at the Grena, Avedore, and Studstrup Power Plants of DONG
Energy in Denmark, demonstrate the need to commit significant resources to dry storage of the
biomass and to extensive grinding systems. Further, because herbaceous biomass is largely a
seasonal product, investments are required to provide this material on a year-round basis.
4.3.3.3 Impacts of Cofiring Biomass in PC Boilers
The impacts of separate injection cofiring solid biomass fuels in PC boilers are distinct and
separate from those for cofiring in cyclone boilers. Biomass cofiring does not reduce boiler
capacity. In winter, when wet coal reduces capacity, biomass cofiring can facilitate recovery
of some of the lost capacity as demonstrated at Seward Generating Station [8, 19, 31].
Efficiency penalties can be very modest or somewhat significant, depending on system design
and operation. Typically, biomass is introduced with ambient air. This reduces the combustion
air passing through the air heater and raises the temperature of the gaseous combustion products
exiting the air heater. Further, like cofiring in cyclone boilers, moisture and hydrogen in the fuel
cause a small boiler efficiency penalty. For example, at Albright Generating Station, the
efficiency penalty was about 3.5 Btu/kWh based on sawdust heat input. This resulted in a penalty
of 35 Btu/kWh at 10% cofiring.
4-44

Biomass Electricity Generation

Emissions impacts are clear and significant. SO2 emissions decline as a function of biomass
Btu substitution in the boiler. Similarly, Hg emissions decrease as a function of Btu substitution.
Assuming the test results at Albright Generating Station can be generalized to all units, CO and
opacity emissions are not affected by cofiring. NOX emissions decrease, particularly as a function
of the location of the biomass injection.
At Seward Generating Station, NOX emissions decreased dramatically when cofiring sawdust,
and the data are represented by Equation 4-1:
NOX = 0.030 + 0.0017(L) + 0.082(EO2) 1.917(Wh)

Equation 4-1

where NOX = oxides of nitrogen, lb/106 Btu, L = load measured as main steam flow in 1000
lb/hr, EO2 = excess O2 reported in the control room (total basis), and Wh = wood cofiring
percentage, heat input basis. The r2 for Equation 4-1 is 0.93. Further, the probability that any
term is a random occurrence is small. These probabilities are respectively L = 2.09 10-5; EO2 =
2.53 10-5; Wh = 8.39 10-7; and the overall probability is 4.36 10-6.
NOX can also be expressed on a mass basis, as is shown in Equation 4-2:
NOX = 0.026 + 0.0017(L) + 0.083(EO2) 0.899(Wm)

Equation 4-2

where Wm = wood cofiring percentage, mass basis; r2 = 0.93. The probabilities in Equation 4-2
are essentially identical to those of Equation 4-1.
The NOX reduction at Albright Generating Station indicated in Figure 4-23 is represented by
the equation:
NOX = 0.361 0.0043(Cm%) + 0.022(EO2%) 0.00055(SOFA)

Equation 4-3

where NOX is measured in lb/106 Btu, EO2% is percent excess oxygen in the flue gas leaving
the furnace (total basis), and SOFA is the total position of the three SOFA damper systems.
Alternatively, the equation is
NOX = 0.157 0.002(W%) + 0.009(O2%) 0.00024(SOFA)

Equation 4-4

where the units of NOX are kg/GJ.


As indicated in these equations, Albright Unit #3 was equipped with a low-NOX firing system,
including three levels of separated overfire air (SOFA). The SOFA values used to create these
equations ranged from 15% to 240%, representing a 0% to 100% opening for each SOFA level.
The mechanisms yielding NOX reduction at Albright included reducing the loading on the
pulverizers, improving their performance, creating a reducing zone in the center of the fireball,
and improving the ability to use the SOFA system. With cofiring, SOFA dampers were opened
up to 240% without increasing unburned carbon or loss on ignition (LOI) in the fly ash.
Experience from the Studstrup plant, which fires 20 tons/hr straw continuously into a 350-MW
unit, indicates few or no impacts on NOX, LOI in ash, corrosion, or heat rate. Other European
cofiring experience, such as those in the UK and Netherlands, similarly suggests little adverse
impact on boiler operations, emissions, or by-products.

4-45

Biomass Electricity Generation

0.6

NOx Emissions, lb/MMBtu

0.5

0.4

0.3

0.2

0.1

0
0.00

2.00

4.00

6.00

8.00

10.00

12.00

Cofiring Percentage, Mass Basis

Figure 4-23
NOX reduction at the Albright Generating Station

4.3.4 Cofiring Gaseous Biomass Fuels


Gaseous biomass fuels can be cofired in coal-fired boilers, natural-gas-fired boilers, or
combined-cycle power plants. In the latter case, the design with the most near-term potential
introduces biomass gas in a duct burner between the combustion turbine and the heat recovery
steam generator (HRSG). Both producer gas from biomass gasification and natural gaseous
biomass fuels can be used in these applications.
The most frequently reported example of producer gas cofiring was installed at the Kymijarvi
power plant in Lahti, Finland. The power station is a district heating plant that also generates
electricity. This coal-fired generating station has capacity to produce 167 MWe and 240 MWth
of district heat. The utility, Lahden Lampovoima Oy, owns and operates the plant. The Lahti
gasification project involved installing a Foster Wheeler atmospheric circulating fluidized bed
gasifier capable of converting sawdust, bark, wood chips, plywood trim, particleboard trim, and
recycled fuel (REF) into a producer gas for firing in the coal-fired boiler. The REF consists of
plastics, paper, cardboard, and wood. The gasifier has also blended tire-derived fuel (TDF) into
the overall feedstock.
Biomass fuels are brought to a central processing facility where they are shredded and prepared
as feedstock for the gasifier. The atmospheric CFB gasifier itself is very simple. The system
consists of a reactor where the gasification takes place, a cyclone to separate the circulating bed
material from the gas, and a return pipe for returning the circulating material to the bottom level
of the gasifier.

4-46

Biomass Electricity Generation

The hot gas temperature is 1520F to 1580F (827C to 860C), and the gas is fed directly to the
boiler to recover both the chemical energy and sensible heat from the gas. In tests, the gasifier
supplied 12% to 16% of the total heat input to the boiler. The gasifier was put on line in
September 1998 and has been operational ever since. It has an average availability approaching
88%, with its highest monthly availability slightly above 95% [47]. A similar system is used by
Electrabel at its Ruein Unit 5 to contribute 17 MWe out of 540 MWe.
Other utilities have studied similar applications, using gasifiers supplied by a variety of vendors.
Various feasibility studies have estimated the capital costs for such systems as approximately
$600/kW supported by biomass. Such capital costs include both the material handling systems
for biomass fuels, the gasifier and the hot gas piping. If gas cooling and clean-up are required,
the capital costs increase.
Use of wastewater treatment gas and landfill gas is also frequently proposed. Such gaseous
products are at low pressure and, at 500 to 650 Btu/scf, they can be attractive if the gas is used in
immediate proximity to a suitable power plant. The gas can be used in coal-fired boilers, naturalgas-fired boilers, or in combined-cycle applications. Again it is most suitable as a duct burner
fuel in a combined-cycle application. One such installation exists south of Sacramento,
California.
Wastewater treatment gas is fired in the Allen Fossil Plant, generating green power for TVA. In
this application, wastewater treatment gas from the Maxson Wastewater Treatment Plant serving
the south side of Memphis, Tennessee. Gas is introduced in separate burners above the cyclone
barrels. Although the Allen Fossil Plant approach does not make special use of the gas, landfill
gas or wastewater treatment gas could well be employed as reburn gas.
The database for these applications is insufficient to provide estimates of capital cost or
operating and maintenance costs. Quantities of gas available, specific application, piping
distances, and the degree of gas clean-up required make each case unique. It is sufficient to note
that this approach can be quite cost-effective depending upon the individual situation.
4.3.4.1 Biomass Pre-Treatments for Cofiring with Coal
There are operational challenges associated with cofiring biomass with coal that can be
addressed either by making modifications to power plants or through downstream actions, such
as replacing corroded equipment. And, to a certain extent, many of the technical challenges
associated with cofiring biomass with coal can be avoided via upstream measures, such as
biomass pre-treatment. However, the more advanced the biomass pre-treatment, the higher the
cost of the feedstock. Common pre-treatment options are sizing, compacting, and drying.
Pelletizing is also applicable in cofiring applications.
The cost of biomass fuel for cofiring with coal, however, depends not only on the procurement
cost but also on the cost of handling, storage, transportation, operability of the boiler, and end
use of the waste. Therefore, it may not necessarily be economical to use the cheapest fuel
available if the negative impact on the boiler operation or fuel operability is significant. The cost
benefit analysis along the complete cofiring chain should be considered in deciding on the
quality and price of the biomass used in cofiring.

4-47

Biomass Electricity Generation

The reasons to consider pre-treatment of biomass include:

Matching the narrow specifications of the feeding system and boiler with a wide variety of
biomass feedstocks

Reducing the costs and improve the characteristics of the biomass fuel for transportation,
storage, and handling

Reducing the costs and outage time for plant retrofits

Minimizing operation and maintenance costs using a homogeneous fuel suitable for
automatic feeding with existing plant systems

Increasing the biomass fraction in the boiler feed

Washing to remove deleterious materials from biomass is not yet commercially ready, and no
known large-scale tests are underway. Torrefaction and pelletizing is nearly commercial and is
expected to be ready for large-scale deployment by as early as 2012 if economic conditions allow.
Table 4-20 lists the pre-treatment options for addressing cofiring constraints.
Table 4-20
Pre-treatment options for addressing cofiring constraints [46]

Biomass pre-treatment via a combination of pelletization and torrefaction is an interesting option


that can mitigate problems related to high transportation and storage costs, handling, and milling
of biomass to an even greater extent than using conventional wood pellets. Compared with
untreated biomass, wood pellets have more attractive properties with regard to heating value,
grindability, storage, transport, and handling. But issues remain regarding the pellets resistance
to moisture uptake, mechanical resistance to crushing and dust formation, and, of course, cost.
4-48

Biomass Electricity Generation

Biological degradation of wood is decreased after pelletization but does still occur. Moreover,
uniformity of pellets is hard to establish, and there are substantial variations between pellets
made out of softwood, hardwood, and different tree species. Additionally, pelletization is
currently mostly limited to sawdust and cutter shavings as economical feedstock, making the
pellet market closely related to the wood processing industry.
Torrefied pellets have advantages over untreated wood pellets. Biomass is very dry after
torrefaction and its moisture uptake is very limited (1%6%). Biological degradation is almost
halted, volumetric energy density is about 30% higher than that of wood pellets, and the
torrefaction process can be used to upgrade most wood and herbaceous biomass, thus increasing
feedstock flexibility. Torrefied pellets would seem to allow high biomass cofiring in existing PC
coal boilers with minimum investment and impact on existing hardware. Additionally, due to
energy densification, torrefied pellets can decrease long distance transport costs compared with
green wood chips and wood pellets. An ongoing EPRI test will confirm the grindability and
combustibility of pelletized and torrefied biomass.
4.3.4.2 The Global Biomass Pellet Market
The wood pellet market is growing rapidlyespecially in North America, Europe, and Russia
but remains immature. To date, its availability is heavily influenced by subsidies, resource
availability, fossil fuel prices, and seasonal influences. Certification of wood pellets is not yet a
hot topic, but it is likely to become one in the years to come, once sustainability requirements for
biomass are provided in Europe and North America.
The following are the main factors driving international pellet trading over the next five years:

Increasing CO2 emissions prices

Promotion of biomass cofiring with coal in existing power plants

RPS policies

Increasing oil and natural gas prices

The primary future producers and exporters of wood pellets are expected to be Canada, the United
States, Russia, Belarus, Ukraine, and Latin America. Meanwhile, the main pellet users and/or
importers are likely to be Canada, the United States, Germany, Austria, the UK, and Ireland.
The anticipated logistical challenges that must be tackled to support the emergence of a global
biomass pellet market include the following:

Development of dedicated pellet terminals at major harbors

Commercial development of advanced treatment options (e.g., torrefied pellets)

Development of loading and unloading equipment

Prevailing high ocean long distance shipping rates

Large facilities for export or ship transport should optimize logistical costs and allow for
international trade in biomass similar to the present coal trade. Coal power plants located close to
the coasts and designed for imported coal may be in the best position to take advantage of large
shipments of pellets by barge at internationally competitive prices from such sources as Brazil
and other Latin American countries, Canada, Russia, Ukraine, among others.
4-49

Biomass Electricity Generation

4.3.5 Conclusions Regarding Biomass Cofiring


Depending upon the biomass available and the technology of choice, biomass can be cofired in
cyclone boilers, wall-fired and tangentially fired PC boilers, natural-gas boilers, and combinedcycle plants. These applications can be cost-effective depending upon the plant, the supply and
cost of the biomass, and the design employed. Cofiring is the least-cost approach to biomass
utilization in electricity generation and can be deployed relatively quickly.
Research is ongoing to address the impact of biomass cofiring on the life of deNOX catalysts and
on fireside corrosion. In both cases, the alkali/alkaline earth constituents prevalent in the biomass
may adversely impact catalyst poisoning or corrosion reactions.

4.4 Stand-Alone Biomass-Fired Systems


Stand-alone biomass-fired systems designed to produce electricity include the following:

Stand-alone wood or wood-waste-fired generating stations producing electricity as an


exclusive product or producing both electricity and process steam in cogeneration or
combined-heat-and-power (CHP) applications.

Spent pulping liquor (e.g., black liquor from Kraft pulping) combustion systems designed to
recover pulping chemicals while producing steam for both electricity generation and process
heat applications.

Bagasse-fired boilers used to generate electricity in the sugar industry.

Other agribusiness systems with biomass either fired directly in a boiler or gasified and gas
fired in a waste heat boiler with steam subsequently used for electricity generation.

Boilers fired with landfill or wastewater treatment gas, designed to produce electricity.

Combustion turbines fired with landfill or wastewater treatment gas.

Internal combustion engines, typically diesel engines, fired with landfill gas, wastewater
treatment gas, or producer gas from woody or herbaceous biomass gasification.

Of these, wood-waste-fired boilers and chemical recovery boilers are the dominant sources of
biomass electricity today. Black liquor from Kraft pulp mills is used to produce more electricity
than any other form of biomass, followed closely by wood waste.
Chemical recovery boilers and associated systems are specific to the pulp and paper industry,
and are not considered further here. The wood-waste-fired boiler is the focus of this discussion,
as it provides the most promise for electricity generation.
4.4.1 Wood-Waste-Fired Boilers
Wood-waste-fired boilers, also known as hog fuel boilers, have long been used in the forest
products and pulp and paper industries to generate process steam and cogenerate both electricity
and process steam. Major integrated forest products mills producing a product mix including
Kraft pulp, plywood, and dimension lumber commonly installed such boilers as power boilers.
Examples abound, including the Longview #11 boiler at the Weyerhaeuser complex in
Longview, Washington and comparable power boilers at Weyerhaeuser facilities in Columbus,
Mississippi, Plymouth, North Carolina, and New Bern, North Carolina. Georgia Pacific,
International Paper, and others have similar installations.
4-50

Biomass Electricity Generation

Electric utilities and independent power producers also constructed wood-waste-fired boilers to
generate electricity as their sole product. Large units have been constructed by Washington
Water Power at Kettle Falls, Washington; Burlington Electric at the McNeil Generating Station
in Burlington, Vermont; and at other locations. Independent power producers built similar
systems to fire wood waste and wood-like biomass fuels. Large numbers of such systems were
built in California, Maine, and elsewhere in response to PURPA, the need for capacity in certain
areas of the country, and localized availability of materials. The Shasta Generating Station in
Anderson, California is a prime example of such facilities.
4.4.1.1 Typical Capacities of Wood-Waste-Fired Plants
While coal is produced at a central mine location and distributed to users, wood waste is
produced as a distributed resource over a large area and must be collected or gathered and
shipped to the user location. This fundamental distinction highlights one capacity limitation
impacting biomass systems. The second limitation results from the typical fuel characteristics:

15 to 20 lb/ft3 bulk density

40% to 50% moisture

8 106 to 10 106 Btu/ton as-received

These factors combine to limit boiler and generating capacity. Given typical transportation
distances for wood fuel of up to 50 miles (80 km), wood-fired boilers have been limited to a
nominal 100 to 125 ton/hr firing rate (nominally 10,000 ft3/hr of fuel), or 50 to 70 MWe
depending on system design and operation. This capacity limitation carries with it the following
significant implications:

It is difficult to economically justify more than three turbine extractions for feedwater
heaters; the three extractions are typically for a low-pressure heater, the deaerater, and a
high-pressure heater. In some cases only one extraction (deaerater) is justified.

Reheat cycles that significantly improve cycle efficiencies are not economical at the small
plants available for wood firing.

Depending on staffing philosophies, such units will achieve ratios of 0.5 to 2.0 MWe per
worker, depending upon unit capacity and staffing approach. This contrasts with
conventional coal-fired staffing ratios of 3 to 6 MWe/worker.

Typical stand-alone wood-fired power plants have been built at capacities ranging from 11 MWe
to 70 MWe, and the larger plants were most successful.
4.4.1.2 Wood-Waste-Fired System Design
Numerous wood-waste-fired systems have been developed for the current market. In order to
evaluate system design, a generic plant is considered. It represents standard practices in woodwaste-fired technology. Variations on this approach are also discussed below.

4-51

Biomass Electricity Generation

4.4.1.2.1 Fuel Handling

Figure 4-24 shows the standard fuel handling approach. Loaded trucks of wood waste are
received and unloaded by hydraulic unloaders that raise the truck sufficiently to dump the load
into a hopper and feed it to a disc screen. Oversized particles are hogged using hammer hogs or
(rarely) knife hogs. The entire fuel pile is then stocked out. Wood waste or hog fuel is then
reclaimed and transported to the plant, where it is discharged into bunkers.

F1

X-110
Truck Unloader/Scale

F-111
Receiving Hopper/Magnet
H-112
Rotary Disc Screen

F2

F3
J-114
RDS Conveyor

F-116 RDS Pile

C-113
Hammer Mill
F4

F1
J-115
HM Conveyor

F-117 HM Pile

X-120 Dozer

F-122 Fuel Storage Pile

F-132
Feed Bin

F1
F-130
Reclaim Hopper/Pit

X-121 Dozer

F6

J-133
Feed Conveyor
B

F5
J-131
Reclaim Conveyor/Scale

F-134 Excess Pile

Figure 4-24
Representative wood fuel handling system

Because waste wood truckloads typically contain 20 to 25 tons, large plants such as McNeil
Generating Station or Kettle Falls Generating Station require a significant commitment to wood
fuel handling. A 50-MWe plant burning 80 to 100 tons/hr requires four to five trucks per hour, 24
hours per day, seven days per week. Burlington Electric chose another approach to keep truck
traffic out of the plant neighborhood. It receives wood waste off site, load it into open hopper
railroad cars, and transport the biomass to the plant. The wood is discharged at a trestle facility.
This has plant benefits, but increases the cost of materials handling.
Weyerhaeuser at Longview, Washington, uses a wood fuel stacker-reclaimer arrangement
instead of bulldozer reclaiming. This helps manage the 100+ tons/hr burned in its #11 boiler.
Numerous studies have addressed the addition of drying systems to the materials handling as a
means for improving boiler efficiency. These are not generally employed. Dryers require energy
to operate, consequently reducing the efficiency gain. Further, drying technologies introduce the
potential for fires, which have occurred at installations that use fuel dryers. Consequently, the
general approach is to use the boiler as the dryer unless site-specific circumstances favor use of a
separate dryer.
4-52

Biomass Electricity Generation

4.4.1.2.2 Boiler and Turbine-Generator Technology

The traditional wood-fired boiler used in electricity generation is a spreader-stoker. Figure 4-25
shows a typical flowsheet for a stoker boiler. Wood is discharged from the surge bin via a twin
counter-rotating auger system and fed to a stoker. Unlike the typical coal stoker, which is a
paddlewheel device, a wood fuel stoker is an air-swept design that throws the wood fuel onto the
grate. Figure 4-26 depicts a typical boiler and turbine-generator system, along with feedwater
heaters, pumps, and fans.

Mechanical
draft cooling
tower
Cooling
water
Turbine/
generator

Moisture losses
Wood
blend

Electricity

Steam (1250 psi, 950 deg F)


Ammonia
710 F

On-site
wood
preparation

308 F
Stoker boiler

Air heater
638 F

Fines
Ferrous

Bottom ash

Fabric filter
80 F
Combustion
air

Stack
Fly ash

Figure 4-25
Schematic of biomass-fired stoker boiler power plant

The plan area heat release for wood-waste-fired boilers is typically on the order of 750 103 to
1 106 Btu/ft2-hr. Most coal-fired stokers are in the 500 103 to 750 103 Btu/ft2-hr range. The
volatility of the wood fuel allows a higher area heat release rate. The furnace volumetric heat
release for wood waste is typically 13,000 to 20,000 Btu/ft3/hr, compared to coal at 20,000 to
25,000 Btu/ft3/hr. The lower calorific value of the wood fuel and higher moisture content reduces
the volumetric heat release. Evaporation of fuel moisture to a vapor can require significant
furnace volume.

4-53

Biomass Electricity Generation


Q-210
Stoker Furnace

5
900
967.4

N-300
STM Turbine

N-310
25 MW Generator

4
275
1105

6A

G-226
Ejector

700
500

7A
X-215
NH4 Grid

400
100

8A

E-217
Economizer

210
10

ATM

125.5
2

E-224
Condenser

250
1110

10A

10B
100
25

80
30

NH4

CA

CA

E1
ATM
E-218
Air Heater

X-211
Oil Burner (4X)

7A
400
100
200
90

WASTE

CA

CA

D-223
Deaerator

6A

8A
210
10

700
500

A2

13
175
140

200

CA
CA

E-221
FW Heater

CA

X-212
Air Swept Burner (4X)
ATM

X-213
Air Cooled Grate

ATM

G-216
Blower

8B
130
4

11

59

H-219
Air Filter

125.5
1.96

12

ATM

200
1160

G-220
FD Fan

J-214
Bottom Ash Screw Conveyor

E-225
FW Heater

6B
300
490

A1

BBD

1
L-222
FW Pump

200
90

DMP

125.5
145

G-227
C Pump

WASTE

Figure 4-26
Typical boiler and turbine-generator approach

Grate selection is of consequence. The generic design presented here uses a water-cooled grate
design such as the Detroit Stoker Hydrograte. Alternatives include air-cooled pinhole grates,
where primary or undergrate air not only provides oxidant to fuel on the grate, but also cools the
grate bars themselves.
Steam conditions are a function of the design capacity of the boiler. For lower-capacity boilers
(e.g., 250,000 lb/h steam), 600 psig/750F conditions are common. For medium-capacity
boilers, the steam conditions are often 850 psig/825F, and higher capacity units typically use
1250 psig/950F and 1450 psig/1000F steam conditions. The larger units use more feedwater
heaters to improve the thermal efficiency of the unit.
Alternative circulating and bubbling fluidized-bed boilers and gasifiers coupled to waste heat
boilers may also be used. Figure 4-27 shows a typical fluidized bed combustion power plant
flowsheet. Selection of a steam generator depends upon the type and quality of biomass available
as well as the site and project-specific economics. For example, the Tacoma Steam Plant #2 was
repowered using two bubbling bed steam generators coupled to the existing boilers, used as
waste heat boilers. Tacoma Public Utilities fueled the units with a mixture of wood waste,
refuse-derived fuel (RDF), and coal.

4-54

Biomass Electricity Generation

Mechanical
draft cooling
tower
Cooling
water
Turbine/
generator

Electricity

Steam (1250 psi, 950 deg F)


Moisture losses
Wood
blend

710 F
On-site
wood
preparation

Fluidized
bed
boiler

312 F
Air heater
642 F

Fines
Ferrous

Sand

Bottom ash

Fabric filter
80 F
Combustion
air

Stack
Fly ash

Figure 4-27
Schematic of biomass-fired atmospheric fluidized boiler power plant

Turbines are selected based on unit capacity and whether the system is designed as a stand-alone
power-only plant, or whether it is a cogeneration or CHP plant. Steam turbines can be designed
with automatic extractions for process steam or as backpressure turbines exhausting process
steam at 50 psig, 150 psig, or other conditions depending upon the plant requirements.
Alternatively, the turbines can be designed and supplied to exhaust steam at 2 to 3 in HgA if
power is the exclusive product. Typical water rates (lb steam/kWh) for Rankine-cycle turbines
used in wood-waste-fired power plants are shown in Tables 4-21 and 4-22. Table 4-21 presents
water rates for condensing turbines, whereas Table 4-22 presents water rates for backpressure
cogeneration turbines.
Table 4-21
Typical water rates for condensing turbines in wood waste-fired plants
Turbine Throttle Pressure

Water Rate

Psig/F

Atm/C

lb steam/kWh

kg steam/kWh

600/750

41/400

9.8

4.4

850/825

58/440

8.7

3.9

1250/950

85/510

8.3

3.8

1450/1000

102/540

8.0

3.6

Note: Assumes 3-inch HgA turbine exhaust.


Source: Tillman, Rossi, and Kitto, 1981.

4-55

Biomass Electricity Generation


Table 4-22
Typical water rates for backpressure turbines in wood waste-fired plants
Turbine Throttle Pressure

Water Rate

TurbineGenerator Cap.

Psig/F

Atm/C

lb steam/kWh

kg steam/kWh

15
15
15
20
20
20
25
25
25

600/750
850/825
1250/950
600/750
850/825
1250/950
600/750
850/825
1250/950

41/400
58/440
85/510
41/400
58/440
85/510
41/400
58/440
85/510

31.6
23.6
18.4
30.9
23.4
18.1
30.6
23.2
17.9

14.3
10.7
8.3
14.0
10.6
8.2
13.9
10.5
8.1

Note: Assumes 10.2 Atm or 150 psig exhaust pressure.


Source: Tillman, Rossi, and Kitto, 1981.

Given the typical boiler efficiencies of well-run wood-waste-fired boilers, the net station heat
rates (NSHR) for small condensing wood-fired plants are typically in the 15,000 to 17,000
Btu/kWh range, whereas the NSHR values for the 40+ MWe units are typically in the 13,000 to
14,000 Btu/kWh range. NSHR values depend on the capital investments made in such equipment
as economizer and air heater surface, feedwater heaters, and like equipment. Balance of plant
becomes a critical concern for these systems. For more severe steam conditions, feedwater
treatment requirements become more substantial.
The incremental net station heat rate for cofiring systems is far more attractive. For
backpressure turbines with throttle steam conditions of 1250 psig/950F and exhaust conditions
of 50 or 150 psig, the NSHR appropriate for the power generation element is on the order of
4,500 to 5,000 Btu/kWh. This is based upon the fact that many process uses of steam capitalize
on the heat of vaporization. Consequently, losses associated with the heat of vaporization are
ascribed to the process user. The losses attributed to cogeneration are related to boiler efficiency
(or boiler thermal losses) and turbine efficiency (or turbine losses). With cogeneration systems
based upon extraction steam, NSHR values can be on the order of 7,500 Btu/kWh or more,
depending on the proportion of steam used in cogeneration applications and the proportion of
steam condensed in power generation application.
4.4.1.2.3 Pollution Control Systems

Air pollution equipment for solid biomass-fired systems includes either fabric filters (FF) or
electrostatic precipitators (ESP), and either selective catalytic reduction (SCR) or selective noncatalytic reduction (SNCR) systems. Acid gas scrubbers are not required due to the compositions
of typical solid biomass fuels. For this analysis, an ESP was assumed, with multi-clones installed
ahead of the ESP. Multi-clones are useful in scavenging stray glowing embers should they leave
the economizer and air heater sections of the steam generator hot. A SNCR system was also
assumed, as firing wood waste typically generates less NOX than firing coal in larger utility boilers.
Figure 4-28 depicts the air pollution control system along with a mechanical cooling tower and a
demineralizer system for feedwater treatment.
4-56

Biomass Electricity Generation


Air Quality Control System
E1

E2
E1

Cooling Tower

210
14.7

230
14.7

ATM

BBD

H-419

C2

J-421

100
25

G-423

J-422

X-424

K-425

C1

C1

WASTE

80
30

AQCS - SNCR System

C3
L-416

L-418

Demineralizer Plant
N3

NH4
F-411
N1

P-417

C2

H-420

N4

H-412
N2

L-410

N6

C3

NH4

M-415

L-413

RW

DW
RW

N5
L-414
P-419

Figure 4-28
Air pollution control system, cooling tower, and demineralizer for a typical wood-wastefired plant

4.4.1.3 Comparisons to Other Biomass-Fired Rankine-Cycle Designs


Boilers comparable to the wood-waste-fired boilers described earlier have been used for
agricultural wastes. Such wastes include nut hulls and shells, stone fruit pits, vineyard prunings,
orchard prunings, rice hulls, corn stover, spent corn cobs, and a host of other materials. Wheat
straws and comparable materials have been fired in stoker and fluidized-bed boilers in both the
United States and Europe. All these materials can be fundamentally different from wood waste
as follows:

Higher nitrogen content than wood waste

Higher chlorine content than wood waste

Higher inorganic matter than wood waste with the wood waste containing significantly
higher concentrations of calcium, potassium, sodium, and phosphorus

Higher bulk densities than the biomass fuel, with the exception of stone-fruit pits
(e.g., peach pits)

Designers of boilers firing such materials must take into account the significantly higher slagging
and fouling propensities of these materials when compared to typical wood waste (see Baxter et
al. [23, 24]). Experience in the Central Valley of California and elsewhere has shown that the
combination of ash chemistries and chlorine has increased ash deposition in the boiler. Some of
these biomass materials have been sufficiently different in characteristics that special systems
have been used. Rice hulls, for example, have been suspension-fired. They have also been
gasified and the product gas fired in a waste heat boiler for steam generation and subsequent
power generation.
4-57

Biomass Electricity Generation

Rankine-cycle generation has also been used in firing landfill gas and wastewater treatment gas.
For example, the 10 million ton/yr Puente Hills landfill has a 50-MWe Rankine-cycle generating
station. Similar boilers are installed at the Spadra and Coyote Canyon landfills, also in the Los
Angeles County area.
Primary differences include fuel handling and pollution control. The gaseous fuel is produced at
low pressure and saturated with water vapor. Further, it can contain solid contaminants and some
chlorinated compounds. Gas treatment for such biogas-burning boilers may include gas
dehydration, particulate scavenging, and other gas treatments depending upon the design
principles used. Flue gas treatment focuses more on NOX removal than particulate capture.
Operating and maintenance costs for wood-waste-fired generation are typically site-specific,
depending on the owner, the location, and its proximity to related facilities. For example, a
project located on the site of an integrated forest products mill can share certain maintenance
facilities and costs. An independent plant must bear the entire cost of operations and
maintenance.
Typically, a wood-fired plant requires about 20 to 30 employees, including a plant manager,
engineer, business manager, secretary, four shifts with three employees/shift, two fuel yard
employees, and a maintenance crew including both millwrights and electricians. Nominally, a
crew of about 25 is adequate for plant operations. However, larger stations (e.g., 40+ MWe)
generally are more heavily staffed to handle increased fuel receipts and the increased complexity
of the unit. Alternatively, outsourcing can be used to manage the staffing complement. For
smaller generating stations up to 15 MWe capacity, staffing is about 0.5 MWe capacity/worker.
Larger units may achieve better ratios such as 1 to 2 MWe capacity/worker.
Maintenance requirements are comparable to those associated with coal-fired generating stations;
the same types of problems occur. Annual maintenance costs can be estimated at 4% to 5% of
the total capital cost of the facility. However, this is a generalized number, and site-specific
maintenance costs vary significantly.
4.4.2 Repowering Existing Coal Units for Biomass Firing
The Schiller station in Portsmouth, New Hampshire is the site of a novel repowering of an
existing coal unit to wood firing. The station consisted of three 50-MW coal-fired units. PSNH
and its subsidiary, Northern Wood Project, opted to decommission one of the boilers and replace
it with a dual-fuel fluid bed boiler. While the unit is designed for both coal and wood, it is
anticipated that only wood will be fired. The new boiler was integrated with the existing steam
circuit, which kept costs down. Additionally, a wood receiving yard was necessary to
accommodate truck delivery of wood. The unit came on line in December 2006 and early reports
suggest the unit is performing up to expectations. The cost of repowering was about $1,400/kW.
Repowering smaller and older coal units may offer significant advantages to utilities. It allows
them to utilize existing sites and local established infrastructure. The wood retrofits would
support the local forestry industry, contribute to the utilitys renewable portfolio, and in general
reduce the site emissions profile.

4-58

Biomass Electricity Generation

4.4.3 Ownership Models and Institutional Issues


Traditionally, biomass-fired generating stations are owned either by a utility (e.g., Burlington
Electric and Washington Water Power), by a non-utility generator (NUG) such as a forest
products company, or by an independent power producer (IPP). IPPs have installed more than
1,000 MWe of wood-waste-fired capacity in California in response to PURPA and the California
Standard Offer #4 (SO4). As SO4 expired, construction of new facilities halted and several
smaller units were shut down.
An alternative ownership model has been used by selected utilities located where large forest
industry and pulp and paper complexes exist. In this situation, a utility owns a turbine-generator
within the site of a forest industry mill and its cogeneration installation. The utility purchases
high-pressure steam from the boiler and sells low-pressure process steam back to the pulp mill or
integrated complex. The utility has a lower capital investment and a lower net station heat rate.
The turbine is dispatched by the forest industry, consistent with its needs. Such ownership
arrangements are niche applications that are mutually beneficial.
Other institutional issues include regulatory constraints and incentives. There is currently
considerable uncertainty in the regulatory and incentive arenas, which makes definitive
observations difficult if not impossible to make. There are some who would, for political
reasons, consider wood waste not renewable. Others argue whether wood waste in various
forms is green or clean power. Regulatory constraints are equally uncertain.
4.4.4 Alternative Systems for Gaseous Biomass Fuels
Alternative power generation systems for gaseous biomass fuels such as landfill gas and
wastewater treatment gas include internal combustion engines, microturbines, and larger (e.g.,
5000 kW) combustion turbines. Wastewater treatment gas includes the product of municipal
wastewater treatment plants as well as anaerobic digestion of animal manures. Traditionally,
wastewater treatment plants have used internal combustion engines if appropriate for power
generation. These systems can be installed in 1-MWe increments, although both smaller and
larger systems are employed for both landfill gas and wastewater treatment gas. Such
installations require gas compression and are frequently backed up with either natural gas or
propane.
More recently, fuel cells and microturbines have been applied to these gaseous biomass fuels on
a somewhat experimental basis. Small 1- to 5-MWe combustion turbines have also been
installed. The Puente Hills landfill pioneered the use of small combustion turbines for electricity
generation. For all systems, attention must be paid to concentrations of hydrogen sulfide (H2S) at
100 to 2,000 ppmv, particularly in wastewater treatment/digester gas. Siloxanes also are a
contaminant with significant potential for causing combustion and environmental issues.
The capital costs of these alternatives have been documented by CH2M Hill and others
(see Kitto and Green [32]):

Internal combustion engine: $3,000/kW (1-MWe system basis)

Fuel cell: $9,000/kW (300-kW system basis)

Microturbine: $4,000/kW (30-kW system basis)

Combustion turbine: $1,000/kW (5-MWe system basis).


4-59

Biomass Electricity Generation

The combustion turbine, in particular, is penalized (relative to natural gas turbines) by the small
scale of the equipment and by the need for gas cleaning and compression.
Operating and maintenance costs for these systems can be both site specific and technology
specific.
4.4.5 Near-Commercial Technologies
Hundreds of biomass gasifiers are deployed worldwide (there are at least 50 suppliers), but
almost all of them are unsuitable for central station type utility power applications. There are
three primary reasons for this. First, almost all of them are too small for utility application. Most
are rated at less than 1 MW thermal, and would consequently have high operating costs. Second,
the gas quality of many of the small gasifiers is quite low and is suitable for direct combustion as
required for process heat or district heating or, in some cases, running small reciprocating
engines with associated high maintenance costs. And finally, while small gasifiers have been
shown to be robust operationally, scale-up to utility sizes (approaching 10 MW and greater) is
problematic.
Demonstrations of larger-scale gasification technology coupled with district heating are
underway throughout Europe and show promise both technically and economically. These are
demonstrations of gasifiers that are either already at utility scale, or could easily be scaled to
larger scale. Larger configurations reduce the operating cost by spreading fixed costs (i.e.,
operations staff) over a larger output. Economies of scale also reduce variable operating unit
capital cost.
4.4.5.1 Types of Gasifiers
Depending on the physical characteristics of the biomass and the type of gasifier, the feedstock
may require pre-processing so that it can be fed into the gasifier. This may include shredding,
pelletizing, screening, and/or drying. The feedstock is then fed into the gasifier. A small portion
of the feedstock reacts with air, oxygen, and/or steam that are injected into the process, and some
low-level partial oxidation reactions occur. This provides heat to drive off the volatile matter in
the feedstock, and initiates the gasification reactions, forming syngas.
Once the gasification reactions are initiated, sufficient heat is generated to maintain the process.
The operating temperature is a function of the feedstock heating value, the configuration of the
gasifier, the amount of air or oxygen added, and the amount of steam injected. In low
temperature gasification, the inorganic ash is removed as a clinker-like bottom ash. In high
temperature gasification, the gasifier operates above the melting points of the inorganic
substances in the ash, forming a molten slag. This slag is tapped from the bottom of the gasifier
into a water bath, where it solidifies into a glassy, inert material.
The types of gasifiers typically used with biomass feedstock are described below and shown in
Figure 4-29.

Moving or fixed bedthe feedstock enters from the top of the reactor, falling onto a grate
(which may be stationary or mechanically moved or agitated to promote contact between the
feedstock and oxidant, and to allow ash to exit). Steam, oxygen or air is injected into the
bottom (updraft) or side/top (downdraft). The syngas exits the reactor. Temperatures are low
to moderate, and ash is removed in solid form.

4-60

Biomass Electricity Generation

Fluid bedthe feedstock enters from the side or bottom of the gasifier, with sufficient air or
oxygen to fluidize a blend of feedstock and a fluidizing agent, typically sand. This provides
turbulence and efficient mixing for thorough gasification, and is advantage for feedstocks
which have low reactivity or are difficult to gasify. Steam may be added into the bottom.
Syngas exits at the top of the gasifier, and typically enters a cyclone for removal of ash,
unreacted char, and any fluidizing sand that are carried over in the syngas stream. The
char/ash/sand mixture may be disposed of, or returned to the gasifier feed to enhance carbon
conversion and to recover fluidizing sand.

Figure 4-29
Gasifier types for biomass (EPRI 1023994)

Selection of the type of gasifier depends on feedstock, ash characteristics, and the type of syngas
desired. For example, the lower temperature of fixed-bed gasifiers tends to form more methane
in the syngas. High methane content is not acceptable for some downstream uses, i.e. production
of liquid fuels and chemicals.
Fixed Bed Gasification

Fixed-bed gasifiers can be classified primarily as updraft and downdraft. Updraft gasifiers
(Figure 4-30) represent the oldest and simplest gasifiers. The updraft gasifier is a counterflow
reactor in which feedstock is typically introduced into the top by means of a lockhopper or rotary
valve and flows downward through the gasifier to a grate where ash is removed. The gasifying
medium, air or oxygen (and possibly steam), is introduced below the grate and flows upward
through the reactor. At the bottom of the reactor (combustion zone) char that remains from
gasification of the feedstock burns to form carbon dioxide (CO2) and steam (H2O), which then
flow upward through the bed countercurrent to the downflowing feedstock material. The partial
oxidation reactions supply the heat needed to drive gasification, pyrolysis and drying. The
maximum temperature in the combustion zone is typically higher than 2,200F. In the reduction
zone, the CO2 and H2O are partially reduced to carbon monoxide (CO) and hydrogen (H2)
through reaction with carbon in the char. In the pyrolysis zone, these gases contact dry biomass
and devolatilize it to produce pyrolysis products (gases, liquids and tars) and residual char.
Above this zone, the gases dry the wet incoming biomass. A wide range of tars and oils, which
4-61

Biomass Electricity Generation

can condense in product lines, are produced in the pyrolysis zone. For this reason updraft
gasifiers are usually operated in a close-coupled mode to a furnace or boiler to produce steam or
hot water. Certain feeds with low-melting ash may have slagging on the combustion grate. In
addition, feedstock particle size needs to be controlled to maintain a uniform bed.

Figure 4-30
Updraft gasifier (EPRI 1023994)

In downdraft gasifiers (Figure 4-31), both the oxidant and syngas flow in the same direction as
the solid biomass. Downdraft gasifiers are specifically designed to minimize tar and oil
production. The biomass and the gases move co-currently downward through the bed. The
pyrolysis products pass through the hot char combustion zone, where they are contacted with air
and the tars are thermally cracked and partially oxidized. This provides a filtering effect to help
to clean the syngas prior to exiting the gasifier. Typical tar conversion is greater than 99%, and is
a function of temperature, combustion efficiency and channeling. The combustion zone
temperature is typically 1,500-2,200F. The hot char in the reduction zone reduces CO2 and H2O
to CO and H2, which are the primary components of syngas. The exit syngas temperature is
typically >1,000F.
Downdraft gasifiers have the same general constraints on feed properties as updraft gasifiers.
The biomass needs to have a fairly uniform particle size distribution with few fines to maintain
bed physical properties and minimize channeling. The biomass needs to have low ash with a high
fusion temperature to prevent slagging. In addition, the biomass moisture content needs to be less
than about 20% to maintain the high temperatures required for tar cracking (too high a moisture
content will cool the syngas).

4-62

Biomass Electricity Generation

Figure 4-31
Downdraft gasifier (EPRI 1023994)

Fluid Bed Gasification

In a fluid bed gasifier, as shown in Figure 4-32, the gas velocity is high enough that the biomass
particles are widely separated and circulate freely throughout the bed. During overall circulation
of the bed, streams of gas flow upward in channels containing few solids, and clumps or masses
of solids flow downward. The fluidized bed looks like a boiling liquid and has the physical
properties of a fluid. For biomass gasification, the fluidizing material is air, oxygen, or steam,
and the bed is usually sand, limestone, dolomite, or alumina. An air inlet distribution manifold or
series of sparge tubes or bubble caps are used to fluidize the bed. Biomass is introduced either
through a feed chute at the top of the bed or through an auger or similar feed system into the bed.
In-bed introduction provides residence time for fines that would otherwise be entrained in the
fluidizing gas and not converted in the bed. As the syngas exits the gasifier, a cyclone is used to
return unreacted biomass, char, and the fluidizing sand to the bed.
The biomass is introduced into the bed to raise the bed temperature to the desired operating
range, typically 1,100 to 1,600F. Bed temperature is governed by the desire to obtain complete
devolatilization of the biomass versus the need to maintain the bed temperature below the ash
fusion temperature of the biomass ash. As biomass is introduced into the bed, most of the
organics vaporize and are partially oxidized in the bed. The exothermic combustion provides the
heat to maintain the bed at temperature and to volatilize additional biomass. Fluidized bed
gasifiers have the advantage of extremely good mixing and high heat transfer, resulting in very
uniform bed conditions. Gasification is very efficient, and 95% to 99% carbon conversion is
typical.

4-63

Biomass Electricity Generation

Figure 4-32
Fluid bed gasifier (EPRI 1023994)

There are two general types of fluidized bed gasifiers: bubbling fluid bed (BFB) and circulating
fluid bed (CFB). Both systems use a sand bed that is maintained in the fluid state by air addition.
Syngas carries some char and sand up into the freeboard area where some solid disengagement
occurs (solids fall back to the sand bed) and then passes on to the cyclone where a very high
fraction of remaining particles are separated and flow by gravity from the bottom of the cyclone
to the sand bed.
The main difference between the two designs is the degree of fluidization that is used. As the
word bubbling implies, the sand bed in the BFB is churning in a relatively gentle fashion; the
sand moves in fashion similar to boiling liquid.
The CFB uses a much more active bed with much more of the interior of the gasifier filled with a
blend of sygas, feedstock and fluidizing sand. As the gas exits the gasifier, a significant amount
of bed material is carried forward to the cyclone for separation. A relatively large stream of sand
(with unreacted char) thus returns to the bed via an intended circulation path
Circulating Pyrolysis

Pyrolysis utilizes externally provided heat in a closed reactor, without the addition of air or
oxygen, to thermally decompose the biomass and drive off the volatile components as syngas.
This typically occurs at temperatures of 400 to 1,400F. Because there is no oxidant added, the
fixed carbon portion of the biomass exits the reactor as a carbon char mixed with the ash from
the biomass. This char/ash mixture is typically disposed of, or, depending on its physical and
chemical characteristics, used as a soil amendment.
Pyrolysis is typically not as efficient as gasification, as heat must be added externally (typically
in the form of combustion of natural gas or even syngas), and a substantial portion of the
biomass can exit as unreacted char. A modification of basic pyrolysis technology, circulating
pyrolysis, directly addresses these issues. Circulating pyrolysis effectively uses a pair of fluid
beds to perform a combination of pyrolysis and generation of the heat needed to initiate the
4-64

Biomass Electricity Generation

pyrolysis reactions by combusting the unreacted char (see Figure 4-33). Biomass is introduced
into the pyrolysis bed, which contacts the biomass with hot sand to generate syngas via
pyrolysis. Steam is typically used to fluidize the pyrolysis bed. The solids (sand and char
primarily) exit overhead and are removed in a cyclone. Separated solids (sand and char) are
moved to the second reactor vessel where oxygen (air can be used) is introduced to combust the
char and heat the fluidizing sand. The hot flue gas exits overhead, carrying sand to another
cyclone; flue gas flows to a stack (or may be used in a heat recovery steam generator to generate
steam for use in the plant or for power generation) and the separated hot sand flows back to the
sand bed in the pyrolysis vessel to provide the external heat required for pyrolysis.
Since no air is added in pyrolysis, the syngas is not diluted with nitrogen from the air and a very
high quality syngas with low tar content is produced. Because steam is used in the pyrolysis
reactor, the syngas contains a relatively high concentration of hydrogen.
Syngas ExhaustGas

Pyrolysis
Reactor

Cyclones

Char
Combustor

Feedstock
Sand/Char
Mixture
Fluidizing
Sand

Fluidizing
Steam

CombustionAir

Figure 4-33
Circulating pyrolysis (EPRI 1023994)

Oxidants: Air, Oxygen and Steam

Gasification requires a form of oxygen to enable the conversion of the organic portions of the
biomass feedstock to syngas. Oxygen (O2) is used to reduce the solid or liquid feedstock to
syngas containing primarily hydrogen (H2) and carbon monoxide (CO), which carry much of the
original feedstocks heat content to a downstream system which can efficiently convert it to
power (via combustion), fuels, or chemicals. Use of oxygen, sometimes in combination with
steam, results in the maximum H2 and CO content in the syngas. Oxygen-blown gasification
typically produces medium-Btu syngas.
Air can also be used since it contains about 21% O2. However the main constituent of air is
nitrogen (N2) at about 79%. Introduction of air into a gasifier results in the O2 being consumed to
form CO. N2 does not react, but dilutes the syngas, resulting in lower heat content per volume.
4-65

Biomass Electricity Generation

Air-blown gasification typically produces low-Btu syngas. The downstream systems thus must
be larger to handle the larger volume of syngas which could contain the same amount of heat as a
much smaller volume of O2-derived syngas. If power generation via combustion of the syngas in
a boiler or thermal oxidizer, or possibly in a reciprocating engine-generator, is intended, syngas
from an air-blown gasifier can be readily handled. However, if a gas turbine will be utilized, the
higher-Btu, lower tar, lower volume syngas from an O2-blown gasifier may be required.
The primary effect of use of O2 is the cost of producing or obtaining the O2. Purchasing O2 from
a third party can be expensive; some savings may be possible by installing and air separation unit
(ASU) as part of the gasification facility, but utility costs, particularly power to drive the
compression systems involved, must be considered. Significant amounts of N2 result from the O2
separation, which can be advantageous if a separate use is available for high purity N2. ASUs are
complex systems to operate and maintain; this may be simplified by purchasing the O2 acrossthe-fence from a supplier that specializes in the operation of ASUs.
When air is used, it is typically supplied via a blower, making it a much lower cost oxidant
4.4.5.2 Syngas Quality for Downstream Use
As noted above, the major components of syngas are H2, CO, CO2, and water vapor. For airblown gasifiers, approximately half of the syngas (by volume) will be N2, which dilutes the
syngas. In addition to these components, syngas produced from biomass gasification will also
contain the following:

Ash

Char

Tars

Vaporized and/or condensed metallic compounds

When the syngas is simply combusted in a thermal oxidizer or boiler to produce steam or for use
in a steam turbine-generator for power generation, little or no syngas cleanup may be required. In
those cases, post-combustion emission control systems are typically used to capture/remove
constituents prior to discharge from the plant stack. Pre-combustion syngas cleanup typically
includes the use of cyclones to remove char and/or ash from the syngas stream. This helps to
protect downstream heat transfer surfaces from plugging, fouling, and corrosion.
When the syngas will be directly combusted in a reciprocating engine or gas turbine, a much
higher level of syngas quality is required. Tars, ash, char and condensed metallic compounds can
quickly damage (through erosion and/or corrosion) the specialized materials and parts in rotating
equipment. Tars can condense onto syngas piping, as well as in control valves in the path to the
engine or gas turbine. As temperatures and pressures change in the syngas line, tars have been
known to form crystalline organic compounds (such as naphthalenes) in syngas filters, control
valves, and then in the engine combustion chamber. Vaporous metallic compounds (i.e., sodium
and potassium compounds that enter with the biomass) can also condense in combustors or on
pistons. Another example of this is when the biomass feedstock contains silica-based organic
compounds, which upon combustion, form silicon dioxide (sand), which quickly erodes pistons
and cylinders.
4-66

Biomass Electricity Generation

In order to protect engines and gas turbines from these contaminants, a range of syngas cleaning
systems may be utilized. The choice depends on the syngas contaminants and the type of
downstream combustion device, as follows:

One or two stages of cyclones for removal of ash and char

Fine filters to capture tars

Catalytic or high temperature systems to crack (decompose) the tars into the simple syngas
constituents

Baghouses to capture ash, char, and condensed tars

Wet electrostatic precipitators to capture fine particulates

Activated carbon beds to remove metallic compounds

Quenching systems to cool the syngas and wash out ash, char, and tars (moisture must then
be removed)

The following section briefly describes some key demonstrations of atmospheric gasification in
Europe, and their role in power production. With the exception of the gasifier at Ruien, the
projects highlighted below (and several more) are first-of-a-kind demonstrations. Consequently,
cost and performance information is difficult to acquire and interpret. Other demonstrations, such
as the biomass gasification/combined-cycle demonstration at Vrnamo, have been completed and
provide a reasonable technical backdrop, but little cost information. At this time, there are no
active biomass gasification power demonstrations in the United States.
4.4.5.3 Wood Gasification for Power and District Heating in Gssing, Austria
The Gssing Gasifier is a 2-MWe, 4.5 MWth, circulating fluid bed gasifier. Wood harvested from
local forests is the fuel. The village of Gssing has 56 km of district heating supplying the local
hardwood flooring industry, other industries, as well as a residential heating network. Repotec
owns the gasifier technology, developed by the Technical University of Vienna, and has built a
larger unit at Heiligenkreuz, Austria. The Heilegenkreuz unit is designed to produce
11 MWe/6 MWth, and was commissioned in late 2006.
Figure 4-34 shows the overall process flow diagram. The dimensions of the gasifier vessel are
2.5-m ID, 5.5-m OD, and 8-m tall. The vessel is refractory lined and has an external combustion
zone or combustion leg, operating at nominally atmospheric pressure, with steam injection in the
bottom of the vessel to keep N2 levels low. The ash is stored and, at some point in the future, is
expected to be cleared by environmental authorities to be spread on fields. It takes about 12 to 15
hours to start up the gasifier and two days to shut down, mostly due to thermal stress limitations
on the refractory. The gasifier could be scaled up easily, according to the staff at the site, but
would require multiple combustion legs around the central gasifier rather than a single
combustion leg as currently configured. The bed temperature is 850C and the gas temperature is
about 830C at the gasifier exit.

4-67

Biomass Electricity Generation

Tar less than 20

150C

830C

850C

Figure 4-34
Schematic of biomass gasification application for combined heat and power

The syngas reports to the gas cooler, where it is cooled to 150C to generate steam, which goes
to the district heating network. The cooled gas goes to the baghouse, where most of the
particulates are removed and reintroduced into the combustion leg. The baghouse is pulsed with
nitrogen. The particulate-free gas reports to the scrubber, which uses biodiesel to scrub tars and
condense H2O. Incoming tar levels are 500 ppm, and less than 20 ppm tar is in the product gas.
Then the cleaning liquid is separated, and the middles (tar/water/biodiesel) are introduced into
the combustion leg. The gas composition is 22% CO, 23% CO2, 40% H2, 10% CH4, 4% N2.
4.4.5.4 Wood Gasification at Harbore
A Vlund-designed updraft gasifier has been installed and operated at Harbore in western
Denmark since 1988. Figure 4-35 shows a process flow diagram of the system. The gasifier
dimensions are about 2.2 m ID and 12 m high. It operates at atmospheric pressure. The feedstock
is nominally 40 mm green wood chips, delivered from local suppliers. The unit consumes 1.5
tons/hr at maximum load. The chips are delivered to the top of the gasifier, where a spreader
turning in the middle distributes the chips. When the torque on the spreader decreases, more
wood is added. At the bottom of the gasifier, humidified air enters the gasifier, which consumes
3 to 4 Nm3 water per day. The temperature is 1200C at the bottom of the gasifier

4-68

Biomass Electricity Generation

vessel and the product gas leaves the top of the gasifier at 75C. The gas composition is 30%
CO, 20% H2, 5% methane, 8% CO2, 3% N2, and the heating value of 6 MJ/nm3, compared to
natural gas at about 40 MJ/nm3. This does not include the dilution of the humidified air. The gas
contains 100 g/nm3 of tar. For firing the gas in IC engines, its tar content needs to be less than
50 mg/nm3. Ash is drawn out of the bottom, and the annual production is about 50 t/year.

Figure 4-35
Process flow diagram for the Vlund Gasifier at Harbore

The gas is cooled in two stages to 35C, and then enters a wet ESP. The product gas is clean
enough to be burned in Jenbacher engines. The ESP effluent, a water/tar/particulate mixture,
reports to a settling vessel. The heavy tar is stored in a recirculated, heated vessel for combustion
during the winter months as peaking heat. The water/soluble tars are heated until vaporized,
then combusted for additional district heating capacity.
The electric capacity of the unit is 1.3 MW and the thermal service is up to 3 MW. The supplier,
Babcock and Wilcox Vlund, has indicated that it should be feasible to scale this gasification
process to about 20 MW total energy output.
4.4.5.5 Gasification of Wood Pellets at Skive
In Skive in northern Denmark, the local utility, Fjernvarme, has an extensive district heating
network, currently fired predominantly by wood pellets and natural gas. There is some electricity
generation as well. The heating network encompasses 8000 homes and is growing. To meet that
growth, Fjernvarme is building a wood gasifier at its Thorvej site. The gasifier will produce
6 MW electric power in three 2-MW Jenbacher engines, and 20 MW of steam heat in two
10-MW boilers. Figure 4-36 shows the process flow diagram for the plant.

4-69

Biomass Electricity Generation

Figure 4-36
Fjernvarme process flow diagram for coproduction of 6 MW electric and 10 MW thermal

The gasifier is a Carbona slightly pressurized gasifier. One of the unique features of this gasifier
is a tar cracker to address the tar problems that other gasifiers have had. Fjernvarme is producing
electricity primarily due to government mandate and would prefer to concentrate on producing
hot water.
The Thorsvej site already has two large wood pellet-fired stoker boilers on site, so the fuel
handling storage and infrastructure is already largely in place.
The gasifier at Skive has completed construction and was undergoing startup activities at the
time of this writing. Originally, the unit was to have been commissioned in summer 2006, but
business issues with the constructor have introduced delays.
4.4.5.6 Cofiring Syngas at Ruien
In western Belgium, Electrabels Ruien power station has begun to embrace biomass fuels as
part of its goal to displace coal-based carbon. A part of that strategy is to install an atmospheric
fluid bed gasifier to provide syngas to the boiler. The gasifier operates on sized green wood. The
feed is generally coarse 50-mm wood chunks. Wood is delivered via walking floor truck,
screened for oversize and tramp material, and conveyed to a covered storage area. The wood is
extracted and conveyed to the gasifier. Figure 4-37 shows a schematic of the gasification facility
at Ruien. Note that the gasifier has only particulate removal for gas cleanup. This is all that is
required, since the gas is directly injected into the boiler and any tars or other organics will be
destroyed in the combustion process. The injection of the fuel gas occurs below the lowest coal
firing level of Unit 5. The syngas contributes about 17 MW to the 190-MW rated capacity of the
unit.

4-70

Biomass Electricity Generation

Figure 4-37
Overall configuration of the biomass gasifier at Ruien

Although there have been a few issues with this unit, it has mostly operated very smoothly.
There are indications that using dry wood causes more operational problems than fresh, green
wood, and consequently, Electrabel strives to have mostly fresh wood delivered.
4.4.5.7 Wood Gasification at Kokemki
The town of Kokemki, Finland, is the site of a demonstration of the Condens Oy Novel
gasifier. The gasifier is a fixed-bed unit, and is fueled by local wood supply. The unit has an
electrical output of 1.8 MWe and a thermal output of 4.3 MWth. The electricity is generated by
syngas-fired reciprocating engines. The unit has been in operation since 2005. The vendor
indicates that capacity could be increased to 10 MW or greater. Figure 4-38 shows a schematic
of the Kokemki gasification system.

4-71

Biomass Electricity Generation

Figure 4-38
Schematic of biomass gasification and power generation system at Kokomki

4.4.6 Non-Commercial Technologies


A host of technologies are commonly proposed for biomass, either to supply fuel or to utilize
that fuel successfully. These technologies include (but are not limited to) the following:

Energy farms to produce woody biomass

Energy farms to produce herbaceous biomass

Direct combustion or gasification of animal manures

Pressurized biomass gasification for use in integrated gasification/combined-cycle


combustion turbine systems with or without coal gasification systems

Technologies for production and utilization of heavy pyrolysis liquids

Indirect liquefaction of biomass fuels by Fischer-Tropsch or other technologies based on


gaseous feedstocks from landfills, wastewater treatment plants, and producer gasifiers

4-72

Biomass Electricity Generation

To date, each of these technologies has encountered significant technical, economic, and
institutional barriers sufficient to inhibit implementation on a commercial utility scale.
Furthermore, the traditional commercial testthe ability of vendors to give guaranteeshas not
been met. Consequently, the technical and economic information available to provide
performance or cost data is somewhat speculative at this time.
Some of these technologies may never achieve commercial status. Many biomass technologies
have already been developed but have not succeeded commercially, including the following:

Cyclone combustion of wood with the addition of inorganic material

Various gasification systems

Mobile pyrolysis in a medium temperature reactor

It is not the purpose of this report to project which technologies will attain commercial status, but
to provide utilities with sufficient information to understand the commercially available
technologies and near-term potentials.

4.5 Cost and Performance Summary


Biomass is unique among renewable resources. Although the capital costs of biomass
technologies can be lower than those of many other renewable technologies, biomass power
technologies must also account for fuel costs that other renewables such as solar, wind, and (for
the most part) geothermal largely avoid.
Except for landfill gas, municipal solid waste (MSW), and a few other special cases in which the
fuel is free or even comes at a negative cost due to disposal credits, the cost of a biomass
feedstock is a significant fraction of the total cost of generating electricity. Also, compared to
coal and natural gas, biomass power plants have relatively high capital costs and low
efficiencies, especially versus natural gas. Hence, the ability to find low-priced, abundant fuel is
critical to the success of biomass power generation.
The capital and O&M cost estimates presented in the following subsections are from the 2010
EPRI report, Engineering and Economic Evaluation of Biomass Power Plants, 100% Biomass
Repowering, Biomass Cofiring, and Bubbling Fluidized Bed Biomass Combustion [43].
4.5.1 100% Biomass Repowering of a Pulverized Coal Boiler
Modification of coal-fired boilers to fire biomass fuels is achieved by removing the lower
furnace section, including the coal burner systems and much of the secondary air systems. This
strategy maintains the existing heat transfer surfaces located in the upper portion of the boiler
and the existing steam cycle. To allow for the combustion of biomass as the sole fuel for the unit,
a grate or fluidized bed system is installed in place of the coal burner systems. Figure 4-39 is a
schematic showing replacement of the bottom of a pulverized-coal boiler by a bubbling fluidized
bed (BFB) combustion system.
Tables 4-22 to 4-27 summarize plant design assumptions, performance, total capital requirement,
fixed and variable operation and maintenance costs, and levelized cost of electricity estimates for
conversion of a 100-MW pulverized coal unit to a bubbling fluidized bed boiler burning woody
biomass.
4-73

Biomass Electricity Generation

Proposed section to be
added to the existing
furnace

Figure 4-39
Example bubbling bed retrofit to a pulverized coal boiler
Source: Metso

4.5.1.1 Plant Design Assumptions


The biomass repowering performance and cost estimates are based on the assumptions in Table
4-23. The unit is originally a wall-fired pulverized coal boiler rated at 100-MW net capacity.
Repowering the unit to burn 100% biomass reduces the net rated capacity to 60 MW. It is
assumed that the biomass fuel is undried woody biomass containing 45% moisture with dry-basis
higher heating value of 8670 Btu/lb. Fuel consumption for both undried and dried wood is given
in Table 4-24.
Table 4-23
Plant design assumptions for biomass repowering base case
Source: EPRI [48]
Parameter

Biomass Repowering Base Case Scenario

Net Capacity (Coal)

100-MW

Original Boiler Type

Wall-fired PC

Net Capacity (Biomass)

60-MW

Biomass Boiler Type

BFB

Capacity Factor

85%

Biomass Properties
Fuel Type
Heating Value (dry)
Moisture Content

4-74

Undried, blended biomass


8670 Btu/lb
45%

Biomass Electricity Generation


Table 4-24
Raw and dried biomass fuel composition
Fuel Quality Parameter

Raw Wood, 45% Moisture

Dried Wood, 30% Moisture

4770

6070

Moisture, %

45.00

30.00

Ash, %

2.18

2.77

Volatile Matter, %

45.32

57.68

Fixed Carbon, %

7.49

9.54

Carbon, %

27.67

35.22

Hydrogen, %

2.53

3.22

Nitrogen, %

0.57

0.73

Sulfur, %

0.06

0.08

Chlorine, %

0.02

0.03

Moisture, %

45.00

30.00

Ash, %

2.18

2.77

Oxygen, %

21.96

27.96

Higher Heating Value, Btu/lbm


Proximate Analysis

Ultimate Analysis

4.5.1.2 Plant Performance Estimate


Table 4-25 compares the performance of the original coal-fired PC unit and the 100% biomass
repowered unit. Repowering reduces the gross turbine output by about 30% to 68.2 MW and net
output from 100 MW to 60 MW. Boiler efficiency decreases from 88% to 78%, and net plant
heat rate increases from 10,690 to 12,600 Btu/kWh. Even with the lower rating, fuel mass input
with biomass repowering is almost double that of the original coal unit.
4.5.1.3 Total Performance and Cost Estimates
Table 4-26 summarizes the performance, total capital requirement, and fixed- and variable-O&M
and costs for the 100% biomass repowered plant. All costs are in October 2010 dollars and are
based on the 60-MW net output of the repowered plant.
The total capital requirement estimate is $1970/kW. The scope includes major equipment (fuel
handling and preparation and boiler modification), direct and indirect balance-of-plant costs,
interest during construction, and owners costs. Fuel handling/preparation and boiler
modification respectively contribute 16% and 47% of the total plant cost before addition of
interest during construction and owners costs.

4-75

Biomass Electricity Generation

The fixed and variable O&M cost estimates are respectively $132/kW-yr and $3.5/MWh. Fixed
O&M costs consist of wages and wage-related overheads for the permanent plant staff, routine
equipment maintenance, and other fees. Variable O&M costs include costs associated with
equipment maintenance during outages, catalyst/reagents, chemicals, water, and other
consumables. Fuel costs are determined separately and are not included in either fixed or
variable O&M costs.
Table 4-25
Plant performance summary for PC coal-fired and repowered biomass-fired bubbling
fluidized bed plants
Source: EPRI [43]
Units

Coal-fired
PC Unit

Woody BiomassFired BFB Unit

kW

113,000

68,200

Btu/kWh

8,218

8,635

kW

13,000

8,200

11.5

12.0

kW

100,000

60,000

MBtu/h

941

589

Boiler Efficiency

88.0

78.0

Boiler Heat Input

MBtu/h

1,069

756

Coal Consumption

tons/h

43.8

Biomass Consumption

tons/h

79.3

Btu/kWh

10,690

12,600

Parameter
Turbine Gross Output
Turbine Heat Rate

Total Auxiliary Power

Net Plant Output

Heat to Steam from Boiler

Net Plant Heat Rate


Notes:

1. Performance is preliminary and for information only. Not to be used for detailed design.
2. Auxiliary power is assumed.
3. Once through cooling is assumed.
4. Average ambient conditions of 62.1F dry bulb temperature and 54.8F wet bulb temperature. Dew point
temperature (49.6F) is assumed to be the cold water temperature.

4-76

Biomass Electricity Generation


Table 4-26
100% Biomass repowering performance and cost estimates (3Q 2010$)
Source: EPRI [43]
100% Biomass Repowering
Rated Capacity
Plant Size (units x unit size, MW)
Physical Plant
Unit Life, Years
Scheduling
Preconst., License & Design Time, Years
Idealized Plant Construction Time, Years
Hypothetical In-Service Date1
Capital Costs
Major Equipment Costs
Fuel Handling/Preparation
Boiler Modification
Total Major Equipment Costs
Direct Balance of Plant Costs
Site Work, Foundations, Roadways
Electrical Equipment, Cable, & Raceway
Instrumentation & Controls
Total Direct Balance of Plant Costs
Indirect Balance of Plant Costs
Facilities, Engineering, and Const. Mgt.
Project & Process Contingency & Fees
Total Indirect Balance of Plant Costs
Total Costs
Total Plant Costs
AFUDC (Interest during construction)
Total Plant Investment (incl. AFUDC)
Owners Cost
Total Capital Requirement
O&M Costs
Fixed, $/kW-yr
Variable, $/MWh
Performance/Unit Availability
Net Heat Rate (Full Load), Btu/kWh
Equivalent Planned Outage Rate, %
Duty Cycle
Minimum Load, %
Emission Rates
NOx, lb/MBtu
SOx, lb/MBtu
Particulate, lb/MBtu
Confidence and Accuracy Rating
Technology Development Rating
Design & Cost Estimate Rating

1 60
20

($1,000)

2.0
0.75
January 1, 2011
($/kW)

(%)

$16,540
$50,000
$66,540

$280/kW
$830/kW
$1,110/kW

16%
47%
63%

$970
$750
$1,670
$3,390

$20/kW
$10/kW
$30/kW
$60/kW

1%
1%
2%
3%

$15,390
$20,680
$36,070

$260/kW
$340/kW
$600/kW

15%
19%
34%

$106,000
$1,200
$107,200
$10,700
$117,900

$1,770/kW
$20/kW
$1,790/kW
$180/kW
$1,970/kW

100%

132
3.50
12,600
4
Baseload
40
0.150.24
Fuel dependent
<0.02
Commercial
Simplified

Notes:
1. The Hypothetical In-Service Date assumes Notice to Proceed with preliminary activities on January 1, 2011 and
assumes no overlap of pre-construction and construction activities.
2. All costs are presented in October 2010 dollars.

4-77

Biomass Electricity Generation

4.5.1.4 Levelized Cost of Electricity Estimates


Table 4-27 summarizes the levelized cost of electricity (LCOE) estimates for the 100% biomass
repowered plant for the base case and two sensitivity cases. The levelized costs are in constant
third-quarter 2010 dollars. The base case assumes that the biomass fuel cost $3.55/MBtu. The
two sensitivity cases assume costs of $2.55 and $4.55/MBtu.
The levelized cost of electricity estimates are $91/MWh for the base case (the delivered biomass
fuel vs. coal is $3.55/MBtu) and $78 and $103/MWh for the two sensitivity cases (biomass fuel
cost is $2.55 and plus $4.55/MBtu.
The other major assumptions include 20-year project life, zero inflation, 55%/45% debt/equity
ratio, 4% interest on debt, 8% return on equity, 15-year debt financing term, zero investment tax
credit, five-year MACRS tax depreciation, and 40% income tax rate. If the 30% investment tax
credit (ITC) is applied, the levelized cost decreases by about $11/MWh. It should be noted that it
is uncertain whether repowering projects would qualify for the ITC.
Table 4-27
LCOE estimates for 100% biomass repowering (3Q 2010 $)
Source: EPRI [48]
Cost
Biomass Fuel Cost ($/MBtu)

Base Case

Sensitivity
Case 11

Sensitivity
Case 21

3.55

2.55

4.55

18

18

18

Levelized Cost of Electricity2 With ITC


($/MWh)
Fixed O&M Component of LCOE ($/MWh)
Variable O&M Component of LCOE ($/MWh)

Fuel Component of LCOE ($/MWh)

45

32

57

Capital Charge Component of LCOE3


($/MWh)

14

14

14

80

67

92

18

18

18

Total4 ($/MWh)
Levelized Cost of Electricity2 Without ITC
($/MWh)
Fixed O&M Component of LCOE ($/MWh)
Variable O&M Component of LCOE ($/MWh)

Fuel Component of LCOE ($/MWh)

45

32

57

Capital Charge Component of LCOE3


($/MWh)

22

22

22

88

76

101

Total ($/MWh)

Notes:
1

Sensitivity cases assume biomass fuel costs are $1.00/MBtu above and below the base cost of $3.55/MBtu.

Estimate of LCOE assumes an 85% capacity factor, 20-year project life, and constant dollars.

Levelized fixed charge is 7.33%/yr with the investment tax credit (ITC) and 8.5%/yr without the credit.

Sum of LCOE component values may not equal total value due to rounding.

4-78

Biomass Electricity Generation

4.5.2 Biomass Cofiring with Coal in a Pulverized Coal Boiler


There are two basic approaches to cofiring. The first is to blend the fuels and feed them together
to the coal processing equipment (i.e., crushers or pulverizers). In a cyclone boiler, generally up
to 10% of the coal heat input can be replaced by biomass fuels. The smaller fuel particle size
requirement of a PC boiler generally limits the fuel replacement to perhaps 3%. Another
approach is to develop a separate biomass processing system, in which high cofiring percentages
(10% and greater) in a PC unit can be accomplished, although at somewhat higher cost.
The section addresses cofiring in a 250-MW wall-fired pulverized coal boiler. In addition to the
base case coal fired plant without cofiring, three biomass fuels are considered, undried blended
biomass, dried biomass, and torrefied biomass. The design data, performance, capital and O&M
cost, and levelized cost of electricity estimates are presented in the following sections.
4.5.2.1 Plant Design and Fuel Assumptions
The biomass cofiring performance and cost estimates are based on base-case assumptions in
Table 4-28 and the coal and biomass fuel compositions in Table 4-29. The unit is a wall-fired
pulverized coal boiler rated at 250 MW net capacity. The baseline coal is Central Appalachian
coal containing 12,207 Btu/lb. In the base case, it is assumed that the biomass fuel is undried
woody biomass containing 45% moisture with wet-basis higher heating value of 4768 Btu/lb, the
biomass is injected separately into boiler, and the biomass provides 10% of the heat input to the
boiler. In addition, two other biomass fuels are addressed in the performance estimates in the
next section: dried woody biomass containing 30% moisture and 6068 Btu/lb (wet-basis, high
heating value), and torrefied biomass containing 6.2% moisture and 12,759 Btu/lb.
Table 4-28
Summary of biomass cofiring base case
Source: EPRI [48]
Parameter
Cofiring Method
Net Capacity
Boiler Type

Biomass Cofiring Base Case Scenario


Separate injection
250 MW
Wall-fired pulverized coal

Capacity Factor

80%

Biomass Heat Input

10%

Biomass Properties
Fuel Type
Heating Value (dry)
Moisture Content

Undried, blended biomass


8,670 Btu/lb
45%

Coal Properties
Type
Heating Value

Central Appalachian
12,207 Btu/lb

4-79

Biomass Electricity Generation


Table 4-29
Woody biomass fuel quality: heat content, proximate, and ultimate analyses
Source: EPRI [48]
Fuel Quality
Parameter

Baseline CAPP
Coal

Raw Wood,
45% Moisture

Dried Wood,
30% Moisture

Torrefied
Wood

12,207

4768

6068

12,759

Moisture, %

8.47

45.00

30.00

6.21

Ash, %

11.24

2.18

2.77

1.78

Volatile Matter, %

29.57

45.32

57.68

19.66

Fixed Carbon, %

50.72

7.50

9.54

72.35

Carbon, %

69.96

27.67

35.22

78.24

Hydrogen, %

4.47

2.53

3.22

3.19

Nitrogen, %

1.36

0.57

0.73

0.56

Sulfur, %

1.03

0.06

0.08

0.01

Chlorine, %**

0.15

0.02

0.03

0.00

Moisture, %

8.47

45.00

30.00

6.21

Ash, %

11.24

2.18

2.77

1.78

Oxygen, %

3.47

21.97

27.96

10.01

Higher Heating
Value, Btu/lbm
Proximate Analysis*

Ultimate Analysis*

Notes:
* As-received values.
** Although not in the ASTM ultimate analysis, chlorine is often reported with it.

4.5.2.2 Plant Performance Estimates


Tables 4-30 through 4-32 present detailed plant performance data for the base coal-fired plant
and cofiring coal with the three biomass fuels: undried, dried, and torrefied biomass fuel, each at
5%, 10%, and 15% heat input.
For undried and dried biomass cofiring, boiler efficiency and net plant heat rate decrease slightly
and net plant heat rate increases slightly as the biomass heat input increases from 0% to 15% of
total heat input. The opposite happens for cofiring co-milled coal and torrefied wood, mainly
because the moisture content of the torrefied biomass is lower than that of coal and thus the
moisture losses are lower.

4-80

Biomass Electricity Generation


Table 4-30
Performance estimates for cofiring coal and undried biomass
Source: EPRI [48]

Coal

Coal +5%
Undried
Biomass

Coal +10%
Undried
Biomass

Coal +15%
Undried
Biomass

Gross Power, MW

261.50

261.50

261.50

261.50

Net Power, MW

250.00

248.46

248.52

247.88

Boiler Efficiency, %

87.38

87.10

86.62

86.13

Net Plant Heat Rate, Btu/kWh

9762.5

9892.5

9951.8

10,037.3

Total Heat Input, MBtu/hr

2440.6

2457.9

2473.3

2488.0

0.0

122.9

247.3

373.2

Coal Burn Rate, ton/hr

99.97

95.65

91.18

86.63

Biomass Burn Rate, ton/hr

0.00

12.89

25.94

39.14

Full Load Performance

Biomass Heat Input, MBtu/hr

Table 4-31
Performance estimates for cofiring coal and dried biomass
Source: EPRI [48]

Coal

Coal +5%
Dried Biomass

Coal +10%
Dried
Biomass

Coal +15%
Dried
Biomass

Gross Power, MW

261.50

261.50

261.50

261.50

Net Power, MW

250.00

248.49

248.58

247.97

Boiler Efficiency, %

87.38

87.28

87.13

86.91

Net Plant Heat Rate, Btu/kWh

9762.5

9867.7

9901.6

9958.1

Total Heat Input, MBtu/hr

2440.6

2452.1

2461.4

2469.3

0.0

122.6

246.1

370.4

Coal Burn Rate, ton/hr

99.97

95.42

90.74

85.97

Biomass Burn Rate, ton/hr

0.00

10.10

20.28

30.52

Full Load Performance

Biomass Heat Input, MBtu/hr

4-81

Biomass Electricity Generation


Table 4-32
Performance estimates for cofiring co-milled coal and torrefied biomass
Source: EPRI [48]

Coal

Coal +5%
Torrefied
Wood
Biomass

Coal +10%
Torrefied Wood
Biomass

Coal +15%
Torrefied
Wood Biomass

Gross Power, MW

261.50

261.50

261.50

261.50

Net Power, MW

250.00

248.40

248.48

248.58

Boiler Efficiency, %

87.38

87.56

87.63

87.69

Net Plant Heat Rate, Btu/kWh

9762.5

9830.7

9822.8

9811.5

Total Heat Input, MBtu/hr

2440.6

2441.9

2440.8

2438.9

0.0

122.1

244.1

365.8

Coal Burn Rate, ton/hr

99.97

95.02

89.98

84.92

Biomass Burn Rate, ton/hr

0.00

4.78

9.56

14.34

Full Load Performance

Biomass Heat Input, MBtu/hr

4.5.2.3 Total Capital Requirement and Operation and Maintenance Cost Estimates
Table 4-33 summarizes total capital requirement, and fixed and variable operation and
maintenance cost estimates for the base case, cofiring coal and biomass at 10% heat input from
undried biomass fuel. All costs are in October 2010 dollars and are based on the contribution of
biomass to the rated capacity with cofiring (10% of 250 MW, or 25 MW).
The total capital requirement estimate at 10% heat input from biomass is $533/kW. The scope
includes major equipment (fuel handling and preparation and boiler modification), direct and
indirect balance-of-plant costs, interest during construction, and owners costs. Fuel
handling/preparation and boiler modification respectively contribute 16% and 47% of the total
plant cost before addition of interest during construction and owners costs.
The incremental fixed and variable O&M cost estimates are respectively $12.44/kW-yr and
$1.72/MWh, based on the contribution of biomass to the 250-MW rated capacity. Fixed O&M
costs consist of wages and wage-related overheads for the permanent plant staff, routine
equipment maintenance, and other fees. Variable O&M costs include costs associated with
equipment outage maintenance, catalyst/reagents, chemicals, water, and other consumables. Fuel
costs are determined separately and are not included in either the fixed or variable O&M costs.

4-82

Biomass Electricity Generation


Table 4-33
Biomass cofiring performance and cost summary (3Q 2010$)
Source: EPRI [48]
Rated Capacity
Coal plant size (units x unit size), MW
Biomass fuel feed system
Fraction of plant output from biomass, %
Equivalent power output, MW
Physical Plant
Unit life, years
Scheduling
Preconst., License & Design Time, Years
Idealized Plant Construction Time, Years
Hypothetical In-Service Date1
Incremental Total Capital Requirement
Major Equipment Costs
Fuel Handling/Prep
Boiler Modification
Total Major Equipment Costs
Direct Balance of Plant Costs
Site Work, Foundations, Roadways
Electrical Equipment, Cable, & Raceway
Instrumentation & Controls
Total Direct Balance of Plant Costs
Indirect Balance of Plant Costs
Facilities, Engineering, and Const. Mgt.
Project & Process Contingency & Fees
Total Indirect Balance of Plant Costs
Total Costs
Total Plant Costs
AFUDC (Interest during construction)
Total Plant Investment (incl. AFUDC)
Owners Cost
Total Capital Requirement
Incremental O&M Costs (cofiring systems only)
Fixed, $/kW-yr
Variable, $/MWh
Performance/Unit Availability
Net Plant Heat Rate, Btu/kWh (HHV)
Equivalent Planned Outage Rate, %
Duty Cycle
Minimum Load, %
Emission Rates
NOx, lb/MBtu
SOx, lb/MBtu
Particulates, lb/MBtu

Biomass Cofiring with Coal


1 250-MW
Separate Injection
10
25
20

($ 1,000)

1.5
0.5
January 1, 2011
($/kW)3

(%)

$8,325
$1,025
$9,350

$333/kW
$41/kW
$374/kW

64%
8%
72%

$587.50
$300.0
$587.50
$1,475

$23.50/kW
$12.0/kW
$23.50/kW
$59/kW

4.5%
2.0%
4.5%
11%

$525
$1,700
$2,225

$21/kW
$68/kW
$89/kW

4%
13%
17%

$13,050

$522/kW

100%

$13,050

$522/kW

$275

$11/kW

$13,325

$533

12.44
1.72
9,760
4
Baseload
40
0.05
0.06
0.006

4-83

Biomass Electricity Generation


Tabe 4-33 (continued)
Biomass cofiring performance and cost summary (3Q 2010$)
Source: EPRI [48]
Confidence and Accuracy Rating
Technology Development Rating

Commercial

Design & Cost Estimate Rating

Simplified

Notes:
1. The Hypothetical In-Service Date assumes Notice to Proceed with preliminary activities on January 1, 2011 and
assumes no overlap of pre-construction and construction activities.
2. All costs are presented in October 2010 dollars.
3. Costs presented on $/kW basis are determined based on biomass-fired capacity (i.e., equivalent power output of 25
MW).

4.5.3 Biomass-Fired Bubbling Fluidized Bed Combustion Boiler Power Plants


Tables 4-34 to 4-37 summarize plant design assumptions, performance, total capital requirement,
fixed and variable operation and maintenance costs, and levelized cost of electricity estimates for
three cases:

50-MW bubbling fluid bed boiler fired by undried woody biomass

100-MW bubbling fluid bed boiler fired by undried woody biomass

50-MW bubbling fluidized bed boiler fired by a mixture of 80% woody biomass and 20%
switchgrass based on heat input.

4.5.3.1 Plant Design Assumptions


The three biomass bubbling bed fluidized bed combustion boiler performance and cost estimate
cases are based on the assumptions in Table 4-34. It is assumed that the biomass fuels are
undried woody biomass containing 45% moisture and 8670 Btu/lb (dry HHV) and as-received
switchgrass containing 15% moisture and 7304 Btu/lb.
Table 4-34
Design assumptions for bubbling BFB boiler plants
50-MW BFB
Woody
Biomass

100-MW BFB
Woody
Biomass

50-MW BFB Switchgrass


/ Woody Biomass

50-MW

100-MW

50-MW

Boiler Type

BFB

BFB

BFB

Capacity Factor

85%

85%

85%

100%

100%

20%Switchgrass/ 80%
woody biomass

Undried
blended
biomass
8,670 Btu/lb

Undried blended
biomass
8,670 Btu/lb

7,304 Btu/lb Switchgrass

45%

45%

15% Switchgrass

Parameter
Net Capacity

Biomass Heat Input


Biomass Properties

Fuel Type
Higher Heating Value (dry)
Moisture Content

4-84

Woody biomass blended


with as-received
switchgrass

Biomass Electricity Generation

4.5.3.2 Plant Performance Estimate


Table 4-34 compares the performance estimates for the BFB boiler cases. The boiler
efficiency is 75% for all three cases. Net plant heat rates are respectively 13,964, 12,894, and
14,119 Btu/kWh for the 50- and 100-MW BFB plants fired by woody biomass and the 50-MW
plant fired by a 80% biomass/20% switchgrass mixture. The principle reason that the 100-MW
BFB plant has a lower heat rate is that the steam cycle is more efficient than that of the 50-MW
BFB plant due to higher steam conditions. In addition, the parasitic power consumption as a
fraction of gross output is lower.
Table 4-35
Plant performance estimates for BFB boiler plants
Source: EPRI [48]
50-MW BFB
Woody
Biomass

100-MW BFB
Woody
Biomass

50-MW BFB
Switchgrass/
Woody Biomass

Turbine Gross Output, kW

58,824

113,636

59,524

Turbine Heat Rate, Btu/kWh

8,902

8,505

8,895

Total Auxiliary Power, kW

8,824

13,636

9,524

Total Auxiliary Power, %

15

12

16

50,000

100,000

50,000

75

75

75

Boiler Heat Input, MBtu/hr

698.2

1289.4

705.9

Biomass Consumption, ton/hr

73.2

135.2

69.8

Net Plant Heat Rate, Btu/kWh

13,964

12,894

14,119

Parameter

Net Plant Output, kW

Boiler Efficiency, %

4.5.3.3 Total Capital Requirement and Operation and Maintenance Cost Estimates
Table 4-36 summarizes the performance, total capital requirement, and fixed- and variable-O&M
cost estimates for the bubbling fluidized bed boiler cases. All costs are in October 2010 dollars
and are based on net output.
The total capital requirement estimates are respectively $5,588, $4,053, and $5,806/kW for the
50- and 100-MW BFB plants fired by woody biomass and the 50-MW plant fired by an 80%
biomass/20% switchgrass mixture. The scope includes major equipment (fuel handling and
preparation and boiler modification), direct and indirect balance-of-plant costs, interest during
construction, and owners costs.

4-85

Biomass Electricity Generation


Table 4-36
Performance and cost estimates for bubbling fluidized bed boiler plants (3Q 2010$)
Source: EPRI [48]
50-MW BFB
Woody
Biomass

100-MW BFB
Woody
Biomass

50-MW BFB
Switchgrass /
Wood

Southeast

Southeast

Southeast

Rated Capacity (units x unit size), MW

1 50 MW

1 100 MW

1 50 MW

Biomass fuel feed system

Mechanical

Mechanical

Mechanical

100%

100%

80%/20%
Wood/Switchgrass

30

30

30

Preconst., License & Design Time, Years

1.5

Idealized Plant Construction Time, Years

2.5

2.5

Jan-11

Jan-11

Jan-11

$1,350

$1,050

$1,350

Turbine Island System

$370

$300

$370

Biomass Handling

$350

$270

$480

Environmental Controls

$110

$70

$110

$2,180

$1,690

$2,310

BOP Facilities

$840

$550

$840

General Facilities & Site Specific

$380

$260

$420

Total Direct Balance of Plant Cost

$1,220

$810

$1,260

$1,330

$890

$1,340

$30

$30

$30

Net Capacity and Boiler Type


Location
Rated Capacity

Fraction of plant output from biomass, %


Physical Plant
Unit life, years
Scheduling

Hypothetical In-Service Date


Total Capital Requirement, $/kW
Major Equipment Cost
Steam Generator System

Total Major Equipment Cost


Direct Balance of Plant Cost

Indirect Balance of Plant Costs


Engineering Fee & Construction
Management
Process Contingency
Project Contingency
Total Indirect Balance of Plant Cost

$170

$120

$180

$1,530

$1,040

$1,550

$4,930

$3,530

$5,120

Total Cost
Total Plant Cost

4-86

AFUDC (Interest during construction)


Total Plant Investment (including
AFUDC)
Total Owners Cost, $/kW

$160

$157

$168

$5,090

$3,684

$5,278

$510

$369

$528

Total Capital Requirement

$5,590

$4,053

$5,806

Biomass Electricity Generation


Table 4-36 (continued)
Performance and cost estimates for bubbling fluidized bed boiler cases (3Q 2010$)
Source: EPRI [48])
50-MW BFB
Woody Biomass

100-MW BFB
Woody Biomass

50-MW BFB
Wood/
Switchgrass

Fixed, $/kWbiomass-yr

113

63

114

Variable, $/MWhbiomass

5.8

5.1

5.85

Net Plant Heat Rate, Btu/kWh (HHV)

13,964

12,894

14,119

Equivalent Planned Outage Rate, %

Baseload

Baseload

Baseload

40

40

40

NOx, lb/MBtu

0.10 0.12

0.1

0.10 0.12

SOx, lb/MBtu

0.04 0.08

0.01

0.04 0.08

0.015 0.035

0.018

0.015 0.035

Commercial

Commercial

Developing

Simplified

Simplified

Simplified

Operation and Maintenance Cost

Performance/Unit Availability

Duty Cycle
Minimum Load, %
Emission Rates

Particulates, total, lb/MBtu


Confidence and Accuracy Rating
Technology Development Rating
Design & Cost Estimate Rating
Notes:

1. The Hypothetical In-Service Date assumes Notice to Proceed with preliminary activities on January 1, 2011 and
assumes no overlap of pre-construction and construction activities.
2. All costs are presented in October 2010 dollars.

The fixed O&M costs are respectively $113, $63, and $114/kW-yr and the variable O&M costs
are respectively $5.8, $5.1, and $5.9/MWh for the 50- and 100-MW BFB plants fired by woody
biomass and the 50-MW plant fired by an 80% biomass/20% switchgrass mixture. Fixed O&M
costs consist of wages and wage-related overheads for the permanent plant staff, routine
equipment maintenance, and other fees. Variable O&M costs include costs associated with
equipment maintenance during outages, catalyst/reagents, chemicals, water, and other
consumables. Fuel costs are determined separately and are not included in either fixed or
variable O&M costs.
4.5.3.4 Levelized Cost of Electricity Estimates
Table 4-37 summarizes the levelized cost of electricity estimates for the 50- and 100-MW BFB
plants fired by undried woody biomass and the 50-MW BFB plant fired by an 80% wood/20%
switchgrass mixture. The costs are levelized over the 30-year project life and are in constant
October 2010 dollars.
The levelized costs are respectively $126, $87, and $148/MWh for the 50- and 100-MW BFB
plants fired by woody biomass and the 50-MW plant fired by an 80% biomass/20% switchgrass
mixture. The major assumptions include 30-year project life, zero inflation, 55%/45%
debt/equity ratio, 4% interest on debt, 8% return on equity, 15-year debt financing term, zero
investment tax credit, five-year MACRS tax depreciation, and 40% income tax rate.
4-87

Biomass Electricity Generation

If the 30% investment tax credit (ITC) is applied, the levelized cost decreases $104, $71, and
$125/MWh for the 50- and 100-MW woody biomass fired plants and the 50-MW plant fired by
the wood/switchgrass mixture. Relative to the repowering and cofiring cases, it is more likely
that stand-alone biomass-fired BFB projects would be eligible for the ITC, provided that the
applying entities (i.e., project owners) are eligible and the project commences operation prior to
the expiration of the ITC.
Table 4-37
Levelized cost of electricity estimates for biomass-fired bubbling fluidized bed plants
(3Q 2010$)
Source: EPRI [48]
50-MW BFB
Woody
Biomass

100-MW BFB
Woody
Biomass

50-MW BFB
Wood/
Switchgrass

3.55

3.55

3.55

5.05

Fixed O&M

15

15

Variable O&M

Fuel

50

46

54

34

24

35

104

84

110

Fixed O&M

15

15

Variable O&M

50

46

54

61

44

63

132

104

139

Cost Component
Biomass Fuel Cost, $/MBtu
Switchgrass Fuel Cost, $/MBtu

Levelized Cost of Electricity2 With ITC ($/MWh)

Capital Charge3
Total

Levelized Cost of Electricity2 Without ITC ($/MWh)

Fuel
Capital Charge
Total3

Notes:
1

Sensitivity cases assume biomass and switchgrass fuel costs are each $1/MBtu above and below the base costs
of $3.55/MBtu and $5.05/MBtu.

Estimate of LCOE assumes an 85% capacity factor, 30-year project life, and constant dollars.

Levelized fixed charge is 6.4%/yr with the investment tax credit (ITC) and 7.42%/yr without the credit.

Sum of LCOE component values may not equal total value due to rounding.

4-88

Biomass Electricity Generation

4.5.4 Biomass-Fired Stoker Boiler Power Plants


Tables 4-38 to 4-41 present the plant design assumptions, performance, total capital requirement,
fixed and variable operation and maintenance costs, and levelized cost of electricity estimates for
the 50-MW stoker boiler plant fired by undried woody biomass. The performance and cost data
were generated by adjusting the 50-MW BFB boiler plant data in Table 4-35 based on data for
stoker and fluidized bed boiler plants in the 2007 EPRI report [42].
4.5.4.1 Plant Design Assumptions
The biomass stoker boiler power plant performance and cost estimates are based on the
assumptions in Table 4-38. It is assumed that the biomass fuel is undried woody biomass
containing 45% moisture and 8670 Btu/lb (dry HHV).
Table 4-38
Design assumptions for stoker boiler plant
Parameter

50-MW Stoker Plant Woody Biomass

Net Capacity

50-MW

Boiler Type

Stoker

Capacity Factor

85%

Biomass Heat Input

100%

Biomass Properties
Fuel Type

Undried blended biomass

Higher Heating Value (dry)

8,670 Btu/lb

Moisture Content

45%

4.5.4.2 Plant Performance Estimate


Table 4-39 presents the performance estimate for the stoker boiler power plant. The boiler
efficiency is 75% and the estimated net plant hear rate is 14,191 Btu/kWh.
Table 4-39
Plant performance estimates for stoker boiler plant
Source: EPRI [42] [43]
Parameter

50-MW Stoker Plant Woody Biomass

Turbine Gross Output, kW

58,824

Turbine Heat Rate, Btu/kWh

8,902

Total Auxiliary Power, kW

8,824

Total Auxiliary Power, %

15

Net Plant Output, kW

50,000

Boiler Efficiency, %

72.8

Boiler Heat Input, MBtu/hr

709.6

Biomass Consumption, ton/hr

74.4

Net Plant Heat Rate, Btu/kWh

14,191

4-89

Biomass Electricity Generation

4.5.4.3 Total Capital Requirement and Operation and Maintenance Cost Estimates
Table 4-40 summarizes the performance, total capital requirement, and fixed- and variable-O&M
cost estimates for the stoker boiler plant. All costs are in October 2010 dollars and are based on
the net output.
The total capital requirement estimate is $5188/kW. The scope includes major equipment (fuel
handling and preparation and boiler modification), direct and indirect balance-of-plant costs,
interest during construction, and owners costs.
Table 4-40
Performance and cost estimates for stoker boiler plant (3Q 2010$)
Source: EPRI [42] [43]
Net Capacity and Boiler Type
Location
Rated Capacity
Rated Capacity (units unit size), MW
Biomass fuel feed system
Fraction of plant output from biomass, %
Physical Plant
Unit life, years
Scheduling
Preconst., License & Design Time, Years
Idealized Plant Construction Time, Years
Hypothetical In-Service Date
Total Capital Requirement, $/kW
Major Equipment Cost
Steam Generator System
Turbine Island System
Biomass Handling
Environmental Controls
Total Major Equipment Cost
Direct Balance of Plant Cost
BOP Facilities
General Facilities & Site Specific
Total Direct Balance of Plant Cost
Indirect Balance of Plant Costs
Engineering Fee & Construction Management
Process Contingency
Project Contingency
Total Indirect Balance of Plant Cost
Total Cost
Total Plant Cost
AFUDC (Interest during construction)
Total Plant Investment (including AFUDC)
Total Owners Cost, $/kW

4-90

50-MW Stoker Plant Woody Biomass


Southeast
1 50 MW
Mechanical
100%
30
1
2.5
Jan-11

$1,209
$370
$350
$110
$2,039
$840
$380
$1,220
$1,330
$29
$165
$1,524
$4,783
$155
$4,938
$247

Biomass Electricity Generation


Table 4-40 (continued)
Performance and cost estimates for stoker boiler plant (3Q 2010$)
Source: EPRI [42] [43]
Net Capacity and Boiler Type
Total Capital Requirement
Fixed, $/kWbiomass-yr
Variable, $/MWhbiomass
Performance/Unit Availability
Net Plant Heat Rate, Btu/kWh (HHV)
Equivalent Planned Outage Rate, %
Duty Cycle
Minimum Load, %
Emission Rates
NOx, lb/MBtu
SOx, lb/MBtu
Particulates, total, lb/MBtu
Confidence and Accuracy Rating
Technology Development Rating
Design & Cost Estimate Rating

50-MW Stoker Plant Woody Biomass


$5,185
110
5.1
14,191
4
Baseload
40
0.10 0.12
0.04 0.08
0.015 0.035
Commercial
Simplified

Notes:
1.

The Hypothetical In-Service Date assumes Notice to Proceed with preliminary activities on January 1,
2011 and assumes no overlap of pre-construction and construction activities.

2.

All costs are presented in October 2010 dollars.

The fixed and variable O&M cost estimates are respectively $132/kW-yr, and $3.5/MWh. Fixed
O&M costs consist of wages and wage-related overheads for the permanent plant staff, routine
equipment maintenance, and other fees. Variable O&M costs include costs associated with
equipment maintenance during outages, catalyst/reagents, chemicals, water, and other
consumables. Fuel costs are determined separately and are not included in either fixed or
variable O&M costs.
4.5.4.4 Levelized Cost of Electricity Estimate
Table 4-41 summarizes the levelized cost of electricity estimate for the stoker boiler plant. The
levelized cost is $122/MWh in constant third-quarter 2010 dollars. The major assumptions
include 30-year project life, zero inflation, 55%/45% debt/equity ratio, 4% interest on debt, 8%
return on equity, 15-year debt financing term, zero investment tax credit, five-year MACRS tax
depreciation, and 40% income tax rate.
If the 30% investment tax credit (ITC) is applied, the levelized cost decreases by about
$20/MWh. Relative to the repowering and cofiring cases, it is more likely that standalone
biomass-fired stoker projects would be eligible for the ITC, provided that the applying entities
(i.e., project owners) are eligible and the project commences operation prior to the expiration of
the ITC.

4-91

Biomass Electricity Generation


Table 4-41
Levelized cost of electricity estimates for 50-MW stoker boiler plant fired by woody
biomass (3rd-Quarter 2010$)
Source: EPRI [47, 48]
Cost Component
Biomass Fuel Cost, $/MBtu

Base Case
3.55

Levelized Cost of Electricity1 With ITC ($/MWh)


Fixed O&M

15

Variable O&M

Fuel Component of LCOE ($/MWh)


Capital Charge

Total3

50
32
102

Levelized Cost of Electricity1 Without ITC ($/MWh)


Fixed O&M

15

Variable O&M

Fuel Component of LCOE ($/MWh)


Capital Charge

Total3

50
52
122

Estimate of LCOE assumes an 85% capacity factor, 30-year project life, and constant dollars.

Levelized fixed charge is 6.4%/yr with the investment tax credit (ITC) and 7.33%/yr without the credit.

Sum of LCOE component values may not equal total value due to rounding.

4.5.5 Biomass Gasification Power Plants


Tables 4-42 to 4-44 present the plant design assumptions, performance, total capital requirement,
fixed and variable operation and maintenance costs, and levelized cost of electricity estimates for
the 42-MW gasification plant.
4.5.5.1 Plant Design Assumptions
The biomass gasification power plant performance and cost estimates are based on the
assumptions in Table 4-42. It is assumed that the biomass fuel is undried woody biomass
containing 45% moisture and 8,670 Btu/lb (dry HHV).

4-92

Biomass Electricity Generation


Table 4-42
Design assumptions for biomass gasification plant
Source: EPRI, 1023994
Parameter

42-MW Gasification Unit

Net Capacity

38-MW

Boiler Type

Stoker

Capacity Factor

80%

Biomass Heat Input

100%

Biomass Properties
Fuel Type
Higher Heating Value (dry)
Moisture Content

Undried blended biomass


8,670 Btu/lb
45%

4.5.5.2 Plant Performance Estimate


Table 4-43 presents the performance estimate for the gasification power plant. The boiler
efficiency is 75% and the estimated net plant heat rate is 14,191 Btu/kWh.
The financial analysis is based on supplier information provided by ANDRITZ Carbona for its
biomass gasification system. The gasification unit itself is a pressurized bubbling fludized bed
system. The bulk of the feedstock is undried, woody biomass reduced to sizes no greater than 1.8
in., followed by pre-drying with a belt dryer system to a maximum moisture of 20%. The
economics of the system is based on a combined cycle configuration with two combustion
turbines, 2 HRSGs, and 1 steam turbine.
The study requirements included evaluation for plant installations of similar sizes in three
geographic regions of the United States and one in Europe. The three regions of the U.S. chosen
were three major quadrants of the U.S.: the Southeast (SE U.S.), Northeast (NE U.S.), and the
Northwest (NW U.S.). These three regions all have extensive woody biomass resources and a
history of biomass-based power plants. Within each of these regions, one U.S. state was selected
as a proxy for the remaining states in those regions. Depending on the boundaries chosen, each
of these regions is comprised of eight to eighteen states, with a land area of about 750,000 square
miles (2,000,000 square kilometers). The specific states used in the evaluation were: SE U.S.
Georgia, NE U.S. Maine, and NW U.S.Oregon. The European site is considered to be in the
United Kingdom.

4-93

Biomass Electricity Generation


Table 4-43
Plant performance estimates for gasification plant
Source: EPRI, 1023994
Parameter
2 - Combustion Turbine Gross Output

42-MW Gasification Plant


13 MW/CT

Steam Turbine Gross Output

16 MW

Total Auxiliary Power

4 MW

Total Auxiliary Power, %


Net Plant Output

9.5%
38 MW

Gasifier Heat Input, MBtu/hr

908

Biomass Consumption, ton/hr

52

Syngas Heating Value,(HHV) Btu/scf

162

4.5.5.3 Total Capital Requirement and Operation and Maintenance Cost Estimates
Table 4-44 summarizes the performance, total capital requirement, and fixed- and variable-O&M
cost estimates for the biomass gasification plant. All costs are in 2011 dollars and are based on
the net output.
The total capital requirement estimate is $5188/kW. The scope includes major equipment (fuel
handling and preparation and boiler modification), direct and indirect balance-of-plant costs,
interest during construction, and owners costs.
Table 4-44
Performance and cost estimates for biomass gasification (2012$)
Source: EPRI
42-MW Gasification
Plant

42-MW Gasification
Plant

Southeast GA

Northeast ME

Rated Capacity (units x unit size), MW

1 42 MW

1 42 MW

Biomass fuel feed system

Mechanical

Mechanical

100%

100%

30

30

2.1

2.1

Fuel Handling/Preparation

$7,590

$7,648

Dryer/Gasification/Power Plant

$87,446

$87,596

Net Capacity and Boiler Type


Location
Rated Capacity

Fraction of plant output from biomass, %


Physical Plant
Unit life, years
Scheduling
Idealized Plant Construction Time, Years
Total Capital Cost, $k
Major Equipment Cost

4-94

Biomass Electricity Generation


Table 4-44 (continued)
Performance and cost estimates for biomass gasification (2012$)
Source: EPRI
42-MW Gasification
Plant

42-MW Gasification
Plant

Southeast GA

Northeast ME

Total Major Equipment Cost

$95,036

$95,244

Direct Balance of Plant Cost

$37,720

$42,435

$37,720

$42,435

Net Capacity and Boiler Type


Location

Total Direct Balance of Plant Cost


Markup of Direct Costs

$17,576

$18,260

Engineering Fee & Construction Mgt.

$13,800

$13,800

Process Contingency

$7,043

$7,820

Total Indirect Balance of Plant Cost

$38,419

$39,880

Total Plant Cost ($k)

$171,175

$177,559

$4,076

$4,228

42-MW Gasification
Plant

42-MW Gasification
Plant

Northwest OR

UK England

Rated Capacity (units x unit size), MW

1 42 MW

1 42 MW

Biomass fuel feed system

Mechanical

Mechanical

100%

100%

30

30

2.1

2.1

Total Plant Cost ($/kW)


Net Capacity and Boiler Type
Location
Rated Capacity

Fraction of plant output from biomass, %


Physical Plant
Unit life, years
Scheduling
Idealized Plant Construction Time, Years
Total Capital Cost, $k
Major Equipment Cost
Fuel Handling/Preparation

$7,648

$7,705

Dryer/Gasification/Power Plant

$87,596

$87,745

Total Major Equipment Cost

$95,244

$95,450

Direct Balance of Plant Cost

$42,435

$47,150

$42,435

$47,150

Markup of Direct Costs

$18,260

$18,948

Engineering Fee & Construction Mgt.

$13,800

$13,800

Process Contingency

$7,820

$8,625

Total Direct Balance of Plant Cost

Total Indirect Balance of Plant Cost

$39,880

$41,373

Total Plant Cost ($k)

$177,559

$183,973

$4,228

$4,380

Total Plant Cost ($/kW)

4-95

Biomass Electricity Generation

The total annual O&M cost (OPEX) assumption is based on ANDRITZ Carbona information and
is $2.8 million per year including labor, supplies, consumables such as water, chemicals, natural
gas, propane, and fuel oil. An annual cost increase of 2.4% is assumed.
Feedstock

The cost for feedstock for biomass gasification is a critical assumption due to the significance of
its cost in relation to total plant costs and due to the volatility in pricing based on feedstock
availability and proximity to the plant.
Feedstock cost was evaluated by using state forestry departments and other forest, wood and
lumber interest websites. The forestry departments are responsible for management and
furtherance of the forest resources within each state and collect and distribute information
regarding biomass types and prices. These representative prices indicate the broad range of
marketable wood materials which are traded within a given state with information with regard to
availability and pricing trends. Between the states there are variations of major uses and
concentrations; for example, Georgia has a greater emphasis on pulpwood than Oregon, which
produces more lumber. Base prices were chosen for this study by reviewing data on the websites
and including an allowance for transportation (wood must be cut, limbs trimmed and loaded onto
a truck, regardless of source; then it must be driven a certain distance to the plant site assumed to
be within about 50 miles).
The feedstock costs assumed for each region are as follows:
SE U.S.Georgia
NE U.S. Maine
NW U.S. Oregon
UK (London)

$ 30/ton
$ 20/ton
$ 45/ton
$ 87/ton

The quantity of feedstock required is based on ANDRITZ Carbona operating information of


1,257 tons per day. The feedstock cost is assumed to increase 2.4% per year.
Property Tax

The cost for property taxes on plant assets is included in annual costs based on an average
property tax rate for each of the regions analyzed. Rates assumed, per $US of economic property
value, are: SE U.S. 1.2%, NE U.S. 1.28%, NW U.S. 1.39% and 0% for the UK. The rate is
applied to a depreciated total plant cost based on 3.33% annual economic depreciation, assuming
a 30-year economic life.
4.5.3.4 Levelized Cost of Electricity Estimate
LCOE for each of the regions are summarized in Table 4-44. Based on the cash flows projected
for capex, opex, feedstock costs, property taxes and incentives over the 30-year plant life, the net
present value (NPV) of these costs was calculated. This NPV, divided by the NPV of the energy
produced over the same 30-year plant life, is the LCOE for the example plant in each region.
Individual contributions to the total LCOE for four major cost components and tax incentives are
included in the table; these were determined using a similar NPV method which used only their
individual cost streams versus generated power values. The LCOE results are shown in Table 445 and the cost components of each LCOE amount are shown graphically in Figure 4-40.
4-96

Biomass Electricity Generation


Table 4-45
LCOE regional costs with component breakdown
LCOE Components - $/MWh

SE

NE

NW

UK

CAPEX

$ 50

$ 52

$ 52

$ 53

OPEX

$ 11

$ 11

$ 11

$ 11

Feedstock

$ 53

$ 36

$ 80

$ 155

Property tax

$ 4

$ 5

$ 5

$ 0

ROC value, net of tax

$ 0

$ 0

$ 0

($ 96)

($ 49)

($ 45)

($ 62)

($ 46)

$ 69

$ 58

$ 87

$ 78

Tax benefits
LCOE

Note: Total LCOE values in each column may not add due to rounding.

Figure 4-40
LCOE cost comparisonscomparison by region

Southeast United States (SE U.S.)

As shown in Table 4-45 and Figure 4-40, the LCOE for the SE U.S. region is in the mid-range of
the four regions evaluated at a total LCOE of $69/MWh. The largest component of the LCOE is
the Feedstock cost, which in this region, comprises $53/MWh of the total LCOE and is
significantly higher than the NE U.S. region, but lower than in the NW U.S. region. The capex
costs amount to $50/MWh which, after being adapted to SE region costs from the Nordic region
base cost, are estimated to be the lowest of the four regions. OPEX and property taxes are
approximately the same as the other U.S. regions and slightly lower than the UK. Before
deducting the tax benefits, the sum of the SE U.S. region costs are $119/MWh. The PTC
incentive included in Tax benefits is $11/MWh and is the same for each U.S. location. While the
UK region does not have this PTC benefit, the ROC benefit of $96/MWh far exceeds the U.S.
4-97

Biomass Electricity Generation

locations PTC benefit. The balance of the SE U.S. regions Tax benefits is due to depreciation
and expense deductions, and this varies from the other locations primarily due to Feedstock cost
differences discussed above.
Northeast United States (NE U.S.)

Table 4-45 and Figure 4-40 show a LCOE in the NE U.S. of $58/MWh, which is the lowest of
the four regions. The largest cost contribution is CAPEX at $52/MWh which is slightly higher
than the SE U.S., but the same as for the NW U.S. Feedstock costs are the lowest of all four
regions by a fairly sizeable margin at $36/MWh which is the key driver in the SE U.S. region
showing the lowest LCOE. OPEX and property taxes are approximately the same as the other
U.S. regions and slightly lower than for the UK. Before deducting the Tax benefits, the sum of
the NE U.S. region costs are $103/MWh. The PTC incentive included in Tax benefits is
$11/MWh and is the same for each U.S. location. Again, while the UK region does not have this
PTC benefit, the ROC benefit of $96/MWh in the UK far exceeds the U.S. locations PTC
benefit. The balance of the NE U.S. tax benefits is due to depreciation and expense deductions,
and this varies from the other locations primarily due to feedstock price differences discussed
above.
Northwest United States (NW U.S.)

Table 4-45 and Figure 4-40 show a LCOE cost of $87/MWh for the NW U.S., which is the
highest of the three U.S. locations, but still lower than for the UK region by $9/MWh. The
$45/ton feedstock cost, which is the highest in the United States, results in the largest
contribution to LCOE at $80/MWh. CAPEX at $52/MWh is about 4% more than for the SE
U.S., the same as for the NW U.S., and about 3% lower than for the UK. OPEX and property
taxes remain at $11/MWh and $5/MW, the same as the other U.S. regions. Total costs before
deducting tax benefits are $148/MW. Tax benefits include the same PTC as the other U.S.
regions, and higher tax deduction benefits than the other U.S. regions due to significantly higher
feedstock costs.
United Kingdom (UK)

Table 4-45 and Figure 4-40 show LCOE projected by the model of $78/MWh for the UK region.
Capex at $53/MWh is 2% to 4% higher than for each of the U.S. locations. Feedstock costs are
significantly higher in this region than any of the U.S. locations, and are by far the largest cost
contribution to LCOE at $115/MWh. OPEX is projected to be $11/MWh, similar to the United
States. There were no property taxes assumed for the UK region, compared to an average of
$5/MWh in the U.S. regions. Total UK costs before reduction for tax benefits and incentives are
$219/MWh, which is almost 48% more than the highest U.S. location, again due primarily to the
high feedstock cost in the UK. Much of this higher UK cost is offset by the ROC incentive,
which amounts to $96/MWh and far exceeds the PTC incentives in the U.S. regions. The Tax
benefits in the UK region are lower than the U.S. locations due to a far lower corporate tax rate
of 24% versus the U.S. average combined federal/state rate of 38%, even though UK deductible
costs of feedstock are so much higher.

4-98

Biomass Electricity Generation

Effect of Production Tax Credit

Due to the uncertainty regarding the extension of the U.S. tax codes PTC, the LCOE results
have been calculated as if the PTC were no longer available, and are shown in Table 4-46.
Table 4-46
LCOE without PTC
LCOE Components - $/MWh
WITHOUT PTC
Capex

SE

NE

NW

UK

$50

$52

$52

$53

Opex

$11

$11

$11

$11

Feedstock

$53

$36

$80

$155

Property tax

$4

$5

$5

$-

ROC value, net of tax

$-

$-

$-

$(96)

Tax benefits

$(38)

$(34)

$(51)

$(46)

LCOE

$80

$69

$98

$78

Note: Total LCOE values in each column may not add due to rounding.

4.6 Biomass to Electricity RD&D Initiatives for the Future


Substantial industrial participation is essential in all but basic research initiatives if biomass
energy is to effectively contribute to GHG reduction policy objectives. Particular attention needs
to be given to issues relevant to fuel supply, conventional combustion, and gasification/pyrolysis.
Primary fuel supply issues that need to be addressed include the following:

Demonstration of innovative equipment for forestry residue systems

Demonstration of machines for establishing, harvesting, transporting, and storing energy crop
systems

Basic and applied R&D on grasses and alternative energy crops for areas not suitable for
wood crops

Continuous breeding and plant improvement programs for energy crops to improve yields

Schemes which can accelerate development of the sustainable biomass fuel supply, such as
standards for certifying sustainable biomass feedstock

Primary conventional combustion issues that need to be addressed include the following:

Demonstration of incremental innovations in technology to improve efficiency, reduce


capital costs, or reduce emissions

Demonstration of technical innovations (e.g. biomass pre-treatment processes, residuederived fuels) that will extend the range of fuels than can be burned

Demonstration of small scale CHP with innovative cycles

4-99

Biomass Electricity Generation

Primary gasification/pyrolysis issues that need to be addressed include the following:

Demonstrations covering the range of gasification and pyrolysis technologies appropriate for
worldwide regions

Applied RD&D of small-scale CHP using advanced systems

Demonstration of pyrolysis systems when the current applied R&D shows sufficient progress

4.7 U.S. Technical Tax Code Modifications to Promote Further Use of


Biomass
Biomass has the potential to generate up to 230 billion kWh of electricity in the United States by
2030. Current U.S. federal tax laws are designed to encourage greater use of biomass to generate
electricity via production tax credits (PTCs), investment tax credits (ITCs), and Treasury grants.
In addition, accelerated depreciation deductions are available to biomass generating facilities.
However, a number of obstacles in U.S. tax law prevent greater biomass development. Several of
these obstacles are identified as follows:

Cofiring: Under IRS Code, an open-loop biomass facility that cofires biomass with fossil
fuels does not qualify for fiscal incentives.

Modification of existing facilities: Facilities that have been converted from fossil fuel units to
open-loop or closed-loop biomass units or cofiring plants do not currently qualify for the
PTC.

Definition of waste: The current definition of open-loop biomass does not assure facility
owners that they can pay for the open loop biomass used as fuel and still qualify for the PTC.

Biomass to create gas or thermal energy: The PTC does not currently apply to facilities that
create biomass-derived renewable energy gas and thermal energy.

4.8 Conclusions
Biomass fuels are many and varied, with distinct characteristics. Although there are institutional
and political questions concerning the classification of wood waste and other waste products as
biomass or renewable, technically they are biomass resources that are derived from living plants.
Most biomass fuels permit utilities to generate dispatchable renewable power. The fuels can
be extracted and/or stored and then used to meet electricity demand. Utilization systems can
capitalize upon the characteristics of these fuels. They are modest in heating value, highly
reactive, low in nitrogen and sulfur, and of varying ash characteristics. Of the biomass fuels
available, woody biomass is the most commonly used material, but agricultural wastes and
dedicated energy crops appear to be the biomass fuels of the future.
Numerous technologies are available for biomass fuel utilization, including both cofiring options
and stand-alone options. Cofiring provides a means for incorporating biomass utilization into
electricity generating facilities with the lowest cost and the least risk. Cofiring can be used to
enhance combustion and, in most cases, reduce SO2 and NOX emissions. The lack of sulfur in
biomass fuels, coupled with low nitrogen concentrations, reactive nitrogen, and reactive fuel as a
whole, provides the basis for this enhanced combustion.
4-100

Biomass Electricity Generation

The forest products and pulp and paper industries, as well as some electric utilities and
independent power producers, have built stand-alone plants. These systems are not inexpensive.
However, they provide a basis for using biomass to generate electricity. They can be
economically justified depending upon localized economics and their use in addressing customer
needs. Institutional arrangements that provide partnerships between industries and utilities can be
particularly useful in such circumstances.
Biomass pre-treatment may open a global biomass trade similar to that of coal. But key to the
expansion of biomass for electricity production will be available tax incentives, RPS mandates,
and GHG-reduction directives that externalize the costs of using fossil fuels. For example,
torrefied biomass may contribute to offsetting fossil-fuel emissions, but consideration of carbon
credits for the char field application will be required to expand its worldwide use.
In summary, biomass is dispatchable renewable power. It can come in solid or gaseous form, and
can be exploited using a wide variety of technologies to supply electricity to internal customers
or to the grid. Moreover, its carbon neutral or negative characteristics position it to potentially
contribute to the target of de-carbonizing electricity production.

4.9 References
1. Walker, J. 1966. Hopewell Village: The Dynamics of a Nineteenth Century Iron-Making
Community.University of Pennsylvania Press, Philadelphia.
2. Szego, G.C. and C.C. Kemp. 1973. Energy Forests and Fuel Plantations. CHEMTECH. May.
3. Bethel, J. et al. 1976. The Potential of Lignocellulosic Material for the Production of
Chemicals, Fuels, and Energy. National Academy of Sciences, Washington, D.C.
4. Henry, J.F. 1979. The Silvicultural Energy Farm in Perspective, in Progress in Biomass
Conversion, Vol. 1. (Sarkanen, K.V. and D.A. Tillman, eds.) Academic Press, New York.
pp. 215256.
5. Walsh, M.E. et al. 2000. Biomass Feedstock Availability in the United States: 1999 State
Level Analysis. Oak Ridge National Laboratory, Oak Ridge, TN.
6. SCS Engineers. 2002. Economic and Financial Aspects of Landfill Gas to Energy Project
Development in California. California Energy Commission. Sacramento, CA. 500-02-020F.
7. Johnson, D.K., D.A. Tillman, B.G. Miller, S.V. Pisupati, and D.J. Clifford. 2001.
Characterizing Biomass Fuels for Co-firing Applications. Proc. 2001 Joint International
Combustion Symposium. American Flame Research Committee. Kuai, Hawaii.
September 912.
8. Tillman, D.A. Final Report: EPRI-USDOE Cooperative Agreement: Co-firing Biomass with
Coal. Contract No. DE-FC22-96PC96252. EPRI, Palo Alto, CA: 2001. 104601.
9. Jenkins, B.M., L.L. Baxter, T.R. Miles, Jr., and T.R. Miles. 1996. Combustion Properties of
Biomass. Proc. Biomass Usage for Utility and Industrial Power. An Engineering Foundation
Conference. Snowbird, UT. April 28May 3.
10. Miles, T.R., et al. 1993. Alkali Slagging Problems with Biomass Fuels. Proc. First Biomass
Conference of the Americas. Aug. 30Sep. 2. Burlington, VT, pp. 406421.
11. Tillman, D.A. Biomass Cofiring: Field Test Results. EPRI, Palo Alto, CA: 1999. TR-113903.
4-101

Biomass Electricity Generation

12. Payette, K., T. Banfield, T. Nutter, and D. Tillman. 2002. Emissions Management at Albright
Generating Station through Biomass Cofiring. Proc. 27th International Technical Conference
on Coal Utilization and Fuel Systems. Clearwater Florida, March 47.
13. Hus, P.J. and D.A. Tillman. 2000. Cofiring multiple opportunity fuels with coal at Bailly
Generating Station. Biomass and Bioenergy 19(6):385394.
14. Tillman, D.A. 1994. Trace Metals in Combustion Systems. Academic Press, San Diego, CA.
15. Opportunity Fuel Cofiring at Allegheny Energy: Final Report. EPRI, Palo Alto, CA: 2004.
1004811.
16. Campbell, A.G. 1990. Recycling and Disposal of Wood Ash. TAPPI. Sep., pp. 141145.
17. EPRI Alternative Fuels Database. EPRI, Palo Alto, CA: 1996. TR-107602.
18. Johnson, D.K., et al. 2002. Report to Foster Wheeler on Pyrolysis of Urban Wood Waste and
Fresh Switchgrass. Private communication. Pennsylvania State University, University Park,
PA.
19. Tillman, D.A. 2002a. Petroleum Coke as a Supplementary Fuel for Cyclone Boilers. Proc.
27th International Technical Conference on Coal Utilization and Fuel Systems. Clearwater
Florida, March 47.
20. Shafizadeh, F., and W. DeGroot. 1976. Combustion Characteristics of Cellulosic Fuels, in
Thermal Uses and Properties of Carbohydrates and Lignins (Shafizadeh, F., K. Sarkanen, and
D. Tillman, eds.). Academic Press, New York, pp. 118.
21. Shafizadeh, F. and W. DeGroot. 1977. Thermal Analysis of Forest Fuels, in Fuels and
Energy from Renewable Resources (Tillman, D., K. Sarkanen, and L. Anderson, eds).
Academic Press, New York, pp. 93114.
22. Broido, A. 1976. Kinetics of Solid-Phase Cellulose Pyrolysis, in Thermal Uses and
Properties of Carbohydrates and Lignins (Shafizadeh, F., K. Sarkanen, and D. Tillman, eds.).
Academic Press, New York, pp. 1936.
23. Baxter, L.L. et al. 1996a. The Behavior of Inorganic Material in Biomass-Fired Power
Boilers: Field and Laboratory Experiences. Proc. Biomass Usage for Utility and Industrial
Power. Engineering Foundation Conference. Snowbird, UT. April 28May 3.
24. Baxter, L.L. et al. 1996b. The Behavior of Inorganic Material in Biomass-Fired Power
BoilersField and Laboratory Experiences: Vol II of Alkali Deposits Found in Biomass
Power Plants. SAND96-8225 Volume 2 and NREL/TP-433-8142.
25. Miller, S.F., B.G. Miller, and D. Tillman. 2002. The Propensity of Liquid Phases Forming
During coal-Opportunity Fuel (Biomass) Cofiring as a Function of Ash Chemistry and
Temperature. Proc. 27th International Technical Conference on Coal Utilization and Fuel
Systems. Clearwater Florida, March 47.
26. Landfill Methane Outreach Program, U.S. EPA:
http://www.epa.gov/lmop/overview.htm#methane.
27. Miller, B.G., S.F. Miller, C. Jawdy, R. Cooper, D. Donovan, and J. Battista. 2000. Feasibility
Analysis for Installing a Circulating Fluidized Bed Boiler for Cofiring Multiple Biofuels and
Other Wastes with Coal at Penn State University. Second Quarterly Technical Progress
Report. Work Performed Under Grant No. DE-FG26-00NT40809.
4-102

Biomass Electricity Generation

28. Miller, S.F. and B. G. Miller. 2002. The Occurrence of Inorganic Elements in Various
Biofuels and its Effect on the Formation of Melt Phases During Combustion. Proc.
International Joint Power Generation Conference. Phoenix, AZ,. June 2427.
(IJPGC2002-26177).
29. Bryers, R.W. 1993. Analysis of a Suite of Biomass Samples. Foster Wheeler Development
Corporation, Livingston, NJ. Report FWC/FWDC/TR-94/03.
30. Bush, P.V. et al. 2001. Evaluation of Switchgrass as a Cofiring Fuel in the Southeast.
USDOE Cooperative Agreement No. DE-FC36-98GO10349. Southern Research Institute,
Birmingham, AL.
31. Harding, N.S. 2002. Cofiring Tire-Derived Fuel with Coal. Proc. 27th International
Technical Conference on Coal Utilization and Fuel Systems. Clearwater, Florida,
March 47.
32. Tillman, D.A. 2002b. Cofiring Technology Review. Report to USDOENETL. Pittsburgh,
PA.
33. Kitto, B. and D. Green. 2003. Electricity Generation Using Digester Gas at Clean Water
Services. Proc. 28th International Technical Conference on Coal Utilization and Fuel
Systems. Clearwater, Florida, March 58.
34. Green, A. 2003. Opening Remarks. International Conference on Co-Utilization of Domestic
Fuels. Gainesville, Florida. February 56.
35. Tillman, D.A., A.J. Rossi, and W.D. Kitto. 1981. Wood Combustion: Principles, Processes,
and Economics. Academic Press, New York.
36. Assessment of Biogas-Fueled Electric Power Systems. EPRI, Palo Alto, CA: 2004. 1009450.
37. Haq, Z. 2002. Biomass for Electricity Generation. U.S. DOE Energy Information
Administration.
38. Perlack, R.D., et al. Biomass Fuel from Woody Crops for Electric Power Generation. Oak
Ridge National Laboratory, Oak Ridge, TN: 1995. ORNL-6871.
39. Campbell, J.E., Lobell, D.B., et al. Greater Transportation Energy and GHG Offsets from
Bioelectricity than Ethanol. Science. 2009.
40. Georgia Pacific. 2009. Processed Engineered Fuel-Effective Waste to Energy Solution.
41. A. Maciejewska. 2006. Cofiring of Biomass with Coal: Constraints and Role of Biomass Pretreatment. European Commission EUR 22462 EN.
42. Engineering and Economic Evaluation of Renewable Energy Technology. EPRI, Palo Alto, CA:
2007. 1012726.
43. Engineering and Economic Evaluation of Biomass Power Plants, 100% Biomass
Repowering, Biomass Cofiring, and Bubbling Fluidized Bed Biomass Combustion. EPRI,
Palo Alto, CA: 2010. 1019762.
44. REN21 2011, Renewables 2011 Global Status Report (Paris: REN21 Secretariat).
45. Annual Energy Review 2011, U.S. Energy Information Administration (EIA): August 2011.
DOE/EIA-0384(2009).

4-103

Biomass Electricity Generation

46. Engineering and Economic Evaluation of Biomass Gasification Power Plants. EPRI, Palo
Alto, CA: 2012. 1023994.
47. Raskin, N. et al. Power Boiler Fuel Augmentation with a Biomass Fired Atmospheric
Circulating Fluid-Bed Gasifier, Bio Bioener, 20(6), 2001, pp. 471-481.

4-104

MUNICIPAL SOLID WASTE

5.1 Introduction
Municipal solid waste (MSW) is an unavoidable by-product of human activities. Across the
globe, countries are grappling with how to manage the waste stream of human existence
properly. The tonnage of glocal post-recycling MSW in urban centers is estimated at
approximately 1.2 billion tons. Of this amount, 16% is landfilled in modern regulation landfills,
67% is disposed of in traditional dumps without methane recovery and about 16% is combusted
in waste to energy (WTE) facilities. Table 5-1 gives an overview of MSW technologies and
Figure 5-1 outlines waste management practices for several countries.
Table 5-1
MSW overview
Installed Capacity
(est.)

2,681 MW in United States (2010)

Technology
Readiness

The core MSW technologies are mature: mass burn, refuse-derived fuel/processed
engineered fuel (RDF/PEF) cofiring, landfill gas (LFG).
Advanced thermal conversion technologies, (that is, pyrolysis, plasma arc) are in
the late demonstration/early deployment stages
MSW pretreatment and MSW-natural gas hybrid cycles are still within the research
and development stages.
High confidence in cost estimates and projections for more traditional combustion
systems.
Low confidence in cost estimates and projections for advanced systems .

Environmental
Impact

Atmospheric and solid emissions from combustion and gasification technologies


controllable using commercial technology.
Furthter emissions performance data is needed for advanced conversion
technologies.
It has been estimated that one ton of MSW combusted rather than landfilled
reduces greenhouse gas emissions by one ton of carbon dioxide, on the average.

Economic Status

Competitive or near-competitive in some markets with some renewable energy


sources when compared on an LCOE basis.
Capital cost ($/kW) is more expensive than other renewable technologies.

Policy Status

Not consistently addressed by state RPS or energy policies; environmental


policies/programs are fragmented and sometimes conflicting.

Trends to Watch

Greater expansion of MSW technologies in foreign countries, especially China and


India.
Continued research and development in advanced thermal applications and its
emissions.
Environmental Protection Agency (EPA) to provide greater guidance in terms of
MSW fuel characterization and emissions limits.

5-1

Municipal Solid Waste

Figure 5-1
MSW management by country

There are currently 87 WTE plants in the U.S., equating to 2.7 GW of electricity generation from
MSW. This does not include facilities for district heating. See Figure 5-2 for a map of facilities
in the United States. In the United States, adversed converstion technologies have been slow to
be deployed and most MSW is landfilled. This is due largely to common misperceptions that
continue to foster unfavorable political conditions and negative public opinions, in addition to
the high up-front capital costs. However, as utilities look for renewable resources to supply
future generation needs, WTE plants have the benefits of being both baseload generation and a
solution to a growing waste management issuea feature that many other renewables lack.
According to previous EPRI analysis and modeling, modern MSW plants could substantially
increase in the deployment in the next 5 to 10 years. This is also true around the world as
countries look to for reduced greenhouse gas emissions. When MSW is processed at a modern
incinerator or advanced energy recovery plant rather than landfill, emissions of methan-rich gas
attributatiole to the decomposition of buried organic materials are avoided. TO the extent that
WTE capapacity relieds on carbon-neutral organic materials, it offsets the CO2 from fossil power
plants. Also, recycling the metals recovered at WTE plants avoids the emissions associated with
the mining and processing of virgin metals.

5-2

Municipal Solid Waste

Although the United States has been rather slow to pursue MSW options, many countries in
Europe and Asia have integrated resource management plans in their energy supply portfolio.
Europe currently has more than 400 WTE factilties to support and EU landfill direction of a 65%
reduction in landfilling of biodegraeable MSW. In addition, China has set a target of meeting
30% of the countrys electricity needs by 2030. India is currently working to determine how
WTE plants can fit in managing their waste stream.

Figure 5-2
States with WTE facilites
Source: Energy Recovery Council, October 2010

There are several WTE technologies that are available and others that are still in the development
phase. The fuel characteristics will define the technology that should be pursued. For instance,
due to the high moisture content in Indian domestic waste, anaerobic digestion is a better
candidate than incineration for a WTE facility. Other WTE technologies include mass burn, RDF
cofiring, landfill gas utilization, pyrolysis, gasification, plasma arc, aerobic composting and
chemical decomposition. WTE technologies are at varying stages of development and
commercial maturity.

5.2 Overview of the Regulatory and Legislative Environment


MSW is processed and energy is generated through a wide range of thermal, biological, and
chemical conversion pathways. Facilities have the potential for interacting with the environment
through the discharge of gaseous emissions, wastewater, and residues. In the interest of
preserving human and environmental health and safety by minimizing these discharges, the U.S.
Environmental Protection Agency (EPA) regulates facilities through permitting programs,
certification of some products as renewable fuels, monitoring and compliance review, delegation
of certain regulatory responsibilities to states, and continued oversight.

5-3

Municipal Solid Waste

5.2.1 Permit Overview


All facilities that handle solid waste are required on both federal and state directive to obtain
permits that hold the facility accountable for their operations and air emissions. Legislation on
the federal level has granted authority to states and the EPA to permit facilities. Table 5-2
displays a summary of permits and review processes to which a facility handling solid wastes
may become subject. The following subsections will provide a brief overview of the air and solid
waste permits.
Table 5-2
Permits and review processes
Permit or Review
Process

Facilities Subject

Permitting Authority

Prevention of
significant
deterioration

Required for new major emission sources and major


facility modifications in attainment areas

USEPA

Non-attainment NSR

Required for new major sources or major sources


making a major modification in a non-attainment area

USEPA

Minor source permits

Required for new minor sources

Title V

Any major source, solid waste incineration units,


municipal solid waste landfills, and others39

State agencies

State solid waste


permits

All facilities handling solid wastes

State agencies

USEPA/state
agencies

5.2.2 Air Emission Permits


Solid waste incinerators (which EPA refers to as municipal waste combustors) and other waste
conversion facilities that have the potential to interact with their environments through the
discharge of gaseous emissions, are regulated under the federal Clean Air Act, originally passed
by Congress in 1963 and updated in 1967, 1970, 1977, 1990, 1995, and 1998. Numerous state
and local governments have enacted similar legislation, either implementing federal programs or
filling in locally important gaps in federal programs. EPA may implement and enforce the
requirements or EPA may delegate such authority to state or local regulatory agencies. Clean Air
Act Sections 111 and 112 allow EPA to transfer primary implementation and enforcement
authority for most of the federal standards to state, local, or tribal regulatory agencies. In general,
the EPA allows state or local agencies to implement more stringent permit requirements or
regulation than those they have established. However they do not delegate to state or local
agencies the authority to make decisions that are likely to be nationally significant, or alter the
stringency of the underlying standard.

39 Full list available in Code of Federal Regulations Title 40, Sec. 70.3 Applicability

5-4

Municipal Solid Waste

A new source review (NSR) will be required for a new municipal waste combustor and other
conversion technologies. There are three types of NSR permitting requirements an emissionsource may have to meet. The NSR requirement will depend on its location, size of the facility,
and classification as a major or minor source (dependent on quantity of emissions). The three
types of NSR requirements are (1) prevention of significant deterioration (PSD) permit, which is
required for new major emission sources or a major emission source making a major facility
modification in an attainment area (an area of heightened emission restrictions); (2) nonattainment NSR permits, which are required for new major sources or major sources making a
major modification in a non-attainment area; and (3) minor source permits, which vary
depending on an individual states requirements and typically aim to prevent emissions that
interfere with regional attainment of National Ambient Air Quality Standards (NAAQS).40
The PSD review and permitting process requires the following:

Existing ambient air quality analysisa detailed analysis of the air quality surrounding the
facility site (at a distance defined by the individual state), which may require installing air
monitoring equipment to collect data for as long as a year.

Best available control technology (BACT) analysisan analysis of all available control
technologies for air emissions in a top down review. Analyses include economic,
environmental and energy costs for each alternative. The criterion for selection is best control
at acceptable cost.

Emission dispersion modelinga detailed analysis, using USEPA-approved models, of the


projected impact of the facility emissions on the ambient air quality, with a focus on the
highest potential impact area.

In addition to NSR and PSD review, Title V of the Clean Air Act of 1990 ensures that for all
emission sources there is facility-wide emission accounting, designed to integrate all local
allowable level restrictions and equipment permits Individual state permits are required to include
an accounting of the plant size, characteristics of the wastes handled, lists of emission sources and
rates, lists of all emission control equipment and reduction characteristics, and descriptions of all
test methods and monitoring plans to ensure continued compliance to emission standards. Each
permit is given on a five-year issue, except for municipal waste combustors, which enjoy a tenyear issue, and permits are re-opened and evaluated for new construction or altered operational
behaviors, as well as when new requirements are passed on the federal or state level. Each
individual permit is required to allow time for extensive EPA and local level public review, and
each is judicially reviewable at the state court level. If facilities are found to experience emissions
at levels higher than for which they are permitted, states are required to levy fines and facilities
are subject to civil and criminal enforcement procedures if they continue to emit above permitted
levels, or are found to be operating without appropriate Title V permitting.41

40
41

Source: USEPA, 2011.


Source: USEPA, 2011.

5-5

Municipal Solid Waste

In April 2007, the U.S. Supreme Court ruled that greenhouse gases (GHGs) are identified as air
pollutants by the Clean Air Act, and so the EPA was responsible for researching the effects of
GHGs on air quality, and in turn, public health.42 Greenhouse gases, including carbon dioxide
(CO2), methane (CH4), nitrous oxide (N2O), hydrofluorocarbons (HFCs), erfluorocarbons
(PFCs), and sulfur hexafluoride (SF6), are natural and anthropogenic gases that absorb and emit
radiation at specific wavelengths within the spectrum of thermal infrared radiation emitted by the
Earths surface, potentially trapping heat within the atmosphere.43 As a result, there has been
increased effort and interest in monitoring and reducing these pollutants. In order to incorporate
monitoring and regulation of GHGs without overburdening the permitting administrative
process, EPA has established a systematic introduction of high emission thresholds, which are
scheduled to be lowered over time to allow small emitters time to come under compliance while
immediately identifying and reducing emissions from the largest sources. This plan for
identification of sources and action against emitters in order of size is known as the Tailoring
Rule, as it tailors air quality and operational permitting requirements to the emission level
experienced.
The Tailoring Rule was finalized on May 13, 2010, and sets thresholds for GHG emissions that
define when Title V operating permits and prevention of significant deterioration (PSD) permits
are required for both new and existing facilities. The largest facilities that are responsible for
70% of all stationary-source GHG emissions will be brought under inspection during the process
of this rule's implementation. The GHG emission threshold is based upon CO2 equivalents, and is
comprised of the EPAs six different GHGs (CO2, CH4, N2O, SF6, HFCs, PFCs), weighted by
global warming potential. The Tailoring Rule is divided into four steps that are scheduled to be
implemented by 2016. As CO2 is identified as a GHG, EPA has responded to comments from a
variety of industry groups as it defines what CO2 emissions are to be included in the thresholds
established by the Tailoring Rule. Initially, no distinction was made between the emission of
anthropogenic (produced by human and industrial activity) and biogenic (naturally derived) CO2.
The EPA has, in the past, counted the methane from LFG and human aggregations of waste to be
anthropogenic, but the CO2 emitted from landfills has been considered biogenic. In the past,
biogenic emissions have not counted toward GHG emission inventories and reporting
requirements, but under the Tailoring Rule, the emission thresholds would count both biogenic
and anthropogenic emissions which would indicate an enormous increase in the apparent
emissions volumes generated. If facilities do not have a standardized and researched manner of
accurately modeling and predicting their biogenic fugitive emissions of CO2, then they would
incur great expense when required to do so. On July 5, 2011, The EPA ruled to postpone the
inclusion of biogenic CO2 for three years. As emissions of CO2 from biogenic sources will not be
counted, only large emitters that must already obtain Clean Air Act permits for other pollutants
will come under review for greenhouse gas emissions in the early years of rule implementation.
As a preemptive action to future reductions of emission thresholds, facilities that have the
capacity should consider energy efficiency and GHG emission reductions in design and
operation, and actively engage in the EPA GHG rulemaking process.44

42

Massachusetts v. EPA, 549 U.S. 497 (2007).


Source: IPCC, 2009.
44
Source: USEPA, 2011.
43

5-6

Municipal Solid Waste

5.2.3 Solid Waste Permits


Under Title 40 CFR Parts 239 through 259, states are required to issue solid waste facility
permits that meet nationwide standards (set forth in Title 40) to all facilities handling solid
wastes (including MSW, non-recycled household appliances, the residue from incinerated
automobile tires, refuse such as metal scrap, wall board and empty containers, and sludge from
industrial and municipal waste water and water treatment plants and from pollution control
facilities).45 States are also given the authority to conduct monitoring or testing to ensure that
facilities are in compliance with state permit requirements. Under authority of the Resource
Conservation and Recovery Act (RCRA), the USEPA is responsible, for the control and
monitoring of hazardous wastes from generation to disposal. To be considered as a hazardous
waste, a material must first be classified as a solid waste (garbage, refuse, sludge, or other
discarded material), and then must display hazardous characteristics and/or be listed in the EPA
register of hazardous materials. Facilities that may accept these materials must meet hazardous
waste inventory and reporting standards set forth by the EPA (including the requirement to hold
a Title V operating permit, and meet facility-specific requirements, including performance of
groundwater monitoring and emergency planning). However, if wastes are not classified as
hazardous, they are considered to be nonhazardous solid waste and regulation of their disposal
and recycling is delegated to state agencies. The tenets of regulation of both hazardous and
nonhazardous wastes may vary depending upon location, and in many cases the state regulation
of hazardous wastes may be more stringent than the federal standard.

5.3 Legislative Overview


Recent legislative action has increased the responsibility and regulatory authority given to the
EPA and to states for management of wastes, control of emissions, and implementation of new
technologies. Although RCRA is the primary federal law governing the disposal of solid
waste and hazardous waste and was first established in 1976, has been implemented throughout
the past 35 years. As of 2011, the EPA has given authority to all 50 states to implement federal
RCRA policies and 48 states (all excepting Iowa and Alaska) have developed state-specific
RCRA regulations for facility permitting to meet federally authorized standards.46
Although not all states have been required to achieve these federal permit standards, many states
have permit standards that exceed federal levels, as bills and statutes have been implemented on
a state level. In 2010, more than 1,600 bills were introduced in state legislatures, many that
addressed and proposed amendments to existing solid waste legislation. Of these bills, 235 were
signed into law in 2010, many addressing disposal methods and waste processing. Floridas state
legislature lifted a ban on yard waste disposal in landfills that extract and utilized landfill gas,
and multiple bills proposed incentives through energy tax credits or other tax credits for entities
that utilize WTE or conversion technologies.47 Although federal guidelines and requirements set
forth by RCRA have set the foundation for state solid waste facility permits and solid waste
programs, rulings made at state legislatures have, and have the continued potential to implement
further permitting and regulatory requirements on facilities handling both hazardous and
nonhazardous solid waste.
45

Source: Non-Hazardous Waste Regulations, USEPA, 2011.


Source: Authorization Status of all RCRA/HSWA Rules, USEPA, 2011.
47
Source: Environmental Industry Associations, 2010.
46

5-7

Municipal Solid Waste

5.3.1 Renewable Fuel Standards (RFS)


In addition to participation in GHG rulemaking, facilities converting MSW to energy through
thermal, biological, and chemical processes have opportunity to apply for classification as a
renewable fuel producer. In February 2010, EPA finalized the Renewable Fuel Standard 2
(RFS2) regulations, which implement new requirements of the Energy Independence and
Security Act (EISA) of 2007 and specify the volumes of biomass-based diesel, cellulosic biofuel,
advanced biofuel, and total renewable fuels that must be incorporated into energy generation in
the U.S. through 2022. The RFS2 regulations aim to increase domestic fuel production,
particularly through development of renewable fuels with low greenhouse gas emission
potentials.
To be considered as a renewable fuel, and thus eligible to receive a renewable identification
number (RIN), fuels must be produced from feedstocks that qualify as renewable biomass, and
be produced through a pathway that has been certified through the RFS2 regulations (or through
an alternative production pathway certified by petition) to ensure that the life-cycle GHG
emissions of the fuel are indeed reduced to a qualifying level. This level of reduction varies
between fuel categories, as shown in Table 5-3, which displays the standard definition of
allowable feedstocks.
Table 5-3
Fuel categories specified by Renewable Fuel Standard 2 (RFS2) regulation
Fuel Category

Definition

Lifecycle GHG Threshold

Cellulosic biofuel

Renewable fuel produced from cellulose,


hemicellulose, or lignin

60%

Biomass-based diesel

Renewable diesel if fats and oils not


co-processed with petroleum

50%

Advanced biofuel

Anything but corn starch ethanol (cellulosic


biofuels and biomass-based diesel)

50%

Conventional
renewable fuel

Ethanol derived from corn starch or any


other qualifying renewable fuel

20%applies only to fuel


produced in new facilities

Source: USEPA, 2011 Renewable Fuel Standards

Several further exceptions and restrictions are defined by the RFS2 regulations. The term
lifecycle greenhouse gas emissions indicates the aggregate quantity of greenhouse gas
emissions (including direct emissions and significant indirect emissions such as emissions from
land use changes), related to the full fuel lifecycle, including all stages of fuel and feedstock
production and distribution, from feedstock generation or extraction through the distribution and
delivery and use of the finished fuel to the ultimate consumer. Greenhouse gas thresholds are
defined as the reduction in lifecycle GHGs for a renewable fuel in comparison to the 2005
baseline gasoline or diesel that it displaces.

5-8

Municipal Solid Waste

All conventional renewable fuel facilities (domestic and international) that commenced
construction prior to EISA are grandfathered, and not required to meet the minimum 20 percent
GHG threshold. EISA restricts types of renewable fuel feedstocks and land that feedstocks can
come from, as it defines renewable biomass48 as the following:

Planted trees and tree residue from actively managed tree plantations on non-federal land
cleared prior to December 19, 2007

Animal waste material and animal by-products

Slash and pre-commercial thinnings from non-federal forestlands

Biomass obtained from certain areas at risk from wildfire

Algae

Separated yard waste or food waste, including recycled cooking and trap grease

All other feedstocks identified in Table 5-4, which displays the three methods named in the
RFS2 regulations as valid for the production of cellulosic biofuels.
Table 5-4
Cellulosic biofuel pathways for use in generating RINs
Fuel Type

Feedstock

Production
Process
Requirements

Ethanol

Cellulosic biomass from crop residue, slash, pre-commercial


thinnings and tree residue, annual covercrops, switchgrass, and
miscanthus; cellulosic components of separated yard waste;
cellulosic components of separated food waste; and cellulosic
components of separated MSW.

Any

Cellulosic
diesel, jet fuel
and heating oil

Cellulosic biomass from crop residue, slash, pre-commercial


thinnings and tree residue, annual covercrops, switchgrass, and
miscanthus; cellulosic components of separated yard waste;
cellulosic components of separated food waste; and cellulosic
components of separated MSW.

Any

Cellulosic
naphtha

Cellulosic biomass from crop residue, slash, pre-commercial


thinnings and tree residue, annual covercrops, switchgrass, and
miscanthus; cellulosic components of separated yard waste;
cellulosic components of separated food waste; and cellulosic
components of separated MSW.

Fischer-Tropsch
process

Source: USEPA, 2011 Renewable Fuel Standards

48

Source: USEPA, 2011 Renewable Fuel Standards

5-9

Municipal Solid Waste

Any producers of biofuels that wish to pursue RINs, yet do not utilize one of the pathways or
feedstocks displayed in Table 5-4, have opportunity to petition for a RIN assignment. Although
the cellulosic components of separated MSW are named as approved feedstocks for renewable
fuel generation, raw MSW that has not been separated would currently require petition for a RIN
assignment. The requirements of the petition process include full documentation of economic
and environmental impacts of the use of the source material, identification of known quantities
and projections of future quantities of source material, as well as full analysis of processing and
conversion technologies.
5.3.2 Renewable Portfolio Standard
The renewable portfolio standard (RPS) requires production of energy from renewable sources to
increase, and obliges utility companies to produce a percentage of their energy from renewable
sources (depending upon specific state requirements and definitions). The entities affected are
most commonly investor-owned utilities and electric service providers, although some states
have included provisions for municipal utilities to voluntarily participate. As of May 2011, 26
states had adopted an RPS and an additional six had adopted a renewable portfolio goal.
Currently, states with RPS requirements have mandated that between 4% and 33% of all
electricity is generated from renewable sources by a specified date.49 Currently, MSW has been
considered as a renewable fuel (and eligible for meeting renewable fuel portfolio standards) in 23
states, shown in Table 5-5.
Table 5-5
States defining MSW as a renewable fuel eligible to meet renewable portfolio standards
States Defining MSW as a Renewable Fuel Eligible to
Meet Renewable Portfolio Standards
(as of 7/5/2011)
California

Massachusetts

Oregon

Connecticut

Mississippi

Pennsylvania

District of Columbia

Minnesota

Puerto Rico

Florida

Missouri

South Dakota

Hawaii

Nevada

Utah

Iowa

New Jersey

West Virginia

Maine

Ohio

Washington

Maryland

Oklahoma

Source: DSIRE Portfolio Standards/Set Asides for Renewables & Efficiency, 2011

49

Source: DSIRE, 2011.

5-10

Municipal Solid Waste

The requirements for RDF utilization as a renewable fuel varies between states. For example, in
California it is not always considered renewable if produced through direct combustion, and in
other locations only a certain percentage of renewable sales may be met through MSW
feedstocks. In Hawaii, Minnesota, and South Dakota, MSW may be classified as biomass.
Although MSW may not be an eligible feedstock for renewable generation in every state and
through every conversion pathway, energy recovered from MSW through mass burn and refusederived fuel (RDF) combustion within WTE processes has been defined as renewable in 27
states (including the District of Columbia), shown in Table 5-6.
Table 5-6
States defining waste-to-energy as renewable in state law
States Defining Waste-to-Energy as Renewable in
State Law (as of 10/1/2010)
Alaska

Maine

Oklahoma

Arkansas

Maryland

Oregon

California

Massachusetts

Pennsylvania

Connecticut

Michigan

Puerto Rico

District of Columbia

Minnesota

South Carolina

Florida

Nevada

South Dakota

Hawaii

New Hampshire

Virginia

Iowa

New Jersey

Washington

Indiana

New York

Wisconsin

Source: Energy Recovery Council (ERC), 2010

In addition to individual state laws designating energy generated through WTE processes as
renewable, the federal government has issued numerous policies, statutes, and regulations
throughout the past thirty years supporting this designation, shown in Table 5-7.
Although MSW may not be considered as a renewable feedstock in all states, landfill gas derived
from the decomposition of MSW may be much more widely used to meet renewable targets.
Landfill gas is named as an eligible feedstock in all states except Indiana (although energy
generated from organic waste biomass may be considered on a case-by-case basis in Indiana).

5-11

Municipal Solid Waste


Table 5-7
Federal statutes and policies defining WTE as renewable (as of 10/1/10)
American Recovery and Reinvestment Act of 2009
Energy Policy Act of 2005
Federal Power Act
Public Utility Regulatory Policy Act (PURPA) of 1978
Biomass Research and Development Act of 2000
Pacific Northwest Power Planning and Conservation Act
Internal Revenue Code (Section45)
Executive Orders 13123 and 13423
Federal Energy Regulatory Commissions Regulations
(18CFR.Ch.I, 4/96Edition, Sec.292.204)
Source: Energy Recovery Council (ERC), 2010

5.4 Technology Considerations for Using MSW


Since the beginnings of its utilization in the 1890s, the burning of MSW with energy recovery
(now known as WTE) has matured into a safe, effective and environmentally acceptable
technology, although it often continues to face economic challenges and local political
opposition. Recently, as federal and state governments have established programs that support
energy independence and promote the use of fossil fuels alternatives, the application of WTE as
a source of renewable energy has gained popularity with many policymakers. This support and
the introduction of possible incentives for use of renewable energy have encouraged
consideration of a variety of energy technologies that utilize MSW.
With mass burning, the most common WTE technology, direct combustion of solid waste at high
temperatures occurs with minimal preprocessing, and heat is recovered in the boiler. Combustion
conditions may vary depending upon the type of stoker/grate utilized, the energy form, electricity
or combined heat and power, is dependent upon the boiler type and energy market addressed.
The basic types of direct MSW combustion technologies include (1) waterwall combustion,
where excess air is introduced to ensure complete burning occurs and waterwall boiler tubes
capture energy generated in the boiler, (2) refractory line-incinerators with excess air and water
heat boilers, and (3) starved air combustion systems, which are more often used in smaller
applications, as a smaller quantity of air is used and a greater quantity of ash is produced per
input ton.
As an alternative to mass burning, MSW may be mechanically processed (shredded, screened,
etc.) to produce refuse-derived fuel (RDF). RDF co-firing with coal was tested by the U.S.
Environmental Protection Agency in St. Louis at Union Electrics Meramec facility in 1972 and
seasoned in 1975. RDF can be fired (1) in a dedicated boiler with 100% RDF firing, or (2) the
RDF can be co-fired with other solid fuel, such as coal or wood. Either dedicated or co-firing
boilers may be operated as suspension boilers where combustion occurs in suspension, semisuspension (or stoker) boilers, where some combustion occurs both in suspension but most on
5-12

Municipal Solid Waste

a traveling grate, or in a fluidized bed where combustion occurs on a bed of sand at the bottom of
a cylindrical furnace. The MSW processing requirements for RDF vary depending on the boiler
style to be used, as suspension boilers, for example, require a smaller particle size then a
fluidized bed.
Another long-term approach to energy recovery from MSW is the tapping into landfill-based
disposal sites to capture the landfill gas (LFG) generated by the decomposition of MSW. Since
1975, with the beginning of operations at the Palos Verdes Landfill LFG Energy Facility in
Rolling Hill Estates, California,50 many landfills have utilized the gases extracted from landfills
to fill demands for energy on-site and at a distance, as an alternative to flaring. According to the
EPA Landfill Methane Outreach Program (LMOP), there are currently more than 500 projects
throughout the United States that generate more than 1,560 MW from LFG, and there are 530
landfills that have the potential to participate in LFG to energy programs.51 LFG may be used onsite or transported via pipeline.
In recent years, the emergence of advanced thermal conversion technologies, such as pyrolysis,
gasification, and plasma arc, have presented additional possibilities for the conversion of waste
into energy and valuable fuel products such as biodiesel or ethanol. Through heating and reacting
waste with oxygen or air to produce a synthetic gas (syngas) that then has the possibility of being
co-fired, these technologies have demonstrated effective at facilities in Europe, Asia, and
Canada. There are multiple configurations of gasifiers and pyrolysis systems, differentiated by
the use of wet or dry feed, gas flow direction, and use of air or oxygen, and these systems are
capable of converting 70% to 85% of the carbon in the feedstock to syngas. Though versions of
these technologies have been used by the electric power industry to generate energy throughout
the world from fossil fuels for over the past 35 years, with nineteen gasification plants in the
United States,52 there are no gasification or plasma arc facilities utilizing MSW as a feedstock
currently operating in the United States.
Other technologies employing MSW for energy recovery include aerobic digestion, anaerobic
digestion, and chemical decomposition. In these methods of energy recovery from MSW, liquefied
and slurried waste feedstocks undergo biological and chemical processes, to be converted into
useful chemical feedstocks, fuels, oils, fertilizers, and other valuable products. Aerobic digestion
utilizes biological processes to decompose the organic fraction of MSW through the use of
microbes that thrive in oxygen-rich environments, producing stable solids that can be used as
agricultural compost. Similarly, anaerobic digestion occurs in oxygen-poor environments,
producing biogas and stable solids (also often used as compost). Through chemical decomposition,
waste is liquefied and processed to promote depolymerization, and ultimate outputs include fuels,
oils, and many chemical products. MSW digestion and decomposition processes have been used
throughout the world to process a wide variety of biomass feedstocks, though no facilities are
commercially operating in the United States to process MSW.

50

National Renewable Energy Laboratory, Managing Americas Solid Waste, 1998


USEPA, 2010
52
Source: World Gasification Database; Gasification Technologies Council
51

5-13

Municipal Solid Waste

5.5 Commercially Available Technologies


The following sections will provide an overview of the current status and maturity of several
MSW conversion technologies or more commonly referred to as waste to energy. The
technologies addressed herein are mass burn, RDF co-firing, landfill gas (LFG) utilization,
syngas co-firing and hybrid cycle, and advance thermal technologies.
5.5.1 Mass Burn Technologies
In mass burn waterwall combustion, MSW is placed by an overhead crane directly into the feed
chute of the system for incineration with no pre-processing, except for the possible removal of
large non-combustible items (refrigerators, washing machines, microwave ovens, etc.) which may
have been collected in the trucks and discharged in the trash loads at the facility. Waste is gravityfed down the chute onto a moving grate at the bottom of a combustion chamber in a furnace with
walls built of water tubes welded together to provide a flat wall, as shown in Figure 5-3.

Figure 5-3
Waterwall furnace section53

Depending on the design, potentially half the heat generated from the burning waste is absorbed
by the waterwalls and the balance heats water flowing through the other components of the
boiler (evaporator, super heater and economizer), as shown in Figure 5-4, or heats the process air
via an air preheater.

53

Source: Babcock and Wilcox.

5-14

Municipal Solid Waste

Figure 5-4
Typical mass-burn waterwall system54

54

Source: Fairfax County, VA.

5-15

Municipal Solid Waste

The off-gas exiting the boiler passes through an air pollution control (APC) system, where
pollutants are removed, and is discharged through a stack to the atmosphere. Waste is burned to
an ash, called bottom ash, in the furnace. Heat extracted from the waterwalls and the boiler
sections generate steam, which, in most facilities, is directed to a turbine generator for electric
power production. Waterwall systems generally have unit sizes of 200 tons per day (TPD) up to
750 TPD, and multiple units are used when higher waste disposal capacity is required. Much of
the equipment is field-erected requiring extended contracting schedules of 28 to 32 months.
5.5.1.1 Typical Mass Burn Technical Parameters
Mass burn waterwall systems are the most efficient of the mass burn technologies in terms of
waste burn-out and in energy generation. They have been built in unit sizes up to 750 TPD.
Through use of multiple units facilities can be built to handle 3000 TPD or more (such as the
four-750 TPD unit plant located in Fairfax County, Virginia). Depending on final design
configuration, typical waterwall boilers operate in the range of 700 to 860 psig and operate at
temperatures in the area of 700 to 830F. Facilities operating at these pressures and
temperatures can generate approximately 680 and 2,000 MWh of electricity each day, through
the use of condensing steam turbine generators. Each facility is coupled with air pollution control
(APC) equipment, including activated carbon injection systems for the reduction of mercury and
dioxins, Ammonia injection, flue gas scrubbers with lime injection, fabric filter baghouses for fly
ash collection, and continuous emissions monitoring (CEM) systems, among other facilityspecific controls.55
5.5.1.2 Mass Burn Experience and Vendors in the United States
No new greenfield mass-burn WTE facilities have been built in the United States for more than
twenty years, although recently there have been several new procurements and expansions of
existing facilities to add additional units. Also, there have been acquisitions where ownership
and operator changes at certain existing facilities have changed. As a result, the majority of firms
associated with mass-burn WTE are either operators or owners/operators of existing facilities. As
shown in Table 5-8, Covanta and Wheelabrator own and operate the majority of privately owned
mass-burn WTE facilities.

55

Source: Covanta, 2011.

5-16

Municipal Solid Waste


Table 5-8
U.S. operating mass-burn/waterwall facilities and vendors
Entity

Owned

Operated

Covanta Energy Corporation

14

30

Veolia Environmental Services

Wheelabrator Technologies, Inc.

10

16

Public

28

Other

Total

54

54

Source: Energy Recovery Council (ERC), 2010 Directory

Some of the mass burn facilities were designed by American firms with their proprietary
technology, such as Detroit Stoker, Combustion Engineering (now Alstom), and Babcock &
Wilcox, but the majority of these existing systems are of European grate design. The two leading
suppliers of WTE grate systems in the United States and overseas are The Martin Company of
Germany and Von Roll of Switzerland, represented in the U.S. by Covanta and Wheelabrator,
respectively.
Although there were no WTE facility procurements in the United States throughout the late 20th
and early 21st centuries, there have been recent expansions and procurements for new facilities,
which indicates a re-emergence of these facilities and the positive environmental impacts they in
comparison to those of a landfill. Table 5-9 summarizes the recent expansions and procurements
of new facilities in the North America. Although the U.S. market may have been stagnant,
throughout this same time period the demand for these facilities has increased in Europe and in
Eastern Asia. European and Japanese technology suppliers are actively marketing their systems,
and they have been consistently improving both their energy production and environmental
performance. The waterwall technology is mature and is used more than any other for large WTE
facilities in the United States and overseas.

5-17

Municipal Solid Waste


Table 5-9
Recent WTE expansions and procurements in North America
Facility

Location

TPD

Status

Expansion1
Hillsborough County Resource Recovery
Facility

Florida

600

Operating

Lee County Resource Recovery Facility

Florida

630

Operating

Honolulu Resource Recovery Venture

Hawaii

900

Under construction

Olmsted County Waste-to-Energy Facility

Minnesota

200

Operating

Frederick and Carroll Energy Recovery


Facility

Maryland

1,500

Design and permit

Palm Beach County Renewable Energy


Facility No. 2

Florida

3,000

Design and permit

Durham York Energy Center

Ontario, CN

470

Design and permit

New Procurements

Note: 1. Tons Per Day (TPD) represent only the expansion capability, not the overall facility processing capabilities

5.5.1.3 Modular Starved Air Combustion Systems


Starved air combustion is another process where unprocessed MSW is fed into a refractory lined
chamber using a ram loader past a firewall. The waste is dropped into a fixed volume opening
and ram-fed into the primary chamber on a continuous batch basisevery seven to 10 minutes,
for example. The primary chamber of the combustion unit includes a series of charging rams
which push the burning waste from one level to another until the residue drops out of the main
chamber at the opposite end into an internal wet ash pit, as in Figure 5-5.
Less than the ideal amount of combustion air is injected into the primary combustion chamber,
and the gas from the burning waste does not fully burn out at this location. The gases are directed
to a secondary combustion chamber where additional overfire air is added to complete the
burning process. Hot gases pass though a separate waste heat boiler for steam generation, and
then through an air pollution control system, before discharge up the stack to the atmosphere.

5-18

Municipal Solid Waste

Figure 5-5
Typical modular combustion system
Source: Consutech Systems; Richmond, VA

A major advantage of this system is operations in the primary combustion chamber occur at substoichiometric airflows, creating more of a pyrolysis type of operation. With less air, the fans can
be smaller and the chamber itself can be smaller than with other systems. Also, with less air
flow, high and uniform temperatures, and turbulent mixing of gases throughout the second
chamber, less particulate matter (soot) and organic contaminants enter the gas stream, resulting
in the air pollution system being sized for a smaller load. The off-gas exiting the boiler then
passes through an the APC system, much like the waterwall boiler, though what is required to
process and treat the off-gas may be less extensive than what is needed in the case of a waterwall
boiler system.
These systems are referred to as modular systems, because they are factory built and can be
brought to a site for final assembly of the modules and set up in a relatively short period of time,
for example, 18 to 24 months. Multiple units, or modules, are used to process larger amounts of
MSW. They are less efficient than waterwall units in waste burnout and in energy generation and
when 24/7 operation is required, are typically built in unit sizes of 120 to 150 TPD. Through use
of multiple modular Enercon refractory-lined modular combustor units, facilities such as
Pittsfield Resource Recovery Facility in Pittsfield, Massachusetts; Pioneer Valley Resource
Recovery Facility in Agawam, Massachusetts; and Covanta Wallingford Energy-from-Waste in
Wallingford, Connecticut are each able to process between 240 and 420 TPD of MSW. The
Pittsfield Resource Recovery Facility produces more than 400 million pounds of steam at 220
psig/ 540F and 3.5 million KW of electricity annually, whereas the Pioneer Valley and
Wallingford facilities produce 9.4 MW and 11 MW of electricity, respectively, through the use
of condensing steam turbine generators, achieving temperatures and pressures as high as 750F
and 700 psig. Modular facilities are coupled with APC systems that are similar to mass burn
facilities, and include activated carbon injection systems for the reduction of mercury and
dioxins, flue gas scrubbers with lime injection, fabric filter baghouses for fly ash collection, and
continuous emissions monitoring (CEM) systems, among other facility-specific controls.56
56

Source: Covanta, 2011.

5-19

Municipal Solid Waste

Modular systems are also used for smaller WTE facilities and industrial applications. There are
only a small a number of American firms supplying such systems in the United States, and are
competitive in overseas markets as well. The more active of these suppliers are Consutech
Systems (formerly Consumat) of Richmond, Virginia; Enercon Systems, Inc. of Elyria, Ohio;
and Basic Environmental Engineering of Chicago, Illinois. They have each been supplying
smaller-scale incineration systems for MSW and other wastes for more than 25 years. As shown
in Table 5-10, most of the modular units in service today are publicly owned and operated.
Table 5-10
Operating modular system facilities and vendors
Entity

Owned

Operated

Covanta Energy Corporation

Veolia Environmental Services

Wheelabrator Technologies, Inc.

Public

Other

Total

12

12

Source: Energy Recovery Council (ERC), 2010 Directory

5.5.2 RDF Co-Firing Technologies


In addition to direct combustion in mass burning WTE plants, MSW has the potential to be
processed into refuse-derived fuel (RDF) and then fired into dedicated boilers or co-fired with
wood waste, coal, or other fuels. Boilers that co-fire coal utilize other varieties of renewable
biomass, often sources derived from agricultural waste and wood waste. However, the definition
of renewable biomass is expanding in multiple states, to include MSW and its products,
including RDF. As previously stated, MSW is considered to be a renewable fuel in 23 states, and
is explicitly defined as biomass in Hawaii, Minnesota, South Dakota, and Maryland.
Of the 147 biomass power facilities currently operational within the United States, five are
reported to co-fire biomass feedstocks (including wood waste, RDF, forest residue, mill
residue and used railroad ties) with coal.57 The installed capacities of these facilities range from
26.5 MW to 585 MW, and one facility additionally employs a combined heat and power system.
The facility that has the capacity to co-fire RDF with coal is the 10 MW Ames Municipal Utility
Power Plant. The facility, located in Ames, Iowa and expanded to cofire RDF in 1975, is
operated by Ames Municipal Electric Utility. Boilers at this facility utilize pulverized coal as
their primary fuel, and RDF is blown into the boiler below the coal inlet as a supplemental fuel.
The RDF is air-classified shredded MSW, produced at the Ames Resource Recovery Plant. The
power plant burns 35,000 tons of RDF annually, and the heat generated from the RDF (which
has a nominal heating value of 6,200 btu/lb of RDF) makes up 10% of the plants annual fuel
requirement. Of the 70 facilities utilizing MSW as a direct feedstock and the four additional
facilities utilizing RDF, The Ames facility is the only one to co-fire with coal at a typical ratio of
57

2011 Biomass Power Map (Biomass Power Magazine)

5-20

Municipal Solid Waste

approximately 10% RDF to coal by mass of incoming feed.58 Of the 49 proposed biomass
facilities and proposed expansions of existing facilities, two are planned to co-fire biomass
feedstocks (both woody biomass). Although not planned to co-fire with coal, one facility will
utilize MSW, however no facilities are planned to utilize RDF. Table 5-11 provides an overview
of the facilities co-firing with coal.
Table 5-11
Biomass power facilities co-firing with coal
State

Biomass
Feedstock

Installed
Capacity (MW)

Cofiring
Biomass
w/Coal

SD Warren Westbrook

ME

Forest residue

65

Yes

Rapids Energy Center

MN

Mill residue,
forest residue

26.5

Yes

ML Hibbard

MN

Mill residue,
forest residue,
railroad ties

72.8

Yes

Ames Municipal Power Plant

IA

RDF

10

Yes

Virginia City Hybrid Energy Center

VA

Wood waste

585

Yes

Cleco Corp. Madison Unit 3

LA

Woody biomass

585

Yes

Hu Honua Bioenergy LLC

HI

Woody biomass

24

Yes

Oneida Seven Generations Energy

WI

MSW

No

Facility Name
Existing Facilities

Planned Facilities

5.5.2.1 RDF Dedicated Boiler


In RDF systems, MSW is mechanically processed in a front end system to produce a more
homogenous fuel. RDF, in its simplest form, as produced by the illustrative process shown in
Figure 5-6, is shredded MSW with ferrous metals removed. However, based on historical
experience with this type of fuel dating back to the 1970s, additional processing must be applied
to the incoming waste stream to remove other noncombustible materials such as glass, rocks,
other inerts, non-burnables, and aluminum. Additional screening and shredding stages can be
placed in the processing line to further enhance the RDF.
The unique feature of RDF systems is in the pre-processing of waste. As seen in the diagram of a
typical RDF processing facility in Figure 5-6, MSW enters the facility and plastic bags of MSW
are broken open in a pre-trommel. Materials then pass through a trommel where smaller fractions
are extracted, and the remaining waste then goes through a shredder. The materials pass through
a secondary trommel, and a magnetic separator removes ferrous metals. The balance of the
material is fired in the furnace.
58

Source: Ames Municipal Power Plant, 2011.

5-21

Municipal Solid Waste

Figure 5-6
Simplistic RDF processing facility

Other configurations may include additional separating equipment, or may not use any trommels,
but the RDF generated is always shredded, so that it is capable of being blown into a furnace.
Although results vary with the processing configuration and unless more aggressive materials
recovery is desired as part of the RDF processing system, in general about 80% of the incoming
MSW stream is converted into RDF for the thermal process. With more aggressive materials
recovery processing for cardboard, plastics, ferrous and non-ferrous metals, and inerts recovery,
the RDF fraction could be 50% to 60%.
For illustration purposes, Figure 5-6 shows that the RDF produced is blown into the furnace from
the left, above the grate. What does not burn in suspension (above the grate) will burn on the
grate, and the hot gases generated will pass through a waterwall section and then a boiler section.
This system is similar to the mass-burn waterwall facility except in the nature of waste charging
and burnout.
An overview of the RDF process is given in Figure 5-7.

5-22

Municipal Solid Waste

Figure 5-7
Typical RDF combustion facility
Source: Energy Answers Corporation

5-23

Municipal Solid Waste

An advantage of this RDF production system is the removal of metals and other recoverable or
inert materials from the waste stream. Although not all these facilities include this step in the
RDF processing line, those that do can realize revenue from the sale of recovered metal. For
instance, at the North County Resource Recovery Project in West Palm Beach, Florida, the
nominal 3,000 TPD RDF processing facility removed and sold over 36,000 tons of ferrous
metals in 2004, which represented over 3 percent of the weight of the incoming waste stream.
With the removal of non-combustibles, the specific heat content per pound of RDF can be
increased by 10% over the original MSW, with the ash and moisture content also reduced, as
shown in Table 5-12.
Table 5-12
Comparative fuel properties of RDF
Type of Fuel

Heating Value of RDF


As Received (J/g)

Heating Value of RDF


As Received (Btu/lb)

Moisture
Content (%)

Ash
Content (%)

RDF

12,000 to 16,000

5,160 to 6,880

15 to 25

10 to 22

Coal

21,000 to 32,000

9,030 to 13,760

3 to 10

5 to 10

MSW

11,000 to 12,000

4,730 to 5,160

30 to 40

25 to 35

Source: CalRecovery, Inc.

When RDF is combusted in existing utility boilers, it is often fired in one of two boiler
configurations, suspension fired boilers or semi-suspension (stoker) boilers. For use in
suspension firing boilers which may have no bottom grates, RDF must be finely processed,
usually below 0.75 inch, greatly increasing the cost of RDF production. Alternatively, the
installation of dump grates at the bottom of furnaces if size reduction cannot be performed
effectively to achieve an acceptable burnout. Although semi-suspension firing, where RDF is
much coarser (and can be of sizes up to six inches) does not always introduce the same
requirements for boiler alteration or increased processing, many potential RDF users have found
challenges to storing, handling, and feeding RDF of this size. As a result, several RDF users have
opted to pelletize or dandify RFD as an alternative to using it in its loose fluff form.
5.5.2.2 RDF Fluidized Bed
For fluidized bed combustion, MSW is shredded to less than 4 inches mean particle size using an
RDF process similar to that previously described to produce the fuel. The RDF is blown onto a
bed of sand at the bottom of a vertical cylindrical furnace, as shown in Figure 5-8.

5-24

Municipal Solid Waste

Figure 5-8
Fluidized bed
Source: Energy Products of Idaho, Coeur dAlene, ID

Hot air is also injected into the bed from below, and the sand has the appearance of a bubbling
fluid as the hot air agitates the sand particles. Moisture in the RDF is evaporated almost
instantaneously upon entering the bed, and organics burn out both within the bed and in the
freeboard, the volume above the bed. Steam tubes are embedded within the bed and a transverse
section of boiler tubes captures heat from the flue gas exiting the furnace; an Energy Products of
Idaho (EPI) system is shown in Figure 5-9. EPIs fluidized bed system in La Crosse, WI
(owned/operated by Xcel Energy) is fueled by RDF and hogged waste wood. By permit, this
facility has a fuel mix of 50% RDF and 50% wood by Btu content. The plant was first fired with
RDF in 1987 and the La Crosse County-Xcel agreement, initially for 20 years, has been extended
through 2023. It consists of two 14-MW units, and approximately 73,000 tons per year (TPY) of
MSW is processed into RDF on-site by Xcel Energy. The wood is procured from a nearby thirdparty processor. The Xcel Energy boilers operate at 450 psig/750F.

Figure 5-9
Typical RDF fluid bed system
Source: Energy Products of Idaho, Coeur dAlene, ID

5-25

Municipal Solid Waste

Fluid bed incineration is more efficient than grate burning-based incineration systems. The fluid
bed is very effective in waste destruction and requires less air flow than mass-burn or modular
systems. The fluid bed, however, does require relatively uniform sized material and removal of
certain slagging materials, therefore RDF preparation is necessary. It is required for operation of
the fluidized bed, not, as with the above systems utilized due to the inherent benefits associated
with materials recovery.
An RDF/gasification/incineration technology, similar to that described earlier, is manufactured
by Ebara Corporation of Tokyo. Ebara has four such systems in operation for MSW and
industrial wastes in Japan, ranging in facility size from 185 TPD to 460 TPD. Its version of the
fluid bed system described previously is called a fluidized-bed gasifier, illustrated in Figure 5-10.

Figure 5-10
RDF fluidized bed gasification system
Source: Ebara Corporation, Tokyo, Japan

The fluidized-bed gasification system exports a burnable gas. RDF is first prepared using a
process similar to the ones illustrated in Figures 5-6 and 5-7. The RDF (called Wastes in
Figure 5-10) is then charged into the fluid bed and the gases generated are directed to a
combustion chamber with molten slag dropping out to a water-cooled sump. The molten slag
solidifies into a glass-like material that can be used as a construction material or fill. Heat from
the gas fired in the combustion chamber is captured in hot water tubes in the boiled to generate
steam which can be used for electric power generation. Without the generation of a usable gas
stream, and with the necessity of a combustion chamber for gas burnout, this system is an
incinerator.

5-26

Municipal Solid Waste

5.5.2.3 RDF Experience and Vendors in the United States


As with mass-burn systems, there have not been any new RDF systems constructed in the
United States in the past decade. The 15 firing RDF plants currently in operation, as shown in
Table 5-13, are operated by some public entities and contractors, including Xcel Energy,
Covanta Energy Corporation, and the Babcock & Wilcox Power Generation Group, Inc.
Table 5-13
U.S. RDF facilities
Entity

Owned

Operated

Covanta Energy Corporation

Babcock & Wilcox Power Generation Group, Inc.

Xcel Energy

Public

Other

Total

15

15

Source: ERC 2010 Directory


Note: Figures include both dedicated and fluidized bed boiler technologies.

Equipment used in RDF/stoker and suspension firing boiler technology is adapted from
equipment provided in coal-fired electricity generation plants, and there are many established
U.S. system and equipment suppliers, such as Foster Wheeler, Riley Power Inc. (a Babcock
Power Inc. company, formerly Riley Stoker Corp.), and Babcock and Wilcox. Some of the
earliest 100 percent fired RDF-boiler systems were installed in the late 1970s through early
1980s by Combustion Engineering, Inc. (now Alstorm). These units in Detroit, Honolulu, and
Hartford, CT are still in operation.
There are several RDF/fluid bed systems operating in Europe (particularly in Scandinavia, where
a number of fluid bed incinerator manufacturers are located). In the United States, fluidized bed
combustion using RDF as a fuel include French Island, WI, owned and operated by Excel
Energy of Minneapolis and Tacoma Washington Municipal Utility. The equipment was supplied
by Energy Products of Idaho in Coeur dAlene, the only U.S. firm currently manufacturing fluid
bed furnaces for RDF firing. Other U.S. firms, Foster Wheeler, Babcock & Wilcox, and others,
have provided fluidized bed units utilizing coal, rice hulls, and other feedstocks.
5.5.3 LFG Utilization Technologies
The concern for the release of methane into the earths atmosphere from MSW landfills has
drawn the attention of regulators and landfill owners/operators in the United States and
internationally for more than 30 years. LFG generation starts from the time when waste is first
put in place and continues for 20 or more years after the landfill is closed. Landfills today
generally need to have active gas collection and control system (GCCS) to control odors and
comply with increasingly more-stringent air-quality regulations. A simplified LFG collection
system is depicted in Figure 5-11.

5-27

Municipal Solid Waste

Gas to Flare or Energy Recovery

Figure 5-11
Simplified landfill gas collection system
Source: USEPA

The number of landfills with a GCCS has increased significantly due to the requirement for
landfills to have a Title V operating permit, and the need for landfill operators to comply with
other air permitting rules and regulations. The USEPA reports that there are about 2,300
landfills59 currently operating in the United States.
There are about 560 LFG energy projects listed as partners with the USEPA in the Landfill
Methane Outreach Program (LMOP), and about 500 other landfills with the potential to turn
their landfill gas into energy. Table 5-14 shows a summary the number of projects and types of
conversion technology, for the projects participating in LMOP. Many more MSW landfills
actively collect and combust LFG with flare system to reduce odors, and maintain compliance
with state and federal regulations. Sites with a consistent gas quality and quantity represent
potential opportunities for recovering energy.
Table 5-14
Summary of LMOP projects, by conversion technology
Conversion Technology

Number of Projects

Comments

Electricity Generation

407

33 of these involve Cogeneration

Direct Heat

129

14 of these are for leachate evaporation

LFG to Pipeline

24

LFG is converted to High Btu

Total

560

Source U.S. EPA Landfill Methane Outreach Program (LMOP), 2011

The three most common options for converting LFG to energy are (1) electricity generation,
(2) direct heating and use by an industry, and (3) treated LFG to a pipeline. The feasibility of
these projects depends on the market value of the product generated, LFG quality and LFG
quantity. These options are described in the following text. As more landfills construct and
operate collection systems, the opportunities for recovering energy from LFG have also increased.
There is a dual benefit in using LFG to produce usable energy because greenhouse gas emissions
are reduced and the need to use nonrenewable sources of energy, such as coal, oil and natural gas,
is reduced.

59

USEPA Landfill Methane Outreach Program-www.epa.gov/outreach.

5-28

Municipal Solid Waste

5.3.3.1 LFG to Electricity


As illustrated in Figure 5-12, the conversion of LFG to electricity (LFGTE) is the most common
option used at landfill sites throughout the United States. Electricity generation projects make up
more than two-thirds of the currently operating energy recovery facilities on landfills.60
Electricity from these facilities can be used on the landfill site, where the gas is produced to
reduce or offset electrical power costs, or it can be sold to the electrical grid.

Figure 5-12
Landfill gas collection and electricity production layout
Source: Enerdyne Power Systems

There are several different technologies used to convert the energy from LFG, including internal
combustion (IC) engines, turbines, micro-turbines, and fuel cells. Table 5-15 presents several
examples of LFGTE projects of varying sizes. As represented here, sites with relatively small gas
flows typically use micro-turbines. Sites with relatively high gas flows have the option to use
either steam turbines, which can combust gas at a high flow rate, or several IC engines.

60

USEPA Landfill Methane Outreach Program-www.epa.gov/outreach.

5-29

Municipal Solid Waste


Table 5-15
List of representative electricity-generating projects
Project Name
Franklin County Landfill
SWACO

Location
Grove City, OH

Generating Capacity
300 KW
Micro-turbine

LFG Flow Rate


0.15 mmscfd

Independent Hill Landfill

Prince William, VA

1.9 MW IC Engines

0.7 mmscfd

Springfield Sanitary Landfill

Willard, MO

3.0 MW-IC Engines

1.58 mmscfd

Orange County Sanitary


Landfill

Orlando, Orange
County, FL

12.4 MW

Coyote Canyon Sanitary


Landfill

Newport Beach, CA

Puente Hills Landfill

Whittier, CA

Steam turbine
21.0 MW
Steam turbine
50 MW-Steam turbine

5.76 mmscfd
7.0 mmscfd
33.1 mmscfd

Source: USEPA Landfill Methane Outreach Program (LMOP) Landfill/Project Database


mmscfd = million standard cubic feet per day

The majority of LFGTE projects use IC engines produced by companies such as Caterpillar or
GE Jenbacher. The evaluation presented in this report will use the IC engine technology as an
example project. One of the most common sizes of an IC engine is the 1MW unit. An engine
with this rating needs about 500 scfm of treated LFG to operate at full generating capacity. he
combination of a gas processing unit, an IC Engine, the generator and a transformer is commonly
called an engine/generator set. Figure 5-12 shows a schematic diagram of gas extraction from a
landfill, gas processing to remove moisture, the engine/generator set and electrical connection to
the power grid. During the development of an LFGTE project using IC engines, it is necessary to
evaluate the historical deposition rate of the MSW, the gas flow rates over time and select an
engine type and size that is suitable for the gas flow rate. A picture of a typical IC
engine/generator set is shown on Figure 5-13.

Figure 5-13
Typical LFG engine/generator set
Source: Caterpillar EnginesSuffolk, VA project

5-30

Municipal Solid Waste

Gas flow rates on landfills vary over time. Depending on gas flow rates, it is common to develop
a project where multiple units are installed. Using a modular or incremental approach, the
amount of electricity produced can be adapted to changes in gas flow rates. Once a project is
established, additional engines can be added as the gas flow rates increase. Similarly, IC engines
can also be removed as flow rates decrease. For example, three 1-MW Jenbacher units were
initially installed at the Winnebago County Snell Road Landfill in 1999. Thereafter in 2007, a
1.06-MW combined heat and power (CHP) unit was installed.
When using LFG in an IC engine to generate electricity, LFG must be processed or treated to
remove water, particulate matter and other trace constituents. The extent of processing depends
on the type of engine used to burn the LFG. Figure 5-12 shows a moisture separator at the inlet
side of the LFG extraction blower. It is common practice to pre-treat LFG by flowing it through
a moisture separator built into the LFG collection system. Additional GCCS treatment may need
to be done to generate gas suitable for certain engines depending on the manufacturer. Figure 512 shows the gas flowing through a condenser and chiller before entering the engine generator
set for conversion to electricity.
The gas treatment process is done to remove water and other impurities. This process involves
compressing and cooling the LFG, as shown in Figure 5-12. Particulates and other trace
constituents such as hydrogen sulfide (H2S), non-methane organic compounds (NMOCs), and
siloxanes that cause silicate deposits in engines are partially removed during the treatment
process. Certain engines need cleaner fuel than others. The decision to select a certain type or
manufacturer of an IC engine is based on weighing the cost of replacement and maintenance that
would need to be done, as compared to the cost for doing additional treatment of the gas.
5.5.3.2 LFG as Direct Heat Source
When an industry utilizing natural gas or other non-renewable fuels in its manufacturing process
is located near a landfill, LFG may be sold to that industry to offset their need for other
traditional fuel sources. LFG can be used as a direct heat source in a boiler, dryer, kiln,
greenhouse or other thermal application. LFG can also be burned and used to evaporate leachate.
LFG is being used as a heat source in the auto manufacturing, chemical production, food
processing, pharmaceuticals, cement and brick manufacturing, and consumer electronics
industries.
When selling gas for use in a direct heat application, the gas must be pretreated at the landfill to
remove water and other trace constituents. As shown in Figure 5-12, pre-treatment in a moisture
separator is done as part of the LFG collection system. Because the pre-treatment only removes part
of the water and other constituents, additional processing or treatment of the LFG is usually required
to make it suitable for use. The treatment process generally involves compressing and dehydrating
the LFG, as shown in Figure 5-14. It is also necessary to remove H2S and volatile organic
compounds (VOCs), which is accomplished with this treatment process. Figure 5-15 shows a picture
of an LFG compression system designed to compress, clean, and deliver processed gas to an end
user at a specified pressure and quality. This type of unit can also be adapted to remove CO2 to
increase the heating value of the gas.

5-31

Municipal Solid Waste

Figure 5-14
LFG cleaning diagram with CO2 wash
Source: Acrion Technologies, Inc.

Figure 5-15
Landfill gas compression and cleaning unit
Source: Lectrodryer, LLC

The size of a direct-heat source project depends primarily on the amount of gas generated at the
landfill, and secondarily on the amount of gas the industrial end-user is willing to accept. The
overall project involves cleaning the gas, compressing the gas at a facility located on the landfill,
and introducing the gas into a pipeline that connects the landfill to the industry. After the gas has
been collected and treated at the landfill, it is compressed and injected into the pipeline so it
meets the pressure specified by the end-user when it arrives at the point of discharge at the
industry. It is usually necessary to construct the pipeline connect from the landfill to the industry
as part of the overall project. The USEPA reports that there are about 129 landfill projects61 in
the United States where treated LFG is delivered to industries that burn the gas in their own
boilers or other similar heating devices. A list of several example direct heat projects is presented
below in Table 5-16.
61

USEPA Landfill Methane Outreach Program-www.epa.gov/outreach

5-32

Municipal Solid Waste

Because the flow of LFG is relatively constant throughout the day, industries often use the LFG
to keep their boilers or other heating units warm during times when production is not at full
capacity. The treated LFG can provide a low-cost source of energy that flows at a relatively
consistent rate.
Table 5-16
List of representative direct heat projects
Project Name

Location

Project Type

LFG Flow Rate

Dane County Landfill #2

Madison, WI

Alternative Fuel

0.028 mmscfd

SPSA Regional Landfill

Suffolk, VA

Industrial Boiler

1.0 mmscfd

Blue Ridge Landfill

Fort Bend, TX

Industrial Boiler

4.32 mmscfd

Burlington Co. Landfill

Bordentown, NJ

Greenhouse

0.216 mmscfd

Source: USEPA Landfill Methane Outreach Program (LMOP) Landfill/Project Database


mmscfd = million standard cubic feet per day

5.5.3.3 LFG to Pipeline


As generated, LFG is generally considered a low-Btu fuel. The average energy content may
range from 400 Btu/scf62 to 500 Btu/scf as collected. However, it can be converted to a high or
medium Btu fuel and pumped into a natural gas pipeline. The process of creating high Btu gas
from LFG (approximately 950 Btu/scf) requires the removal of CO2. Similar to the other types of
conversion technology projects, water and other constituents such as VOCs, H2S, and particulate
matter must also be removed from the LFG. CO2 and other constituents are removed by
compressing and cooling the gas to a temperature where the CH4 and CO2 molecules can be
separated. CO2 removal can also be removed using one of several different chemical processes.
When CO2 has been removed, the processed gas has a higher heating value. Figure 5-14 shows a
gas cleaning process where CO2 is removed after water and H2S are removed.
A list of several example projects where high-Btu gas is being produced is shown in Table 5-17.
The USEPA reports that there are about 24 high-Btu gas projects63 participating in the LMOP.
Table 5-17
List of representative high-Btu projects
Project Name

Location

Project Type

LFG Flow Rate

City of Fort Smith Landfill

Fort Smith, AR

High Btu to Pipeline

3.3 mmscfd

Carter Valley Landfill

Church Hill, TN

High Btu to Pipeline

1.44 mmscfd

Cedar Hill Regional Landfill

Maple Valley, WA

High Btu to Pipeline

10.6 mmscfd

Source: USEPA Landfill Methane Outreach Program (LMOP) Landfill/Project Database


mmscfd = million standard cubic feet per day

62
63

Btu/scf = British thermal units per standard cubic foot of gas.


USEPA Landfill Methane Outreach Program-www.epa.gov/outreach.

5-33

Municipal Solid Waste

Once the treated gas has been processed to meet the pipeline companys specifications, it is
compressed and injected into the pipeline at a location approved by the pipeline company.
According to the definition for Pipeline Natural Gas in 40 CFR 72.2, the treated gas must have
at least 70% methane by volume or have a gross calorific value between 950 and 1100 Btu per
cubic foot, and contain less than 0.5 grains of total sulfur per 100 standard cubic feet of gas. The
extent of gas treatment needed on a project will depend on the pipeline companys required
specifications, and the quality and quantity of available LFG.
Payment for the gas is usually based on measurements of Btu. Some projects convert LFG to
either compressed natural gas (CNG) or liquefied natural gas (LNG) for use as fuel in trucks or
other applications where these types of fuel are needed.
5.5.4 Syngas Co-Firing and Hybrid Cycles
A combustible gas can be produced from organic compounds, including coal and those in solid
waste, directly through various processing systems. The resulting gas product is referred to as
syngas (from synthetic gas or synthesis gas) In the late 1800s, syngas was produced by the
gasification of coal for lighting and industrial use. During WWII, Germany produced large
quantities syngas from coal, which was further processed into diesel and aviation fuels. When
the feedstock used to produce syngas is recently living organisms or products made from them,
the syngas is termed biogas. The feedstocks, biomass, used to produce biogas include
agricultural residues, wood, solid waste and crops grown for feedstock, such as switch grass. It
can be directly produced from biomass feedstocks by a variety of methods and productions
techniques. Landfill gas, discussed in the previous section, is an indirect method of biogas
production from solid waste.
The current research and development activities are concentrated on the production of biogas
from biomass and/or solid wastes containing biomass. The focus on biomass for biogas
production is because it is carbon neutralthat is, the CO2 released in the biomass utilization is
recaptured in growing the biomass. Once produced, biogas can be used as an intermediate in
producing synthetic petroleum for use as a fuel or lubricant via the FischerTropsch process,
similar to what Germany did in World War II.
Biogas consists of a mixture of gases including methane, hydrogen, carbon monoxide, and
carbon dioxide. It varies in its energy density, from 50% of the energy density of natural gas to
15%, depending on the feedstock and the gas production technology. Biogas is combustible and
may be used as a fuel in internal combustion engines or as an intermediate for the production of
other chemicals.
The composition of biogas varies depends upon the nature of the feedstock and the production is
what produces LFG, which typically has an energy density, typically measured as Btu/scf, of
approximately half that of natural gas. Biogas is made up of a combination of gases, as shown in
Table 5-18.

5-34

Municipal Solid Waste

Moisture or water is endemic to organic feedstocks and can cause difficulty in the production of
biogas. Also, the moisture produced in the gas production can become part of the biogas, which
reduces the energy density of the gas. Practical and cost-effective technologies to remove
moisture are available. In addition to moisture, biogas may contain a number of trace
constituents that cause problems in combustion systems including inorganic silicon-based
particles, tars, and low-melting-point metals such as aluminum and magnesium.
Table 5-18
Primary constituents of biogas
Compound

Chemical Symbol

Percentage Range

Methane

CH4

5575%

Carbon Dioxide

CO2

3045%

Carbon Monoxide

CO

Trace

Hydrogen

H2

01%

Nitrogen

N2

05%

Sulfur Dioxide

H 2S

12%

Water

H 2O

010%

Source: Arizona Energy, 2009

Production of biogas can be accomplished utilizing three mains technical approaches:


gasification, pyrolysis and anaerobic digestion. These techniques generally require some
feedstock preparation prior to biogas production, such as size reduction and inert removal.
This is particularly applicable to heterogeneous feedstocks such as MSW.
Co-firing of biogas in existing coal fired boilers is in the pilot and demonstration phase. This
practice is referred to as indirect co-firing of biomass, as contrasted with direct co-firing of biomass.
In Europe, Japan, Taiwan, and Korea a number of indirect co-firing projects have been conducted.
In Taiwan, research was conducted on a coal fired boiler, 154,000 lb/hour steam production,
replacing 5% to 25% of the heat input with biogas made from paper sludges. The biogas used had
a low heat content ranging from 3.99 to 6.24 MJ/Kg (megajoules/kilograms). The results varied
with the location of the port feeding the biogas and there was a significant reduction in NOx
emissions when firing biogas.
Foster Wheeler developed a biogas co-fired pulverized coal power plant for the town of Lahti in
Finland. The circulating fluidized bed gasifier utilizes a variety of biomass including bark, wood
chips, saw dust and RDF. The power plant has a steam production of 275 lb/s, produces
electricity and district heat to power and heat the town. The area around Lahti has an abundance
of biomass and only small modifications were required to co-fire the biogas. The environmental
advantages were found to include lower SOx and NOx emissions. A similar plant was built by
Foster Sheeler for Electrabel in Ruien, Belgium (shown in Figure 5-16). Foster Wheeler built a
commercial plant co-firing biogas produced from MSW and plastics in 2001 in Varkaus,
Finland, in which 50 MW of the energy production came from the co-fired biogas.
5-35

Municipal Solid Waste

Figure 5-16
Electrabel Plant, Ruien, Belgium
Source: Vergokan Group

Amergas in Amer, Netherlands has developed a gasifier, Lurgi design, firing demolition wood.
The biogas produced is co-fired in Unit 9 of EPZs Amer power station, which has a net capacity
of 600 MW and 350 MWth (Megawatt Thermal). The use of biomass saves the combustion of
70,000 TPY of coal. The biogas provides output equivalent to 29 MW, or 26 MW plus 15
MWth64 or about 5% of the facilitys energy. The power consumption of the gasifier is 0.75 MW.
The net efficiency of wood conversion is 34%, or 31% for electricity production and 18 percent
for heat in cogeneration mode. The product gas is cleaned before being fired in the existing
boiler, which has to comply with stringent SO2 and NOx-emission standards, as well as providing
a certified fly ash. Although Amergas started operations in 2000, several problems arose during
commissioning, with major modifications scheduled for completion by mid-2002. The capital
costs for the Amer gasification plant cost approximately $36.2 million (in 2000 US$).
The SilvaGas (previously FERCO) system for production of biogas was operated at 350 TPD dry
of wood at the McNeil facility in Burlington, Vermont, beginning in 1997, with the biogas
successfully co-fired in the wood combustion boiler. The initial work had US DOE funding but
additional support of full integrated gasification combined cycle (IGCC) implementation
(including gas cleaning and a new high efficiency gas turbine to replace the boiler) did not occur;
the existing plant proved to be uneconomic for electricity production, and was shut down in
2001. FERCO also failed to raise further capital with disputes between investors, and filed for
bankruptcy in November 2002. SilvaGas has planned several larger commercial plants for some
time (again only for heat and power production), but construction has not started.

64

Source: Coal Online, Cofiring-coal-with-other-fuels.

5-36

Municipal Solid Waste

In the United States, there is at least one project that have co-fired LFG in existing coal fired
boilers. The landfill gas is produced by anaerobic digestion and averages approximately 500 Btu
per cubic foot, after condensate removal. The Los Reales Landfill owned by the City of Tucson
produces LFG which is captured for co-firing. The LFG is transported to the Tucson Electric
Power, Sundt Generating Station via a 3.5-mile pipeline. It is co-fired in the Unit #4 boiler
replacing approximately 20,000 tons of coal per year.
Co-firing biogas can reduce the carbon footprint of an existing coal fired plant and will utilize
local biomass and waste materials. The gasification system will allow a facility to utilize local
carbon-neutral fuels and waste streams resulting in reduced environmental impact of the plant,
while meeting energy needs of the community. It provides the additional benefit of providing
disposal for those wastes used as feedstock for the biogas production.
Co-firing biomass via a gasification process holds great promise from an environmental
perspective. NOx, SO2, CO2, and heavy metals emissions can be reduced through displacement of
coal with cleaner biogas while improving the performance and reducing operational risk for the
existing boiler when compared to direct co-firing of biomass. In general the proposed solution
would represent a low cost alternative for NOx abatement with the incentives of green power
production when compared to other solutions available. Co-firing of biogas in existing coal fired
boilers has the advantage over the co-firing of RDF. These include a much lower tendency for
fouling and slagging and subsequent problems in the superheaters with the resulting drop in
efficiency. Also, by not introducing biomass ash into the boiler eliminates the problems
associated with ash removal and utilization. By co-firing of biogas produced from MSW and
other wastes, a number of problems related to deterioration of ash quality may be avoided.
5.5.5 Advanced Thermal Conversion Technologies
There are many technologies currently being proposed for the treatment and disposal of MSW
throughout the world. Most of these involve thermal processing, but some others comprise the
biological or chemical decomposition of the organic fraction of the waste to produce useful
products like compost or energy products, notably gas for downstream combustion or use as
chemical feedstock.
Thermal processing refers to a number of different types of technologies utilizing heat as the
mode of waste treatment. Currently there are more than 100 thermal processing firms, including
developers, offering gasification, pyrolysis, plasma arc, and anaerobic digestion technologies.
(Note: pyrolysis, or starved air combustion, was mentioned earlier in Section 5.5.1.3 associated
with smaller modular incineration systems.)
5.5.5.1 Pyrolysis
In pyrolysis, an organic waste, such as MSW, is heated without oxygen (or air), similar to the
generation of coke from coal or charcoal from wood. This generates a gas, char and inorganic
residue. The gas is burned out in a gaseous phase, requiring much less oxygen than incineration.
Metals, glass and other inorganic residues will usually melt at the temperatures within the
pyrolysis chamber and will be discharged as a black gravel-like substance, termed frit.
Advantages of this process are in the lack of air entering the chamber and the resulting smaller
size of system components. Without air, there is little nitrogen oxide generation, and low
particulate (soot) formation. There have been many attempts to develop this technology outside a
laboratory or a pilot plant.
5-37

Municipal Solid Waste

In demonstrations in the 1970s, it was difficult to maintain a sealed chamber to keep air out, and
waste variability created problems in maintaining consistent operation. When the pyrolysis gas is
fired in a combustion chamber that is part of the system, the system is classified as an
incinerator, and facilities were required to obtain the appropriate permits as incinerators.
Between March 1975 and February 1977, the city of Baltimore utilized pyrolysis to process
30,000 tons of solid waste, but ceased operation after numerous unsuccessful attempts to achieve
even partial capacity sustained operation. The plant was redesigned and began operating again in
1979, and was finally closed in 1980, as its emissions far exceeded permit limitations.65 Through
the 1980s, permit requirements became continuously more stringent and while facilities were
able to install air quality control equipment at great expense to meet emission standards, the
byproducts of pyrolysis at that time could not be treated easily, and often were tested to be above
permit limits for low volatile metals. Although pyrolysis has been used throughout the world to
process a variety of feedstocks, including plastic wastes and biomass, and treatment methods
have been developed for byproducts, limited performance data is available for pyrolysis systems
that process hazardous wastes that may contain dioxins and PCBs and the treated byproducts
from hazardous waste pyrolysis may still require special handling and disposal.66 Currently, there
are no full-scale pyrolysis systems in commercial operation using MSW in the United States.
A 50-TPD pilot demonstration system began operating in southern California in 2005. Shown in
Figure 5-17, it was built and is operated by International Environmental Solutions (IES) of
Romoland, CA. The process shreds MSW down to a uniform size capable of feeding into the
thermal converter, or pyrolysis chamber. The pyrolysis gas generated is fired in a secondary
combustion chamber, or thermal oxidizer, and passes through a waste heat boiler for heat
recovery. Char drops out the bottom of the pyrolysis chamber for disposal or further processing
for recovery of metals and other constituents. Although this system is marketed as a pyrolysis
system, a combustion chamber is necessary for its operation (for destroying organics in the offgas) and the presence of this chamber classifies the system as an incinerator.

65

Source: Gagliardo.Baltimores Resource Recovery HistoryPyrolysis Demonstration to Proven Mass-Burn


Technology. Rockville, Maryland: Government Institutes, Inc. 1984, p. 513.
66
Source: Center for Public Environmental Oversight, 2008.

5-38

Municipal Solid Waste

Figure 5-17
Process diagram of a pyrolysis system
Source: International Environmental Solutions, Inc., Romoland, CA

5.5.5.2 Gasification
Gasification is the heating of an organic waste, such as MSW, to produce a synthesis gas, syngas,
which consists primarily of hydrogen, carbon monoxide, carbon dioxide, and some trace
compounds. It varies in its heating value from 200 to 500 Btu per cubic foot. It can be used as
fuel or as feedstock or for production of other chemicals.
Whereas pyrolysis systems are primarily focused on waste destruction, a gasifier is designed
primarily to produce a usable gas. As shown in Figure 5-18, Thermoselect, a European firm
represented in the U.S. by Interstate Waste Technologies (IWT) of Malvern, PA, has developed a
system composed of 400-TPD modules processing MSW.

5-39

Municipal Solid Waste

H2, CO, CO2, H2O

Wastes

Synthesis Gas
Production of

High
Temperature
Reactor

1,200C
Quench

Synthesis gas scrubbing

Hydrogen
Methanol
Ammonia
or

Power
generation

Scrubber

Sulfur
Process water
treatment

Degassing Channel

Press

Clean water
Salt

O2

Homogenization reactor
Zinc Concentrate
2000C

1600C

Metals and
Minerals

Oxygen generation
facility

Figure 5-18
Typical gasification system
Source: Interstate Waste Technologies, Malvern, PA

Waste is fed into a gasification chamber to begin the heating process, first having been
compressed to remove entrapped air. Some oxygen, sufficient only to maintain the heat
necessary for the process to proceed, is injected into the reactor, where temperatures in excess of
3,000F are generated. At this high temperature, organic materials in the MSW will dissociate
into hydrogen, methane, carbon dioxide, water vapor, and so on, and non-organics will melt and
form a glass-like slag. The gas is cleaned, water is removed, and it can be used for power
generation, heating or for other purposes. The glass-like slag can be processed to remove metals
and used as fill, or as a building material for roads and the like.
Seven plants with IWT technology are currently operating in Japan, with at least two of them
firing MSW. The largest of these plants in Kurashibi has a reported furnace size of 200 TPD,
with three units of this size the plant is firing up to 600 TPD of MSW.
Another gasifier marketed for MSW is built by EnTech of Devon, England. It has constructed
approximately 20 of these facilities, which are in operation on MSW in Europe and Asia, many
of which are relatively small (less than 10 tons per day), with none designed for more than 70
TPD throughout. This system generates, in addition to a salable gas (synthesis gas, or syngas),
recyclable plastics and other potential revenue streams such as metals, sulfur and salt. As shown
in Figure 5-19, MSW is classified by a combination bag breaker and gravity separator process,
termed a Kinetic Streamer.
5-40

Municipal Solid Waste

Figure 5-19
EnTech process schematic
Source: EnTech

In the EnTech system oversize materials, which are basically inorganic, are directed either to a
plastics recycler or a non-plastics recycling station, while the majority of waste (presumably
organic) is directed to a dryer to remove entrained moisture. The dryer utilizes the latent heat
inherent in the organic content of the waste to produce the heat necessary to drive the
gasification process. The syngas can be fired in a waste heat boiler for steam and subsequent
electric power production.
Two Canadian firms have advanced gasification. Enerkem, headquartered in Montreal, Quebec,
has an operating pilot gasification facility in Sherbrooke, Quebec, and a commercial
demonstration facility in Westbury, Quebec which began operation in 2009. Also, in 2008,
Enerkem signed an agreement with Edmonton, Alberta to build a 100,000 TPY facility, with
completion and commissioning scheduled for 2012. These facilities produce ethanol from the gas
using a thermal/chemical process. The Plasco Energy Group, which has a five-TPD research
facility in Spain, has operated a 100-TPD pilot plant in Ottawa, Ontario. Plasco received final
approval from the Ontario Ministry of the Environment to complete construction at a
400-TPD commercial facility for the City of Ottawa. The demonstration plant is in operation and
3600 tons of trash were processed at the facility in 2012. The final 400-TPD facility is scheduled
to be completed in 2016. Plascto was in agreement with CWMC in Red Deer, Alberta for a 200TPD facility, scheduled for completion by 2012, but this plan was scrapped because the city was
unable to guarantee enough trash to feed the plant.

5-41

Municipal Solid Waste

5.5.5.3 Plasma Arc


Plasma arc refers to the means of introducing heat into the process. Essentially a plasma arc
system is a pyrolysis or starved air process generating heat by firing the waste with a plasma
torch using electric current to produce a syngas, which is then combusted to produce steam
and/or electricity, and is typically classified as an incinerator. If the system generates an off-gas
that contains burnable gases (e.g., hydrogen and carbon monoxide) that can be used off-site, it
can be classified as a gasifier. A typical unit is shown in Figure 5-20.

Figure 5-20
Cross-section of a plasma arc furnace
Source: Westinghouse Plasma Corporation

Plasma is a collection of free-moving electrons and ions across a gas volume at reduced pressure.
The gas molecules, losing one or more electrons, become positively charged ions capable of
transporting electric current and generating heat when the electrons go into a stable state and
release energy similar to lightning in the atmosphere. Plasma can reach temperatures exceeding
7,000F. Molten slag generated by the process is about 3000F. The by-products of plasma
gasificationslag or glassy aggregate and metalsare similar to those produced in other hightemperature gasification technologies. As with other gasification technologies, plasma
gasification requires the pre-processing of the MSW feed to reduce the particle size before its
introduction into the plasma reactor. (Note: although Figure 5-20 indicates MSW enters in the
upper left side of the gasifier, the actual feedstock is similar to RDF.)
5-42

Municipal Solid Waste

One of the primary drawbacks of plasma arc technology is the huge parasitic load of the plasma
torches. Therefore, the net electric output of the conversion process, if generating electricity for
sale from the system, would be substantially reduced. There are no commercial-scale plasma arc
facilities processing MSW in the United States, although several companies are marketing some
form of this technology and proposing facilities. There are three small plasma arc facilities
processing MSW and/or auto-shredder residue in Japan reportedly using the Westinghouse
plasma reactor. Few, if any of the plasma arc pilot facilities have been able to generate a fuel gas,
and air emissions have been found to be no better than conventional incineration systems.
5.5.6 Aerobic Composting
Composting is a natural process that depends on the action of microscopic organisms to break
down organic matter. Composting has been used for hundreds of years, if not thousands, to process
a variety of agricultural wastes. There are two types of micro-organisms that digest the organic
materials: aerobic and anaerobic. The first need oxygen or air to function and the latter work
without oxygen. There are five factors that influence the composting process: (1) moisture,
(2) oxygen or air, (3) temperature, (4) chemical balance of carbon and nitrogen and (5) particle size.
Large-scale mixed waste composting facilities are industrial plants that receive MSW and grind
the material in large shredders, and remove inert materials by screening and other processes. The
feed material is then moved to the composting vessel where the organic materials are digested by
the micro-organisms. The process and factors 1 through 3 above are controlled by computer.
After initial processing through the vessel, the resulting compost product is stored to cure and
then it is ready to be sold. Using California post-recycling waste composition data,67 it is
estimated that aerobic composting would reduce the waste landfilled to 25% of the initial feed.
There would be 43% recovered as compost and material products and 32% released to the
atmosphere as gases (mainly CO2 and water vapor). The products produced from MSW
processing facility include materials: ferrous metals, nonferrous metals, and various grades of
plastic, which are recovered in front end processing. Glass may be recovered or may be
pulverized and become part of the compost product. The compost product has a low economic
value and competes with peat moss as a soil conditioner.
There are several hundred mixed waste composting plants in Europe, both aerobic and anaerobic.
BioCycle reports68 that there are 13 MSW composting facilities operating in the United States, as
shown in Table 5-19.
In 1995, there were 17 MSW composting facilities operating in the United States.69 Of these
plants only three are currently operating. In 1995, the number of technology suppliers included
Bedminster, Danaco, Ryan/OTVD, Buhler and others. Large-scale plants with daily processing
capacities of 200 TPD or greater, were built in Portland, OR, Baltimore, MD, Miami, FL, Cobb
County, GA, Sevier County, TN, Sumter County, FL and Pembroke Pines, FL, all of which
failed for technical reasons, such as odor control, or financial difficulties. A key problem has
typically been that the quality of the compost being produced was lower than expected, which
reduced the revenues and made the projects too costly and/or non-competitive with other
available waste management alternatives.
67

Statewide Waste Characterization Study, California Integrated Waste Management Board, December 1999.
BioCycle Magazine, JG Press, Inc., November 2008.
69
Municipal Solid Waste Composting A Review, Gershman, Brickner & Bratton, Inc., for Solid Waste Association of North
America and U.S. EPA, April 1995.
68

5-43

Municipal Solid Waste

The trend in composting seems to be toward segregating bio-wastes and then composting to
produce compost and/or biogas. In the United States, composting is used primarily to process
yard waste and sewage sludge (biosolids), and there are thousands of successful projects. These
are generally small units processing less than 10 dry TPD of biosolids, with two large facilities
processing more than 200 dry TPD.

5-44

Municipal Solid Waste


Table 5-19
Operating mixed waste compost facilities
Locations

State

Ownership/Operator

System

Throughput Tons/Day
(unless noted)

Gilroy

CA

Private/Z-Best Composting

Enclosed ASP (CTI Systems)

350

Mariposa County

CA

Municipal

In-vessel (SV Composter ECS)

60

Cobb County

GA

Municipal

Rotating drum/aerated windrow


(Bedminster)

200

Marlborough

MA

Private/WeCare Environmental

Rotating drum/aerated windrow


(Bedminster)

1,500-2,000 tons/month
(with bio solids)

Nantucket

MA

Private/Waste Options, Inc.

Rotating drum/aerated windrow


(Bedminster)

100 peak/15 off-peak


(with bio solids)

Truman

MN

Municipal

In-vessel (OTVD)

65

West Yellowstone

MO

Municipal

In-vessel (SV Composter ECS)

40 peak/20 off-peak

West Wendover

NV

Municipal

Rotating drum/aerated windrow

12 (with 2 tons biosolids)

Delaware County

NY

Municipal

Rotating drum with agitated bays


(Conporec/IPS-Siemens)

24,000 tons/year (with


7,000 tons biosolids)

Medina

OH

Municipal/Norton Environmental

Windrow

33

Rapid City

SD

Municipal

Rotation drum/aerated windrow


(Bedminster/Backhus turner)

250 (with 50 wet TPD


biosolids)

Columbia County

WI

Municipal

Rotating drums/windrows

80

Source: BioCycle Magazine, JG Press, Inc., 2008.

5-45

Municipal Solid Waste

Anaerobic digestion is a natural process where micro-organisms break complex organic molecules
into smaller ones and produces combustible biogas composed of methane and carbon dioxide as a
product. It has been applied to a variety of organic feedstocks including sewage sludge, animal
wastes and purpose grown plants. As applied to the processing of MSW, anaerobic digestion is a
treatment process where waste is first presorted and then fed into closed vessels. The control of the
process has several key factors including temperature, amount of moisture and number of stages
which determine the residence time to accomplish the digestion. Most common are high-moisture
processes. These use agitators, pumps, conveyors and other materials handling equipment, MSW is
hydrated and forms a slurry. Metals, glass and other constituents of MSW that have no affinity for
water are eventually discharged from the system into dedicated containers for recycling, further
processing or final disposal. The paper, garbage, soluble components,and the like generate black
water, which has a relatively high organic content. This stream is processed in a series of sealed
digesters without air where microorganisms break down the solids and generate gas containing
methane. The time in the chamber and the residence time will be sufficient to generate the gas. The
process is shown in the schematic in Figure 5-21.

Figure 5-21
Process flow for anaerobic digestion system
Source: Arrow Bio

When the moisture level is between 25% and 40%, the process is refered to as dry anaerobic
digestion. At this moisture level the prepared MSW feedstockinorganics removed and
shreddedcannot be pumped but is stacked in a closed vessel for digestion. The process is one
of plugged flow through a verticle or horiznotal vessel. The advantages of dry anaerobic
digestion is the facilities are compact and require low water and energy inputs.
All anaerobic digestions systems produce a residue at the end of digestion. This is composed of
mineral and non-organic material plus organic material that was not digested. Some organic
materials, such as lignin, do not break down in the anaerobic digestion process. This solid
residue has application as soil conditioners. Suitability would be determined by the analysis of
the residue which may have contaminents from the MSW.
This gas is rich in methane and other organics and can be burned as a fuel for heating or for
electric power generation. The solid residual from the digestion process is similar to compost and
can be used as a soil amendment. The process also separates out recyclable materials, such as
glass and metals. There are many such facilities processing sewage sludge, manure and other
homogeneous wastes.
5-46

Municipal Solid Waste

ArrowBio of Haifa, Israel, is a vendor offering to construct anaerobic digestion facilities to


process MSW in the United States. It has responded to procurements in Los Angeles and
New York. It operates a 100 TPD, full-scale MSW demonstration process line in Tel Aviv and
has a 270 TPD, commercial scale plant for MSW operating in Sydney, Australia, pictured in
Figure 5-22.
The system operates without high temperatures or pressure. In theory, it is extremely simple,
relying on non-specialized mechanical equipment (pumps, screens, macerators, tanks, conveyors,
etc.) for operation. Digestion occurs through the presence of natural micro-organisms in MSW,
so charging with specialty or unique bacteria is not necessary.

Figure 5-22
ArrowBio facility in Sydney, Australia
Source: Arrow Bio

5.5.7 Chemical Decomposition


Chemical decomposition, also referred to as depolymerization, is a process whereby waste
feedstocks are directly liquefied into useful chemical feedstocks, oils and/or gases. The oils are a
replacement for fuel oil and the gases consist of carbon monoxide, hydrogen and methane. The
process generally utilizes medium temperature and pressure to break large complex molecules
into smaller ones. If higher temperatures are employed, chemical decomposition becomes
indistinguishable from gasification.
The solid waste feedstock for chemical decomposition will generally be pre-processed to remove
recyclable and inert materials and to reduce the particle size. Moisture is favorable to the process
and may need to be added to create steam reforming reactions. The process is multi-step: gas
recovery, liquid separation to isolate the oil product, and processing the solids to separate carbon
char from inerts. Chemical decomposition processes require an external energy source to make
the reactions take place.

5-47

Municipal Solid Waste

One form of chemical decomposition is used to break cellulose into sugars for fermenting to
produce ethanol. This is the hydrolysis process, of which two types have been applied to the
organic components of solid waste: acid hydrolysis and enzyme hydrolysis. They have also been
used in combination. The National Renewable Energy Laboratory developed and has operated
pilot processes, which have demonstrated technically feasibility. No production plants, however,
have been built to date. The city of Los Angeles received nine submissions for hydrolysis
processes, including those from Arkenol and Iogen, but no hydrolysis process was selected by
Los Angeles.
The U.S. Department of Energy (DOE) is working to move cellulosic ethanol to commercial
production. In this effort, DOE selected six companies for demonstration grants in 2007,
including the following:

Abengoa Bioenergy Biomass of Kansas, LLC of Chesterfield, MOThe proposed plant will
be located in Kansas and will produce 11.4 million gallons of ethanol annually.

ALICO, Inc. of LaBelle, FLThe proposed plant will be in LaBelle, Florida and will
produce 13.9 million gallons of ethanol a year.

BlueFire Ethanol, Inc. of Irvine, CAThe proposed plant will be in Southern California and
will produce about 19 million gallons of ethanol a year.

Broin Companies of Sioux Falls, SDThe plant is in Emmetsburg, Iowa, and after
expansion, it will produce 125 million gallons of ethanol per year.

Iogen Biorefinery Partners, LLC, of Arlington, VAThe proposed plant will be built in
Shelley, Idaho and will produce 18 million gallons of ethanol annually.

Range Fuels (formerly Kergy Inc.) of Broomfield, COThe proposed plant will be
constructed in Soperton, Georgia and will produce about 40 million gallons of ethanol and
9 million gallons methanol per year.

INEOS New Planet BioEnergy Lisle, ILThe proposed plant will be constructed in Vero
Beach, Florida and will produce about 8 million gallons of fuel grade ethanol and 6 MW of
power.

Enerkem Montreal, QuebecThe proposed plant will be constructed in Edmonton, Alberta,


Canada and will produce about 10 million gallons of fuel grade ethanol.

Enerkem Montreal, QuebecThe proposed plant will be constructed in Pontotoc Mississippi


and will produce about 10 million gallons of fuel grade ethanol.

These facilities plan to use a variety of cellulosic feedstocks including: corn stover, wheat and
rice straw, wood waste, green/yard wastes, and switch grass. Vereniums 1.4 million gallon per
year cellulosic ethanol plant in Jennings, LA is the nations first operating demonstration. It uses
bagasse and other wastes as feedstocks and received DOE funding.
Microwaves can be used as the external heat source for chemical decomposition or
depolymerization. Microwave systems have been built to decompose some special wastes,
particularly tires. Goodyear obtained a patent to de-vulcanize tires and built a facility in
Lincoln, NE to process in-plant scrap in the late 1970s. Several small units have been operated

5-48

Municipal Solid Waste

on tires. The application of microwaves to drying and decomposition of various wastes,


including medical waste and nuclear waste, is proven, but its application to municipal solid waste
has not been proven but is being promoted by Molecular Waste Technologies, Inc. Global
Resource Corporation has also proposed microwave plants for MSW, but has not constructed
one.
5.5.8 Technology Summary
In summary, the use of MSW as a fuel and source of energy has been practiced for more than
100 years. Significant interest and investment have been made throughout the 1980s and the past
several years, due to increased energy costs and environmental focus. Recently, and as
previously noted, there has been an increasing interest in proven mass-burn waterwall
technology; there has also been increasing interest in alternative conversions technologies. Table
5-20 summarizes the maturity of each of the technologies previously presented in this overview.
Table 5-20
Technology maturity
Technology

Maturity

Mass Burn Technology

Mature

RDF Co-Firing Technology

Mature

Note

LFG Utilization:
LFG to Electricity

Mature

LFG to Direct Heat

Mature

LFG to NG Pipeline

Mature

Syngas Co-Firing and Hybrid Cycles

PilotLab

Advanced Thermal Conversion Technology

PilotLab

Source: GBB 2011


Notes:
1. RDF co-firing with woody waste and other biomass material, limited quantities and current experience co-firing
RDF with coal.
2. Several projects utilizing these technologies are in the process of being developed or built and are yet to be
commercially proven.

5.6 Evaluation of a 25-MW Mass Burn WTE Facility


Although MSW is a heterogeneous mix of societys waste, there is a degree of uniformity in its
properties. About half of its weight is composed of paper waste, with the remainder containing
plastics, glass, ferrous and non-ferrous metals, food waste and yard waste and has a significant
moisture content of 25% or more. A key element in its treatment and disposal is the fact that in
the United States, MSW will burn to produce useful energy. Based on typical United States
waste profiles, one ton of MSW can generate over 500 KW of electric power when properly
combusted.

5-49

Municipal Solid Waste

In this chapter, the results of an engineering and economic evaluation of a representative power
plant firing MSW is presented. This evaluation includes a mass burn 25-MW (net output)
waterwall boiler with a moving grate, fired by typical raw (unprocessed) MSW.
5.6.1 The MSW Stream
MSW includes materials discarded by households, institutions (schools, libraries, prisons, and so
on), commercial establishments (stores, offices, etc.), and other members of society with,
increasingly, a significant amount of the recyclable materials removed. This amount varies from
community-to-community due to several factors. Although MSW is heterogeneous, it is possible
to establish a typical estimate of its characteristics that can be used for its evaluation. It is
assumed that white goods (refrigerators, washing machines, etc.) will be removed from the waste
stream and waste materials such as yard wastes (grass clippings, tree trimmings, dead leaves,
etc.) will be composted or otherwise recycled rather than being comingled with MSW. On this
basis, the analysis in Table 5-21 was developed, showing MSW as a fuel compared to coal, as
well as for RDF, discussed later in this report.
Table 5-21
Typical waste fuel parameters compared with coal
Unit

Baseline Coal
Illinois/Central

MSW @31%
Moisture2

RDF @25%
Moisture2

Btu/lb

12,000

5,300

5,700

Moisture

17.6

31

25

Ash %

8.9

16

9.6

Volatile Matter

36.4

44.2

54.6

Fixed Carbon

37.1

8.8

10.8

Total

100

100

100

Fuel Parameter
Heating Value
Proximate Analysis

Ultimate Analysismoisture free


Carbon

77.4

39

47.9

Hydrogen

5.5

5.3

6.5

Nitrogen

1.1

0.4

0.3

Sulfur

4.8

0.2

0.2

Chlorine

Not detected

1.5

1.2

Oxygen

11.2

53.6

43.9

Total

100

100

100

5-50

Municipal Solid Waste


Table 5-21 (continued)
Typical waste fuel parameters compared with coal
Unit

Baseline Coal
Illinois/Central

MSW @31%
Moisture2

RDF @25%
Moisture2

Silica (SiO2)

41.7

50.5

56.3

Alumina (Al2O3)

20

18.2

20.1

Titania (TiO2)

0.8

Not detected

Not detected

Iron Oxide (Fe2O3)

19

12

Lime (CaO)

7.8

8.7

Magnesia (MgO)

0.8

0.7

0.8

Potassium Oxide (KO)

1.6

5.9

6.6

Sodium (Na2O)

1.6

3.1

3.5

Sulfur Trioxide (SO3)

4.4

Not detected

Not detected

Undetermined

2.1

1.8

Total

100

100

100

Fuel Parameter
Ash Mineralpercent of ash

Sources:
1. Coal: Steam, Its Generation and Use, Babcock & Wilcox, 41st Edition, pp. 9-10.
2. MSW, RDF: Comprehensive Waste Characterization by Glaub, Savage, et al., Proceedings of the Eleventh
Biannual Conference of the ASME Waste Processing Division, p. 95.

5.6.2 Advantages and Disadvantages of Waste-to-Energy Systems for MSW


MSW must be removed from residential, commercial, institutional and other sources, and either
directly disposed of, or processed for recovery of energy and/or materials. As of 2009,
approximately 12%, 29 million tons, of MSW generated in the United States was processed
through WTE systems, and approximately 54%, 132 million tons was disposed of in landfills, as
shown in Table 5-22.

5-51

Municipal Solid Waste


Table 5-22
Municipal solid waste generation and disposal
Thousands of Tons
2007

2008

2009

Generation

254,980

251,020

242,960

Recovery for recycling

63,090

61,750

61,270

Recovery for composting1

21,710

22,100

20,750

Total materials recovery

84,800

83,850

82,020

Discards after recovery

170,180

167,170

160,940

Combustion with energy recovery

31,970

31,550

29,010

Discards to landfill, other disposal3

138,210

135,620

131,930

Percent of Total Generation


2007

2008

2009

Generation

100

100

100

Recovery for recycling

24.8

24.6

25.2

Recovery for composting1

8.5

8.8

8.6

Total materials recovery

33.3

33.4

33.8

Discards after recovery

66.7

66.6

66.2

Combustion with energy recovery

12.5

12.6

11.9

Discards to landfill, other disposal3

54.2

54

54.3

Source: U.S. EPA, 2009 Municipal Solid Waste in the United States, Table 29
Notes:
1. Composting of yard trimmings, food scraps and other MSW organic material, and does not include backyard composting.
2. Includes combustion of MSW in mass burn or refuse-derived fuel form, and combustion with energy recovery of source
separated materials in MSW (for example, wood pallets and tire-derived fuel).
3. Discards after recovery minus combustion with energy recovery. Discards include combustion without energy recovery.

A mass-burn WTE facility converts waste into useful energy. The energy can be in the form of
heat, which can then be used to generate steam or steam and electricity, or it can be in the form
of a combustible gas, which can be used to supplement natural gas supplies. The majority of
WTE processes used in the United States today are direct combustion (incineration) technologies
and they offer several benefits when compared to land application, or landfilling, as discussed in
the following text.

5-52

Municipal Solid Waste

Benefits of WTE incineration include the following:

WTE is immediate. Waste is processed within hours or days of its delivery to the facility. By
contrast, landfills are designed for the permanent deposition of MSW and its decomposition
over the years. Modern landfills are sealed from external influences, such as rainwater, and
the natural degradation of MSW is slowed, creating a potential source of leachate and
volatile off-gases that can contaminate water sources, create odor, and cause other safety
hazards.

WTE systems require a relatively small land area, allowing the facility to be located close to
the waste source. Landfills need many acres of land, making them unsuitable for locations
near population centers, resulting in longer trucking routes and associated safety and
pollution concerns.

By definition, WTE systems produce steam and/or electric power. They can be connected to
an existing electric power grid to generate revenue for the facility. Landfill gas can be
collected and used to generate useful energy, but a landfill typically has to be of a minimum
of 10 acres in size and must have been in operation for at least 8 to 10 years to generate
sufficient energy to make landfill gas recovery for energy production economical. After the
landfill has been filled, the quantity of landfill gas generation will decline over time until it is
exhausted. WTE systems will generate energy whenever they are charged with waste.

Regulations in many states consider MSW as a renewable resource when combusted in a


WTE system, and this can translate into tax and regulatory advantages.

Modern air emissions control for WTE is effective and efficient. Troublesome pollutants
such as dioxins, nitrogen oxides, hydrogen chloride, heavy metals, and so on, can be
effectively removed from the flue gas before being discharged into the atmosphere. This also
includes nuisance emissions such as smoke and odor.

WTE systems are reliable, reflecting annual operating availability of some individual units
being up to 95%, and an overall system availability being approximately 90%. There nearcertainty not only that MSW will be disposed of, but also that power generation goals will be
met.

Ash from WTE systems is approximately 25% of the MSW by weight, and 10% by volume.
Where ash is found to contain troublesome pollutants, such as heavy metals in dangerous
concentrations, admixtures, such as Dolomitic lime (CaMg(CO3)2), are available that will
bind these constituents and render them harmless, allowing the ash to be disposed of in a
permitted landfill, or in some instances recycled as alternative landfill cover, road fill, cinder
blocks or other similar products. In some WTE facilities, the ash is processed through screens
and magnetic separation equipment to recover ferrous metals, and in some cases, non-ferrous
metals.

5-53

Municipal Solid Waste

There are a number of issues with WTE plants that should be considered, including the following:

WTE requires skilled operators and maintenance personnel, with the same education and
experience as utility boiler operators and maintenance personnel, whereas a landfill labor
pool is predominantly equipment operators who operate the waste compaction equipment and
LFG fueling equipment.

Fossil fuel is required for WTE startup and in the rare instances where MSW is wet (waterlogged), requiring supplemental fuel to maintain operating temperatures for CO control.

At this time, the cost of disposal of MSW at a WTE facility is generally higher than at a
landfill in many parts of the country, which is one reason that WTE plants currently handle
only 12% of the MSW stream.

The capital cost of a WTE facility typically ranges between $220,000 and $300,000 per
installed daily ton of design capacity. Factors such as facility size and location are the driving
factors for the cost differences. Although landfills can be expensive, depending on their
location, in general, the capital cost of a WTE facility is higher.

More operators and other personnel (electricians, mechanics, other maintenance workers),
and with a greater level of skills, are required for WTE facilities than for landfill operations
of a similar daily capacity.

5.6.3 MSW Receipt and Storage


MSW is typically delivered to the WTE facility via truck in quantities ranging between 8 to 12
tons in each load. If the MSW is compacted, either prior to loading or within a compactor truck,
a transfer trailer truck may carry from 18 to 20 tons of MSW.
Vehicles are typically weighed prior to entering the facility tipping area, and then again upon
exiting the facility, to establish the net weight of waste discharged. In some instances, the empty
weight of the truck (tare weight) will have been established and recorded in the scale system, so
the tare weight is automatically subtracted from the gross weight upon arrival, yielding the net
weight, or the weight of the waste, which eliminates the need for weighing the vehicle upon exit.
At most plants, the delivery trucks discharge the waste directly on the tip floor for inspection
prior to being pushed into the storage pit. A direct pit dump delivery example is indicated in
Figure 5-23. Depending on the facility design and throughput, the number of vehicle or truck
bays can vary between four and 10 bays for waste receipt and movement to the storage pit. The
storage pit is typically designed to store enough waste for three or four days to allow for
continuous operation of the facility when waste receipts fall off during a three-day weekend, or
with severe winter weather, for example.
Overhead cranes with grapples move MSW from the storage pit to the incinerator charging
hoppers.

5-54

Municipal Solid Waste

5.6.4 Waste to Energy Technologies for Combustion


As indicated in Figure 5-23, MSW is brought to the facility, discharged, temporarily stored in a
pit, and charged directly from the storage pit. The MSW may be compacted and handled at
remote transfer stations prior to being sent to the WTE facility to decrease trucking requirements.
Most garbage collection trucks include a compaction system within their chassisfor example,
rear-loader, side loader and front-end loader waste collection vehicles. Other than compaction,
no processing is applied to the MSW. The MSW is charged just as received; this process is
referred to as mass-burning. MSW is chute fed, falls onto moving grates within the incinerator,
and burns as the grates move the waste through the unit. The advantage of mass burn technology
is that no front-end processing of MSW is required.

Figure 5-23
Simplified mass burn waste-to-energy facility schematic

5-55

Municipal Solid Waste

5.6.5 Mass Burn of MSW


As previously noted, mass burn of MSW requires no pre-processing of the MSW. As shown in
Figure 5-23, waste is unloaded into a storage pit and is charged to an incinerator system. This
system includes the following features, referring to Figure 5-23:

Charging hopperWaste is dropped into the charging hopper by the crane. With a negative
pressure in the furnace, air will tend to leak into the charging hopper, and it is important to
keep the hopper full of waste. This amounts to a plug that prevents air flow that can
prematurely ignite waste before it gets into the furnace.

GratesA series of moving grates convey the MSW through the furnace. The first section of
grates serve to dry out the material. Burning will begin on the second grate section, when
sufficient moisture has evaporated from the MSW. After the volatiles in the MSW have been
released and burned out, the remaining combustible content of the waste burns out to an ash
on the third grate section. At the end of this section, the ash is dropped by gravity into an ash
quench.

Waterwall systemThe combustion chamber (waterwall section) is kept at a negative


pressure through an induced draft in the gas clean-up system. Maintaining a negative
pressure will assure that any air leakage is into the furnace rather than out of the furnace.
This prevents the unwanted discharge of hot, dirty gases to the area around the outside of the
furnace, where such leakage can cause hazards to personnel and other safety hazards.

Ash dischargeAsh residue from the MSW falls off the last grate segment and is water
quenched and transferred up an incline to an ash storage pit. Traveling up the incline, usually
on a drag-chain conveyor system, the Ash loses much of the quench water, which flows by
gravity back down to the quench pit for reuse. The ash pit is emptied on a regular basis, and
the ash is taken to a landfill for ultimate disposal. As noted earlier, at some WTE facilities
the ash is processed with screens and magnets for recovery of metals and/or further managed
for beneficial use. Beneficial use application usually requires that the ash test non-hazardous.

Off-gasGases released from the MSW pass through the waterwall section of the furnace
where they burn, generating temperature in excess of 1,800F. The waterwall is built of water
tubes and about half of the heat contained in the burning gas is absorbed in this area of the
furnace. The balance of the usable heat in the hot gas stream is removed in longitudinal
sections of the boiler, including the superheater and economizer. The gas is reduced in
temperature from 1,800F, or higher, to a temperature of less than 400F before exiting to a
gas cleanup system.

5-56

Municipal Solid Waste

5.6.6 Air Emissions Control


The gas exiting the incinerator proper (leaving the economizer) will include pollutants which
must be removed before discharging the gas to the atmosphere. Typical air emission criteria are
shown in Table 5-23.
Table 5-23
Air emission criteria
Emissions

Unit(s)

Basis

NOx

(ppmdv)

205

HCl

(ppmdv)

29

SO2

(ppmdv)

80

CO

(ppmdv)

30

Mercury

(g/dscm)

44

Cadmium

(g/dscm)

27

Dioxins/Furans

(g/dscm)

29

Lead

(g/dscm)

100

Particulates

(mg/dscm)

40

Source: U.S. EPA Emission Standards FAR 40, Part 60, for Municipal Solid Waste Combustors, at 7% Oxygen

These include the following:

Carbon monoxideThis pollutant is not affected by the air emissions control system. It is
controlled in the burning process. With the proper temperature in the combustion chamber,
with sufficient air flow and with effective turbulence, carbon monoxide levels should be less
than the maximum allowable value

Hydrogen chloride and sulfur oxides (acid gases)The generation of these gases is based on
the amount of chlorine and sulfur in the waste. A reagent, usually lime, is injected into the
flue gas and will absorb these gases.

Nitrogen oxidesThe generation of nitrogen oxides increases with temperature in the


combustion chamber as well as with other factors. It is controlled in the air emission control
system with the addition of ammonia, urea, or another reagent. The nitrogen oxides are
converted to innocuous nitrogen gas and water vapor.

CDD/CDFThese are dioxins and furans, and along with mercury, they are controlled
through the addition of activated carbon injection into the flue gas stream. The activated
carbon will adsorb mercury, CDD/CDF, and other unburned organics on their surface,
removing them from the flue gas.

5-57

Municipal Solid Waste

Particulate matter (PM), heavy metalsSome of the particulate matter and most heavy
metals (including cadmium and lead), will be adsorbed in the surface of the lime reagent
injected into the flue gas stream, as discussed above for acid gas control. A fabric filter
(baghouse) is normally included in the air emissions control system. This will remove
particles from the gas stream, including the spent lime and activated carbon reagents already
introduced into the flue gas.

OpacityOpacity is a visual criterion. It is an indication of the presence of visible particles


(or visible gases) in the off-gas stream entering the atmosphere. By complying with the
standards for the pollutants previously listed in Table 5-23, there should be no visible
opacity.

5.6.7 Energy Generation


For the postulated 25-MW (net) plant, with 20% power for internal use (25 MW 0.8 =
31.25 MW gross power generation), assuming steam generated at 825 psig and 850F with a
turbine exhaust of 2.0 inches Hg, and a 97% turbine-generator efficiency:

The theoretical steam rate is 6.597 lbs/ KWh

Applying T-G efficiency: 6.596 0.97 = 6.801 lbs/ KWh actual steam rate

For production of 31.25 MW (31,250 KWh):

Steam generation = 31,250 KW 6.801 lbs/ KWh = 212,531 lbs/hr

For 25 MW (net): 25,000 KW 503 KWh /ton = 49.70 tons/hr MSW

49.7 tons/hr MSW 24 hrs = 1,192 tons/day MSW

The critical parameters for process evaluation from the above calculation are 1,192 ton/day
(assume 1,200 tons/day) MSW, a total steam flow (212,531 lb/hr) and a net electrical generation
of 25 MW.
5.6.8 Facility Sizing
The WTE facility must operate continuously because it receives waste on a continuous basis.
Two units would be provided, each for half the design capacity. When one unit is down for
scheduled (or unplanned) maintenance, the other unit will be able to operate. For the indicated
capacity in this evaluation, two 600 tons/day units would be provided for the 1,200 tons/day
of MSW.
5.6.9 System Maintenance and Reliability
A waste-to-energy facility is designed to operate continuously, 24 hours per day, and except for
maintenance shutdowns, 365 days per year. Planned (scheduled) maintenance will require
approximately two weeks a year per boiler and will involve refractory inspections, tube
inspections, replacement of grate components and refractory sections, burner and boiler
maintenance. Unplanned maintenance can occur at any time, to repair boiler tubes, refractory or
moving parts within the grate system. The boiler will require two days for shutdown and another
two days to heat up to operating temperature, so unplanned maintenance, where access into the
boiler is necessary, will likely require at least a week of downtime.
5-58

Municipal Solid Waste

With a total of two weeks scheduled outage per unit and an additional week of unscheduled
downtime, the availability of each WTE unit is approximately 94%. Planned maintenance is
usually done in the fall and/or winter when waste flows are lower. So, the facility will need to be
designed with enough storage capacity to accommodate waste receipts during the outage(s) or an
alternative method of disposal will need to be found for approximately 5% of the MSW the year.
An alternative to this is to oversize the facility to accommodate this expected downtime, which
can represent a significant increase in cost.
5.6.10 Mass Burn Process
As previously described, mass burning of MSW refers to the charging of refuse to the furnace
without pre-processing. It is charged as it comes off the truck except for the removal of white
goods (refrigerators, washing machines, etc.) and certain other large bulky items. As depicted in
Figure 5-24, 50 tons per hour of MSW is charged to the loading hopper of the furnace. With the
addition of appropriate combustion air, the MSW will burn out to 12 tons/hour of ash residue
which must be collected and removed for processing or for final disposal in a landfill. Overfire
air is injected into the furnace to burn off the gases generated by the waste combustion process.
The hot gases will release heat into the furnace boiler sections to generate over 300,000 lb per
hour of steam. The cooled gases will continue to an air emissions control system where all but
trace amounts of contained pollutants will be removed. The cleaned gas will be drawn through an
induced draft fan, then into the stack and discharged to the atmosphere, meeting the strict air
emissions regulatory requirements for the facility.
Steam from the furnace will be directed to a turbine generator for the generation of over 31-MW
of electric power. Extraction points along the turbine will discharge a low pressure steam supply
for feedwater heating and other in-plant use. In addition, some of the generated power will be
used for in-plant operations. The net power generation for sale is assumed to be 25-MW. The
turbine, a condensing turbine, will discharge through a condenser and condensate will be
returned to the steam cycle. Condenser cooling water is provided through a conventional cooling
tower on-site. Makeup water is required for losses in both boiler blowdown and cooling tower
evaporation and blowdown.

5-59

Municipal Solid Waste

Figure 5-24
Waste-to-energy system schematic

5-60

Municipal Solid Waste

5.6.11 Capital and O&M Related Scope and Equipment


The major equipment or equipment categories affecting the capital costs of the 25-MW mass
burn system evaluation include those listed in Table 5-24:
Table 5-24
Total capital equipment
Equipment

Comment

Two (2) 600 ton per day boilers

Includes stoker/grate equipment

Two (2) APC lines (SNCR)

Includes CEM, baghouse, I.D. fan and stack

One (1) turbine generator set

30.5 MW, 10 stage turbine

Balance of Plant Equipment:


Scales and scale house

Two (2) scales and one (1) scale house

Cranes

For material feed and maintenance

Compressors

For maintenance and control

Distributed control system


Ferrous and non-ferrous removal equipment

Ash processing equipment

Equipment and administrative buildings


Mobile equipment

The property size requirements for a 1,200-ton per day facility range between 5 and 10 acres.
Available traffic patterns play a key role in final site size needs. The site would typically be
located in or near a community capable of meeting the daily and annual feedstock requirements.
Once the general location has been identified, other setup factors include but are not limited to
proximity to major roads for waste vehicle access and egress; availability of utilities, including
natural gas, power and water; utility substation for the energy produced by the facility; and
applicable zoning laws and ordinances. In preparing the economic evaluation for the 25-MW
mass burn facility, we have assumed an 8-acre site at an estimated land cost $350,000 per acre.
As stated in the introduction, fixed and variable O&M costs for the mass burn facility evaluation
have been developed based on a generic location within the mid-Atlantic region of the United
States. Fixed and variable O&M cost categories are defined in Table 5-25.

5-61

Municipal Solid Waste


Table 5-25
Representative fixed and variable O&M cost categories
Fixed Costs
Wages (including overtime and benefits)
Staff supplies, training, and safety
Routine building and equipment maintenance
Other costs
Variable Costs
Major equipment outage maintenance
Utilities
Chemicals and reagents
Ash transportation and disposal
Other consumables

The assumptions used to develop the O&M costs are as follows:

Performance
1,200 TPD of MSW at 5,000 Btu/lb
Electric Power Output
Gross 31.5 MW
Net 25 MW
MSW moisture content is 30%

Operations Related
42 personnel at an average hourly rate of $26.40 are necessary to manage, operate,
and maintain the facility on a 24/7 basis.
The fringe benefits burden rate on all labor is 35% of the base salary.
A 10% overtime factor for all operations and maintenance staff is applied.
Ash transportation and disposal costs are assumed at $40 per ton.
Plant consumables and supplies are included.
Chemicals and reagents
Ammonia (assumed cost: $850 per ton)
Carbon (assumed cost: $1,700 per ton)
Utility costs have been included.
Purchased electricity at $70 per MWh
Potable water at $3.40 per 1,000 gallons
Mobile equipment and auxiliary fuel costs are included.

5-62

Municipal Solid Waste

Maintenance
Annual maintenance costs have been developed per equipment manufacturer
recommendations or GBB experience

General
Management and overhead is included at 12% of the total O&M costs.
Insurance and property tax allowances have been included.

5.6.11.1 MSW Fuel Costs


As previously calculated, a 1,200 TPD mass burn WTE facility should be able to handle at least
400,000 tons per year of MSW. This assumes a conservative 91% annual system capacity factor.
Mass burn waste-to-energy facilities do not pay for their fuel; they are paid fees, commonly
referred to as a tipping fee, for the waste disposal services. Tipping fees charged must be
competitive, and are usually based on the local waste disposal market. Therefore, GBB has
adjusted for this factor in our calculations by entering the MSW fuel cost as a negative () value.
The waste tipping fees in the mid-Atlantic region vary between $40 and $70 per ton depending
on the disposal options available to the local waste disposal companies. Based on each of these
factors, GBB has estimated the facility will be paid an average of $50 per ton, which equates to
approximately $5.00/MBtu.
5.6.11.2 Cost Estimate Summary
Table 5-26 summarizes the marginal capital, operations and maintenance costs on a $/KW basis,
for the 25 MW mass burn system.
Table 5-26
Mass burn capital operations and maintenance costs
Facility Type

25 MW (net) Mass Burn

Rated Capacity
Plant size (units unit size), MW (net)

1 25

Gross capacity, MW

31.25

Net capacity, MW
Capacity factor, % (Net: Gross Power)

25
80%

Physical Plant
Plant life, Years

50

Scheduling
Preconstruction, license and design time,
Months

16.5

Idealized plant construction time, months

16.5

Hypothetical in-service date

2012

5-63

Municipal Solid Waste


Table 5-26 (continued)
Mass burn capital operations and maintenance costs
Total Capital Requirement

$1,000

$/KW

Steam generation system

$63,400

$2,536

24%

Turbine island system

$10,000

$400

4%

Environmental controls

$17,500

$700

7%

$90,900

$3,636

34%

BOP Facilities

$27,700

$1,108

10%

General facilities and site-specific

$108,600

$4,344

41%

$136,300

$5,452

51%

Engineering fee and construction management

$18,000

$720

7%

Contingency

$21,200

$850

8%

$39,200

$1,570

15%

Total Plant Cost

$266,400

$10,658

100%

AFUDC (Interest during construction)

$11,606

$464

Total Plant Investment (including AFUDC)

$278,006

$11,122

Total owner costs

$26,644

$1,066

Total Capital Requirement

$304,600

$12,188

Major Equipment Costs

Total Major Equipment Cost


Direct Balance of Plant Costs

Total Direct Balance of Plant Cost


Indirect Balance of Plant Costs

Total Indirect Balance of Plant Cost


Total Costs

Operation and Maintenance Cost


Fixed, $/KW-yr

480.72

Variable, $/MWh

52.35

Performance/Unit Availability
Net plant heat rate, Btu/KWh (HHV)
Equivalent Planned (Scheduled) Outage Rate, %

4,000
4%

Confidence and Accuracy Rating


Technology development rating

Mature

Design and cost estimate rating

Preliminary

5-64

Municipal Solid Waste

5.6.11.3 Levelized Cost of Electricity


The levelized cost of electricity (LCOE) is the fixed charge of the total capital, fixed and variable
O&M, and fuel life-cycle costs. The costs are presented on a levelized basis over 20 years
divided by the marginal annual MWh generated at the facility. The LCOE estimates presented in
Table 5-27 represent the marginal change to the cost of electricity caused by the mass burn
facility in relation to a facility firing coal. Table 5-27 shows the LCOE estimates with and
without the investment tax credit (ITC) and with fuel cost sensitivities. The LCOE is further
broken down to show the fixed O&M, variable O&M, fuel, and capital charge components.
Table 5-27
Mass burn facility levelized cost of electricity
Cost

Base Case

Sensitivity
Case 1

Sensitivity
Case 2

5.00

6.00

4.00

Fixed O&M component of LCOE ($/MWh)

54.88

54.88

54.88

Variable O&M component of LCOE ($/MWh)

52.35

52.35

52.25

Fuel component of LCOE ($/MWh)

20.00

24.00

16.00

Capital charge component of LCOE ($/MWh)

44.31

44.31

44.31

Total3 ($/MWh)

131.54

127.54

135.54

Fixed O&M component of LCOE ($/MWh)

54.88

54.88

54.88

Variable O&M component of LCOE ($/MWh)

52.35

52.35

52.25

Fuel component of LCOE ($/MWh)

20.00

24.00

16.00

Capital charge component of LCOE ($/MWh)

52.48

52.48

52.48

Total3 ($/MWh)139.71

139.71

135.71

143.71

MSW Fuel Cost, $/MBtu


Levelized Cost of Electricity2 with ITC ($/MWh)

Levelized Cost of Electricity2 without ITC ($/MWh)

Notes:
1. Sensitivity cases assume the MSW fuel costs differential is based on a $1/MBtu change above and below the
difference between base case fuel cost. However, the actual difference between the cost of $3.40/MBtu and the
disposal fee charged to haulers of $5.00/MBtu is $8.40/MBtu.
2. Estimate of LCOE assumes an 80% capacity factor, 20-year project life, and constant dollars.
3. Sum of LCOE component values may not equal total value due to rounding.

5.6.11.4 Levelized Fixed Charge Rate


The levelized fixed charge rate (LFCR) was calculated on an annual basis and applied to the total
capital requirements for the retrofit. As noted in EPRIs Economic Methodology and
Assumptions, the LFCR includes the applicable depreciation, interest on debt, income and
property taxes, return on equity, insurance and administrative costs. Table 5-28 shows LFCR and
the fixed capital charge component of cash flow estimates with and without the investment tax
credit (ITC).
5-65

Municipal Solid Waste


Table 5-28
Mass burn facility levelized fixed capital charge component
Levelized Fixed Charge
Rate (LFCR)

Fixed Capital Charge Component


of Cash Flow, $/kW-yr1

25 MW mass burn with ITC

3.54%

$388

25 MW mass burn without ITC

3.77%

$460

Scenario

Note:
1. Fixed capital charge component of cash flow calculated by multiplying the LFCR by the total capital requirement.
The components of the annual fixed charge excludes charges associated with fixed O&M costs and includes estimates
for the amortization, depreciation, return on equity, income and property taxes, insurance, and other costs.

5.7 Evaluation of Cofiring of Coal with RDF


Refuse-derived fuel (RDF) is a combustion feedstock derived from MSW. This chapter presents
the results of an engineering and economic evaluation of a representative boiler at a power plant
modified to co-fire coal and RDF in an existing pulverized coal (PC) steam boiler. This
evaluation is based on the following assumptions:

Base facility: a 250-MW wall-fired PC boiler fired by Illinois/Central Appalachian coal,


including air emissions control equipment to meet existing regulations on NOx, SOx,
particulate, and so on. The facility includes truck scales, as well as other balance-of-plant
equipment, necessary for plant operation.

Evaluation: at a heat level of 10%, the equivalent input of 10% from RDF (25-MW as
co-fired refuse derived fuel with a heat content of 5,700 Btu/lb) and 90% coal (225-MW).

As previously described in this report, MSW can be processed to produce a feedstock typically
called RDF, which has improved properties for subsequent combustion and will also have other
advantages over directly firing MSW.
5.7.1 RDF Processing and Features
The processing of MSW into RDF is typically done for the following reasons:

RecyclingThere may be an economic value in the removal and recycling of materials


within the MSW stream. These materials would include items such as ferrous and nonferrous metals, plastics, cardboard and glass. Recycling programs across the United States
have targeted removal/recycling to as much as 40% or 50% and use of RDF as a fuel will
help them achieve their goals.

ReductionRemoving recyclables from MSW reduces the amount of MSW to be managed


and disposed of.

Heat contentMany of the materials removed from MSW in preparing an RDF are
inorganic, such as metals, aluminum and glass. As a result, the net heating value of RDF is
usually higher than that of the original MSW. Where removal of plastics is also included in
the process, the increase in heat content would be less because plastics generally have a
relatively high heating value. Even in this case, the heat content (measured on a Btu/lb basis)
will not decrease because of expected high fraction of removed inorganic materials.

5-66

Municipal Solid Waste

Inorganics removalLow melting point components of MSW (mainly glass and aluminum)
will tend to increase slagging during the combustion process, so removal of these materials
through RDF preparation will reduce the incidence of slag formation in the boiler. Also, ash
quantity will decrease.

Uniformity of fuel materialRDF processing creates a physically more uniform size product
which can be conveyed, blown, and fired in suspension, which would have been impossible
with the original MSW.

To be compatible with conventional power systems for co-firing with coal, the MSW must be
converted to RDF to allow it to be handled, conveyed, and fired in an effective manner.
5.7.2 RDF Production Technologies
RDF is a generic term for a variety of treatment options which can include any of the following
processes:

Basic: shred MSW and remove ferrous and non-ferrous metals from the shredded material
with magnetic and eddy current magnet systems.

Intermediate: use an automated bag breaker to open any plastic bags present, pass the MSW
through a trommel screen to remove glass and other very small and larger materials by
gravity (or screen size), shred the rest of the waste, and remove ferrous and non-ferrous
metals from the shredded material with magnetic and eddy current magnet systems.

Manual: use an automated or semi-automated bag breaker system to open any plastic bags
present, pass the MSW through a manual sorting line where identifiable and bulky
components of the waste are removed, pass the balance of the material though a shredder and
remove ferrous and non-ferrous metals from the shredded material with magnetic and eddy
current magnet systems.

Advanced: Pass the shredded material in any of the above options through an air density
separator (air classifier), optical sorting system, salt bath or other equipment to remove most
of the glass, aluminum and other metals and inorganics from the shredded waste.

There are a number of options within these examples depending on the ultimate size of the RDF
desired and the kinds of materials desired to be removed from the waste stream. In some
applications, double shredding (two shredders in series) is used to reduce the ultimate size of the
RDF more efficiently than with a single shredder.
Any of these treatment options will remove some level of inorganics from the waste stream (metals,
glass, ceramics, stones, and so on) Common, readily available equipment, such as shredders and
magnets in the Basic and Intermediate options noted earlier will reduce the inorganic content of
MSW to the range of 8% to 12%, but to get much above this range would require the air classifiers,
optical sorting systems, and other equipment as noted in the Advanced option.
The maximum size of RDF that can be fired reliably in suspension (noted below) is minus 2 inch
(maximum dimension of 2 inches); however, the smaller this size, the more efficient will be the
combustion process. Also, with a smaller size the ability to remove inorganics increases. Air
classification, for example, is much more efficient with smaller size materials. The cost of
processing RDF necessarily increases as the complexity of the process increases. The smaller the
RDF size, the less the inorganic fraction, the higher the heat content, but the more it costs.
5-67

Municipal Solid Waste

5.7.3 Heating Value of RDF


MSW, depending on the locality and/or country in which it is collected, may have a moisture
content of up to 40%, and contain approximately 25% noncombustible solids such as metals and
glass. MSW, as-received, in the United States typically has a moisture content of 30% and
calorific value of approximately 5,000 Btu per pound. A basic processing system, which shreds
the MSW and then solely removes ferrous metals with a magnet and other inorganics with a
trommel screen, will typically produce an RDF with a moisture content of 25% and a heat value
of 5,700 Btu per pound. This relatively simple RDF production system can remove
noncombustibles down to the range of 8% to 10% of the as-received MSW stream. Therefore,
the net moisture content of the fuel produce as it passes through the system producing an RDF
product with is approximately 15% greater than the initial MSW.
It is possible to reduce the moisture content and the noncombustible fraction even further, with a
corresponding increase in specific heat content, but this requires a significant amount of
additional equipment and processing. Equipment such as air classifiers, and additional eddy
current magnets, and screens would be needed to remove additional inorganics. Drying, and if
desired pelletizing, the RDF further would also contribute to a higher calorific vale, on a per
pound basis. This would require a rotary drum dryer and possibly pelletizing equipment. All of
this equipment, beyond a shredder and ferrous magnets, is costly, and additional labor would
also be required for its operation and maintenance, so the price of RDF would increase as
additional equipment and labor is provided.
Factors influencing the calorific value of a waste or other material include its moisture content
and ash (or non-combustible) fraction. Moisture represents an endothermic load which means
that heat is required to raise the temperature to the combustion temperature within the furnace, as
opposed to the heat produced by the burning waste, an exothermic reaction. Non-combustibles
do not draw as much heat from the waste as moisture, but it too is endothermic and takes heat
away from the process. The higher the moisture and non-combustible content of a waste, the
lower is its heat content. For instance, fuel oil has no moisture content and very little ash (noncombustibles) and it, therefore, has a relatively high calorific value. The same is true for coal and
tires, both of which have little or no moisture and minimal inorganic content. By comparison,
RDF produced with very limited processing system, has over 20% moisture and over 10%
inorganics and its calorific value, likewise, is much lower than these other materials.

5-68

Municipal Solid Waste

Table 5-29 outlines typical heating values of waste and other materials.
Table 5-29
Typical heat content of waste and other materials
MSW (in England)

3,800-4,800

Btu/lb

MSW (in the USA)

4,000-5,500

Btu/lb

Wood, hardwood, average

5,100

Btu/lb

RDF (shredded, mag separation)

5,700

Btu/lb

Hospital waste

6,000

Btu/lb

RDF (max1 processing in USA)

7,800

Btu/lb

RDF (max1 processing in England)

8,000

Btu/lb

Coal

11,500

Btu/lb

Tires

13,800

Btu/lb

Fuel oil

18,750

Btu/lb

Note: 1. Maximum processing refers to the use of eddy current magnetic separation, air classification, screens,
drying and additional processing equipment stages in addition to the conventional shredding, trommel and ferrous
metal magnetic separation used for basic RDF production. Higher heat content RDF may not be available in all
locations but if a reliable supply is established, the per-ton cost of RDF compared to the advantages of higher
calorific waste would have to be evaluated on a case-by-case basis depending upon the boiler system to determine
cost effectiveness.

5.7.4 SlaggingA Key Combustion Parameter


The effective design of a combustion system requires consideration of a number of factors,
including heat content, ash fraction, material consistency and slagging potential. Slagging
potential is crucial to system design, whether for a power boiler or a fluid bed boiler burning
either coal or some other material. Slagging will occur when melting occurs, and when the melt
contacts a wall, a tube, or another surface of slightly lower temperature. When this happens the
molten material will harden into a slag or a clinker and can build up, over time, to create a dead
mass interfering with air flow, gas flow, and heat transfer, as well as potentially causing other
problems.
Slagging potential is a function of melting temperature, which is itself determined by material
characteristics and the presence (or absence) of oxygen as well as other factors. Typical melting
(or ash fusion) temperatures crucial to coal, and to other materials, are as follows:

Coal ash fusion temperature: 1,900F to over 2,200F

Ash residual from paper combustion fusion temperature: 1,800F to 2,100F

Aluminum melting point: 1,220F

Melting point of glass: from 900F to over 2,500F depending on the type of glass and its
additives

5-69

Municipal Solid Waste

Table 5-30 outlines the typical moisture and ash ranges for RDF, and the baseline coal, used in
this evaluation. RDF contains a slightly higher ash fraction than coal, which may increase
slagging.
Table 5-30
Moisture and ash percentages of RDF and coal
Fuel
Baseline Coal Illinois/Central Appalachian
RDF

Moisture, %

Ash, %

17.6

8.9

25

9.6

Sources: Portions of: Steam, its Generation and Use, Babcock & Wilcox, 41st Edition, pp. 9-10; Comprehensive
Waste Characterization by Glaub, Savage, et al, Proceedings of the Eleventh Biannual Conference of the ASME
Waste Processing Division, p. 95.

As discussed in the following section, some furnace systems can be designed to accommodate
the presence of most of these materials, with certain restrictions.
5.7.5 RDF Co-Fire Energy Generation and Feedstock
For the 250-MW (net) plant considered in this evaluation, with 20% power for internal use (250
MW 0.8 = 312.5 MW gross power generation), assuming steam generated at 825 psig and
850F with a turbine exhaust of 2.0 inches Hg, and a 97% turbine-generator efficiency:

Theoretical steam rate: 6.597 lbs/KWh

Applying T-G efficiency: 6.597 0.97 = 6.801 lbs/KWh actual steam rate

For production of 312.5-MW (312,500 KW):

Steam generation = 31,250 KW 6.801 lbs/KWh = 2,125,312 lbs/hr

With this PC boiler burning 12,000 Btu/lb coal with a typical equivalency of 9,000 net Btu/hr per
KW70 the following coal feed rate is required:

12,000 Btu/lb 2,000 lbs/ton 9,000 Btu/KWh = 2,666 KWh/ton

For 250-MW: 250,000 KW 2,666 KWh/ton = 93.8 tons/hr coal

93.8 tons/hr coal 24 hours/day = 2,252 tons/day coal

Using RDF for 10% (25 MW) of the heat capacity of the net 250-MW boiler, and with a net heat
content of RDF of 5,700 Btu/lb, equivalent to 562 KWh/ton71:

For 25 MW: 25,000 KW 562 KWh/ton = 44.5 tons/hr RDF

44.5 tons/hr RDF 24 hours = 1,068 tons/day RDF

The two critical parameters for process evaluation from the above calculation are total steam
flow (2,125,312 lb/hr) for generating a net of 250 MW and 1,068 tons/day RDF with 10% of the
calorific requirement derived from the combustion from RDF.
70
71

B&W Steam, 41st edition, Chapter 29.


Mid-Connecticut WTE Facility.

5-70

Municipal Solid Waste

5.7.6 RDF Co-Firing Advantages and Disadvantages


There are both advantages and disadvantages associated with RDF co-firing:
RDF co-firing advantages are as follows:

Reduces coal consumption.

RDF co-firing may reduce base facility operating costs for ammonia due to lower NOx.

Potential financial incentives or revenues, such as carbon credit sales, may be available to the
base facility for the RDF fired.

MSW is ubiquitous, although RDF may not be. RDF processing facilities can be built and
placed in operation in a relatively short time to process the MSW as the market requires.

There are no noticeable adverse effects on air emissions due to the addition of 10% RDF to
the power plant.

RDF has a very low sulfur fraction, decreasing the sulfur oxides footprint from a coal-fired
power plant.

RDF has a low nitrogen content compared to coal and will generate less NOx.

If the cost of RDF becomes uneconomical as a fuel, the boiler can return to 100% coal firing.

Firing RDF for power generation is a form of using a waste for the public good. It could be
expected to have a positive public relations impact, particularly on the local community.

Firing MSW/RDF for power generation will reduce the need for landfills or other disposal
facilities for this waste.

The following are potential RDF co-firing disadvantages:

Modifications are required for the power plant, including additional equipment for RDF
receipt, storage and conveyance, additional burners, boiler tube re-routing, possibly a dump
grate in the boiler, and requisite RDF firing system controls.

RDF must be stored and handled separately from coal.

Operators, and management, will have to deal with a completely different material than that
for which they are accustomed. RDF is a heterogeneous fuel that degrades over time and may
produce odor, it can be difficult to store long-term and may increase the potential for fire at
the facility.

Additional use of soot blowers may be required to minimize slagging and remove potential
accretions from increased particulate loading from RDF firing.

The quality of the fly ash from a facility firing RDF may be different from what is typically
seen in 100% coal firing facility, so the difference must be thoroughly understood before the
ash is utilized in process, such as a cement admixture.

As a different type of fuel and originating from MSW, it will require separate receiving and
handling, and may require separate environmental permitting.

5-71

Municipal Solid Waste

5.7.7 RDF Co-Firing Strategy


Differences in the physical, chemical, and thermal characteristics of coal and RDF require that
they be stored and handled separately, and that they be fired separately, as shown in Figure 5-25,
a simplified system schematic.

Figure 5-25
Simplified system schematic

The RDF stream will have dedicated burners, firing concurrently in the existing boiler with the
existing PC burners. RDF has a higher ash content than coal, but at 10% input, the additional ash
should not require additional equipment in the gas stream. Soot blower operation frequency will
increase, and there will likewise be additional residual in the existing flyash hoppers.
If not already provided with the PC furnace, a dump grate would be desirable to increase the
combustion of the heavier fraction of RDF particles that will not burn in suspension within the
boiler freeboard.
The air pollution control system within the boiler facility must control pollutants from the
burning coal, and with the exception of chlorine, RDF combustion will generate essentially the
same categories of pollutants such as particulate, NOx and SOx. The reagent used for SOx control
(usually lime or a derivative product) would be sufficient to remove the hydrogen chloride from
the RDF combustion process.
As shown in Figure 5-25, a continuous emissions monitoring system is provided with existing
boiler systems. It would be desirable to add a hydrogen chloride (HCl) analyzer if not already in
place. This will monitor for HCl, which is not presently a major concern with coal burning plants.

5-72

Municipal Solid Waste

5.7.8 Preliminary Design Parameters


There will be changes to the existing coal-fired boiler that are necessary for RDF firing, as
follows:

RDF receiptThe RDF fuel will be delivered to the power plant by truck and weighed prior
to being off-loaded into the RDF storage system. Truck scales are assumed to already exist at
the plant.

RDF conveyance equipmentTrucks would unload RDF into a dedicated hopper, which will
convey it to an on-site storage/unloading hopper. From the storage hopper the RDF will be
conveyed and controlled into metering bins for charging the boiler.

Storage equipmentA live-bottom storage hopper for at least two days storage of RDF is
required. With an estimated density of 2-inch RDF of 405 pounds per cubic yard, and a daily
consumption of 1,100 tons, the storage requirement for one day is approximately
6,000 cubic yards.

Storage capacityFor two days storage, and applying a contingency factor of 25% to the
required volume, the storage bin would have an estimated storage capacity of 15,000 cubic
yards.

Boiler introductionThe RDF would be introduced into the boiler at three locations on each
of two opposing sidewalls, using three burners or nozzles each boiler wall, at approximately
slightly higher elevation then the existing PC burners. A typical RDF burner is shown
in Figure 5-26.

RDF burnoutMost of the RDF will burn in suspension in the boiler. Heavier portions of the
RDF, however, will drop to the bottom of the furnace, and a dump grate is advisable to
complete the burnout of this material. A typical dump grate is shown in Figure 5-27.

Boiler modificationsThe boiler wall tubes will have to be modified to provide for the
installation of the RDF burners, along with modifications to accommodate the installation of
the dump grate, if not already present.

ControlsThe RDF supply system must be integrated into the boiler fuel feed control
system. This requires a change in the existing fuel controls in addition to provision of RDF
measuring and metering equipment.

5-73

Municipal Solid Waste

Figure 5-26
RDF burner
Source: B&W Steam 41st edition

Figure 5-27
Dump grate
Source: B&W Steam 41st edition, Chapter 29

5-74

Municipal Solid Waste

5.7.8.1 Material Handling


As previously noted, RDF will be delivered via truck in quantities sufficient to replace 10%, on a
BTU basis, of the coal typically fired by the base facility previously described. The anticipated
RDF material handling equipment required for co-firing 10% includes the following:

Truck dumper with hopper

Conveyor to storage

Storage (silo or bunker)

Feedstock conveyor to boiler nozzles

RDF flow measurement

RDF flow metering and bins

Additional fire detection and protection

The largest equipment would be the live-bottom storage hopper, approximately 90 feet in
diameter by 90 feet high. RDF, because of its higher concentration of combustible materials and
its lower moisture content compared to MSW, represents a higher flammability danger than
MSW. This is an important factor in the design and selection of an RDF storage facility with
supplemental safety equipment, including fire detection instruments and fire protection
equipment included and integrated in the existing facility.
5.7.9 Performance Parameters
By limiting the RDF feed rate to 10% of the total heat input to the boiler, there should be
minimal effect of RDF firing on many of the existing systems and equipment as follows:

Fuel feedingCoal and RDF are charged into the furnace in completely separate feed lines.
Coal should be used to bring the boiler up to at least 50% of the rated load.

Boiler controlOnce the boiler is at or above 50% load, RDF can be introduced. Control of
boiler parameters should be switched to RDF for the base (constant) load with coal feeding to
vary with fuel demand.

Boiler unburned carbonRDF has a lower fixed carbon content than coal. The firing of coal
in the boiler, however, will likely be more efficient than firing of RDF because all of the coal
is fired in suspension while some of the RDF will not burn until it falls onto the dump grate,
which is a less efficient burning process. With these two opposing issues, it is likely that
there will be no, or negligible, change in unburned carbon due to RDF firing.

Boiler NOx generationCoal contains a higher percentage of fuel nitrogen than RDF, which
represents only a fraction of the total NOx generated. If there is a change in NOx generation,
there should be a decrease when firing with RDF.

Mercury (Hg) Coal contains a higher percentage of mercury than RDF, which represents
only a fraction of the total Hg generated. If there is a change in Hg generation, there should
be a decrease when firing with RDF.

5-75

Municipal Solid Waste

Boiler slagging and foulingDepending on the amount of inorganics present in the RDF, it
is expected that there will be some additional slagging and fouling with RDF firing. Many of
the inorganics (aluminum, glass, etc.) will have lower melting and ash fusion temperature
than the fired coal, and it is likely to melt at boiler temperatures. This will require more
frequent use of wall and tube soot blowers with RDF than with coal alone. However, the
majority of these materials should be removed in the RDF production process.

Boiler corrosionAs noted earlier, at 10% RDF firing, there should be no significant
increase in boiler corrosion.

Ash characteristicsWith the majority of inorganics removed from the MSW stream during
the RDF production process, the majority of heavy metals, including mercury, should also be
removed. It is likely that the ash inorganic content, including heavy metals, will be no more
than what is currently in coal ash.

Air pollution controlThe control of SOx and particulate in a power boiler, although not
necessarily the preferred design for hydrogen chloride removal from RDF, will remove over
90% of the HCl generated by RDF combustion at the 10% RDF loading rate. Likewise,
particulate removal from the increased load expected from the RDF should also be satisfactory
without additional modifications from the existing control system. RDF contains less sulfur
than coal, and there should be a slight decrease in reagent requirement when firing it.

Continuous emissions monitoring (CEM) With relatively little chlorine or HCl in the gas
stream for a coal burning power plant, there has been no need for HCl monitoring in the
CEM package. As a recognized pollutant, however, and with chlorinated materials expected
in the RDF, HCl monitoring of the stack discharge would be necessary.

Out of serviceWhen the boiler is out of service it should be kept at a temperature above the
hydrogen chloride dew point. As shown in Figure 5-28, the acid gas dew point for HCl for
moisture concentrations to about 12.5% and HCl concentrations up to 1,000 PPM is 60C
(140F). Although the amount of HCl is very low at 10% RDF firing, there will be some HCl
in the flue gas, and deposited on boiler walls and tubes. If allowed to remain at a temperature
below the HCl dew point for even a few days, hydrogen chloride can form and would be very
destructive to any steel surfaces within the equipment.

Fly ash and bottom ashRDF has a higher ash content than coal, as shown in Table 5-30.
The ash quantities are expected to increase slightly to reflect the additional ash in the
RDF feed.

5-76

Municipal Solid Waste

Figure 5-28
Hydrogen chloride dew point72
Source: Anti-Corrosion Methods and Materials, Vol. 51, 3 (2004, W.M.M. Huijbregts, R. Leferink)

5.7.9.1 Impact of RDF Firing on Boiler Performance


The firing of 10% RDF should have little impact on boiler performance or air emissions, as follows:

Boiler deratingBoiler derating is not required. It would be necessary if the feed would
create a severe problem such as excessive slagging for a particular feed rate, but this is not
expected with only 10% RDF fraction, and with many inorganics removed from the RDF
stream (inorganics are the main source of furnace slagging).

Air emissionsAs previously noted, emissions due to RDF firing are not expected to have a
significant impact on overall emissions from a coal fired plant that is equipped with a modern
air pollution control system. Specifically, note the fate of some of the major pollutants
expected in a power plant firing coal and RDF:

72

Sulfur oxides: RDF has little sulfur, less than coal. Less sulfur oxides are generated from
the burning of RDF than the burning of coal.

Hydrogen chlorideRDF will have a higher concentration of chlorides and a portion of


these will be in the organic form which will convert to hydrogen chloride gas in the
combustion process. A coal power plant will have a pollutant control system using lime or
sodium bicarbonate based reagent. Either of these reagents is more effective in removal of
HCl than sulfur oxides. Over 95% of any HCl that is generated will be removed compared to
less than 90% of sulfur oxides. It is expected that the amount of reagent normally used to
clean up emissions from coal firing would reduce any HCl in the gas stream to insignificant
amounts from 10% RDF firing.
Leferink, R., Latest Advances in the Understanding of Acid Gas Dewpoints.

5-77

Municipal Solid Waste

Dioxin/furansDioxins/furans can be generated in an air pollution control system where


chlorine is present and where gas cooling is relatively slow, compared to a water or air
quench. They, as well as other unburned organic compounds, are absorbed by the same
reagent that is used for mercury control, normally a form of activated carbon. Therefore, it is
expected that dioxin/furans will be removed from the gas stream with existing systems and
equipment.

Heavy metalsWith the majority of inorganics removed from the MSW stream when RDF
is produced, the majority of heavy metals, including mercury, should also be removed. Of the
remaining inorganics, it is likely that they will amount to no more than what is found in the
fired coal stream. The systems and equipment for emissions control in a coal plant, generally
reagent with a bag house or electrostatic precipitator, will remove any heavy metals brought
to the combustion chamber with the RDF.

Nitrogen oxidesNitrogen oxides are generated through thermal NOx (formed as a function
of temperature and oxygen present in the combustion air stream), fuel NOx (a function of the
amount of nitrogen in the fuel) and prompt NOx (from hydrocarbon compounds within the
feed). Thermal NOx will comprise the greatest fraction of NOx produced, but it is a function
of temperature. The greater the temperature the greater the NOx; however, the temperature
within a coal burner is higher than that of a RDF burner. Less thermal NOx will be generated
through RDF firing than coal firing.

Particulate matterIt is possible that RDF will generate more particulate matter during
combustion than coal because most of the coal particulate will be captured in the molten ash
generated in a PC burner whereas the geometry of an RDF burner will not provide the same
absorption mechanism. With 10% RDF burning, it is likely that additional ash will be
generated than 100% coal burning, and this will lead to additional service for the wall and
tube soot blowers. Once past the boiler and economizer tube banks, the majority of any
particulate remaining in the flue gas stream should be removed by the air pollution control
system.

5.7.10 Coal and RDF Quality Summary


The estimated fuel quality characteristics for both coal and the RDF product assumed for this
report are listed in Table 5-31.

5-78

Municipal Solid Waste


Table 5-31
Fuel quality characteristics
Unit

Baseline Coal
Illinois/Central Appalachian

RDF @25% Moisture2

Btu/lb

12,000

5,700

Moisture

17.6

25

Ash %

8.9

9.6

Volatile Matter

36.4

54.6

Fixed Carbon

37.1

10.8

Total

100

100

Fuel Parameter
Heating Value
Proximate Analysis

Ultimate Analysismoisture free


Carbon

77.4

47.9

Hydrogen

5.5

6.5

Nitrogen

1.1

0.3

Sulfur

4.8

0.2

Chlorine

Not detected

1.2

Oxygen

11.2

43.9

Total

100

100

Silica (SiO2)

41.7

50.5

Alumina (Al2O3)

20

18.2

Titania (TiO2)

0.8

Not detected

Iron Oxide (Fe2O3)

19

12

Lime (CaO)

7.8

Magnesia (MgO)

0.8

0.7

Potassium Oxide (KO)

1.6

5.9

Sodium (Na2O)

1.6

3.1

Sulfur Trioxide (SO3)

4.4

Not detected

Undetermined

2.1

1.8

Total

100

100

Ash Mineralpercent of ash

Sources: Steam, Its Generation and Use, Babcock & Wilcox, 41st Edition, pp. 9-10; Comprehensive Waste
Characterization by Glaub, Savage, et al., Proceedings of the Eleventh Biannual Conference of the ASME Waste
Processing Division, p. 95

5-79

Municipal Solid Waste

5.7.11 Capital and O&M Related Scope and Equipment


The major capital-related scope and equipment involved in developing the economic evaluation
are based on the assumption that the RDF is appropriately sized (<2 in.), and includes the RDF
receiving, unloading, feedstock storage and conveying, boiler equipment and modification, and
controls. The equipment associated with weighing the feedstock upon receipt, along with the
property necessary for the RDF receipt and storage, has not been included, since it is assumed
the site of the existing power plant is adequate to support the RDF development project, and it
also has the necessary weighing equipment and personnel to manage the receipt of the RDF.
Figure 5-29 presents a simplified process diagram for the RDF co-firing process evaluated in
this report.

5-80

Municipal Solid Waste

Figure 5-29
RDF co-firing processing diagram

5-81

Municipal Solid Waste

There are several alternatives for the receiving and storing of RDF, such as (1) the material
being received and stored in a dedicated building with boiler feed conveyors being fed through
the use of front end loaders; (2) the material being received in a dedicated building with a pit and
automated feeding system to feed the boiler; and (3) through the use of truck tipping equipment,
dedicated storage building or silo and automated boiler feed equipment. The latter option
described is the basis of estimates contained within this report. The major equipment categories
affecting the capital costs of this evaluation include those listed in Table 5-32.
Table 5-32
RDF capital equipment
Equipment

Comment

RDF truck bay/receipt

Includes truck tipper and hopper

RDF conveyor to storage

220 TPH maximum

RDF storage hopper

2 days of RDF storage @ 12,000 cubic yards

RDF transfer conveyors

2 conveyors @ 50 TPH

RDF flow metering/measurement equipment

2 systems @ 40 TPH

RDF burners

6 burners @ 10 TPH

Boiler modifications for burners

Includes tube panels

Boiler dump grate

Includes boiler modifications

CEM equipment modification

HCl monitoring

Note: Boiler modifications including the number of RDF injection nozzles and grate requirements are based on
assumptions described herein. However, the actual modifications may vary significantly once the actual
facility/boiler is identified. GBB highly recommends the owners engineer and the original equipment manufacturer
(OEM) participate in this engineering review.

As stated earlier, marginal fixed and variable operation and maintenance (O&M) costs for the
RDF co-fire evaluation have been developed based on a generic location within the mid-Atlantic
region of the United States. Fixed and variable cost categories are defined in Table 5-33.
Table 5-33
Fixed and variable cost categories
Fixed Costs
Wages (including overtime and benefits)
Staff supplies, training, and safety
Routine building and equipment maintenance
Other costs
Variable Costs
Equipment outage maintenance
Marginal ash disposal
Utilities
Other consumables

5-82

Municipal Solid Waste

The RDF co-firing assumptions used to develop the marginal O&M costs are as follows:

Performance
1,100 TPD of RDF at 5,700 Btu/lb
Displaces 225 TPD of Illinois/Central Appalachian coal at 12,000 Btu/lb
10% or 25 MW of the plants power produced by co-firing RDF
RDF moisture content is 25%

Operations-Related Costs
Management or administrative labor is assumed to be included in the base facility
operations.
Six new personnel at an average hourly rate of $24 will be necessary to operate and
maintain the new equipment.
The fringe benefits burden rate on all labor is 35% of the base salary.
A 10% factor is used for overtime for operations and maintenance staff.
Marginal plant consumables and supplies are included.
Marginal ash disposal and transportation costs are included at $40 per ton.
Utility costs for the new equipment is included.
Equipment electrical loads estimated at $35 per MWh
Potable water: $3.40 per 1,000 gallons

Maintenance
Annual maintenance costs have been developed utilizing equipment manufacturer
recommendations or GBB experience.

General
Management and overhead is included at 12% of the marginal O&M costs.
A marginal insurance allowance is included.

5-83

Municipal Solid Waste

5.7.11.1 RDF Fuel Costs


As previously calculated, 1,068 TPD, or 400,000 TPY of RDF, will be required to replace 10%
of the daily and annual coal requirements of the base facility. The annual receipt and firing of
400,000 tons of RDF means the base facility will need to be located in an area where MSW is
generated in excess of 600,000 tons annually, and one large RDF or multiple RDF production
facilities exist within a 50-mile radius. The cost of producing RDF from MSW can vary between
$40 and $60 per ton depending on the process equipment utilized and materials recovered, so the
cost of the RDF for use as fuel will be affected by these and other factors, such as availability
and transportation costs. In preparing the RDF fuel cost estimate for use in this report, we have
estimated the RDF production facilitys revenues and costs to operate a 600,000 TPY RDF
production facility in the mid-Atlantic United States, then added transportation costs from an
RDF production facility assumed to be within 50 miles of the base power plants location. Based
on each of these factors, we have estimated the RDF fuel costs will be approximately $35 per
ton, including transportation, which at 5,700 Btu/lb equates to approximately $3.00/MBtu.
5.7.11.2 Cost Estimate Summary
Table 5-34 summarizes the marginal capital and operations and maintenance costs, on a $/kW
basis, for the RDF co-firing evaluation.
Table 5-34
Marginal capital and operations and maintenance costs
Facility Type

RDF Co-firing

Rated Capacity
Plant size (units unit size), MW (net)

1 250

Gross capacity, MW

250

Net capacity, MW (RDF)

25

Capacity factor, %

10%

Physical Plant
Plant life, Years

20

Scheduling
Preconstruction, license and design time, Years

0.5

Idealized plant construction time, Years

0.5

Hypothetical in-service date


Total Capital Requirement

2012
$1,000

$/KW

RDF receiving and storage

$4,362

$174

19%

Boiler modifications and metering

$6,450

$65

28%

Total Major Equipment Cost

$10,812

$239

47%

Major Equipment Cost

5-84

Municipal Solid Waste


Table 5-34 (continued)
Marginal capital and operations and maintenance costs
Facility Type

RDF Co-firing

Direct Balance of Plant Cost


BOP Facilities

$1,620

$258

7%

General facilities and site-specific

$4,730

$189

21%

$6,350

$447

28%

Engineering fee and construction management

$3,430

$137

15%

Contingency

$2,210

$88

10%

$5,640

$226

25%

$22,802

$912

100%

$358

$14

--

Total Plant Investment (including AFUDC)

$23,160

$926

--

Total Owner Costs

$2,280

$91

--

Total Capital Requirement

$25,400

$1,018

--

Total Direct Balance of Plant Cost


Indirect Balance of Plant Cost

Total Indirect Balance of Plant Cost


Total Cost
Total Plant Cost
AFUDC (Interest during construction)

Operation and Maintenance Cost


Fixed, $/KW -yr

47.65

Variable, $/MWh

18.63

Performance/Unit Availability
Net plant heat rate, Btu/KWh (HHV)

4,560

Equivalent planned outage rate, %

4%

Confidence and Accuracy Rating


Technology development rating

Demo

Design and cost estimate rating

Preliminary

5.7.11.3 Levelized Cost of Electricity


The levelized cost of electricity (LCOE) is the fixed charge of the total capital, fixed and variable
O&M, and fuel life-cycle costs. The costs are presented on a levelized basis over 20 years
divided by the marginal annual MWh generated at the facility. The LCOE estimates presented in
Table 5-35 represent the marginal change to the cost of electricity caused by the RDF co-firing in
relation to the base facility. Table 5-35 shows the LCOE estimates with and without the
investment tax credit (ITC) and with fuel cost sensitivities. The LCOE is further broken down to
show the fixed O&M, variable O&M, fuel, and capital charge components.
5-85

Municipal Solid Waste


Table 5-35
Marginal RDF co-fire levelized cost of electricity
Base Case

Sensitivity
Case 1

Sensitivity
Case 2

-0.4

-1.4

0.6

Fixed O&M component of LCOE ($/MWh)

5.67

5.67

5.67

Variable O&M component of LCOE ($/MWh)

18.63

18.63

18.63

Fuel component of LCOE ($/MWh)

2.00

6.00

3.00

Capital charge component of LCOE ($/MWh)

7.79

7.79

7.79

Total3 ($/MWh)

30.09

26.09

35.09

Fixed O&M component of LCOE ($/MWh)

5.67

5.67

5.67

Variable O&M component of LCOE ($/MWh)

18.63

18.63

18.63

Fuel component of LCOE ($/MWh)

2.00

-6.00

3.00

Capital charge component of LCOE ($/MWh)

9.13

9.13

9.13

Total3 ($/MWh)

31.42

27.42

36.42

Cost
MSW Fuel Cost1, $/MBtu
Levelized Cost of Electricity2 with ITC ($/MWh)

Levelized Cost of Electricity2 without ITC ($/MWh)

Notes:
1. Sensitivity cases assume MSW fuel costs different than the base case, at $1/MBtu above and below the base cost
of $3.40/MBtu.
2. Estimate of LCOE assumes an 80% capacity factor, 20-year project life, and constant dollars.
3. Sum of LCOE component values may not equal total value due to rounding.

5.7.11.4 Levelized Fixed Charge Rate


The levelized fixed charge rate (LFCR) was calculated on an annual basis and applied to the total
capital requirements for the retrofit. As noted in EPRIs Economic Methodology and
Assumptions, the LFCR includes the applicable depreciation, interest on debt, income and
property taxes, return on equity, insurance and administrative costs. Table 5-36 shows LFCR and
fixed capital charge component of cash flow estimates with and without the ITC.
Table 5-36
Marginal RDF co-fire levelized fixed capital charge component
Levelized Fixed
Charge Rate (LFCR)

Fixed Capital Charge Component


of Cash Flow, $/kW yr 1

Marginal 25 MW RDF co-fire with ITC

7.15%

$65

Marginal 25 MW RDF co-fire without ITC

7.54%

$77

Scenario

Notes: 1. Fixed capital charge component of cash flow calculated by multiplying the LFCR by the total capital
requirement. The components of the annual fixed charge excludes charges associated with fixed O&M costs and
includes estimates for the amortization, depreciation, return on equity, income taxes, insurance, and other costs.

5-86

Municipal Solid Waste

5.8 Landfill Gas Evaluation


Landfill gas (LFG), and the methane quantity contained therein, has become a marketable energy
source over the past 25 years. For the purpose of this evaluation, 1,000 standard cubic feet per
minute (scfm) of LFG with a higher heating value (HHV) of 500 Btu/ft3 is the feedstock. As
previously described, LFG can be used to produce or be converted to several different forms of
energy. This section presents the results of a conceptual engineering and economic evaluation of
energy recovery from LFG for the following conversion technologies:

LFG to electricityLFG is cleaned and fired in internal combustion engine/generator sets to


produce 2.8 MW of net power for distribution and sale into the power grid.

LFG as direct heat sourceLFG is cleaned and dried for use as of fuel for an industry or user
that consumes a relatively large amount of natural gas fuel.

LFG to pipelineLFG is cleaned, dried, and purified to become a high-Btu fuel for injection
into a natural gas pipeline.

5.8.1 LFG Characteristics


By volume, LFG is generally composed of about 45% to 55% methane (CH4) and 45% to 55%
carbon dioxide (CO2), as it is produced from the decomposition of the waste mass of a landfill.
However, under certain conditions it can be as low as 35% methane and as high as 65% carbon
dioxide. There are also traces of other constituents comprising less than 1% of the volume, with
the most common trace constituents or elements being hydrogen sulfide (H2S), non-methane
organic compounds (NMOCs), sediment, and siloxanes. In addition to containing trace
constituents or elements, LFG is also saturated with water vapor or condensate, which must be
removed to avoid operational problems such as obstructing the gas flow in the collection piping
network.
To extract LFG from a landfill, a vacuum is applied through a pipe network connected to gas
extraction wells or collection trenches installed in the waste mass. Depending on the amount of
vacuum applied to withdraw the LFG, ambient air can be drawn into the landfill and into the
collection system, which introduces oxygen and nitrogen into the waste mass and LFG. Small
amounts of oxygen drawn into the waste mass are usually consumed by organic matter before
entering the collection system; however, nitrogen is inert and is collected along with the CO2,
CH4, and previously mentioned trace constituents. A balanced vacuum on and many wells
associated with the LFG collection system provides for the collection of up to 75% of the LFG
produced. In utilizing a vacuum to collect and remove the gas, sediment and other particulate
matter are also collected with the gas. When the gas undergoes changes in temperature and
pressure, while being pulled to a common collection point, condensate forms. As the collection
pipeline increases in length, it is common design practice for the gas to be routed through a
condensate removal system installed as part of the gas collection system. The condensate is then
disposed along with the landfill leachate.
When condensate is removed, it usually contains some of the trace constituents such as NMOCs
that are water soluble and sediment.

5-87

Municipal Solid Waste

5.8.2 LFG Treatment


LFG is typically unsuitable as a fuel unless treated. In the case of creating high-Btu gas, CO2
must be removed because LFG is considered a low- to medium-Btu fuel with a heating value of
about 500 Btu per scfm at roughly 50% methane by volume. The remainder of the gas is largely
CO2. When LFG is utilized as a fuel for energy recovery projects, such as the types being
evaluated in this chapter, treatment is required to remove condensate and other trace constituents.
Table 5-37 shows the characteristics of untreated LFG.
Table 5-37
Landfill gas characteristics
Unit

Average Untreated
LFG

Methane (CH4)

% v/v

50

Carbon Dioxide (CO2)

% v/v

45

Nitrogen (N2)

% v/v

Oxygen (O2)

% v/v

<1

Hydrogen Sulfide (H2S)

ppmv

21

Halides

ppmv

132

Non-Methane Organic Compounds (NMOC)

ppmv

2,700

Water Vapor (H2O)

lb/MMscf

3,500

Representative Heating Value

Btu/Scfm

450550

Fuel Parameter

Source: Energy Information Administration. US Department of Energy, Growth of Landfill Industry 1996

The typical methods for treating LFG gas to remove condensate and trace constituents are to
compress and cool and/or filter it. The cleaning process is determined by the technology, project
design and type of energy being produced. A single cycle system, primarily compressing and
cooling the gas, is typically adequate for allowing the LFG to be combusted in an internal
combustion engine for the production of electricity or used in most direct heat projects.
Additionally, in the case of a high-Btu conversion project, it will be necessary to remove CO2 to
increase the methane content of the gas and thus increase the heating value of the gas. If the LFG
contains relatively high amounts of nitrogen (N2), it may not be suitable for conversion to a highBtu fuel because the nitrogen will reduce the methane concentration and heating value of the end
product. N2 may also cause operational problems with certain types of CO2 removal processes.
5.8.3 LFG Basic Design
The design elements for converting LFG in an energy recovery project are basically the same.
The most important aspect of a project is the conceptual design phase. The technological
approach needs to be selected at this stage of the project. The type of project will depend on an
evaluation of the quality, quantity and long-term trend of gas production from the landfill. The
size of the landfill, the amount of waste buried, and future disposal quantities expected are
critical to evaluate. The type of project also depends on the long-term priority and commitment
5-88

Municipal Solid Waste

of the entity that owns and operates the gas collection system. In certain situations such as a
high-Btu project, the gas collection system must be operated in a manner to minimize the amount
of air drawn through the waste field and into the gas. If the gas collection system is operated in a
manner to maintain compliance with air-quality permit requirements, where the goal is collect
the maximum amount of methane, the introduction of air is an unavoidable secondary impact, so
the gas quality, including amount of air it contains, must be considered during the project
development and design stage.
5.8.3.1 System Maintenance and Reliability
An LFG-to-energy facility is designed to operate continuously, 24 hours per day and, except for
maintenance shutdowns, 365 days per year. Planned (scheduled) maintenance will require
approximately two weeks a year and will involve equipment inspections and routine maintenance
of the engine/generators, compression and cleaning skids, and scheduled outages can occur at
any time, due to problems with the landfill collection system, gas quality issues or unforeseen
issues with the applicable LFG to energy components.
With a total of two weeks scheduled outage and an additional week of unscheduled downtime, the
availability of the LFG-to-energy facilities is approximately 94%. Therefore, a flare system will
typically be needed to be included with each facility to manage the disposal of LFG during outage
periods.
5.8.4 LFG Issues
The potential issues associated with using LFG as an energy source are as follows:

LFG contains corrosive compounds not typically found in natural gas.

Poor maintenance or operation of the LFG collection system (wells and collection header
system) affects the gas quality.

Inconsistent maintenance or operation of the collection system affects the gas pressures.

LFG has a lower Btu value than natural gas, which may require modification of the end
users equipment.

5.8.5 LFG Benefits


The potential benefits associated with using LFG as an energy source are as follows:

Use of landfill gas has a positive effect on the environment through the reduction of
greenhouse gases (GHGs).

In most state and federal renewable energy programs, LFG is considered a renewable fuel.

Use of LFG in the production of electricity, direct heat and injection into a gas pipeline can
generate revenues in the carbon market, and provide the project owner with valuable tax
credits.

5-89

Municipal Solid Waste

5.8.6 LFG-to-Electricity
In a typical LFG-to-electricity project, electrical power is generated using an internal combustion
(IC) reciprocating engine or combustion turbine to power the generator(s) and produce electricity
for in-house use and sale. The basic capital elements of a LFG-to-electricity project of this type
are gas cleaning equipment, engine-generator set, switchgear and electrical interconnection yard
for connection to the power grid. The gas cleaning equipment is installed to remove water vapor,
H2S and siloxanes. The assumed LFG flow of 1,000 scfm will supply enough fuel to power two
IC engines, such as CAT 3520 or GE Jenbacher engines. Other engine suppliers that provide
equipment to the industry include Deutz and GE Waukesha. Engines designed for this
application typically require the methane content of the LFG be in the range of 40% to 50% by
volume. Table 5-38 illustrates the typical post treatment gas characteristics for IC engines
designed for this application. These engines are purchased as generator sets along with the gas
cleaning equipment and switch gear.
Table 5-38
Summary of gas supply parameters for LFG engines
Fuel Parameter

Illustrative Range

Calorific value and variability

14.4 23.6 MJ/Nm3

Total sulfur content

<2,140 mg H2S per Nm3 CH4 (total S as H2S)3

H2S content

<0.15% v/v

Total Cl content

<100 mg/Nm3 CH4

Total F content

<50 mg/Nm3 CH4

Silicon (Si)

<10 mg/Nm3 CH4

Dust

<50 mg/Nm3 CH4 (particles <3 m)

Oil/residual oil

<5 mg/Nm3 CH4

Moisture

<80% with zero condensate

Inlet gas temperature

1050 C

Methane (CH4) (% v/v)

>40%

Hydrogen (% v/v)

<12%

Source: Data from four LFG engine manufacturers integrated into this table by GBB

The use of gas turbines, such as those manufactured by Solar Turbines, is an option for the IC
units. However, these typically require a higher gas pressure and quality than an IC engine, so
gas turbines are used when individual sites might produce 3 to 5 MW without going to multiple
generating units. Individual IC engines are typically used at sites where the LFG quantity could
produce 0.8 to 3.0 MW.73

73

LFG Energy Product Development Handbook, USEPA.

5-90

Municipal Solid Waste

With the output of this example facility being electricity, and the output from an IC
engine/generator usually at 480 volts, a transformer will be needed to increase the voltage to the
level in the transmission lines at the interconnection point to the power grid. It is usually
necessary and desirable to make the connection to transmission lines with three-phase power.
The interconnection location and input power specifications must be determined through the
local power company before beginning the design for a project of this type. The voltage and
distance to the interconnection point can have a significant impact on capital and maintenance
costs. This evaluation does not include an estimate for the distribution of power beyond the
transformers 230/13.8 KV connection. Depending on site location, the interconnection
requirements with the local utility, and associated costs, could be extensive.
5.8.7 Landfill Gas to Direct Heat
In the case of a direct heat application, the final product gas must be clean, relatively dry and
delivered to the end-users facility at a specific pressure. There are no standard or minimum fuel
quality requirements when using LFG in direct heat applications. Fuel quality in these
applications will be determined by the end user, pipeline length and untreated gas characteristics.
For the purpose of this evaluation, it is assumed the gas qualities for LFG-to-electricity
previously identified in Table 5-38 would meet the minimum requirements for most direct heat
applications.
The basic capital elements for this type of project are the gas cleaning equipment, one or more
compressor stations, the pipeline constructed between the landfill and the end users facility, and
the cost to modify the end users plant to burn the LFG. It is desirable to remove as much
condensate as possible to increase the heating value of the gas and increase the efficiency. This
also reduces the cost of conveying the gas to the end user by eliminating or reducing the amount
of condensate in the gas.
After the LFG is dried and cleaned at or near the landfill, it is compressed and discharged into a
pipeline leading to the end user. A compressor skid is used to pressurize the gas for conveyance.
The compressor skid will need to be designed to meet the pressure required at the point of
delivery of the end user, allowing for whatever pressure drop occurs in the system as the pipeline
size and length between the generation site and user increases.
There is usually no economical way to store treated LFG at the landfill. The ideal gas customer
has a steady daily gas demand that is consistent with consuming the gas production of the
landfill, so the gas volume can be used as it is collected (generated) and delivered.
The pipeline constructed between the LFG cleaning and compression equipment and the end
users facility, would typically be a nominal 4-inch-diameter HDPE pipe rated for high pressure
use. One of the most important aspects of the design and construction of this technology is to
identify the pipeline distance and acquire a right-of-way for the proposed pipeline. A single stage
compressor with about 75-horsepower rating would be used to compress that gas to about 250
psi. By compressing the gas to a pressure of 250 psi at the property line of the landfill, the gas
would flow at the rate of 1,000 cfm and maintain a pressure of about 100 psi at adistance of
about 5 miles. The designer would work with the end user to learn the required discharge
pressure at the point where the gas enters the end-users facility. The size of the compressor
would be based on the required discharge pressure, length of pipeline needed to reach the end
used, and the pipe size used to convey the required volume of gas.
5-91

Municipal Solid Waste

The cost of designing and constructing the gas pipeline ranges from about $100 to $200 per
lineal foot, or about $500,000 to $1,000,000 per mile. The piping material used would be high
density polyethylene, which is suitable for use with pressures up to about 250 psi. Property
acquisition or easements for the rights to install the gas pipeline must also be factored into the
capital and operating costs. Due to this uncertainty and variability of costs and site conditions
associated with different projects and locations, these costs have not been included in our
analysis.
5.8.8 Landfill Gas to Pipeline
In preparing landfill gas for injection into a natural gas pipeline, the Btu value of the gas will
have to be increased from 50% methane (by volume, as generated) to a minimum of 80% to 90%
methane, by volume. Table 5-39 illustrates the typical post treatment gas characteristics for LFG
in a pipeline application.
Table 5-39
Landfill gas quality characteristics for injection into pipeline
Fuel Parameter

Unit

LFG to Pipeline (High Btu)

Methane (CH4)

% v/v

>90

Carbon Dioxide (CO2)

% v/v

<3

Nitrogen (N2)

% v/v

<3

Oxygen (O2)

% v/v

<1

Hydrogen Sulphide (H2S)

ppm v/v

<4

Water Vapor (H2O)

lb/MMscf

<7

Representative Heating Value

Btu/Scfm

>950

Source: GBB research integrated into this table.

For marketing as pipeline quality gas, the LFG must be cleaned and virtually all of the CO2
removed to create a pure methane product. Certain projects are using MEDAL membranes for
CO2 removalfor example, where the design is to reduce the CO2 content from 45% down to
1%. An illustrative schematic of the MEDAL system, which can be skid-mounted, is depicted in
Figure 5-30.

5-92

Municipal Solid Waste

Figure 5-30
MEDAL CO2 removal system
Source: Renewable Solutions Group 9/21/09, D. Wentworth

The basic capital elements for this type of project are gas cleaning equipment for H2S removal, a
CO2 removal system, a compressor station, and the pipeline constructed between the LFG
treatment plant at the landfill and the point of injection into the natural gas pipeline. The
processing equipment for a high-Btu process will include cleaning equipment to remove H2S,
NMOCs, and other trace constituents, which is typically done before the gas enters the CO2removal components of the system. Similar to the direct heat technology, a compressor skid must
be installed, typically at the landfill, after the point where CO2 removal is completed and the
LFG gas has been converted to a methane-rich quality (such as >90%). Natural gas pipelines
typically operate at what is called the maximum allowable operating pressure (MAOP), which is
dictated by factors such as pipe material, pipe size and location of the line. MAOP typically
ranges from 200 psi to 1,500 psi74 and normally is at approximately 90% of the systems designed
MAOP, so a post treatment compressor skid is typically necessary to prepare the LFG gas for
injection into the pipeline.
The pipeline constructed between the LFG cleaning and compression equipment and the
injection point of the high pressure natural gas pipeline would typically be a nominal 4-inchdiameter carbon steel pipe rated for high pressure use. One of the most important aspects of the
design and construction of this technology is to identify the pipeline distance and acquire a rightof-way for the proposed pipeline. A two stage compressor with a 250-horsepower rating would
be used to compress the gas to a pressure of about 900 psi. This was assumed to be the pressure
required to discharge into the distribution pipeline. By compressing the gas to a pressure of 900
psi at the property line, there would be no significant drop in pressure over a distance of up to 5
miles at the required flow rate of about 475 cfm considering a 5% loss. It is assumed the designer
would work with the natural gas company to select the pipeline design, materials of construction
and location for the gas to be discharged into the natural gas line. The designer would also obtain
information on the distribution pipeline material, size and operating pressure from the gas
company, and use this information to select the appropriate compressor.
74

American Gas Association.

5-93

Municipal Solid Waste

The cost of designing and constructing the gas pipeline ranges from about $120 to $200 per
lineal foot, or about $600,000 to $1,200,000 per mile. The piping material would carbon steel,
which is needed due to withstand the 900 psi pressure of the high-Btu gas when discharged into
the pipeline used to convey the gas. Property acquisition or easements for the rights to install the
gas pipeline must also be factored into the capital and operating costs. Due to this uncertainty
and variability of the costs and site conditions associated with different projects and locations,
we have not included these costs in our analysis.
5.8.9 Capital and O&M Related Scope and Equipment
The technologies involved in each of the LFG-to-energy projects evaluated in this report are
similar in many respects, with the differences being dictated by the desired end use of the gas.
The property requirements for each project configuration are all similar in size, approximately
1 acre, and located adjacent too and within the borders of the permitted landfill. Figure 5-31
presents a simplified process diagram for each landfill gas applications that was evaluated.

5-94

Municipal Solid Waste

Figure 5-31
Landfill gas utilization process diagram

5-95

Municipal Solid Waste

Each of the LFG technologies evaluated is conceived to be operated and monitored remotely,
thus minimizing the necessity for dedicated staff and associated buildings or facilities. In
consideration of this, costs of enclosures were estimated for only the engine/generator set and
associated electrical equipment, with the standby flare and gas clean-up and compression skids
and associated equipment designed for outdoor operation, with no building or enclosures
required. The major equipment supply categories, affecting the capital costs of the evaluation,
include those listed in Table 5-40.
Table 5-40
LFG capital equipment requirements1
LFG to
Electricity

LFG to Direct
Heat

LFG to Pipeline
(High Btu)

Post LFG blower

Stand-by flare

Compressor/cleaning skid

Internal combustion/generator set

Equipment enclosures

Electrical interconnection to grid

Equipment

Compressor for pipeline pressure

H2S siloxane removal

CO2 removal equipment

Other Equipment (not in GBB estimate)


LFG to Electric:
Transmission lines beyond transformer output
LFG to Direct Heat:
Pipeline cost beyond compressor skid output
Modifications to end users equipment
LFG to Pipeline:
Pipeline cost beyond compressor skid output
Note: Final equipment selection and need is project-specific and must also consider local and state regulatory
requirements.

As stated earlier, fixed and variable O&M costs for the LFG evaluation have been developed
based on a generic location within the mid-Atlantic region of the United States. Applicable
variable costs include cost estimates for outage maintenance, utilities, and reagent costs
associated with the equipment. Normal O&M or fixed costs include wages, overhead, and
equipment maintenance. Table 5-41 defines the fixed and variable cost categories used in the
evaluation.

5-96

Municipal Solid Waste


Table 5-41
Fixed and variable cost categories
Fixed Costs
Wages (including overtime and benefits)
Staff supplies, training, and safety
Routine building and equipment maintenance
Other costs
Variable Costs
Equipment outage maintenance
Marginal ash disposal
Utilities
Other consumables

The assumptions used to develop the O&M costs are as follows:

Performance
LFG to Electricity
Two (2) 1.6-MW internal combustion engine generator sets
Net plant output 2.8 MW
LFG to Direct Heat
Output of 250 psi LFG @ 500 Btu v/v
LFG to Pipeline
Output of 900 psi LFG @ 900 Btu v/v

Operations-Related Costs
The average labor rate is $26 per hour and based on staffing levels typical for projects of
these types.
The fringe benefits burden rate on all labor is 35% of the base salary.
A 10% factor for overtime for operations and maintenance staff is applied.

5-97

Municipal Solid Waste

Utility costs have been included.


Purchased electricity: $70 per MWh

Maintenance
Annual maintenance costs have been developed based on manufacturer recommendations or
GBB experience.

General
Management and overhead has been included at 10% of the total O&M costs.

Property
The annual facility property lease expense is included in the fuel cost.

Because each of the LFG technologies evaluated is assumed to be designed to be operated and
monitored remotely, minimizing the necessity of dedicated staff, the staffing levels and expenses
are estimated accordingly.
5.8.9.1 LFG Fuel Costs
As previously discussed, LFG is composed of 45% to 55% methane (CH4) with the balance
being carbon dioxide (CO2) and other trace elements, so effectively, LFG has approximately half
the heating value of natural gas; therefore, the cost of LFG is adjusted accordingly. Aside from
the heating value affecting the cost of LFG, other factors such as maintenance of the gas
collection system, agreement term with landfill owner, and the property lease are all accounted
for in the gas cost. In some cases, the gas users do not pay for the gas because they invested in,
and operate, the LFG collection system. In developing the LFG fuel cost in our evaluation, we
utilized the U.S. Energy Administration, 2011 average U.S. natural gas wellhead price as the
base fuel cost and adjusted it for the difference in heating value of approximately 50% and
property lease and collection system costs. Based on these factors, we have estimated LFG fuel
costs to be approximately $2.50/MBtu.
5.8.9.2 Cost Estimate Summary
Tables 5-42 through 5-44 summarize the capital and O&M costs, on a $/KW basis, for the LFGto-electricity project. For the LFG to direct heat and LFG to high pressure pipeline evaluations,
the costs are presented on a $/KW equivalent basis.

5-98

Municipal Solid Waste


Table 5-42
LFG to electricity capital operations and maintenance costs
Facility Type

3.2 MW LFG to Electricity

Rated Capacity
Plant size (units unit size), MW (net)

2 1.6

Gross capacity, MW

3.2

Net capacity, MW

2.8

Capacity factor, %

88%

Physical Plant
Plant life, Years

20

Scheduling
Preconstruction, license and design time,
Years

0.5

Idealized plant construction time, Years

0.5

Hypothetical in-service date


Total Capital Requirement

2012
$1,000

$/KW

$500

$179

7%

Internal combustion/generator and


interconnection

$2,900

$1,036

42%

Total Major Equipment Cost

$3,400

$1,214

49%

$680

$243

10%

$1,728

$617

25%

$2,408

$860

35%

Engineering fee and construction


management

$430

$154

6%

Contingency

$655

$234

10%

$1,085

$388

16%

$6,893

$2,462

100%

Major Equipment Costs


Landfill gas blower, flare compression and
cleaning

Direct Balance of Plant Costs


BOP Facilities
General facilities and site-specific
Total Direct Balance of Plant Cost
Indirect Balance of Plant Costs

Total Indirect Balance of Plant Cost


Total Costs
Total Plant Cost

5-99

Municipal Solid Waste


Table 5-42 (continued)
LFG to electricity capital operations and maintenance costs
Facility Type

3.2 MW LFG to Electricity

AFUDC (Interest during construction)

$108

$39

Total Plant Investment (including


AFUDC)

$7,001

$2,500

$689

$246

$7,600

$2,747

Total owner costs


Total Capital Requirement
Operation and Maintenance Cost
Fixed, $/KW-yr

164

Variable, $/MWh

3.24

Performance/Unit Availability
Net plant heat rate, Btu/KWh (HHV)

438

Equivalent Planned (Scheduled) Outage


Rate, %

4%

Confidence and Accuracy Rating


Technology development rating

Mature

Design and cost estimate rating

Preliminary

Table 5-43
LFG to direct heat capital operations and maintenance costs
Facility Type

LFG to Direct Heat

Rated Capacity
Plant type
Net energy output BTU/yr
Energy output KWh/yr (Btu equivalent)

LFG to Direct Heat


250,000,000,000
22,185,600

Net capacity, MBtu

250,000

Capacity Factor,1 %

100%

Physical Plant
Plant life, years

5-100

20

Municipal Solid Waste


Table 5-43 (continued)
LFG to direct heat capital operations and maintenance costs
Facility Type

LFG to Direct Heat

Preconstruction, license and design time,


Years

0.5

Idealized plant construction time, years

0.5

Scheduling

Hypothetical in-service date


Total Capital Requirement

2012
$1,000

$/MBtu

$500

$2.00

47%

$500

$2.00

47%

BOP Facilities

$100

$0.40

9%

General facilities and site-specific

$300

$1.20

28%

$400

$1.60

38%

Engineering fee and construction


management

$65

$0.26

6%

Contingency

$101

$0.41

10%

$166

$0.67

16%

$1,066

$4.27

100%

AFUDC (Interest during construction)

$17

$0.07

Total Plant Investment (including


AFUDC)

$1,083

$4.33

$107

$0.43

$1,100

$4.76

Major Equipment Costs


Landfill gas blower, flare compression and
cleaning
Total Major Equipment Cost
Direct Balance of Plant Costs

Total Direct Balance of Plant Cost


Indirect Balance of Plant Costs

Total Indirect Balance of Plant Cost


Total Costs
Total Plant Cost

Total owner costs


Total Capital Requirement
Operation and Maintenance Cost
Fixed, $/KW-yr

$0.48

Variable, $/MWh

$0.23

5-101

Municipal Solid Waste


Table 5-43 (continued)
LFG to direct heat capital operations and maintenance costs
Facility Type

LFG to Direct Heat

Performance/Unit Availability
Net plant heat rate, Btu/KWh (HHV)

500

Equivalent Planned (Scheduled) Outage


Rate, %

4%

Confidence and Accuracy Rating


Technology development rating

Mature

Design and cost estimate rating

Preliminary

Note: 1. The capacity factor for the direct heat application is 100% because the gas collected from the landfill gas
collection and control system is typically treated only to remove condensate and trace constituents. This treatment
process does not reduce the amount of methane in the processed gas. The gas delivered to the end user is assumed to
maintain its heating value at 500 Btu per cubic foot.
Table 5-44
LFG to high-pressure pipeline capital operations and maintenance costs
Facility Type

LFG to Pipeline

Rated Capacity
Plant type

LFG to Pipeline

Net Energy output, Btu/yr

237,500,000,000

Capacity Factor,1 %

95%

Physical Plant
Plant life, years

20

Scheduling
Preconstruction, license and design time,
Years

0.5

Idealized plant construction time, years

0.5

Hypothetical in-service date


Total Capital Requirement

2012
$1,000

$/MBtu

Landfill gas blower, flare compression and


cleaning

$500

$2

6%

Carbon (CO2) and siloxane (H2S) removal

$2,150

$9

25%

$600

$3

7%

$3,250

$14

38%

Major Equipment Costs

High pressure compression skid


Total Major Equipment Cost

5-102

Municipal Solid Waste


Table 5-44 (continued)
LFG to high-pressure pipeline capital operations and maintenance costs
Facility Type

LFG to Pipeline

Direct Balance of Plant Costs


BOP Facilities

$1,300

$5

15%

General facilities and site-specific

$2,275

$10

27%

$3,575

$15

42%

Engineering fee and construction


management

$870

$4

10%

Contingency

$808

$3

10%

$1,678

$7

20%

$8,503

$36

100%

AFUDC (Interest during construction)

$133

$1

Total Plant Investment (including


AFUDC)

$8,636

$36

$850

$4

$9,400

$40

Total Direct Balance of Plant Cost


Indirect Balance of Plant Costs

Total Indirect Balance of Plant Cost


Total Costs
Total Plant Cost

Total owner costs


Total Capital Requirement
Operation and Maintenance Cost
Fixed, $/MBtu

$1.97

Variable, $/MBtu

$1.16

Performance/Unit Availability
Net plant heat rate, Btu/KWh (HHV)

475

Equivalent Planned (Scheduled) Outage


Rate, %

4%

Confidence and Accuracy Rating


Technology development rating

Mature

Design and cost estimate rating

Preliminary

Note: 1. The term capacity factor in a high Btu application is associated with the HHV loss as a result of the gas
being treated to increase the quality of the gas for injection into the pipeline. The loss, which can be 5% of the
methane in the gas stream, cannot be recovered.

5-103

Municipal Solid Waste

5.8.9.3 Levelized Cost of Electricity


The levelized cost of electricity (LCOE) and levelized cost of energy (LCOEn) is the fixed
charge of the total capital, fixed and variable O&M, and fuel life-cycle costs. The costs are
presented on a levelized basis over 20 years divided by the marginal annual MWh or MBtu
equivalent generated in each LFG evaluation. The LCOE and LCOEn estimates presented in
Tables 5-45 through 5-47 represent the marginal change to the cost of electricity caused by
changes in fuel costs, along with the estimated LCOE and LCOEn with and without the
investment tax credit (ITC). The LCOE and LCOEn calculations are further broken down to
show the fixed O&M, variable O&M, fuel, and capital charge components.
Table 5-45
LFG to electricity levelized cost of electricity
Base Case

Sensitivity
Case 1

Sensitivity
Case 2

2.50

1.50

3.50

Fixed O&M component of LCOE ($/MWh)

19.53

19.53

19.53

Variable O&M component of LCOE ($/MWh)

3.24

3.24

3.24

Fuel component of LCOE ($/MWh)

1.10

.70

1.50

Capital charge component of LCOE ($/MWh)

20.07

20.07

20.07

Total3 ($/MWh)

43.94

43.54

44.34

Fixed O&M component of LCOE ($/MWh)

19.53

19.53

19.53

Variable O&M component of LCOE ($/MWh)

3.24

3.24

3.24

Fuel component of LCOE ($/MWh)

1.10

.70

1.50

Capital charge component of LCOE ($/MWh)

24.45

24.45

24.45

Total3 ($/MWh)

48.32

47.92

48.72

Cost
LFG Fuel Cost,1 $/MBtu
Levelized Cost of Electricity2 with ITC ($/MWh)

Levelized Cost of Electricity2 without ITC ($/MWh)

5-104

Municipal Solid Waste


Table 5-46
LFG to direct heat levelized cost of energy
Base Case

Sensitivity
Case 1

Sensitivity
Case 2

2.50

1.50

3.50

Fixed O&M component of LCOEn ($/MBtu)

0.50

0.50

0.50

Variable O&M component of LCOEn ($/MBtu)

0.23

0.23

0.23

Fuel component of LCOEn ($/MBtu)

2.50

1.50

3.50

Capital charge component of LCOEn ($/MBtu)

0.31

0.31

0.31

Total3 ($/MBtu)

3.54

2.54

4.54

Fixed O&M component of LCOEn ($/MBtu)

0.50

0.50

0.50

Variable O&M component of LCOEn ($/MBtu)

0.23

0.23

0.23

Fuel component of LCOEn ($/MBtu)

2.50

1.50

3.50

Capital charge component of LCOEn ($/MBtu)

0.38

0.38

0.38

3.61

2.61

4.61

Base Case

Sensitivity
Case 1

Sensitivity
Case 2

2.50

1.50

3.50

Fixed O&M component of LCOEn ($/MBtu)

2.05

2.05

2.05

Variable O&M component of LCOEn ($/MBtu)

1.16

1.16

1.16

Fuel component of LCOEn ($/MBtu)

2.50

1.50

3.50

Capital charge component of LCOEn ($/MBtu)

2.45

2.45

2.45

Total3 ($/MBtu)

8.16

7.16

9.16

Fixed O&M component of LCOEn ($/MBtu)

2.05

2.05

2.05

Variable O&M component of LCOEn ($/MBtu)

1.16

1.16

1.16

Fuel component of LCOEn ($/MBtu)

2.50

1.50

3.50

Capital charge component of LCOEn ($/MBtu)

2.98

2.98

2.98

8.69

7.69

9.69

Cost
LFG Fuel Cost,1 $/MBtu
2

Levelized Cost of Energy with ITC ($/MBtu)

Levelized Cost of Energy without ITC ($/MBtu)

Total ($/MBtu)

Table 5-47
LFG to high-pressure pipeline levelized cost of energy
Cost
LFG Fuel Cost,1 $/MBtu
2

Levelized Cost of Energy with ITC ($/MBtu)

Levelized Cost of Energy without ITC ($/MBtu)

Total ($/MBtu)

Notes: 1. Sensitivity cases assume the LFG fuel costs are $1/MBtu above and below the base case.
2. Estimate of LCOE assumes an 80% capacity factor, 20-year project life, and constant dollars.
3. Sum of LCOE and LCOEn component values may not equal total value due to rounding.

5-105

Municipal Solid Waste

5.8.9.4 Levelized Fixed Charge Rate


The levelized fixed charge rate (LFCR) was calculated on an annual basis and applied to the total
capital requirements for the retrofit. As noted in EPRIs Economic Methodology and
Assumptions, the LFCR includes the applicable depreciation, interest on debt, income and
property taxes, return on equity, insurance and administrative costs. Tables 5-48 through 5-50
indicate the LFCR and fixed capital charge component of cash flow estimates, with and without
the investment tax credit (ITC), for the three LFG recovery projects reviewed.
Table 5-48
LFG to electricity levelized fixed capital charge component
Levelized Fixed Charge
Rate (LFCR)

Fixed Capital Charge


Component of Cash Flow,
$/kW-yr1

2.8 MW net LFG to Electricity with ITC

6.83%

$169

2.8 MW net LFG to Electricity without ITC

7.49%

$206

Scenario

Table 5-49
LFG to direct heat levelized fixed capital charge component
Levelized Fixed Charge
Rate (LFCR)

Fixed Capital Charge


Component of Cash Flow,
$/kW-yr1

LFG to Direct Heat with ITC

7.30%

$0.31

LFG to Direct Heat without ITC

8.00%

$0.38

Scenario

Table 5-50
LFG to high-pressure pipeline levelized fixed capital charge component
Levelized Fixed Charge
Rate (LFCR)

Fixed Capital Charge


Component of Cash Flow,
$/kW-yr1

LFG to Pipeline with ITC

6.81%

$2.45

LFG to Pipeline without ITC

7.47%

$2.98

Scenario

Note: 1. Fixed capital charge component of cash flow is calculated by multiplying the LFCR by the total capital
requirement. The components of the annual fixed charge excludes charges associated with fixed O&M costs, and
includes estimates for the amortization, depreciation, return on equity, income and property taxes, insurance, and
other costs.

5-106

Municipal Solid Waste

5.8.10 Conclusions
Biomass fuels are many and varied, with distinct characteristics. Although there are institutional
and political questions concerning the classification of wood waste and other waste products as
biomass or renewable, technically they are biomass resources that are derived from living plants.
Most biomass fuels permit utilities to generate dispatchable renewable power. The fuels can
be extracted and/or stored and then used to meet electricity demand. Utilization systems can
capitalize upon the characteristics of these fuels. They are modest in heating value, highly
reactive, low in nitrogen and sulfur, and of varying ash characteristics. Of the biomass fuels
available, woody biomass is the most commonly used material, but agricultural wastes and
dedicated energy crops appear to be the biomass fuels of the future.
Numerous technologies are available for biomass fuel utilization, including both cofiring options
and stand-alone options. Cofiring provides a means for incorporating biomass utilization into
electricity generating facilities with the lowest cost and the least risk. Cofiring can be used to
enhance combustion and, in most cases, reduce SO2 and NOx emissions. The lack of sulfur in
biomass fuels, coupled with low nitrogen concentrations, reactive nitrogen, and reactive fuel as a
whole, provides the basis for this enhanced combustion.
The forest products and pulp and paper industries, as well as some electric utilities and
independent power producers, have built stand-alone plants. These systems are not inexpensive.
However, they provide a basis for using biomass to generate electricity. They can be
economically justified depending on localized economics and their use in addressing customer
needs. Institutional arrangements that provide partnerships between industries and utilities can be
particularly useful in such circumstances.
Biomass pre-treatment may open a global biomass trade similar to that of coal. But key to the
expansion of biomass for electricity production will be available tax incentives, RPS mandates,
and GHG-reduction directives that externalize the costs of using fossil fuels. For example,
torrefied biomass may contribute to offsetting fossil-fuel emissions, but consideration of carbon
credits for the char field application will be required to expand its worldwide use.
In summary, biomass is dispatchable renewable power. It can come in solid or gaseous form, and
can be exploited using a wide variety of technologies to supply electricity to internal customers
or to the grid. Moreover, its carbon neutral or negative characteristics position it to potentially
contribute to the target of de-carbonizing electricity production.

5-107

SOLAR PHOTOVOLTAICS

6.1 Introduction
Solar photovoltaic (PV) modules are solid-state semiconductor devices that convert sunlight into
direct-current electricity. PV was initially developed for the space program in the 1950s to power
the first satellites, an application for which simplicity and reliability were tremendously
important and cost was of much lesser consequence. In the early 1970s, PV began to be used for
terrestrial power applications, which led to a shift in R&D goals and funding sources away from
the space program to private/public sector collaborative efforts. This work, led primarily by the
United States, Europe, and Japan, has resulted in a generally steady, dramatic decrease in module
costs and a concurrent near-doubling in energy conversion efficiency. The average module sales
price (in 2011 dollars) dropped from $80,700/kW in 1976 to $1,320/kW in 2011 to under
~$900/kW in 4Q2012. Over the same period, the performance of high-end commercial modules
has increased from 8% conversion efficiency to about 20%. As the market for solar panels grew,
the demand for panels during 2003 to 2008 outpaced the supply. This was primarily a result of
constraints in solar-grade silicon, a key raw material in crystalline panels. This economic
imbalance was the principal reason module prices rose after 2003 and remained higher than
expected after years of declining prices.
Although the PV market growth has continued, the impact of the recent global economic
downturn combined with a very significant increase in poly-silicon supply and module
manufacturing capacity have driven prices significantly downward. Expanded Chinese
manufacturing has been a primary cause of the large increase in solar-grade silicon and
crystalline module manufacturing capacity. In 2012, Chinese manufacturers now account for
60% of world PV shipments, and nine of the top 10 PV manufacturers are predicted to be
crystalline-based Chinese companies within the next several years (the lone exception being
CdTe producer First Solar). First Solar, the thin film market leader, reached commercial scale
production capacity of 2,300 MWp in 2011 and now holds approximately 11% of global market
share. These factors, combined with the recessionary downturn and revisions of various national
tariffs and incentives, have created the current volatile market in which supply outpaces demand
and margins are squeezed.
As of the end of 2011, there were approximately 70,000 MW (70 GW) of PV capacity installed
worldwide, with about 4,000 MW deployed in the United States (Table 6-1) [37]. The
incremental global capacity additions in 2011 totaled approximately 30,000 MW [38], a 75%
increase over 2010 (Figure 6-1).

6-1

Solar Photovoltaics
Table 6-1
Overview of solar photovoltaics
Installed Capacity
(Jan. 1, 2012 est.)[36]

70,000 MW worldwide
4,000 MW in United States
World Leaders:
Germany: 28,000 MW
Italy: 12,800 MW
Japan: 4,900 MW
Spain: 4,500 MW

Technology Readiness

Flat-plate crystalline and some thin-film types are commercial


Concentrator photovoltaic modules and most thin-film products are in
pilot-scale production, while third-generation PV concepts are in the
research phase
Moderate-high confidence in cost estimates and projections for
commercial technologies

Environmental Impact

Manufacturing: Manageable waste streams


Installation: Minimal; potential visual and land-use issues
Operation: Minimal
Decommissioning: Material recycling and disposal needs

Economic Status

Currently competitive only in high-value applications or with subsidies;


cost (without subsidy) remains above that of grid power in most areas,
but the gap has decreased considerably with recent price reductions
Tax and cash incentives still necessary to facilitate most projects

Policy Status

6-2

Countries with generous feed-in tariffs (FITs) have been successful in


spurring market growth, but meeting the financial costs of these
policies has been challenging. Many countries, including Spain and
Germany reduced FITs in 2012 and revised policies. However, the
German FIT reduction was less dramatic than anticipated, and a new
FIT in Japan and revised (higher) PV targets in China acted to buoy
the market in 2012.
The US Federal 30% investment tax credit will continue through 2016
and then revert to a 10% credit. However, the Section 1603 Cash
Grant program, which allowed developers to opt for a cash grant in lieu
of the tax credits, expired at the end of 2011.
The U.S. PV industry is encouraged by state-level RPS and energy
policies that target increasing renewable energy production,
particularly those that favor solar energy, i.e., solar carve-outs and
solar renewable energy credits (SRECs).

Solar Photovoltaics
Table 6-1 (continued)
Overview of solar photovoltaics
Trends to Watch

Increasing automated manufacturing leading to lower cost.


Continued research in the areas of: emerging materials, efficiency
increases, and manufacturing processes.
Some regions, such as North Africa, have not been highly engaged in
the solar PV market despite high solar insolation because of high
capital costs. As cost reduction trends continue, deployment in those
regions will increase.
In May 2012, the U.S. Department of Commerce announced that it
would assign trade duties on imported Chinese modules, alleging that
Chinese government subsidies allowed Chinese solar panel
manufacturers to sell panels at prices significantly lower than the
market value. The ruling was finalized in October. Since the antidumping investigation was filed, the Chinese ministry of commerce
has opened its own investigations to file an anti-dumping claim against
American and European polysilicon manufacturers.
More consolidation and liquidation is expected, particularly for lowertier crystalline silicon (c-Si), thin-film and concentration PV (CPV)
technologies, in response to a lengthy over-supply situation and
subsequent narrowing (if not negative) operating margins.
More building-integrated PV (BIPV) and energy-efficiency bundling.
Continued growth as a distributed resource, including microgrids
Growth of third-party solar PV vendors like Solar City that install and
maintain residential and commercial systems at no cost and sell the
electricity to home and building owners and to the grid.
Large corporations, some with expertise in different but related
businesses such as electronics or chemicals, are angling for strategic
position and investing heavily in PV as the market has grown.
Examples include GE, Dow Chemical, Sharp, Kyocera, and Norsk
Hydro. At the same time, some corporations such as BP and Shell that
were once significant players in PV have retreated amidst the current
state of unsustainable PV equipment prices.

6-3

Solar Photovoltaics

Figure 6-1
Worldwide PV industry growth
Source: EPRI, Navigant Consulting, Solar Power Consulting

6-4

Solar Photovoltaics

The PV deployment growth rate has subsided greatly (but is still over 30%) due to the current
supply-demand mismatch, but as the global economy recovers and capital for PV investments
becomes more readily available, it is expected that more robust growth rates will resume. The
forecast for 2012 global deployment is likely slightly below 2011 levels of 30,000 MW. In the
United States, slightly over 3 GW is expected [42]. Given the increased supply capacity and
tempered demand-pull incentives, it is reasonable to expect continued consolidation in the
industry. The recent collapse in module prices has come at an inopportune time for many thin-film
(a-Si, CdTe, CIGS) companies that are just starting commercial manufacturing. As newcomers,
they faced not only higher actual costs on average than more-established c-Si competitors, but it
has become more difficult to offer the significantly lower price points necessary to win over new
customers from the incumbents. Many companies are exiting the conventional a-Si marketmost
notably Uni-Solaror are changing focus, as with Moser Baers hybrid wafered-silicon cell with a
thin film of amorphous silicon. Only two CdTe technologies remainFirst Solar and Xunlight
26following the exit of Abound, Calyxo GmbH, GE (Primestar) and WK Solar. Some CIGS
companies, such as Heliovolt and Nuvosun, have opted to delay production ramp-up altogether to
limit near-term losses. Others, such as Ascent, Global Solar, and SoloPower, have adopted nichemarket strategies in which they offer a specialized product that commands a higher price.
These survival approaches may succeed in the likely event that the oversupply subsides as the
total PV market continues its historic expansion, even if at a slower pace than the past decades
over-50% compound annual growth rate. Thin film technologies are expected to continue
competing with conventional silicon-based products. Silicon cells at approximately 15%
efficiency (roughly twice as efficient as older a-Si thin-film PV technologies at 7% efficiency)
are well suited to space-constrained rooftop applications (roughly twice as efficient as older a-Si
thin-film PV technologies at 7% efficiency). Although crystalline cells still have an efficiency
advantage, thin-film technologies are moving into the 12% to13% efficiency range, which lowers
racking costs in utility-scale applications and makes them power dense enough for many rooftop
applications. Global thin-film panel production is also increasing and driving down costs. It has
risen from 400 MWp in 2007 to near 6,000 MW in 2012; however, it has dropped a few
percentage points in market share in the past year, now accounting for 11% of total PV panel
production. Thin film production and installation at this scale has allowed this portion of the PV
industry to build expertise and lower costs on the production side through improved
manufacturing equipment. Also, this volume makes them significant participants in the utilityscale, commercial and residential markets. First Solar continues to be the industry leader with its
low-cost, higher-efficiency (now at 12.6%) CdTe thin-film modules [48] (see Figure 6-2).

6-5

Solar Photovoltaics

Figure 6-2
View down one row in 12-MW section of EdF Energies Nouvelles 115-MW plant at Toul,
France using First Solar modules

6.2 Cost and Economic Issues


The economics of the photovoltaic sector have changed rapidly. For about 20 years, up to 2004,
the cost of manufacturing and installing a PV solar power system consistently decreased by about
20% with each doubling of installed capacity. Between 2004 and 2008, demand outpaced supply,
causing a deviation in this trend as prices remained high. However, for the past four years, PV
price reductions have not only returned to the historic trend, but fallen faster, with over 70%
reduction in average module selling price between 2008 and 4Q2012. The cost of electricity
generation from conventional sources, by contrast, varies with the price of fuel, and natural gas has
varied significantly over that time period. This heavily influences electricity prices in regions that
have large numbers of gas-fired power plants, such as California, Texas, and the northeastern
United States, as well as Italy and Spain. Although PV costs continue with significant declines,
recent development of seemingly abundant new supplies of natural gas have created a more
competitive power generation market in key PV sales areas. EPRI analysis shows that utility-scale
centralized PV could prove increasingly practical as the cost of PV falls and becomes competitive
generally with other intermediate and peaking supply technologies.
Although the fundamental PV technology may be the same, cost and economic considerations
related to distributed PV are significantly different from those related to centralized utilityowned PV. In the former case, some or all of the cost may be born by customers interested in
benefiting from on-site distributed generation and supporting a clean, renewable technology.
Depending on local electricity prices and any rebates, subsidies, or incentives available, the
payback period for a typical residential-scale PV system may be as short as five years or could
well exceed a decade, depending on variables such as technology, site latitude and climate, and
system location and orientation. A distributed PV system may also be owned by an energy
company interested in providing grid support in critical areas, exploiting new business
opportunities, or complying with policy demands such as renewable portfolio standards. Because
PV is highly modular, distribution companies can precisely deploy it in optimal locations and
capacities or work with customer-owners to do so. Other purchase or lease arrangements may be
beneficial to customers, utilities, or both.
6-6

Solar Photovoltaics

The U.S. government, acting through the DOE, has instituted several programs through the years
to support PV research and development, subsidize its use, and plan for its future. The federal
Million Solar Roofs initiative, begun under the Clinton administration, aimed to install one
million solar energy systemsincluding PV, water heating, and space heatingon the rooftops
of American homes and businesses by 2010. The initiative included federal procurement
programs, technology grants, and lending programs. Briefly renamed Solar Powers America,
the Million Solar Roofs program effectively concluded in 2006 after investing $16 million to
help install approximately 370,000 solar projects comprising 200 MW of grid-connected PV and
another 200 MW of solar water heating capacity. A final report on the program was issued in
October 2006 [33]. Since then, states such as California and Delaware have established their own
Million Solar Roofs programs and goals. In July 2010, Senator Bernie Sanders (Ind.-VT)
introduced the Ten Million Solar Roofs Act of 2010 (S. 3460) to support similar efforts
through 2021.
In 2006, the Bush administration launched the DOE Solar America Initiative (SAI), whose goals
were to accelerate PV deployments and conduct RD&D targeted to bring PV costs to grid
parity by 2015. The DOE PV budget was roughly doubled as a result of the SAI program and
many new projects were launched, consisting primarily of industry-led short-term efforts to more
rapidly scale up current PV technologies. They also included public-outreach educational
projects, a technology incubator program to help startup companies move from laboratoryscale pilot lines into early commercial-scale production, and a small amount of more exploratory
research into novel concepts. The SAI concluded in 2009, but the activities developed during its
tenure were incorporated into the portfolio of DOEs Solar Energy Technologies Program.
Funding provided through the American Recovery and Reinvestment Act (ARRA) of 2009
greatly expanded DOEs budget for solar energy research, development, and demonstration. In
addition, the DOE Loan Guarantee Program was established to support qualified projects in the
belief that accelerated commercial use of these new or improved technologies will help to sustain
economic growth, yield environmental benefits, and produce a more stable and secure energy
supply. In March 2009, DOE awarded a $535 million loan guarantee to PV developer Solyndra
to advance its cylindrical PV panel design and build a 500-MW-per-year module factory in
Fremont, California. In October 2009, DOE announced that it would draw on ARRA funds to
provide guarantees for up to $8 billion in loans for conventional renewable energy projects such
as solar, wind, biomass, geothermal and hydropower. At the same time, DOE pledged to
streamline the selection standards and lending process.
In early September 2011, Solyndra ceased all business activity, filed for Chapter 11 bankruptcy,
and laid off nearly all its employees. The Solyndra business model was flawed in that it assumed
continued high costs for silicon. Its design utilized CIGS PV cells. When the cost of silicon
sharply decreased, the custom equipment designs for the Solyndra product were too expensive
compared to the more traditional standardized PV panels. Since the announcement of the
bankruptcy, an investigation into the DOE loan process has begun and vigorous debate on the
role of government in supporting new technologies has ensued. In addition, the solar energy
sector has also come under severe scrutiny. However, despite the criticisms, the DOE continues
to provide funding to organizations in the solar industry. Dozens of PV companies are currently
in the queue for DOE support.

6-7

Solar Photovoltaics

In 2011, the Obama administration introduced the SunShot Initiative. The program aims to
reduce the cost of installed solar to $1 per watt, bringing the levelized cost of electricity to 6
cents/kWh without subsidies, by 2020. Under the SunShot Initiative, the U.S. Department of
Energy is funding selective research and loan guarantees for high risk, high payoff concepts in
all facets of solar energy project installation, ranging from the photovoltaic modules and balance
of plant (BOP) components to the reduction of soft costs. Grid integration costs are also being
addressed.
Several states also have programs encouraging the development of renewable energy. In many
cases, a combination of state and federal incentive, rebate, or loan programs can pay a significant
share of a PV systems cost. Most notably, beginning in 1998 Californias Emerging Renewables
Program (ERP) and Self-Generation Incentive Program (SGIP) led to an estimated installed gridconnected PV capacity totaling one-quarter of the nations installed PV in 2005. As of January 1,
2010, SGIP had helped fund more than 890 PV projects representing nearly 136 MW of
generating capacity [34]. Californias PV momentum was further boosted in 2006 by the passage
of SB-1 (Senate Bill 1) and the formation of the California Solar Initiative (CSI), a $3 billion
program to incentivize the installation of one million solar rooftops by 2017. Through midNovember 2012, the CSI had helped fund nearly 106,000 installed PV projects with a cumulative
capacity over 1540 MW (Figure 6-3) [38].

Figure 6-3
California Solar Initiative installed projects by month [38]

As of the end of 2011, California led the United States in cumulative installed PV capacity with
1,564 MW, followed by New Jersey with 566 MW, Arizona with 398 MW, Colorado with 197
MW, and New Mexico with 165 MW [48]. The Interstate Renewable Energy Council (IREC)
concluded that while non-residential (i.e., commercial) systems made up the bulk of new
installations in 2011, utility-scale systems had the highest growth. Residential, non-residential,
and utility installed capacity in 2011 compared to 2010 increased by 24%, 137%, and 146%,
respectively. The non-residential segment in 2011 had the highest installed capacity at 822 MW
(45%), followed by the utility segment at 698 MW (38%) and residential segment at 324 MW
(18%) (Figure 6-4).
6-8

Solar Photovoltaics

Figure 6-4
Annual U.S. installed grid-connected PV capacity by sector [38]

In addition, solar requirements embedded in renewable portfolio standards (RPSs) are also
fueling PV development. The National Database of State Incentives for Renewable Energy
(DSIRE; www.dsireusa.org) provides up-to-date information on state financial and regulatory
incentives that are designed to promote renewable energy technologies.
Other nations, such as Germany, Italy, Spain, and Japan, have had active and generous
government-sponsored and -mandated incentive programs. As a result of its aggressive
programs, Germany leads the world in both annual grid-connected PV installations, with almost
7,500 MW added in 2011 and a total installed capacity of approximately 28 GW, representing
nearly half the worlds total. Italy installed over 9 GW in 2011, making it the second largest
market in the world with almost 13 GW of cumulative installed capacity. Japan installed almost
1,300 MW for a total of nearly 5 GW. The fourth place position belonged to Spain, which added
about 400 MW for a total of 4.5 GW. The United States was fifth with 4 GW [36].

6-9

Solar Photovoltaics

Historically, government subsidies have strongly influenced the growth of the solar power
industry. For example, in the United States producers of renewable energy receive tax credits. In
Germany, electricity distributors are required to pay above-market rates, called feed-in tariffs, for
electricity generated from renewable sources. Similar feed-in tariff incentives have been adopted
by Spain, Italy, Ontario, Canada, and dozens of other countries around the world. Figure 6-6 shows
historic growth in different regions of the world. Asia, North America, and, more recently, Oceania
have ramped up deployment over the past few years, whereas growth in Europe has slowed.

Figure 6-5
Historic PV market growth by country
Source: Navigant Consulting [41]

6.2.1 Japan
Japan had approximately 320 MW of PV in place at the end of 2000, with more than two-thirds
of it installed through a residential capital cost buy-down program that began in 1994 and was
administered by the New Energy Foundation, part of the Ministry of Economy, Trade, and
Industry (METI) until the programs termination in 2005 [15]. In the following five years,
another 1.1 GW was installed, nearly all under the METI program, launched in the late 1990s
and run by METIs New Energy and Industrial Technology Development Organization (NEDO).
The purpose of the METI program is to develop a full range of new energy technologies,
including PV, by subsidizing both R&D and market penetration to bring costs down.

6-10

Solar Photovoltaics

METI PV systems are grid-connected and net-metered, which increases their attractiveness given
Japans high-priced electricity. The maximum subsidy provided to system buyers was initially
about 50% of installed costs. It has decreased annually, reaching 10% in 2003, and was phased out
completely in 2005. Despite the decline and eventual termination of end-user incentives, the
program remained popular and successful to its end, in part because net prices in Japan have
continued to fall, aided by ongoing Japanese government market-development assistance to the
industry.
After three stagnant years, new incentives that took effect in January 2009 paid 70,000/kW
(approximately $820/kW U.S.) for residential PV systems and led to Japans installing more PV
than ever in 2009, approximately 484 MW [32, 40]. In November 2009, METIs New Purchase
System for Solar Power Generated Electricity established a new feed-in tariff of 48/kWh
($0.56/kWh U.S.) for electricity generated by residential PV smaller than 10 kW and 24/kWh
($0.28/kWh U.S.) for non-residential PV, which is expected to further Japans refined goals to
install 28 GW of PV by 2020 and 56 GW by 2030 [40]. After the Fukushima Dai-Ichi nuclear
power plant radiation leaks following the March 2011 earthquake and tsunami, Japanese officials
pledged to reduce reliance on atomic energy, including an aggressive plan that would make it
compulsory for all new buildings and houses to come fitted with solar panels by 2030. The
countrys new FIT plan, ratified in June 2012, pays 42/kWh, or $0.53/kWh over a 20-year
timeframe.75 Operators must sign contracts with power companies by March 2013 to lock in the
FIT purchase price.
6.2.2 Germany
Germany had approximately 110 MW of PV installed as of the end of 2000. Following a 1000
Solar Roofs program in 1989, Germany initiated a 100,000 Solar Roofs program in 1999 with
the goal of installing 300 MW of PV by 2004. The program provides 10-year low-interest loans
that feature no money down and no interest payment for two years toward the installation of
rooftop PV systems. Over the term of the program that ended in June 2003, loans were approved
for approximately 350 MW.
Germanys Renewable Energy Sources Act, which took effect in April 2000, also encouraged
PV development by offering high financial incentives. The act established a PV feed-in tariff
which initially offered solar power producers roughly 0.5062 per kWh ($0.66 at April 2006
exchange rates, about $0.75 in late 2007, $0.69 in 2008), guaranteed for 20 years. Since 2002,
this guaranteed price has been reduced by 5% to 6.5% per year, depending on the type of system,
to encourage cost reductions. New regulations that took effect in January 2004 provide a tiered
incentive structure ranging from 0.469 to 0.632/kWh of PV generation, depending on the
system size and whether it is free-standing, rooftop, or building integrated. In addition to national
incentives, some individual German municipalities offer production incentives funded by
surcharges on utility bills. All told, these programs have helped catapult the countrys PV
capacity from 76 MW in 2000 to nearly 8900 MW in 2009 [36], with about 3000 to 3800 MW of
that total installed in 2009 alone [25]. Much of the year-end rush was due to a steep reduction in
Germanys feed-in tariff (Figure 6-6).

75

A slight variation in compensation under Japans new FIT exists for PV systems above and below 10 kW.

6-11

Solar Photovoltaics

Figure 6-6
German PV installations between January 2009 and March 2010 showing surge of projects,
particularly larger than 10 kW, at years end in anticipation of declining feed-in tariff
Source: Solarenergie-Frderverein Deutschland e.V.

In January 2010, the feed-in tariffs were 0.4301/kWh ($0.56/kWh) for a typical 30-kW rooftop
installation, 0.4092/kWh ($0.53/kWh) for a 30- to 100-kW rooftop system, and 0.3194/kWh
($0.42/kWh) for ground installations. Due to the strong growth of solar PV installations in
Germany, the government lowered solar PV tariffs by 20-30 percent in June 2012, with 1%
digressions every month and a cumulative 52-GW capacity cap. In addition, subsidies were
eliminated for systems greater than 10 MW. The German PV market stayed steady through 2011
despite the declining feed-in tariff and political uncertainty, and over 5 GW were installed
through August 2012, putting the country on track for a third year of approximately 7 GW of
incremental new capacity.
6.2.3 Technology Performance and Cost Tables
As solar photovoltaic technology continues to evolve, significant cost reductions are expected to
continue as a result of improving power conversion efficiencies, development of low-cost cell
fabrication processes, increasing cell production volume with attendant economies of scale, and
reduced balance-of-plant costs through standardized modular power blocks.
The following subsections address the performance and cost of PV technologies using a twotrack approach. Distributed and utility-scale PV systems demand very different economic and
technical considerations. In particular, for distributed PV systems, the costs are typically covered
by a building owner or operator with significant support in the form of rebates or tax credits in
most cases. These systems require a more flexible and tailored approach. The information and
examples that follow are intended to illustrate some of the variables involved and an approach to
estimating system performance and cost through a range of options.
6-12

Solar Photovoltaics

6.2.3.1 Distributed PV Performance and Cost Estimates


Many customers considering installation of a distributed PV system are relatively sophisticated
energy consumers. In many cases, they have investigated their energy consumption patterns,
taken measures to improve energy efficiency and reduce usage, and developed a working
knowledge of PV operating principles, utility interconnection practices, net metering tariffs, and
financial incentive programs for solar electric power.
Nonetheless, because of the complexity (and novelty) of PV production models and financial
calculations, their investment decisions are often made without a complete understanding of the
true economic effects. To accurately evaluate the costs and benefits of a possible PV project, the
consumer must do the following:

Obtain detailed cost and performance data from the PV system or module manufacturer.

Acquire meteorological data from a representative ground station.

Simulate the production of the PV system, taking into account sun angles and module
orientation.

Determine hourly impacts on utility consumption.

Calculate monthly energy and/or demand savings using their specific utility tariff structure.

Calculate the financial impacts and applicability of various buy-down programs, tax credits,
and financing terms.

In recent years, the rooftop PV market has evolved so that many capable installers take
responsibility for such concerns on behalf of their clients, who need do little more than sign a
contract and write a check. Such customer service has minimized the effort that might have
deterred potential PV owners in the past and helped increase sales. Because the factors involved
vary so widely, it is impractical to attempt to construct a single table representative of all
distributed PV systems. One consumer tool that has been designed to overcome these difficulties
is the Clean Power Estimator. Section 6.2.5.2 provides sample calculations using the
Estimator.
6.2.3.2 Utility-Scale PV Plant Performance and Cost Estimates
The technology performance and cost tables for utility-scale central-station PV power plants are
based on the performance and cost estimates presented in the March 2012 EPRI report,
Engineering and Economic Evaluation of Central-Station Solar Photovoltaic Power Plants [35].
Tables 6-2 through 6-8 summarize the design and site assumptions and performance, capital and
operation and maintenance costs, and levelized cost of electricity estimates for conceptual 10MW central-station PV power plants for 22 combinations of six PV technologies and four sites
in the United States. The assumptions and results are addressed in the following sections.
Although the capital costs have changed significantly since the time of the study, the relative cost
and performance for different technologies and locations still offer useful comparisons.

6-13

Solar Photovoltaics

6.2.3.2.1 Design and Site Assumptions

Table 6-2 describes the six PV technologies that were analyzed. They are fixed-tilt flat plate
modules using crystalline silicon (c-Si), thin-film amorphous silicon (a-Si) , thin-film cadmium
telluride (CdTe) and thin film copper indium gallium (di)selenide (CIGS); single-axis tracking
modules using c-Si cells; and two-axis tracking, multijunction concentrating PV modules (CPV).
The system designs are based on current best practices for North American utility-scale
photovoltaic plants.
Table 6-3 provides information on the four PV project sites: Las Vegas, Nevada; Alamosa,
Colorado; Jacksonville, Florida; and Columbus, Ohio. They were chosen to be representative of
a range of site conditions in different regions of the United States. Region 1 is a location with
high direct normal insolation (DNI), little cloud cover and high ambient temperatures, typical of
the arid zones lower than 5000 feet in the desert Southwest (Arizona, California, Nevada, Texas
and New Mexico). Region 2 is a location with high DNI, little cloud cover and low ambient
temperatures, typical of Rocky Mountain areas with locations above 5000 feet in Arizona,
Colorado, Utah and New Mexico. Region 3 is a location with high global horizontal insolation
(GHI), dense cloud cover, high humidity and high ambient temperature. This region is
considered typical of southeast-coastal Texas, the southern States and Florida (excluding the
southernmost coastal areas). Region 4 is a location with low GHI, cold winters with variable
skies and hot humid summers with frequent dense cloud cover. This region is considered typical
of New England and the Northern and Midwestern states.
Table 6-2
PV technology summary
PV Technology
Fixed-Tilt Thin Film
Amorphous Silicon
Fixed-Tilt Thin FilmCadmium
Telluride
Single-Axis Tracking
Crystalline Silicon
Fixed Tilt Thin FilmCIGS

Abbreviation

Description

a-Si

Amorphous silicon modules mounted at a fixed 30degree tilt facing south

CdTe

Cadmium telluride modules mounted at a fixed 30degree tilt facing south

SAT c-Si

Monocrystalline modules on a north-south axis


tracker with backtracking

CIGS

CIGS modules mounted at a fixed 30-degree tilt


facing south

Fixed-Tilt Polycrystalline Silicon

c-Si

Polycrystalline modules mounted at a fixed 30-degree


tilt facing south

Concentrating Photovoltaic

CPV

Two-axis tracked high concentration photovoltaics

6-14

Solar Photovoltaics
Table 6-3
Photovoltaic site assumptions
Region #

Description

Site

Irradiance1

Ambient
Temperature

Cloud Cover

Desert Southwest

Las Vegas,
Nevada

High DNI

High

Low

Rocky Mountain

Alamosa,
Colorado

High DNI

Low

Low

South-Central U.S.

Jacksonville,
Florida

High GHI

High and humid

High

Northeastern and
Midwestern U.S.

Columbus,
Ohio

Low GHI

High in summer,
Low in winter

High in summer,
low in winter

DNI is direct normal insolation, GHI is global horizontal insolation

Although not analyzed in this study, coastal areas of the Pacific Northwest are expected to have
less favorable economics than the other four locations as they present low insolation and dense
cloud cover for a significant part of the year. However, as PV prices become more competitive
this region also is experiencing increased PV deployment activity.
The terrain conditions of each project site are assumed to be either flat or a location with a
topography and cover that does not incur a significant cost to clear and grade flat in preparation
for the construction of the PV power plant.
6.2.3.2.2 Plant Performance Estimates

Table 6-4 summarizes the estimated first-year power generation and annual capacity factors of
each of the 22 conceptual PV power plants. The annual capacity factor estimates represent the first
year of operation, after which they decline gradually as the PV modules age. They range from
14.4% for the fixed-flat plate CdTe and c-Si systems in Columbus, Ohio to 27.9% for the
concentrating PV system in Alamosa, Colorado. The assumed degradation rate is 0.75% per year.

6-15

Solar Photovoltaics
Table 6-4
First-year electricity generation and capacity factors
Technology

Fixed a-Si

Fixed CdTe

Single Axis Tracking


Crystalline

Fixed CIGS

Fixed Crystalline

Concentrating
Solar PV

Location

First Year MWh

AC Capacity Factor

Las Vegas, NV

21,513

24.6%

Alamosa, CO

20,052

22.9%

Jacksonville, FL

16,434

18.8%

Columbus, OH

13,963

15.9%

Las Vegas, NV

23,220

26.5%

Alamosa, CO

21,833

24.9%

Jacksonville, FL

18,145

20.7%

Columbus, OH

15,445

17.6%

Las Vegas, NV

24,517

28.0%

Alamosa, CO

24,862

28.4%

Jacksonville, FL

18,828

21.5%

Columbus, OH

16,408

18.7%

Las Vegas, NV

22,340

25.5%

Alamosa, CO

20,833

23.8%

Jacksonville, FL

17,120

19.5%

Columbus, OH

14,577

16.6%

Las Vegas, NV

21,153

24.1%

Alamosa, CO

19,760

22.6%

Jacksonville, FL

16,325

18.6%

Columbus, OH

14,099

16.1%

Las Vegas, NV

24,294

27.7%

Alamosa, CO

24,416

27.9%

6.2.3.2.3 Total Capital Requirement

Table 6-5 summarizes the total capital requirement estimates for the conceptual solar PV plants.
The reference date of the cost estimates is 4th quarter 2011.
The total capital requirement consists of the total plant cost, escalation and interest during
construction, and owners costs. The total plant cost includes PV equipment procurement and
transportation, balance-of-plant direct labor and material, and indirect and EPC construction
costs. The total plant investment is the sum of total plant cost and escalation and interest during
construction (also known as allowance for funds used during construction, or AFUDC). The total

6-16

Solar Photovoltaics

capital requirement is the sum of total plant investment and owners costs (due diligence,
permitting, legal, development, land, taxes, and fees). Larger projects generally benefit from
economies of scale and this is reflected in the cost estimates. The assumed land cost is $10,000
per acre.
The total capital requirements of the six PV technologies are assumed to be the same at the four
locations. As shown in Table 6-5, they range between $3,275/kW76 (net ac) for the fixed-tilt poly
crystalline (c-Si) systems and $3,900/kW for the single-axis tracking mono crystalline (c-Si)
systems.
6.2.3.2.4 Operation and Maintenance Cost

Operation and maintenance (O&M) cost consists of scheduled maintenance and cleaning,
unscheduled maintenance, inverter replacement reserve, and insurance, property taxes and
owners costs. As shown in Table 6-6, the O&M costs range from $19.7/kW-yr for the fixed-tilt
poly-crystalline (c-Si) system to $24.6/kW-yr for the two-axis tracking system (CPV). Scheduled
maintenance and module cleaning are the largest components of the O&M cost.
6.2.3.2.5 Levelized Cost of Electricity

The levelized cost of electricity (LCOE) is a measure of the life-cycle cost of electricity, and
provides a consistent basis for comparing the economics of the projects evaluated across all of
the technologies and locations. The LCOE is calculated using a model that sums the present
values of the annual O&M and capital-related costs and amortizes the result to yield the levelized
cost. The capital component of LCOE is a function of the total capital requirement and the
economic assumptions.
Table 6-7 summarizes the LCOE estimates for each of the technologies and locations, and both
with and without the 30% federal investment tax credit (ITC). The 30% ITC expires at the end of
2016 while the parallel non-expiring 10% investment tax credit for solar facilities will continue
beyond 2016. With the 30% ITC, the LCOE varies between $112/MWh for the two-axis tracked
high concentration PV/Alamosa, Colorado case and $195/MWh for the fixed tilt thin film
CdTe/Columbus, Ohio case. Without the tax credit, the LCOE varies between $175/MWh and
$305/MWh for the same technologies and locations.
All costs are in 4th quarter 2011 U.S. dollars, and the levelized costs assume 20-year plant life,
55/45 debt-equity ratio, 7.0%/yr interest on debt, 11%/yr return on equity, 3%/yr, inflation and
escalation of operation and maintenance costs, five-year MACRS tax depreciation, 40% income
tax rate, 30% investment tax credit, and no production tax credit.

76

According to market research analysts and PV installers, average megawatt-scale PV installation prices have
fallen to approximately $2,600/kW as of 4Q2012.

6-17

Solar Photovoltaics
Table 6-5
Utility-scale solar photovoltaic power plant total capital requirement estimates (4th Quarter 2011$)

PV Technology

Fixed-Tilt
a-Si $/kW

Fixed-Tilt
CdTe $/kW

Fixed-Tilt
CIGS $/kW

Single
Axis
Tracking
C-Si $/kW

Fixed Tilt
C-Si $/kW

2-Axis
Tracking
Conc. PV
$/kW

Direct Balance of Plant Costs


Site Prep

$363

$381

$312

$159

$263

Inverter + Installation

$303

$303

$303

$303

$303

DC Wiring

$517

$484

$437

$383

$349

Substation + MV System

$145

$145

$145

$145

$145

$15

$15

$15

$15

$15

$498

$606

$383

$910

$315

$1,842

$1,935

$1,595

$1,916

$1,390

Engineering + Management

$141

$148

$141

$158

$135

Contingency + Profit

$177

$186

$176

$198

$169

Total Indirect Balance of Plant Costs

$318

$334

$316

$356

$304

$1,014

$1,191

$1,338

$1,379

$1,379

$1,502

$3,173

$3,460

$3,250

$3,651

$3,074

$3,318

$0

$0

$0

$0

$0

$0

$3,173

$3,460

$3,250

$3,651

$3,074

$3,318

$239

$244

$224

$250

$202

$241

$3,412

$3,704

$3,474

$3,900

$3,275

$3,559

DAS/SCADA
Mounting System
Total Direct Balance of Plant Costs

$1,815

Indirect Balance of Plant Costs

Modules
Total Plant Cost
Escalation and Interest During Construction
Total Plant Investment
Total Owner's Cost
Total Capital Requirement

6-18

Solar Photovoltaics
Table 6-6
Utility-scale solar photovoltaic power plant operation and maintenance cost estimates (4th Quarter 2011$)
Fixed-Tilt
Plate a-Si
$/kW-yr

Fixed-Tilt
Plate CdTe
$/kW-yr

Fixed-Tilt
Plate c-Si
$/kW-yr

Fixed-Tilt
CIGS
$/kW-yr

Single Axis
Tracking
C-Si $/kW-yr

2-Axis
Tracking
Conc. PV
$/kW-yr

Scheduled Maintenance and Cleaning

$11.0

$10.5

$9.7

$10.6

$11.1

$14.6

Unscheduled Maintenance

$0.5

$0.5

$0.5

$0.5

$0.5

$0.5

Inverter Replacement Reserve

$6.0

$5.9

$5.9

$6.0

$5.9

$5.9

Subtotal O&M

$17.5

$16.9

$16.1

$17.1

$17.5

$21.0

Insurance, Property Taxes, Owner's Costs

$3.6

$3.6

$3.6

$3.6

$3.6

$3.6

Total O&M

$21.1

$20.5

$19.7

$20.7

$21.1

$24.6

PV Technology
Operation and Maintenance Cost

6-19

Solar Photovoltaics
Table 6-7
Levelized cost of electricity with and without 30% investment tax credit (4th Quarter 2011$)

Technology

Fixed a-Si

Fixed CdTe

Single-Axis
Tracking
Crystalline

Fixed CIGS

Fixed
Crystalline

Concentrating
Solar PV

6-20

LCOE With 30% ITC


($/MWh)

LCOE Without 30% ITC


($/MWh)

Las Vegas, NV

$134

$208

Alamosa, CO

$136

$210

Jacksonville, FL

$160

$249

Columbus, OH

$192

$299

Las Vegas, NV

$133

$206

Alamosa, CO

$132

$206

Jacksonville, FL

$156

$244

Columbus, OH

$195

$305

Las Vegas, NV

$131

$203

Alamosa, CO

$119

$188

Jacksonville, FL

$157

$246

Columbus, OH

$183

$287

Las Vegas, NV

$131

$202

Alamosa, CO

$131

$204

Jacksonville, FL

$156

$243

Columbus, OH

$185

$289

Las Vegas, NV

$128

$198

Alamosa, CO

$129

$201

Jacksonville, FL

$154

$240

Columbus, OH

$181

$282

Las Vegas, NV

$120

$183

Alamosa, CO

$112

$175

Location

Solar Photovoltaics

A probabilistic analysis of the LCOE is performed to generate a cumulative probability


distribution of the LCOE for each combination of PV technology and site. The first step in the
process is to perform a sensitivity analysis to identify the parameters that have the most impact
on the LCOE. The five highest sensitivity parameters are the capacity factor, capital cost, return
on equity, interest on debt, and O&M cost. Then, a probability distribution is developed for each
of the selected parameters based on the range of uncertainty of the parameter. Finally, a 10,000iteration Monte Carlo simulation is performed using the Crystal Ball risk analysis software and
the LCOE model to generate the cumulative probability as a function of the LCOE. Figure 6-7 is
an example cumulative probability distribution for the fixed flat-plate, crystalline-silicon cell/Las
Vegas, Nevada case. The chart shows the minimum, maximum and 25th, 50th, and 75th
percentile values of the LCOE.

Figure 6-7
LCOE probabilistic analysis: fixed a-Si modules, Las Vegas, NV (4th Quarter 2011$)

6.2.3.2.6 Performance and Cost Summary

Table 6-8 summarizes the performance, capital, and O&M cost estimates for the six PV
technologies.

6-21

Solar Photovoltaics
Table 6-8
Utility-scale solar photovoltaic power plant performance and cost estimate summary (4th Quarter 2011$)

Fixed-Tilt
c-Si $/kWyr

Fixed Tilt
Thin Film
CIGS
$/kW-yr

Single
AxisTracking
c-Si $/kW-yr

2-Axis
Tracking
CPV $/kW-yr

10 1 MW

10 1 MW

10 1 MW

10 1 MW

10 x 1 MW

2012

2012

2012

2012

2012

Fixed-Tilt
Thin Film
a-Si $/kWyr

Fixed-Tilt
Thin Film
CdTe $/kWyr

Plant Size (no of units unit size, MW)

10 1 MW

Available for Commercial Orders, Year

2012

PV Technology

2012

2012

2012

2012

2012

2012

Jan-12

Jan-12

Jan-12

Jan-12

Jan-12

Jan-12

$1,842

$1,935

$1,390

$1,595

$1.916

$1,815

$318

$334

$304

$316

$356

PV Modules

$1,014

$1,191

$1,379

$1,338

$1,379

$1,502

Total Plant Cost

3,173

$3,460

$3,074

$3,250

$3,651

$3,318

$0

$0

$0

$0

$0

$0

$3,173

$3,460

$3074

$3,250

$3,651

$3,318

$239

$244

$202

$224

$250

$241

$3,412

$3,704

$3,275

$3,474

$3,900

$3,559

$21.1

$20.6

$19.8

$20.8

$21.1

$24.6

Las Vegas, NV

22.6%

22.2%

22.4%

25.0%

26.9%

27.7%

Alamosa, CO

19.1%

20.7%

21.1%

25.3%

26.8%

27.9%

Jacksonville, FL

17.0%

16.9%

16.9%

18.4%

19.6%

Columbus, OH

14.6%

14.6%

14.4%

15.8%

16.9%

First Commercial Service, Year


Hypothetical In-Service Year
Total Capital Requirement ($/kW)

Direct Balance of Plant Cost


Indirect Balance of Plant Cost

Escalation and Interest During Construction


Total Plant Investment
Owner's Cost
Total Capital Requirement
Operation and Maintenance Costs

Costs for Hypothetical In-Service Year


Fixed, $/kW-yr
Power Generation
First-Year Annual Capacity Factor

6-22

Solar Photovoltaics
Table 6-8 (continued)
Utility-scale solar photovoltaic power plant performance and cost estimate summary

Fixed-Tilt
c-Si
$/kW-yr

Fixed Tilt
Thin Film
CIGS
$/kW-yr

Single
AxisTracking
c-Si $/kW-yr

3.8%

3.8%

3.8%

3.8%

1.0%

1.0%

1.0%

1.0%

2.0%

95%

95%

95%

95%

94%

Preconst., License & Design Time, Months

Idealized Plant Construction Time, Months

Unit Life, Years

20

20

20

20

20

Commercial

Commercial

Commercial

Commercial

Commercial

Simplified

Simplified

Simplified

Simplified

Simplified

Fixed-Tilt
Thin Film
a-Si $/kWyr

Fixed-Tilt
Thin Film
CdTe $/kWyr

Equivalent Planned Outage Rate

3.8%

Equivalent Unplanned Outage Rate, %


Equivalent Availability, %

PV Technology

2-Axis
Tracking
CPV $/kW-yr

Unit Availability

Technology Development Rating


Design & Cost Estimate Rating

6-23

Solar Photovoltaics

6.2.4 Performance and Cost History and Projections


Figure 6-8 shows the evolution of worldwide PV module average selling price from 1976
through 2011 in constant year-2011 dollars [20, 22]. The figure presents these data in the form of
an experience curve, where the average sales price is plotted versus cumulative capacity of
modules sold. This format was chosen to show the clear power-law behavior of historical module
prices, where the average selling price has declined by about 20% with each doubling of
capacity.
In the 35-year timeframe of the historic data in Figure 6-8, the total market size has grown about
100,000-fold while prices have fallen more than 97% and typical system efficiency has nearly
tripled. In contrast to the four orders of magnitude of historical progress, looking ahead to
achieving 1,000 GW of installed capacity entails only about 25-fold new growth. Straightforward
extrapolation of this trend for the next 30 years, while by no means guaranteed to be an accurate
forecast, implies another efficiency doubling and order-of-magnitude cost reduction while
reaching terawatts of PV deployments. Figure 6-9 shows the general trend of PV module selling
prices, including 2011 and 2012. For specific projections of future costs and performance, see
Tables 6-16 to Table 6-18.

PV Power-Module Global Average Sales Price


Average Sales Price
Year 2009 Dollars per Peak Kilowatt

$100,000

$10,000

$1,000

$100
0.0001

0.001

ASP data source: P. Mints, Navigant Consulting PV Service Practice

0.01

0.1

Cumulative Sales, GWP

Figure 6-8
Worldwide average PV module selling price vs. cumulative sales

6-24

10

100

1000
T.M. Peterson 09/23/10

Solar Photovoltaics

Figure 6-9
Trends in worldwide average PV module selling price vs. cumulative sales
Source: Navigant Consulting

6.2.5 PV Performance and Cost Estimate Model


The procedure and data tables that follow provide solar resource estimates and cost and
performance information for both distributed and central-station solar PV power generation.
6.2.5.1 Resource Estimates
Location is a fundamental consideration when estimating PV performance and cost. Table 6-9
presents site data for 27 representative locations in the continental United States.
The global latitude-tilt insolation rates (kWh/m2/yr) represent the annual solar energy that
reaches an equator-facing solar array oriented at the latitude tilt angle relative to horizontal; i.e.,
the angle equals the latitude coordinate of the site.
The global one-axis tracking insolation rates are 20% higher than the latitude-tilt rates, to
include the additional energy collected as the solar panel rotates around a single axis to follow
the sun. The axis is oriented in the north-south direction at the latitude tilt angle relative to
horizontal.
The direct-normal insolation rate excludes diffuse solar energy and includes only the direct
solar energy that can be collected by a concentrating lens and solar module that tracks the sun
via a two-axis tracking system.
Other site conditions include: dry-bulb temperature, which affects the PV cell operating
temperature and therefore conversion efficiency; maximum wind speed, which affects design
of the structural components; and the city cost factor, which affects capital and O&M costs.
6-25

Solar Photovoltaics

6.2.5.2 Distributed PV Cost and Performance Estimates


As described earlier, the Clean Power Estimator evaluates energy investments for a variety of
technologies, including PV, wind, solar thermal, energy efficiency, and fuel cells. Although the
discussion that follows focuses on the Estimator, EPRI is not endorsing this particular tool over
similar products that serve the same purpose. However, it provides a useful illustration of the
methods and variables involved.
The Estimator takes into account the specific characteristics of the customer purchasing the
system (such as the utility rate structure, local solar and environmental conditions, and local,
regional, and federal financial incentives) to provide a directly relevant personalized analysis.
Table 6-9
Site data for 27 locations in the continental United States
Solar Insolation

Location

Global
Latitude Tilt
(kWh/m2/yr)

Average Ambient Temperature

Global
Maximum
Direct
1-Axis
Wind Speed
Normal
Tracking
(mph)
(kWh/m2/yr)
(kWh/m2/yr)

Dry
Bulb
(C)

Wet Bulb
(C)

City
City
Cost
Index
Factor

El Paso, TX

2434

2921

2631

70

17.4

20.5

0.850

Albuquerque, NM

2401

2881

2566

90

13.8

18.9

0.927

Phoenix, AZ

2394

2873

2496

75

21.3

24.4

0.977

Carrisa Plains, CA

2345

2814

2445

70

20.1

26.5

1.100

Ely, NV

2230

2676

2377

74

6.7

15.6

1.062

Fresno, CA

2169

2603

2243

43

16.8

22.2

1.112

Dodge City, KS

2055

2466

2103

78

12.7

23.3

0.950

Santa Maria, CA

2039

2447

1947

60

13.8

18.3

1.130

Fort Worth, TX

1877

2252

1764

68

19.0

25.6

0.951

Columbia, MO

2058

2470

1552

54

12.5

25.6

0.987

10

Apalachicola, FL

1866

2239

1639

67

19.8

26.6

0.900

11

Brownsville, TX

1868

2242

1587

106

18.2

26.7

0.860

12

Omaha, NE

1767

2120

1652

109

9.7

25.6

0.950

13

Great Falls, MT

1715

2058

1641

82

7.2

17.8

0.952

14

Bismarck, ND

1707

2048

1641

72

5.2

22.8

0.880

15

Medford, OR

1699

2039

1582

60

11.6

21.1

1.080

16

Cape Hatteras, NC

1747

2096

1504

110

16.5

25.5

0.820

17

Miami, FL

1802

2162

1418

132

24.1

26.1

0.930

18

Charleston, SC

1690

2028

1354

76

18.2

27.2

0.836

19

Lake Charles, LA

1652

1982

1285

69

20.2

26.7

0.920

20

Nashville, TN

1589

1907

1279

73

15.2

25.6

0.843

21

Madison, WI

1554

1865

1304

77

7.2

25.0

0.939

22

6-26

Solar Photovoltaics
Table 6-9 (continued)
Site data for 27 locations in the continental United States
Solar Insolation

Location

Global
Latitude Tilt
(kWh/m2/yr)

Average Ambient Temperature

Global
Maximum
Direct
1-Axis
Wind Speed
Normal
Tracking
(mph)
(kWh/m2/yr)
(kWh/m2/yr)

Dry
Bulb
(C)

Wet Bulb
(C)

City
City
Cost
Index
Factor

Washington, DC

1559

1871

1287

78

12.1

25.6

0.970

23

Caribou, ME

1418

1702

1183

76

3.8

21.7

0.885

24

Boston, MA

1421

1705

1160

87

10.7

23.9

1.073

25

New York, NY

1402

1682

1087

113

12.5

24.4

1.127

26

Seattle, WA

1299

1559

999

65

10.6

19.7

1.051

27

Source: EPRI [2]

A critical aspect of evaluating economic feasibility is obtaining the correct solar resource, utility
tariff, and incentive data applicable to the specific customer. The Estimator contains databases
for this purpose that are regularly updated to ensure that Estimator licensees are using accurate
data. Table 6-10 describes the contents of the Estimator databases.
Many of the Estimator databases are indexed to locations. Once a location is selected, the
databases provide available utility tariffs for that location. The Estimator is normally configured
to use a default tariff, but it allows the user to select from all tariffs offered by the utility. In
addition, the selected location determines the available federal, state, and utility economic
incentive programs, the solar resource data, income tax rates, and air emissions rates.
Table 6-10
Information included in Clean Power Estimator databases
Database

Description

Solar Resource

Hourly typical solar irradiance by month (12 24 values); U.S. database is


derived from the TMY-2 data and includes 237 locations

Utility Incentives

The buy-down and incentives provided by utilities; can include a variety of


cost, size, and tax treatment constraints

State Incentives

Incentives provided by state governments or other jurisdictions below the


federal level

Federal Incentives

Incentives provided at the federal level

Electricity Tariffs

Over 1300 tariffs for residential and commercial customers, including net
metering tariffs; provides wide flexibility in capturing the range of tariff
structures, including seasonal variations, time of day variations, metering
charges, energy pricing, demand pricing, energy and demand tiers, sell-back
pricing, fuel cost adjustments, public programs, and other special charges;
tariffs are updated using the Estimators semi-automated monitoring tools

Electricity Load Profiles

Hourly profiles based on independent utility research showing consumption


patterns for various locations and customer types, such as all-electric heating,
gas water heating, etc.

State Income Tax Rate


Schedules

Tax rate schedules reflect tax filing status, such as personal income for
married couples filing jointly, individuals, corporations, etc., for all states

6-27

Solar Photovoltaics
Table 6-10 (continued)
Information included in Clean Power Estimator databases
Database

Description

Federal Income Tax


Rate Schedules

Similar information as above, except for federal taxes

Depreciation Schedules

Depreciation schedules such as the Modified Accelerated Cost Recovery


System (MACRS) of the U.S. tax code used by businesses to depreciate their
PV capital investment

Environmental
Emissions Factors

Air emissions (CO2, SO2, and NOX) offset rates by state based on regional mix
of sources, such as coal, gas, diesel, hydro, and nuclear

The case studies below show how cost estimates are derived for rooftop PV systems of both
residential and commercial sizes in five different geographical, regulatory, and tariff situations in
the United States. These examples demonstrate that factors such as incentives and rate structures
can be as important as local solar resource to customer economics when deploying a distributed
PV system.
6.2.5.3 Hypothetical Case Studies
Assumed capital costs are based on the median total system cost (per kWdc nameplate rating) for
completed California Solar Initiative (CSI) incentive applications with a reservation request
review date of April through September 2010, rounded to the nearest $100 per kW [47].
Residential systems are assumed to have a total installed cost of $6,500 per kWdc, and
commercial systems have a cost of $6,000 per kWdc. (Table 6-17 illustrates the cost impacts of
various utility, state, and federal incentives.) The net capital cost on a CEC-rating basis ($/kWac)
is also included for reference. The tabulated values are the net costs to consumers making the
investments, after all incentives and tax impacts are taken into account.
The Estimator methodology was applied to 10 case studies located in five utility territories across
the country. These represent a range of meteorological conditions, electric rates, net metering
availability, and incentive policies. The 10 cases are defined in Table 6-11.
Table 6-11
Rooftop case-study electric-rate scenarios

6-28

State

Case Name

Tariff

NY

LIPA Residential

Residential Service (Codes 180, 183, 186)

NY

LIPA Commercial

General Service-Large Secondary (Code 281)

OH

FirstEnergy Residential

Residential Service

OH

FirstEnergy Commercial

General Service

AL

Alabama Power Residential

Residential Service (Rate FD)

AL

Alabama Power Commercial

Medium Light and Power (Rate LPM)

AZ

SRP Residential

Basic Plan (Rate E-23)

AZ

SRP Commercial

General Service (Rate E-36)

CA

PG&E Residential

Residential Service (Rate E1 Area X)

CA

PG&E Commercial

Medium General Service (Schedule A-10-Secondary)

Solar Photovoltaics

The assumed annual electric bill for each case-study customer was $1200 for residential
systems and $20,000 for commercial systems. Using these values along with specific rate details
(fixed charges, demand charges, energy charges), seasonal variations, and hourly load profiles,
the corresponding annual consumptions were was calculated.77 Table 6-12 shows annual
consumption and comparative average electricity prices (including all cost components) for the
10 scenarios.
Systems are assumed to have a dc rating (i.e., the sum of all the PV modules Standard Test
Condition [STC] nameplate ratings) of 5 kW for residential systems and 10 kW for commercial
systems. System performance was modeled using actual equipment characteristics assuming
south facing module orientation with tilt angle equal to the local latitude and no shading.78
CEC-AC system ratings and estimated annual performance by location are shown in Table 6-13,
and capital costs are given in Table 6-14.

77

Load profiles by region are included in the Clean Power Estimator database. The data include relative hourly
values by month (12 month 24 hour matrices). All days within a given month are assumed to be identical.
78
PV modules selected were BP Solar model MST-50, rated at 50 W (100 modules for residential and 200 modules
for commercial). Inverters were Xantrex model PV-5208, rated at 5 kW, and PV-10208, rated at 10 kW, for
residential and commercial systems, respectively.

6-29

Solar Photovoltaics
Table 6-12
Rooftop case-study utility rates
Residential

Commercial

LIPA

First
Energy

Alabama
Power

SRP

PG&E

LIPA

First
Energy

Alabama
Power

SRP

PG&E

Assumed Annual Electric Bill ($)

1,200

1,200

1,200

1,200

1,200

20,000

20,000

20,000

20,000

20,000

Annual Energy Consumption


(kWh)

6,215

6,576

10,319

10,620

7,511

126,434

142,407

199,559

219,838

130,435

Average Price ($/kWh)

0.193

0.182

0.116

0.113

0.160

0.158

0.140

0.100

0.091

0.153

Net Metering Offered

Table 6-13
Rooftop case-study system performance
Residential

Commercial

LIPA

First
Energy

Alabama
Power

SRP

PG&E

LIPA

First
Energy

Alabama
Power

SRP

PG&E

42

40

33

33

37

42

40

33

33

37

System Rating
(kWac, CEC-AC*)

4.47

4.47

4.47

4.47

4.47

8.57

8.57

8.57

8.57

8.57

Annual Energy Produced (kWh)

6,162

5,545

6,624

8,322

7,274

11,805

10,623

12,691

15,997

14,056

Capacity Factor (%)

15.7

14.2

16.9

21.3

18.6

15.7

14.2

16.9

21.3

18.7

Latitude (N)

* The CEC-AC rating is calculated from the PV-USA Test Conditions (PTC) module dc rating (1000-W/m2 insolation, 20C ambient temperature, 10-m
above grade, 1-m/s wind) and the system inverter efficiency. Although the CEC-AC rating does use the more real world PTC measure of module
performance, rather than the industry-standard STC rating, it ignores other typical system losses, such as module mismatching and field-wiring resistance.
Therefore, it is still an optimistic measure of total system output in most cases.

6-30

Solar Photovoltaics
Table 6-14
Rooftop case-study capital costs
Residential

Commercial

LIPA

First
Energy

Alabama
Power

SRP

PG&E

LIPA

First
Energy

Alabama
Power

SRP

PG&E

Assumed Capital Cost


($/kWdc) (STC basis)

6,500

6,500

6,500

6,500

6,500

6,000

6,000

6,000

6,000

6,000

Assumed Capital Cost


($/kWac) (CEC-rating basis)

7,270

7,270

7,270

7,270

7,270

7,000

7,000

7,000

7,000

7,000

Capital Cost ($)

32,500

32,500

32,500

32,500

32,500

60,000

60,000

60,000

60,000

60,000

Incentives
LIPA Buydown

(17,500)

OH PV Rebate

(15,000)

(30,000)

CSI Buydown-PG&E EPBB

(1,549)

(2,993)

AZ SRP SolarWise Energy

(15,000)

(10,037)

AZ Tax Credit

(1,000)

(6,000)

NY Tax Credit

(5,000)

Federal Tax Credit

(9,750)

(5,250)

(9,750)

(8,250)

(9,285)

(18,000)

(18,000)

(18,000)

(18,000)

(18,000)

Taxes on Incentives
LIPA Buydown

13,633

CSI Buydown-PG&E EPBB

4,457

AZ SRP SolarWise Energy

11,778

AZ Tax Credit

2,465

NY Tax Credit

1,650

Net Capital Cost


Net Capital Cost ($)

19,400

12,250

22,750

18,580

21,666

31,316

23,883

42,000

31,877

40,025

Net Capital Cost ($/kWac)

4,340

2,740

5,089

4,157

4,847

3,654

2,787

4,901

3,720

4,670

6-31

Solar Photovoltaics

Finally, Table 6-15 shows the economic results, in terms of both simple payback and the internal
rate of return. These results vary enormously, with simple payback ranging more than a factor of
4, from 13 to 57 years.
The best possible results would benefit from all four primary driving factors:

High retail electricity price to offset

Availability of net metering

High system performance (good local solar resource)

Availability of incentives

The least economic residential case of those studied is the one at Alabama Power. Although its
performance (capacity factor of 16.9%) is in the middle of the range, it has nearly the lowest
average retail electricity price ($0.116/kWh), net metering is not offered, and there are no
incentives other than the federal investment tax credit that is available to all systems.
Table 6-15
Rooftop case-study economic-analysis results
Residential

Commercial

LIPA

First
Energy

Alabama
Power

SRP

PG&E

LIPA

First
Energy

Alabama
Power

SRP

PG&E

Utility Bill Savings


First Year ($)

1122

970

402

779

1146

2024

1659

1406

1419

2285

Simple Payback
(years)

17

13

57

24

19

15

14

30

22

18

IRR (%)

7.0

10.4

<0

4.2

6.1

15.3

20.9

7.6

11.8

11.7

The most economically attractive case studied is the commercial system in First Energys
territory. Although its performance is the lowest of the group (14.2% capacity factor), the utility
does offer net metering, the incentives available are the most generous, and the retail electricity
price is above the commercial-case average. It is particularly interesting to compare the LIPA
and SRP commercial cases. They have comparable incentive levels so that their net costs are
similar, and they both offer net metering. However, SRPs far better solar resource (21.3%
versus LIPAs 15.7% capacity factor) fails to offset the advantage of LIPAs higher electric
rate so that the LIPA system is a rather more attractive investment.
6.2.5.4 Central Station PV Cost and Performance Methodology
This section presents and illustrates a methodology to estimate the performance and cost of
central-station PV power plants using fixed flat-plate, one-axis, and two-axis tracking. Tracking
systems maximize the power output of a particular PV array but also add to its cost and
mechanical complexity. The source of the 2010 PV cost and performance estimates is the 2010
EPRI report, Engineering and Economic Evaluation of Central-Station Solar Photovoltaic
Power Plants [42].

6-32

Solar Photovoltaics

The resulting estimates represent technology available for commercial order in a given year and
site conditions that include annual insolation, average ambient temperature, and a city cost index.
Because solar PV technology is continually evolving, the projections are based on an assumed
future technology development and market scenario. Hence, these performance and cost
projections are subject to considerable positive and negative uncertainty, especially in the out
years. To evaluate the economics of future central-station PV, three possible technologies are
evaluated below and each is evaluated for various years. For fixed flat-plate PV systems, the
example assumes a scenario of thin-film PV technology evolution, based initially on singlejunction copper indium-gallium diselenide or cadmium telluride technology and evolving to
multi-junction technology involving copper indium diselenide, cadmium telluride, and perhaps
other materials in later years. The overall system efficiency (solar energy to ac power) will
increase from approximately 10% in 2010 to 14.2% by the year 2030. This assumes that the
commercial module efficiency increases from 12% in 2010 to 17% by 2030.
For the one-axis tracking flat-plate PV systems, the development scenario assumes crystalline
silicon PV technology. The overall system efficiency (solar energy to ac power) will increase
from 14.7% in 2010 to 20% by the year 2030. This assumes that the commercial module
efficiency increases from 17% in 2010 to 24% by 2030. Note that the addition of one-axis
tracking also increases the solar input by 20%, relative to the fixed flat-plate system.
For the two-axis tracking, high-concentration PV system, the development scenario assumes that
the technology uses a point-focus, gallium arsenide cell system with 1000x concentration. The
overall system efficiency will increase from 21.7% in 2010 to 40% by 2030. The corresponding
cell efficiencies range from 27% in 2010 to 50% in 2030. The rated outputs are 17,000 kWac in
2010 and 68,000 kWac in 2030.
The capital cost of solar PV power plants is a function of the performance of the solar collectors,
the plant rating and number of modules, and the delivered cost of the modules, arrays, trackers,
controls, power conversion system, and the other balance-of-plant components. O&M cost is a
function of the solar collector area, the complexity of the solar tracking and control system, and
the failure and replacement rates of the plant components. As the technology advances over time,
performance improves, component costs drop, and the market for PV technology will grow,
significantly reducing unit capital and O&M costs.
Tables 6-16 through 6-18 provide estimates of cost and performance for each the three PV
system examples as a function of the year of commercial order. The data are extrapolated from
the May 2010 Engineering and Economic Evaluation of Central-Station Solar Photovoltaic
Power Plants report [42] using staff judgment based on historic PV cost and performance trends.
Performance data include system conversion efficiencies and rated outputs (Wac/m2). Cost data
include collector area and projections of total plant cost ($/m2 and $/kWac), O&M cost ($/m2/yr
and $/kWac/yr), and land lease cost (% of revenues, $/kWh) for two kWac output ratings: (1) the
nominal output rating of plants to be installed in a given year; and (2) scaled to an output rating
of 5 MWac. Table 6-19 lists scaling factors that are used to scale the solar collector area, total
plant cost, and O&M cost estimates to different plant ratings.

6-33

Solar Photovoltaics
Table 6-16
Solar PV power plant capital, O&M, and lease cost projections for
fixed flat-plate thin-film photovoltaic power plants (Dec 2009$)
Commercial Order Year

2020

2030

System Efficiency

13%

15%

130

150

50,000

200,000

380,000

1,300,000

$180

$110

TPC ($/kWac)

$1,400

$730

O&M ($/m2/yr)

$3.0

$2.0

O&M ($/kWac/yr)

$23

$13

Collector Area (m2)

38,000

33,000

Total ($/kWac)

$1,500

$810

O&M ($/kWac/yr)

$32

$23

Land Lease, % of Revenues

3%

3%

Rated Output (Wac/m )


Plant Rating (kWac)
2

Collector Area (m )
Total Installed Cost ($/m2)

Costs Scaled to 5 MW

Table 6-17
Solar PV power plant capital, O&M, and lease cost projections for one-axis tracking
crystalline silicon flat-plate photovoltaic power plants (Dec 2009$)
Commercial Order Year

2020

2030

System Efficiency

18%

20%

180

200

50,000

200,000

280,000

1,000,000

$250

$150

TPC ($/kWac)

$1,400

$750

O&M ($/m2/yr)

$4.50

$3.00

$25

$15

Collector Area (m2)

28,000

25,000

Total ($/kWac)

$1,500

$830

O&M ($/kWac/yr)

$35

$26

Land Lease, % of Revenues

3%

3%

Rated Output (Wac/m )


Plant Size (kWac)
2

Collector Area (m )
Total Installed Cost ($/m2)

O&M ($/kWac/yr)
Costs Scaled to 5 MW

6-34

Solar Photovoltaics
Table 6-18
Solar PV power plant capital, O&M, and lease cost projections for two-axis
tracking high concentration photovoltaic power plants (Dec 2009$)
Commercial Order Year

2015

2020

2030

System Efficiency

28%

32%

40%

Rated Output (Wac/m2)

240

270

340

20,000

50,000

200,000

84,000

180,000

590,000

1000

1000

1000

$600

$390

$250

$2,500

$1,400

$740

$10.00

$6.50

$5.00

$42

$24

$15

Collector Area (m2)

21,000

18,000

15,000

Total ($/kWac)

$2,800

$1,700

$920

O&M ($/kWac/yr)

$51

$33

$25

Land Lease, % of Revenues

3%

3%

3%

Plant Size (kWac)


2

Collector Area (m )
Concentration Ratio
2

Total Installed Cost ($/m )


TPC ($/kWac)
2

O&M ($/m /yr)


O&M ($/kWac/yr)
Costs Scaled to 5 MW

Table 6-19
Solar PV power plant capital and O&M cost scaling factors vs. rated plant output relative
to a 5-MWac plant
Total Plant Cost Factor
Rated
Output
(kWac)

O&M Cost Factor

Fixed
Flat Plate

1-Axis
Flat Plate

2-Axis
HCPV

Fixed
Flat Plate

1-Axis
Flat Plate

2-Axis
HCPV

10

2.33

2.33

2.55

2.75

2.75

2.51

50

1.87

1.87

2.12

2.12

1.98

100

1.7

1.7

1.8

1.89

1.89

1.78

500

1.37

1.37

1.41

1.45

1.45

1.41

1000

1.24

1.24

1.27

1.3

1.3

1.27

5000

10,000

0.97

0.97

0.95

0.9

0.9

0.91

20,000

0.95

0.95

0.89

0.82

0.82

0.82

50,000

0.92

0.92

0.83

0.72

0.72

0.73

100,000

0.9

0.9

0.8

0.63

0.63

0.65

200,000

0.9

0.9

0.8

0.57

0.57

0.59

6-35

Solar Photovoltaics

6.2.5.4.1 Example Performance Calculations

As described in the steps below, the annual power generation is calculated by multiplying the
annual insolation estimate for the site from Table 6-9 by the system efficiency from Tables 6-16
through 6-18. The system efficiency allows for soiling, energy collection, power conversion,
step-up transformer, availability, and other losses within the power plant, but excludes substation
losses.
The performance estimating procedure involves the following steps:
1. Obtain insolation and other site data from Table 6-9 and the system efficiency and rated
output corresponding to the PV technology and year of commercial order from Table 6-16
for fixed flat-plate, Table 6-17 for one-axis tracking flat-plate, and Table 6-18 for two-axis
tracking concentrator technologies.
2. Calculate the required solar collector area (m2) by dividing the kW rating of the plant at
standard conditions (kWac) by the rated output (Wac/m2) corresponding to the year of
commercial order:
Collector Area (m2) = [(Plant Rating, kWac)/(Rated Output, Wac/m2)] x 1000
3. Calculate the annual power generation by multiplying the annual insolation (kWh/m2/yr)
by the system efficiency (%) and the collector area (m2):
Annual Power Generation (kWh/yr) =
4. (Annual Insolation, kWh/m2/yr) (System Efficiency, %) (Collector Area, m2)
For a 1000-kWac one-axis tracking flat-plate solar PV power plant, ordered in 2010 and a plant
site near El Paso, Texas, the following steps are used:
1. The one-axis tracking insolation rate for El Paso is 2921 kWh/m2/yr % (from Table 6-8).
For the one-axis tracking flat-plate PV system ordered in 2010, the Rated Output is 147
Wac/m2 and the system efficiency is 14.7% (from Table 6-17).
2. The required solar collector area is: Collector Area (m2) = (1000 kWac)/(147 Wac/m2) 1000
= 6,803 m2
3. The net annual power generation is:
Annual Power Generation (kWh/yr) =
(2921 kWh/m2/yr) (14.7%) x (6,803 m2) = 2,921,000 kWh/yr
6.2.5.4.2 Example Capital and O&M Cost Calculations

The total plant cost includes all direct and indirect construction costs, sales tax, engineering and
procurement costs, and contingencies. The O&M cost includes all labor and material required to
operate and routinely maintain the plant as well as allowances for periodic repair and
replacement of failed modules and other components. If applicable, the lease cost covers the cost
of leasing the land for the plant site.

6-36

Solar Photovoltaics

The cost estimating procedure involves the following steps:


1. Estimate the total plant cost (TPC, $/kWac) by multiplying the base total plant cost
(base TPC, $/kWac) corresponding to the PV technology and year of commercial order
(from Table 6-16 for fixed flat-plate, Table 6-17 for one-axis tracking flat-plate, and Table
6-18 for two-axis tracking concentrator technologies) by the city cost index from Table 6-9
and the TPC scale factor from Table 6-19:
TPC ($/kWac) = (Base TPC, $/kWac) (City Cost Index) (TPC Scale Factor)
2. Estimate the total O&M Cost ($/kWac/yr) by multiplying the base O&M Cost ($/kWac/yr)
from Tables 6-16, 6-17, or 6-18 by the City Cost Index from Table 6-8 and the O&M Scale
Factor from Table 6-19: O&M ($/kWac/yr) = (Base O&M, $/kWac/yr) (City Cost Index) x
(O&M Scale Factor)
3. If the land is leased by the project, estimate the annual land lease cost by multiplying the
average power revenue rate ($/kWh) by the lease rate from Tables 6-13 through 6-15: Lease
Cost ($/kWh) = (Power Revenue, $/kWh) (Lease Rate, %)
Continuing with the example for a 1000-kWac one-axis tracking flat-plate solar PV power plant,
ordered in 2010, for a plant site near El Paso, Texas, the cost calculations are as follows:
1. The City Cost Index for El Paso is 0.850 (from Table 6-9) and the TPC and O&M scale
factors are 1.24 and 1.30 (from Table 6-19)
2. For a 2010 commercial order, one-axis tracking flat-plate technology, and 10,000-kWac
uniform plant rating, the Total Plant Cost is $4,171/kWac, the O&M cost is $60/kWac/yr,
and the annual land lease cost is 3% of power revenue (from Table 6-18).
Thus: TPC ($/kWac) = ($4,171/kWac) (0.85) (1.24) = $4,396/kWac
3. O&M ($/kWac/yr) = ($4.43/kWac/yr) (0.85) (1.22) = $4.59/kWac/yr. Assuming the power
revenue averages $0.05/kWh or $50/MWh, the land lease cost is:
Land Lease ($/MWh) = ($50/MWh) x (3%) = $1.50/MWh

6.3 Environmental Issues


Although PV operation emits no gases or sounds and generally presents less environmental
impact per deployed megawatt than any other known generation technology, it is not completely
free from impacts and potential hazards. Most notably, some impacts may be expected during
system manufacture involving handling of potentially toxic or flammable materials. These
resemble the hazards encountered in the semiconductor industry, and the PV industry has taken
advantage of many of the approaches practiced there. Other issues exist for thin-film systems in
terms of ultimate disposal or recycling. These issues may prove problematic for deployment of
those specific PV technologies in BIPV applications in the long term, but for large-scale utility
deployments they may be readily addressed in cradle-to-cradle recycling schemes [4].
One significant benefit of PV is its lack of noise. PV is a solid-state technology that operates
completely silently, making it unique among generation options and particularly among
competing distributed generation technologies. The fact that PV does not require noise
abatement could, in some cases, help offset its cost or argue for its use when other generation
technologies would not be permitted or desirable.
6-37

Solar Photovoltaics

Land-use concerns for most PV deployments are minimal. In operation, PV produces no liquid
or solid waste and requires no fuel handling. In most applications, PV is installed on a building
rooftop, a parking structure, or available nearby land. Building-integrated PV is incorporated
directly into a buildings roofing, siding, or glass surfaces.
Only large utility-scale, industrial, or military PV installations demand large areas of dedicated
land. In most cases, such land consists of inhospitable, desert-like terrain of little use for
alternative development but which nevertheless raises concerns about habitat disruption. In late
2009, potential conflicts between solar power and wildlife habitat were spotlighted with Sen.
Dianne Feinsteins (D-CA) introduction of the California Desert Protection Act of 2010 (S.2921),
which at this writing in December 2012 is pending. The act sets aside approximately 2.5 million
acres of the Mojave Desert for conservation by expanding Death Valley National Park, Joshua
Tree National Park, and the Mojave National Preserve, as well as establishing the new Mojave
Trails National Monument and Sand to Snow National Monument. Several solar power projects
mostly solar thermal but also photovoltaichad been slated for siting on land that will be off
limits to development if the bill passes. Press reports claimed that as many as 13 solar power
project developers cancelled proposals rather than work around the restrictions. However, of
possible benefit to solar projects, Section 203 of the bill addresses renewable energy and requires
the Bureau of Land Management, Department of Defense, and U.S. Forest Service to undertake
environmental studies and identify federal land where such projects could be expedited. In July
2012, the Bureau of Land Management released a roadmap for solar energy development on
federal lands in seventeen zones across six states. The Programmatic Environmental Impact
Statement (PEIS) intends to streamline permitting and provide other incentives for large-scale
solar projects in regions identified as having good solar resource, transmission, and other favorable
project development characteristics. In general, environmental impacts on native plants and
wildlife can be minimized with proper planning and management.
Interestingly, the land-use impacts of PV can compare very favorably with those of other
generation options such as hydroelectric power. For example, one can contrast the output of
Hoover Dam to that of a hypothetical PV installation occupying the same land area as Lake
Mead (which is formed by the Colorado River behind Hoover Dam and provides the water to
turn its hydro turbines):
(Typical Annual Insolation Value) (PV Module Efficiency) (Area of Lake Mead) =
(2000 kWh/m2) (10%) (688.6 million m2) = 138 billion kWh per year
Hoover Dam generates an average of four billion kWh per year; its maximum annual net
generation was 10.35 billion kWh in 1984. Thus, on average, Hoover Dam/Lake Mead produces
just 3% as much electricity as would PV occupying an equivalent amount of land. This comparison
is merely illustrative of the potential energy that could be generated using a given amount of land
area, and obviously does not take into account the many other costs and benefits involved.

6-38

Solar Photovoltaics

6.4 Design and Deployment Issues


Because distributed and utility-scale PV systems employ essentially the same technology at
different scales, many of their design and deployment issues are the same. Several fundamental
details must be considered when planning a PV project of any scale:

Project sizeFor most power-generating technologies, large systems are more cost-effective
per unit capacity. However, due to its modular construction, economies of scale have a much
smaller effect on PV system cost than on the cost of conventional power plants such as
combustion turbines.
Demonstration versus production systemsDemonstration projects are more expensive per
unit of capacity than production systems and, because they are research oriented and employ
new or untested equipment, they can have lower reliability as well.
Overall project planAlthough most structures with the proper orientation and design (roof
support, drainage, etc.) can be readily retrofitted to accept PV, it is preferable to design a
building, subdivision, or business park to optimize the placement of PV from the start.
Structures can be designed with the future deployment of PV in mind even if it is not included
at the time of construction. Although it is also possible to retrofit structures for buildingintegrated PV, incorporated into roofing materials, window glass, facades, and so on, its full
economic and technological benefits are best realized when it is included from design through
completion.
Type of installationGround-mounted systems may benefit from economies of scale
because they can be larger and heavier than roof-mounted systems. However, roof-mounted
systems or building-integrated PV may benefit from savings in land costs or provide
ancillary benefits such as shade or insulation. More sophisticated mounting systems that
move a PV array to track the suns motion on either one axis or two axes can increase the
arrays power output but also add to its cost and mechanical complexity.

PV technologyAlternative technologies should be evaluated to determine which best suit


the system needs.

Inverter technologyBecause PV devices generate dc power, electronic interfacing devices


(inverters) are used to convert the power into ac for connection to the grid. Inverter reliability
has been an issue for grid-connected PV systems for the past two decades, making inverter
replacement or repair a leading O&M cost component. This situation has been improving,
but careful selection of the specific inverter for any large installation is warranted.

Use of storageBecause PV is an intermittent resource, storage technologies are necessary


to produce dispatchable or firm solar power, but they add cost to the system. Where net
metering policies are in place, grid-tied systems often forego storage because they can
transfer excess energy to the grid, essentially using it as a 100%-efficient storage system.

Type of developerPV projects can be designed and installed by electric utilities, energy
service companies, solar companies, and private individuals. The system developers
background and experience can make a significant impact on a projects success. For
example, using high-quality components and proper installation practices, an experienced
system integrator can improve the likelihood that a project will be trouble free.

6-39

Solar Photovoltaics

Ownership and advocacyPV has a variety of champions. It makes a difference who


initiated the project under consideration and who are the project owners and team members.

Non-energy benefits (ancillary benefits)Beyond power, PV can provide additional value in


some applications. Building-integrated photovoltaics can provide insulation, shading, hot
water, and extended roof life among other benefits.

Distribution system issuesUtility distribution companies have concerns about the safety
and power-quality impact of connecting PV to distribution feeders. These interconnection
questions involve such issues as islanding protection, fault contributions, and voltage
regulation. For most small PV systems, recent IEEE and state utility commission standards
have made interconnection more straightforward.

Interconnection and inspection requirementsAlthough they are increasingly standardized,


such requirements still vary greatly among utilities and jurisdictions. Some critics contend
that utility distribution companies and some jurisdictions create barriers to PV by imposing
unnecessarily stringent interconnection and inspection requirements.

Availability and competitive cost of grid-supplied electricityPV is recognized as costeffective for many off-grid energy applications that are remote from a distribution line. The
technology has advanced to the point where some on-grid applications may also prove costeffective in high-price utility locations when coupled with various tax incentives.

Financial assumptionsPV system energy cost is impacted by issues involving insurance,


taxes, incentives, rebates, green pricing, return on investment, O&M, and capital costs. For
example, depending on the lifespan assumptions, maintenance assumptions, and the discount
rate used in an analysis, PV energy costs can vary by as much as a factor of two for the same
basic system design at the same location. Costs must be compared on an equivalent basis. Of
course, the local cost of utility-supplied electricity is also frequently part of the economic
analysis as mentioned above.

6.4.1 Utility-Scale Issues


The design and deployment of utility-scale PV systems resembles a simple distribution system
construction project in terms of wiring complexity and civil engineering. The main differences
are that most of the wiring involved is at lower voltages and currents than those in typical
distribution systems. Modern inverters typically do not require an additional transformer, at least
up to tens of kilowatts; however, for larger-scale and higher-than-distribution-level voltages, a
transformer is desirable.
6.4.2 Operating and Maintenance Labor Requirements
Distributed and utility-scale PV systems share many of the same O&M requirements. In general,
PV is the lowest-maintenance generation technology available. However, PV power plants are
not maintenance free; they require a regimen of continual monitoring, periodic inspection,
scheduled preventive maintenance, and service calls, among other tasks. Tracking systems
require periodic inspections to ensure proer operation of a few moving parts, and inverter
replacement or repair is a leading contributor to O&M cost. Broadly speaking, the specifics of
effective O&M strategy are varied, and depend upon a number of environmental, policy-related,
and organizational factors. Investment in labor and maintenance activities is dependent on
6-40

Solar Photovoltaics

system size and location (e.g., water availability, climate and weather conditions, travel
distances, customer vs. utility property), plant technology and architecture (e.g., panel and
inverter types, fixed vs. tracked, performance ratio thresholds), ease of site access (e.g., ground
mount vs. roof mount), as well as the extent that meters, inverters, and monitoring equipment are
deployed at a site, among other factors.
Although occasional panel washing may be necessary, in most climates and locations natural
precipitation limits losses due to dust and grime accumulation to about 10% to 15%. PV modules
are manufactured to withstand a hail-impact test. For glass-encapsulated modules, the primary
environmental hazard is vandalism in cases where its potential is not minimized by the site
design. Lightning is also an infrequent hazard. A typical rooftop PV system located on a home or
business can operate with literally no maintenance for years (as most do). PV O&M strategies
focus on the following three areas:

Preventive maintenance, which entails routine inspection and servicing to prevent


breakdowns and unnecessary production losses. Although increasingly popular, preventive
maintenance programs entail moderate upfront costs and extra labor expense.

Corrective or reactive maintenance, which addresses equipment breakdowns after the fact.
The current industry standard break-fix strategy allows for low upfront costs but also
greater risk of component failure and higher back-end costs. It can be lessened by effective
preventive maintenance and condition-based maintenance strategies.

Condition-based maintenance (CBM), which uses real-time data to prioritize and optimize
maintenance and resources. Increasing numbers of third-party providers are developing CBM
plans to improve O&M efficiency, but with some upfront costs for communications and
monitoring software and hardware.

A major distinction among the three approaches is that preventive maintenance requires O&M
action perhaps once or twice per year, whereas corrective or reactive maintenance as needed is
typically less frequent and condition-based maintenance is continuous. Overall, the PV industry
is trending toward O&M approaches that promote greater oversight and management capability.
An EPRI survey of PV O&M knowledge and practices [42] concluded that the O&M costs for
PV systems smaller than 1 MW have ranged from $6/kW to $27/kW, or from less than 1% to 7%
of total system cost over the life of a PV project. While the sample size for the survey was small,
the wide cost range indicates diverse approaches to PV O&M. For utility-scale PV plants, O&M
costs were estimated to vary with the type of technology and tracking schemes involved. Those
estimates are summarized in Table 6-20.

6-41

Solar Photovoltaics
Table 6-20
Utility-scale PV power plant O&M cost estimates (2011$)
O&M Costs
($/kW-yr)

Fixed-Tilt
c-Si

Fixed-Tilt
CdTe

Fixed-Tilt
a-Si

Single-Axis
Tracking c-Si

Fixed Tilt
CIGS

Scheduled
Maintenance/Cleaning

9.7

10.5

11.0

11.1

10.6

Unscheduled
Maintenance

0.5

0.5

0.5

0.5

0.5

Inverter Replacement
Reserve

5.9

5.9

6.0

5.9

6.0

16.1

16.9

17.5

17.5

17.1

3.6

3.6

3.6

3.6

3.6

19.7

20.5

21.1

21.1

20.7

SUBTOTAL O&M
Insurance, Property
Taxes, Owners Costs
TOTAL O&M

2-axis
Tracking
CPV
14.6
0.5
5.9
21.0
3.6
24.6

The optimal and most cost-beneficial means for operating and maintaining PV systems to reduce
downtime and maximize output are still being developed as the industry gains experience. As
this information is documented and disseminated best practices can be identified and adopted.
6.4.3 Grid Connection
Only a very small proportion of PV-equipped facilities operate completely disconnected from the
electric grid, most often for reasons of distance and cost to connect or the owners personal
philosophy of self reliance. Grid-connection provides significant technical and economic
advantages for a PV power system and consequently the majority, more than 95%, of residential,
commercial, industrial and third party PV plants are grid connected.
The electric grid enables renewable power systems to deliver energy and meet varying demand
without the need for a separate energy storage system. Connection also provides a path to the
larger electricity market for PV energy. By virtue of many interconnected sources and loads the
grid provides stability of supply with the capacity to give and take energy when demand changes
or overloads occur. It simplifies the balancing of variations in supply and demand of individual
distributed generators over a wide area. It also provides the quality of supply needed for PV and
end-use equipment, maintaining and regulating the voltage level and power frequency.
A healthy grid is good and necessary for widespread PV deployment; however, because the grid
has been designed, built, and operated based on a centralized generation approach, much of it is
not ready for a high penetration of distributed generation. With continued PV deployment there
is a need to define what a future and more PV-ready distribution grid will look like and how it
will evolve. Grid modernization, concepts that are often labeled as smart grid, can enhance the
value of distributed renewable power while enhancing distributed load control and overall
electric system operating efficiency.
With this evolution in PV penetration, grid planning and operation must also evolve. From
todays requirements in IEEE 1547-2003, Standard for Distributed Resources Interconnected
with Electric Power Systems, interconnection and operating rules are expected to change with
PV penetration level. From a few percent of demand, where PV is not expected to affect
6-42

Solar Photovoltaics

distribution, to significant deployment where PV is need to provide voltage and energy


regulation support, operating requirements are evolving and connection standards are being
revised. In between there remain a few issues for integrating PV into the distribution grid.
Unintentional islanding is perhaps the most significant and commonly raised concern with
respect to PV, primarily because it can endanger the safety of line workers and the public.
Islanding is a situation in which a portion of the utility system operates separately from the rest
of the system. Line crews working on a section of line they believe to be de-energized may
unexpectedly encounter line voltage and could be electrocuted. Similar danger extends to the
public in situations with downed conductors or other live wires within reach that would have
normally been de-energized by upstream utility switchgear had an island not developed. The
most common way to prevent unintentional islanding is to use voltage and frequency relays on
DG units, set to trip whenever voltage or frequency migrate outside a selected window. This
form of islanding protection is called passive protection, and prevents islanding in most cases.
Most existing and proposed interconnection requirements stem from fundamental issues related
to providing a compatible interface. These include safety, reliability, quality, and potentially
damaging interactions. Some requirements originate with distribution system operators and take
the form of interconnect agreements, permits, or public service commission rulings. Industry
standards, such as those promulgated or being developed by the IEEE, may be adopted in whole
or in part through such rulings. Local inspection authorities may impose other requirements to
fulfill their obligation to protect the public and comply with building and electric codes. Five
basic functions include electrical isolation via power transformer, controlled connection and
disconnection, a visible and secure disconnect, short-circuit protection, and surge protection.
These functions are common to most interconnection policies or agreements.
Other potential issues related to interconnection include changes in radial feeder power flows,
possible reverse power flow in distribution networks, loss of effective voltage regulation,
harmonic injection and distortions, voltage fluctuation and flicker, and overcurrent protective
device coordination. These issues are addressed in existing connection standards and will evolve
with new standards and practices designed for higher penetration levels.
The idea of PV operating without the grid, albeit only once in a while, is receiving more
attention. In the face of a number of recent weather-related blackouts, the question Why cant
my PV system provide backup power when the grid is unavailable? is being asked. Such
intentional islanding or backup micro-grid operation is being considered. However,
implementation will require a step change in operating rules and requirements if PV were to
provide some limited energy without batteries and without the stability of the grid. For this
separation both the distributed PV system and the grid must go through a paradigm shift in
operating philosophy and requirements. The good news is that technology evolutions on both
sides of the meter may enable the additional service of distributed PV.
In the meantime, for interconnected PV, local, state, and federal policies increasingly allow and
even encourage interconnecting distributed generation with the grid. Net metering programs in
many states allow PV owners to feed short-term excess power they generate to the grid, which is
colloquially called running the meter backward. It is important to note that net metering
policies vary greatly among jurisdictions and that only some allow customers to sell electricity
back to their host utility should they generate more power than they consume. Some programs
offer payment or credit at rates less than the retail price, such as the avoided cost rate.
6-43

Solar Photovoltaics

6.5 Equipment Markets and Key Participants


In the United States, the PV market is also evolving rapidly. State government mandates for PV,
the promulgation of uniform interconnection standards that in most cases allow net metering, and
growing utility involvement in the sector are making PV more attractive to manufacturers,
investors, and potential buyers. Some private companies have established business units to target
specific markets, such as California or New Jersey. In some cases, demand had outstripped
supply and resulted in transient delays and price increases. This situation was repeated on a
global scale starting in 2005 and continued into 2008 as module demand outstripped the worlds
current production capacity of polysilicon feedstock for wafers. That situation was fully resolved
by 2009, and led to oversupplies of PV inventory at the beginning of the year that have since
contributed to greatly reduced module prices.
6.5.1 Equipment Markets
In the domestic PV market, five major market sectors can be identified:
1. Grid-connected residential, commercial, and industrial rooftop or building-integrated
applications that rely on incentives offered by state programs as well as the 30% ITC. In the
United States, the commercial sector alone now accounts for nearly half of annual PV sales,
and remains dependent on incentives [16].
2. U.S. DOE plus other civilian government agencies committed to using PV and renewables
through efforts such as the Federal Energy Management Program.
3. U.S. Department of Defense, which has been installing both grid-independent and gridconnected PV systems at military facilities for years. DOD has deployed PV at several military
bases, including a 14-MW system at Nellis Air Force Base near Las Vegas. The department
also procures several megawatts of PV per year for mobile applications for use by personnel in
the field. In late 2009, Secretary of the Navy Ray Mabus set the goal of producing at least 50%
of Navy and Marine Corps shore-based energy from renewable sources by 2020.
4. Grid-independent buyers of PV systems for off-grid homes, ranches, or farms. An estimated
100,000 off-grid residential PV installations are operating today, most of them supplied by
retail companies, service providers, system integrators, or mail-order organizations.
However, as the number of grid-connected commercial, industrial, and residential
installations grows, this sector is becoming much smaller in proportion to the entire market.
5. Grid-independent commercial buyers, which typically use PV to power remote
communications installations, environmental measurement and monitoring sites, and security
systems.
Figure 6-10 illustrates how PV cells and modules made in the United States are distributed by
market sector and end use, according to the most recent data available. According to the most
recent data from EIA, the total amount of PV produced for all uses in the United States in 2009
was 601.1 MW, a 15% increase above the 2008 output of 524.3 MW [16]. The electric power
sector had the greatest increase of PV usage at 50% above 2008 levels. Both the industrial and
transportation sectors showed decreases, with transportation having the greatest decrease at 94%
below 2008 levels.

6-44

Solar Photovoltaics

Figure 6-10
U.S. shipments of PV by market sector (top) and end use (bottom) in 2009
Source: EIA [16-53]

6-45

Solar Photovoltaics

Internationally, the World Bank, the Global Environmental Facility, regional development
banks, and other national and international programs have made funds available for developing
renewable technology systems, particularly in developing countries such as Indonesia and Brazil.
This financial and technical support has helped provide PV-based systems to areas where electric
service has never been available. The mix of PV applications in different countries varies
according to their particular electricity infrastructures, local demand, incentives and policy
support, and other variables as demonstrated by Figure 6-11. For example, note the
overwhelming fraction of PV dedicated to grid-connected applications in Italy (ITA) due to the
country aggressive solar program, whereas countries such as Australia (AUS) have many offgrid communities and fewer policies supporting deployment.

(MEX = Mexico, CAN = Canada, ISR = Israel, AUS = Australia, SWE = Sweden, DNK =
Denmark, FRA = France, AUT = Austria, CHE = Switzerland, PRT = Portugal, JPN = Japan, ITA
= Italy, CHN = China)
Figure 6-11
Distribution of installed PV applications in countries participating in the International
Energy Agency (IEA) Photovoltaic Power Systems Program in 2011
Source: IEA [35]

6.5.2 Key Participants


The PV industry is made up of a large number of organizations that can be separated into two
categories: component manufacturers and infrastructure organizations. The first category designs
and fabricates PV cells, modules, and balance-of-system hardware. The second category consists
of integrators that help identify customers, design and install systems, and provide maintenance
services. There may be a few hundred such companies nationwide. Most serve off-grid
customers, but the number and size of those involved in the grid-connected market is growing.
Today, hundreds of PV manufacturers and suppliers are providing components to customers
around the world. The majority are located in China, Europe, Japan, and the United States. In
2009, for the first time, China took the lead as the worlds leading PV supplier, led by companies
such as Suntech Power, Yingli Solar, and JA Solar. China and Taiwan now produce the majority
6-46

Solar Photovoltaics

Capacity (MW)

of the worlds PV capacity (Figure 6-12) with 20 GW of capacity in China in 2011, almost
double the previous year [29]. Note that shipments by region do not coincide with demand. The
majority of installations are in Europe. Many of the leading PV producers are vertically
integrated organizations that refine basic cell materials, produce cells, and assemble modules.
Another type of organization is the module assembly firm, which purchases PV cells from
various manufacturers and adds value by producing modules. A list of the leading PV
manufacturers is provided in the next section.

Figure 6-12
Annual solar cell production by country
Source: Data compiled by the Earth Policy Institute from GTM Research,
http://earth-policy.org/indicators/C47 [43]

6.5.3 Resources
6.5.3.1 Internet Sites
The reader is referred to the following Internet sites for further information. Although these sites
contain information and contacts that may be of interest, EPRI cannot vouch for the accuracy of
their content.
U.S. DOE Energy Office of Efficiency and Renewable Energy
www.eere.energy.gov
U.S. National Renewable Energy Laboratory (NREL)
www.nrel.gov/csp/

6-47

Solar Photovoltaics

Solar Electric Power Association (SEPA)


Formerly the Utility PhotoVoltaic Group (UPVG), SEPA is a nonprofit collaboration comprising
more than 700 utilities, energy service providers, and the PV industry working to create and
encourage commercial use of new solar electric power technology and business models. SEPA
also created and maintains the Solar Data and Mapping Tool, a web-based utility that provides
project data for solar power projects in the United States (www.solarelectricpower.org/solartools/solar-data-and-mapping-tool.aspx).
www.solarelectricpower.org
American Solar Energy Society (ASES)
ASES is a national organization dedicated to advancing the use of solar energy for the benefit
of U.S. citizens and the global environment. The society sponsors conferences and publishes a
magazine and white papers on the subject. ASES is the U.S. section of the International Solar
Energy Society (ISES).
www.ases.org
www.ises.org
National Center for Photovoltaics (NCPV)
NCPV is headquartered at the National Renewable Energy Laboratory (NREL) in Colorado.
Its primary members comprise PV researchers affiliated with NREL and Sandia National
Laboratories.
www.nrel.gov/pv/ncpv.html
Solar Energy Industries Association (SEIA)
Established in 1974, SEIA is the national trade association of the U.S. solar energy industry. As
the voice of the industry, SEIA works with its 1000 member companies to make solar a
mainstream and significant energy source by expanding markets, removing market barriers,
strengthening the industry and educating the public on the benefits of solar energy.
www.seia.org
Energy Information Administration
Part of DOE, the EIA collects, analyzes, and disseminates information about energy sources,
usage, forecast, and other data, generally focused on the United States.
www.eia.doe.gov
University Centers of Excellence for Photovoltaics Research and Education
The U.S. DOE established PV centers of excellence at Georgia Tech and the University of
Delaware to improve the fundamental understanding of the science and technology of advanced
PV devices, create breakthrough PV technology, provide training, and give the United States a
competitive edge in PV development.
www.ece.gatech.edu/research/UCEP
www.udel.edu/iec/

6-48

Solar Photovoltaics

6.5.3.2 PV Manufacturers
Note: In rankings compiled by Solar Outlook and Photon International, the companies listed
below were among the largest PV manufacturers in 2010, worldwide. Worth noting is that PV
cell production increased 118% from 2009 levels with almost half of the cell manufacturing
being produced in China. The ranges of production estimates reported here reflect differences in
the source data. These companies are included in this list of contacts for informational purposes
only; no endorsement is implied.
Suntech Power
Suntech, based in Wuxi, China, surpassed 2009 industry leader First Solar to become the top PV
cell producer at 1.58 GW in 2010. In 2009, Suntechs production was only 704 MW, and in 2011
it is 2.4 GW [35].
www.suntech-power.com
JA Solar
JA solar is also based in China, specifically Shabei, China. In 2009 the company produced
approximately 509520 MW of multi-crystalline PV cells. According to the IEA, the 2011
production was 2.8 GW [35]. JA Solars rise is particularly impressive, as the company rose
from being sixth largest manufacturer in 2009 to the second largest in 2010.
www.jasolar.com
First Solar
A manufacturer of CdTe cells/modules, First Solar increased its production from 60 MW in 2006
to approximately 1981 MW in 2011 [35]. It became the first pure thin-film manufacturer to top
the PV production list in 2009. The company is based in Tempe, Arizona.
www.firstsolar.com
Yingli Solar
Yingli is headquartered in Baoding, China. Although founded in 1998, the company only began
producing PV modules in 2002, and in 2009 produced around 525 MW. Production in 2010
increased to over 1,000 MW.
www.yinglisolar.com
6.5.4 Applicable Codes and Standards
Both customers and distribution companies anticipate that adoption of universal standards will
address interconnection requirements, ease the introduction of PV into the marketplace, and
eliminate the complex and sometimes time-consuming or expensive tangle of practices currently
followed. Several respected organizations are leading the drive to create such standards,
including IEEE, Underwriters Laboratories (UL), and the Federal Energy Regulatory
Commission (FERC).
A trend in this area is growing consensus that earlier codes and standards addressing PV
inverters and interconnection were better suited for an environment in which PV on the grid was
rare. As PV becomes commonplace, it appears that more robust and intelligent inverters could be
very helpful in stabilizing the grid during disturbances, rather than immediately disconnecting as
most standards require. This perspective is expected to inform future work on relevant codes and
standards, although it has not yet emerged in the form of new drafts or working group activities.
6-49

Solar Photovoltaics

The following list summarizes the content and status of key interconnection standards already
promulgated or under development. Each standard relies on the precedents the others have set. In
many instances, standards issued by different organizations have been developed cooperatively.
IEEE 929-2000: the IEEE Recommended Practice for Utility Interface of Residential and
Intermediate Photovoltaic Systems, which describes the interface, functions, and requirements
necessary to interconnect a PV power system with the electric grid. It also describes acceptable
and safe practices for accomplishing those functions.
UL 1741: the Standard for Safety for Static Inverters and Charge Controller for Use in
Photovoltaic Power Systems, is closely related to IEEE 929-2000. In fact, UL and IEEE worked
together to ensure that testing procedures described in UL 1741 ensure inverter compliance with
the guidelines established in IEEE 929-2000.
IEEE 1547: the Standard for Distributed Resources Interconnected with Electric Power
Systems. It is seen as a critical milestone for the DG industry. IEEE 1547 provides a uniform
interconnection standard that states and energy companies throughout the U.S. and the world are
expected to use as the basis for their own interconnection practices. The recent Energy Policy
Act of 2005 also includes language encouraging states to adopt IEEE 1547 as the basis for their
interconnection standards.

6.6 References
1. Renewable Energy Technology Characterizations. EPRI and U.S. Department of Energy:
1997. TR-109496.
2. Engineering and Economic Evaluation of Central-Station Photovoltaic Power Plants.
EPRI, Palo Alto, CA: 1992. TR-101255.
3. Capital Requirement and Economic Summary, 50 MW Photovoltaic Fresnel Lens
Concentrator Plant, Scientific Analysis, Inc., EPRI Project RP 2948-8, March 5, 1991.
4. Moskowitz, P.D., National PV Environmental, Health and Safety Information Center:
Bibliography, Brookhaven National Lab, Upton, NY: 1993. Report 11973.
5. Renewable Energy 2000: Issues and Trends, U.S. Energy Information Administration:
February 2001. DOE/EIA-0628.
6. Renewable Energy Annual 2000, U.S. Energy Information Administration: March 2001.
DOE/EIA-0603.
7. Renewable Power Industry Status Overview. EPRI, Palo Alto, CA: 1998. TR-111893.
8. Photovoltaic Five-Year Market Forecast 20002005, Strategies Unlimited: April 2000.
Report PM-48.
9. 2000 Photovoltaic Industry Competition Analysis, Strategies Unlimited: July 2000.
Report PC-11.
10. Annual Solar Thermal and Photovoltaic Manufacturing Activities Tables, 2000,
U.S. Energy Information Administration, www.eia.doe.gov.
11. Distributed Renewable Energy Generation Impacts on Microgrid Operation and Reliability.
EPRI, Palo Alto, CA: 2002. 1004045.
6-50

Solar Photovoltaics

12. CERTS Customer Adoption Model, Consortium for Electric Reliability Technology
Solutions, prepared for the Transmission Reliability Program, Office of Power Technologies,
U.S. Department of Energy by Lawrence Berkeley National Laboratory, Berkeley,
California: March 2001. LBNL-47772.
13. Assessment of Rooftop and Building-Integrated PV Systems for Distributed Generation.
EPRI, Palo Alto, CA: 2003. 1004204.
14. Photovoltaic Manufacturer Shipments 20002002, Strategies Unlimited: August 2002.
Report PM-51.
15. Bolinger, M., and R. Wiser, Support for PV in Japan and Germany, Case Studies of State
Support for Renewable Energy, Lawrence Berkeley National Laboratory, September 2002.
16. Annual Energy Review 2009, U.S. Energy Information Administration (EIA), August
2010. DOE/EIA-0384(2009). See also
http://tonto.eia.doe.gov/cneaf/solar.renewables/page/solarreport/solarpv.html.
17. Hoff, T.E., TAG Case Studies for Clean Power Estimator Model of Distributed PV.
Draft report to EPRI, Clean Power Research, October 2002.
18. Sun Screen II: Investment Opportunities in Solar Power, Credit Lyonnais Securities Asia:
July 2005.
19. Case Studies of Grid-Connected Photovoltaic Systems, Volume 2. EPRI, Palo Alto, CA:
2003. 1004203.
20. Analysis of Worldwide Markets for Photovoltaic Products & Five-Year Application Forecast
2006/2007, Navigant Consulting: August 2007, Report NPS-GLOBAL2.
21. Sharp, T., New Energy for Japan, Cogeneration & On-Site Power Production, James &
James Ltd., London, U.K.: September-October 2003.
22. Photovoltaic Manufacturer Shipments & Profiles, Strategies Unlimited: September 2003.
SUPM 53.
23. Photovoltaic Balance-of-Systems and U.S. Grid-Connected System Price Analysis 2005,
Strategies Unlimited: June 2005. Report PM 56.
24. Green, M.A., Third Generation Photovoltaics: Advanced Solar Energy Conversion,
Springer-Verlag, Berlin (2003).
25. Solar Outlook, Navigant Consulting: February 29, 2010. Issue SO2010-1.
26. U.S. Solar Industry Year in Review, Solar Energy Industries Association (SEIA), April 2010.
27. Laying the Foundation for a Solar America: The Million Solar Roofs Initiative, DOE Office
of Energy Efficiency and Renewable Energy (EERE), October 2006.
28. Engineering and Economic Evaluation of Central-Station Solar Photovoltaic Power Plants.
EPRI, Palo Alto, CA: 2012. 1025005.
29. Development of Renewable Energy Sources in Germany in 2009, Federal Ministry for the
Environment, Nature Conservation and Nuclear Safety: March 2010.
30. California Solar Initiatives website,
http://www.californiasolarstatistics.ca.gov/reports/monthly_stats/.
6-51

Solar Photovoltaics

31. National Survey Report of PV Power Applications in Japan 2009, International Energy
Agency Cooperative Program on Photovoltaic Power Systems and New Energy and
Industrial Technology Development Organization (NEDO): May 2010.
32. Addressing Solar Photovoltaic Operations and Maintenance Challenges: A Survey of
Current Knowledge and Practices. EPRI, Palo Alto, CA: July 2010. 1021496.
33. Mills, A. and R. Wiser, Implications of Wide-Area Geographic Diversity for Short-Term
Variability of Solar Power, Lawrence Berkeley National Laboratory, LBNL-3884E.
September 2010.
34. Hoff, T. and R. Perez, Quantifying PV Power Output Variability, conditionally accepted for
publication in Solar Energy, 2010.
35. Trends in Photovoltaic Applications: Survey Report of Selected IEA Countries between 1992
and 2011, International Energy Agency, Paris, France: August 2012. IEA-PVPS T1-21:2012.
36. Engineering and Economic Evaluation of Central-Station Solar Photovoltaic Power Plants.
EPRI, Palo Alto, CA: 2012. 1025005.
37. REN21 2012, Renewables 2012 Global Status Report (Paris: REN21 Secretariat).
38. U.S. Solar Market Trends 2011, Interstate Renewable Energy Council (IREC): August 2012.
39. Thin-Film 20122016: Technologies, Markets, and Strategies for Survival.
http://www.greentechmedia.com/research/report/thin-film-2012-2016/. GTM Research:
2012.
40. Annual Energy Review 2010, U.S. Energy Information Administration (EIA): August 2010.
DOE/EIA-0384(2009).
41. Analysis of Worldwide Markets for Solar Products and Five-Year Application Forecast:
2011/2012. Navigant Consulting, Report NPS-Global7: August 2012.
42. Engineering and Economic Evaluation of Central-Station Solar Photovoltaic Power Plants.
EPRI, Palo Alto, CA: May 2010. 1021320.
43. U.S. Solar Market Insight: 2nd Quarter 2012. GTM Research: 2012.
44. U.S. Solar Photovoltaic Manufacturing: Industry Trends, Global Competition, Federal
Support. Congressional Research Service: June 2012.

6-52

GEOTHERMAL ENERGY

7.1 Introduction
Geothermal energy is literally energy in the form of heat contained within the Earth. It arises
from the Earths molten interior and occurs mainly in geologically active areas where the
planets continental plates meet. It is very accessible in certain areas, such as the Ring of Fire
that bounds the Pacific Ocean (Figure 7-1). These regions are often characterized by volcanism,
earthquakes, hot springs, and/or geyser activity. However, in recent years geothermal resources
are also being used in central China, Utah and other areas considerably distant from geologic
plate boundaries. Table 7-1 presents an overview of geothermal energy.
Table 7-1
Geothermal energy overview
Installed Capacity (est.
as of Q1 2012)

11,224 MW worldwide
3,187 MW in the United States

Technology Readiness

Ranging from commercial for shallow hydrothermal to early research for


hot dry rock
Moderate-low confidence in cost estimates and projections beyond
commercial range of technology

Environmental Impact

Economic Status

Some atmospheric emissions and solid waste that can be controlled


Possible long-term impacts on local hydrology and possible related
subsidence without careful reservoir management
Can help meet local wastewater disposal challenges by pumping treated
water to recharge subterranean aquifers
Significant water use may be necessary to stimulate and charge EGS
reservoirs
Hydrothermal is competitive in many markets, although in the United
States it is limited geographically to the Western states
Binary technology has potential to be more widespread and economical,
but needs further experience
Subsurface technology needs improvement to move beyond prime areas
in the West
Re-injection of treated water into geothermal reservoir has been shown
to be economically feasible to prolong life of the wells and mitigate
subsidence
High-grade enhanced geothermal systems (EGS) (hot dry rock) are likely
to become competitive in the near future, with reasonable investments in
R&D for improvements and cost reductions in resource development and
deep drilling

7-1

Geothermal Energy
Table 7-1 (continued)
Geothermal energy overview
Policy Status

Trends to Watch

Currently being assessed in most states where potential growth of


available resources adds significant capacity to state RPS
Included with other non-hydro renewables
Public RD&D funds in critical areas to promote geothermal technologies
Improved drilling and new exploration technologies
Lower cost for binary-cycle conversion technologies, lower effective
temperatures
Better characterization and identification of geothermal resources.
Enhancement of geothermal systems' waste water re-injection to
increase capacity of current plants
Re-injection of treated water into geothermal reservoir
Limit to near-term growth potential driven primarily by economics,
permitting challenges, and the lack of economically feasible resources
near transmission facilities
Development of plant designs for higher resource temperatures to the
supercritical water region that would lead to an order-of-magnitude gain
in both reservoir performance and heat-to-power conversion
Short-term development of shallow, high-grade EGS sites at the margins
of hydrothermal reservoirs
Use of co-produced hot water from existing oil and gas operations using
binary cycle technology
Development of first of kind EGS reservoirs in hot dry rock settings

Figure 7-1
Map illustrating the Ring of Fire [38]

7-2

Geothermal Energy

Geothermal power production is a renewable technology with a worldwide operating capacity of


more than 11,000 MW [42]. Geothermal reservoirs have been a commercial reality in Italy,
Japan, the United States, Iceland, New Zealand and Mexico for many decades. According to the
Energy Information Administration (EIA), the United States is the world leader in electricity
production from geothermal resources with approximately 15,364 GWh (in 2011). California
alone has approximately 2,615 MW of geothermal power plants that provide roughly one-third of
the states non-hydro renewable capacity [43], exceeding that of every country in the world.
However, the known high-grade pure steam (i.e., not water-steam) reservoir (the Geysers) is
already largely exploited. Future geothermal power generation will depend on locating and
tapping new water-steam, hot pressurized water, enhanced (or engineered) geothermal systems
(EGS), and hot dry rock (HDR) resources.
Geothermal power plants are base loaded, dispatchable and range in size from 0.02 to 180 MWe,
with 30 to 60 MW considered standard for steam or flashed-steam plants. For binary cycle
plants, smaller sizes are more common, in the range of 15 to 45 MW. Energy production is not
affected by daily or seasonal resource supply fluctuations, but may degrade over time due to
reservoir degradation. However for dry cooled binary plants, both daily and seasonal
fluctuations related to ambient temperature swings are observed. Atmospheric emissions are low
or non-existent, and solid wastes are readily managed by conventional techniques. In addition to
use for electric power generation technologies, which are the subject of this report, hot
geothermal fluids can be used directly in solution mining, aquaculture, greenhouses, space
heating, wood drying, and vegetable drying. Today the U.S. non-electric energy uses consists of
a very small portion of geothermal usage compared to electric power applications.
Two limiting factors in geothermal power development are the high risk of resource exploration
and development, particularly when geothermal power competes against lower-cost fossil fuels,
and time delays from leasing and permitting requirements, which can add significant costs to
projects.

7.2 Resources and Current Installed Capacity


As a general rule, geothermal resources are classified into three categories: hydrothermalconvection, geo-pressured, and HDR resources. Hydrothermal-convection resources are
subdivided further into vapor and liquid-dominated resources, which produce mostly steam
and hot water, respectively. They occur as a result of heat transfer from geologically active
high-temperature belts to aquifers in close proximity. Geo-pressured resources are hot water
containing dissolved methane under a high subsurface pressure about twice that of normal
hydrostatic pressure. HDR resources are hot rock masses that lack fluid content but are close
enough to the surface to have potential for commercial heat extraction.
7.2.1 Hydrothermal Geothermal Resource Assessment
Currently, hydrothermal-convection resources account for all electricity generation from
geothermal energy in the United States. Its major physical characteristics are the resource
temperatures, depths, and subsurface permeability, which controls the flow of hot water from
the hydrothermal aquifer into wells drilled into a subsurface geothermal reservoir.

7-3

Geothermal Energy

7.2.1.1 Resource Temperature


How a resource is developed depends on its temperature: hot, defined as greater than 400F
(204C); moderate, 300 to 400F (149204C); or low, 194 to 300F (90149C). Hot resources
are generally developed via dual flash (two-stage flash) technology and all low-temperature
and many moderate-temperature resources will be developed via binary technology. Both
technologies are described in detail in Section 7.3 of this report, the geothermal technology
section.
In 2005, a new proposed temperature-based geothermal resource classification system was
developed at the request of the Geothermal Energy Association (GEA) and DOE [29]. With a
new national resource assessment on the horizon, this proposal would further subdivide the
resource temperatures into seven categories as shown in Table 7-2.
The seven resource classes in Table 7-2 are based on temperature as the primary criterion, and
steam fraction in the mobile fluid phase in the reservoir (but not necessarily steam saturation in
the reservoir) as the secondary criterion. The attributes of the reservoirs in the various classes, as
listed in Table 7-2, reflect general industry experience.

7-4

Geothermal Energy
Table 7-2
Proposed geothermal resource temperature classification
Class of
Resource

Reservoir
Temperature

Mobile
Fluid
Phase in
Reservoir

Production
Mechanism

Fluid State at
Wellhead

Well Productivity and


Controlling Factors Other
than Temperature

Applicable
Power
Conversion
Technology

Unusual
Development or
Operational
Problems

Class 1:
Non-electrical
Grade

< 100C

Liquid
water

Artesian
self-flowing
wells;
pumped
wells

Liquid water

Well productivity dependent


on reservoir flow capacity
and static water level

Direct Use

Class 2:
Very Low
Temperature

100C to
< 150C

Liquid
water

Pumped
wells

Liquid water
(for pumped wells);
steam-water mixture
(for self-flowing wells)

Typical well capacity 2 to 4


MWe; dependent on reservoir
flow capacity and gas content
in water; well productivity
often limited by pump
capacity

Binary

Class 3:
Low
Temperature

150C to
< 190C

Liquid
water

Pumped
wells; selfflowing wells
(only at
the higher
temperature
end of the
range)

Liquid water
(for pumped wells);
steam-water mixture
(for self-flowing wells)

Typical well capacity 3 to 5


MWe; dependent on reservoir
pressures, reservoir flow
capacity and gas content in
water; productivity of pumped
wells typically limited by
pump capacity and pump
parasitic power need;
productivity of self-flowing
wells strongly dependent on
reservoir flow capacity

Binary; Twostage Flash;


hybrid

Calcite scaling
in production
wells and stibnite
scaling in binary
plant are
occasional
problems

Class 4:
Moderate
Temperature

190C to
< 230C

Liquid
water

Self-flowing
wells

Steam-water mixture
(enthalpy equal to
that of saturated
liquid
at reservoir
temperature)

Well productivity highly


variable (3 to 12 MWe);
strongly dependent on
reservoir flow capacity

Single-stage
Flash; Two
stage Flash;
Hybrid

Calcite scaling in
production wells
and occasional
problem; aluminosilicate scale in
injection system
a rare problem

7-5

Geothermal Energy
Table 7-2 (continued)
Proposed geothermal resource temperature classification
Class of
Resource

7-6

Reservoir
Temperature

Mobile
Fluid
Phase in
Reservoir

Production
Mechanism

Fluid State at
Wellhead

Well Productivity and


Controlling Factors Other
than Temperature

Applicable
Power
Conversion
Technology

Unusual
Development or
Operational
Problems

Class 5:
High
Temperature

230C to
< 300C

Liquid
water;
Liquiddominated
two phase

Self-flowing
wells

Steam-water mixture
(enthalpy equal to or
higher than that of
saturated liquid at
reservoir
temperature)
saturated steam

Well productivity highly


variable (up to 25 MWe);
dependent on reservoir flow
capacity and steam
saturation

Single-stage
Flash; Hybrid

Silica scaling in
injection system
occasionally
corrosion;
occasionally high
NCG content

Class 6:
Ultra-High
Temperature

300C +

Liquiddominated
two phase

Self-flowing
wells

Steam-water mixture
(enthalpy equal to or
higher than that of
saturated liquid
at reservoir
condition); saturated
steam; super heated
steam

Well productivity extremely


variable (up to 50 MWe);
dependent on reservoir flow
capacity and steam
saturation

Single-stage
Flash

High NCG content;


silica scaling in
injection system;
occasionally
corrosion; silica
scaling potential
in production wells
at lower wellhead
pressures

Class 7:
Steam Field

240C
(33.5 bar-a
pressure;
2,800 kJ/kg
enthalpy)

Steam

Self-flowing
wells

Saturated or
superheated steam

Well productivity extremely


variable (up to 50 MWe);
dependent on reservoir flow
capacity

Direct Steam

Occasionally high
NCG content or
corrosion

Geothermal Energy

Figure 7-2 illustrates the calculated distribution of geothermal energy as a function of the
resource temperature worldwide. Based on this distribution, it is possible to evaluate the
potential for low-temperature electricity production and direct heat utilization.

Figure 7-2
Distribution curve of geothermal energy as a function of worldwide temperature [39]

7.2.1.2 Depth
Resource development also depends on the depth to which wells must be drilled to tap
underground reservoirs of hot water or steam as well as the permeability of the rock formation
that constitutes the reservoir and, therefore, the rate of flow from the reservoir rock into the
wells. Depth may range from less than 1,000 feet (300 m) to over 7,000 feet (2,000 m). The
combined effects of permeability and well design/performance can be measured as flow rate per
unit of pressure difference that drives the flow. Some new projects are designed to produce from
depths of 10,000 feet or more.
7.2.1.3 Resource Permeability
Permeability of a hydrothermal reservoir has a significant impact on the economics of
geothermal energy projects. Low resistance to fluid flow (hot water or steam) from the
hydrothermal aquifer to the well bore(s) results in high well production rates and therefore,
lowers unit cost to deliver hot fluid to the geothermal power plant or heat-using facility.
Permeability is a function of the degree of fractures in the rock and the porosity of the rock
matrix. Geothermal resources are associated with volcanic and seismic regions of the Earth.
Therefore, fractures due to seismic (i.e., earthquake) faults are often associated with
hydrothermal reservoirs and can often significantly contribute to their permeability. Often, the
pattern of productive wells in a hydrothermal geothermal field follows a fault linefor example,
at Roosevelt Hot Springs near Milford, Utah, where PacifiCorp has its 38-MWe Blundell
Geothermal Plant. The limit of the production zone of a geothermal field may also be defined by
a rather pronounced line where the change in rock type at a fault ends the zone of a permeable
rock formation containing the water and/or steam.
7-7

Geothermal Energy

7.2.2 Geo-Pressured Geothermal Resource Assessment


Geo-pressured resources are natural deposits of hot brine found in conjunction with oil and gas
deposits. Methane gas dissolved in the hot brine, plus the heat of the brine itself, gives these
deposits value as potential energy and power resources. Use of geo-pressured resources depends
on the economics of natural gas. This is because the low temperature of the hot brine produced
and the great depth, pressure, and hence cost of the producing wells drilled into geo-pressured
zones make electricity generation alone uneconomic. In the United States, geo-pressured
resources are found for the most part in the TexasLouisiana coastal region near the Gulf of
Mexico. These geo-pressured resources are expensive to exploit and the power generation
potential associated with them is only a fractionperhaps 20%of the estimated U.S.
hydrothermal potential [5, 12].
7.2.3 Hot Dry Rock Resource Assessment
Hot dry rock (HDR) resources are still in the experimental stage. The resource is characterized
by man-made reservoirs of hot water created by fracturing geothermally heated crystalline rock
formations at depths of 2,000 to 10,000 meters [36]. Surface water is then pumped into the hot
fractures and most of that water is recovered through production wells, as in a natural
hydrothermal power system. The heated water transfers its heat to a secondary fluid or working
fluid and is then recirculated and pumped down an injection well. There are currently no HDR
systems generating electricity on a commercial scale anywhere in the world, but an enormous
amount of potential geothermal energy is available from HDR. By some estimates, just 1% of the
heat contained in the uppermost 10 km of the Earths crust equals 500 times the energy contained
in all oil and gas resources.
In 1979, the U.S. Geological Survey (USGS) estimated that the known accessible hydrothermal
resource in the United States is about 25,000 MW of capacity for 30 years and the as-yetundiscovered accessible hydrothermal resource is approximately 95,000 MW to 150,000 MW for
30 years [46]. These USGS estimates did not consider deeper geothermal resources below |
3 km at the time, which advanced drilling technology for HDR will potentially achieve.
Figure 7-3 is a map produced by DOEs Office of Energy Efficiency and Renewable Energy
(EERE) indicating geothermal resource temperatures at a depth of 6 km and a 2004 assessment
of heat flow in the United States by Southern Methodist University (SMU).

7-8

Geothermal Energy

Figure 7-3
U.S. geothermal resource potential map (estimated temperatures [C] at 6 km depth) (top),
2004 SMU heat flow assessment (bottom)

7-9

Geothermal Energy

A 2006 study from the Massachusetts Institute of Technology (MIT) estimated that 100,000 MW
of electricity could be installed by 2050 using EGS technology. This projection is based in part
on an abundance of available geothermal resources and readiness of technology elements to
capture the resource. However, a DOE evaluation of the MIT study indicates that significant
advances are needed to achieve the goals of large-scale (100,000 MW) use of cost-competitive
geothermal energy. Needed advances include site characterization, reservoir creation, well field
development and completion, and system operation, as well as improvements in drilling and
power conversion technologies [37]. These technology improvements will also support ongoing
development and expansion of the hydrothermal industry.
Figure 7-4 shows the geographic distribution of estimated Engineered Geothermal Systems
(EGS) resources in the western United States.

Figure 7-4
Estimated EGS resources in the western United States
Source: USGS 2008

Several EGS projects are currently producing power while several others are expected to produce
power in the near term. These include the following:

Soultz (France): 1.5-MW plant (in operation).

Landau (Germany): 2.5-MW plant (in operation).

Paralana (Australia): 7- to 30-MW plant (in drilling stages).

Cooper Basin (Australia): 1-MW showcase plant by Geodynamics is scheduled to be


commissioned in 2013, and a 25- to 500-MW plant (in drilling stages) expected to begin as
the worlds first 25-MW EGS plant.

Desert Peak (Nevada): Stimulation of existing non-commercial well in existing geothermal


field with goal of converting it to a commercial grade well with a target flow rate of 1
gpm/psi.

In October 2008, the DOE selected four new cooperative projects for EGS system
demonstrations in the United States which it hopes will lead to technology readiness by 2015.

7-10

Geothermal Energy

7.2.4 Installed Geothermal Capacity


In 2007, the World Geothermal Congress (WGC2005) estimated that the total geothermal
capacity in the world was 9,929 MW. As of March 2012, total installed capacity in the world and
the United States was 11,224 MW and 3,187 MW respectively according to the Geothermal
Energy Association [42].
Table 7-3 presents data from countries that currently generate geothermal electricity for 2000,
2005, and 2007, as well as incremental capacity installed since 2005. The forecast for 2010
considered the existing projects in advanced stage of development. Figure 7-5 shows the
installed world capacity in 2007 [34].

Figure 7-5
Total installed capacity in 2007
GHC Bulletin, September 2007

7-11

Geothermal Energy
Table 7-3
Installed geothermal power worldwide [28, 34]
Installed
Capacity
In 2000 (MW)

Installed
Capacity
In 2005 (MW)

Installed
Capacity
In 2007 (MW)

Running
Capacity
In 2007 (MW)

2010 Forecast
(MW)

2010 Actual (IGA)

United States

2,228

2,564

2,884

1,935

2,817

3,093

Philippines

1,909

1,930

1,969.7

1,855.6

1,991

1,904

Indonesia

589.5

797

992

991.8

1,192

1,197

Mexico

755

953

953

953

1,178

958

Italy

785

791

810.5

711

910

843

546.9

535

535.2

530.2

535

536

New Zealand

437

435

471.6

373.1

590

628

Iceland

170

202

421.2

420.9

580

575

El Salvador

161

151

204.2

189

204

204

Costa Rica

142.5

163

162.5

162.5

197

166

Kenya

45

129

128.8

128.8

164

167

Nicaragua

70

77

87.4

52.5

143

88

Russia

23

79

79

79

185

82

Papua New Guinea

56

56

56

56

Guatemala

33.4

33

53

49

53

52

Turkey

20.4

20

38

29.5

83

82

China
Portugal

29.2
16

27.8
16

27.8
23

18.9
23

28
35

24

France

COUNTRY

Japan

4.2

14.7

14.7

14.7

35

29
16

Germany

0.2

8.4

8.4

52

Ethiopia

7.3

7.3

7.3

7.3

7.3

Austria

1.1

1.1

0.7

1.4

Thailand

0.3

0.3

0.3

0.3

0.3

.3

Australia
Total

0.2
7,973

0.2
8,933

0.2
9,929

0.1
8,590

0.2
10,993

10,717

7-12

1.1

Geothermal Energy

Note that geothermal plants can typically achieve annual capacity factors of 85% to 90% or
higher [3, 10, 13]. In fact, dry steam resources such as the Geysers typically run at a lower
capacity factor than possible due to different business arrangements (i.e., there is no requirement
that the utility buy all the electricity that can be generated).
Table 7-4 provides data from the Geothermal Energy Association detailing total U.S. installed
capacity as of March 2009. Figure 7-6 plots the trajectory of U.S. geothermal development.
Table 7-4
U.S. installed capacity as of March 2009 [42]
Alaska

California

Hawaii

Idaho

Nevada

Oregon

Utah

Wyoming

TOTAL

0.73MW

2,615MW

43MW

16MW

469MW

0.28MW

42MW

0.25MW

3,187MW

Figure 7-6
Total U.S. installed geothermal capacity as of March 2009 [42]

7.3 Technology Description


As noted, hydrothermal power generation involves well-established systems and technologies
with decades of practical experience. Most of this discussion will therefore focus on
hydrothermal technologies.

7-13

Geothermal Energy

7.3.1 Hydrothermal Power


Hydrothermal power is generated from steam and hot water drawn from geologic structures
relatively close to the Earths surface. A geothermal hydrothermal system consists of a
geothermal reservoir, wells, and a power plant (Figure 7-7). The set of producing wells are often
referred to as the field, or the well field. Sites using direct steam for power generation are located
in the United States, Italy, and Japan. Geothermal hot water or brine is used for generation in the
United States, Philippines, New Zealand, Mexico, Japan, El Salvador, Nicaragua, Kenya,
Turkey, Iceland, and the other countries listed in Table 7-3 [2, 10].

Figure 7-7
Hydrothermal power plant at the Geysers, California

In general, geothermal fluids are tapped through wells, also referred to as bores or bore
holes, and although akin to oil and gas wells which use similar drilling technology geothermal
wells are larger in diameter and typically drilled through hard crystalline rock at much higher
temperatures than the sedimentary environments encountered for oil and gas wells. Well depths
typically range from 200 to 3,500 meters. The fluids surging out of the wells are piped out to the
power plant. One or more wells may be connected to feed steam to the turbine. Geothermal hot
water, steam, or vaporized secondary working fluids then impel a turbine, which in turn transfers
the energy by spinning a generator to create electricity. Waste heat from the process, mainly due
to heat rejection from the steam cycle, is then recovered by a heat exchanger or directly ejected
to the atmosphere through condensers and cooling towers. Remnant geothermal liquids,
including any excess condensate, are generally pumped back into the reservoir through injection
wells. In some systems outside the United States, cooled liquid leaving the plant is disposed of
on the surface or in streams. This cooled liquid is, of course, hot relative to ground surface,
stream, or river temperatures.
If present, non-condensable gases are removed from the system by gas ejection equipment and
released to the atmosphere after treatment mandated by emission regulations. Some emission
control systems may produce sludge or solids that are disposed of either in reclaim fields or in
landfills.
7-14

Geothermal Energy

The nominal plant size used in most system characterizations and industry comparisons is 50
MWe. Real-world systems range in size from 0.075 MWe to 180 MWe [3, 10]. The technology
driver affecting design, performance, and cost of these systems is directly linked to the reservoir
temperature. In general, higher temperatures facilitate higher thermodynamic cycle efficiencies,
and result in a lower cost of the produced electricity. This is because higher-temperature fluid
has more energy per unit mass and because the second law of thermodynamics states that higher
temperatures are converted into useful work more efficiently than lower temperatures. To allow
for these variations, this discussion addresses high-temperature reservoirs (flashed-steam
systems) and moderate-temperature reservoirs (binary systems).
In high-temperature geothermal systems with temperatures greater than 302F (150C), the
reservoir fluid consists of hot water and/or steam. These higher-temperature systems typically
flash the hot water into steam to drive steam turbines and are typically the better candidates for
power plant electricity generation.
Intermediate-temperature systems have resources between 194F and 302F (90C and 150C)
and usually require the use of a secondary fluid for power production. These binary power plants
require a closed-loop heat exchanger technology that allows transfer of the heat from the
geothermal fluid to a secondary fluid that vaporizes at a lower temperature, which in turn drives
the turbine.
The 1996 DOE/EPRI Next Generation Geothermal Power Plant (NGGPP) study [13] provides
substantial performance data and a robust set of cost data for a wide range of reservoir conditions
and power plant technologies, and is still used today. Additional general background information on
geothermal electric technologies and resources is available in several references [9, 15, 16, 27].
7.3.1.1 Major Common System Components and Features

Reservoir: A geothermal hydrothermal reservoir consisting of hot rock with substantial


permeability, and aqueous fluid in situ. The temperature of the fluid ranges from 194F to
752F (90C to 400C). The fluid may contain substantial amounts of dissolved solids and
non-condensable gases (particularly carbon dioxide and hydrogen sulfide).

Exploration/confirmation: An exploration and reservoir confirmation process to identify


and characterize the reservoir. This process is usually risky, complex and can add substantial
front-end cost to a hydrothermal project. For every successful reservoir, there are three failed
exploration/confirmation efforts. Such costs are usually borne out of developers equity and
can be a large barrier to exploration projects. Those costs are accounted for in this discussion
but are not represented in the system schematics. In the long run, the costs of exploration and
confirmation must be included in payments made to developers of geothermal fields or to
producers of geothermal fluids or power.

Wells: Wells for production and injection of geothermal fluids. These range in total depth
from 200 to 3,500 m at existing producing U.S. hydrothermal reservoirs. The wells are
drilled and completed using technology for deep wells that has been incrementally adapted
from oil and gas well technology since the 1960s, and are typically larger diameter than oil
and gas wells. The produced fluids range from totally liquid to liquid-vapor mixtures (with
two-phase flow at the wellhead). In a dry steam field, the wells produce only steam (and
perhaps very small amounts of liquid condensate). Injection wells at some geothermal fields
return spent (heat removed) geothermal fluid to the underground reservoir.
7-15

Geothermal Energy

Reservoir design and management: The goal of a reservoir design and management
process is to optimize the harvesting of heat, and thus the production of electricity from the
reservoir, at least cost over the life of the plant. These costs are accounted for in this
discussion but are not represented in the system schematics.

Surface piping system: Insulated surface piping that transports fluid between the wells and
the power plant equipment.

Power conversion system (power plant): A powerhouse that converts heat (and other
energy) from the geothermal fluid into electricity. Power plants consist of (a) one or more
turbines connected to one or more electric generators; (b) a condenser to convert the vapor
exiting from the turbine (water or other working fluid) to a liquid; (c) a heat rejection
subsystem to transfer waste heat from the condenser to the atmosphere; most systems use
cooling towers (wet or dry), and others use cooling ponds; (d) electrical controls and
conditioning equipment, including a step-up transformer to match transmission line voltage;
and (e) an injection pump that pressurizes the spent geothermal liquid from the power plant
to return it to the geothermal reservoir through the injection wells. Representative powerconversion (power plant) technologies are described below.

Operation and maintenance: Activities and costs related to the operation and maintenance
(O&M) of the system over a typical 30-year useful life of an individual power plant and a 40to 100-year production life of a reservoir as a whole. Relative to fossil-fuel-fired power
generation, simple-cycle hydrothermal power plants are much less complex, operate at
relatively low pressure and temperature, and have fewer auxiliaries. Nonetheless, geothermal
plants incur a significant penalty with respect to the severe equipment operating environment
created by saturated steam containing relatively high levels of impurities of non condensable
gases, dissolved and suspended solids. Geothermal maintenance costs are about twice that of
fossil power plants, due principally to corrosion of well casings, surface pipelines, and so on;
deposition of geothermal mineral phases and corrosion products inside production and
injection wells, steam field piping and turbines; and erosion of surface piping, valves, and
turbine blades.

7.3.1.2 Direct Steam


Dry steam (also called direct steam or vapor-dominated) power plants draw from
underground reservoirs of steam. Conventional steam turbines are used with hydrothermal fluids
that are wholly or primarily steam. The plant gathers steam from wells and routes it directly to a
turbine, which drives an electric generator. It rejects waste heat to the atmosphere through a
condenser and cooling tower, and returns remaining geothermal liquids to an underground
reservoir via injection wells. Dry steam resources are rare. The Geysers geothermal field in
northern California (Figure 7-7) is an important example of such a reservoir and accounts for
approximately half of existing U.S. geothermal capacity. However, geothermal developers do not
expect to find any more such fields in the United States. The only other major dry steam
geothermal field in the world is the Larderello field in Italy, site of the worlds first geothermal
power plants, originally installed in about 1904.

7-16

Geothermal Energy

In some situations, vapor-dominated reservoirs can produce two-phase water/steam mixtures.


The vapor-dominated term applies to the governing pressure and flow driver in the reservoir.
Sometimes the transition from nearly all-liquid production to increasingly higher fractions of
steam entering along with the liquid water (or brine) can indicate that the reservoir is reaching
an advanced stage of depletion. The higher steam fraction may increase the power generating
capacity for a brief time, but the well is near depletion and a replacement well will soon be
needed. EPRI supported development of a method of assessing changes in reservoir conditions
and predicting changes in the steam/water ratio [26]. The technique is based on the higher
mobility of radon gas in the steam phase than in the liquid phase.
7.3.1.3 Flash Steam
Facilities that use hydrothermal fluids above 360F (180C), primarily water, usually use flash
steam technology. This method uses very hot water that flows through wells under pressure.
Flash steam systems include one or two (for single-flash or double-flash, respectively) large
vessels called flash tanks, in which part of the geothermal fluid vaporizes into steam at pressures
below that of the reservoir. These vessels can be located at the power plant or at the individual
wells (at the wellhead) or at both, depending on whether single-phase or two-phase
transmission pipelines move the fluids from wells to power plant. The steam flow is typically
18% to 25% of the mass of the fluid from the reservoir (for double-flash plants) and is sent to the
high-pressure and low-pressure inlets of a conventional turbine-generator. Steam volume
depends on conditions in the reservoir and the designs of the production wells and power plant.
The remaining liquid (brine) from the second flash tank, containing 75% to 82% of the original
mass, is injected through a well back into the geothermal reservoir.
Three-stage flash power plants that were built and operated in New Zealand in the 1970s were shut
down in the 1980s due to resource depletion. Three stages of flash enable heat to be converted to
electricity with higher efficiency than the double-flash approach, which is, in turn, more efficient
than the single-flash approach. The added costs of each additional stage of flash, dealing with
lower-pressure/higher-volume steam flows, must be traded off against efficiency and
environmental considerations. One figure of merit that improves as a design goes to more stages of
flash and lower temperatures of spent fluid is the hot-water rate or brine rate, measured in
mass (tons) of geothermal fluid extracted per unit of electricity generated (MWh). Lower rates
mean that less geothermal fluid is required to generate a unit of electricity. In general, the less fluid
produced, the lower the total cost of the production system and the less its environmental impact.
Equipment present in all or most flashed-steam systems includes the following:

One or two large flash tanks in which part of the geothermal fluid vaporizes (flashes) into
steam at pressures less than the pressure in the reservoir. The turbine in the dual-flash system
shown in Figure 7-8 has dual inlets to admit high-pressure steam from the first flash tank and
low-pressure steam from the second flash tank. The flash tanks can be located at the
wellhead, at the power plant, or at both locationsfor instance, the first stage at the
wellhead(s) and second stage at the power plant. At Cerro Prieto in Mexico, just south of the
U.S.Mexico border near Mexicali, the original single-flash system uses flash separators at
the wellheads. A subsequent modification generates power from a second-stage flash and
features a central low-pressure flash steam facility and adjacent power plant. The liquid from
the first separation (wellhead separators) is flashed to generate the low-pressure steam that
drives the added turbine/generator.
7-17

Geothermal Energy

Figure 7-8
Dual-flash (two-flash) geothermal power plant

Special features related to minimizing the deposition of silicate scale. The plant depicted in
the system diagram is similar to some, but not all, U.S. flash plants, and geothermal brine
often contains substantial amounts of dissolved silica, which tends to precipitate on
equipment walls as hard scale if not treated. Mitigation measures include the following:

Elevating the conversion cycles brine exit temperature above that optimal for maximum
power production, which tends to keep some of the silica in solution (this is the method
of choice when silica presents a small to moderate problem.)

Using a crystallizer-clarifier system, which consists of a brine solids clarifier and a


return line and injects silica seeds into the first flash tank to prevent the precipitation of
amorphous silica on the walls of the vessel and connecting pipes

Using a pH modification system, which injects small quantities of acid upstream of the
first flash tank to reduce the pH of the geothermal fluid, thus reducing the rate at which
silica precipitates out of solution allowing the fluid to be re-injected before scaling
occurs.

Environmental control measures. At reservoirs where the concentration of non-condensable


gases (e.g., CO2) is high, substantial gas ejection equipment is attached to the condenser.
Steam or electricity drives the ejectors. If hydrogen sulfide in the gases requires abatement,
H2S control equipment is added downstream of the ejectors. H2S removal is a significant
environmental control system when such removal is necessary to prevent odor at plant sites.
Other environmental control measures can include settling and cooling ponds, water and
solid waste containment and disposal system, and other measures if appropriate or required
for local situations.

7-18

Geothermal Energy

Descriptions of flashed steam geothermal power plants are presented in the Renewable Energy
Technology Characterizations (RETC) [10]. The NGGPP report [13] provides a range of process
and cost information. A 2012 update to the NGGPP report provides updated information on
flashed steam technologies [44].
7.3.1.4 Binary Power Plants
For hydrothermal fluids with temperatures below 360F (180C), binary cycle technology is
generally most cost-effective. Binary cycles are also applied together with a flash cycle in
hybrid or bottoming cycle configurations, where the binary cycle generates power from the
lower-temperature liquid that remains after one or two flash steps upstream have exploited the
higher-temperature heat. In this application, the binary cycle is an alternative to the second or
third stage of flash. Thus, the binary cycle is another way to use more of the heat and increase
the megawatt-hours of electricity generated per ton of geothermal fluid brought to the power
plant from the well field.
In a binary cycle power system, the natural geothermal fluid flows in one of two loops that make up
the binary system. Figure 7-9 is a schematic diagram of a binary cycle geothermal power plant.

Figure 7-9
Binary cycle geothermal power plant (air-cooled design)

Hydrothermal fluid passes through a heat exchanger where it vaporizes a secondary fluid, usually
an organic compound with a low boiling point such as isopentane, which drives a specially
designed turbine. Having given up its heat, the water is returned to the geothermal reservoir
using a deep injection well. Low-pressure vapor from the turbine is liquefied in the condenser
and re-pressurized by the hydrocarbon pump. Waste heat is ejected to the atmosphere through a
condenser and cooling tower. The water and working fluid are confined to separate closed loops
during the whole process, so there are few or no air emissions.
7-19

Geothermal Energy

Makeup water is required for the heat rejection system if wet cooling towers are used, but not if
dry cooling towers are used. The binary system characterized here is one with dry cooling (aircooled). However, wet cooling will usually be less expensive than dry cooling, if cooling water
is available. Due to the lower operating temperature of a binary geothermal power plant, more
water will be needed for the heat rejection system than for an equivalent sized fossil fuel plant.
Technical descriptions of binary-cycle or organic Rankine cycle (ORC) power systems and other
systems proposed for moderate-temperature reservoirs can be found in the NGGPP report [13] as
well as in the 2010 EPRI report Engineering and Economic Evaluation of Low-Temperature
Geothermal Power Plants [49]. The RETC report [10] references others such as binary systems,
vacuum-flash, and ammonia-based cycles. A 2012 update to the NGGPP report provides updated
information on binary technologies [44].
Equipment present in most binary systems includes the following:

Down-hole production pumps in the production wells. These keep geothermal fluid from
vaporizing in the wells or in the power plant, and enhance the production well flow rate.

A working fluid pump, the main cycle pump, that pressurizes the low-boiling-temperature
liquid working fluid to drive it around the power-conversion loop.

A primary heat exchanger, where heat is extracted from the geothermal fluid and transferred
to vaporize the binary working fluid (isobutane in the case shown in Figure 7-9).

A turbine that converts energy in the high-temperature high-pressure working fluid vapor to
shaft energy, driving a generator.

A condenser and cooling system, receiving the low-temperature low-pressure isobutane


vapor exhausted from the turbine and condensing it to a liquid for re-pressurization by the
working fluid pump. The heat extracted from the working fluid in the condenser is rejected to
the atmosphere via a wet or dry cooling system (the cooling tower).

Most geothermal binary plants consist of a number of smaller modules, each having a capacity of
1 to 12 MWe net. EPRI has extensively documented the largest binary power system ever built,
the nominal 48-MWe (net) Heber Binary Project, built by DOE and San Diego Gas and Electric
Company as a demonstration project in the 1980s. [17].
The trend in binary-cycle technology since the mid-1980s has been toward binary cycles of
relatively small size, despite the lower cost of electricity that would be expected to result from
economies of scale at larger sizes. The reasons for this are that binary cycles lend themselves
well to smaller sizes and to modular construction techniques, and the economics of financing
favors smaller sizes. For instance, in New Zealand, the large unit size that launched Wairakei at
over 100 MWe more than 40 years ago is no longer seen as the preferred path, because field
development in smaller increments of 20 MWe or less enables a well developer to establish a
revenue stream sooner, without the cost and risk of proving a 50- or 100-MWe resource in the
ground. The 100-MWe unit size was developed in New Zealand when there was a national
government-owned power generation and distribution corporation. Now, with smaller, private,
competing power generation companies in charge of generating electricity, smaller increments of
capital and less time from first exploration to power sales take priority. The new priorities have
led to the choice of binary cycles at 20- to 24-MWe or smaller unit size as the preferred
technology during the past decade.
7-20

Geothermal Energy

7.3.1.5 Low Enthalpy/Reverse Air Conditioning Cycles


For low-enthalpy geothermal resources, a cycle based on mass-produced air conditioning
components (Carrier Chiller) can be utilized. The cycle uses a single-stage centrifugal
compressor which runs in reverse as a radial inflow turbine, a heat exchanger originally designed
for large chiller applications, and low-cost R-134a fluid. A plant based upon low-enthalpy cycles
has been in operation in Alaska since 2006 and has demonstrated the ability to produce
electricity with high availability (98%) from a 75C (165F) geothermal resource and a 5C
(40F) river water heat sink at competitive cost. Furthermore, modules can be added to expand
the power of the geothermal plant's low-temperature resources.
In general, this low-cost technology can expand the minimum temperature range from 105C
(225F) to around 80C (175F) for producing power from lower-temperature, shallow
geothermal hot springs systems. As a partial result, technologies that harness reverse air
conditioning cycles may promote future short-term small-scale geothermal distributed cogeneration.
7.3.2 Geo-Pressured
Geo-pressured resources are natural deposits of hot brine found in conjunction with oil and gas
deposits. Methane gas dissolved in the hot brine, plus the thermal energy of the brine itself, give
these deposits a role as potential energy and power resources. As mentioned above, the extent of
this resource amounts to only about 20% of the hydrothermal resource and appears to be
expensive to exploit, except for a niche involving wells drilled for other purposes and used to
provide geothermal heat and dissolved methane as an alternative to being plugged and
abandoned [12]. Geo-pressured resources are found for the most part in the Texas-Louisiana
coastal region near the Gulf of Mexico [5, 12]. Use of geo-pressured resources depends on the
economics of natural gas. This is because the low temperature of the hot brine produced and the
great depth and pressure (and, hence, expense) of the producing wells drilled into geo-pressured
zones make the production of electricity alone uneconomic.
7.3.3 Hot Dry Rock (Enhanced Geothermal Systems)
Hot dry rock (HDR) resources are relatively deep masses of rock that contain little or no steam
or water and are not very permeable. They exist where geothermal gradientsthe vertical profile
of changing temperatureare well above average (>50C/km). Rock temperature reaches
commercial usefulness at depths of about 3 km or more. Traditional hydrothermal systems rarely
require drilling deeper than 3 km, while the technical limit for current drilling technology is to
depths greater than 10 km [36]. The potential energy to be derived from HDR is enormous. The
U.S. Geological Survey estimates that 3 to 18 million EJ of accessible HDR energy exists in the
United Statesenough to satisfy current U.S. energy consumption for tens of thousands of years
[4]. Figure 7-10 illustrates HDR geothermal production process.

7-21

Geothermal Energy

Figure 7-10
HDR geothermal production process

To exploit HDR resources, site developers first create a permeable reservoir by hydraulic shear
fracturing in a construction phase [3]. Then, in the operating phase, water from the
surface must be pumped through the fractures to extract heat from the rock [10]. At sufficient
temperature, the water can be flashed to steam and used to generate power. Alternatively, the
injected water can be held at sufficient pressure to maintain it in liquid form for use in a binary
or a flash/binary hybrid cycle. The technical challenges lie in the sub-surface elements of the
system, not the power cycle, notably related to deep drilling and in situ fracturing [9, 15].
Flash steam or binary cycle technologies could both be used with HDR resources depending on
the temperature of geothermal fluid produced. However, because of constraints imposed by high
well costs, most of the accessible HDR resource would likely produce wellhead fluids in the
moderate temperature range best suited for binary cycle technology [9].
To date, HDR resources have not been developed commercially in the United States. Well costs
increase exponentially with depth; because HDR resources are much deeper than hydrothermal
resources, they are much more expensive to develop. Also, although the technical feasibility of
creating HDR reservoirs has been demonstrated at experimental sites in the United States,
Europe, and Japan, operational uncertainties regarding the resistance of the reservoir to flow,
thermal drawdown over time, and water loss have so far made commercial development risky.
Lower-cost resource assessment and drilling technologies will help bring these HDR systems
into commercial use [3, 4, 9]. Significant progress has been achieved in recent tests at Soultz,
France under European Union (EU) sponsorship, and in Australia under largely private
sponsorship. However, HDR systems lack a demonstration of their capability at the present time.
DOE has aptly named its effort to investigate the potential for increasing the amount of
electricity generated from geothermal sources as enhanced geothermal systems (EGS). As the
name suggests, EGS technology could be used to enhance and extend hydrothermal geothermal
production zones. The same technology that fractures HDR formations and allows them to heat
water that can flow to a geothermal power plant on the surface could also be applied to extend

7-22

Geothermal Energy

the life, flow rate, and size of the production zone of a hydrothermal geothermal field. Such
technology could enhance or restore the permeability of rock formations and bring in new
sources of water, such as the treated wastewater from the fringes of the Geysers field in northern
California. It could also help maintain production at sites where natural geothermal steam and
water are depleted or insufficient natural permeability is encountered.
7.3.3.1 Supercritical Cycles
Supercritical fluids reside at a temperature and pressure that can allow the fluid to diffuse
through solids. A supercritical fluid such as supercritical carbon dioxide (SCCO2) can be pumped
into an underground formation to fracture the rock, thus creating a reservoir for geothermal
energy production and heat transport. The supercritical fluid used to form the reservoir can also
be used as a geofluid to heat-up, expand, and be pumped out of the reservoir to transfer heat to a
surface power plant or other application.
Utilizing SCCO2 for geothermal power production in an EGS system has promise. Although the
heat capacity of SCCO2 is 40% that of water, it has a much lower viscosity and large density
differential between hot and coldthat is, extraction and reinjection points that make this
technology worth further exploration. The large density differential has the ability to form a
thermal siphon in which the heavier cold SCCO2 at the injection point tends to sink while the
lighter hot SCCO2 at the extraction point tends to rise. This combined with the low viscosity of
SCCO2 could greatly reducing pumping needs, a critical consideration given that EGS systems
are envisioned residing several to tens of kilometers beneath the surface. This advantage of
SCCO2 would have an offsetting effect on its relatively low heat capacity compared to water
vapor. Another consideration is that subsurface minerals are insoluble in SCCO2, thus reducing
mineral scaling issues on well lining and surface equipment [40].
Supercritical cycle technology could use SCCO2 as the geofluid (for geothermal heat exchange in
the reservoir) and remove the need for a binary plant heat exchanger by directly expanding the
SCCO2 through a specifically designed turbine, increasing efficiency while lowering capital
investment. However, the produced SCCO2 stream needs to be dried before entering the turbines
to avoid condensation of liquid water during decompression and cooling. Brown [40] has
postulated that the SCCO2 would eventually dissolve all in situ water in the EGS reservoir,
resulting in pure SCCO2 at the extraction point.
Access to large quantities of CO2 is essential for supercritical plants, first to charge the reservoir
and then to make-up for underground capture. The cost of such access could be defrayed by the
CO2 sequestration benefits. Overall, the concept of SCCO2 as a heat transmission fluid for EGS
looks promising for future applications and deserves more study to develop the scientific basis
for a field demonstration.
7.3.4 Hot Sedimentary Aquifer (HSA)
Hot sedimentary aquifers are reservoirs in which rainwater that has been absorbed into the
ground is heated at temperatures that increase with depth or by contact with hot rocks. The water
collects in porous rocks between two impermeable sedimentary layers, creating an aquifer from
which hot fluid can be extracted, usually by drilling. HSA typically requires a binary cycle for
electricity production due to the temperature of the brine. The focus of HSA research is to find
shallow systems that reduce development costs and allow the use of proven hydrothermal
7-23

Geothermal Energy

systems and supporting technology. Secondary reservoir stimulation techniques, known as


secondary enhancement of sedimentary aquifer play (SESAP), have also been researched as a
way to increase permeability and production rates of HSA.
Sedimentary basins in Australia (Otway, Murray) and in the United States (Northern Gulf of
Mexico areas of Texas and Louisiana) are areas with high potential HSA resource. Figure 7-11
illustrates the HSA production process.

Figure 7-11
Hot sedimentary aquifer production process

7.3.5 Down-Hole Closed-Loop Systems


Down-hole closed-loop systems place a heat exchanger deep in the earth within the geothermal
resource for extracting geothermal energy. The working fluid cycles through the heat exchanger
in the ground and then returns to the surface for use in the power block. Because this system is a
closed loop, it removes the risk of contaminating the aquifer and uses limited water and space,
produces zero emissions and has low visual impact. Other advantages include no risk of induced
seismic instability and chronic pipe corrosion. The system is targeted for use in new drilled or
existing depleted or abandoned oil and gas wells, which are numerous in the United States, and
may be near existing transmission and distribution (T&D) infrastructure sites. Recent EPRI
studies [45, 46] have shown that a single-well closed-loop enhanced geothermal system
(SWCLEGS) shows some promise in limited applications.
7.3.6 Geothermal Hybrid Plants
From an energy efficiency viewpoint, energy production requires the use of appropriate heat
sources optimized for the temperature/pressure requirements of the turbines. The hybrid
approach, in which more than one heating source is employed for a given application, offers
perhaps the best way to exploit geothermal technologies.

7-24

Geothermal Energy

7.3.6.1 Geothermal-Solar Hybrid Cycles


Solar-geothermal hybrid plants combine a concentrated solar thermal field with a geothermal
plant. The arrangement can include either binary or flash steam plants. In either design, heat is
collected in the solar field and transferred to a heat transfer fluid (HTF). In a flash steam plant,
the HTF passes through a heat exchanger with geothermally heated water coming from a
production well. The HTF further heats the geothermal water before it enters the separator,
where it is flashed and then expanded through the turbine. In a binary system, the working fluid
is first heated by geothermal water before it is further heated in a HTF heat exchanger and
expanded through the turbine.
Solar-geothermal hybrid plants (Figures 7-12 and 7-13) offer improved performance over pure
geothermal systems owing to the additional heat generated from a solar field. Additionally, they
can offset the risk of premature resource depletion, provide operating flexibility, and take
advantage of high peak summer electricity prices by generating additional electricity during hot
summer days when the solar resource is strongest.
EPRI considered two scenarios in which a plant with reduced output due to a partially depleted
geothermal resource is augmented with solar thermal heat. The economics of a solar augmented
flash plant look better than the economics of a lower temperature solar augmented binary plant
primarily due to the higher underlying cycle efficiency of the flash plant. In both cases drilling a
new production well is the most economic method of increasing lost capacity should the resource
exist [47, 48].

Figure 7-12
Geothermal-solar hybrid production process for a geothermal flash steam plant

7-25

Geothermal Energy

Figure 7-13
Geothermal-solar hybrid production process for a hot water/brine non-flash resource

7.3.6.2 Geothermal + Gas Peaker Cycles


Geothermal plants paired with gas turbine peaker plants offer potential value opportunities. The
reason is that gas peakers can be dispatched by the utility when market prices are high while the
geothermal plants are operated at base load. Moreover, when the gas turbines are on-line, power
produced by the geothermal plants can be enhanced by the heat exhaust from the gas turbines,
thus increasing efficiency. Figure 7-14 is a schematic of a geothermal/gas peaker arrangement.

Figure 7-14
Geothermal-gas peaker production process

7-26

Geothermal Energy

7.3.6.3 Geothermal-Fossil Hybrid Cycles


Hybrid steam power plants with geothermal feed-water heating enable the utilization of
geothermal energy in areas with moderate enthalpy resources. This scheme may be applicable to
hydrothermal and EGS-HDR geothermal technologies. The process of geothermal feed-water
preheating as a means of improving performance through an improved plant heat rate forms both
an alternative and an extension to the existing electricity generation methods based on renewable
energy.
7.3.6.4 Geothermal-Biomass Hybrid Cycles for Moderate-Low Enthalpy Sources
Figure 7-15 shows a possible geo-biomass hybrid cycle with energy transfer accomplished by
means of three heat exchangers: a pre-heater, a steam generator, and a super-heater. The energy
source for the first two heat exchangers is geothermal; the energy source for the third one may be
geothermal, or could alternatively be biomass, natural gas, or other available fuels.
In addition to steam cycles, organic cycleswhich have operating points that may be more
adequate for geothermal productioncan be used as working fluids. The organic working fluid
may generate more power than water due to increased heat transfer in the steam generator, but it
may also have a slightly lower efficiency and probably decrease the geothermal percentage of
power generation. The application of the optimal working fluid needs to be carefully planned and
adjusted for individual sites.

Figure 7-15
Geo-biomass hybrid cycle

7.3.7 Non-Electric (Direct Use) Geothermal


In addition to the electric power generation technologies addressed in this report, hot geothermal
fluids can be used directly in aquaculture, greenhouses, space heating, wood drying, and
vegetable drying. In the United States, non-electric energy uses are very small relative to electric
power applications. District-heating projects exist in Idaho, Oregon, and other states, and there
are a few new projects in the proposal stage. If the price of energy, especially natural gas,
continues to rise, similar projects will evolve.
7-27

Geothermal Energy

Geothermal heat pumps (GHPs) are one of the fastest-growing applications of renewable energy
in the world. This form of direct use of geothermal energy is based on the relatively constant
ground or groundwater temperature in the range of 4C to 30C (39F to 86F). GHPs provide
space heating, cooling, and domestic hot water for homes, schools, factories, and public and
commercial buildings.
Until recently, almost all ground source heat pump installations occurred in Europe and North
America. However, installations expanded to more than 40 countries in 2007, up from 26
countries in 2000. China is the most significant newcomer in the application of heat pumps to
space heating.
The potential of GHP is very large, as space heating and water heating are significant parts of the
energy budget in large parts of the world. In industrialized countries, 35% to 40% of the total
primary energy consumption is used in buildings. Future development predictions show
exponential worldwide growth in the heat pump sector: up to 200,000 MWth for 2020 to about
750,000 MWth by 2050. In contrast, estimated capacity worldwide in 2008 for geothermal heat
pumps was about 30,000 MWth.

7.4 Technology Status


While geothermal power production (hydrothermal, not HDR) is generally considered a
commercial technology, there are many opportunities for improvement through research and
development. Geothermal projects are very capital intensive. Continued research and
development will be useful on many fronts, particularly in developing improved geophysical
methods to detect geothermal reservoirs and resources, decreasing the cost of drilling geothermal
wells, decreasing the cost of power production by reducing maintenance cost, increasing the
conversion effectiveness of power plants sited at low-temperature reservoirs, and decreasing
O&M costs of wells, field equipment, and power plant equipment performance.
Technologies for exploration, drilling, and reservoir analysis and management are essentially the
same for all types of geothermal systems. There is substantial room for improvement in most
aspects of these technologies, including both the fluid-production (exploration, wells, and
reservoir management) and electricity-conversion (power plant) components. Improvement in
drill bit technology is expected to help reduce the cost of drilling deep geothermal wells by about
20% over the next decade. The cost of conversion technologies should decrease substantially
over the next decade for lower-temperature systems such as binary-cycle plants, but may not
decrease much for higher-temperature flash systems. The main thrusts for reducing power plant
costs involve the following [10, 13]:

Substantial changes in the basic conversion cycle designs used in the plants, including the
addition of topping and bottoming cycles, improved working fluids, and the used of
various hybrid cycles. An example of a topping cycle is the rotary separator turbine (RST)
developed by Douglas Energy and previously by Biphase Energy Systems, with major testing
support from EPRI during 19781984 and from Utah Power & Light in 19811987.
Subsequent testing has been done at Cerro Prieto, the geothermal field in Mexico just south
of the California border, with support from the California Energy Commission and DOE. The
RST extracts added electric power at the high-temperature side of a flash process by having
the flash and separation take place in a rotary separator designed to capture the kinetic energy
that the flashing steam imparts to the liquid phase. Examples of bottoming cycles include the

7-28

Geothermal Energy

third stage of flash in use in New Zealand, a binary cycle added at the low-temperature side
of a flash cycle, and a special low-temperature water-ammonia binary cycle (the Kalina
Cycle79 of Exergy Inc.) added below a flash or a binary cycle.

Efforts on the part of system operators to reduce O&M costs, especially by reducing the
number of staff employed at each system and site. This is being accomplished through
increased automation of controls and monitoring. An example is the operation and
maintenance of the Wairakei and Ohaaki-Broadlands fields by Contact Energy in New
Zealand. The central control system at the Wairakei plant also controls the Ohaaki plant 15
km away, and a roving operator performs monitoring and maintenance tasks for both plants.

Gradual reduction in complex instrumentation and controls as operations are streamlined and
engineers learn what components are safe to omit.

The consensus in the industry is that characterizing the commercial potential of identified
geothermal reservoirs is a high priority [3]. Techniques such as fracture mapping, more accurate
thermal gradient wells, and other untested methods should be evaluated and refined if
appropriate. In the near term, such work will focus on the following:

Techniques for improving the cost-effective definition of resources, as discussed previously

Reducing the costs of drilling production wells

Improving the efficiency of geothermal facilities

Understanding the results of wastewater injection and improving this process

Investigating environmental impacts of geothermal energy developments

7.4.1 Recent U.S. Developments


Recent interest in geothermal energy in particular and renewable energy in general has
precipitated expansion of the geothermal electricity market, with a number of states seeing
developers working on projects. Altogether, there are a total of 147 geothermal projects in the
United States in various stages of development [42].
Geothermal electric power plants are located in eight states: Alaska, California, Hawaii, Idaho,
Nevada, New Mexico, Utah, and Wyoming. The total installed capacity was 3,187 MW as of
December 2011, with 15 states with geothermal projects currently under consideration or
development. A Geothermal Energy Association survey identified up to 4,320MW of new
geothermal power plant capacity under development in the United States, which includes
projects in the initial development phase [42].
7.4.1.1 Alaska
The first geothermal power plant in Alaska was installed in 2006 at Chena Hot Springs. The
plant pairs the direct use of geothermal thermal energy with on-site geothermal electricity
generation, and uses a binary cycle plant to produce 200 kW gross from the coldest geothermal
resource worldwide, just 74C. Additional units were added increasing the total capacity to
730kW. Alaska has 85 MW of additional geothermal power in early project development.
79

Kalina Cycle is a trademark of Exergy Inc.

7-29

Geothermal Energy

7.4.1.2 Arizona
Arizona has one small project in development. Greenfire Energy is developing a DOE funded
2MW EGS plant utilizing SCCO2 along with metal organic heat carriers (MOHC) as the working
fluid in the St. Johns near Springerville in northern Arizona. The MOHC are intended to improve
the heat capacity of the SCCO2 thus improving its performance relative to steam. Greenfire
presented its results to date at the DOE Geothermal Technologies Program peer review in May
2012.
7.4.1.3 California
In Californiasite of the vast majority of U.S. geothermal power productiongeothermal
development continues at a high level of activity. Currently 31 projects from 11 different
companies are being developed representing an additional 1,766 MW of geothermal capacity.
Geothermal power plants operating in California include the following:

The Geysers: 22 dry steam units, totaling 1,531 MW of installed capacity (but only 932 MW
operating due to a steam shortfall)

Imperial Valley-East Mesa: 70 units totaling 100 MW of installed capacity

Imperial Valley-Heber: 14 units totaling 94 MW of installed capacity

Imperial Valley-Heber South: one unit totaling 14 MW of installed capacity

Imperial Valley-Salton Sea: 10 units totaling 327 MW of installed capacity

COSO: nine units totaling 270 MW

North Brawley one unit of 50 MW

Others: 29 units

The total installed geothermal capacity in the state as of December 2011 was 2,615.4 MW.
Geothermal power supplied about 42% of Californias in state renewable electric generation in
2010. With an aggressive renewable portfolio standard, California expects the state to at least
double its geothermal capacity in the next decade.
7.4.1.4 Colorado
Colorado is experiencing some geothermal activity with two early phase projects being
developed. One is at Mt. Princeton and the other is at Poncha Hot Springsboth of these
locations are in the mountains of central Colorado. Current project estimates are 20 to 25 MW
combined.
7.4.1.5 Hawaii
Currently, one geothermal power plant operates on Hawaii. The existing 10 flash + binary units
provide 35 MW installed capacity (30 MW running, after rehabilitation and work over). This
power plant supplies approximately 20% of the total electricity need of the Big Island and its
160,000 inhabitants. Recently an 8-MW expansion was completed, with permitting in place to
add an additional 22 MW thereafter.
7-30

Geothermal Energy

Hawaiian Electric Company (HELCO) is actively seeking new geothermal resources, and Ormat
currently has three projects in developmenttwo on the Big Island and one in Maui.
7.4.1.6 Idaho
In early 2008, Idahos first geothermal power plant came on line. The Raft River binary plant has
a nameplate rating of 15.8 MW, although currently its net electrical power output is between
10.5 and 11.5 MW. The plants owner, U.S. Geothermal, is in the process of expanding the
project to 45 MW. There are several other projects in the early development phases that could
potentially add 250 to 325 MW of electricity in Idaho.
7.4.1.7 Louisiana
A small 50-kW geothermal project that plans on using hot geothermal brine co-produced with oil
and gas production is currently being developed by Gulf Coast Green Energy.
7.4.1.8 Mississippi
A small DOE funded 50kW co-produced demonstration project was completed in 2011 at a
Denbury Resources producing oil well in Laurel, MS. The project was developed by Gulf Coast
Green Energy. The six month demonstration used technology from ElectraTherm (the Green
Machine). The project concluded in November 2011 with 1,136 total runtime hours, and
provided insight for future installations.
7.4.1.9 Nevada
Currently, 22 geothermal plants operate in Nevada, with a collective nameplate capacity of 469
MW. Between confirmed and unconfirmed projects, a total of 59 projects and 17 geothermal
prospects are under development in Nevada, which could eventually supply between 1,915 and
2,125 MW of electricity.
7.4.1.10 New Mexico
Raser Technologies 240-kW pilot installation project at Lightning Dock went on line in July
2008, making it New Mexicos first geothermal power plant. The pilot has since been shut down,
and Utah-based Cyrq Energy is redeveloping the site for a total planned capacity of 15 MW.
Gradient Resources has a project in early development in southern New Mexico at a resource
estimated to be as high as 100 MW. A direct-use aquaculture facility is also operating in New
Mexico, with the potential for a 1-MW plant to be built at that site in the future.
7.4.1.11 North Dakota
The University of North Dakota is working with several entities to put in a 250-kW coproduction demonstration plant at an existing oil field.
7.4.1.12 Oregon
The Oregon Institute of Technology continues to host the only geothermal plant in the state. The
280-kW plant is located on its main campus and serves as a learning platform for students.
7-31

Geothermal Energy

Another 15 projects and two geothermal prospects are currently in development with a total
capacity of up to 300 MW.
7.4.1.13 Texas
Texas currently has 800 kW of co-produced projects in some phase of development in the
southeast part of the state.
7.4.1.14 Utah
Utah had three geothermal power plants in operation. The Blundell Plant has two units for a total
of 32 MW. Utahs second geothermal power plant, Hatch, has a gross capacity of 10 MW.
Eleven projects and eight geothermal prospects are currently in development with a total
capacity of up to 195 MW.
7.4.1.15 Washington
Gradient Resources is currently developing one geothermal prospect in the state near Mt. Baker.
The estimated capacity available from this resource is up to 100 MW.
7.4.1.16 Wyoming
The first geothermal power plant in Wyoming came on line in September 2008. The coproduction demonstration project consists of a 250-kW organic Rankine cycle power unit.
7.4.2 Hot Dry Rock Technology Status
Hot dry rock (HDR) field investigations began in the early 1970s at Fenton Hill, New Mexico,
near the Los Alamos National Laboratory. This project was the first attempt in the world to make
a deep, full-scale engineered reservoir. Experiments from 1973 to 1996 aimed to develop
methods to economically extract energy from HDR systems located in basement rock at suitably
high temperatures. Support for the project declined during the later phases of work and by 2000,
all experiments were terminated and the site was decommissioned.
Building on the experience of the Fenton Hill project, experiments were performed at two sites
in Japan during the 1980s and 1990s. These experiments included work similar to that of the
DOE EGS program, namely enhancing a poor or declining conventional hydrothermal hot-waterproducing field through targeted rock fracturing in zones where more permeability was needed.
The most significant experiment in the world since 1994 has been the European project at Soultz,
France, where two fracture zones have been completed, one deeper and hotter than the other. The
deep zone could be the source of sufficiently high-temperature hot water to drive an electric
power generation station. Circulation tests have been performed, from injection well to hot rock
fractured by high-pressure injection to the production well and back to the surface. The wells
were stimulated by adding salt and nearby lake water to provide initial brine. During the 2001 to
2004 period, a pilot plant was built using the previously drilled and stimulated three-well system.
During the second phase, which was to last from 2005 to 2008, an initial 1.5-MW power plant
was constructed for operation using the medium circulation loop of 26 gal/s [100 L/s]. The plant
went on-line in June 2008.
7-32

Geothermal Energy

In mid-2003, Geodynamics Ltd. of Australia began drilling to build the worlds first commercial
HDR plant. The Habenero 1 well extends just under 3 miles (4.4 km) under South Australias
Cooper Basin to tap a promising geothermal region in which rocks achieve temperatures of about
550F (290C). HDR exploration in South Australia alone covers an area of 991 km2 and has
potential for future capacities in the thousands of megawatts. The nearly 1000 km2 HDR resource
is equivalent to 50 billion barrels of oil or 10.3 billion tonnes of coal20 times larger than all
the known Australian oil reserves and equivalent to 40 years of the nations current black coal
production.
Geodynamics has secured the rights to another HDR field in the Cooper Basin and two others
in the Hunter Valley in New South Wales. After developing an initial demonstration HDR plant
in the 1 MWe range, the company plans to build a commercial-scale power plant capable of
generating hundreds of megawatts.
Habanero 1 and 2, Geodynamics first two geothermal wells, have produced good results. The
company has successfully completed the riskiest task of the project, hydraulic stimulation of the
underground wells and heat exchanger. In 2005, the firm reported that using a larger than
expected heat exchanger had exceeded expectations. An initial water circulation test in spring
2005 lasted 40 hours and reached temperatures of 389F (198.5C). Habanero 2 produced about
10 MW of thermal power during diagnostic flow tests. The flow test in April 2005 found some
potential for power production, but concluded that a dropped plug was restricting the connection
to the geothermal reservoir. The companys attempt to drill another sidetrack encountered
problems and was concluded in June 2006 when the drill pipe became stuck. Multiple other wells
have since been drilled including Habenero 3 (2008, 4.2km, 242C), Jolokia 1 (2008, 4.9 km,
263C), Savina 1 (2009, 3.9km, suspended) and the recently completed Habenero 4 (2012,
4.2km). Habenero 4 will be paired with Habenero 1 (as an injector well) to demonstrate a 1
MWe demonstration power plant expected to come on-line in 2013.
HDR development to date indicates that there is a very good potential for economic energy
extraction in the future. This potential has been considerably enhanced by the discovery of
overpressures in the granite fracture network, which could add a large convective heat
component into the original design. Overall, EGS technology has the potential to provide a
sizeable addition to current geothermal production.
In 2008 Davenport Newberry drilled two exploration wells to over 3 km on the shoulder of
Newberry Volcano near Bend, OR encountering 316C rock with no permeability and thus no
geofluid. Davenport partnered with AltaRock Energy to develop a proof of concept HDR project
at Newberry using $23.45M of DOE funding out of a total budget of $43.81M. Utilizing one of
the original wells, the goal of the project is to stimulate the rock by pumping cold water from the
surface thus creating a fractured network of rock with in situ water. In order to create more
fractures per well, AltaRock plans on using proprietary diverters to create up to three fracture
zones per well.
To date the project has encountered significant permitting delays; however recently all
permitting was completed and AltaRock begin its first stimulation test in October 2012. Of note
is an extensive seismic monitoring array which was installed prior to the stimulation in order to
get a seismic base line. As the system is stimulated seismic events will be compared to the
baseline in order to better understand induced seismicity in EGS projects.
7-33

Geothermal Energy

7.4.3 Geothermal Technologies for Electricity Production Deployment Curve


(Grubb Curve)
Figure 7-16 illustrates the Grubb curve, which communicates the maturity of different
geothermal technologies. The curve can be used to understand the cost required to complete
research from concept to finished product or to complete a particular stage of technology
development. While early-stage development does not require as much capital relative to
development at middle stages, technologies in the last stages of development may require
significant capital to purchase major components and conduct full-scale test over many months
or years.
Technical risk can also be assessed using the Grubb curve. In the first development stages, little
is known about the core technology with regard to the size, details, and costs of its final form. As
a result, the technical risk is at its highest in the early stages of development. As the technology
matures, the technical risk lessens but the overall risk to the development project can remain high
when considering the likelihood of failure.

Figure 7-16
The Grubb curve

7.4.4 Geophysical Methods in Geothermal Exploration


This section explains the diverse geophysical methods available for geothermal exploration, as
well as their respective limitations; advocates the need to integrate the different geophysical
methods to minimize drilling risk and development costs; and emphasizes the importance of
complementing geophysical information with geological information to produce reliable results.

7-34

Geothermal Energy

Geophysical methods in geothermal exploration can be divided into four main groups depending
on the physical parameters measured, as follows:

Potential methods based on the density and magnetic properties of rocks as well as two of the
Earths potential fieldsmagnetism and gravity

Electric and electromagnetic methods, based on the electromagnetic properties of rocks


(conductivity, permittivity) and the Maxwell equations

Seismic methods, based on the elastic properties of rocks and the equations of wave
propagation in continuous media

Radiometric methods, based on the radioactive emission of rocks and atomic physics
equations (These methods are most commonly used in well logging.)

Each method has a specific application, depending on the physical properties of the target and
how precisely these properties can be detected by the technology available.
Gravimetric methods are comparatively easy to use and fairly economical. They provide a good
estimate of the extent of bodies with certain density. The resolution and quality of data, however,
decrease considerably with depth. Gravimetric studies therefore provide a useful tool for shallow
reservoirs.
Similarly, magnetic methods have been very popular during the past 30 years because of the
rapidity with which the measurements can be made and the low cost of the operation. Restrictions
are the depth of resolution, the complexity of the interpretation which compromises reliability for
structures with complicated geometric shapes, and insensitivity to the actual presence of water.
Methods for measuring the electrical resistance of the sub-surface can be divided in two general
groups:

Those that measure the difference in electrical potential

Those that measure an electromagnetic field, naturally or artificially created

Electrical potential has been used mainly for shallow depths; for example, for groundwater
aquifers or very shallow geothermal reservoirs.
The most common geophysical methods used today for geothermal exploration are
electromagnetic. They can be induced actively via the transient electromagnetic (TEM) method,
which is now routinely applied to depths of 2000 m. For greater depths, the magnetotelluric
(MT) method, which measures the Earths impedance to naturally occurring magnetic waves, has
become the standard approach.
Seismic methods use the propagation of elastic waves, which are either generated artificially by
an explosive source or occur naturally due to earthquake activity. Active seismic methods are the
standard tool for hydrocarbon prospecting, as they can be used to supply a detailed image of the
subsurface structure in the sedimentary environment of most oil and gas reservoirs. Passive
seismology, if recorded appropriately, can be used to help understand the structural context or to
give an outline of the actual fluid/geothermal reservoir.

7-35

Geothermal Energy

Electrical and seismic methods are also used for down-hole tools. Some specific tools were
specifically developed for well logging, such as to perform radiometric measurements of the rock
units accessed by the well via neutron and gamma ray measurement.
The principal methods used for geothermal regional studies are MT and seismic surveys,
whereas gravity and aeromagnetic approaches are often applied to increase the reliability of
conclusions based on the former two methods. Seismic tomography of the Earths crust, which is
quite expensive, reveals geometrical boundaries. Without geological information, none of the
geophysical methods can produce reliable results upon which to base decisions on the existence
of a geothermal reservoir and its boundaries. Therefore, seismic and electromagnetic methods are
often supplemented by gravity and/or aeromagnetic surveys of a different scale that depends on
the required resolution. Magnetic anomalies correlate quite well with horizontal zonation of the
structure, while gravity maps are used to locate deep faults and delineate fracture zones
characterized by low density.
Because the mining cost (exploration and drilling) to access geothermal resources represents
over 60% of a projects total investment, a reduction in mining cost would increase the
competitiveness of geothermal energy significantly. This goal can be achieved by developing
innovative detection methods prior to drilling fluid bearing zones in naturally and/or artificially
fractured geothermal reservoirs.
There is substantial RD&D associated with geophysical methods improvement and application,
as follows:

Before and during production: Identify prospective reservoirs without drilling, define
boundaries (lateral and vertical), and improve methods to identify drilling targets (productivity).

During and after production: Continuously characterize the reservoir, predict fluid
circulation during stimulation, track injected fluids and predict reservoir
performance/lifetime (effectiveness and sustainability).

7.4.5 Geothermal Drilling Technology


Well costs are a significant economic component of any geothermal development project. Insight
into geothermal well costs is gained by examining trends from experience in the oil and gas
industry. The similarity between oil and gas wells and geothermal wells makes it possible to
normalize any geothermal well cost from the past three decades to present current values, so that
the well cost can be compared on a common dollar basis.
Wells are classified into the following three categories:

Shallow wells (1,5003,000 m)

Mid-range wells (4,0005,000 m)

Deep wells (6,00010,000 m)

7-36

Geothermal Energy

Shallow wells supply the majority of current hydrothermal plants. Unfortunately, however, a
scarcity of geothermal well cost data makes it impossible to estimate statistically meaningful
well costs via a comparison to oil and gas wells costs. There is simply too large a degree of
uncertainty; an average completed geothermal well may cost up to 30% more than oil and gas
well for the same depth (up to 3,000 m). Well design concepts and predictions for the deeper
categories6,000 m, 7,500 m, and 10,000mwould be even more speculative, as there have
been only two or three wells drilled close to depths of 10,000 m in the United States.
Emerging technologies, which have yet to be demonstrated in geothermal applications and are
still going through development and commercialization, have the potential to significantly reduce
well costs, especially those at 4,000 m and deeper.
Technologies with the potential to reduce well costs include the following:

Expandable tubular casing

Drilling with casing that may permit longer casing intervals

Well design changes, particularly involving the use of smaller increments in casing diameters
with depth

Under-reaming, a key enabling technology for almost all of the EGS drilling applications

Three revolutionary drilling technologies that may have a particularly large impact on lowering
EGS drilling costs are hydrothermal flame spallation and fusion drilling, chemically enhanced
drilling, and metal shot abrasive-assisted drilling. Each of these methods augments or avoids the
traditional methods of penetration based on crushing and grinding rock with a hardened material
in the drill bit itself, thereby reducing the tendency of the system to wear or fail.
In addition, rate of penetration (ROP) issues can significantly affect drilling costs in crystalline
formations by increasing costs by as much as 20% above those for more easily drilled basin and
range formations.

7.5 Cost and Economic Issues


Geothermal power systems combine fuel supply and power conversion systems into one system.
The geothermal fluid serves as the equivalent of fuel. Unlike fuel-fired power systems, in a
geothermal plant the fuel supply (the geothermal resource) and electricity generation (the power
plant) are integrated and physically connected. As a result, the power plant cost is even more
site-specific than for other power generation technologies. The cost is affected not only by the
size and design of the power plant, but also by the geothermal resource temperature and pressure,
steam, impurity and salt content, and well depth.
This close and direct link is emphasized in the cost breakdowns provided here by assuming a fuel
cost of zero, and including the capital and operating costs of the geothermal field as part of the
costs of the power plant itself. In this way, geothermal is treated with the same type of cost
breakdown as solar and wind energy. In contrast, biomass is treated as having a fuel supply that
is separate from the power plant system, analogous to a coal-, gas-, or oil-powered plant, which
results in a fuel cost line item in the cost breakdowns provided.

7-37

Geothermal Energy

Geothermal power plants are able to generate energy at all times, except during planned and
unplanned maintenance and repair episodes. Experience has led to longer periods between major
overhauls, with many operators achieving five-year time spans. Because of the strong economic
incentive to stay on line, unplanned maintenance outages are kept to a minimum through the
application of highly reliable equipment and adequate spare parts inventories. Some facilities are
achieving capacity factors of 85% to 95%. Plant operators prefer to run their plants as baseload
units because the fixed costs of operation far exceed the variable costs. An exception occurs
when the fuel supply (steam) is less than that needed to satisfy the power plants capacity. In
such a case, the plant will operate at or near full capacity during peak-use periods (highest price
for electricity) and then back down during off-peak periods.
Experiments have shown that some geothermal power plants, such as the dry steam plants
at the Geysers facility in northern California, can be cycled to follow system load in the
intermediate-baseloaded area of the utility time-demand curve, thereby increasing their value in
certain applications. It is likely that load following would be more difficult for flash and binarycycle plants than for dry steam plants. Current contract capacity factors are on the order of 60%
to 80%, but capacity factors for many operating plants are close to 100%. A capacity factor
greater than 100% is possible, because some plants can operate above their design and nameplate
rating nearly all the time. Recent cost reductions in flash power plants may enhance their
competitiveness in the United States. In the international market, geothermal development in
regions such as Indonesia and the Philippines is returning to the strong growth that began in the
early 1990s [2].
During the 1990s, competition from low-cost electricity generated from natural gas hurt the
economic competitiveness of geothermal electric systems, slowing their development and
deployment. During the early 1990s, geothermal power developers expected to compete
successfully against 6 to 7 cent/kWh power in the western United States. By the end of the
decade, they instead found themselves competing against 2.5 to 3.5 cent/kWh electricity. This
situation may be repeating because of the drop in natural gas prices due to the success of
fracking. As of the publication of this report natural gas prices were at an average of well below
$4/MBTU at most of the hubs in the United States and corresponding power prices were
averaging 3.7 cents/kWh. Mitigating low power prices are changing public and private
preferences for more diverse sources of electricity supply, particularly from renewable and fossil
carbon-free sources. Current and future renewable portfolio standard (RPS) demands in the
western states will likely drive geothermal power development, and correspondingly geothermal
may significantly contribute to the renewable energy power supply.
The two factors most frequently cited for limited geothermal development are permitting issues
and the high cost of resource exploration and development. Several proposed projects have been
severely delayed in the permitting process, and market prices have just recently been sufficient to
encourage more substantial project developments [3, 27]. Another lesser factor is the limited
number of economically feasible resource sites located near transmission grids. Although market
participants do not generally cite transmission constraints as a restriction on development, the
U.S. Forest Services denial of permits for transmission lines and the high costs of alternate
routing have delayed several planned projects in California [3].

7-38

Geothermal Energy

The current state of many aspects of geothermal technology was fairly well documented in the
1990s. The 1996 Next Generation Geothermal Power Plants (NGGPP) study, prepared for DOE,
EPRI, BPA, and SCE by the Ben Holt Company, characterizes conventional flash and binary
technology and evaluates new technologies proposed for the next generation of geothermal
power plants [13]. Prior to that study, it had been difficult to obtain current cost and performance
data for geothermal power plants because of the proprietary nature of such information. Newer
cost models still reference this document for their base case comparisons [1, 24, 30]. For HDR,
the cost estimates provided for creating a HDR reservoir, as well as its performance, were based
on assessments made by scientists at Los Alamos National Laboratory, where HDR had been
studied in previous years [11]. Table 7-5 provides an updated cost breakdown, performed by
EPRI, of four different 50-MW geothermal plants in 2009 U.S. dollars.
Table 7-5
Estimated costs for a 50-MW geothermal plant (Dec 2009 $)
Flash Steam
Capital Cost
($/kW)

Break-Even
Generation
(/kWh)

Binary

Reverse
Refrigeration
Cycle

EGS

Plant: $1230
$1540

Plant: $154$1850

Plant: $1850
$2050

Plant: $1230
$2460

Well field: $770


$1540

Well field: $970


$2050

Well field: $770


$1540

Well field: $1850


$6660

3.6 5.6

4.6 9.2

6.2 12.3

7.7 16.4 +

7.5.1 Example Geothermal Project Costs


Data specific to California and Nevada from a 2005 assessment by the California Energy
Commission [32], suggest that the capital cost of incremental generation capacity averaged about
$3,100/kW installed (see Table 7-6). For California sites alone, the average capital cost of
incremental generation capacity was somewhat lower, about $2,950/kW installed. These cost
estimates include the following components:

Exploration (up to siting of the first deep, commercial-diameter well)

Confirmation drilling (up to achieving 25% of required capacity at the wellhead)

Development drilling (up to achieving 105% of required capacity at the wellhead)

Construction of the power plant (including ancillary site facilities)

Transmission-line costs

7-39

Geothermal Energy
Table 7-6
Estimated geothermal development costs (Dec 2009 $)
Most Likely
Incremental Capacity
(Gross MW)

Average
Development Cost
Per kW Installed

California

3,000

$3,024

Western Nevada

1,300

$3,536

Total

4,300

$3,178

Region

7.5.2 Factors that Influence Geothermal Power Plant Costs


Many factors influence the cost of a geothermal power plant. In general, geothermal power
plants are affected by the cost of steel, other metals, and laborfactors universal to the power
industry. However, drilling cost may vary as well. Geothermal projects are site-specific, thus the
cost to connect to the electrical grid varies from project to project. Also, whether the project is
the first in a particular area or reservoir will impact both risks and costs. The acquisition and
leasing of land also varies, because to fully explore a geothermal resource a developer is required
to lease the rights to 2,000 acres or more. Challenges to leasing and permitting vary from project
to project; especially on federal lands. These factors include the following:

Size of the plant

Power plant technology

Knowledge of resource

Temperature of the resource

Chemistry of the geothermal water

Resource depth and permeability

Environmental policies

Tax incentives and subsidies

Markets

Financing options and costs time delays

7.5.3 Engineering and Economic Evaluation


Three geothermal power plant cases are addressed: 50-MW and 40-MW single-flash steam and
50-MW single-pressure binary-cycle systems. Table 7-7 summarizes the design assumptions for
the three cases.
Table 7-8 presents performance and cost estimates for 50- and 100-MW flash steam and 50-MW
binary-cycle geothermal power plants. The data are from the 2010 EPRI report, Engineering and
Economic Evaluation of Geothermal Power Plants [41]. Table 7-9 gives the levelized cost of
electricity for the three cases.
7-40

Geothermal Energy
Table 7-7
Design assumptions for geothermal plants
Design Assumption
Flash stages

Rated capacity and


configuration

Flash Steam

Binary Cycle

Comments

Single

N/A

Single is more common.


Dual flash systems have
slightly higher efficiencies
and outputs.

50 MWn

50 MWn

Bottom-exhaust, dual-flow
turbine.

40 MWn

N/A

Top-exhaust, dual-flow
turbine.

Turbine inlet pressure


(flash)

6 bara
(1 bara = 1 bar
absolute)

NCG System

Two-stage
hybrid system

None

Type of binary cycle

N/A

Single pressure,
recuperated or
unrecuperated; working
fluid optimized for
conditions; 50 MWn

Not considered/compared:
mixed working fluids;
combined cycles; or mixed
cycles.

Working fluids considered

N/A

Iso-pentane

Preferred for higher


temperature applications.

Working fluid composition

N/A

Pure

No impurities or NCGs
considered.

Cooling

Wet cooling
tower

Air-cooled condenser

Makeup water

Condensate
from steam

H2S abatement

Incinerator

Midrange for many new


flash plants.

None

Flash based on use of


condensate and cooling
towers. Binary based on
air-cooled condensers.

None

Assumes combustible
mixture of H2S, methane,
hydrogen that might be
present in noncondensable gases, to
allow use of an incinerator.
Other abatement options
would likely be more costly.

7-41

Geothermal Energy
Table 7-8
Performance and total capital requirement estimates for geothermal power plants (December
2010 $)
Source: EPRI [41]
Flash Steam 50
MWn

Flash Steam, 40
MWn

Binary Cycle 50
MWn

Rated Capacity
Net Capacity, kWn

50,000

40,000

50,000

Gross Capacity, kWg

54,560

43,880

57,580

Brine Flow Rate, kg/s

380

307

458

Resource Temperature, C

280

280

280

Physical Plant
Plant Life, Year
Land Required, m2

30

30

30

30,000

25,000

40,000

December, 2010

December, 2010

December, 2010

Capital Costs, $/kWg


Month/Year Dollars
Identification Phase

$15

$15

$15

$1,558

$1,595

$1,737

$501

$504

$573

$1,598

$1,628

$1,650

Total Plant Indirect Construction


Cost

$786

$868

$665

Project and Process Contingency

$4,458

$4,565

$4,639

$293

$300

$305

$4,752

$4,865

$4,945

$238

$243

$247

$4,989

$5,109

$5,192

76.4

76.8

80.6

Makeup Drilling, $/MWh

3.6

4.5

3.6

Total Variable, $/MWh

9.6

10.5

7.6

37.1

36.8

30.8

96

96

96

Technology Development Rating

Commercial

Commercial

Commercial

Design & Cost Estimate Rating

Preliminary

Preliminary

Preliminary

Well Field
Gathering System
Total Plant Direct Construction Cost

Total Plant Cost (TPC)


AFUDC (Interest during
construction)
Total Plant Investment (inc.
AFUDC)
Owners Costs, $/kWn
Other owners costs and
contingency
Total Capital Requirement (TCR)
O&M Costs
Fixed, $/kW-yr
Nominal Variable, $/MWh*

Performance/Unit Availability
Exergetic Efficiency, %
Capacity Factor, %
Confidence and Accuracy Rating

*This is the nominal annual VOM cost, which excludes makeup drilling.

7-42

Geothermal Energy
Table 7-9
Levelized cost of electricity estimates for geothermal power plants (30-year project life,
constant December 2010 $)
Source: EPRI [41]
Levelized Cost of Electricity ($/MWh)

Flash Steam
50 MWn

Flash Steam,
40 MWn

Binary Cycle
50 MWn

Capacity Factor

96%

96%

96%

Capital Charge @ 8.26%/yr based on


Total Plant Cost

$47.7

$49.1

$52.4

Fixed O&M

$9.1

$9.1

$9.6

Variable O&M

$9.6

$10.5

$7.6

Total

$66.4

$68.7

$69.6

7.6 Environmental Issues


Although geothermal energy is a renewable energy technology, it is not free of environmental
concerns. These concerns include land use issues in addition to emissions and chemicals that
may be entrained in the steam and must be captured and processed. Furthermore, geothermal
energy generation has resulted in land impacts such as subsidence and induced seismicity.
7.6.1 Land Use
Land use issues are similar to those experienced by conventional power plants. Geothermal
power plants are sized according to the volume, temperature, and chemistry of the geothermal
reservoir underlying the project.
Land use patterns for geothermal power plants vary with site design and type of installation.
While a well field may cover 100 to 200 acres, heat is tapped by drilling slant wells from a single
aboveground location. A typical 50-MW plant and related well sites may occupy one or two
acres of land. In some cases, a single plant can require multiple wellhead locations. In that case,
aboveground lines carry the resource to the power plant.
A table developed for the EPRI report Greenhouse Gas Reduction with Renewables [8] and
based on land use figures in the DOE/EPRI RETC report [10], compares and quantifies the land
use requirements of the various renewable energy resources and technologies, as shown in Table
7-10. Note that geothermal energy and solar energy are more land-area-effective, measured in
kilowatts per acre, than wind and biomass. The area needed for a geothermal power plant itself is
very small, about two acres for a 50-MW plant, but the value most relevant to the comparison is
the area of the entire geothermal field, not just the site of the power plant. When doing such
comparisons of land use, two points should be kept in mind:
1. Both wind and geothermal energy production are compatible with other simultaneous uses of
much of the same land. This is especially true of wind, as shown in photographs of agricultural
land with wind turbines on islands within fields of crops or grazing cattle. In the geothermal
case, the pipelines to move geothermal fluids from wellhead to power plant and the wellhead
sites themselves disrupt other land uses. In New Zealand geothermal fields, cattle or sheep can
be seen grazing on land between the well sites and pipelines (see Figure 7-17).
7-43

Geothermal Energy

2. In the United States and in many other countries as well, land is not the limiting factor in
most settings. This is especially important in considering biomass energy, where the land
comparison should be made not to the land use of other energy sources but rather to the land
used for agriculture and forestry.
Table 7-10
Land area requirements for current renewable technologies
Land Area for 100 GWe

Unit Size
(MWe)

Area to
Support Unit
Size (Acres)*

Power Per
Unit of Area
(kW/acre)

Area
(106 Acres)

Fraction
of U.S.**

Solar Thermal

75

408

180

0.54

0.028%

PV, Residential

0.0026

0***

NA

PV, Utility-Scale

2.4

24

100

0.053%

Wind

50

3,707

14

7.4

0.39%

Geothermal

50

420

120

0.84

0.044%

Biomass

50

39,400

1.3

78.7

4.1%

* To convert acres to hectares, multiply by 0.4.


** Fraction is of the 1.9 billion acres of the continental U.S.
*** No land is required for residential PV systems, which are typically installed on existing structures.

7.6.2 Subsidence
Land above or adjacent to a subsurface geothermal reservoir may sink or subside due to the
extraction of geothermal fluid from the reservoir. Subsidence is not always easy to detect and
may be of no consequence.
Several areas of subsidence have been detected in the New Zealand fields. At Wairakei, the
oldest geothermal field in New Zealand and one that has operated at close to 170-MWe capacity
for approximately 40 years, one area of subsidence has created a pond at a low spot in a stream.
Another has been detected over an area of approximately 1 km2 and could eventually lead to
rerouting a section of roadway. A series of elevation surveys, conducted every several years
throughout the region of the Wairakei field and environs, made it possible to detect this rather
inconspicuous subsidence.
In anticipation of geothermal development expansion, biennial surveys of elevations in and near
the Imperial Valley (California) geothermal fields were performed in the 1970s. Because of a
gravity flow ditch irrigation system in a low desert valley and the agriculture-based economy,
the surveys were needed to ensure that no harm was caused by geothermal production and to
alert and correct the situation if adverse trends developed. Several hundred megawatts of
generating capacity have been installed and operated in the Imperial Valley over the past two
decades, and there has been no significant subsidence.

7-44

Geothermal Energy

(Evan Hughes for EPRI)

Figure 7-17
Steam pipelines for the Ohaaki, New Zealand plant run through land used for
grazing and agriforestry

7.6.3 Emissions
The environmental impacts of generating electricity from geothermal resources are benign
relative to conventional power generation options. Geothermal power generation does not
produce the federally regulated air contaminants commonly associated with other power
generation options, including sulfur dioxide, particulates, carbon monoxide, hydrocarbons, and
photochemical oxidants. Some hydrothermal fluids contain hydrogen sulfide and/or high levels
of dissolved solids such as sodium chloride (salt). Hydrogen sulfide emissions are abated, when
necessary, with environmental control technology. Emission factors are typically in the range of
45 kg/MWh for carbon dioxide and 0.0145 kg/MWh for hydrogen sulfide. Hydrogen sulfide is
now routinely abated at geothermal power plants, resulting in the conversion of over 99.9% of
the hydrogen sulfide from geothermal non-condensable gases into elemental sulfur, which can be
converted to a non-hazardous fertilizer feedstock. Since 1976, hydrogen sulfide emissions have
declined from 1,900 lb/hr to 200 lb/hr or less, while at the same time the geothermal power
production has increased from 500 MW to over 2,500 MW. Another hazardous element found in
a few geothermal plants is mercury. Although mercury is not present in every geothermal
resource, where it is present mercury capture equipment typically reduces emissions by 90% or
more. The relatively highest mercury emitters are two facilities at the Geysers in California.
They release mercury at levels that do not trigger any health risk analyses under California
regulations. The Soultz EGS project in northeastern France near the German border has
encountered radioactive scaling due to a radiogenic source within the geothermal reservoir.
Plant operators have implemented polyphosphonates and other products from the oil and gas
industry to inhibit scaling. It is not known whether this will be an issue in future EGS projects;
research has been proposed to assess the level of probable occurrence in other locations.
7-45

Geothermal Energy

The visible plumes seen rising from some geothermal power plants are, in fact, water vapor
(steam). A case study by GEA [27] showed that a coal plant updated with scrubbers and other
emissions control technologies emits 24 times more carbon dioxide, 10,837 times more sulfur
dioxide, and 3,865 times more nitrous oxides per megawatt-hour than a geothermal plant.
Averages of four significant pollutants emitted by geothermal and coal facilities are presented in
Table 7-11. Following the table is a brief discussion of other emissions that have sometimes been
associated with geothermal development. Issues revolving around groundwater contamination
are avoided through protective well completion practices. The geothermal industry has
developed various methods to mitigate chemical emission concerns.
Table 7-11
Averages of four significant pollutants, as emitted from geothermal and coal facilities
Source: GEA 2005 [27]
Nitrogen
Oxides (NOx)

Sulfur Dioxide
(SO2)*

Particulate
Matter (PM)

Carbon
Dioxide (CO2)

00.35

060

Coal Emissions (lb/MWh)

4.31

10.39

2.23

2191

Emissions Offset by
Geothermal Use (per yr)

32,000 tons

78,000 tons

17,000 tons

16 million tons

Emission
Geothermal Emissions
(lb/MWh)

* Although geothermal plants do not emit sulfur dioxide directly, once hydrogen sulfide is released as a gas into the atmosphere,
it eventually converts to sulfur dioxide and sulfuric acid. Therefore, any sulfur dioxide emissions associated with geothermal
energy derive from hydrogen sulfide emissions.

Generally, there is less likelihood of adverse environmental impacts from binary-cycle


generation than from flash steam generation because the hotter geothermal fluids (steam and
water) used in flash plants tend to contain greater concentrations of chemical contaminants than
do the cooler geothermal fluids (water) typically used in binary plants. In general, hotter water
will dissolve more mineral from a rock formation than will cooler water. Also, in binary plants
that employ dry rather than wet cooling systems, the geothermal fluid remains in a closed system
and is never exposed to the atmosphere before it is injected back into the reservoir. The potential
environmental impact of a HDR binary system is likely to be even lower. This is because the
water used in such a system originates from a shallow groundwater well or other source with low
levels of dissolved solids and no hydrogen sulfide to begin with, and because the limited contact
time of the injected surface water with the hot rock will in most cases limit the level of dissolved
mineral (including dissolved gas) to below that of chemical equilibrium.
In some situations, geothermal operations offer novel benefits to the environment. For example,
operators of the Geysers complex in northern California are injecting treated municipal
wastewater into underground steam-generating reservoirs whose natural stores are being
depleted. This technique helps meet the generation facilitys resource challenges in addition
to mitigating the waste disposal problems of nearby communities.

7-46

Geothermal Energy

7.6.4 Thermal Discharge


Geothermal power plants are intrinsically less efficient than fossil-fuel-fired power plants. This is
because high-temperature steam can be converted into work at higher thermal efficiency than lowtemperature steam. Temperatures of steam produced in fossil fuel boilers are in the range of 800 to
1000F (400500C), much higher than the 300 to 600F (150300C) of geothermal steam and
hot water. Less heat converted to work and electricity means more heat wasted, much of which
is discharged into the environment. Fossil fuel steam plants have thermal efficiencies in the 32% to
38% range, nuclear plants close to 30% to 32%, and geothermal plants as low as 5% to 17%.
For this reason, geothermal power plants have very large cooling systems relative to other power
generation options. The Ohaaki Power Station in New Zealand installed a natural-draft cooling
tower, one that appears comparable to those seen at nuclear power plants and at some fossil fuel
power plants located on rivers (Figure 7-17). The reason for the cooling tower at Ohaaki is the
same as that at U.S. nuclear or fossil plants: to avoid the discharge of relatively large amounts of
waste heat into a stream, river, or estuary of limited capacity and the resulting increased
temperature of water returned to the source.
In many geothermal locations, water is scarce and air cooling is the only option. Air cooling can
be done by using fans or natural-draft cooling towers. The fan-driven cooling system can be built
quickly and at relative low cost, but its operating costs are high due to maintenance costs and
parasitic losses associated with the fans. Innovative cooling concepts are being developed to
improve the efficiency of heat exchangers and reduce the capital cost of natural cooling towers
by using hybrid air condensers and storing cool capacity (as either sensible or latent storage) to
extend night time cooling through the day.
7.6.5 Induced Seismicity
Seismicity has been correlated with underground fluids extraction and injection. Seismic activity
has steadily increased as production has increased at the Geysers geothermal field in northern
California. In most geothermal fields, fluids are being withdrawn and injected in the same
region, thermal and chemical effects are also occurring, and all of these effects can combine to
produce a variety of mechanisms for inducing seismicity.
Water injection into geothermal systems, which has become a nearly universal and often required
strategy for extended and sustained production of geothermal resources, seems to be a common
cause of induced seismicity. However, the data available do not support the hypothesis that
induced seismicity in geothermal regions is caused by injection.
In many geothermal fields and potential geothermal fields in the United States, the potential for
induced or naturally occurring seismicity (less than grade 5) should not cause significant
damage. This is the case for two reasons:
1. Faults are not close enough to the geothermal plants to create a large damaging event
2. If faults are nearby, geothermal production and injection activities are occurring at depths
less than 5 km, which is shallower than most seismic activity (which tends to occur at depths
of between 5 and 10 km).

7-47

Geothermal Energy

Overall, the impact of induced seismicity on the implementation of various different enhanced
geothermal activities will depend on the risk associated with the activity and the cost-benefit
ratio. All experience to date has shown that the risk, while not zero, has been either minimal or
can be handled in a cost-effective manner. Recently a protocol for assessing induced seismicity
risks for EGS projects was released by the Department of Energy Geothermal Technologies
Program [50].

7.7 Design, Deployment, and O&M Issues


Many geothermal developers acquire the rights to a geothermal resource and rely on service
companies for exploration, drilling, well testing, reservoir engineering, and power plant design
and construction. However, most developers handle their own power plant operation.
For domestic power projects, coated corrosion-resistant pipe and heat exchangers are typically
obtained from U.S. manufacturers. However, virtually all steam turbines are purchased from
Japanese manufacturers. A few U.S. companies manufacture their own small binary and hybrid
power plant equipment. One important developer, Ormat, is the U.S. arm of an Israeli company
that uses the Israeli Ormat assembled, skid-mounted, and modular equipment.
Most of the necessary drilling and well-testing equipment is adapted from the oil and gas
industry. However, some of this equipment has been specifically refined for the hotter and more
corrosive environments of geothermal drilling. While DOEs geothermal research and
development program is working on geothermal-specific improvements to some power plant
equipment, in general the pipe, heat exchangers, turbines, and cooling towers come from the
fossil fuel power plant industry.
Technology is expected to improve incrementally over the coming years. Industry is working
with DOEs geothermal R&D program to improve exploration, drilling, reservoir engineering,
and energy conversion technology. R&D success in fracture detection and fracture permeability
studies could give the industry important new tools to improve geothermal exploration and
reservoir engineering effectiveness, and also drive down project costs. As discussed earlier under
EGS, these fracture and permeability enhancements of natural hydrothermal hot water reservoirs
could be the start of commercial activity in the HDR field. Better materials and coatings for
power plant equipment are being researched to reduce high maintenance costs. On the energy
conversion front, R&D also continues on binary geothermal systems that use a secondary
working fluid to produce energy from the more abundant moderate- to low-temperature
geothermal resources.
7.7.1 Industry Trends
Geothermal project activity is escalating significantly after relatively slow growth over the past
two decades. The recent renaissance is a reflection of geothermal's successful track record and
performance and cost attributes. Even following the significant downturn in global energy prices
during the end of 2008 and 2009, there remain great expectations for geothermal's long-term cost
competitiveness as global demand for energy escalates. Indeed, the global installed geothermal
capacity is forecasted to more than triple its current installed base to over 31 GW by 2020.

7-48

Geothermal Energy

Today, there are more than 200 commercial geothermal projects operating in 24 countries with a
cumulative installed capacity of approximately 10.7 GW worldwide. The United States is
currently the worlds top geothermal market, with more than 3 GW of installed capacity. With a
pipeline of more than 4GW of confirmed projects, the United States is poised to more than
double existing capacity over the next five years. U.S. geothermal activity will be primarily
concentrated in a few Western states that are home to vast hydrothermal resources. Developers
are expected to focus on low-hanging fruit opportunities located in California and Nevada.
Internationally, Indonesia and the Philippines are poised to ramp up their geothermal capacity in
the near term. Meanwhile, market activity in Japan, Iceland, Italy, New Zealand, and Mexico,
which account for over 65% of the remaining global installed geothermal capacity, should also
increase. Although these markets have seen relatively limited growth over the past few years,
urgency to advance low-carbon baseload power generation is helping to re-ignite new capacity
growth in these markets.
Separately, attention is turning to new markets. Chile, Turkey, Russia, Nicaragua, Germany,
Australia, and East Africa are all poised to initiate projects as greater effort is made to integrate
renewables, diversify energy portfolios, and hedge against natural gas price volatility.
7.7.2 Research Needs and Recommendations
Although geothermal is considered a mature technology, there are many opportunities for
improvement through research and development. Geothermal projects are very capital intensive.
Discussions with industry leaders suggest that that future geothermal R&D programs should
focus on reducing the cost of finding and developing new geothermal resources.
The industry consensus is that characterizing the commercial potential of identified geothermal
reservoirs is a high priority. Techniques such as fracture mapping, more accurate thermalgradient wells, and other, untested methods should be evaluated and refined, if appropriate. The
objective is to be able to measure the temperature, fluid characteristics, and permeability of the
resource prior to committing to expensive production wells and generation equipment.
Industry comments suggest that structured, integrated tests and experiments by industry in
collaboration with the California Energy Commission and DOE are essential to ensure further
development of commercial geothermal power. In the near term, this work should focus on the
following areas:

Reducing life-cycle costs of electrical power generation

Improving geothermal facility efficiencies

Finding techniques to more cost-effectively define resources

Supporting funding to reduce production drilling costs and develop direct-use demonstration
and commercial projects

Improving technology and technical support for geothermal direct-use projects

Reducing geothermal heat pump (GHP) initial costs

Improving professional training and consumer education campaigns for GHP

7-49

Geothermal Energy

Supporting financing programs for residential applications for GHP

Understanding and improving wastewater injection

Improving public awareness about the benefits of geothermal power, direct-use, and GHP

Enhanced geothermal systems are still widely experimental. A 2008 DOE survey identified
the following three critical assumptions for EGS technology that require evaluation and testing
before the economic viability of EGS can be confirmed:

Demonstration of commercial-scale reservoirs: Stimulation and maintenance of a large


volume of rock is required in order to minimize temperature decline in the reservoir. Actual
stimulated volumes have not been reliably quantified in previous work.

Sustained reservoir production: Recent studies conclude that 200C fluid flowing at
80 kg/sec (equivalent to about 5 MWe) is needed for economic viability. No EGS project to
date has attained flow rates in excess of approximately 25 kg/sec.

Replication of EGS reservoir performance: EGS technology has not been proven to work at
commercial scales over a range of sites with different geologic characteristics.

These assumptions can be tested with multiple EGS reservoir demonstrations using current
technologies. In the long term, significant reduction in drilling costs will be necessary to access
deeper resources, and the cost of converting geothermal energy into electricity must be reduced.
These improvements will move EGS technology forward as an economically viable means of
tapping the nations geothermal resources.

7.8 Key Research Participants


The following includes brief descriptions of government and non-governmental organizations
involved in geothermal energy research. Note that although the websites for these organizations
contain information and contacts that may be of interest, EPRI cannot vouch for the accuracy or
quality of all their content.
DOE Geothermal Technologies Program
The DOE administers its geothermal energy efforts through its Energy Efficiency and Renewable
Energy Network.
www.eere.energy.gov/geothermal
NREL Geothermal Technologies Program
This National Renewable Energy Laboratory program supports activities of DOEs Geothermal
Energy Program and works with DOE and industry partners in geothermal R&D efforts.
www.nrel.gov/geothermal/
Sandia Geothermal Research Department
Sandia National Laboratory hosts a Geothermal Research Department focused on developing
well construction technologies to reduce the cost of geothermal wells. Hard rock drill bit
technology, high-temperature instrumentation, rig instrumentation, and data management are
among the projects it pursues.
www.sandia.gov/geothermal/

7-50

Geothermal Energy

Idaho National Engineering and Environmental Laboratory


This laboratorys Geothermal Energy website serves as a repository for some DOE technical data
related to geothermal energy, including maps and graphics, information about U.S. laws and
standards, annual research program summaries, proposal solicitations, and other information.
https://inlportal.inl.gov/portal/server.pt/community/energy_resources_recovery/418/energy_reso
urces_recovery___sustainability_%28home%29
Stanford University Geothermal Program
Stanfords geothermal program has conducted geothermal reservoir engineering research for
more than 20 years, completing some 150 projects. Study areas have included well test analysis
of fractured and multiphase reservoirs, design and interpretation of tracer tests in fractured
reservoirs, experimental measurements of fluid flow parameters, and optimization of production
and reinjection strategies.
https://pangea.stanford.edu/researchgroups/geothermal/
Southern Methodist University
SMU has the most up-to-date data currently available for resource assessment based on the 1978
USGS data and other information collected since then.
www.smu.edu/geothermal/
European Deep Geothermal Energy Programme
This website describes a hot dry rock research program funded by a consortium of European
government, research, and utility organizations in Soultz-sous-Forts, France. The multiyear
project aims to develop a scientific pilot plant by 2006 and a subsequent commercial plant that
could produce approximately 25 MW.
http://www.soultz.net/version-en.htm
Geothermal Energy Association
GEA is a trade association composed of U.S. companies that support the expanded use of
geothermal energy and are developing geothermal resources worldwide for both electric power
generation and direct-heat uses.
www.geo-energy.org.
Geothermal Resources Council
The GRC is a nonprofit educational association formed in 1970 to encourage geothermal
research and development, and to serve as an informational conduit among researchers, industry,
and the public.
www.geothermal.org.
International Geothermal Association
Based in Italy with members in 63 nations, IGAs objective is to encourage research,
development, and utilization of geothermal resources worldwide by compiling, publishing, and
disseminating scientific and technical information.
http://www.geothermal-energy.org/index.php
Center for Renewable Energy and Sustainable Technology
The Center for Renewable and Sustainable Technology (CREST) provides excellent background
information and extensive links to other websites. CREST has been a part of the Renewable
Energy Policy Project since 1999.
http://www.repp.org/geothermal/index.html
7-51

Geothermal Energy

Calpine Corp.
Founded in 1984, Calpine operates the Geysers in northern California and is the largest producer
of geothermal electricity in the world.
www.calpine.com
Unocal Geothermal
Unocal operates three of the worlds largest geothermal electricity projects, Tiwi and Mak-Ban
in the Philippines and Gunung Salek in Indonesia, which together have a combined installed
generating capacity of 1,100 MW. Unocal merged with Chevron Corporation in 2005.
http://www.chevron.com/deliveringenergy/geothermal/
The World Bank
The World Bank provides a source of financial and technical cost data for geothermal
development in countries around the world.
http://econ.worldbank.org
GeothermalBiz.com
The Geothermal Biz website provides general information on state government agencies, for
example state energy offices, relevant legislation, e.g., for renewable portfolio standards (RPS)
and public benefits funds. Also included is information on state, local, and utility incentives that
promote geothermal energy, such as tax incentives, grants, loans, rebates, and utility green
pricing programs.
http://www.geothermal-biz.com

7.9 References
1. Renewable Energy Annual 2003. Available
http://www.eia.doe.gov/cneaf/solar.renewables/page/rea_data/rea_sum.html. EIA,
Washington, D.C.: December 2004. DOE/EIA-0603(99).
2. Huttrer, G., The Status of World Geothermal Power 19952000. Paper prepared for the
World Geothermal Conference and the DOE Geothermal Power Program, as an update of
the 1995 status report on geothermal power deployment, based on the Country Update
papers submitted to the World Geothermal Conference held in Japan, MayJune 2000.
Paper prepared by Gerald W. Huttrer of Geothermal Management Co., Frisco, Colorado:
May 2000.
3. Market Assessment of Renewable Energy in California. EPRI, Palo Alto, CA: 2001.
1001193.
4. Duffield, W.A., J.H. Sass, and M.L. Sorey, Tapping the Earths Natural Heat, U.S.
Geological Survey Circular 1125, 1994.
5. Assessment of Geothermal Resources of the U.S.1978. Ed. L.J.P. Muffler, USGS Circular
790, 1978.
6. Geothermal Energy in the Western U.S. and Hawaii: Resources and Projected Electricity
Generation Supplies. EIA: September 1991. DOE/EIA-0544.
7. Annual Energy Outlook 2007, EIA: February 2007.
8. Greenhouse Gas Reduction with Renewables. EPRI, Palo Alto, CA: 2000. TR-113785.
7-52

Geothermal Energy

9. Tester, J.W., and H.J. Herzog, Economic Predictions for Heat Mining: A Review and
Analysis of Hot Dry Rock (HDR) Geothermal Energy Technology, Massachusetts Institute
of Technology, July 1990.
10. Renewable Energy Technology Characterizations, EPRI and DOE. EPRI, Palo Alto, CA:
1997. TR-109496.
11. Brown, D. and D. Duchane, Los Alamos National Laboratory, personal communication
to Lynn McLarty (one of the authors of the RETC [10]), November 5, 1996.
12. Assessment of Geopressured Energy Resources on the Gulf Coast of the U.S. Report to EPRI
by Southwest Research Inst. (San Antonio, TX). EPRI, Palo Alto, CA: 1983. AP-3109-WE.
13. Next Generation Geothermal Power Plants. EPRI, Palo Alto, CA: 1996. TR-106223.
14. Proceedings of the 10th EPRI Geothermal Conference and Workshop, San Diego, June 1986.
EPRI, Palo Alto, California: 1986. AP-5059-SR.
15. Armstead, H.C.H. and J.W. Tester, Heat Mining: A New Source of Energy, E.&F.N.
Spon Ltd. University Press, London: 1987.
16. Mason, T.R., Improving the Competitive Position of Geothermal Energy, Proceedings of
Geothermal Program Review XIV, DOE: April 1996. Report DOE/EE-0106.
17. Heber Binary Cycle Geothermal Demonstration Power Plant, EPRI, Palo Alto, CA: August
1998. AP-5787-SR. (Note: This was the last in a series of several reports published by EPRI
during the development and implementation of the Heber Project between 1978
and 1988.)
18. Los Alamos National Laboratory. Proceedings of the 3rd International Hot Dry Rock
Energy Forum, Santa Fe, NM, May 1316, 1996.
19. EPRI Technical Assessment Guide Electricity Supply Volume 1, Rev. 7. EPRI, Palo Alto, CA:
1993. TR-102275-V1R7.
20. Annual Energy Outlook 1996, EIA: August 1996. DOE/EIA-0603(96).
21. Entingh, D.J., Technology Evolution Rationale for Technology Characterizations of U.S.
Geothermal Hydrothermal Electric Systems, BNF Technologies Inc., Alexandria, Virginia,
Draft Report: March 1993.
22. Renewable Power Industry Status Overview. EPRI, Palo Alto, CA: 1998. TR-111893.
23. Annual Energy Outlook 1998 with Projections Through 2020, EIA, Office of Integrated
Analysis and Forecasting, Washington, D.C. 20585: December 1997. DOE/EIA-0383 (98),
Distribution Category UC-950.
24. Annual Energy Outlook 2002 with Projections Through 2020, EIA, Office of Integrated
Analysis and Forecasting, Washington, D.C., December 2001.
25. Hunt, T., geophysical and geothermal consultant in Taupo, New Zealand. Personal
communication to Evan Hughes, October 2003.
26. Radon as an In Situ Tracer in Geothermal Reservoirs, Project 1992-1, Final Report to
EPRI by Professor Paul Kruger of Stanford University: August 1987. EPRI AP-5315.

7-53

Geothermal Energy

27. Kagel, A., D. Bates, and K. Gawell, A Guide to Geothermal Energy and the Environment,
Geothermal Energy Association, Earth Day 2005.
28. Bertani, R., World Geothermal Generation 20012005 State of the Art, 2005. Prepared for
World Geothermal Congress 2005, Turkey.
29. Sanyal, S.K., Classification of Geothermal SystemsA Possible Scheme, 2005. Prepared for
Thirtieth Workshop on Geothermal Reservoir Engineering, Stanford University, Stanford,
California: January 31February 2, 2005. SGP-TR-176.
30. Sanyal, S.K., Cost of Geothermal Power and Factors that Affect It, 2005. Prepared for
Thirtieth Workshop on Geothermal Reservoir Engineering, Stanford University, Stanford,
California: January 31February 2, 2005. SGP-TR-175.
31. Geothermal Resource Council, Geothermal Bulletin Volume 33 Number 3, June 2004.
32. Lovekin, J., New Study Highlights Geothermal Resources Available for Development in
California and Nevada, December 2004. GRC Bulletin, p. 242.
33. New Geothermal Site Identification and Qualification: April 2004. CEC Publication No.
P500-04-051.
34. Bertani, R., World Geothermal Generation in 2007, GHG Bulletin, September 2007.
35. U.S. Geothermal Power Production and Development Update, March 2009, Geothermal
Energy Association.
36. The Future of Geothermal Energy: Impact of Enhanced Geothermal Systems (EGS) on the
United States in the 21st Century, Massachusetts Institute of Technology, 2006.
37. An Evaluation of Enhanced Geothermal Systems Technology, Geothermal Technologies
Program, Energy Efficiency and Renewable Energy, DOE, 2008.
38. Sanyal, S.K. et al. Future of Geothermal Energy, Geothermal Energy Conference,
Canadian Geothermal Energy Association, April 2009.
39. Bertani, R., Long Term Projections of Geothermal-Electric Development in the World,
GeoTHERM ConferenceOffenburg, March 2009.
40. Brown, D.W., A Hot Dry Rock Geothermal Energy Concept Utilizing Supercritical CO2
Instead of Water Proceedings, Twenty-Fifth Workshop on Geothermal Reservoir
Engineering, Stanford University, Stanford, California, January 2426, 2000.
41. Engineering and Economic Evaluation of Geothermal Power Plants. EPRI, Palo Alto, CA:
2010. 1019761.
42. Annual Geothermal Power Production and Development Report: April 2012, Geothermal
Energy Association.
43. Renewable Energy in California, ACORE, updated 09/2012.
44. Next Generation Geothermal Power Plants: 2012 Update. EPRI, Palo Alto, CA: 2012.
1024006.
45. Single-Well Enhanced Geothermal System Front End Engineering and Design: Optimization
of a Renewable Geothermal System for Harvesting Heat from Hot, Dry Rock. EPRI, Palo
Alto, CA: 2011. 1021665.
7-54

Geothermal Energy

46. Program on Technology Innovation: Modeling of Single-Well Closed-Loop Enhanced


Geothermal Systems. EPRI, Palo Alto, CA: 2012. 1023438.
47. Geothermal/Solar Hybrid Applications. EPRI, Palo Alto, CA: 2011. 1024675.
48. Solar Thermal Augmentation of a Flash Geothermal Plant. EPRI, Palo Alto, CA: 2012.
1026879.
49. Engineering and Economic Evaluation of Low-Temperature Geothermal Power Plants.
EPRI, Palo Alto, CA: 2010. 1019775.
50. Majer, E., J. Nelson, A. Robertson-Tait, J. Savy, and I. Wong, Addressing Induced
Seismicity Associated with Enhanced Geothermal Systems, US DOE Geothermal
Technologies Program, January 2012.
51. Mississippi Oilfield Generates ow-Temperature, Emission Free Geothermal Energy at the
Wellhead. Electrotherm publication, November 2012,
http://electratherm.com/docs/Denbury%20White%20Paper.pdf.

7-55

SOLAR THERMAL

8.1 Introduction
Solar thermal electric power technology has received substantial development attention since the
early 1970s. Government energy research programs in the developed countries have invested
several billions of dollars (U.S. equivalent) over the past three decades aimed at its advancement
and commercial introduction. U.S. development efforts were particularly strong during the
1980s, and resulted in deployment of a significant amount of experimental and commercial
hardware during that decade.
Attention subsided in the 1990s primarily because conventional energy prices seemed to stabilize
at affordable levels. In addition, the primary interest of the solar power community shifted from
large-scale wholesale power to smaller-scale retail power, due in part to the emerging electric
sector restructuring movement. Key elements of that movement included increased emphasis on
dispersed, distributed generation and retail competition for electricity supply. In addition,
development issues had surfaced for the solar thermal options, and the resources to resolve those
issues were not forthcoming because of the markets characteristics at the time. Nonetheless,
several major field deployment and advancement activities were carried out in the 1990s,
resulting in important new contributions that enhanced the prospects for these technologies.
Driven by looming capacity shortfalls, the new century found the restructuring movement
slowing and industry interest in central-station power plants increasing. Growing public interest
in renewable energy, fueled by pending carbon emissions constraints and renewed concerns over
energy prices and security, has put solar thermal electric generation back into serious
consideration. In the past few years, there have been indications that the time has come for
significant new investments in solar thermal. New plants are now operating in the United States,
Spain and elsewhere, with thousands of more megawatts of solar thermal capacity on the
drawing boards. Real-world operating experience is bolstering confidence that solar thermal
plants can run reliably, and further understanding of their engineering challenges, operating
issues, and O&M costs will be possible as such data are made public. Anecdotal evidence says
that industry experience with operational plants has been generally favorable.

8-1

Solar Thermal

Overviews of the various solar thermal technologies are given in Tables 8-1 through 8-4.
Table 8-1
Central receiver technology
Installed Capacity (est.)

64 MW worldwide (as of December 2012)

Technology Readiness

Demonstration to early-commercial
Moderate-low confidence in cost estimates and technology projections

Environmental Impact

Minimal during manufacturing and operation


Most modern designs use steam or molten salt for heat transfer and
storage, which are easily contained if spilled
Significant land usage may impact species
Water usage concern in some desert locations; most plants will use dry
cooling

Economic Status

Potentially high financing costs for first utility-scale systems, mitigated in


the United States by DOE loan guarantees
Potential for lower cost with higher operating temperatures
Long-term performance remains to be demonstrated, with significant
uncertainty about component and O&M costs

Policy Status

Along with other solar technologies, encouraged by RPS and energy


policies in most jurisdictions

Trends to Watch

Large projects planned in the U.S., Europe, South Africa, Turkey, and
China
Thermal storage to allow dispatchable solar power; value of storage
depends on whether local time-of-use rates and other pricing and policy
circumstances make it worth the investment
Molten salt and direct steam projects in Spain, South Africa, Turkey,
China, Nevada, Arizona, and California

Table 8-2
Dish/engine technology
Installed Capacity (est.)

2.6 MW worldwide (as of December 2012)

Technology Readiness

Early demonstration
Low confidence in cost estimates and development projections

Environmental Impact

Minimal during manufacturing


None in operation

Economic Status

Long-term performance remains to be demonstrated, with component


and O&M costs largely uncertain

Policy Status

Along with other solar technologies, encouraged by RPS and energy


policies in most jurisdictions

Trends to Watch

Increasing field experience leading to better understanding of long-term


performance and O&M costs
R&D to adapt thermal energy storage to dish/engine system
New materials and techniques to reduce manufacturing and O&M costs

8-2

Solar Thermal
Table 8-3
Linear Fresnel reflector technology
Installed Capacity (est.)

42 MW worldwide (as of December 2012)

Technology Readiness

Demonstration
Low confidence in cost estimates and development projections

Environmental Impact

Minimal during manufacturing


None in operation
Significant land usage may impact species; somewhat lower footprint than
other solar thermal technologies
Water usage concern in some desert locations

Economic Status

Successful early demonstrations suggest economic competitiveness in


some markets and applications
The 2009 acquisition of Ausra by large French company Areva, and Arevas
stated intent to make CLFR the centerpiece of its solar energy business,
may bode well for future investment and prospects. In addition, other
corporations are expressing quiet interest in developing CLFR projects
Long-term performance and reliability remains to be demonstrated, with
great uncertainty about component and O&M costs

Policy Status

Along with other solar technologies, encouraged by RPS and energy


policies in most jurisdictions

Trends to Watch

Large-scale plants in Australia, India, Jordan, Spain, and France


SkyFuel and AREVA Solar plan for molten salt as heat transfer fluid and
storage media, supported by DOE research grant

Table 8-4
Trough technology
Installed Capacity (est.)

2,270 MW worldwide (as of December 2012)


526 MW in U.S.

Technology Readiness

Commercial
Moderate confidence in cost estimates and technology projections

Environmental Impact

Minimal during manufacturing and operation


Significant land usage may impact species
Water usage concern in some desert locations; plants may use dry
cooling
Synthetic oil heat transfer fluid is a volatile organic compound (VOC)

Economic Status

Competitive in some markets with high government subsidies; aided in


United States by DOE loan guarantee
Resurgence in new system construction after two-decade layoff; lost
manufacturing capability and experience being regained

Policy Status

Along with other solar technologies, encouraged by RPS and energy


policies in most jurisdictions

8-3

Solar Thermal
Table 8-4 (continued)
Trough technology
Trends to Watch

Hybrid operation and thermal storage


New solar trough projects in Arizona, California, Spain, UAE, India, South
Africa, Egypt, Morocco, and elsewhere
DOE loan guarantees backed by ARRA stimulus funding are providing
significant incentive in the United States
New materials and techniques to reduce manufacturing and O&M costs
Lower-cost mirror support structures
Evident interest in non-glass mirrors and reflectors even while glass
manufacturers are emerging to support the solar thermal market
Interest in direct steam generation remains high
Research continues on advanced, alternative heat transfer fluid materials,
including molten salts that remain liquid over a wider range of
temperatures

8.1.1 Solar Thermal Technologies


There are four common types of solar thermal power systems: parabolic trough, central receiver
or power tower, dish/engine, and linear Fresnel reflector (Tables 8-1 through 8-4). Another
technology, solar chimney, has been proposed but has not yet been demonstrated for electricity
production at large scales. Except for the solar chimney, all of these technologies involve a heatdriven engine and most can be readily hybridized with fossil fuel. Some can be adapted to use
thermal storage. The primary advantage of hybridization and thermal storage is that the
technologies can provide dispatchable power and operate during periods when solar energy is not
sufficient. Thus, hybridization and thermal storage can enhance the economic value of the
electricity produced.
Parabolic trough systems use banks of trough-shaped mirrors with a parabolic cross-section to
focus sunlight on highly absorbing receiver tubes that contain a heat-transfer fluid (HTF). This fluid,
typically a synthetic oil, is heated and pumped through a series of heat exchangers to produce
superheated steam that powers a conventional turbine generator to produce electricity. Nine trough
systems, built in 1984 to 1990, are generating 354 MW in southern California. These 14- to 80-MW
systems are hybridized to derive up to 25% of their output from natural-gas firing and provide
dispatchable power independent of the solar energy available. In addition, a 1-MW trough system,
commissioned in 2006, is operating in Arizona, and a 64-MW trough facility in Nevada became
operational in early 2007, and a 75-MW solar field is augmenting a natural gas combined-cycle plant
in Florida. The first of several Spanish 50-MW trough plants was completed in late 2008 and 32
such plants are now operational. The Spanish plants are especially notable because more than a
dozen include 7.5 hours or more of molten-salt thermal energy storage.
Central receiver systems, also referred to as power towers, use a field array of large mirrors
called heliostats that track the sun and focus its light onto a central receiver mounted on top of a
tower. The first central receiver in the United States, Solar One, was installed in southern
California and operated in the mid-1980s. It used water as the HTF and generated steam directly
to power a 10-MW steam turbine. In 1992, a consortium of U.S. utilities, DOE, and EPRI formed
to retrofit Solar One and demonstrate a system, aptly named Solar Two, which incorporated
8-4

Solar Thermal

molten salt as the HTF and energy storage medium. In this system, molten salt was pumped from
a cold tank and cycled through the receiver, where it was heated and returned to a hot tank.
The hot salt could then be used to generate electricity when needed. Current designs allow
storage times ranging from 2 to 15 hours. Such thermal storage capability makes central
receivers (and troughs, when so equipped) the most flexible of solar technologies, promising
dispatchable power with load factors of up to 75%in principle, even 100%. Solar Two
operated from April 1996 to April 1999. The EUs first commercial 11-MW central receiver
project, PS10, became operational in Spain in August 2007, and the second, PS20, began
operation adjacent to PS10 in April 2009. Also in Spain, the 17-MW Gemasolar central receiver
project began operation in May 2011. Meanwhile, a number of new power tower projects are
being proposed for construction over the next five to seven years in the U.S., EU, South Africa,
Turkey, and China.
Dish/engine systems use an array of mirrors made from glass facets to form a parabolic dish that
focuses solar energy onto a receiver located at the focal point of the dish. A HTF, typically
helium or hydrogen, is heated in the receiver tube and used to generate electricity in a small
engine attached directly to the receiver. Current designs employ a Stirling engine, but future
designs could use Brayton-cycle (turbine) engines or dense arrays of high-efficiency
photovoltaic cells. The former technology leader, Stirling Energy Systems (SES), demonstrated a
six-dish 150-kW system with Sandia National Laboratories and signed power purchase
agreements (PPAs) in 2005 for up to 1,750 MW, but in 2011 SES declared bankruptcy, and those
large projects were converted to PV generation. Meanwhile, the 60-dish, 1.5-MW Maricopa
dish/engine plant in Arizona which began providing electricity to Salt River Project in January
2010 continues to operate. Another company, Infinia has developed a little over 1 MW of 3-kW
systems and is under DOE contract to incorporate thermal energy storage into its receiver.
Near- and long-term cost projections for trough technology are higher than those for power
towers and dish/engine systems due in large part to the lower solar concentration involved,
which results in lower temperatures and efficiency. However, with more than 20 years of
operating experience, continued technology improvements, and O&M cost reductions, troughs
are currently the most reliable and lowest risk solar thermal electric technology for near-term
deployment.
Linear Fresnel reflector (LFR) is conceptually similar to the parabolic trough, but instead of
using curved mirrors in an actual parabola, it uses an array of narrow, nearly flat mirrors
individually tilted and turned on their axes, forming a Fresnel approximation of a parabola to
reflect sunlight to the receiver. Using flat mirrors and water/steam as the HTF translates to lower
costs, but performance is expected to be inferior to trough, central receiver and dish/engine
technologies. Future plant designs will target operation at temperatures similar or greater than
synthetic oil-based trough systems, and molten salt HTF designs are under development.
LFR has gained more attention in recent years after AREVA Solar (formerly Ausra) deployed a
5-MW system in Bakersfield, CA in 2008 and Novatec Biosol commissioned a 30-MW plant in
Spain in 2012. LFR has also proven its value in solar thermal hybrid configurations with roughly
6 MW operating and 44 MW under construction in Australia. Large commercial standalone CSP
projects are under construction in Jordan and India.The more novel solar chimney is not yet as
well understood as the first four technologies above. Therefore, they are only briefly described in
this section, as no significant economic details are yet known.
8-5

Solar Thermal

8.1.2 Hybridization
All solar thermal technologies have the potential to operate as hybrids, with collected solar
energy introduced into a gas turbine or coal boiler system in concert with conventional fossil
fuels or other sources of heat. For example, a parabolic-trough solar field can be integrated with
a conventional combined-cycle plant to make up an integrated solar combined-cycle (ISCC)
system. Such hybridization can effectively increase the dispatchability of solar electricity and
thereby minimize grid integration issues. For retrofit applications, this approach is typically
limited to receiving about 10% to 20% of the total heat input from sunlight due to the limitation
of the existing equipment to accommodate larger amounts of solar. Central-receiver solar fields
are better suited to hybridization with natural gas combined-cycle and Rankine-cycle plants, and
can potentially accommodate higher levels of solar input due to the higher steam temperatures
that more closely match the existing steam turbine conditions. Eight solar thermal hybrid projects
representing 170 MW of solar capacity are now operational, and eleven more with 286 MW of
planned capacity are slated to come online in the next few years. The existing and planned
projects incorporate a range of solar technologies, fossil plants, and integration designs. The
majority of the projects use parabolic-trough technology, although a few have selected linear
Fresnel reflector, and a project in Turkey will use central-receiver technology.
As originally proposed by Luz Solar International, the ISCC configuration consists of a
parabolic-trough field, a combined-cycle power plant fueled by natural gas or other fossil fuel,
and a solar steam generator. During solar-powered operation, water from the cooler end of the
steam cycle is diverted and heated into saturated steam by the solar steam generator. The
saturated steam is then returned to the heat recovery steam generator (HRSG), where the
combined fossil- and solar-heated steam flow is then superheated by fossil-fueled heating. Other
integration options are feasible depending on the steam conditions available from the solar field.
When sunlight is unavailable, an integrated plant can operate as a conventional fossil-fueled
combined-cycle facility.
Advantages of solar thermal hybrid plants include the high efficiency with which they convert
solar energy to electrical energy compared to stand-alone solar plants, the relatively low
incremental unit cost for the larger steam turbine in an integrated plant compared to a solar-only
plant, and reduced thermal inefficiencies associated with the daily start up and shut down of the
steam turbine. Projects up to 75 MW in size are now operational in the United States, Algeria,
Egypt, Iran, Italy, Australia and Morocco. Other projects ranging from small demonstration units
up to 92.5 MW are being developed in Australia, Mexico, Spain, Chile, Turkey, Portugal,
France, China and Canada.

8-6

Solar Thermal

A novel hybridization option well-suited to central-receiver plants is one in which a pressurized


volumetric air receiver contributes heat energy to the compressed air of a gas turbine before it
enters the combustor. Additional heat input from conventional combustion is also required. A
receiver module, consisting of a secondary concentrator and a volumetric receiver unit, was
tested at Spains Plataforma Solar de Almeria, where air temperatures up to 1000C and receiver
efficiencies of more than 70% were achieved. This European Commission-funded projects
participants included Ormat in Israel, CIEMAT in Spain, DLR in Germany, Solucar in Spain,
and others. It was conducted in two stages, termed REFOS and SOLGATE.80 This concept
involves very high operating temperatures and the need to use window materials such as quartz
to contain pressurized air while admitting solar flux. This may limit the practical size of such
units to a few tens of megawatts. A few modular, air-based receiver concepts are under
development by industry.

8.2 Resources
This section of the Renewable Energy Technology Guide addresses solar resource assessment,
technology overview, and status for utility-scale solar thermal electric applications. It also
discusses likely development pathways, performance and cost history and projections, operating
and maintenance labor requirements, environmental issues, and potential for greenhouse gas
reduction.
The solar energy resource at a given location is characterized by the solar radiation per unit area
(or insolation) expressed in units of kilowatt-hours or megajoules per square meter per year
(kWh/m2/yr or MJ/m2/yr). The insolation reaching the Earths surface varies with latitude,
altitude, time of day and season, as well as with local weather and atmospheric conditions arising
from natural particulates or air pollution.
As sunlight passes through the atmosphere, some is reflected, some is absorbed, and some is
scattered. Because of these losses, the amount of energy that actually reaches the Earth never
exceeds about 70% of that present outside our atmosphere. Latitudes closer to the equator, and
especially regions with dry climates, typically exhibit the highest average insolation.
Insolation reaches a solar collector either directly (direct-normal radiation), after being scattered
(diffuse radiation), or after being reflected from the ground. These are important distinctions
because all solar thermal designs employ optical concentration and can use only direct-normal
solar radiation.

80

See www.ec.europa.eu/research/energy/pdf/solgate_en.pdf for the 2005 European Commission report Solgate:


Solar Hybrid Gas Turbine Electric Power System, which summarizes recent work in this area.

8-7

Solar Thermal

Figure 8-1 shows the distribution of direct-normal insolation in the world and United States. The
areas with the highest insolation include parts of the United States, Australia, Africa, and the
Middle East. In the United States, direct-normal insolation is most abundant in the Southwest,
and the largest regions of highest-quality resource are in Nevada, Arizona, California, and New
Mexico. Average insolation values for a given location, such as those shown in Figure 8-1, are
adequate for rough calculations, but to accurately predict the output and, therefore, the energy
cost of a solar thermal system, measurements should be taken over the course of at least one year
at a potential power plant location.
To assess the potential solar resource overall, it is important to know the geographic regions
where insolation is generally high, but the truly available resource is also constrained by land
topology, prior claims on it for other purposes, and other considerations. Solar thermal plants
cannot be easily installed in the mountains and they would be unwelcome in many places such as
national parks, military reservations, and urban areas. An NREL study used GIS screening
techniques to identify the land comprising the premium sites suitable for solar thermal power
plant deployment in the U.S. Southwest [11]. The study began with all land having an average
insolation greater than 6.75 kWh/m2 and filtered out unsuitable areas using the following criteria:
military bases with a one-mile buffer; national wilderness areas with a five-mile buffer; Fish and
Wildlife Service land with a one-mile buffer; National Park Service land with a five-mile buffer;
National Forest Service land; cropland; major highways with a half-mile buffer; navigable
waterways with a half-mile buffer; lakes with a two-mile buffer; major urbanized areas with a
four-mile buffer; railroads with a 500-foot buffer; and locations 9,000 feet above sea level with a
4.5-mile buffer around each plant. The study concluded that using only 3% of the land area that
remained after filtering could provide over 1 million GWh of solar thermal-generated electricity
annually, representing about 25% of net U.S. electricity generation in 2008. The recent SunShot
Solar Vision Study published by the DOE outlined the feasibility of achieving 10% and 20% of
power generation from solar (PV and solar thermal) in the U.S. by 2030.

8-8

Solar Thermal

Figure 8-1
Distribution of direct-normal insolation worldwide (top, kWh/m2/yr) and U.S.
(bottom, Wh/m2/day)
Sources: U.S. DOE and ISET

8-9

Solar Thermal

8.3 Technology Description


This section describes the general function and operating characteristics of the most significant
solar thermal technologies: parabolic trough, central receiver, dish/engine, linear Fresnel
reflector, and solar chimney. These descriptions focus on the history and engineering
technologies themselves. More information on current and future projects is provided in Section
8.4, Technology Status.
8.3.1 Parabolic Trough
Parabolic trough technology is currently the most proven solar thermal electric technology. To
date, the parabolic trough is the only solar thermal technology that has achieved commercial
operation at significant scale. Some 354 MW of trough solar-electric systems, hybridized to use
natural-gas firing for up to 25% of the necessary heat input, have been operating in southern
California since 1990. An additional 1-MW trough system was commissioned in Arizona in
2006, a 64-MW Nevada facility began operating in 2007, a 75-MW hybrid project began
operating in 2010 and thirty-two 50-MW units have begun operating in Spain as of December
2011. Overall, approximately 2,270 MW of parabolic trough capacity exists worldwide.
A parabolic trough collector field consists of several thousand single-axis tracking solar
collectors, each with an area of 100 m2 to 500 m2 (see Figures 8-2 and 8-3). The solar field is
highly modular, comprising many parallel rows of solar collectors aligned on a north-south axis.
Each collector has a linear parabolic-shaped reflector that focuses the suns direct-beam radiation
onto a linear receiver located at the focus of the parabola. The collectors track the sun from east
to west during the day to ensure that sunlight is continuously focused on the receiver.
The sun heats the HTF from about 560F (290C) to 735F (390C) as it circulates through the
receiver tube. The HTF then returns to a series of heat exchangers in the power block, where it
is used to generate high-pressure superheated steam at 1,450 psia (10 MPa) and 700F (370C).
The superheated steam is then fed to a conventional reheat steam turbine/generator to produce
electricity. Spent steam from the turbine is condensed in a standard condenser and returned
to the heat exchangers via condensate and feedwater pumps to be transformed back into steam.
Condenser cooling is typically provided by mechanical draft wet cooling towers, although dry
cooling is an option. After passing through the HTF side of the solar heat exchangers, the cooled
HTF is recirculated through the solar field.
Historically, parabolic trough plants use solar energy as the primary energy source to produce
electricity. Given sufficient solar input, such plants can operate at full rated power using solar
energy alone. During summer months, the plants typically operate for 10 to 12 hours a day at full
rated electric output. However, all plants built to date have been solar/fossil hybrids, meaning
that they have a backup fossil-fired capability that can be used to supplement solar output during
periods of low solar radiation. In light of the relatively high heat rates of the smaller turbines at
these plants, actual gas usage has been lower than the historical and contractually allowed 25%
rate. Thermal storage is another strategy that can provide dispatchability without a fossil-fuel
backup system.

8-10

Solar Thermal

Figure 8-2
Solar trough collector field at Kramer Junction, California (Note the receiver tube at the
reflectors focal point that transports heat-transfer fluid.)

Figure 8-3
Andasol 1 and 2 parabolic trough plants, Spain
Source: Solar Millennium

8-11

Solar Thermal

In the United States, the solar collector field typically provides maximum solar output in late
spring or summer. The gas-fired heater is typically used to extend operation to late afternoon and
early evening hours during the summer peak period and to augment the solar collector field
during cloudy periods. Although the gas-fired heater is physically capable of more extended use,
power plants operated as qualifying facilities (QFs) under PURPA are limited to generating 25%
of total energy from gas. Utility-owned plants are not subject to the same limitation on gas use at
the present time. The California plants principals formerly included Sunray Energy, the Kramer
Junction Co., KJC Operating Co., and FPL Energy, but now FPL Energy is the principal owner.
The plants benefit from federal and state investment incentives, solar property tax exclusion, and
accelerated depreciation intended to encourage solar projects.
8.3.2 Central Receiver
Central receiver plants use hundreds or thousands of sun-tracking mirrors called heliostats to
focus concentrated solar radiation on a tower-mounted heat exchanger (receiver). Because they
perform better at larger sizes, these plants are best suited for utility-scale applications of many
tens to a couple hundred of megawatts.
The heliostat field surrounding the tower is laid out to optimize either the annual or seasonal
performance of the plant. In a typical installation, solar energy collection occurs at a rate that
exceeds the maximum required to provide steam to the turbine. Consequently, a thermal storage
system can be charged at the same time that the plant is producing power at full capacity.
Without energy storage or fossil-fuel hybridization, solar technologies are limited to annual
capacity factors of less than approximately 30%. However, with sufficient storage capability, a
central receiver could potentially operate all of the time without a backup fuel source. The Solar
Two and Gemasolar projects have demonstrated such 24-hour operation.
Although central receivers are commercially less mature than parabolic trough systems, several
component and experimental systems have been field tested around the world in the past 25
years to demonstrate the engineering feasibility and economic potential of the technology (Table
8-5). These experimental facilities were built to show that central receivers can produce
electricity and to test and improve individual system components.
The first large-scale test of the central receiver system was the 10-MWe Solar One project at
Daggett, California, which operated from 1982 to 1988 and was then the worlds largest central
receiver plant. It proved that large-scale central receiver power production was feasible. In that
plant, water was converted to steam in the receiver and used directly to power a conventional
Rankine-cycle steam turbine. The thermal storage system stored heat from solar-produced steam
in a thermocline tank filled with rocks and sand, using oil as the heat-transfer fluid. The storage
system extended the plants power-generation capability into the night and provided heat for
generating low-grade steam for keeping parts of the plant warm during off-hours and for
morning startup. Unfortunately, the storage system was complex and thermodynamically
inefficient. Although Solar One successfully demonstrated central receiver technology, it also
revealed the disadvantages of a water/steam system, such as intermittent turbine operation due
to short-term cloud cover and lack of effective thermal storage.

8-12

Solar Thermal
Table 8-5
Central receiver demonstration projects
Country

Power
Output
(MWe)

Heat Transfer
Fluid

Storage Medium

Operation
Began

Spain

0.5

Liquid Sodium

Sodium

1981

EURELIOS

Italy

Steam

Nitrate Salt/Water

1981

SUNSHINE

Japan

Steam

Nitrate Salt/Water

1981

Solar One

USA

10

Steam

Oil/Rock

1982

CESA-1

Spain

Steam

Nitrate Salt

1983

MSEE/Cat B

USA

Molten Nitrate

Nitrate Salt

1984

THEMIS

France

2.5

Hi-Tec Salt

Hi-Tec Salt

1984

SPP-5

Russia

Steam

Water/Steam

1986

TSA

Spain

Air

Ceramic

1993

Solar Two

USA

10

Molten Nitrate Salt

Nitrate Salt

1996

PS10

Spain

11

Steam

Water/Steam

2007

Solar Energy
Development Center

Israel

Steam

na

2008

PS20

Spain

20

Steam

Water/Steam

2009

Eureka

Spain

Steam

Water/Steam

2009

Germany

1.5

Air

Air

2009

Spain

20

Molten Nitrate Salt

Nitrate Salt

2011

Project
SSPS

Jlich
Gemasolar
(aka Solar Tres)

During the operation of Solar One, research began on a more advanced molten-salt central
receiver design that would store energy more efficiently. This led to the Solar Two project
(Figure 8-4). In the Solar Two power tower, liquid salt at 550F (290C) was pumped from a
cold storage tank through the receiver, where it was heated to 1,050F (570C), then on to a
hot tank for storage. When power was needed from the plant, hot salt was pumped to a steam
generating system that produced superheated steam for a conventional Rankine-cycle
turbine/generator system. From the steam generator, the salt was returned to the cold tank, where
it was stored and eventually reheated in the receiver. Determining the optimum storage size to
meet power-dispatch requirements was an important part of the system design process.

8-13

Solar Thermal

Figure 8-4
Solar Two central receiver in operation (The two tanks to the left of the tower base are the
storage tanks for the hot and cold molten salt solutions.)

The Solar Two storage medium was a 60/40 mixture of sodium nitrate and potassium nitrate,
which melts at 428F (220C). It was maintained in a molten state at 554F (290C) in the cold
storage tank. Molten salt can be difficult to contain and transport because it has a low viscosity
(similar to water) and it wets metal surfaces extremely well. An important consideration in
successfully implementing this technology is identifying pumps, valves, valve packing, and
gasket materials that will work with molten salt. Accordingly, Solar Two was designed with a
minimum number of gasketed flanges, while most instrument transducers, valves, and fittings
were welded in place. Designers specified stainless steel for all pipes, valves, and vessels
because of its corrosion resistance in the molten-salt environment. In addition, the design must
allow the receiver to rapidly change temperature during cloud passage, for example, without
being damaged. The Solar Two receiver could safely change from 554F to 1058F (290C to
570C) in less than one minute.
The developers of Solar Two envisioned the project primarily as a demonstration of a complete
molten salt central receiver power plant, albeit at a less than commercial scale. In reality, the
program was more a development effort than a demonstration effort. Operators used many of the
programs resources to improve system components and devise solutions for problems that arose.
The result is that Solar Two was quite successful as a development program. Much was learned
and documented that should substantially reduce the technical risks associated with future molten
salt central receiver plants.

8-14

Solar Thermal

Solar Twos major contributions toward resolving key issues included the following:

It clearly demonstrated that a solar central receiver can operate as a dispatchable power plant
as part of a utility generating mix. On several occasions the plant operated for many
consecutive days, providing power to the grid during peak load periods even after sundown,
and in some cases even through the night. During one such period, the plant operated
continuously for 154 hours, although at reduced power because of limited thermal storage
capacity.

Although trace heating was a major unresolved concern at the conclusion of earlier central
receiver projects, substantial progress toward resolving this issue was made at Solar Two.
Trace heating is needed to prevent the molten salt from solidifying while the plant is not
operating. After the total rework of the original trace heating system during 1996, the system
operated with good reliability through the end of the project. Effective trace heating in the
vicinity of bends and unusual shapes such as those present in the receiver piping is still an
issue of some concern, but reasonable solutions have been proposed that can be employed in
future plants.

Except for the heliostat field, which had known deficiencies, Solar Two actually achieved
design-point efficiencies for all its major subsystems. Of particular importance were the
receiver and thermal storage performance parameters. The receiver demonstrated efficiencies
in the range of 86% to 88%, and the effective efficiency of the storage subsystem was 99%
after taking into account that some of the collected thermal energy is transported quickly to
the steam generator. Even if all of the energy were to reside in thermal storage for a 24-hour
period before transport to the steam generator, efficiency would have been about 97%.

Storage performance is highly significant. That subsystem worked under normal operating
conditions for essentially three years, and was particularly well characterized. It clearly
demonstrated the ability to store energy for several days with minimal losses. In addition,
post-experiment examination of the tanks (carbon-steel for the low-temperature tank and
stainless steel for the high-temperature tank) showed little evidence of corrosion, which
suggests not only excellent performance but also long service life for molten-salt storage
systems. No other renewable energy generation option can offer storage with performance
approaching this level. In addition, the incremental cost of this storage is relatively low, by
some estimates, approximately 10% of the total plant cost assuming five hours of storage.

Project funding limitations, coupled with concerns about the condition of the heliostat field,
prevented the operation of Solar Two for another year or two. This is unfortunate, because a
number of equipment and operating strategy issues had been resolved during the project, and
it may have been possible to obtain operating experience under conditions characteristic of
commercial operation. Data on annual operating efficiencies obtained under such conditions
could have been adjusted to account for known deficiencies in the plant such as turbine
generator limitations and heliostat field condition.

8-15

Solar Thermal

In addition to annual energy production potential, several other key technical issues remain
unresolved and cause risk levels to remain high. These include the following:

Long thermal time constants still result in limitations on useful energy collection. Morning
warmup periods for the thermal collection and electricity generation subsystems are still
measured in hours. Warmup periods might be reduced by improving warming strategies
using stored thermal energy and by increasing plant capacity factors by increasing field size
and storage capacity. The effectiveness of such measures remains to be demonstrated in
future plants.

Lifetime and maintenance requirements for molten-salt valves are still uncertain. The Solar
Two evaluation team recommended redesigning the salt loop to minimize the use of valves.
Another recommendation for future plants was to use pumps with long shafts so that they can
be placed on top of the salt tanks.

Acceptable parasitic power consumption needs to be demonstrated. Substantial related


progress was made with Solar Two, but long-term power production is needed to resolve this
concern.

System startup in high winds is still an issue. As with earlier central receiver field tests, the
inability to operate during high winds was a significant cause of lost energy at Solar Two.

Heliostat field availability at Solar Two was less than expected. In contrast, availability for
Solar One was very high. While inattention to the entire plant facility between the two
projects was likely a major cause for the drop in availability, the issue of long-term heliostat
life is still open. It also appears that earlier high hopes for potentially low-cost stretched
membrane heliostats have diminished and that, for the foreseeable future, plants will rely on
glass-metal designs. This has implications for commercial plant costs because the glass-metal
designs are more expensive. On the other hand, experience with trough systems demonstrates
that glass reflectors perform very well under real-world operating conditions.

The 11-MW PS10 plant near Seville, Spain was the worlds largest operating central receiver
plant in 2008. The system operates at saturated steam conditions of 482F (250C). After
roughly a year and a half of commercial operation, the plant owners reported that the
performance had improved by 10% due to optimization of the temperature gradient across the
receiver surface and other operational improvements. The 20-MW PS20 plant sited next to PS10
came online in 2009. The larger tower was designed with lower auxiliary loads to boost
performance. Eureka, a 2-MW high-temperature demonstration tower, is also operational at the
same site to validate superheated steam operation. In May 2011 the 19.9-MW Gemasolar project
was commissioned in Spain, and several larger-scale projects are scheduled to come online
between 2012-2015. SolarReserve is developing 150-MW, 110-MW, 150-MW and 50-MW
plants in California, Nevada, Arizona and Spain, respectively. BrightSource has approximately
900 MW of projects in California and 400 MW in Nevada to built in the same timeframe, and a
pipline of an additional 1800 MW to be built in California by 2020.

8-16

Solar Thermal

8.3.3 Dish/Engine
Dish/engine systems employ a parabolic reflector to focus concentrated sunlight onto a receiver
located at the focal point of the dish (Figure 8-5). The sunlight heats a working fluid in the
receiver tube, which transfers heat to a small engine used to generate electricity. Many
thermodynamic cycles and working fluids have been considered for dish/engine systems, but the
Stirling and open Brayton (gas turbine) cycles are generally favored. Conventional automotive
Otto and Diesel engine cycles are not feasible in this application because of the difficulties in
integrating them with concentrated solar energy.

Figure 8-5
25-kW SAIC dish/engine system at the DOE Mesa Top Thermal Test Facility
Source: NREL

The key components that affect the cost and performance of dish/engine systems are the
concentrator, the receiver, and the engine itself. To achieve the high temperatures required to
efficiently convert heat to work, the dish must precisely track the sun on two axes.
Dish/engine systems in operation have typically been scattered demonstration units, including
several installed at Sandia National Laboratories Solar Thermal Test Facility. In January 2010,
SES and Tessera Solar commissioned the 1.5-MW Maricopa Solar Plant, composed of 60 25-kW
SunCatcher units, in Peoria, AZ. However, SES declared bankruptcy in 2011 and the Maricopa
project assets were auctioned in April 2012.
Infinia has thus far survived the solar industry shake-out. Under new management since 2011
and with a recent $15.5M VC investment in October 2012, the company is developing a 1.5-MW
installation (430 units) at the Tooele Army Depot outside Salt Lake City, UT.

8-17

Solar Thermal

8.3.3.1 Concentrators
The size of a dish/engine systems solar concentrator is determined by the heat input
requirements of the engine. At a nominal maximum direct-normal solar insolation of 1,000
W/m2, a 25-kW dish/Stirling system has a concentrator diameter of approximately 10 meters.
Infinias 3.5-kW PowerDish81 design has a 6.7-meter diameter dish. Concentrators use a
reflective surface of aluminum or silver deposited on glass or plastic. The most durable reflective
surfaces have been silver/glass mirrors. Because dish concentrators have short focal lengths,
relatively thin glass mirrors approximately 1 mm are required to accommodate the required
curvatures. In addition, for rear-surface-reflector mirrors, glass with low iron content helps to
reduce absorbance losses in the glass. Depending on the glass thickness and iron content,
silvered mirrors have solar reflectance values in the range of 90% to 94%. Although glass
mirrors achieve high performance and the technology is well-understood, they drive up the cost
of the concentrator. For this reason, ongoing work has centered on developing low-cost reflective
elements, such as those based on polymer films and stainless-steel membranes.
Typically, concentrators approximate a paraboloid of rotation by using an array made up of
multiple mirrors supported by a truss structure. One innovative solar-concentrator design used
stretched membranes, in which a thin reflective membrane is stretched across a rim or hoop. A
second membrane is used to close off the space behind. A partial vacuum is drawn in this space,
bringing the reflective membrane into an approximately spherical shape. This design proved
problematic in operation, however, and is not being currently pursued.
A concentrators optical design and accuracy determine the concentration ratio, defined as the
average solar flux through the receiver aperture divided by the ambient direct-normal solar
insolation. Typical concentration ratios are over 2,000. The intercept fraction, defined as the
fraction of the reflected solar flux that passes through the receiver aperture, is usually greater
than 95%.
The receiver absorbs energy reflected by the concentrator and transfers it to the engines working
fluid. The absorbing surface is usually placed behind the focus of the concentrator to reduce the
flux intensity incident on it. Stirling and Brayton engines each have their own interface issues.
Stirling engine receivers must efficiently transfer concentrated solar energy to a high-pressure
oscillating gas, usually helium or hydrogen. In Brayton receivers, the flow must be steady and at
relatively low pressure.
There are two general types of Stirling receivers: direct-illumination receivers (DIRs) and
indirect receivers which use an intermediate heat-transfer fluid. DIRs adapt the heater tubes of
the Stirling engine to absorb the concentrated solar flux. Because of the high heat-transfer
capability of high-velocity, high-pressure helium or hydrogen, DIRs are capable of absorbing
high levels of solar flux (approximately 75 W/cm2). However, balancing the temperatures and
heat addition between the cylinders of a multiple-cylinder Stirling engine remains an integration
issue.

81

PowerDish is a trademark of Infinia.

8-18

Solar Thermal

Liquid-metal heat-pipe solar receivers may help to solve this problem. In a heat-pipe receiver,
liquid sodium metal is vaporized on the absorber surface of the receiver and condensed on the
Stirling engines heater tubes. This results in a uniform heater-tube temperature, thereby enabling
a higher engine working temperature for a given material, and therefore higher engine efficiency.
Although not yet demonstrated, longer-life receivers and engine heater heads are also
theoretically possible through the use of a heat pipe. Stirling receivers are typically about 90%
efficient in transferring energy delivered by the concentrator to the engine.
Solar receivers for dish/Brayton systems are less developed. In addition, the poor heat-transfer
coefficients of relatively low-pressure air, combined with the need to minimize pressure drops in
the receiver, make receiver design challenging.
8.3.3.2 Engines
The engine in a dish/engine system converts heat to mechanical power in a manner similar to
conventional engines by compressing a working fluid when it is cold, heating the compressed
working fluid, and then expanding it through a turbine or with a piston to produce work. The
mechanical power is converted to electric power by an electric generator or alternator. Electric
output in current dish/engine prototypes using a Stirling engine is about 25 kWe.
The Stirling engine is a leading candidate for dish/engine systems because of its high efficiency
and because its characteristic external heating makes it adaptable to concentrated solar flux. The
Stirling engines used in solar dish/Stirling systems use hydrogen or helium working gas and
operate at high temperatures and pressures, up to 700C (1,300F) and 20 MPa (2,900 psia),
respectively. In the Stirling cycle, a working gas is alternately heated and cooled by constanttemperature and constant-volume processes. Stirling engines usually incorporate an efficiencyenhancing regenerator that captures heat during constant-volume cooling and replaces it when
the gas is heated at constant volume. There are a number of mechanical configurations that
implement these constant-temperature and constant-volume processes. Most involve the use of
pistons and cylinders. Some use a displacer (a piston that displaces the working gas without
changing its volume) to shuttle the working gas back and forth from the hot region to the cold
region of the engine. For most engine designs, power is extracted by a rotating crankshaft. An
exception is the free-piston configuration, where the pistons are not constrained by crankshafts or
other mechanisms. They bounce back and forth on gas springs and the power is extracted from
the power piston by a linear alternator or pump. The best Stirling engines today achieve thermalto-electric conversion efficiencies of about 40%.
The Brayton engine, also called the jet engine, combustion turbine, or gas turbine, is an internalcombustion engine similar to an Otto-or Diesel-cycle engine in that air is compressed, fuel is
added, and the mixture is burned. In the Brayton cycle, however, the hot exhaust gases expand
through a turbine, rather than against a piston. In a dish/Brayton system, solar heat is used to
replace (or supplement) the fuel. As in the Stirling engine, recuperation of waste heat is
a key to achieving high efficiency. Therefore, waste heat exhausted from the turbine is used to
preheat air from the compressor.

8-19

Solar Thermal

8.3.4 Linear Fresnel Reflector


Linear Fresnel reflector (LFR) technology is designed to reduce capital costs compared to
parabolic trough and central receiver systems. The unresolved question is whether the capital
cost is sufficiently low to compensate for the lesser performance of LFR systems.
LFR technology is conceptually similar to the parabolic trough, except instead of using curved
mirrors it uses a field of nearly flat mirrors individually tilted and turned on their axes to reflect
sunlight to the receiver (Figure 8-6). It has some similarities to central receiver technology as
well, in that its reflectors are separately mounted from a stationary receiver.
A complete LFR system uses arrays of mirror strips in a large field with several receivers; each
individual mirrored reflector has the option of directing reflected solar radiation to at least two
different receivers. This minimizes shading losses, allows arrays to be much more densely
packed, and permits the receiver tubes to be lower than would otherwise be possible. The
Fresnel in LFR refers to the optical arrangement of reflectors in rows. Whereas a conventional
parabolic trough solar thermal system has one curved reflector for each receiver line, the LFR
system typically has 10.

Figure 8-6
LFR Array at Liddell Power Station, Australia
Source: SolarPACES

The tower-mounted linear receivers consist of a series of steel tubes surrounded by a trapezoidal
reflective surface. Nominal sun concentration is less than 80 suns. Within the receiver,
pressurized water is converted to either saturated or superheated steam conditions (up to
approximately 480C has been demonstrated), and the steam from this process drives
conventional Rankine cycle steam turbines and generators. AREVA Solar is currently
constructing a 44-MW solar field to augment the Kogan Creek supercritical coal plant in
Queensland, Australia. Separately, it is building a 250-MW stand-alone LFR solar project in
Rajasthan, India to be completed in phases from 2013 to 2015.

8-20

Solar Thermal

8.3.5 Solar Chimney


A solar chimney generates electricity when air heated by the sun rises through a very tall
convection tower, thereby driving wind turbines (Figure 8-7). The air is heated in a large
greenhouse-like structure surrounding the base of the chimney. Turbines could be installed in a
ring around the base of the tower or within the chimney itself.
The generating capacity of a solar chimney depends on the size of the collector area and the
chimney height. Solar chimneys have been proposed that are as tall as a kilometer (0.6 mi) with
7-km (4.4-mi) collector diameter and would generate up to 200 MW.
The German government funded a research prototype chimney in Spain in 1982 that was 195
meters (640 feet) high, 10 meters (33 feet) in diameter. The plant produced a maximum electrical
output of about 50 kW and operated successfully until its decommissioning in 1989.

Figure 8-7
Cross-section diagram of solar chimney (Sunlight heats trapped air that then rises through
the cylindrical tower. Water tubes inside the greenhouse-like structure surrounding the
chimney would help retain heat.)
Source: C. Pietschiny

Architect Christian Pietschiny has suggested a novel strategy for gaining broader support and
economic benefit from a solar chimney: make it the worlds largest work of art. Pietschiny
proposes that a solar chimney presented as an enormous public expression of aesthetics, ecology,
and economy could attract attention, investors, and even tourists who would not otherwise take
an interest in an electricity generation project. Figure 8-8 is a painting showing one approach to
this vision. In October 2010 the Southern California Public Power Authority approved a power
purchase agreement to buy the output of a 200-MW solar chimney to be built in Arizona by
EnviroMission of Australia. If completed, the $700 million plant, with a 3-mile (4.8-km)
diameter greenhouse and 800-plus meter-tall chimney containing 35 wind turbines, would
become the worlds second tallest man-made structure. A 27-MW solar chimney is reportedly
under development in China as well.
8-21

Solar Thermal

Figure 8-8
Solar chimney sculpture
Source: C. Pietschiny

8.4 Technology Status


After decades during which solar thermal technologies were hindered by an underdeveloped
industrial base, opportunities in the field are attracting investment and growth. Policies and
public funding to stimulate solar energy have successfully drawn the attention of industrial
players such as General Electric, AREVA, Alcoa, and Rio Glass.
Progress in the 1980s, most evident with parabolic trough systems, was largely fueled by public
policies that encouraged deployment of solar thermal and other renewable technologies. After
many fallow years, similar strong legislative backing has returned. In the United States, ARRA
commitments are funding solar R&D through the DOE and its national laboratories and backing
the DOE loan guarantee program for clean energy projects. In Spain and Germany, interest
among industrial players has grown in recent years. China and India have made politically
backed overtures into solar technologies, notably the Indian Solar Mission and programs to
subsidize solar power in China. Recent CSP initiative announcements in South Africa, Morocco,
and Saudi Arabia indicate that solar thermal deployment regions are becoming increasingly
diverse. Assuming that this trend continues, the extensive existing industrial infrastructure can
quickly become engaged in solar thermal technology and business development, as most of the
equipment in solar thermal plants consists of glass and metal formed by routine industrial
processes.
Table 8-6 is a technology monitoring guide of the leading developers and technical issues in
solar thermal power generation. Table 8-7 is a technology process development map for solar
thermal power generation.

8-22

Solar Thermal
Table 8-6
Technology monitoring guide for solar thermal power plants
Leading Developers of the Science or Technology
R&D
Intensity
Moderate

Government
Nonprofit
Developers and
Organizations Organizations
Vendors
FlabegSolar
DOE (Sandia, DLR
Acciona Solar
(Germany)
NREL),
Abengoa/Solucar
CIEMAT
SENER
(Spain)
Cobra
Schott Solar
SkyFuel
Iberdrola
Flagsol
TSK

Central
Receiver

Moderate

DOE (Sandia,
NREL),
CIEMAT

Parabolic
Dish/Stirling
Engine

Low

DOE (Sandia),
CIEMAT

Linear
Fresnel
Reflector

Low

DOE,
CIEMAT

Technologies
Parabolic
Trough

DLR

Changes to
Watch for
Direct steam
Lower-cost
generation, moltenmirror support
salt or gas HTFs,
structures,
thermal storage, non-glass reflective
film collectors,
improvedsingle-vessel
performance
storage, lower-cost
receiver tubes,
storage strategies,
larger plants,
improved receiver
hybrid plants
tube designs
Major Trends

Nexant (Bechtel)
Solar Reserve/USRG
(UTC)/Rocketdyne
Abengoa/Solucar
SENER
BrightSource
eSolar

Molten-salt
receiver,
superheatedsteam receiver,
air receiver,
smaller/cheaper
heliostat
designs

Schlaich-Bergermann
and Partner (SBP)
Infinia
HelioFocus
AREVA Solar (Ausra)
TSK

Lower-cost dish
and mirror
structures
Early demos,
hybrid
operation, proof
of economics

Thermal storage,
decoupling energy
collection from
electricity
generation,
multiple distributed
towers vs. single
tower,
supercritical-CO2
Brayton engines
Durable free-piston
engine, phasechange thermal
storage
Scale-up to 100megawatt scale,
stand-alone
operation

Unresolved
Issues
Steam or gas
flow control,
cost reduction
potential of
reflective film
collectors, freeze
protection of
molten-salt HTF
in collector field,
operation of
thermal-gradient
storage tank
Energy
production,
high temperature
operation, cost

Engine
availability, O&M
costs, cycling
impacts
Cost, efficiency,
proven
performance,
long-term O&M
cost.

8-23

Solar Thermal
Table 8-7
Technology process development map: solar thermal power plants
Parabolic
Trough/Gas
Hybrid

Parabolic
Trough with
Storage

Molten Salt
Central
Receiver

Direct Steam
Central
Receiver

Dish/Stirling
Engine

Features and Advantages


Technology
characterization

Parabolic
trough
Heat transfer
from oil to
steam cycle

Efficiency (net)1

Parabolic trough

Heliostat field

Heliostat field

Parabolic dish

Heat transfer
from oil to
molten salt to
steam cycle

Heat transfer
from molten
salt to steam
cycle

Direct steam
cycle
(saturated or
superheated)

Heated gas
(helium/hydrogen)
powers Stirling
engine

1517%

1314%

1520%

1522%

1630%

Flat terrain with


high annual
direct insolation
and some
natural gas
availability
(depending on
amount of
hybridization)

Flat terrain with


high annual
direct insolation
and some
natural gas
availability (for
freeze
protection or
backup)

Flat terrain with


high annual
direct insolation
and some
natural gas
availability (for
freeze
protection or
backup)

Relatively flat
terrain with
high annual
direct insolation
and some
natural gas
availability (for
backup)

High annual direct


insolation

Capital and
O&M cost

O&M cost

Capital cost

Capital and O&M


cost

Optimum control
of storage
subsystem, high
freezing point of
salt

High freezing
point of salt

Steam cycle
control, high
parasitic losses

Uncertain
maintenance
costs

Direct steam
generation
or other hightemperature
HTF,
inexpensive
reflectors with
resistance
to soiling

Direct steam
generation or
other hightemperature
HTF,
inexpensive
reflectors with
resistance to
soiling, lower
cost storage

Reliable, costeffective trace


heating
capability

Steam control
system, energy
storage,
parasitic loss
reduction

Need for
highly reliable
reflectors,
trackers, and
engines

Relative Capital

Medium

High

Medium

Uncertain

Uncertain

Relative O&M

Medium

High

High

Uncertain

Uncertain

Busbar2

From 13/kWh

From 13/kWh

From 11/kWh

Uncertain

Uncertain

Resources

Challenges and Disadvantages


Cost

Capital cost

Operation

Technology
Needs

Costs

Expected values for new plants

Does not reflect potentially high financing costs for first-of-a-kind systems or subsidies

8-24

Solar Thermal
Table 8-7 (continued)
Technology process development map: solar thermal power plants
Parabolic
Trough/Gas
Hybrid

Parabolic
Trough with
Storage

Molten Salt
Central
Receiver

Direct Steam
Central
Receiver

Dish Stirling
Engine

Development Timeframe
Research

Late 1970s

Current

1980s

Late 1970s

1970s1980s

Development

1980s

Current

1980s

Early 1980s

1980s1990s

Demonstration

19861987

Current

19931998

19841987

19831990s

Commercialization

1988

2009

2011

2009

2010

Examples
operating
successfully

More than a
dozen
operating in
Spain

High first
cost hurts
market
prospects

High first cost


hurts market
prospects

Dispatchability
helps market
prospects, but
temperature
limitations may
make it
uncompetitive
relative to other
CSP options

Limited
competition
for moltensalt
equipment;
strong
competition
for other key
components

Successful
first year of
operation at
Gemasolar
20-MW plant

Markets
Key Issues

High first cost


hurts market
prospects

Several 100150 MW-scale


plants to
begin
operating in
2013

Limited
competition
for Stirling
engine and
parabolic dish
collectors

Several 50150 MWscale plants


to begin
operating in
2013
Impact on
Technology Outlook

Recent interest,
driven in part
by RPS in CA,
NV, AZ, NM,
CO, TX, and
Spanish solar
incentive may
spur very largescale
development in
next 510
years
Competition
from PV has
slowed
momentum

Recent
interest, driven
in part by RPS
and Spanish
solar incentive
may spur very
large-scale
development in
next 510
years
Competition
from PV has
slowed
momentum, but
storage
capability is
differentiator

Interest
initially
lagged
trough and
direct steam
tower due to
high cost
and need for
relatively
large system
sizes
(>100 MW).
Pricing
below trough
options and
inherent
storage
capability is
differentiator

PS10 and
PS20 designs
appear to
be limited
to very high
feed-in tariff
market
Unclear if
Brightsource
thermal
energy
storage option
will be
competitive
with direct
molten salt
systems

Infinia has
begun
megawattscale
deployments
and
prospects for
learning more
about largescale dish
operation still
appear bright
Addition of
thermal
storage may
be key

8-25

Solar Thermal

8.4.1 Parabolic Trough


In 1984, Luz International installed Solar Electric Generating System I (SEGS I) in Daggett,
California. It originally had an electric capacity of 13.8 MW, incorporated six hours of thermal
storage, and used natural-gas-fueled superheaters to supplement the solar energy. The storage
system was taken out of service after a fire in 1999, but the rest of the plant is still operating. Luz
also constructed additional plants, SEGS II through VII, each with 30-MW capacity. In 1990,
Luz completed construction of SEGS VIII and IX in Harper Lake, each with 80-MW rated
capacity. The total generating capacity of these plants is 354 MW; individually and combined,
they are the largest solar power plants in the world. From the mid-1990s until about 2000, they
had generated more solar-electric kilowatt-hours than the total of all other worldwide sources.
As a result of regulatory and policy obstacles, as well as marginal economics without ongoing
tax credits, Luz International and four subsidiaries filed for bankruptcy on November 25, 1991.
Several companies assumed operation and maintenance of the facilities as independent power
generators. As of February 2005, FPL Energy held a significant minority ownership in all but
SEGS I and II, making it the largest holder of solar-electric generation facilities. In its later
years, Luz had begun developing direct steam generating solar collectors and integrating the
solar collector field into a gas-fired combined cycle plant. Other firms that continue involvement
in this technology include Acciona Solar Power, Abengoa/Solucar, Schott Solar, SENER, Cobra,
Iberdrola, SkyFuel, Flagsol (a subsidiary of Ferrostaal), TSK, and FlabegSolar. Specifically in
the field of mirrors, Rio Glass and Saint Gobain are very active. In December 2002, Sierra
Pacific Resources of Las Vegas announced plans for its two Nevada utility subsidiaries, Nevada
Power and Sierra Pacific Power, to buy the output of what later became a 64-MW parabolic
trough plant near Boulder City, Nevada. The project was built by Acciona Solar Power and is
part of the utilitys strategy to comply with the states renewable portfolio standard.
Groundbreaking for the plants construction occurred in February 2006 and the plant was
commissioned in June 2007.
The Spanish governments alternative energy program has been testing trough systems at its
solar test facility in Almera. The version of the Spanish solar feed-in tariff begun in 2007
provides two options for payments to solar plants up to 50-MW capacity: a flat 269/MWh or a
premium over market of 254/MWh, capped at 344/MWh total price. This offer was valid for
up to 500 MW of solar generation and would-be solar-plant builders applied for permits totaling
perhaps twice that amount. In the beginning in 2013, Spain will halt new feed-in tariff contracts
for renewable energy, citing the electric system deficit the amount the system is paying out via
the tariff program has exceeded the income. In addition, Spain's government will enact a 7% tax
to incomes of power producer, whatever they use renewable sources or not. Particularly concerning
to concentrated solar power producers, is a provision within the new law that reduces the amount
they receive through FIT in a percentage according to the use of natural gas.
Natural gas is used in the CSP plants to increase performance during start-up times, keep the heat
transfer fluid above a minimum temperature, and as a backup fuel.
The first to be finished was Andasol 1, a 50-MW plant with over seven hours of storage. It began
commercial operation in early 2009. Thirty-one additional 50-MW trough plants are now
operational.

8-26

Solar Thermal

The World Banks Global Environment Facility (GEF) has pursued government and commercial
interests in building the next wave of solar thermal trough power plants. The builders envision
these installations as fossil-hybrid combined-cycle plants, with the solar contribution limited to
10% to 20%. Egypt built a hybrid solar fossil thermal power plant that generates 140 MW, 20
MW of which is provided by solar power. The Moroccan national utility, ONE, built an
integrated solar combined cycle (ISCC) plant with a net output of 470 MW, 20-MW of which is
solar trough. Separate from the GEF, Algeria developed a 150-MW total, 25-MW solar plant and
GEF studies for similar systems are under way in Mexico.
In the United States, there is one operating ISCC plant and one more planned. The first to be
commissioned was FPLs $476 million Martin Solar Energy Center (Figure 8-9), which added
75-MW of solar capacity to an existing combined-cycle plant. It was completed at the end of
2010. Another project, the Palmdale (Calif.) Hybrid Plant, is conceived as a city-owned 570-MW
combined-cycle plant with about 50 MW of solar capacity. As of August 2011, the project was
approved by California Energy Commission staff.

Figure 8-9
Martin Next-Generation Solar Energy Center
Source: FPL

Roughly 4,700 MW of additional parabolic trough systems are planned to be on-line by 2020 in
the United States, Spain, Abu Dhabi, Morocco, South Africa and other locations worldwide.
Development efforts are under way to improve solar collector assembly technology. A
consortium in Europe has developed and tested an advanced collector, dubbed the EuroTrough,
that is intended to contain the best features of the second and third generations of Luz collectors
(LS-2 and LS-3). Several of the parties to the EuroTrough development have also continued to
evolve variations of their own that minimize support frame material and offer simpler, faster
installation features. In addition, Acciona in the United States developed a collector based on the
LS-2 design that offers advantages in manufacturing, installation, and resistance to high winds.
Schott Solar and the former Siemens/Solel (recently divested) have each developed improved
vacuum-insulated receiver tubes.
8-27

Solar Thermal

8.4.2 Central Receiver


After operating for three years and successfully demonstrating the feasibility of molten-salt
central receiver technology, Solar Two shut down in 1999. Since the Solar Two project, the locus
of central receiver activity has moved to Europe, where interest in such concepts has continued
for more than two decades. A consortium formed by Boeing, Bechtel subsidiary Nexant, and
their Spanish partner Ghersa invested significantly to investigate design options, prepare cost
estimates, and negotiate contract opportunities for a new central receiver plant in Spain,
Gemasolar (formerly Solar Tres). After some delays and corporate realignment, Torresol Energy
developed the project.
As of May 2011, the 17-MW Gemasolar project near Seville, Spain became operational (Figures
8-10 through 8-12). Since then it has demonstrated close to full-load operation for an entire 24hour period and has run for complete weeks at full- or partial-load without any shutdowns. The
plant makes use of several advanced technologies, including 2650 glass-metal heliostats (120 m
each) with higher reflectivity glass and lower manufacturing costs, and a large molten-salt
storage system that provides up to 15 hours of capacity at 565C (1050F) and enables capacity
factors of about 75%. The central-receiver tower is 147 meters (480 feet) tall. Table 8-8
compares Solar Two and Gemasolar.
Table 8-8
Comparison of the Solar Two and Gemasolar projects
Solar Two

Gemasolar

Steam Turbine Net Power (MW)

10

19.9

Annual DNI (MWh/m2)

2.7

2.1

Surround

Surround

1.1

>2.5

Field Mirror Area (m2)

82,000

318,000

Receiver Size (MWt)

40

120

Cylindrical

Cylindrical

15

Mature annual capacity factor

20%

~75%

Mature annual efficiency

8%

unknown

Field Layout
Solar Multiple

Receiver Configuration
Storage (hrs)

In January 2009, Torresol Energya joint venture of SENER and Masdar of Abu Dhabi
invested 171 million to begin construction on the project. In November 2009, the European
Investment Bank (EIB) loaned an additional 80 million. The overall project is also backed with
funds provided under the terms of the Fifth European Community Framework Research Program
(Contract NNE5-2001-369).

8-28

Solar Thermal

Figure 8-10
Gemasolar plant under construction, May 2010
Source: SENER

Figure 8-11
Gemasolar molten salt storage tanks under construction, June 2010
Source: SENER

8-29

Solar Thermal

Figure 8-12
Gemasolar receiver tower and heliostat
Source: T. Peterson

The 5-MW Sierra SunTower built by eSolar began generating electricity in August 2009 (Figure
8-13). Constructed in less than one year, the plant features 24,000 heliostats, approximately 1 m2
each, that concentrate sunlight onto a central tower. The Sierra SunTower facility achieves 5
MW by combining two 2.5-MW modules. Power generated is being sold to Southern California
Edison.

8-30

Solar Thermal

Figure 8-13
Sierra SunTower plant using modular tower approach
Source: T. Peterson

The project, on private land zoned for heavy industrial use, was fully financed and developed by
eSolar, which hopes to use it as a template for future plants. In February 2009, eSolar announced
an agreement with NRG Energy to develop three plants in California and New Mexico that
would generate up to 465 MW. In March 2009, the company licensed its technology to the
ACME Group to develop up to 1 GW of solar-thermal capacity in India, and in January 2010 it
signed a memorandum of understanding to develop up to 2 GW of solar power in China. More
recently, the eSolar technology was licensed by GE for solar thermal hybrid applications. The
first such project will be located in Turkey.
In September 2007, BrightSource filed an application with the California Energy Commission
(CEC) to build three plants totally 377 MW on federal land, to be called Ivanpah 1, Ivanpah 2,
and Ivanpah 3. It won approval for a DOE loan guarantee, and began construction in 2010.
Electricity generated will be sold to PG&E. Contracts with PG&E and SCE total approximately
1800 MW.
The Ivanpah plants will employ BrightSources proprietary Luz Power Tower LPT 550
technology, whose working fluid operates at 550C (1020F). The company is also developing a
next-generation central receiver called the LPT 650 that would operate at 650C (1200F).
BrightSource expects each 100 to 250-MW unit to employ a single tower surrounded by tens of
thousands of heliostats, each about 10-m2 in area. BrightSource plans to deploy two units per
500-MW plant in future project designs. These plants could in principle also incorporate thermal
energy storage to increase their flexibility and dispatchability, and in fact, in November 2011
BrightSource announced that storage systems would be added to certain projects to be developed
for SCE. As a consequence one fewer plant will need to be built to satisfy the energy production
requirements of the PPA. The plants will use dry air-cooled condensers, rather than wet cooling,
to minimize water usage in the desert environment.
8-31

Solar Thermal

BrightSource selected Bechtel, an equity investor in the project, to build the Ivanpah plant in
September 2009. In February 2010, BrightSource requested the CEC to consider a revised design
that has only three towers and covers about 12% less land to lessen the plants environmental
impact, specifically to reduce incursion on desert tortoise habitat. Ivanpah is one of several
projects that BLM is fast-tracking to expedite their approval to make them eligible for cash
refunds in lieu of investment tax credits under ARRA. Construction began in late 2010 and
operations are scheduled to commence in 2013.
In late 2009, SolarReserve announced three large molten salt central receiver projects, one in
Spain and two in the United States. One of them, the 110-MW Crescent Dunes Solar Energy
Project with NV Energy to be located near Tonopah, Nevada, won approval from the Public
Utilities Commission of Nevada in July 2010, and the Rice Solar Energy Project was approved
by the California Energy Commission in December 2010. This is good news for the solar thermal
development community, as such plants represent an order-of-magnitude size increase over Solar
Twoand nearly as large a step beyond Gemasolar. As such, they could demonstrate actual
commercial-scale central-receiver-with-storage operation. However, such a large one-step scale
increase entails considerable risk, which SolarReserve addresses through a technology guarantee
from United Technology Corporation (UTC), a $50 billion company that licenses the technology
to SolarReserve. SolarReserve also has project contracts in Spain.
South Africas major electric utility, Eskom, has also been actively developing plans for solar
thermal projects for use in that country. Unlike the solar energy resources in Spain, where the
best resources are still 10% to 20% below that of the best Southwestern U.S. sites, some South
African sites are among the best in the world. The government of South Africa has set the goal of
generating 10,000 GWh of energy from clean renewable resources by 2013, representing about
5% of the nations current electricity consumption. In March 2009, the National Energy
Regulator of South Africa (NERSA) established a feed-in tariff for renewables based on the
levelized cost of electricity. South Africa has also established a National Solar and CSP research
program through which it provides funding to university researchers in the field. In late 2010 the
South African government announced plans for a 5-GW solar park to be developed in Upington
by 2020. So far two CSP projects with Abengoaa 100-MW trough with two hours of storage
and a 50-MW central receiver with three hours of storagehave been announced.
8.4.3 Dish/Engine
The dish/engine designs of Stirling Energy Systems, Inc. (SES) stem from the McDonnell
Douglas work of the 1980s. SES holds the rights to the McDonnell Douglas concentrator
technology, and continues the relationship with Kockums to use its 4-95 25-kWe engine, which
SES has the exclusive right to market in North America. The companys SunCatcher product
is a 25-kW dish/engine unit with radial curved glass mirror facets arranged in a paraboloid to
focus sunlight onto a four-cylinder reciprocating Stirling engine. In February 2008, a SunCatcher
system at Sandia National Laboratories set a world record for dish/engine solar-to-electric
efficiency of 31.25%. In 2004, SES began working with Sandia on a six-dish 150-kW
demonstration program to clarify and mitigate scale-up issues for the SES design. Research with
Sandia and others led to improvements in design, materials and weight, which should translate
into higher efficiency and lower cost.

8-32

Solar Thermal

SES was the leading producer of parabolic dish technology in the world until declaring
bankruptcy in November 2011. The companys future is uncertain, and the pipeline of projects in
southern California has been converted to PV. In the meantime, Infinia has made great strides in
recent years and has promising ideas for incorporating thermal energy storage into its
dish/engine designs. Infinia broke ground in August 2012 on a 1.5-MW project at Tooele Army
Depot near Salt Lake City, UT. The installation will use 430 of Infinias 3.5-kW PowerDish units
(see Figure 8-14 for an example of the PowerDish units). A 10-MW installation was expected to
be on-line by the end of 2012, and earlier in 2012 the company announced a 25.5-MW project in
Cyprus.
One strategy pursued by dish/engine developers is development of technologies and
manufacturing process with production partners from the automotive industry that are already
familiar with similar materials and requirements. Infinia, based in Ogden, Utah, claims multiple
years of maintenance-free operation at a cost moderately lower than photovoltaic technology.
They report a net peak efficiency of 24%. One unique aspect of the Infinia system is the
possibility of thermal energy storage using phase-change materials. The company received $9.4
million in 2008 and another $3 million in 2009 from DOE to develop a thermal energy storage
system for its dish/Stirling unit.

Figure 8-14
Thirty Infinia PowerDish units demonstrated in Spain
Source: DOE EERE

A European program called EuroDish is a joint venture of several German and Spanish firms,
with government support from the European Community and the German and Spanish
governments. This program is headed by Schlaich-Bergermann and Partner (SBP), and uses the
SOLO Kleinmotoren V-161 engine. Twelve units have been built and tested, most on the order
of 10 kW in size. SBP has focused attention on achieving fully automatic operation. There
appears to have been little EuroDish activity since about 2007.
8-33

Solar Thermal

In addition, a number of small industrial firms have emerged to develop and/or commercialize
Stirling systems of various sizes ranging from tens of watts to tens of kilowatts. Many of these
have originated since 1995, apparently in response to growing interest in distributed generation.
Owing to their use of external, continuous combustion, Stirling engines are expected to have
environmental characteristics superior to those of internal combustion engines. In fact the SOLO
unit mentioned above is being developed primarily for on-site cogeneration.
8.4.4 Linear Fresnel Reflector
Until it was acquired by AREVA Solar in February 2010, Ausra was the worlds leading
proponent and developer of compact LFR (CLFR) technology. The first significant prototype
demonstration of CLFR was the Liddell Solar Thermal Station in Australia, where a 3-MW
CLFR system was built to augment the steam cycle of a conventional 2000-MW coal-fired steam
turbine plant. In October 2008, the 5-MW Kimberlina pilot plant was commissioned in central
California to become the first commercial demonstration of CLFR technology. AREVAs stated
intention is to build an international solar steam-generating business around Ausras CLFR
technology, which may stimulate additional interest in the approach and investment by others.
They are currently constructing a 44-MW project that will augment a 750-MW supercritical coal
plant in Australia and a 250-MW standalone plant in India.
Puerto Errado 1 (PE1) is a 1.4-MW CLFR plant commissioned in April 2009 near the town of
Calasparra, Spain. Built by Novatec Biosol and Prointec, the plant consists of two 800-meterlong solar fields providing steam for a Rankine-cycle turbine-generator. It was considered a
prototype for larger plants to follow.
In February 2010, the international engineering firm TSK announced plans to build the 30-MW
Puerto Errado 2 (PE2) CLFR plant in Spain, consisting of Novatec Biosols solar field
technology and two 15-MW GE steam turbines. Each turbine is served by a solar field consisting
of 14 receivers 940 meters long, each flanked with eight rows of flat mirrors with a total area of
302,000 m2 to concentrate sunlight on the receivers. The project minimizes water consumption
by using air-cooled condensers and a robotic mirror washing method. Groundbreaking took place
in April 2010, and the project was completed in August 2012.

8.5 Cost and Economic Issues


The cost of electricity from solar thermal power plants depends on a multitude of factors,
including capital and O&M cost and system performance. It is important to note that the
technology cost and the eventual cost of electricity generated will be significantly influenced by
factors independent of the technology itself. For example, for parabolic troughs and central
receivers, small stand-alone projects would be very expensive. To reduce the technology costs to
be competitive with those of conventional fossil or nuclear technologies, solar projects must be
scaled up to larger plant sizes. One strategy already in practice is to create a solar power park, in
which multiple projects are built at a single site over time. Project infrastructure and mobilization
costs can be shared across all of the projects, and often the same maintenance crew can service
all of the projects. In addition, because solar thermal technologies are capital-intensive, the cost
of capital and taxation issues will have a significant impact on their competitiveness.

8-34

Solar Thermal

Except for the parabolic trough, none of the solar thermal technologies described in this report
has ever been available commercially. Although the PS10, PS20, and Gemasolar central receiver
projects are now providing electricity to the Spanish grid, they remain heavily dependent on
generous feed-in tariffs and have not released information about their capital expenses or O&M
costs. Therefore, real cost data based on recent commercial production and plant operation do not
exist (except, again, for parabolic trough plants).
Because of the limited activity in building new parabolic trough plants until the past five years
and the demonstration status of central receiver, dish/Stirling, and LFR technologies, EPRI notes
that they stand to benefit significantly from learning-by-doing cost reductions during commercial
scale up, and this effect cannot be represented accurately in the cost projections. EPRI therefore
recommends that use of these data be restricted to order-of-magnitude cost comparisons in farfuture planning, because actual costs will depend greatly on the cumulative deployments
achieved by a given year for each of the technologies.
8.5.1 Engineering and Economic Evaluation
The performance and cost data presented in Tables 8-9 and 8-10 are from the 2012 EPRI Report,
Solar Thermal Technology Status and Performance and Cost Estimates2011 [19]. Four solar
thermal technologies are addressed, parabolic trough with no thermal storage, parabolic trough
with six-hour thermal storage, central receiver with 10-hour thermal storage, and dish/Stirling
with no thermal sotorage. The levelized cost of electricity estimates in Table 8-10 are in
December 2010 dollars and are levelized over the 30-year project life.
Table 8-9
Performance and cost estimates for solar thermal technologies (December 2010$)
Source: EPRI [19]
Parabolic
Parabolic
Central
TroughTroughReceiverDish/Stirling
No Thermal 6-hr Thermal 10-hr Thermal
1
1
1
Storage
Storage
Storage
Rated Capacity
2 125 MW

2 125
MW

100 MW

4,000 25
kW (100
MW)

Unit Life, years

30

30

30

30

Land Required, acres/MW

5.6

7.7

14

Cooling Type

Wet

Wet

Dry

NA

Plant size (no. of units unit size, MW)


Physical Plant

8-35

Solar Thermal
Table 8-9 (continued)
Performance and cost estimates for solar thermal technologies (December 2010$)
Source: EPRI [19]
Parabolic
Parabolic
Central
TroughTroughReceiverDish/Stirling
No Thermal 6-hr Thermal 10-hr Thermal
Storage1
Storage1
Storage1
Scheduling
Hypothetical In-Service Year

Jan 2011

Jan 2011

Jan 2011

Jan 2011

Preconst., License & Design Time, Months

18

18

18

18

Idealized Plant Construction Time, Months

24

24

24

18

EOY 2010

EOY 2010

EOY 2010

EOY 2010

$2,168

$3,068

$3,410

$2,200

$1352

$1240

Capital Costs
Month/Year Dollars
Collectors, and Concentrators
Thermal Storage
Power Generation and Balance of Plant

$852

$852

$950

$920

Project and Process Contingency

$300

$528

$480

$620

Total Plant Cost


AFUDC (Interest during construction)

$3,320

$5,796

$6,150

$3,740

$72

$128

$140

$40

Total Plant Investment (includes AFUDC)

$3,392

$5,924

$6,290

$3,780

$680

$1,184

$1,260

$760

$4,072

$7,108

$7,550

$4,540

$64

$68

$66

$62

96

96

96

98

Plant Annual Output MWh/yr

556,000

892,000

430,000

208,000

Assumed DNI, kWh/m2-day

7.65

7.65

7.65

7.65

Commercial

Early
Commercial

Early
Commercial

Early
Commercial

Simplified

Simplified

Simplified

Simplified

OwnersCosts
Total Capital Requirement, Hypothetical InService Year (Includes AFUDC)
Operational and Maintenance Costs
Fixed O&M, $/kW-yr
Equivalent Planned Outage Rate, %
Equivalent Unplanned Outage Rate, %
Equivalent Availability, %

Confidence and Accuracy Rating


Technology Development Rating
Design & Cost Estimate Rating

8-36

Solar Thermal
Table 8-10
Constant-dollar levelized cost of electricity for solar thermal power plants (December 2010$)
Source: EPRI [19]
Parabolic
Parabolic
Central
TroughTroughReceiveNo Thermal 6-hr Thermal 10-hr Thermal
Dish/Stirling
Storage
Storage
Storage
Plant Size (no. of units unit size)

2 125 MW

2 125 MW

100 MW

4,000 25 kW
(100 MW)

25.4%

40.7%

49.1%

23.7%

152

147

129

175

Fixed O&M portion of LCOE

39

26

21

40

Capital Charge Portion of LCOE

113

122

109

135

224

225

198

262

Fixed O&M Portion of LCOE

39

26

21

40

Capital Charge Portion of LCOE

185

199

178

221

Capacity Factor
LCOE with Investment Tax Credit, $/MWh

LCOE without Investment Tax Credit, $/MWh

Constant-dollar evaluation assumes zero escalation for general inflation and fixed O&M inflation is excluded
from the debt interest and equity return rates.

8.5.2 Parabolic Trough


The data presented in Table 8-11 are derived from studies by EPRI and others [1, 79, 12, 19]
and will be used to estimate upper and lower bounds for cost and performance of a trough plant
with 6 hours of storage.
The lower bound on annual solar efficiency can be inferred from the SEGS VI column of the
table. The 10.7% efficiency value shown is somewhat higher than the 9.8% value calculated for
the entire Kramer Junction park in 1998. This can be explained entirely by the SEGS VI
collector aperture area, which, at 188,000 m2, is less than one-fifth or about 14% of the total
collector area of the park. Owing to higher steam-cycle efficiency, this plant outperformed the
earlier plants and proved that parabolic trough installations could achieve higher efficiency.
Contrary to earlier assumptions, it has been demonstrated that the later SEGS plants, including
SEGS VI, use little if any gas for plant warming. Thus, for the current evaluation, a value of 11%
will be used as the lower bound on this efficiency.
The column labeled Advanced assumes a good measure of optimism and includes assumptions
about cost reductions that are debatable and may prove unrealistic. For this evaluation, the data
presented for the near-term plant appear reasonable as a lower bound on cost for current
technology, and with experience and expanded production it seems reasonable to expect those
costs to drop by 10%. Hence, $490/m2 will be used as the lower bound cost estimate.

8-37

Solar Thermal
Table 8-11
Selected data for solar trough systems
SEGS VI

Near Term

Advanced

30

125

200

188,000

1,157,560

2,000,000

None
(fossil hybrid)

12

Capacity Factor (total) (%)

34

35.8

53

Solar Capacity Factor (%)

22.51

35.8

53

Insolation (DNI) (kWh/m2-yr)

2519

2519

2519

Annual Energy Production (GWh/yr)


(total) (gross)

89.4

427.2

928.6

Annual Solar Energy (GWh/yr) (gross)

59.0

427.2

928.6

O&M Costs ($/kW-yr)

107

65

281

O&M Costs (/kWh)

3.61

1.91

0.6

Collector Costs ($/kWe)

3048

2,710

14701

Collector Cost ($/m2)

4861

2931

147

Storage Cost ($/kWe)

NA

754

324

Total Capital Cost ($/kW)

3972

5,028

2535

Annual Solar-Electric Efficiency (net)

10.7

13.5

16

Plant Cost1 ($/m2)

633.8

542.9

253.5

Plant Size (MWe)


Solar Field Area (m2)
Thermal Storage (hours)

Data in italics are calculated values inferred from other entries or data discussed in references.

The near-term plant also provides the basis for estimating the upper bound on efficiency, again
because the apparent optimism of the advanced plant has been tempered. However, it seems
reasonable to expect that at least half of the 20% receiver efficiency improvement assumed in the
advanced plant will actually occur, which would increase plant efficiency by about 10% over that
of the near-term plant. Hence, 15% will be used for the current evaluation as the upper bound.
For the upper bound on cost, the value achieved with SEGS VI provides a firm basis. It is
reasonable to assume that additional experience would reduce that cost somewhat, perhaps by
5% to approximately $600/m2. Because that value is given in 1997 dollars, it is adjusted from
December 1997 to December 2012 dollars. Assuming a 2.5%/yr escalation rate, the increase is
about 45% and the resulting cost is about $870/m2.
A 90%-confidence levelized cost of electricity (LCOE) range can now be calculated for trough
systems. This analysis assumes a 10.5% fixed charge rate, covering all annual charges related to
the plants capital investment, and annual insolation of 2,500 kWh/m2. This insolation is
somewhat less than that reported earlier for Kramer Junction, but is felt to be more representative
of typical locations for trough plants. For O&M costs, the upper bound assumes the value
8-38

Solar Thermal

demonstrated at the Kramer Junction 150-MWe park: 2.5/kWh. For the lower bound, it is
reasonable to assume that mature plants would approach a value between those shown for the
advanced and near-term plants. For this evaluation, 1.0/kWh will be used. The resulting upper
and lower values for levelized COE, not including any financial incentives, are:
Levelized COEU = (100)(0.105)(800)/((2500)(0.11)) + 2.5 = 33.0 /kWh
Levelized COEL = (100)(0.105)(490)/((2500)(0.15)) + 1.0 = 14.7 /kWh
Thus, the LCOE for trough plants having modest technology advances is estimated to lie
between about 15 /kWh and 33 /kWh. These values are based on the assumed fixed charge
rate and insolation level, and can be easily adjusted using the preceding relationships to reflect
different values for those parameters if desired.
8.5.3 Central Receiver
With respect to the cost of a central receiver project, there is little basis to make substantial
changes from the projections EPRI developed in 1989 [7]. Solar Two, the only major central
receiver hardware activity in the United States since 1989, provided no new cost information,
although it did provide solid, positive experience with molten salt storage equipment and
operation. It therefore provides a qualitative reduction in prudent contingency factors for some of
the major plant components, and it strengthens the basis for a portion of the lower-bound cost
estimate from the 1989 evaluation. However, the lower bound also assumed significant cost
reduction of heliostats, based on the use of stretched polymer membranes rather than glass
reflectors. In 1989, the expectation was that the unit-area cost of heliostats would eventually drop
to $75/m2 if polymer reflectance and lifetime were acceptable. However, experience with
polymer membranes over the past decade for heliostats, dish reflectors, and troughs has not been
encouraging. All current and foreseeable system development work is now based on glass
reflectors. Hence the heliostat portion of the lower-bound cost is now somewhat suspect.
Cost data from the PS10, PS20, and Gemasolar plants are currently not publicly available, and
the saturated steam systems are not likely to become economical commercial products. Generous
Spanish feed-in tariffs were required to overcome the economic disadvantages of saturated steam
and make PS10 and PS20 feasible in the context of their unique market. Until large-scale
superheated steam systems are developed, the economics of commercial direct steam towers will
not be well-understood.
Additional perspective on cost projections can be obtained from EPRI and DOEs 1997
Renewable Energy Technology Characterization report [1] and EPRIs 2008 New Mexico
Central Station Solar Power: Feasibility Study [12]. These studies offer projections for a 2010
central receiver plant with 13 hours of molten-salt thermal storage and current cost estimates for
a 2011 plant with three to nine hours of storage, respectively. Projections in the former report are
also presented for 2020 and 2030, which include modest improvements in cost and operating
efficiencies. The improvements in operating efficiencies are judged to be speculative. The results
of the study in New Mexico are considered the most realistic near-term cost and performance
figures. It should also be noted that $75/m2 in current dollars has been a long-term goal for
heliostat costs for more than two decades. Thus, the goal in constant-dollar terms has actually
been decreasing. This reduces the credibility of the goal, especially because experience suggests
that the hoped-for reductions in materials costs are becoming less likely.
8-39

Solar Thermal

The 2011 central receiver plant cost, expressed in dollars per unit area of heliostats, was
projected to be $474/m2 in December 2009 dollars. In EPRIs 1989 evaluation, the corresponding
cost was about $296/m2 in 2009 dollars. In light of the preceding discussion of heliostat costs
and, in particular, the speculative nature of the $75/m2 goal, this evaluation will retain the lowerbound cost estimate of $296/m2. The upper-bound cost estimate will be changed from the value
calculated for the 1989 evaluation to 10% above the New Mexico estimate, or $520/m2.
The upper and lower bounds on the efficiency were assumed to be 10% above and below the
16% estimate in the New Mexico study. O&M was assumed to be similar to that of the trough.
Using the foregoing values in the same formula pair as the parabolic trough calculation yields an
LCOE range, expressed in December 2012 dollars, of approximately 8.6/kWh to 19.3/kWh.
8.5.4 Dish/Engine
Progress with dish/engine systems over the past 15 years has come primarily from testing
individual units fielded by SES, SAIC/STM, SBP, Sandia, and Infinia. In aggregate, including
the earlier experience of the 1980s, these units logged nearly 80,000 operating hours through
mid-2002. However, there is as yet no large body of operating experience on which to base
projections for mature-system energy costs that have substantially more validity than those
offered in EPRIs 1989 evaluation [7].
Similarly, because these programs were all heavily supported by government sources and
commercial offerings do not yet have public visibility, there is little basis for developing new
cost estimates with increased confidence. However, one factor is evident that would tend to
increase the lower bound cost estimate described in the 1989 evaluation: it now appears that all
firms pursuing dish systems are employing glass reflectors. As with central receiver heliostats,
experience with polymer membranes has been disappointing. This is expected to increase the
cost of the dish concentrator component by as much as 100%, which in turn would increase the
cost of the entire unit by about 30%. The impact on the lower-bound LCOE would be an increase
of about 30%, assuming no change in the upper-bound efficiency value.
Consequently, the LCOE range for dish systems today is estimated to be 9.7 to 37.7/kWh, in
2012 dollars.

8.6 Environmental Issues


Like solar PV technologies, solar thermal electric technologies are environmentally benign
relative to other forms of electricity generation. Without energy storage, most large-scale solar
plants have similar footprints, roughly 5 acres/MW (2 hectares/MW). LFR technology reportedly
uses as little as 3.5 acres/MW (1.4 hectares/MW). Except for dish/Stirling plants, solar thermal
electric generators also have water requirements unless they use dry cooling (which most new
plants in the U.S. Southwest plan to use) and potentially damaging outcomes of operational
mishaps. Of course, if the plant is hybridized, operation of the fossil-fueled section will be
accompanied by the usual emissions and hazards of fossil-fuel use.

8-40

Solar Thermal

8.6.1 Land Use


As mentioned, solar thermal plants without storage require a minimum of 3.5 to 5 acres/MW of
peak capacity in good solar-resource locales (greater than 2,200 kWh/m2/yr). However, plants
that incorporate storage will require more land per peak megawatt. Large-scale solar thermal
plants have footprints in the range of 5 to 10 acres/MW (2 to 4 hectares/MW) depending on the
number of hours of storage capacity (10 acres corresponds to about 9 hours of thermal energy
storage in good solar-resource locales of over 2200 kWh/m2/yr). Parabolic trough and central
receiver plants with thermal energy storage systems have an oversized collector field to capture
energy for the storage system during the peak hours of the day. The generating units will
typically not be designed to use the entire peak thermal output of the collector field. Therefore, a
more meaningful metric of land use would be the area required per annual megawatt-hour of
output, which would range between 2.7 and 3.0 10-3 acres/MWh/yr (1.11.2 10-3
hectare/MWh/yr). Note that, although conceptually distinct, these two quantities actually have
the same fundamental units of area per unit power output.
Dust suppression must be considered due to the large uncovered areas that are created by solar
thermal projects. Such consideration would include an economic evaluation of alternative control
methods for the collector field including compaction, dust suppressants, and permanent
stabilization.
In most cases, land appropriate for solar thermal projects consists of inhospitable, desert-like
terrain of little use for alternative development but which nevertheless raises concerns about
habitat disruption. In late 2009, potential conflicts between solar power and wildlife habitat were
spotlighted with Sen. Dianne Feinsteins (D-CA) introduction of the California Desert Protection
Act of 2010 (S.2921), which at this writing in late 2012 still has not made its way through Senate
committees. The act sets aside approximately 2.5 million acres of the Mojave Desert for
conservation by expanding Death Valley National Park, Joshua Tree National Park, and the
Mojave National Preserve, as well as establishing the new Mojave Trails National Monument
and Sand-to-Snow National Monument. Several solar power projectsmostly solar thermal but
also photovoltaichad been slated for siting on land that will be off limits to development if the
bill passes. Press reports claimed that as many as 13 solar power project developers canceled
proposals rather than work around the restrictions. However, of possible benefit to solar projects,
Section 203 of the bill addresses renewable energy and requires the Bureau of Land
Management, Department of Defense, and U.S. Forest Service to undertake environmental
studies and identify federal land where such projects could be expedited. In July 2012, the
Bureau of Land Management released a roadmap for solar energy development on federal lands
in seventeen zones across six states. The Programmatic Environmental Impact Statement (PEIS)
intends to streamline permitting and provide other incentives for large-scale solar projects in
regions identified as having good solar resource, transmission, and other favorable project
development characteristics. In general, environmental impacts on native plants and wildlife can
be minimized with proper planning and management.

8-41

Solar Thermal

8.6.2 Water Use


Water requirements of trough and tower solar thermal generating plants are similar to those of
other steam plants of equal nameplate capacity using wet cooling towers. Dry cooling is a viable
water-conserving alternative, but at a cost of up to 10% lower operating efficiency, depending on
ambient conditions. Increasingly, dry cooling is being required for thermal plants in desert
locations. In wet-cooled plants some 2 to 4 m3 of cooling water is needed for optimum efficiency
for every megawatt-hour generated. Dish/Stirling engines require no water for cooling. In all
cases, a small amount of water is needed for mirror cleaning.
8.6.3 Molten Salt for Central Receiver
No hazardous gaseous or liquid emissions are released during operation of a solar central
receiver plant. If a molten salt spill occurs, the salt will freeze before significant soil
contamination occurs. The salt may be picked up with a shovel and can be recycled if necessary.
8.6.4 Oil for Heat Transfer in Trough Applications
The current heat transfer fluid (Monsanto Therminol VP-1) is an aromatic hydrocarbon,
biphenyl-diphenyl oxide. The oil is classified as non-hazardous by U.S. standards but is a
hazardous material in California. When spills occur, contaminated soil is removed to an on-site
bio-remediation facility that uses indigenous bacteria in the soil to decompose the oil until HTF
concentrations have been reduced to acceptable levels. In addition to liquid spills, there is some
level of HTF vapor emissions from valve packing and pump seals during normal operation, but
such emissions are well within currently permissible levels.
8.6.5 Potential for Greenhouse Gas Reduction
Solar thermal power plants that are not hybridized with fossil fuel generate no direct emissions
of CO2, methane, or other greenhouse gases. Even when hybridized, the solar-generated portion
of the plants output is emissions-free. Consequently, all solar-thermal power plants provide
greenhouse gas emissions reductions. In addition, should a CO2 emissions-reduction mandate be
enacted in the future, solar thermal power could become an important component of a CO2
emissions-reduction strategy and could participate in CO2 emissions trading.
The CO2 emissions-reduction potential of a renewable energy power plant is a function of the
generation mix of the existing generation system, whereas the effective CO2 emissions-reduction
cost is a function of the CO2 emission rate and the average generation costs of the base system
and the renewable energy power plant.

8.7 Design, Deployment, and O&M Issues


Issues associated with design, deployment, and O&M include the reliability and maintainability
of sun-tracking drive systems and balance-of-plant systems, operation under extreme weather
conditions (wind, precipitation, dust), and labor requirements. As noted in the Cost and
Economic Issues discussion, there are currently few solar thermal systems in commercial
operation yielding limited real-world O&M experience.

8-42

Solar Thermal

8.7.1 Parabolic Trough


Through much of the 1990s, the DOE-Sandia-NREL solar thermal program worked closely with
the operators of the trough plants in southern California to reduce O&M costs for these systems.
That work focused primarily on the five 30-MWe plantsSEGS III through VIIlocated at
Kramer Junction, California. Through that effort, participants learned a great deal about
optimizing O&M procedures and about hardware improvements that would increase energy
production and reduce O&M costs. As a result, these costs have been substantially reduced, and
the O&M requirements for mature commercial plants based on similar hardware can be
estimated with improved confidence.
The solar collector assembly (SCA) (i.e., the mirrored trough unit) of a trough system rotates
around the horizontal north/south axis to track the sun through the sky during the day. The axis
of rotation is located at the collector center of mass to minimize the tracking power required, and
the drive system uses hydraulic rams to position the collectors. The SEGS plants were designed
for a nominal lifetime of 30 years, and the SCAs on those systems were designed for normal
operation in winds up to 25 mph (40 km/h) and with somewhat reduced accuracy in winds up to
35 mph (56 km/h). They are designed to withstand a maximum wind of 70 mph (113 km/h) in
their stowed position, aimed 30 degrees below the eastern horizon.
Actual experience in California has resulted in many lessons learned. During the period 1992
through 1998, the Sandia-industry O&M improvement program provided a 31% increase in
annual solar energy production from the 150-MWe solar park at Kramer Junction. In addition,
annual O&M costs per megawatt were reduced by 19%. Together, these advances resulted in
a reduction in O&M costs/kWh of 38%, from about 4/kWh in 1992 to about 2.5/kWh in 1998,
not including costs for natural gas fuel. The advances in collector design and receiver elements
described previously are also likely to increase annual energy performance through
improvements in wind damage resistance, integrity and protection of metal-to-glass seals,
stability of optical alignment of collector-reflectors and receiver elements, and increased
efficiency of receivers.
The following are some specific advances at Kramer Junction:

Although the mirrors have proven to be among the trough systems most reliable
components, differential thermal expansion caused some separation of the mirrors from their
mounting pads. The problem was resolved through the use of new mounting materials and
adhesives. In general, mirrors can be maintained without degradation through simple periodic
washing. Front-surface mirrors have the potential to yield higher reflectivity, but must
demonstrate long-term performance, durability, and washability.

Mirror breakage at the edges of the collector fields in high winds was a major problem that
often caused cascading breakage of encapsulated glass heat collection elements and other
mirrors. A means of strengthening the edge-mounted mirrors with a backing made of
fiberglass-resin successfully reduced breakage of these mirrors. Additional work developed
operating strategies for high wind conditions, with the aim of allowing plant operation in
high winds to continue and avoid loss of solar energy while also ensuring that vulnerable
collectors would be stowed in a safe position before wind damage could occur. Although
some valuable insights were developed, results of this work were inconclusive.

8-43

Solar Thermal

By 1992, approximately 15% of the 50,000 heat collection elements (HCE) at the Kramer
Junction park were damaged, either through loss of vacuum in the space between the receiver
and the concentric glass tube or through loss of the glass tube entirely. Because loss of the
glass reduced performance to a much greater extent than loss of the vacuum, workers
developed a temporary repair procedure in which a new glass tube could be affixed to the
HCE. Then after several adjacent HCEs developed problems, all could be replaced at the
same time. In addition, strategies to protect the glass-to-metal seal have reduced the
incidence of vacuum loss in the HCEs. Finally, Solel and Schott have developed new HCE
designs that use an improved selective receiver-tube coating with reduced emissivity and
greatly increased resistance to oxidation and the associated degradation of optical and
thermal properties in the event of vacuum loss. Protection of the vacuum seal is also
enhanced in these new designs. Future HCE designs should use new tube materials to
minimize bowing problems, allow broken glass sleeves to be replaced in situ, and continue to
improve coating materials that efficiently transmit solar energy to the working fluid.

The Luz LS-3 collector initially showed higher performance than earlier Luz collectors,
owing to higher concentration ratio and higher peak optical efficiency. Over time, however,
these units would lose their alignment, causing significant performance degradation.
Subsequent changes in the mechanical design of the LS-3 have corrected this problem, and
LS-3 collectors used in SEGS VIII and IX showed substantial improvement in alignment
maintainability. Operators also devised procedures for identifying collectors that were
developing alignment problems.

Because the solar collector assemblies track the sun, piping connections to the receiver tubes
must tolerate motion. Originally, corrugated flex hoses served this purpose. Reliability of
these was marginal, and occasionally a catastrophic failure would occur, resulting in a fire
fed by the heat-collection oil. The corrugations also caused noticeable pressure losses in the
heat-collection loop. Many of these hoses have been replaced with a ball-joint arrangement
that has greatly improved the reliability of these connections. Ball joints are highly likely to
be used in future trough plants. Better seals for pumps in the thermal collection loop
represent another advance, which has greatly improved pump reliability.

Several key issues remain for trough systems, including the following:

Despite substantial progress, O&M costs are still quite high. The trough community projects
that these costs can be reduced to 1/kWh or less, but this remains to be demonstrated.

Although incremental advances have been made, substantial reductions in capital costs have
not been demonstrated.

Increasing plant size, including thermal storage, and further increasing total collector area to
fill the thermal storage and increase plant capacity factor are likely to result in significant
cost reductions. Early experience with the molten-salt storage system of the Andasol-1 plant
appears very promising in this regard.

8-44

Solar Thermal

Molten salt storage appears to have good prospects for use in trough systems, particularly
because of its success with the central receiver approach. However, the storage subsystem
contributes more cost to a trough system than to a central receiver system, both because lower
temperature differences are involved, which increases the salt inventory per unit of energy
storage, and because salt cannot, as yet, be used as the heat collection fluid. Therefore, oil-tosalt heat exchangers are required that add a cost component beyond the salt storage tanks.

Alternative heat collection fluids have been proposed for trough systems, such as molten salt
and water-steam. These offer higher peak performance due to higher operating temperatures,
which could translate into lower costs. However, the technical difficulties of handling these
fluids without leaks or freezing throughout large collector fields in a wide range of
environmental conditions are significant. This indicates that credible cost projections for
near-term trough systems are limited to those based on close derivatives of the systems
deployed in southern California.

Other O&M issues include tracking maintenance via computerized maintenance management
software, performing periodic collector alignment, developing system design specifications and
operating procedures to allow quick start-up, and developing an effective O&M organization to
handle the special demands of operating a solar trough plant.
Prior to commercial operation, large solar systems in utility-size power plants need to pass a
performance acceptance test. NREL has undertaken the development of interim guidelines for
parabolic trough solar plants to recommend test procedures that can yield results of a high level
of accuracy consistent with good engineering knowledge and practice. The report, Utility-Scale
Parabolic Trough Solar SystemsPerformance Acceptance Test Guidelines [14] was published
in May 2011.
8.7.2 Central Receiver
The Solar Two plant was constructed during the mid-1990s. Initial operation began in June 1996,
and the plant was operated in various experimental modes through April 1999. The original
intent was to complete system checkout during the first several months, then conduct a one-year
experimental evaluation program during which the characteristics of the plant would be fully
explored and documented without emphasis on power production. This was to be followed by a
two-year period of simulated commercial operation in which optimum power production was the
only goal.
In practice, initial operation uncovered serious problems with the system design. The trace
heating subsystem had to be completely reworked, and the steam generator developed an early
tube rupture. These repairs and others set back the schedule by about 18 months. Consequently,
the experimental evaluation and the power production phases were conducted concurrently over
about a 12-month period. During this period, the design-point (i.e., peak) performance of the
system and major components was well characterized. However, operators did not collect longerterm operating performance information, so they could not determine actual annual system
performance. Some performance projections were possible, based on a combination of modeling
and actual data.

8-45

Solar Thermal

Peak component and system efficiencies for Solar One and Solar Two are shown in Table 8-12.
In addition to measured values for the two plants, this table shows goals for both Solar Two and
a commercial plant. The table shows that the net peak efficiency for Solar Two was projected to
be essentially the same as that actually observed for Solar One. For Solar Two, analysts factored
some acknowledged deficiencies in the heliostat field into the projected goal, but these were to
be offset by a higher receiver efficiency (which arises from a reduction in physical size, owing
to improved fluid heat transfer and, thus, reduced radiation losses). Parasitic losses would also
be higher for Solar Two, primarily because of the need for trace heating.
Table 8-12
Peak central-receiver system efficiency components
Solar One
Actual (%)

Solar Two
Goal (%)

Solar Two
Actual (%)

Commercial
Plant Goal (%)

90.3

90.7

90.7

94

Field Efficiency

76

72

66

71

Field Availability

99

98

94

99

Mirror Corrosion

100

97

97

100

Mirror Cleanliness

95

95

90

95

Receiver Efficiency

78

87

88

88

Storage Efficiency

N/A

99

99

99

Steam/Turbine Generator

33

34

34

42

Gross Peak Efficiency

16.6

17.2

14.7

23.2

Parasitic Loss Factor

90

88

87

93

Net Peak Efficiency

15

15

13

22

Category
Mirror Reflectivity

Solar Twos actual peak performance was somewhat less than the goal. This resulted primarily
because of deficiencies in the heliostat field, as discussed earlier. It is reasonable to assume that
those deficiencies would be eliminated in a new power plant. It is remarkable that the inclusion
of energy storage in the Solar Two plant produced an almost negligible impact on the overall
plant efficiencies.
Of greater significance are annual system efficiencies. Although these were not measured for
Solar Two, an attempt to estimate them was made by the Sandia evaluation team. These
estimates are shown in Table 8-13, along with measured values for Solar One and goals for Solar
Two and a commercial plant. The table also shows values adjusted for the known deficiencies of
Solar Two.

8-46

Solar Thermal
Table 8-13
Annual average central receiver system efficiency components
Solar One
Actual (%)

Solar Two
Goal (%)

Solar Two
Adjusted (%)

Commercial
Plant Goal (%)

Plant Availability

90

90

90

91

Overall Field Performance

58

50

51.5

56

Receiver Efficiency

65

76

76

79

Piping Efficiency

99

99

99

99

Storage Efficiency

N/A

97

97

99

Steam/Turbine Generator

30

32

41

41

10.1

10.5

13.9

16.3

Parasitic Loss Factor

72

73

73

90

Net Annual Efficiency

7.3

7.7

10.1

14.7

Category

Gross Annual Efficiency

The heliostat field represents the largest single capital investment in a central receiver power
plant. Relatively few heliostats have been manufactured to date, and their cost remains high.
In contrast to solar trough drive systems, which move on only one axis, heliostats must move
in two axes and reflect the suns rays onto the central tower with a high degree of accuracy.
Thus, related O&M is more complex and costly.
Although it is a suitable working fluid in many respects, molten salt has proven problematic in
others. It can be troublesome to maintain in a molten state, requiring a fairly complex heat
trace system employing electric wires attached to the outer surface of pipes. Valves can also
be a problem in a molten salt system, requiring special packing and extended bonnets. Design
improvements and standardization in valve design would reduce risk and ultimately reduce
O&M costs. Based on O&M experience at Solar Two, subsequent central receiver projects
are likely to employ new materials or thermal management practices that ease some of the
problems observed.
Since Solar Two was decommissioned and disassembled, the most ambitious and noteworthy
central receiver projects are the currently operating 11-MW PS10, 20-MW PS20, and 20-MW
Gemasolar systems. All are located in Spain. PS10 and PS20 use saturated steam as a heat
transfer fluid, whereas Gemasolar uses molten salt. Abengoa, which owns PS10 and PS20, has
not released detailed operational cost or performance information for those plants. More
generally, Abengoa has shared that day-to-day operation of the plants has gone as expected, and
that overall their performance has been satisfactory. Likewise, SENER and Torresol have shared
plant operational highlights [20], but not sufficient cost and performance data to provide insights
on subsystem costs and efficiencies.

8-47

Solar Thermal

8.7.3 Dish/Engine
Like central receiver heliostats, dish/engine systems must track the sun in two axes and with
extreme accuracy. This introduces increased complexity and potential for O&M problems.
Dish/engine systems must be stowed when wind speeds are too high, typically about 35 mph
(16 m/s). If intended for remote or autonomous operation, a dish/engine system must therefore be
equipped with reliable sun and wind sensors to determine whether conditions warrant operation.
Dish/engine systems need a parabolic reflective surface to focus sunlight on a receiver; however,
that does not necessarily require a mirror. Although a reflective surface of silver deposited on
glass is the most durable option, low-cost and lightweight polymer films stretched over hoops
and drawn into a spherical shape with a vacuum have also been used successfully on a small
scale, with hope that their relatively lower cost will offset their greater fragility.
Other systems of concern include the receiver, the enginemost commonly a Stirling engine,
although Brayton cycle engines are also being investigatedengine cooling systems, and
the alternator. In particular, the Stirling engine must demonstrate acceptable efficiency and
reliability before commercial deployment can be seriously considered. In general, dish/engine
systems would benefit from the introduction of a commercial engine system produced in volume.
Detailed operating experience and O&M data from Infinias installations are not yet available.
8.7.4 Linear Fresnel Reflector
The only significant demonstrations of LFR technology to date have been the Liddell Solar
Thermal Station in Australia, the 5-MW Kimberlina project in California, and the Puerto Errado
(PE) 1 and 2 projects in Spain. All except PE2 entered service in 2008. Detailed operating
experience and O&M data have not yet been released for any of them. However, all reportedly
operated as expected and provided their owners with sufficiently encouraging information to
develop more ambitious projects.

8.8 Equipment Markets and Key Participants


The high-concentration solar thermal electric technologies discussed in this chapter will
prove most practical and economical in regions with abundant direct-normal sunlight. This
requirement somewhat constrains suitable locations for solar thermal in comparison to other
solar technologies such as flat-plate photovoltaics, which can make use of diffuse or indirect
sunlight. In addition to the systems successfully operating in 13 countries, other nations pursuing
solar thermal power development include Canada, Portugal, Mexico, Greece, Turkey, Jordan,
UAE, India, Chile, Saudi Arabia, and South Africa.
The 1991 bankruptcy of Luz International and its subsidiaries had a large impact on the viability
of solar thermal systems for over a decade and considerably affected the pool of expertise
available. However, the recent rebuilding of interest in central-station solar power plants has
begun attracting both investors and engineering companies. The following list of developers,
consultants, and manufacturers includes those currently engaged in developing and providing
solar thermal technologies. This list is provided for informational purposes only; no endorsement
is implied.

8-48

Solar Thermal

Abengoa Solar
Based in Spain, Abengoa is committed to developing PV, solar thermal and industrial heat
technologies for commercial, industrial and utility applications.
www.abengoa.com
Acciona Energia
C/Yanguas y Miranda, 1-5, 31002-Pamplona, Spain. Accionas mission is to be leaders in
the management of infrastructures, services and renewable energy, actively contributing to social
well-being and sustainable development. Its annual revenue exceeds $5 billion and it operates on
five continents. Solargenix Energy, an Acciona subsidiary based in Sanford, North Carolina, is
developing a range of solar thermal systems using compound parabolic collectors for broad use
in buildings and electricity generation. Through Solargenix, Acciona also acquired intellectual
property from the former Duke Solar Energy.
www.acciona-energia.com
www.solargenix.com
Arizona Public Service Co.
P.O. Box 53999, Phoenix, AZ 85072. Through its Solar Test and Research (STAR) Center, APS
is a utility leader in researching and implementing many types of solar energy systems.
www.aps.com
AREVA
Headquartered in Paris, France, AREVA acquired U.S.-based Ausra in February 2010 with the
stated intention of using Ausras experience in compact linear Fresnel reflector (CLFR) systems
to anchor an international solar energy business. The new AREVA Solar continues to market
CLFR as a solar steam generator for power and industrial applications.
www.areva.com/
Black & Veatch
With over 30 years experience with assessing renewable energy technologies, Black & Veatch
has participated in hundreds of projects throughout the world. They provide engineering,
consulting and construction services.
www.bv.com
BrightSource Energy, Inc.
1999 Harrison Street, Suite 500, Oakland, CA 94612. BrightSource Energy designs and builds
large scale solar plants that deliver solar energy in the form of steam and/or electricity
to industrial and utility customers worldwide.
www.brightsourceenergy.com
CIEMAT
(Research Centre for Energy, Environment and Technology), Carretera Sens s/n-04200
Tabernes (Almera), Spain. CIEMAT is a Research Public Institution attached to the Spanish
Ministry of Education and Science. The Solar Platform of Almera, jointly managed by CIEMAT
and DLR, is one of the most important European Centers on Solar Thermal Energy.
www.ciemat.es/portal.do

8-49

Solar Thermal

Cobra
Cardenal Marcelo Spinola, 10, 28016-Madrid, Spain. Cobra is an engineering, operations,
installation and maintenance provider that supports several industries, including energy projects.
The company has over sixty year of experience and employs more than 22,000 professionals.
www.grupocobra.com
DLR (German Aerospace Center)
Cologne, Germany. This public non-government center focuses on aviation, space flight, and
energy technology in close cooperation with partners from academia and industry. An
experienced solar plant contractor and operating agent, DLR is involved in solar trough
feasibility studies in Spain and elsewhere.
www.dlr.de
eSolar Inc.
130 West Union Street, Pasadena, CA 91103. eSolar is a developer of a novel, modular, 46-MW
system for building distributed central-receiver plants.
www.esolar.com
FlabegSolar
Also known as Flabeg Solar International, this company is a subsidiary of the Pilkington Group,
one of the worlds leading glass manufacturers. FlabegSolar manufactures a highly-specialized
glass called Optisol that is used for solar reflectors, flat-plate collectors, and PV applications.
Pilkington supplied the parabolic trough reflectors for SEGS I through IX and is active in solar
thermal feasibility studies and development activities.
www.flabeg.com
Flagsol
Flagsol GmbH is a German joint venture between Solar Millennium AG and Ferrostaal AG to
develop parabolic trough solar thermal plants 50 to 250 MW in size, including the Andasol
plants in Spain and the Kuraymat plant in Egypt.
www.flagsol-gmbh.com/flagsol/index.html
Fluor
6700 Las Colinas Blvd., Irving, TX 75039. Fluor is active in solar, wind, geothermal, and
biomass activities around the world. In the solar field, it develops both photovoltaic and solar
thermal projects. In August 2009, Fluor won a contract to design eSolars 46-MW central
receiver plants, and, in February 2010, signed an agreement to develop a 50-MW CSP plant in
Spain.
www.fluor.com
Infinia
6811 West Okanogan Place, Kennewick, WA 99336, is a developer of small-scale
(3-kW) dish-Stirling systems and has recently begun marketing them for utility-scale
applications.
www.infiniacorp.com

8-50

Solar Thermal

IST Abengoa Solar Inc. (formerly, Industrial Solar Technology Corp.)


11500 West 13th Ave, Lakewood, CO 80215. IST designs, manufactures, installs, and operates
large-scale parabolic trough collector systems for industrial and commercial water heating, steam
generation, and absorption cooling.
www.industrialsolartechnology.com
Nexant Corp.
44 Montgomery St., San Francisco, CA 94104. In addition to its California headquarters, Nexant
has offices in Colorado, London, and Bangkok. Nexant was a key partner in early solar power
tower projects, including Solar Two, and designed the molten-salt storage system for the
Andasol trough plants in Spain.
www.nexant.com
NextEra Energy Resources
P.O. Box 14000, Juno Beach, FL 33408. Formerly FPL Energy and renamed in January 2009,
NextEra Energy Resources owns the SEGS parabolic trough plants in California.
www.nexteraenergyresources.com
Paneltec Corp.
Lafayette, CO 80026. Paneltec is involved in the development and manufacture of mirror facets
using sandwich panel technology. The company custom-laminates honeycomb and foam core
panels for a variety of scientific and industrial applications.
www.panelteccorp.com
Parsons
100 West Walnut St., Pasadena, CA 91124. Engineering and construction.
www.parsons.com
R.W. Beck
1125 17th St., Suite 1900, Denver, CO 80202.
www.rwbeck.com
Schott Solar
SCHOTT AG, Hattenbergstr. 10, 55122 Mainz, Germany.
www.schott.com/solar/
SkyFuel Inc.
10701 Montgomery Blvd. NE, Ste. A, Albuquerque, NM 87111. SkyFuel is the developer of
parabolic trough systems using low-cost non-glass reflectors.
www.skyfuel.com
SENER
With offices in Spain, Portugal, Argentina, Mexico, the U.S., and Poland, is an engineering,
consultancy and systems integration company currently involved in several major solar thermal
plant projects in Spain and the United States.
www.sener.es

8-51

Solar Thermal

SolarReserve, LLC
2425 Olympic Blvd, Ste 6040W, Santa Monica, CA 90404. SolarReserve has developed a
reference plant configuration for a concentrating solar power tower project using the proprietary
molten salt receiver, thermal storage and delivery system developed by Pratt & Whitney
Rocketdyne, Inc. and Hamilton Sundstrand Corporation, both of which are wholly owned
subsidiaries of United Technologies Corporation. SolarReserve is owned by USRG Power &
Biofuels Fund II, L.P. and Hamilton Sundstrand Corporation and was formed to develop utilityscale solar power projects using this technology.
www.solar-reserve.com
WorleyParsons
With offices in 38 countries, WorleyParsons provides professional engineering, construction and
operation services to solar projects throughout the world. The company is technology-neutral,
with over 5000 MW of solar thermal under development.
www.worleyparsons.com
8.8.1 Internet Resources
In addition to the vendors and organizations listed in the preceding section, the reader can visit
the following websites. Note that although these sites contain information and contacts that may
be of interest, EPRI cannot vouch for the accuracy or quality of all of their content.
U.S. DOE Energy Office of Efficiency and Renewable Energy
www.eere.energy.gov
U.S. DOE Concentrating Solar Power Program
The U.S. DOE administers the SunLab program through two of its national laboratories:
Sandia National Laboratory in New Mexico, and the National Renewable Energy Laboratory
in Colorado. SunLab operates the National Solar Thermal Test Facility.
http://www.sandia.gov/csp/csp_r_d_sandia.html
U.S. DOE/USA Solar Trough Initiative
Funded by the U.S. DOE and operated by SunLab, TroughNet is a comprehensive source for
government efforts in solar thermal technology, projects, research and development, and other
related resources.
www.nrel.gov/csp/troughnet/
U.S. National Renewable Energy Laboratory (NREL)
www.nrel.gov/csp/
American Solar Energy Society (ASES)
ASES is a national organization dedicated to advancing the use of solar energy for the benefit
of U.S. citizens and the global environment. The society sponsors conferences and publishes a
magazine and white papers on the subject. ASES is the U.S. section of the International Solar
Energy Society (ISES).
www.ases.org
www.ises.org

8-52

Solar Thermal

Solar Energy Industries Association (SEIA)


Established in 1974, SEIA is the national trade association of the U.S. solar energy industry. As
the voice of the industry, SEIA works with its 1000 member companies to make solar a
mainstream and significant energy source by expanding markets, removing market barriers,
strengthening the industry and educating the public on the benefits of solar energy.
www.seia.org
Solar Electric Power Association (SEPA)
Formerly the Utility PhotoVoltaic Group (UPVG), SEPA is a non-profit collaboration
comprising more than 700 utilities, energy service providers, and the PV industry working to
create and encourage commercial use of new solar electric power technology and business
models. SEPA also created and maintains the Solar Data and Mapping Tool, a web-based
utility that provides project data for solar power projects in the United States
(www.solarelectricpower.org/solar-tools/solar-data-and-mapping-tool.aspx).
www.solarelectricpower.org
SolarPACES
Solar Power and Chemical Energy Systems (SolarPACES) is an international cooperative
organization managed by the International Energy Agency. Member countries (currently 16)
develop solar thermal technologies by addressing technical problems associated with
commercialization and market development.
www.solarpaces.org
ESTELA
The European Solar Thermal Electricity Association (ESTELA) is an organization of developers,
manufacturers, utilities, engineering firms, and research institutions created to support the
emerging solar thermal industry in Europe.
www.estelasolar.eu

8.9 References
1. Renewable Energy Technology Characterizations. EPRI and U.S. Department of Energy:
1997. TR-109496.
2. Zavoico, A., W. Gould, B. Kelly, and I. Pastoril, Solar Power Tower (SPT) Design
Innovations to Improve Reliability and PerformanceReducing Technical Risk and Cost,
presented at the ASME 2001 International Solar Energy Conference, Washington, D.C.:
April 2001.
3. Kelly, B., Herrmann, U., and Hale, M.J., Optimization Studies for Integrated Solar
Combined Cycle Systems, presented at the ASME 2001 International Solar Energy
Conference, Washington, D.C.: April 2001.
4. Stone, K., Liden, B., Ellis, E., Sattar, T., and Mancini, T., SES/Boeing Dish Stirling
System Operation, presented at the ASME 2001 International Solar Energy Conference,
Washington, D.C.: April 2001.
5. Diver, R., Andraka, C., Rawlinson, K., Thomas, G., and Goldberg, V., The Advanced
Dish Development System Project, presented at the ASME 2001 International Solar Energy
Conference, Washington, D.C.: April 2001.
8-53

Solar Thermal

6. Mayette, J., Davenport, R., and Forristall, R., The Salt River Project SunDish-Stirling
System, presented at the ASME 2001 International Solar Energy Conference, Washington,
D.C.: April 2001.
7. Holl, R.J., Status of Solar-Thermal Electric Technology. EPRI, Palo Alto, CA: 1989.
GS-6573.
8. DeMeo, E., Solar-Thermal Electric Power: 2003 Status Update. EPRI, Palo Alto, CA:
2003. 1008463.
9. Price, H., Lupfert, E., Kearney, D., Zarza, E., Cohen, Gl, Gee, R., and Mahoney, R.,
Advances in Parabolic Trough and Solar Power Technology, Journal of Solar Energy
Engineering, Vol. 124: May 2002.
10. Stirling Engine Assessment. EPRI, Palo Alto, CA: 2002. 1007317.
11. Leitner, A. Fuel from the Sky: Solar Powers Potential for Western Energy Supply,
NREL/SR-550-32160. Golden, CO: National Renewable Energy Laboratory, July 2002.
12. New Mexico Central Station Solar Power: Final Summary Report. EPRI, Palo Alto, CA
and the Public Service Company of New Mexico, El Paso Electric, San Diego Gas &
Electric, Southern California Edison, Tri-State Generation & Transmission Association,
and Xcel Energy: 2008. 1016344.
13. Case Study: Gemasolar Central Tower Plant, SENER Engineering and Systems Inc.,
presentation at Concentrated Solar Power Summit conference, San Francisco, June 2010.
14. Utility-scale Parabolic Trough Solar SystemsPerformance Acceptance Test Guidelines,
NREL: August 2010. Draft report.
15. Solar Thermal Electric Technology: 2009. EPRI, Palo Alto, CA: 2010. 1021386.
16. Concentrating Solar Thermal Technology. EPRI, Palo Alto, CA: 2009. 1018577.
17. Australian Sustainable Energy: Zero Carbon Australia Stationary Energy Plan, Energy
Research Institute, University of Melbourne: July 2010.
18. Solar Energy Technologies Program Peer Review, DOE Energy Efficiency and Renewable
Energy (EERE), Washington D.C.: May 2010.
19. Solar Thermal Technology Status and Performance and Cost Estimates2011. EPRI, Palo
Alto, CA: 2012. 1025007.
20. Case Study on Thermal Energy Storage: Gemasolar. EPRI, Palo Alto, CA: 2012. 1025405.

8-54

OCEAN TIDAL ENERGY

9.1 Introduction
The history of tidal power goes back to at least the Middle Ages. Primitive tidal mills were
operated in England in the 11th century, and numerous tidal mills were built in Western Europe,
Canada, and the United States during the 18th century. As a result of the Industrial Revolution
and subsequent widespread use of electric power, tidal mills became obsolete.
Actual construction of first-generation modern tidal power plants (dam or barrage systems) has
been limited to one large, 240-MW facility at LaRance, France; a small, 20-MW experimental
plant at Annapolis, Nova Scotia that is still operating today; and a number of small experimental
plants in Russia and China. During the summer of 2011, a tidal power plant began operation at
Lake Shihwa on the west coast of South Korea. When fully operational it will be the worlds
largest tidal power plant, at 254 MW.
Second-generation or tidal in-stream tidal energy conversion (TISEC) technology, which
captures the kinetic energy from unimpounded tidal currents, is a new technology that has
heightened interest in the commercial development of tidal energy conversion technology in
Europe, Korea, Australia, New Zealand, Canada, and the United States over the past five years.
This second-generation TISEC is the focus of this chapter (see Table 9-1 for an overview).
Table 9-1
Ocean tidal energy overview
Installed Tidal
Capacity
(as of November
2011)

About 6 MW worldwide; 60 kW in the United States.


Estimated annual incremental U.S. capacity additions:
2012: 260 kW
20132015: 10.6 MW (33 MW globally)
Estimated cumulative capacity by 2025: 500 MW.
Total theoretical available tidal power in U.S. is 50GW
U.S. tidal energy hot spots: Alaska, Maine, Washington, Oregon, California,
New Hampshire, Massachusetts, New York, New Jersey, North and South
Carolina, Georgia, and Florida (descending order)

9-1

Ocean Tidal Energy


Table 9-1 (continued)
Ocean tidal energy overview
Tidal Energy
Conversion
(TISEC)
Technology
Readiness
(as of August
2012)

Economic Status

9-2

TISEC is an emerging technology. As of May 3, 2012 Open Hydro, Voith


Hydro, Atlantis Resources Corporation, Bluewater Energy Services,
Scotrenewables, Andritz Hydro Hammerfest, Tidal Generation Ltd, and
Kawasaki Heavy Industries are each occupying one of the eight available
berths at the European Marine Energy Center (EMEC) to test their large-scale
prototype devices. On December 22, 2011 the NYS Department of
Environmental Conservation granted Verdant Power a Water Quality
Certification Permit for the Roosevelt Island Tidal Energy (RITE) Project. On
January 23, 2012 FERC granted Verdant Power a 10-year hydrokinetic pilot
project license, making it the first licensed tidal power project in the U.S. The
project is being developed in a phased approach to include up to 30 turbines
providing 1 MW of power.
On February 27, 2012 ORPC received a pilot project license from FERC for
the Cobscook Bay project in Maine. On April 26, 2012 the Maine Public
Utilities Commission approved the primary contract terms of power purchase
agreements (PPAs) for the ORPC Maine Tidal Energy Project. These will be
the first long-term PPAs for tidal energy in the United States. On July 25, 2012
ORPC held the dedication ceremony for the project, making it the first
commercial tidal energy project in the U.S. On September 13, 2012, the
underwater turbine generator came on-line. The TidGen is designed to have a
peak output of 180 kilowatts. Two additional TidGen devices will be installed in
the fall of 2013.
On August 12, 2011 Neptune Renewable Energy Ltd announced that it has
been given preliminary approval for the deployment of the Neptune Proteus
NP1000 demonstrator device and is presently developing its first commercial
array of five advanced Neptune Proteus NP1500s in the UK. Commissioning of
the demonstrator device began in November 2011.
EPRI economic feasibility studies indicate TISEC technology is competitive at
a half dozen or so very good sites in the United States (average annual
power density greater than 2 kW/m2 and meeting other required site
attributes).
The DOE is funding a large multi-organizational project from May 2010 to
September 2013 to develop Reference Models. According to a DOE
presentation, the goal of the project is to, Develop a representative set of
Reference Models (RM) for the MHK industry to develop baseline costs of
energy (COE) and evaluate key cost component/system reduction pathways.

Ocean Tidal Energy


Table 9-1 (continued)
Ocean tidal energy overview
Environmental
Impact

Regulatory Status

Government
Support of Tidal
Energy
Technology

Proper care in siting, installation, operation, and decommissioning may enable


tidal energy technology to be one of the more environmentally benign
electricity generation technologies.
The DOE provides funding to six national laboratories to conduct
environmental research in six critical areas: Physical Interactions with
Devices, Electromagnetic Fields, Acoustics, Toxicity, Benthic Habitat
Alteration, Data Aggregation and Risk Modeling. Additionally, the DOE
funded the creation of three National Marine Renewable Energy
Centers in partnership with local universities, including: Northwest
National Marine Renewable Energy Center (NNMREC), Hawaii
National Marine Renewable Energy Center (HINMREC), and
Southeast National Marine Renewable Energy Center (SNMREC).
Several numerical modeling, laboratory flume, and field monitoring studies
have been performed to determine the environmental impacts from tidal power
devices. Preliminary results indicate that there is minimal environmental impact
associated with these devices. One such study was published in February of
2012 by the Pacific Northwest National Laboratory and Sandia National
Laboratory, titled, Assessment of Strike of Adult Killer Whales by an
OpenHydro Tidal Turbine Blade, and determined that a Southern Resident
Killer Whale is not likely to experience significant tissue injury from impact by
an OpenHydro turbine blade.
The U.S. Department of Energy and the Pacific Northwest National Laboratory
developed TETHYS, a database and knowledge management system that
provides access to information and research pertaining to the potential
environmental effects of marine and hydrokinetic (MHK) and offshore wind
development. TETHYS also hosts data from Annex IV, an international
collaboration to gather information on MHK environmental research
worldwide.
The Federal Energy Regulatory Commission (FERC) has primary jurisdiction
for licensing tidal in-stream power plants under the Federal Power Act (FPA).
The time, cost, and complexity of the U.S. regulatory process can be difficult
for tidal project developers.
There are 26 active and 3 pending FERC-issued preliminary permits for tidal
projects as of August 2, 2012, nine of which were issued in 2012.
As of August 2, 2012 a FERC decision is still pending on an application for a 5MW exemption from licensing for the TideWorks Project, Sasanoa River,
Maine.
European governments (particularly in the UK, Ireland, Portugal, and
Denmark) as well as those in Canada, Japan, New Zealand, and Australia
support the development of TISEC technology and are now providing
subsidies to stimulate a commercial market.
Nova Scotia is currently in the process of finalizing its Tidal Array Feed-In
Tariff, which will greatly assist projects such as Fundy Tidal Inc.s Petit
Passage project.
DOE initiated a Waterpower R&D Program in FY 2008 with a congressionally
mandated $10 million. FY 2012 received a budget increase totaling $59 million
($34 million for MHK and $25 million for conventional hydro). The FY 2013
budget proposal for the water program is only $20 million, which represents a
66% reduction from the FY 2012 budget.

9-3

Ocean Tidal Energy


Table 9-1 (continued)
Ocean tidal energy overview
Trends to Watch

SnoPUD filed a Final License Application with FERC on March 1, 2012 and is
awaiting FERC approval for the Admiralty Inlet project.
Additional megawatt-scale demonstration and early commercial projects are
expected to be deployed over the next decade in Canada, Europe, Korea,
Australia, New Zealand, and the United States.
Major players in the wind and conventional hydropower industries are
acquiring significant stakes in hydrokinetic device development and
manufacturing companies. The technical and business experience they bring
to hydrokinetic generation may accelerate the development of the tidal
generation industry.
Final reports for DOE Reference Models are expected by September 30, 2013.

The tides are the result of gravitational forces exerted by the Moon and the Sun on the oceans of
the Earth, with the Moon having the predominant influence. The changing relative position of
these bodies causes the surface of the oceans to rise and fall periodically, as illustrated in
Figure 9-1. The gravitational forces of the Sun and the Moon and the centrifugal forces due to
the rotation about the Earth-Moon center-of-gravity create two bulges in the Earths oceans:
one closest to the Moon, and the other on the opposite side of the globe, as illustrated in Figure
9-2. These bulges result in two tides per day, called semi-diurnal tidesthe dominant tidal
pattern in most of the worlds oceans.
TISEC sites are typically located where there is a narrow channel or passage between two land
masses or an inlet into a bay, through which substantial volumes of tidal water must flow at high
speed. Tidal power sites should also have suitable seabed geology for proper anchoring of the
TISEC devices, and should be situated reasonably close to existing grid interconnection points.

Figure 9-1
Earth, Moon, and Suns influence on tides

9-4

Figure 9-2
The Earths tidal bulge

Ocean Tidal Energy

An illustration of a typical open-rotor, axial-flow water turbine TISEC device is shown in Figure
9-3. For such a water turbine, the hydrokinetic power per unit cross sectional area (in this case,
the area of the rotor blades) is equal to one-half the density of water multiplied by the cube of the
water speed. This relationship is illustrated in Figure 9-4.

(P/A)flow = 0.5 x density of water x V3

Figure 9-3
Typical hydrokinetic water turbine

Figure 9-4
Tidal power density

9.2 U.S. Tidal In-Stream Energy Highlights: Late 2011 to Mid-2012


Interest in ocean tidal in-stream renewable energy continues to grow in the United States. This
section provides a brief account of notable federal and state developments as well as device
developer activities within the sector from late 2011 through mid-2012.
9.2.1 U.S. Federal Highlights
The following sections present the latest marine and hydrokinetic (MHK) federal developments
with regard to congressional appropriations, awards issued, and regulatory updates.
9.2.1.1 Federal Appropriations
In FY 2012, the U.S. Congress appropriated $59 million to the Department of Energy Water
Power Program to conduct research and development (R&D) on advanced water power energy
generation technologies, including both MHK technologies ($34 million for wave, tidal, ocean
current, in-stream hydrokinetic and ocean thermal), as well as conventional hydropower ($25
million for any hydropower technology that uses a dam or diversionary structure). This marks a
significant increase in funding from the $10 million appropriated in 2008, which was the first
year of the Water Power Program. However, the FY 2013 budget proposal announced in
February 2012 allotted only $20 million to the Water Power Program, representing a 66%
reduction from the previous year. Fortunately, in April 2012 the House and Senate
Appropriations Committee passed the FY 2013 Energy and Water Development Funding Bill.
The Senate funding measure provides $59 million to the program ($44 million for MHK and $15
million for conventional hydropower) and the House measure provides $45 million to the
program ($25 million for MHK and $20 million for conventional hydropower).
9-5

Ocean Tidal Energy

The DOEs marine and hydrokinetic energy research is focused on assessing the potential
recoverable energy from MHK resources in the United States, and facilitating the development
and deployment of technologies to fully realize this potential. MHK technologies represent an
opportunity for the United States to engage directly in an emerging area of science and
discovery, while developing an entirely new suite of renewable energy technologies available to
reduce emissions and help states meet renewable energy portfolio standard (RPS) targets.
The DOEs priorities for tidal power include the following:

Facilitating the deployment of prototypes, and collecting data on the energy conversion
performance and environmental impacts of the devices

Determining the available, extractable, and cost effective resources in the United States

Characterizing and comparing the wide variety of existing marine hydrokinetic technologies

Improving technology performance and reliability, and reducing technology development


costs

Minimizing the cost, time, and negative impacts associated with siting projects

9.2.1.2 Funding Awards


In 2008, the majority of DOE funding for tidal power was awarded to specific technology and
project development efforts, selected through a competitive process. Those awarded efforts
included the following:

The design, fabrication and testing of an improved turbine blade design structure for Verdant
Power, Inc.

A program to conduct in-water testing and demonstration of tidal flow technology as a first
step toward deploying a commercial tidal power facility by the Snohomish Public Utility
District (SnoPUD), a municipal utility in Washington State.

The selection and funding of two National Marine Renewable Energy Centersone at the
University of Hawaii, and a second run jointly by Oregon State University and the University
of Washington. Further, DOE funded a market acceleration program, consisting of
nationwide resource wave and tidal hydrokinetic resource assessments and a collaborative
project to address navigation and environmental issues as well as clarify the permitting
process.

In 2009, DOE support for tidal power included funding of the following:

An EPRI project to conduct desktop modeling studies of strike probability and flume testing
of fish interactions with hydrokinetic turbines

Various projects related to siting and evaluation of tidal energy projects

DOE funding in 2010 focused on advancing the development of marine and hydrokinetic
systems and components.

9-6

Ocean Tidal Energy

Throughout 2011 and 2012 the DOE offered multiple financial opportunities to MHK industry
businesses, including special loan guarantees and grants. The main types of awards offered to
MHK developers included Small Business Innovation Research (SBIR) programs and Small
Business Technology Transfer (STTR) programs. In FY 2010 the program selected 27 projects
for funding, with individual awards ranging from $160,000 to $10 million. In April 2012 the
DOE announced that there is $9 million available this year to fund approximately 50 small
businesses. Selected projects will receive 1-year awards of up to $150,000. Awardees with
successful projects will also have the opportunity to compete for follow-on funding in excess of
$1 million. These funding programs are intended to help emerging MHK technologies advance
along the DOE Technology Readiness Level (TRL) chain. There are 10 TRL levels that are
organized as follows:

TRL 13

Concept Definition

TRL 4

Proof of Concept

TRL 5/6

System Integration and Laboratory Demonstration

TRL 7/8

Open Water System Testing, Demonstration, and Operation

TRL 9

Array Testing

TRL 10

Commercialization

For the tidal industry specifically, the DOE funded Georgia Tech to develop a tidal resource
assessment. The report was released in June 2011 and is titled Assessment of Energy
Production Potential from Tidal Streams in the United States. The DOE is also funding a large
multi-organizational project to develop MHK reference models. The project began in May 2010
and according to a DOE presentation, the goal of the project is to develop a representative set of
Reference Models (RM) for the MHK industry to develop baseline costs of energy (COE) and
evaluate key cost component/system reduction pathways. The presentation notes that there is a
need [to identify] COE targets with regard to technology type and future innovation
opportunities to prioritize research and cost reduction pathways. The reference model
completion date is slated for September 2013.
So far in 2012 five funding opportunity announcements have been issued for water power
technologies:
1. Open Funding Opportunity Announcement
Advanced Research Projects Agency-Energy (ARPA-E)
2. Hydropower Advancement Project (HAP)Standard Assessments to Increase Generation
and Value
U.S. Department of Energy Office of Energy Efficiency and Renewable Energy
3. In-Water Wave Energy Conversion (WEC) Device Testing Support
U.S. Department of Energy Office of Energy Efficiency and Renewable Energy Water Power
Program
4. FY12 SBIR/STTR Funding Opportunity Announcement
Small Business Funding Opportunity Announcement (FOA)
9-7

Ocean Tidal Energy

5. DE-FOA-0000747: RFIImproving Marine and Hydrokinetic and Offshore Wind Energy


Resource Data
U.S. Department of Energy Office of Energy Efficiency and Renewable Energy
9.2.1.3 Federal Regulations
On August 3, 2009, Minerals Management Service (renamed the Bureau of Ocean Energy
Management, Regulation, and Enforcement [BOEMRE]) and the Federal Energy Regulatory
Commission (FERC) issued guidance on the regulation of hydrokinetic projects on the Outer
Continental Shelf (OCS).
On July 29, 2012 the newly renamed Bureau of Ocean Energy Management (BOEM) (formerly
BOEMRE) and FERC announced revised guidelines for MHK developers pursuing permitting
and licensing on the OCS. The revisions further clarify the regulatory process and help
streamline the process for authorizing research and testing of MHK devices.
9.2.2 U.S. State Highlights
9.2.2.1 Projects in Alaska
EPRI estimates that over 90% of the U.S. tidal hydrokinetic energy resource exists in Alaska. In
2011, the DOE and Georgia Tech released a tidal resource assessment that estimated
approximately 47 GW of tidal power is available in Alaska. A key limitation to the use of these
resources for electricity generation is that most of the resource is found in remote areas and not
adjacent to any electric transmission infrastructure necessary to provide power export
capabilities at significant scales. As of February 15, 2012 five tidal projects in Alaska hold
preliminary permits issued by FERC.
ORPC Cook Inlet Project (P-12679). Ocean Renewable Power Company (ORPC) was issued a
preliminary permit for the proposed Cook Inlet Tidal Energy Pilot Project on October 13, 2010.
At that time the project had an authorized capacity of 1 MW. As of March 30, 2012 no other
permits have been issued to the project. FERC has granted ORPC time extensions to respond to
requests for additional information, and three progress reports have been completed. According
to the U.S. DOEs Marine and Hydrokinetic Technology Database the project is in Phase 2Site
Development, as of July 2012.
Turnagain Arm Tidal Energy Project (P-13509). Turnagain Arm Tidal Energy Corporation was
issued a preliminary permit for its project of the same name on February 5, 2010. The proposed
project consists of 8-mile-long and 7.5-mile-long tidal fences containing 220 10-MW Davis
turbines (a modification of the Darrieus cross-flow turbine); total capacity is 2200 MW. The
project site is located in Cook Inlet on the Kenai Peninsula, Alaska. On January 20, 2012 FERC
notified the project that their proposed study plan is deficient and request corrections and new
proposed study plan to be submitted with 30 days. On February 15, 2012 the project withdrew
its Notice of Intent and Preliminary Application Document. Progress report #5 was submitted on
July 20, 2012 and the project remains in Phase 1Siting/Planning in the DOE MHK database.

9-8

Ocean Tidal Energy

Icy Passage Tidal Project (P-13605). Natural Currents Energy Services was issued a preliminary
permit for its Icy Passage Tidal Project on April 30, 2010. This 300-kW project would be located
in Icy Passage, near Gustavus, Alaska. The project has continued to file progress reports with
FERC up through March 29, 2012. The project remains in Phase 1Siting/Planning in the DOE
MHK database.
Gastineau Channel Tidal Project (P-13606). Natural Currents Energy Services was issued a
preliminary permit for its Gastineau Channel Tidal Project on April 30, 2010. This 400-kW
project would be located in Gastineau Channel, near Juneau, Alaska. The project has continued
to file progress reports with FERC up through March 29, 2012.
East Foreland Tidal Energy Project (P-13821). ORPC received a FERC preliminary permit for
its proposed East Foreland project on March 11, 2011. The project has a combined capacity
between 5 and 100MW, and would be located in Cook Inlet near Nikiski, Alaska. The project
has continued to file progress reports with FERC up through February 27, 2012.
Killisnoo Tidal Energy Project (P-13823). Natural Currents Energy Services received a
preliminary permit from FERC on January 21, 2011. The 250-kW project would be located in
Kootznahoo Inlet, northeast of Killisnoo Island, near the City of Angoon, in southeastern Alaska.
The project has continued to file progress reports, but as of July 26, 2012 the projects third
progress report is overdue.
9.2.2.2 Projects in Washington
There are three tidal power projects under consideration or development in Washington state.
Two have been issued preliminary permits by FERC, and the other is a U.S. Navy (federal)
project.
Deception Pass Tidal Energy Project (P-12687). FERC issued a preliminary permit for
SnoPUDs Deception Pass Tidal Energy Project on August 4, 2010. The proposed project would
consist, in part, of four 20-m, 1.6-MW horizontal-axis turbines and would be located in Puget
Sound within Deception Pass, between Whidby Island and Fidalgo Island. The preliminary
permit is a reissuance of a preliminary permit held by SnoPUD. Under the previous preliminary
permit, SnoPUD filed a pre-application document (PAD) for the project on January 31, 2008.
SnoPUD engaged in other activities, such as consultation with resource agencies and other
stakeholders, that were directed toward submission of a license application. Under the reissued
Preliminary Permit, SnoPUD proposed to submit a draft license application for the Deception
Pass Tidal Energy Project by August 4, 2012. As of July 19, 2012 progress reports continue to be
submitted but the draft license application has not been submitted.
Admiralty Inlet Tidal Energy Project (P-12690). On March 2, 2010, the Public Utility District
No. 1 of Snohomish County, Washington (SnoPUD) filed an application with FERC to renew its
preliminary permit for its Admiralty Inlet Tidal Energy Project. FERC issued the renewed
preliminary permit on July 8, 2010. This project would be located in Admiralty Inlet in the
northwestern portion of Puget Sound, between the Olympic Peninsula and Whidbey Island,
Washington. The proposed project consists, in part, of two 10-m, 500-kW capacity turbines to be
manufactured by Open Hydro. In October 2008, the DOE provided a $1.2 million grant to the
utility to support the tidal energy effort at Admiralty Inlet, and on September 9, 2010, the DOE
announced that it had awarded the project an additional grant of $10 million in funding, which
9-9

Ocean Tidal Energy

will be matched by $10.1 million in non-federal funding. SnoPUD submitted its draft license
application with FERC on December 28, 2009. On July 7, 2011, FERC published the notice
concluding the pre-filing process. The final license application was expected by August 31,
2011, and finally submitted on March 1, 2012. As of July 18, 2012 the project continues to
perform environmental analyses and work with agencies and various stakeholders to address
their comments. According to the projects presentation at the Global Marine Renewable Energy
Conference in April 2012, the future scheduled activities are as follows:

FERC pilot plant license was anticipated in September 2012.

OpenHydro will be awarded a supply an installation contract in October 2012.

The turbines will ship in July 2012.

Onshore construction, foundation fabrication, and subsea cable lay will occur in late 2012 to
early 2013.

Turbines will be installed on foundations and testing and tow trials will commence in
September 2013.

Deployment is expected in October or November 2013.

Commissioning will take place in January 2014.

U.S. Navy Demonstration Project in Puget Sound. Through a congressionally directed and
funded project, the Navy will install turbines from Verdant Power for a demonstration
installation in Puget Sound. Unlike the Snohomish PUD project, however, the primary goal of
the Navys pilot project is aimed at furthering the state of tidal energy technology by gathering
operational and environmental assessment data. The Navy plans to temporarily place three
Verdant Power tidal turbines in the waters on the west side of Admiralty Inlet off Marrowstone
Island. The turbines, which are to be installed in early 2013, will power two buildings and a
parking lot at the Navys ammunition depot on Indian Island, near Port Townsend. On October 4,
2012, the Federal Communications Commission (FCC) sent a letter to FERC requesting that
FERC review the project and determine that it does not present material risk to underwater
cables owened by Nippon Telegraph & Telephone Corp. of Tokyo. Verdant has modified its
installation plan to not include anchors during the installation plan.
9.2.2.3 Projects in California
San Francisco Bay Tidal Energy Project II (P-12585). On September 30, 2008, Golden Gate
Energy (GGE) filed with FERC an application for a pilot project license, which FERC denied.
On October 1, 2008, GGE was reissued a preliminary permit to evaluate the feasibility of the
project and prepare a license application. Once again, on February 4, 2010, GGE was reissued a
preliminary permit. As of June 21, 2012 FERC has granted GGE a time extension for filling the
draft license application.

9-10

Ocean Tidal Energy

9.2.2.4 Projects in New York


Roosevelt Island Tidal Energy (RITE) Project (P-12611). Since 2002, Verdant Power has been
gathering operational and environmental data from turbines at its Roosevelt Island site in the
East River in New York City. The current preliminary permit, issued February 17, 2009,
encompasses a proposed array of 30 turbines with an installed capacity of 1 MW in the East
Channel, and a proposed array of 100 turbines with an installed capacity of 4 MW in the West
Channel. Verdant submitted a final license application for the East Channel 1 MW commercial
pilot project on December 20, 2010. On May 3, 2011, FERC recommended licensing the project.
On December 22, 2011 the New York Department of Environmental Conservation granted
Verdant Power a water quality certification permit. On January 23, 2012 FERC granted Verdant
Power a 10-year hydrokinetic pilot project license, making it the first licensed tidal power project
in the United States. The project is being developed in a phased approach to include up to 30
turbines providing 1 MW of power.
The RITE project has received key support from the New York State Energy Research and
Development Agency (NYSERDA).
East River Tidal Energy Pilot Project (P-12665). New York Tidal Company was issued a
preliminary permit on January 10, 2011 for the East River Tidal Energy Tidal Project. The
location of the proposed project is east of Wards Island at Hell Gate in the East River, southwest
of the Triborough Bridge, in New York City. The proposed project would have a capacity of 20
kW. As of July 12, 2012 the project continues to file progress reports to FERC.
Wards Island Tidal Power (P-12718). Natural Currents Energy Services received a preliminary
permit for its Wards Island Tidal Power Project on April 17, 2009. The project would be a
hybrid, comprising solar, wind, and tidal power generators. The tidal power generators would
consist of eight 12-kW vertical-axis cross-flow turbines. On July 7, 2011 the project filed a draft
pilot license application, but after several correspondences FERC informed the project of the
applications dismissal on October 25, 2011. The project applied to reverse the dismissal on
November 11, 2011 and provided additional comments on January 4, 2012.
Astoria Tidal Energy Project (P-13730). On January 10, 2011, New York Tidal Energy
Company received a preliminary permit for its Astoria Tidal Energy Project. The project would
be located in an area of the East River in New York extending from north of Roosevelt Island
through an area known as Hell Gate, to Stony Point. The permit application specifies 50 to 150
turbine units of 0.5 to 2 MW each, for a maximum installed capacity of 300 MW. As of July 12,
2012 the project continues to provide progress reports to FERC.
Fishers Island Tidal Energy Project (P-14395). Natural Currents Energy Services applied for a
preliminary permit for its Fishers Island Tidal Energy Project on April 23, 2012. FERC approved
the preliminary permit on June 27, 2012 after adjustments were made to the application. The
project is located in Fishers Island, NY on the Long Island Sound. The project proposes to
install 50 Sea Dragon 100-kW units or approximately 5 MW of tidal power.
Orient Point Tidal Energy Project (P-14233). On May 17, 2012 Natural Currents Energy
Services received a preliminary permit from FERC for a project on the eastern end of the Long
Island Sound in Suffolk County, New York. The project would deploy 45 Natural Currents Sea
Dragon or Red Hawk 110-kW units.
9-11

Ocean Tidal Energy

9.2.2.5 Projects in Massachusetts


New England Marine Renewable Energy Center, Massachusetts Dartmouth Marine Energy
Research Center. On September 15, 2010, the New England Marine Renewable Energy Center
announced that it had received a federal grant of $750,000 for annual operations and $750,000
for study of advance techniques for assessing offshore wind and hydrokinetic renewable energy
resources. The researchers come from several local universities and institutions, including the
University of Massachusetts, MIT, the Woods Hole Oceanographic Institute, the University of
New Hampshire, and the University of Rhode Island.
Part of the centers goal is also to support early-stage ventures to commercialize technology spun
out of university research. It opened a clean energy laboratory in July 2009 to measure and test
prototype marine generation technologies. The center provides incubation space to three marine
energy firms: wave power technology developer Resolute Marine Energy Inc., tidal power
developer Ocean Renewable Power Corp., and underwater unmanned vehicle developer Ocean
Server Technology Inc.
In April 2012 the research center received a $535,000 state grant from the Massachusetts Clean
Energy Center. The grant will be used to support research of ocean technologies at multiple
stages of development in the centers 27-square-mile offshore area south of Massachusetts.
Muskeget Channel Tidal Energy Project (P-13015). The town of Edgartown, Massachusetts,
was issued a preliminary permit for the Muskeget Channel project on August 2, 2011. The
proposed project would comprise thirteen 380-kW OCGen (Ocean Renewable Power Company)
turbines, with a total installed capacity of 4.94 MW. As of February 2, 2012 the project continues
to submit progress reports to FERC.
Cape Cod Tidal Energy Project (P-13828). Free Flow Power received a preliminary permit for
its Cape Cod Tidal Energy Project on December 9, 2010. The project area is the Cape Cod
Canal, near the Massachusetts towns of Sandwich, Sagamore, and Buzzards Bay. The project has
a proposed capacity of up to 20 MW, comprising up to 2,000 10-kW turbines, each 3 m in
diameter. On October 31, 2011 Free Flow Power decided not to pursue licensing the project and
requested to surrender its preliminary permit. On February 2, 2012 FERC issued the notice of
surrender for the preliminary permit. Natural Currents Energy Services, LLC applied for a
preliminary permit on April 23, 2012 under the same project name (P-14394), but the permit was
rejected by FERC on July 3, 2012.
9.2.2.6 Projects in Maine
On April 9, 2009, a $951,500 federal appropriation to the University of Maine was announced.
Researchers there will lead a collaborative effort to develop Maines tidal power resource. The
funding comes from a congressional initiative developed by the Maine delegation, and will be
used to assess current prototypes and turbine models that can be submerged in the ocean to
produce power using tidal currents. Although the University of Maine will lead the project,
Maine Maritime Academy and Portland-based ORPC are partners in the ongoing research.
Researchers will also move forward to evaluate the potential environmental impact of harnessing
tidal energy off the coast of Eastport in the Western Passage of Passamaquoddy Bay.

9-12

Ocean Tidal Energy

In September, 2010, the U.S. Department of Commerce awarded $1.4 million to the city of
Eastport and Washington County to establish the Maine Marine Energy Center, an advanced
composite material manufacturing facility that can produce the complex components and subassemblies required by the ocean energy industry.
Ocean Renewable Power Company (ORPC) received $10 million in funding from DOE (to be
matched by $11.1 million in non-federal funding) to build, install, operate, and monitor a
commercial-scale array of five grid-connected devices in Cobscook Bay, Eastport, Maine. An
award of $1.2 million to ORPC from the Maine Technology Institute to support that work was
announced on October 14, 2010.
ORPC Cobscook Bay Tidal Energy Project (P-12711) and Western Passage Tidal Energy
Project (P-12680). ORPC is planning a three-part project, called the Maine Tidal Energy Project,
at the mouth of the Bay of Fundy (off Eastport and Lubec, Maine). The three-part project
includes Cobscook Bay, Kendall Head, and the Western Passage. ORPC filed a draft
hydrokinetic pilot license application with FERC on July 24, 2009 for both the Cobscook Bay
and Western Passage projects. In preparation for writing this application, ORPC developed
environmental study plans for its Eastport sites and has been collecting data from the sites
regarding tidal current velocity, marine geophysical characteristics, benthic characteristics and
terrestrial site review. ORPC was reissued preliminary permits for these two projects on
January 13, 2011. The Cobscook Bay Tidal Energy Project (P-12711) has a proposed capacity of
750 kW, and the Western Passage project (P-12680) has a proposed capacity of 1.2 MW. ORPC
completed fabrication, assembly and shop testing of its 60-kW, beta pre-commercial tidal
generator unit (TGU) in early 2010 and deployed the device in Cobscook Bay in March 2010.
On September 1, 2011 ORPC applied for a pilot project license for the Cobscook Bay project,
and on February 27, 2012 the license was issued by FERC. On April 26, 2012 the Maine Public
Utilities Commission approved the primary contract terms of power purchase agreements (PPAs)
for the ORPC Maine Tidal Energy Project. These will be the first long-term PPAs for tidal
energy in the United States. On July 25, 2012 ORPC held the dedication ceremony for the
project, making it the first commercial tidal energy project in the U.S. The project is under way
and ORPC installed a commercial TidGen82 Power System in the second half of 2012. On
September 13, 2012, the underwater turbine generator came on-line. The TidGen is designed to
have a peak output of 180 kilowatts. Currently in pilot phase, the project will eventually add two
additional TidGen turbines, each with a peak output of 180kw. The second phase of installation
will take place in the fall of 2012. The Western Passage Tidal Energy Project (P-12680) remains
in the preliminary permit phase. As of July 11, 2012 FERC granted the project a time extension
for filing the draft license application.
Town of Wiscasset Tidal Resources Project (P-13329). FERC issued the town of Wiscasset a
preliminary permit for the Wiscasset Tidal Resource Project on May 28, 2009. The preliminary
permit cites 4 to 40 OCGen hydrokinetic turbines with a total installed capacity of 1 to 10 MW;
however, the project proponents anticipate applying for a pilot project license, which limits the
project capacity to 5 MW. The project received a time extension from FERC for filing a draft
license application on June 15, 2011 and the most recent progress report was submitted on
November 1, 2011.
82

TidGen is a trademark of Ocean Renewable Power Company.

9-13

Ocean Tidal Energy

Homeowner Tidal Power Electric Generation Project (P-13345). On July 1, 2009, FERC issued
a preliminary permit for the Homeowner Tidal Power Electric Generation Project to be located
on the Kennebec River in Sagadahoc County, near Phippsburg, Maine. The proposed project
would comprise six hydrokinetic turbines with a combined capacity of 60 kW, in addition to
transmission lines connected directly to individual homes and appurtenant facilities. As of the
end of 2009, the project applicant anticipated pursuing a license exemption. On July 2, 2012 the
project applied for a subsequent preliminary permit for Kennebec River Pleasant Cove
Road/Woodland Lane Homeowners by Dot Kelly. On August 1, 2012 FERC notified the project
that the permit application was deficient.
Damariscotta River Hydrokinetic Tidal Energy Project (P-13646). FERC issued a preliminary
permit for the Damariscotta Tidal Project on May 19, 2010. The proposed project would be
located on the Damariscotta River in Lincoln County, Maine. The project would have a total
installed capacity of 250 kW provided by 10 to 20 EnCurrent hydrokinetic generator units.
FERC cancelled the preliminary permit on July 28, 2011, because the project applicant failed to
submit the required notice of intent (NOI) and pre-application document (PAD). As of August 2,
2012 no further updates have occurred.
Kendall Head Tidal Energy Project (P-13801). On January 13, 2011, FERC issued ORPC a
preliminary permit for the Kendall Head Tidal Energy Project. The project would consist of four
OCGen hydrokinetic tidal devices, each consisting of two 150-kW turbine generator units for a
combined capacity of 1.2 MW; an anchoring support structure; a mooring system; a 2,700-foot
submersed cable connecting the turbine generating units to a shore station; an 8,500-foot, 34.5kV transmission line connecting the shore station to an existing distribution line; and appurtenant
facilities. The project would be located in the Western Passage, Washington County, Maine. On
July 11, 2012 FERC issued a time extension to ORPC for filing the draft license application.
Treat Island Tidal Project (P-14330) and Lubec Narrows Tidal Project (P-14331). On April
19, 2012 FERC issued preliminary permits to ORPC for the Treat Island Tidal Project in
Passamaquoddy Bay, between the town of Eastport and Treat Island, Washington Count, Maine.
The project will consist of 15 TidGen devices, each 150 kW. On that same day FERC issued
another preliminary permit to ORPC for the Lubec Narrows Tidal Project to be located in Lubec
Narrows and Johnson Bay, near the Town of Lubec, Washington County, Maine. The project
will consist of 10 RivGen83 devices, each 60 kW.
9.2.2.7 Projects in New Hampshire
General Sullivan and Little Bay Bridges Tidal Energy Project (P-13503). The University of
New Hampshire Center for Ocean Renewable Energy (UNH/CORE) filed a preliminary permit
application for the General Sullivan and Little Bay Bridges Tidal Energy Project to be located on
the Piscataqua River in Rickingham and Stafford counties, New Hampshire. FERC issued the
preliminary permit on September 30, 2009. The proposed project comprises a single 10-footwide by 35-foot-long test platform suspending a variety of hydrokinetic devices into the river,
including a Gorlov Helical turbine generation unit, plus appurtenant facilities. Because electricity
generated by the system would power the test platform, the project would require no
transmission line. UNH/CORE intends to develop the site as a nationally recognized testing and
evaluation facility for marine hydrokinetic turbines, with no intention to connect with the electric
83

RivGen is a trademark of Ocean Renewable Power Company.

9-14

Ocean Tidal Energy

grid. Thus, UNH/CORE expects to seek an exemption from a FERC license. FERC requires that
projects receiving preliminary permits also intend to develop the site for commercial use. The
project is located in the town of Newington, NH, where the local Conservation Commission
refused to allow the project to be developed commercially. Therefore, on March 19, 2012 a
notice of surrender for the preliminary permit was issued by FERC.
9.2.2.8 Projects in New Jersey
Kingsbridge (Wills Hole) Tidal Electric Project (P-13247). On December 22, 2010, Natural
Current Energy Services filed a draft application for an original license for the Wills Hole
project. That draft application was dismissed by FERC on April 4, 2011. Reasons cited for
dismissal included incomplete consultation, no proposed post-license monitoring and safeguard
plans, and a general lack of detail in the draft application. On July 12, 2011, the applicant filed a
regular six-month progress report. The preliminary permit expired on November 30, 2011. It is
not clear whether the applicant has applied for a renewal of the preliminary permit. As of May 3,
2012 the project continues to file progress reports with FERC.
Hoffmans Marina Tidal Energy Project (P-13682). Natural Currents Energy Services filed an
application for a FERC-issued preliminary permit on May 5, 2010, for the Hoffmans Marina
Tidal Energy Project, which would be located on Manasquan River, in Monmouth County, New
Jersey. The proposed project would consist of 10 20-kW turbines with a combined capacity of
200 kW and would be operated to power the marina and associated office building on site. A
notice of surrender of preliminary permit was issued on September 26, 2011 for the project.
Highlands New Jersey Tidal Energy Project (P-13725). Natural Currents (NC) Energy Services
was issued a preliminary permit on January 11 2011, for the Highlands New Jersey Tidal Energy
Project, which would be located on the Shrewsbury River, in Monmouth County, New Jersey.
The project would utilize 20 150-kW Natural Currents Red Hawk Tidal In-Stream Energy
Conversion (TISEC) modules with a total combined capacity of 3 MW. The preliminary permit
was surrendered on January 11, 2012 following overdue progress report submissions. Following
this, Natural Currents Energy Services filed a new permit (P-14393) on April 23, 2012 for the
same site. FERC rejected the permit application on July 9, 2012 based on the projects previous
failure to submit progress reports on time.
Salem Tidal Energy (P-13849). Natural Currents Energy Services received a preliminary
permit for its Salem Tidal Energy Project on May 2, 2011. The proposed capacity of the project
is 3 MW, utilizing between 10 and 30 NC Sea Dragon or Red Hawk tidal turbines (rated capacity
100kW each). On June 25, 2012 FERC issued a letter to the project stating that the notice of
intent and pre-application document for licensing was overdue.
Cohansey River Tidal Energy (P-14127). FERC issued a preliminary permit to Natural Currents
Energy Services for its Cohansey River project on September 1, 2011. The authorized capacity
of the proposed project is 3 MW, utilizing between 10 and 30 NC Sea Dragon or Red Hawk tidal
turbines (rated capacity 100kW each). As of March 7, 2012 the project continues to submit
progress reports to FERC.

9-15

Ocean Tidal Energy

BW2 Tidal Energy Project (P-14222). Natural Currents Energy Services filed an application for
a preliminary permit on July 13, 2011, for a 1-MW capacity project on the Maurice River. The
proposed project consists of 2 NC Sea Dragon or Red Hawk tidal turbines, each with a rated
capacity of 150 kW. On February 3, 2012 the application for preliminary permit was accepted.
As of February 6, 2012 the project has submitted its 45-day report to FERC.
DorchesterMaurice Tidal Energy Project (P-14223). Natural Currents Energy Services filed
an application for a preliminary permit on July 13, 2011, for a 2-MW capacity project on the
Maurice River. The proposed project would consist of 1 to 10 NC Sea Dragon or Red Hawk tidal
turbines, at a rated capacity of 100 kW each. On February 3, 2012 the application for a
preliminary permit was accepted. As of February 6, 2012 the project has submitted its 45-day
report to FERC.
Margate Tidal Energy Project (P-14224). On July 14, 2011, Natural Currents Energy Services
filed an application for a preliminary permit for a 3 MW project north of Margate City. The project
would deploy 10 to 30 NC Sea Dragon 100-kW tidal turbines. The preliminary permit was issued
on January 18, 2012. On July 5, 2012 the project submitted its six-month progress report.
Avalon Tidal Energy Project (P-14228). Natural Currents Energy Services filed an application
for a preliminary permit on July 15, 2011, for a 3-MW capacity tidal project in Ingram
Thorofare, near Avalon, NJ. The proposed project would utilize between 10 and 30 NC Sea
Dragon or Red Hawk tidal turbines (rated capacity 100kW each). On February 3, 2012 the
preliminary permit was issued by FERC. On February 6, 2012 the project submitted its 45-day
report.
Cape May Tidal Energy Project (P-14232). On July 18, 2011, Natural Currents Energy Services
filed an application for a preliminary permit for a 3 MW capacity project on the Cape Bay Canal,
Cape May, NJ. The project would deploy 10 to 30 Natural Currents Sea Dragon 100-kW units.
The permit was issued on February 13, 2012 and on February 14, 2012 the project submitted its
45-day report.
Maurice River Tidal Energy Project (P-14234). Natural Currents Energy Services submitted an
application for a preliminary permit to FERC on July 18, 2011, for a 3-MW tidal energy project
on the Maurice River. The project would utilize between 10 and 30 NC Sea Dragon or Red
Hawk tidal turbines (rated capacity 100kW each). On March 12, 2012 FERC issued the
preliminary permit, and on March 16, 2012 the project submitted its 45-day report.
9.2.3 U.S. Developer and Project Deployment Highlights
9.2.3.1 Verdant Power
In December 2010, Verdant Power submitted a final license application to FERC to install and
operate 30 turbines (each turbine rated at about 34 kW for a total of roughly 1 MW). On May 3,
2011, FERC recommended licensing the Roosevelt Island Tidal Energy (RITE) Project. On
December 22, 2011 the NYS Department of Environmental Conservation granted Verdant Power
a water quality certification permit. On January 23, 2012 FERC granted Verdant Power a 10-year
hydrokinetic pilot project license, making it the first licensed tidal power project in the U.S. The
project is being developed in a phased approach to include up to 30 commercial class Free Flow
System turbines (5th generation) providing 1 MW of power.
9-16

Ocean Tidal Energy

9.2.3.2 Vortex Hydro Energy


Founded in 2004, Michigan-based Vortex Hydro Energy LLC manufactures a technology called
VIVACE (for vortex-induced vibration aquatic clean energy) that is based on the extensively
studied phenomenon of vortex-induced vibration (VIV). Usually something naval engineers seek
to suppress, VIVACE exploits this phenomenon to extract useful energy from ocean, river, tidal
and other water currents. VIV results from vortices forming and shedding on the downstream
side of a bluff body in a current. Vortex shedding alternates from one side to the other, thereby
creating a vibration or oscillation, as illustrated in Figure 9-5.

Figure 9-5
Vortex-induced vibrations oscillate objects in fluid currents

A prototype device was field-tested in the St. Clair River in spring 2010, and the company is
developing its next generation device for testing. The company is striving to commercialize the
technology in 2013 or 2014.
9.2.3.3 Ocean Renewable Power Company
ORPC is developing MHK tidal projects at two locations in the United States (Maine and
Alaska) and one location in Nova Scotia, Canada.
Maine updates: ORPC fabricated its 60-kW beta turbine generating unit in 2009 to early 2010 and
deployed it in March 2010 on a floating barge in Cobscook Bay, near Eastport, Maine. On October
14, 2010, ORPC was awarded $1.2 million to facilitate development of the TidGen Power System
from a beta pre-commercial status to a grid-connected and revenue generating commercial product.
This follows the award of $10 million from DOE to support a commercial-scale tidal project in
Cobscook Bay on September 9, 2010. On September 1, 2011 ORPC applied for a pilot project
license (which differs from a preliminary permit) for the Cobscook Bay project (part of the Maine
Tidal Energy Project), and on February 27, 2012 the license was issued by FERC. On April 26,
2012 the Maine Public Utilities Commission approved the primary contract terms of power
purchase agreements (PPAs) for the ORPC Maine Tidal Energy Project. These will be the first
long-term PPAs for tidal energy in the United States. On July 25, 2012 ORPC held the dedication
ceremony for the project, making it the first commercial tidal energy project in the U.S. The
project is under way and ORPC installed a commercial TidGen Power System in the second half of
2012. On September 13, 2012, the underwater turbine generator came on-line. The TidGen is
designed to have a peak output of 180 kilowatts. Currently in pilot phase, the project will
eventually add two additional TidGen turbines, each with a peak output of 180kw. The second
phase of installation will take place in the fall of 2012. The Western Passage Tidal Energy Project
(P-12680) remains in the preliminary permit phase. As of July 11, 2012 FERC granted the project
a time extension for filing the draft license application.
9-17

Ocean Tidal Energy

Alaska updates: ORPC is developing a project at the Cook Inlet in Alaska. Owing to special
environmental concerns about endangered species in the Cook Inlet, ORPC was awarded
$600,000 in 2009 by the U.S. DOE to monitor beluga whales.
Nova Scotia updates: ORPC is collaborating with Canadian independent power producer, Fundy
Tidal Inc., to develop a tidal energy project in the Bay of Fundy. The project plans to supply up
to 1.5 MW of power and begin delivering electricity to the local grid by mid-2013. The project
hopes to benefit from the worlds first Community Feed-In Tariff for tidal energy, which is
currently being finalized by the Province of Nova Scotia.
9.2.3.4 Free Flow Power
The first field demonstration of a tidal energy device in Massachusetts took place in Muskeget
Channel on five consecutive days between August 8 and 15, 2011. Free Flow Power also tested a
prototype during the summer of 2011 in the Mississippi River, near Baton Rouge, Louisiana.
9.2.3.5 UEK Corporation
University of Manitoba and Manitoba collaborated to test UEK Corporations Underwater
Electric Kite in the Winnipeg River, Pointe du Bois, Manitoba, Canada. The turbine was
deployed from a barge in early October 2011.

9.3 Worldwide Tidal In-Stream Energy Highlights: Late 2011 to Mid-2012


This section provides a brief summary of notable TISEC activities outside the United States
between late 2011 and mid-2012.
9.3.1 Canada
9.3.1.1 Fundy Ocean Research Centre for Energy Developments
The Fundy Ocean Research Centre for Energy (FORCE) has received $20 million from the
government of Canada, and signed an $11 million contract for the production and installation of
four subsea cables at its Minas Passage test site in 2011. The cables, with a combined capacity of
64 MW, will connect tidal devices to the Nova Scotia grid.
This facility will allow for the testing of in-stream turbine devices in tidal waters that are known
to possess the most demanding conditions in the world. Throughout the course of the test period,
the devices will be monitored for wear, environmental impacts and performance that will lead to
further technological developments.
Four tidal energy developers have been selected to install commercial prototype turbines at this
test facility in 2012: Clean Current of British Columbia, Open Hydro of Ireland, Marine Current
Turbines of the UK, and Atlantis Resources, also of the UK. The first device to be installed was
a 1-MW Open Hydro turbine, which was deployed in the fall of 2009. The device was removed
from the water after it was discovered that two of its blades had broken off.
The project sites were selected via a one-year, $1-million study. They lie on sediment-free
bedrock and have a linear current flow at speeds of up to 10 m/s during ebb and flood tide. The
estimated cost to deploy each project is $15 to $18 million and total project cost is estimated to
be $70 million. Project partners include Nova Scotia Power and Minas Pulp and Power.
9-18

Ocean Tidal Energy

In recognition of the common challenges faced at FORCE and the European Marine Energy
Centre (EMEC), as well as the benefits of collaboration, representatives of the two centers singed
a strategic agreement in May 2011.
On November 7, 2011 FORCE opened its $1.3 million visitor center in Parrsboro. The 3,000
square-foot facility overlooks the Cape Split tidal turbine demonstration site in the Minas
Passage. During that same month FORCE released the results of its first environmental
monitoring report. Additional studies were performed, including acoustic Doppler current
profiles, electromagnetic field risks to marine organisms, and observation surveys of seabirds
and mammals. Final reports for these studies will be available in 2012.
9.3.1.2 The Canoe Pass Tidal Energy Corporation (CPTEC)
CPTEC will apply a $2 million award from the Innovative Clean Energy Fund toward the
installation of two 250-kW EnCurrent turbines, developed by New Energy Corp of Calgary,
Alberta. The Canoe Pass site is adjacent to the foot of Seymour Narrows north of the Campbell
River. The project was undergoing environmental review in 2011. Device deployment is
anticipated in 2012. On May 31, 2012 CPTEC filed an application to the Minister of Transport,
Infrastructure and Communities under the Navigable Waters Protection Act for the approval of
plans and site work.
9.3.1.3 Ocean Renewable Power Company
In July, 2011, Maine-based ORPC formed ORPC Nova Scotia in order to take advantage of the
proximity and tremendous tidal energy potential of Nova Scotias Bay of Fundy waters. A
planned project at Petit Passage is just 50 miles from ORPCs Eastport, Maine, projects. ORPC
has formed a strategic partnership with Canadian independent power producer, Fundy Tidal Inc.,
to develop a 1.5 MW tidal energy project in the Bay of Fundy utilizing the robust tidal resource
of Digby Gut at the entrance to the Annapolis Basin. The project will use ORPCs TidGen Power
System and expects to be delivering electricity to the local grid by mid-2013. The project also
expects to benefit from the worlds first Community Feed-In Tariff for tidal energy development,
currently being finalized by the Province of Nova Scotia.
9.3.2 United Kingdom (UK)
The UK is maintaining its stature as the global leader in tidal energy technology development. In
an effort to solidify its leadership position in TISEC development, the UK has established an
installed marine energy capacity goal of 2 GW by 2020 and adopted an energy policy designed
to attract and support TISEC developers and equipment testing.
On March 16, 2010, the Scottish government and crown estate announced a 4 billion project to
build 10 tidal and wave power sites with a combined capacity of 1.2 GW in the Orkney Islands
and Pentland Firth. The projects are expected to require an additional investment of 1 billion to
construct new grid connections, ports, and other support facilities. These projects were described
as the worlds first commercial wave and tidal power schemes, with capacity evenly divided
between wave and tide-based generation. The tidal projects are listed in Table 9-2.

9-19

Ocean Tidal Energy


Table 9-2
Offshore in-stream tidal energy conversion device developers
Tidal Site

Developer

Technology Supplier

MW

Westray South

SSE Renewables
Developments

TBC

200

Cantick Head

SSE Renewables
Holdings

Open Hydro

200

Brough Ness

SeaGeneration Ltd.

Marine Current
Turbines

100

Ness of Duncansby

ScottishPower
Renewables

TBC

100

Inner Sound

MeyGen

Atlantis Resources

400

European Marine Energy Center Developments


The UK has a prototype test facility at the European Marine Energy Center (EMEC) in the
Orkney Islands that hosts many tidal energy developers. Marine currents at the test site reach
almost 4 m/sec (7.8 knots) at spring tides. The facility offers eight tidal test berths at depths
ranging from 12 m to 50 m. From each developer berth, subsea cables follow the seabed and then
pass under the beach and into an external housing next to the substation. An adjacent building on
site holds a Scottish and Southern Energy transformer, where 11 kV is transformed to 33 kV.
The Orkney Islands of Eday, Westray, and Sandy are linked by subsea cables that form a ring
through the Northern Isles and feed into the grid.
During 2011, EMEC added two wave and two tidal nursery sites that provide at-sea test
locations that are not exposed to full ocean conditions. The tidal nursery sites are located in
Shapinsay Sound. Unlike the tidal test site at Fall of Warness that is cabled to shore, the Shapinsay
Sound site employs test support buoys containing onboard load banks and uninterruptible power
supplies.
During November 2011, acoustic surveys were carried out at the Fall of Warness tidal site and
Shapinsay Sound to assess the effects that acoustic emissions may have on wildlife.
The first developer to use the site was OpenHydro, whose open center turbine was installed in
late 2006 (Figure 9-6). Testing took place throughout 2007 and grid connection occurred in late
2007. In November 2007, OpenHydro successfully changed out its turbines at the Fall of
Warness. In May 2008, the company successfully supplied electricity to the National Grid during
its testing procedure, marking the first time electricity had been produced and delivered to the
National Grid by a tidal energy device in the UK. An important upgrade to the tidal and wave
energy infrastructure at EMEC occurred with the delivery of 11 km of 20-kV subsea cable,
which allowed EMEC to expand its test sites to include three additional grid-connected test
berthstwo at the tidal test site and another at the wave test site.

9-20

Ocean Tidal Energy

On July 30, 2012 the Marine Energy Park was launched at EMEC to heighten the international
profile of the region and its reputation as a world leader in marine energy. In March of 2012
representatives from EMEC announced that they had signed a memorandum of understanding
(MOU) with the Ocean Energy Association of Japan (OEAJ) to develop a marine energy test
center in Japan. EMEC will provide design, development, and operation support for the new test
center.
As of April 2012 EMEC has 14 clients and remains the worlds only accredited, grid connected,
wave and tidal test site. Recent developer-specific activities at EMEC are described in the
following subsections, which summarize the recent activities of selected UK TISEC developers.

Figure 9-6
OpenHydro device at the EMEC Fall of Warness tidal test site

9.3.2.1 Marine Current Turbines (MCT)


Siemens acquired a stake in Marine Current Turbines in early 2010, and in late 2011 expanded
that stake to 45%. MCT successfully completed the deployment of its 1.2-MW SeaGen84 Tidal
System. The MCT prototype was deployed in April 2008 in Strangford Narrows (UK) and was
the worlds first demonstrated TISEC unit with a power rating exceeding 1 MW. The SeaGen
system (Figure 9-7) delivered electricity into the UK grid for the first time in early July 2008,
and has since delivered more than 2.7 GWh of power to the grid, thereby achieving a 66%
capacity factor. The power generated by SeaGen is being purchased by the Irish energy company
ESB Independent for its customers in Northern Ireland and the Republic of Ireland. SeaGen has
the capacity to generate power to meet the average electricity needs of around 1,000 homes.

84

SeaGen is a trademark of Marine Current Turbines.

9-21

Ocean Tidal Energy

Figure 9-7
Marine Current Turbines SeaGen

9.3.2.1.1 Marine Current Turbines - SeaGeneration (Kyle Rhea) Ltd.

SeaGeneration (Kyle Rhea) Ltd. is a development company set up by Marine Current Turbines
(MCT). The company is proposing to develop a tidal stream array at a site between the Isle of
Skye and the west coast of Scotland called Kyle Rhea. The project is currently in the process of
performing an environmental impact assessment (EIA) necessary before applying for a lease
from the Crown Estate. The process is expected to take one full year.
9.3.2.1.2 Marine Current Turbines -SeaGeneration (Wales) Ltd.

SeaGeneration (Wales) Ltd. is a development company set up by MCT and RWE npower
renewables. The company is proposing to develop a tidal stream array at a site off the northwest
coast of Anglesey referred to as the Skerries. Extensive feasibility assessments and
environmental studies have been carried out looking at the site and SeaGen Wales submitted an
application in early 2011 to the Welsh Assembly Government for consent to build and operate
the tidal stream array.
9.3.2.2 Pulse Tidal
The Pulse Tidal technology is based on twin hydrofoils positioned across the tidal flow. Moving
water pushes the foils either up or down according to the angle of the foil in the water. The
vertical forces act in the same way that air moving over a wing provides lift. A conventional
generator above the water surface is driven by the foils moving below the surface.
Pulse Tidal received funding from the UK Governments Technology Program to research and
develop a 100-kW Pulse generator in the Humber estuary close to Immingham (Figure 9-8). The
generator fed power ashore directly to supply a large chemical works.

9-22

Ocean Tidal Energy

The mean water level the Pulse Stream 100 operates in at Humber is only 9 m; with 4 m of tidal
range either side of that, the Pulse Tidal concept proved for the first time the potential for tidal
stream energy from shallow waters. Pulse Tidal has begun environmental studies at the planned
site of the worlds first commercial shallow water tidal energy site at Kyle Rhea, the strait
between the Isle of Skye and the Scottish mainland, and plans to start generating there in 2012.
On January 20, 2012 Pulse Tidal was selected to receive a $50,000 feasibility study grant from
USDA Rural Development for a hydrokinetic turbine project in Matanuska Susitna Borough,
Alaska.
In the spring of 2012 the Crown Estate awarded Pulse Tidal a lease agreement for a 1.2 MW fullscale demonstration tidal project to be deployed off Lynmouth in Devon, at the South West
Marine Energy Park. The deployment, planned for 2014, will be Englands largest tidal
generation machine.

Figure 9-8
Pulse Tidal generator

9.3.2.3 Swanturbines
A 300-kW demonstration device is being assembled to be installed at the European Marine
Energy Centre (EMEC) in Orkney. The Cygnet device is illustrated in Figure 9-9. This
7 million project will demonstrate the robust design and installation philosophy of the
Swanturbines technology. In early April 2010, trials confirmed the feasibility of deploying
Swanturbines Cygnet device at EMEC. On October 19, 2010 the project successfully completed
a test of installation methodology with Jumbo Offshore.

9-23

Ocean Tidal Energy

Figure 9-9
Swanturbines Cygnet

The GlaxoSmithKline Montrose project is a proposed tidal array project in the River South Esk
Estuary. Each of the 0.7-MW turbines will be part of a 15 turbine array that will be installed in
two phases over a period of 14 months. The project is intended to provide power to the GSK
Montrose site, which is a primary manufacturing and supply facility and a key provider of active
pharmaceutical ingredients for some of GSKs most therapeutically and commercially important
products. The project began in 2010 and has since completed environmental impact statements,
feasibility studies, and has secured a lease with the Crown Estate (November 30, 2011). On
March 20, 2012 Swanturbines submitted a license application for the project.
9.3.2.4 Tidal Generation Limited
Tidal Generation Limited (TGL), a wholly owned subsidiary of Rolls-Royce, deployed and
connected a prototype of its 500-kW tidal turbine at EMEC in 2010 (Figure 9-10). As of late
October 2011, the device had supplied over 100 MWh of electrical power to the national grid. In
March 2012 the company announced that its device has surpassed 200 MWh of power generation
fed into Scotlands national grid. TGL is currently constructing a 1-MW device to be installed at
EMEC in 2012. TGL then plans to install a 10 MW demonstration array in 2013/2014.

9-24

Ocean Tidal Energy

Figure 9-10
Tidal Generation Limited DeltaStream

9.3.2.5 Atlantis Resources Corporation


Atlantis Resources unveiled its AK1000 tidal turbine on August 12, 2010, and deployed it 12
days later at EMEC in 35-m water. The tandem arrangement of twin rotors on a single hub is
18 m in diameter and produces 1 MW at a water velocity of 2.65 m/s. The turbine was intended
to be deployed at EMEC and tested for three years; however, the device was removed from the
water later in 2010 because manufacturing defects in the blades caused them to delaminate.
Atlantis deployed its AR1000, a 1-MW single-rotor device, at EMEC during the summer of
2011. The device was connected to the grid on August 11, 2011, making it the worlds first gridconnected commercial scale tidal turbine. The AR1000 is also one of the largest turbines every
built, at 22.5 m high and weighing 1,5000 tonnes. The device will be tested at EMEC over the
next two years, and deployed commercially at Scotlands phased 400-MW MeyGen project in
Scotlands Pentland Firth. The $3 billion tidal array is due to be installed in 2014. The project
secured a 20-year lease from the Crown Estate in October 2010.
Atlantis Resources Corporation has been selected by the Energy Technologies Institute (ETI), a
public-private partnership, to head a world leading team to identify ways of providing cost
effective deployment of tidal stream technologies at commercial scale in UK water. The first 16month phase of the Tidal Energy Converter (TEC) System Demonstrator project will cost up to
3.2 million. The second phase will take approximately three years and cost approximately 10
million.
9.3.2.6 Hammerfest Strm
On August 17, 2010, Hammerfest Strm, a Norwegian company, announced contracts totaling
4 million to construct the first of 10 of its HS1000 tidal turbines at EMEC in 2011. The HS1000
is a 1-MW pre-commercial demonstrator. The company deployed an Acoustic Doppler Current
Profiler (ACDP) and turbulence meter at the site in preparation for deployment of their device. In
October the company commenced work on the subsea cable in preparation for deployment of
their device.

9-25

Ocean Tidal Energy

Hammerfest Strm was selected by ScottishPower Renewables for their Islay and Duncansby
Head sites totaling 105 MW. Tidal resource and seabed surveys have been completed, and the
environmental impact assessment has been submitted. The Scottish government approved the 10MW Islay project in March 2011. The 1 MW demonstration project on the island of Eday was
installed in December 2011 and on May 17, 2012 the project completed its initial testing period.
9.3.2.7 Bluewater
In November 2011, Bluewater Energy secured a test berth at EMEC for a full-scale version of its
floating BlueTEC tidal power turbine system.
9.3.2.8 Kawasaki
In late October 2011, Kawasaki Heavy Industries announced it had signed a contract for a test
berth at EMEC. It also announced it was about to begin testing a sub-scale device off of Okinawa
to support development of a 1-MW prototype for deployment at EMEC. The prototype is
being certified by GL Renewable Certification (GL RC) and the process began in April 2012.
The 1-MW device will be tested at EMEC in 2013.
9.3.2.9 Voith Hydro
In late July 2011, Voith Hydro began installing a monopole for its tidal turbine at EMECs Fall
of Warness tidal test site. The company plans to deploy a full-scale prototype of its HyTide
device in August or September 2012.
Voith Hydros first tidal current device prototype is in operation off the coast of South Korea in
the Sea Turtle Tidal Park. The device has a rated capacity of 110kW. When complete, the park
will have a total capacity of 150 MW.
9.3.2.10 Flumill
Flumill AS, a Norwegian company, delivered its first commercial tidal energy system at EMEC
in late 2011. The device will be tested at EMECs nursery tidal test site in Shapinsay Sound over
a period of three to six months. The device uses 30-40 m long glass fiber Helix Screws to
capture energy from tidal currents.
On March 28, 2012 Flumill received approval from the Norwegian Water Resources and Energy
Directorate to build a 5 MW tidal power pilot project outside Troms, at Rystraumen, a tidal
current in the fjord dividing the island Kvalya from the mainland.
9.3.2.11 Scotrenewables
The SR250 was constructed at Harland & Wolff in Belfast toward the end of 2010. Following
tow tests in the sheltered waters of Shapinsay Sound to simulate tidal flow, Scotrenewables
successfully generated power with its floating SR250 device at the Falls of Warness test site in
December 2011. The device is designed to facilitate deployment and retrieval by a modest-sized
boat. The device was successfully connected to the national grid at the European Marine Energy
Centre at the end of March 2011, making it the worlds first floating tidal turbine to export power
to the national grid. A two-year test program has now commenced and a 4-day deployment was
successfully completed in May 2012.
9-26

Ocean Tidal Energy

9.3.3 Ireland
9.3.3.1 OpenHydro
The development of OpenHydros Open-Centre technology began in the United States during the
early 1990s, and the company has over a decade of experience developing and testing the OpenCentre Turbine in marine conditions. OpenHydro was formed in 2005 after negotiating world
rights to the Open-Centre technology in late 2004. The Open-Centre Turbine is designed to be
deployed directly on the seabed. It can be located at depth and presents no navigational hazard.
OpenHydros installation in September 2008 (Figures 9-11 and 9-12) is the worlds first
successful deployment of a tidal turbine mounted directly on a gravity base on the seabed. The
installation took place at EMEC in Scotland.

Figure 9-11
OpenHydro Gravity Base installation

Figure 9-12
OpenHydro Gravity Base photo

To solve the problem of vessel availability and enable launches of the tidal power devices,
OpenHydro commissioned the worlds first specialist barge to install seabed-mounted tidal
turbines (Figure 9-13).

9-27

Ocean Tidal Energy

Figure 9-13
OpenHydro deployment vessel

OpenHydro unveiled its next-generation 10-m diameter, 1-MW Open Centre turbine in
September 2009 and deployed that turbine in the Bay of Fundys Minas Passage in November
2009, in partnership with Nova Scotia Power and the Fundy Ocean Research Centre for Energy
(FORCE). The device was fabricated and shipped to the Bay of Fundy from OpenHydros
Technical Center at Carlingford Lough, Ireland. The turbine was pulled from the water in May
2010, prior to its planned removal in the fall of 2010, when it was discovered that two of the
turbine blades had broken off.
On October 5, 2009, Open Hydro received notice of a 2 million grant from Sustainable Energy
Irelands (SEI) Ocean Energy Prototype Research and Development Programme. The grant will
be used to design and develop OpenHydros next generation 16-m Open-Centre turbine, subsea
base, and installation barge.
On March 16, 2010, OpenHydro and its partner, SSE Renewables, were awarded exclusive rights
to develop a 200-MW tidal energy farm at the Cantick Head site in the Pentland Firth off the
north coast of Scotland. On July 6, 2010, OpenHydro was awarded a 1.85 million grant from
Scotlands Wave and Tidal Energy: Research, Development and Demonstration Support
(WATERS) fund. The funds will be used to design, develop, manufacture, and test a power
conversion and control system for connecting arrays of tidal turbines to the grid.
Open Hydro also has demonstration projects under planning and development in Admiralty Inlet,
Puget Sound, Washington; at FORCE in the Bay of Fundy, Nova Scotia, Canada; and Alderney,
Channel Islands (a British Crown dependency), and Paimpol-Brehat in Brittany, France
(discussed in the next sextion).
9.3.4 France
Electricite de France (EDF) Group has selected Irish company OpenHydro Group to build the
first underwater tidal turbines at Paimpol-Brehat in Brittany, France. The effort is part of its pilot
project to build a tidal turbine farm to generate electricity from tidal energy.

9-28

Ocean Tidal Energy

Announced in October 2009, the OpenHydro partnership involves installing four 16-m, 2-MW
turbines. The first of the turbines to be assembled was towed from the DCNS shipyard in Brest
on August 31, 2011, for commissioning tests at sea prior to installation. The turbines will be
connected to the electricity grid beginning in 2012. The test unit at Paimpol-Brehat will
reportedly enable the technology to be tested in real-world conditions and permit assessment of
impacts on the marine environment. In the future, other technologies tested by EDF could be
deployed to take advantage of the Paimpol-Brehat regions powerful tidal currents.
In July of 2012 the European utility GDF Suez announced that they are exploring the feasibility
of installing two tidal test sites off the coast of France. The company hopes to install up to six of
Voith Hydros HyTide turbines at Raz Blanchard by 2015, providing up to 12 MW of generating
capacity. The company also hopes to establish a second test site at the Passage du Fromveur
using Sabellas D10 prototype by 2016.
9.3.5 Australia
Tenax Energy is developing tidal energy at sites in Australia. These locations were selected
based on high tidal velocity, water depth, and proximity to existing grid infrastructure. The initial
focus is on Clarence Strait in the Northern Territory. Two other sites are in Victoria and
Tasmania.
In early September 2010, Australias Northern Territory granted a site license to Tenax Energy
for a proposed 200-MW project in Clarence Strait between the northern coast of Australia and
the Tiwi Islands, near the city of Darwin. The license allows Tenax to undertake environmental
studies, which are needed to obtain other regulatory approvals. Tenax is considering OpenHydro
turbines for this project. As of late 2011, the only tidal power installation in Australia was
Atlantis Resources 150-kW AN400 device deployed at San Remo, Victoria, Australia.
9.3.6 New Zealand
In March 2011, New Zealands Minister of Conservation approved a proposed tidal power
project in the mouth of Kaipara Harbour in Northland, New Zealand. The project put forth by
Crest Energy would deploy up to 200 completely submerged devices with a total maximum
generating capacity of 200 MW.
9.3.7 India
Atlantis Resources Corporation is developing a tidal power project with Gujarat Power
Corporation Limited in the Gulf of Kutch in Gujarat India. The project team is current assessing
the viability of a 250 MW tidal turbine farm. The Gulf of Kutch is an eastwest-oriented
indentation north of Saurashtra Peninsula. It is known for its tidal range of up to 8 m and charts
indicate flow rates of up to 5 knots. The project was ratified by the Chief Minister of Gujarat in
November 2010 and pilot project funds were approved in April 2012. Deployment is expected to
begin in 2013.

9-29

Ocean Tidal Energy

9.3.8 Africa
In May 2012 the Florida-based MHK developer, Hydro Alternative Energy Inc. (HAE),
announced its plan to develop a 1 MW tidal power project in the South African Agulhas Current
near Durbin. The project will be in partnership with the South African municipality e Thekwini.
HAE plants to installed one of its Oceans tidal power units at a cost of approximately US$20
million.
9.3.9 Japan
In March 2012 representatives from EMEC announced that they had signed a memorandum of
understanding (MOU) with the Ocean Energy Association of Japan (OEAJ) to develop a marine
energy test center in Japan. EMEC will provide design, development, and operation support for
the new test center.

9.4 Tidal Power and Energy Resources


EPRI has researched available data sources and developed a methodology for estimating instream tidal current resource and performance [1], the results of which are available at
http://oceanenergy.epri.com/streamenergy.html. With DOE funding, Georgia Tech completed a
tidal energy resource assessment for the United States in July 2011. A geographic information
system-based website (www.tidalstreampower.gatech.edu/) provides access to the resulting data.
9.4.1 Data Sources
The National Ocean Service (NOS) Center for Operational Oceanographic Products and Services
(CO-OPS) collects and distributes observations and predictions of water levels and currents to
ensure safe, efficient, and environmentally sound maritime commerce. The center manages the
National Water Level Observation Network (NWLON) and a national network of Physical
Oceanographic Real-Time Systems (PORTS) in major U.S. harbors.
The CO-OPS NOS predictions page (http://co-ops.nos.noaa.gov/tide_pred.html) is separated
into two sections. The first, Tidal Predictions, provides tidal predictions at a few thousand
stations around the United States, Caribbean, and Pacific Islands. The second, Tidal Current
Predictions, provides tidal current prediction for a few hundred U.S. and Canadian tidal current
reference stations, as well as time and speed adjustment to allow the calculations to predict tidal
currents at a few thousand secondary stations. Both tide and tidal current predictions are
provided for the present calendar year. The predictions are updated in late October or early
November to include the following calendar year.
Current velocity tables may be found at http://co-ops.nos.noaa.gov/curents04/currpred.html.
Both tide and tidal current predictions are based on analyses of observations at reference stations.
Unlike tide stations, which are normally located along the shoreline, most tidal current stations
are located offshore in channels, rivers, and bays. Tidal current stations are often named for the
channel, river, or bay in which they are located or for a nearby navigational reference point.
Figure 9-14 shows the locations of the CO-OPS reference stations for tidal current prediction in
the United States.

9-30

Ocean Tidal Energy

Figure 9-14
United States tidal current reference stations

9.4.2 Tidal Power Flux and Average Annual Energy


The power density of ocean tides is expressed in kilowatts per square meter (kW/m2) of area
perpendicular to the direction of flow. Annual average power density ranges up to about
4 kW/m2 for a very fast-moving tidal stream.
To estimate mean annual power density, it is first necessary to know the speed of the surface
currents as a function of time. The power density calculation must account for change in current
speed with depth, and horizontal variability across the width of the channel. This calculation
yields an estimate for the mean annual, depth-averaged, width-averaged tidal stream power
density. Figure 9-15 shows a tidal current frequency distribution for a single transect in Maines
Western Passage.
In 2005, EPRI performed a site survey and assessment study for two U.S. states (Maine [2]
and Massachusetts [3]), two Canadian provinces (Nova Scotia [4] and New Brunswick), and
three other U.S. sites: Knik Arm, Alaska [5]; Tacoma Narrows, Washington [6]; and Golden
Gate, California [7]. In 2006, EPRI performed a tidal energy resource assessment for southeast
Alaska [8].

9-31

Frequency (hours per


year)

Ocean Tidal Energy

1000
800
600
400
200
0

Depth-Averaged Velocity
(mid-point of speed bin, m/sec)

Figure 9-15
Annual average tidal current speed probability distributions for Dog Island Transect,
Western Passage, Maine

Figure 9-16 shows the available tidal energy resource for the five U.S. feasibility study sites
and one additional site, North Inian Pass in southeast Alaska. North Inian Pass may be the largest
tidal energy site in North America.

Figure 9-16
Annual tidal energy resource for six U.S. tidal current sites

Following the site surveys, EPRI performed design, performance, cost, and economic feasibility
studies for five U.S. sites and two Canadian sites. The results of this work are documented in
reports available from EPRIs Ocean Energy website (http://oceanenergy.epri.com).

9-32

Ocean Tidal Energy

9.4.3 Extractable Tidal Power and Energy


Estimating the portion of tidal energy at a given site that can be extracted without causing
unacceptable ecological impact is far more complex than calculating the available tidal energy at
a given site. Since tidal energy technology is modular, energy extraction is scalable and can be
limited to avoid significant ecological impact. Bryden and Crouch [9] suggest that 10% of the
available energy can be extracted without undue modification of the flow characteristics,
whereas Black and Veatch [10] suggested a limit of 20%.
Recent research by Dr. Ian Bryden [9, 11], now at the University of Edinburgh, Scotland,
indicates that the extractable resource may be significantly greater than earlier reported,
depending on the characteristics of the site. Extraction of kinetic power from tidal streams alters
the tidal regime in an estuary by reducing flow volumes, constricting the tidal range, and altering
the timing of tidal events. However, the magnitude of these impacts depends strongly on the
level of power extraction, estuary geometry, the nonlinear dynamics of in-stream turbines, and
the natural tidal regime. An understanding of the various fluidic effects of large-scale kinetic
power extraction is an essential first step in a more detailed investigation of ecosystem impacts.
The modular nature of tidal energy technology also has an effect on natural processes. It would
be counterproductive to over-develop any particular tidal current resource because, at a certain
point, adding more turbines ceases to generate additional kinetic energy. The tidal stream is in
effect a self-regulating energy supply and any significant reduction in water transfer reduces the
energy capture.
EPRI sponsored a Ph.D. dissertation at the University of Washington on the Effect of Kinetic
Power Extraction [12] that resulted in publication of a peer-reviewed paper [13]. The
dissertation, published in early 2009, noted:
The hydrodynamic effects of extracting kinetic power from tidal streams present unique challenges
to the development of in-stream tidal power. In-stream tidal turbines superficially resemble wind
turbines and extract kinetic power from the ebb and flood of strong tidal currents. Extraction
increases the resistance to flow, leading to changes in tidal range, transport, mixing, and the kinetic
resource itself. These far-field changes have environmental, social, and economic implications that
must be understood to develop the in-stream resource.
This dissertation describes the development of a one-dimensional numerical channel model and its
application to the study of these effects. The model is applied to determine the roles played by site
geometry, network topology, tidal regime, and device dynamics. A comparison is also made
between theoretical and modeled predictions for the maximum amount of power which could be
extracted from a tidal energy site. The model is extended to a simulation of kinetic power
extraction from Puget Sound, Washington.
In general, extracting tidal energy will have a number of far-field effects in proportion to the level
of power extraction. At the theoretical limit, these effects can be very significant (e.g., 50%
reduction in transport), but are predicted to be immeasurably small for pilot-scale projects.
Depending on the specifics of the site, far-field effects may either augment or reduce the existing
tidal regime. Changes to the tide, in particular, have significant spatial variability. Since tidal
streams are generally subcritical, effects are felt throughout the estuary, not just at the site of
extraction.

9-33

Ocean Tidal Energy

9.4.4 Tidal Power Forecasting


Because tides can be computed from the relative positions of the Sun, Earth, and Moon, the
available tidal power as a function of time can be computed years in advance. Weather has only a
minor effect on tidal currents. The timing of the peak tidal power advances by 50 minutes each
day.

9.5 TISEC Research and Development


Tidal power research programs in industry, government, and at universities in the UK, Ireland,
France, Italy, Korea, Canada, Australia, New Zealand, and the United States over the past decade
have established an important foundation for the emerging tidal power industry.
9.5.1 Harnessing Tidal Energy
Tidal energy extraction is complex and many device designs have been proposed. It is helpful
to introduce these designs in terms of their physical arrangements and energy conversion
mechanisms. A hydrokinetic energy conversion device can be classified as one of the following
three types:

Axial flow: the axis of rotation is parallel to the direction of water flow.

Cross flow: the axis of rotation is perpendicular to the water stream and may be oriented at
any angle, from horizontal to vertical, with respect to the water surface.

Non-turbine: oscillatory hydrofoil, vortex induced motion, or hydro Venturi device.

Figure 9-17 illustrates axial and cross flow-types of turbines.

9-34

Ocean Tidal Energy


Axial and Cross-Flow Turbines

Marine Current Turbines Axial Turbine


(Courtesy Marine Current Turbines)

Lucid Energy Cross-Flow Vertical Turbine


(Courtesy Lucid Energy)

ORPC Cross Flow Horizontal Turbine


(Courtesy ORPC)

Figure 9-17
Tidal energy conversion system configurations

The subsystems for a hydrokinetic turbine typically include a blade or rotor, which converts the
energy in the water to rotational shaft energy; a drive train, which usually comprises a gearbox
and a generator; a support structure for the rotor and drive train; and other equipment, including
controls, cables, and interconnection equipment. Additional ways of classifying these devices
include the following:

Support structures: These may be either gravity-foundation-based, attached to a monopile


foundation, or anchored or moored and allowed to fly in the stream using buoyancy and/or
dynamic pressure forces. TISEC devices may also be hung from a floating platform that is
anchored and moored to maintain its location.

Open vs. ducted: Turbine rotors may be either open, much like wind turbines, or enclosed in
a duct or shroud. Because the energy is a function of the cross-sectional area of the flow
directed through the swept area of the turbine, using a duct of a given cross sectional area is
the equivalent of using an open rotor with the same cross-sectional area. The wind energy
industry has found that adding length to a rotor blade is more economical than adding a duct
to increase the cross-sectional area and power of a wind machine. However, the economics
may differ for wind turbines given the lower density of the fluid medium.
9-35

Ocean Tidal Energy

Fixed vs. variable pitch blades: Pitch control is used to limit power, maximize the efficiency
of the turbine in variable flows, and enable bi-directional operation. There are other ways to
accomplish these three functions with fixed-blade turbines and many different design
concepts for implementing them.

Closed-center vs. open-center hubs: Instead of a fixed hub and rotating blades, a design
variation uses an outer fixed rim and an inner rotating bladed disc. The potential benefit of an
open-center design is eliminating the need for a gearbox by encapsulating the stator of a
generator on the rim of the machine.

Savonius vs. Darrieus vertical-axis turbines: First invented in Finland, the Savonius turbine
is S-shaped when viewed along the axis of rotation. This drag-type turbine turns relatively
slowly, but yields a high torque. The Darrieus turbine was invented in France in the 1920s.
Often described as looking like an eggbeater, this type of turbine has vertical blades that
rotate into and out of the flow. Using hydrodynamic lift, these turbines can capture more
energy than drag devices.

Helical vs. cycloidal: Aerodynamic lift-type vertical axis turbine blade configuration. A
helical blade traces a three-dimensional curve that lies on a cylinder or cone, such that its
angle to a plane perpendicular to the axis is constant. A cycloidal blade has the shape of a
curve traced by a point on the circumference of a circle that rolls on a straight line.

There are other types being investigated, such as hydro Venturi and hydrofoil or vortex-induced
oscillation. However, they are not of sufficient practical importance at this time to be described
in this chapter. Information on these devices is contained in an EPRI report [14] on TISEC
devices available at EPRIs Ocean Energy website (http://oceanenergy.epri.com).
9.5.2 TISEC System Developers
Today, a number of companies are leading the commercialization of technologies to generate
electricity from tidal currents. Table 9-3 lists most known device developers as of November
2011; it excludes those with only conceptual designs and patents.

9-36

Ocean Tidal Energy


Table 9-3
Offshore in-stream tidal energy conversion device developers
Device Developer(1) & Website
AeroHydro Research and Technology
www.ahrta.com
Alstom-Clean Current
www.cleancurrent.com
Aquamarine Power
www.aquamarinepower.com
Atlantis Resources
www.atlantisresourcescorporation.com

Device Name
Unknown

Type
Oscillatory Wings

Development Status(2)
Experimental

Tidal Turbine
Generator
Neptune

Ducted Horizontal Axis

Early Commercial

Dual-rotor Axial Flow

Technology
Demonstration
1. Early Commercial
2. Laboratory Testing
3. Laboratory Testing

1. AR series
2. AS series
3. AN series

Not disclosed

1. Unducted axial flow


2. Ducted axial flow
3. Chain-linked
Aquafoils
Cross-Flow, Horizontal
Axis
Not disclosed

Technology
Demonstration
Experimental

BioStream

Oscillatory Biomimetic

Experimental

Vertical Axis Hydro


Turbine
BlueTEC

Ducted Cross-Flow

Experimental
Technology
Demonstration
Experimental

Aquantis

Vertical-Axis CrossFlow
Dual Ducted AxialFlow
Vertical-Axis CrossFlow
Axial Flow

Stingray

Oscillatory Wing

Sea Caisson and


Turbine System
Unknown

Ducted Axial-Flow

Technology
Demonstration
Laboratory Testing

Ducted Horizontal Axis

Early Commercial

Osprey
HS300
HS1000
Morild

Vertical Axis
Axial-Flow

Hydro-Gen

Axial-Flow

Experimental
Technology
Demonstration
Commercial
Demonstration
Technology
Demonstration
Early Commercial

Atlantisstrom
www.atlantisstrom.de
Aquascientific
http://aquascientific2.moonfruit.com
BioPower Systems
www.biopowersystems.com
Blue Energy Canada Inc.
www.bluenergy.com
Bluewater
www.bluewater.com
BluStream

Atlantisstrom

C-Energy
www.c-energy.nl
Ecomerit Technologies
www.ecomerit.com
Engineering Business
www.engb.com
Firth Tidal Energy
http://firthtidalenergy.com/home/
Free Flow Power
www.freefllowpower.com
Free Flow 69
Hammerfest STROM AS
www.hammerfeststrom.com
Hydra Tidal
www.hydratidal.com
Hydro-Gen

Wave Rotor

Hydrovolts
http://hydrovolts.com
Kepler Energy
www.keplernergy.co.uk
Lucid Energy
www.lucidenergy.com
Lunar Energy Ltd
www.lunarenergy.co.uk
Marine Current Turbines Ltd
www.marineturbines.com
Minesto Tidal Energy Solutions
www.minesto.com

Transverse
Horizontal Axis
Water Turbine
Gorlov Helical
Turbine (GHT)
Rotech Tidal Turbine
Seaflow and
SeaGen
Deep Green

Axial-Flow

Darrieus, Savonius,
and FlipWing blades
Horizontal Cross-Flow
Vertical Axis
Ducted Horiz. Axis
Axial-Flow
Kite-Mounted AxialFlow Turbine

Experimental
Experimental

Experimental
Technology
Demonstration
Technology
Demonstration
Commercial
Demonstration
Laboratory Testing

9-37

Ocean Tidal Energy


Table 9-3 (continued)
Offshore in-stream tidal energy conversion device developers
Device Developer(1) & Website
Natural Currents Energy Services
www.naturalcurrents.com
Nautricity
www.nautricity.com
Neptune Renewable Energy
www.neptunerenewableenergy.
com/tidal_technology.php
New Energy Corporation
www.newenergycorp.ca
Ocean Flow Energy
www.oceanflowenergy.com/
Oceana Energy
www.oceanaenergy.com/
Ocean Renewable Power Corp
www.oceanrenewablepower.com
OpenHydro
www.openhydro.com
Ponte de Archimede
www.pontediarchimede.it
Pulse Tidal
www.pulsegeneration.co.uk
Sabella
www.sabella.fr
Scotrenewables Tidal Power
www.scotrenewables.com
Seapower Intl AB
www.seapower.se
SMD Hydrovision
www.smdhydrovision.com
Swan Turbines
www.swanturbines.co.uk
Tidal Energy Ltd
www.tidalenergyltd.com
Tidal Energy Pty Ltd
http://tidalenergy.net.au
Tidal Generation
www.tidalgeneration.co.uk
Tidal Hydraulics
www.dev.onlinemarketinguk.net?THG?i
ndex.html
Tidal Sails
www.tidalsails.com
TidalStream
www.tidalstream.co.uk
Tocardo
www.tocardo.com
UEK
www.uekus.com
University of Southampton
www.epsrc.ac.uk/newsevents/news/200
6/Pages/tidalgenerator.aspx

9-38

Device Name
Sea Dragon
Red Hawk
CoRMaT
Proteus

Type

Axial-Flow Tandem
Contra-Rotating Rotor
Vertical-Axis CrossFlow

EnCurrent Turbine

Vertical Axis

Evopod

Axial-Flow

TIDES

Unknown

TidGen

Cross-Flow HorizontalAxis
Ducted Axial-Flow

Open Centre
Turbine
Kobold

Vertical Axis

Pulse Stream 100

Oscillating Wing

Sabella D03
Sabella D10
SR 250

Axial-Flow

EXIM

Vertical Axis

TidEL

Twin Axial-Flow

Swan Turbine

Axial-Flow

DeltaStream

Axial-Flow

Unknown

Ducted Vertical-Axis
Cross-Flow
Axial-Flow

Twin Axial-Flow

Development Status(2)
Commercial
Demonstration
Experimental
Commercial
Demonstration
Technology
Demonstration
Technology
Demonstration
Experimental
Technology
Demonstration
Commercial
Demonstration
Commercial
Demonstration
Technology
Demonstration
Experimental
Technology
Demonstration
Technology
Demonstration
Technology
Demonstration
Technology
Demonstration
Unknown
Experimental
Technology
Demonstration
Technology
Demonstration

Tidal Hydraulic
Generator

Horizontal Axis

Tidal Sail

Cross-flow Sails

Experimental

Triton

Axial-Flow

Tocardo

Axial-Flow

Underwater Electric
Kite
Integrated Tidal
Generator

Ducted Axial-Flow

Technology
Demonstration
Technology
Demonstration
Commercial
Demonstration
Experimental

Ducted Axial-Flow

Ocean Tidal Energy


Table 9-3 (continued)
Offshore in-stream tidal energy conversion device developers
Device Developer(1) & Website
Verdant Power
www.verdantpower.com
Voith Siemens
www.voith.com/press/546530.htm
Vortex Hydro Energy
www.vortexhydroenergy.com
Water Wall Turbine
www.wwturbine.com

Device Name
Free Flow Turbine

Type
Axial-Flow

Unknown

Horizontal Axis

VIVACE

Oscillatory Wing

Development Status(2)
Commercial
Demonstration
Technology
Demonstration
Laboratory Testing

Unknown

Not Disclosed

Laboratory Testing

Notes:
1. This list excludes individual inventors with conceptual-level technology.
2. The following definition of development status was used:

Laboratory testing stage

ExperimentalSubscale at-sea testing

Technology demonstration: Large engineering prototype at-sea testing whose purpose is to test for function and
performance

Commercial demonstration: Large manufacturing prototype at-sea testing whose purpose is to test for
commercial viability

Early commercial: Offering many units of large size for purposes of generating and selling the electricity
produced.

9.5.3 Survival in Storms and Hostile Marine Environments


Tidal energy conversion devices can be deployed entirely below the waters surface and thereby
avoid exposure to the full brunt of atmospheric storms. Other types of devices have
superstructures that pierce the surface; while such structures are exposed to atmospheric storms,
servicing of above-water equipment may be facilitated. The offshore oil and gas industry
provides valuable technological knowledge and experience that is applicable to the tidal energy
industry. The offshore oil and gas industry has demonstrated lifetimes of 50 years or more for
equipment employing anti-corrosion and anti-biofouling technologies.
9.5.4 Effect of Tidal Power Plants on the Environment
Given proper care in siting and scaling of projects, tidal in-stream power can be one of the more
environmentally benign electrical generation technologies. Early demonstration and commercial
tidal in-stream power plants should include rigorous monitoring to record both environmental
impacts as well as natural variation at nearby undeveloped sites.
Extensive in situ monitoring was conducted in support of Verdant Powers license application
for its RITE project. Anecdotal information from numerous temporary testing activities in the
United States, Canada, UK, and other countries has not observed any harm to aquatic life. TISEC
device blades rotate very slowly (around 10 rpm for an 18-meter diameter rotor) and the speed at
the tip of the blade is about 10 to 12 m/s (2227 mph). The devices are self-limiting in that any
faster speeds result in cavitation.
9-39

Ocean Tidal Energy

Studies of potential effects on marine life were initiated in 2008 for the MCT SeaGen project in
Strangford, UK. More recently, environmental studies have been undertaken to assess fish
interactions with Ocean Renewable Power Corporations beta TidGen turbine in Cobscook Bay.
While results of many environmental studies are proprietary, a body of literature is developing
that specifically addresses environmental impacts of tidal and other marine renewable energy
technology. A project undertaken by DOEs Pacific Northwest National Laboratory called
TETHYS seeks to compile and provide access to this developing body of knowledge. TETHYS
can be accessed online at: http://mhk.pnnl.gov/wiki/index.php/Tethys_Home.
Jacobson [15] described the challenges associated with environmental assessment of tidal and
wave energy technologies, and outlined a strategy for addressing them.
EPRI conducted desktop and laboratory flume studies of interactions between hydrokinetic
turbines and fish at Alden Research Laboratories in Holden, Massachusetts, and at the USGS
Conte Anadromous Fish Research Laboratory in Turners Falls, Massachusetts. The results of
these studies [16, 17], sponsored by industry and the DOE, are publicly available and
downloadable from EPRIs website. These studies found that the majority of fish in the
laboratory setting passed around the area swept by the turbine rotor if allowed to do so. Injury
and mortality rates were low for fish that were forced through the turbines.
9.5.5 Permits for Tidal Power Plants
The novelty of TISEC technology has to date triggered conservative evaluations and an
extensive approval process at the federal, state, and local levels. As of November 2011, there
were no FERC licenses issued for operation of tidal energy projects. Table 9-4 lists the pending
preliminary permits for tidal hydrokinetic projects; Table 9-5 lists the active, issued preliminary
permits. Many of these preliminary permits are have been reissued either because the prior
permit expired or to accommodate substantive changes in the project description. Preliminary
permits are valid for 36 months from the month of issuance. Of the permits listed in Table 9-5,
the Margate Tidal permit through the Orient Point Tidal permit were issued in 2012. Additional
information can be obtained at the FERC website: www.ferc.gov/industries/hydropower/indusact/hydrokinetics/permits.asp).
Table 9-4
Pending FERC preliminary permits for tidal current projects (as of August 2, 2012)
Docket No.

9-40

Project Name

State

Authorized
Capacity (kW)

Homeowner Tidal Power Electric


Generation Project

ME

P-13345
P-14394

Cape Cod Canal Tidal Energy Project

MA

5,000

P-14395

Fishers Island Tidal Energy Project

NY

5,000

60

Ocean Tidal Energy


Table 9-5
Issued FERC preliminary permits for tidal current projects (as of August 2, 2012)
Docket No.

Project Name

State

Authorized Capacity (kW)

P-12585

San Francisco Bay Tidal Energy Project

CA

10,000

P-13509

Turnagain Arm Tidal

AK

2,200,000

P-13605

Icy Passage Tidal

AK

300

P-13606

Gastineau Channel Tidal

AK

400

P-12690

Admirality Inlet Tidal Energy

WA

1,000

P-12687

Deception Pass Tidal Energy

WA

6,400

P-12679

Cook Inlet Tidal Energy

AK

1,000

P-12704

Half Moon Tidal Energy

ME

9,000

P-12665

Astoria Tidal Energy

NY

200

P-13730

Astoria Tidal Energy

NY

2,000

P-12680

Western Passage Ocgen

ME

1,200

P-13801

Kendall Head Tidal Energy

ME

1,200

P-13884

Pennamaquan Tidal Power Plant

ME

21,100

P-13821

East Foreland Tidal Energy

AK

100,000

P-13849

Salem Tidal Energy

NJ

3,000

P-13015

Muskeget Channel Tidal Energy

MA

4,940

P-14127

Cohansey River Tidal Energy

NJ

3,000

P-14224

Margate Tidal

NJ

3,504,000

P-14223

Dorchester-Maurice Tidal

NJ

1,000

P-14222

BW2 Tidal

NJ

300

P-14228

Avalon Tidal

NY

3,000

P-14232

Cape May Tidal Energy

NJ

3,000

P-14234

Maurice River Tidal

NJ

100

P-14330

Treat Island Tidal

ME

2,250

P-14331

Lubec Narrows Tidal

ME

600

P-14333

Orient Point Tidal

NY

110

9-41

Ocean Tidal Energy

9.5.6 Overview of Regulatory Status for Tidal Power Plants


The regulatory permitting and licensing processes associated with U.S. ocean wave and tidal
projects can be quite complex, lengthy, and costly. Indeed, regulatory issues may be the primary
barrier to ocean energy development in the United States.
Since 1920, construction and operation of a non-federal hydroelectric project in the United States
has required a license issued by FERC in accordance with the Federal Power Act. In 2004, as a
result of legal interpretation from FERC, unconventional tidal and wave energy hydroelectric
projects were placed under FERC jurisdiction. An EPRI report [18], available at EPRIs Ocean
Energy website (http://oceanenergy.epri.com/streamenergy.html), provides a detailed description
of the U.S. regulatory environment.
9.5.6.1 FERC Pilot Project License
In 2008, FERC rolled out a licensing process for hydrokinetic pilot projects tailored to meet the
needs of entities interested in testing new technology, including connection with the interstate
grid, while minimizing the risk of adverse environmental impacts. The goal of the pilot project
licensing process (PPLP) is to allow developers to test new hydrokinetic technologies, determine
appropriate siting of these technologies, and confirm their environmental effects, while
maintaining FERC oversight and agency input and reducing the licensing requirements
compared to what they would be for a commercial license of 30 to 50 years duration. Pilot
project licenses are restricted to projects of 5 MW or less. Licensed pilot projects must be
removed at the end of the license term (nominally 5 years) unless a commercial license is
obtained. Pilot projects must be thoroughly monitored and may be terminated prior to expiration
of the license term if unacceptable effects of the project are observed. A Pilot Project License is
not an option for proposed projects located at environmentally sensitive sites. On July 29, 2012
the Bureau of Ocean Energy Management (BOEM) and FERC announced revised guidelines for
MHK developers pursuing permitting and licensing on the Outer Continental Shelf (OCS). The
revisions further clarify the regulatory process and help streamline the process for authorizing
research and testing of MHK devices.

9.6 Design, Performance, Cost, and Economic Feasibility


9.6.1 TISEC Sites
EPRI has investigated the attributes required for a good tidal energy site. Reports that document
the EPRI site assessments for Maine [2], Massachusetts [3], Nova Scotia [4], and New
Brunswick [19] are available at EPRIs Ocean Energy website
(http://oceanenergy.epri.com/streamenergy.html).
There are many factors to consider when evaluating potential sites for a tidal energy project.
Primary factors include the following:

High annual current flow measured by both flow volume and velocity

Proximity to a harbor/marina with sufficient depth, size, and infrastructure to support


assembly and deployment of the plant as well as maintenance operations

9-42

Ocean Tidal Energy

Proximity to a transmission and distribution system that can transfer power from the tidal
plant to a substation interconnection point

Water depth that allows the device to be mounted on the seabed with sufficient overhead
clearance for navigation, if required

Minimal potential for conflicting use of sea space

Receptivity of the local community

Availability of a local labor force to be trained for employment in this new industry

9.6.2 TISEC Device Assessments


For the purpose of this assessment, a number of cost centers were defined that would allow
sufficient detail to provide a solid understanding of the impacts of the major cost drivers. The
following are cost centers that contribute to capital cost estimates in this analysis:

Installation: Includes all activities associated with the plant construction, including:
transport, underwater cable installation, mooring installation, device deployment and
commissioning. The plant is assumed to terminate on shore. No cost is included for the
interconnection of the plant with the electrical grid.

Permitting and environmental compliance: Includes all cost components required to permit a
plant. Uncertainties in this cost category are high because there is limited experience to draw
from. These costs include site baseline assessments, environmental baseline studies, and
consultant fees.

Infrastructure: All costs except for those of the devices and moorings themselves, including
underwater cables, cable landing to shore, dockside improvements, and specialized vessels
used to operate the project.

Mooring: Includes all components required to keep the device on-station. These may involve
mooring chains, anchors, piled foundations, shackles and other mooring components.

Structural: Includes the main structural components, typically built from structural steel,
concrete or fiberglass.

Power take-off: Includes all components and subsystems required to convert the primary
mechanical energy into electricity that can be fed into the electric grid. Subsystems include
items such as a generator, gearbox, hydraulic system, frequency converter, step-up
transformer, and electrical riser cable.

9-43

Ocean Tidal Energy

Operating and maintenance (O&M) costs are divided into the following cost centers. To
normalize costs from operational activities, such costs are expressed as present value, divided by
the plants operational life and rated capacity to yield a unit of dollars per kilowatt capacity per
year ($/kW-year). The following are cost centers that contribute to O&M cost estimates in this
analysis:

Insurance: Insurance costs for one-off off-shore installations are typically on the order of 2%
of capital cost. As technology matures and technology-related risks are reduced, such
insurance costs are expected to decline to a level similar to that of wind energy today.

Marine monitoring: It is expected that many early-adopter projects will require ongoing
monitoring of the plants environmental effects to satisfy regulatory agencies. This may
include measures such as active and passive acoustic monitoring, fish studies, and sediment
transport studies.

Operations: This cost center contains all expenses associated with the operation, maintenance
and repair of the plant. This category includes labor cost, fuel cost, management expenses
and facility leases.

Replacement parts: Includes the cost of all parts that require replacement of the life of a
project.

The first analysis summarized here assesses the cost and performance of a specific TISEC device
in a variety of applications and environments. That analysis is followed by another completed in
2011 for a generic wave energy conversion device located off the coast of northern California.
When possible, the cost, performance and economic assessments presented are based on actual
testing and pilot-plant data to inform and drive techno-economic models.
9.6.2.1 Assessment for Selected Tidal Power Plants in Maine and Washington
EPRI performed a point design for both a single-unit demonstration and a commercial wave
power plant for sites in five U.S. states (Alaska [5], Washington [6], California [7],
Massachusetts [20], and Maine [21]) and two Canadian provinces (New Brunswick [22] and
Nova Scotia [23]). These designs were used to estimate cost and performance. Performance
estimates were developed using tidal current data predictions obtained from NOAA. Cost
estimates were developed by creating a detailed breakdown of the various cost centers and
outlines of installation and operation procedures, and by cross checking them with a variety of
sources, including local operators, the design team, local manufacturers, and similar offshore
projects in the oil and gas and offshore wind industries.
EPRI independently estimated plant system costs based on the Marine Current Turbine SeaGen
dual 18-m diameter rotor device design for the five tidal sites in the United States and the two
tidal sites in Canada. Table 9-6 summarizes the cost estimates for both small- and large-scale
TISEC plants at two of the seven sites: Tacoma Narrows in Washington State and the Western
Passage between Maine and New Brunswick.

9-44

Ocean Tidal Energy


Table 9-6
Cost and performance estimates for selected tidal power plants
Rated Capacity
Plant Capacity (number of units size, MW)
Annual Energy (MWe/year)
Plant Nameplate Rating (MW)

0.83 MW
Western
Passage ME

10 MW
Western
Passage ME

0.72 MW
Tacoma
Narrows WA

45.8 MW
Tacoma
Narrows WA

1 0.83

12 0.83

1 0.716

64 0.716

3,480

41,732

1,875

120,000

0.83

10

0.75

45.8

SeaGen

SeaGen

SeaGen

SeaGen

Seabed, m2

500

125,000

500

750,000

Unit Life, years

20

20

20

20

Note 1

Note 1

Note 1

Note 1

12

24

12

24

Dec 2009

Dec 2006

Dec 2006

Dec 2006

On shore transmission and grid I/C

266

592

524

84

Subsea cables

144

23

28

20

1,576

987

1,576

729

571

558

1,018

932

1,806

345

1,042

229

1,921

655

2,223

496

Construction mgmt and commissioning

Note 3

Note 3

Note 3

Note 3

Contingencies 3

Note 3

Note 3

Note 3

Note 3

Total Plant Cost (TPC)

6,284

2,679

6,411

2,493

535

195

544

186

6,819

2,874

6,955

2,679

Technology Description Axial Open Rotor MCT


Physical Plant

Scheduling
Development time, months1
Construction time, months
Capital Cost ($/kW)
Month/Year Dollars2

Power conversion modules


Structural sections
Subsea cable installation
Turbine installation
3

AFUDC (interest during construction)


Total Plant Investment
Due diligence, eng & permitting, etc.

Total Capital Requirement

341

7,160
5

144

348

3,018

7,303
5

134

2,813

N/A

4.2

N/A

3.7

95

98

95

98

Technology Development Rating

Pre Coml

Pre Coml

Pre Coml

Pre Coml

Design & Cost Estimate Rating

Simplified

Simplified

Simplified

Simplified

Yearly O&M Costs (% of TPC per yr)


Unit Availability (%)
Confidence and Accuracy Rating

Notes:
1. Development time for permitting is an unknown at this early point with emerging ocean energy technology.
2. The costs are in December 2005 dollars in References 10 through 16 and were adjusted by 2.5%% inflation to
Dec 2009.
3. Construction management, commissioning, and contingency costs are built into each of the subsystems.
4. Assumed to be 5% of TPI; cost of permitting is an unknown at this early point with emerging ocean energy
technology.
5. O&M costs for a pilot plant cannot be estimated.

9-45

Ocean Tidal Energy

These cost estimates are in December 2009 constant dollars, are based on an assumed set of
financial incentives, and do not include permitting and engineering costs. To avoid significant
environmental impacts, the energy extraction was conservatively limited to 15% of the available
energy resource.
9.6.2.2 Assessment for Typical TISEC Device Installed in Washington State
The following cost and economic assessments are based on a hypothetical installation at a
reference site and the following table shows the relevant parameters of the reference deployment
site [29]. In contrast to the assessment above, this analysis assumes generic wave power arrays of
5 MW, 20 MW, and 50 MW operating in an ocean environment with the following
characteristics:

Location: Tacoma Narrows, Washington

Mean current velocity at hub height: 1.12 m/s

Mean power density at hub height: 1.47 kW/m2

Mean water depth at deployment site: 60 m

Vertical power law profile: 1/7

Distance to Shore: 1 km.

Table 9-7 summarizes the results of the assessment, with costs for each size plant broken down
by capital expenses (CAPEX in $/kW), O&M expenses (OPEX in $/kW-year), plant
performance, and the resulting cost of electricity (cents/kWh). CAPEX includes the costs of
permitting and environment, installation, infrastructure, mooring, structural components, and
power take-off equipment. OPEX includes the costs of insurance, marine monitoring,
replacement parts, and operations.
Figure 9-18 graphically summarizes the CAPEX results for the three sizes of TISEC plants,
whereas Figure 9-19 does the same for O&M.

9-46

Ocean Tidal Energy


Table 9-7
Typical cost, performance and economic profiles for tidal power plants (2011$)
Plant Capacity
5 MW

20 MW

50 MW

Power Take-off

$1493

$1382

$1315

Structural

$789

$657

$583

Infrastructure

$3005

$800

$344

Installation

$2966

$1108

$580

Permitting & Environment

$1276

$349

$140

$9528

$4295

$2962

Operations

$94

$48

$24

Replacement Parts

$19

$19

$19

Marine Monitoring

$197

$49

$20

Insurance

$161

$69

$38

$470

$185

$100

4.9%

4.3%

3.4%

Capacity Factor

30%

30%

30%

Availability

95%

95%

95%

CAPEX Contribution

40

18

12

O&M Contribution

24

64

28

18

CAPEX ($/kW)

TOTAL
O&M ($/kW-year)

TOTAL
O&M Percent of CAPEX
Performance

Cost of Electricity (cents/kWh)

TOTAL

9-47

Ocean Tidal Energy

$12,000
$10,000

CAPEX($/kW)

$8,000

Permitting&Environment
Installation

$6,000

Infrastructure
Structural

$4,000

PowerTakeOff
$2,000
$0
5MW

20MW

50MW

Figure 9-18
Capital cost breakdown for a typical tidal power technology
$500
$450

O&M($/kWyear)

$400
$350
$300

Insurance

$250

MarineMonitoring

$200

ReplacementParts

$150

Operations

$100
$50
$0
5MW

20MW

50MW

Figure 9-19
Operational cost breakdown for a typical tidal power technology

Tidal current velocity and corresponding power density are highly localized and can vary
significantly over a few hundred meters. To illustrate the sensitivity of a plants levelized cost of
electricity (LCOE) to the power density at the site, the velocity distribution during the
performance assessment was simply scaled to evaluate the technologys sensitivity to the
resource strength (Figure 9-20). This assessment does not consider regional differences in
diurnal inequalities and other site-specific resource differences, but can serve as an indicator to
evaluate the sensitivity of LCOE to resource strength.

9-48

Ocean Tidal Energy


30

CostofElectricity(cents/kWh)

25
20
15
10
5
0
0.5

1.5

2.5

3.5

4.5

PowerDensity(kW/m^2)

Figure 9-20
Cost of electricity as a function of power density

9.6.3 TISEC Economic Feasibility


EPRIs models assume that initial early-adopter issues have been resolved and reliability similar
to that of comparable technologies has been obtained. Because real-world experience is limited,
typical cost and performance estimates must also be developed from technical specifications
supplied by leading device manufacturers. It is important to recognize that there is significant
diversity in technical approaches in this emerging field. It is also important to note that system
performance is heavily influenced by site conditions; resource power density tends to be the
dominant driver of the COE.
Manufacturers typically underestimate cost in the early stages of development. As the
technologies approach commercial maturity, such cost projections increase. The actual
construction and operational costs of a pilot device reveal a more complete economic picture and
provide a solid starting point for further cost studies. Once a technology reaches commercial
maturity, economy of scale applies and volume production begins to drive down costs. This
typical projected cost trajectory (Figure 9-21) can make it difficult to compare early-stage
technologies with more developed options. Based on its extensive experience with emerging
energy technologies, EPRI has developed a cost estimate rating methodology which assesses the
likely range of uncertainty based on a technologys design maturity and the amount of detail
available for the cost estimate. To reduce the uncertainty in this assessment, only data from
devices that are at a fairly advanced stage of development are used. As a result, all the cost data
given in this report have an estimated uncertainty range of 30%.

9-49

Ocean Tidal Energy

Lab/Idea

Prototype

Volume
Production

Commercial

Cost

Stage of Development

Figure 9-21
Cost projection as a function of development status

Table 9-8 describes the economic results for the five U.S. feasibility evaluation sites. These
cost and COE estimates were escalated from 2005 to December 2009, assuming an escalation
rate of 2.5%/year. The LCOE is in 2009 constant dollars and is based on an assumed set of
financial incentives. The capital cost estimate includes a 5% allowance for due diligence and
permitting costs and excludes engineering costs. In two cases, California and Maine, the 15%
extraction limit could not be reached because of the relatively small high-current area limiting
the number of existing technology turbines which could be deployed.
Table 9-8
Cost estimates for selected U.S. feasibility evaluation sites (Dec 2009$)
AK
Knik Arm

WA
Tacoma
Narrows

CA
Golden
Gate

MA
Muskeget
Channel

ME
Western
Passage

50

46

44

10

14.6

13

16.5

1.6

4.6

Number of Turbines

66

64

40

12

Total Plant Cost ($M)

122

114

100

19

27

Level O&M Costs ($M/yr)

4.5

4.2

4.0

0.6

1.1

Annual Energy (GWh)

128

120

129

1.5

41.7

Utility Gen COE (cents/kWh)

10.2

10.0

7.3

9.5

6.2

Muni Gen COE (cents/kWh)

7.8

7.9

5.4

6.6

4.6

Rated Power (MW)


Annual Average Power (MW)

9-50

Ocean Tidal Energy

TISEC technology has similarities with wind technology and has benefited from learning
experiences and cost reductions that have occurred in that industry. Additional reductions from
the TISEC cost shown in the table above will be realized through value engineering and
economies of scale.

9.7 Environmental Impact Issues


Given proper care in scale, siting, deployment and operation, EPRI expects that tidal in-stream
power could be one of the most environmentally benign electricity generation technologies.
Early demonstration and commercial offshore tidal power plant projects should include rigorous
monitoring of the environmental effects of the plant and similar rigorous monitoring of a nearby
undeveloped site in its natural state (so that natural variation can be distinguished from TISECinduced effects).
Between late 2009 and late 2011, three major reports on the environmental effects of tidal instream energy have been published in the United States [2426]. One of the reports, Potential
Environmental Effects of Marine and Hydrokinetic Energy Technologies, was prepared in
response to the Energy Independence and Security Act (EISA) of 2007. Section 633(b) of the
EISA called for a report to be provided to Congress that addresses (1) the potential
environmental impacts of marine and hydrokinetic energy technologies; (2) options to prevent
adverse environmental impacts; (3) the role of monitoring and adaptive management; and (4) the
necessary components of an adaptive management program.
The EISA-mandated report to Congress was prepared based on a review of peer-reviewed
literature, project documents, and U.S. and international environmental assessments of new
technologies. The information was supplemented by contacts with technology developers,
experts in state resource and regulatory agencies and non-governmental organizations, and input
and reviews by federal agencies (NOAA Fisheries, Minerals Management Service, U.S. Fish and
Wildlife Service, National Park Service, Bureau of Indian Affairs, Federal Energy Regulatory
Commission). Excerpted findings from the report follow.
There are numerous conceptual designs for converting the energy of waves, river and tidal currents,
and ocean temperature differences into electricity. Most of these technologies remain at the
conceptual stagethey have not yet been tested in the field or as prototype, full-scale devices.
Consequently, there have been few studies of their environmental effects. Most considerations of
the environmental impacts have been in the form of predictive studies and environmental
assessments that have not yet been verified.
The assessments have identified common elements among these technologies that may pose a risk
of adverse environmental effects. These potential impacts include the alteration of currents and
waves; alteration of substrates and sediment transport and deposition; alteration of habitats for
benthic organisms; noise during construction and operation; emission of electromagnetic fields;
toxicity of paints, lubricants, and antifouling coatings; and interference with animal movements
and migrations. Project installation and operation will change the physical environment. Effects on
biological resources could include alteration of the behavior of animals, damage and mortality to
individual plants and animals, and potentially larger, longer-term changes to plant and animal
populations and communities. Some effects are expected to be minor, but the potential significance
of many of the environmental issues cannot yet be determined owing to a lack of experience with
operating projects.

9-51

Ocean Tidal Energy

Although there have been few environmental studies of these new concepts, a preliminary
indication of the importance of each of these issues can be gained from published literature related
to other technologies, e.g., noises generated by similar marine construction activities, EMF
emissions from existing submarine cables, and environmental monitoring of active offshore wind
farms. Experience with other, similar activities in freshwater and marine systems will also provide
clues to effective impact minimization and mitigation measures that can be applied to these new
renewable energy technologies. However, some aspects of the environmental impacts are unique to
the technologies, and will require operational monitoring to determine the seriousness of the
effects. This is particularly true for the cumulative effects of large numbers of ocean energy or
hydrokinetic devices that will comprise fully built-out projects. Impacts to bottom habitats,
hydrographic conditions, or animal movements that are inconsequential for a few units may
become serious if large, multi-unit projects exploit large areas in a river, estuary, or nearshore
ocean. For some environmental issues it will be difficult to extrapolate predicted effects from small
to large numbers of units because of complicated, non-linear interactions between the placement of
the machines and the distribution and movements of aquatic organisms. Assessment of these
cumulative effects will require careful environmental monitoring as the projects are deployed.
Evaluation of monitoring results might be usefully conducted in an adaptive management
framework. There are numerous state and federal agencies and environmental laws and regulations
that will influence the development of marine and hydrokinetic technologies. Federal licensing of
these renewable energy projects is the responsibility of the Federal Energy Regulatory Commission
and the Minerals Management Service. Their licensing decisions will include input from other
federal and state agencies, tribes, environmental groups, and other stakeholders. After a licensing
decision has been made and operation of the energy project has begun, the identification (and
correction) of environmental impacts will depend on appropriate monitoring.
The ability to modify the project in order to mitigate unacceptable environmental impacts identified
by operational monitoring might be based on application of adaptive management principles
reflected in the project license conditions. In the context of marine and hydrokinetic energy
technologies, adaptive management is a systematic process by which the potential environmental
impacts of installation and operation could be evaluated against quantified environmental
performance goals during project monitoring. Early information about undesirable outcomes could
lead to the implementation of additional minimization or mitigation actions which are subsequently
re-evaluated. An adaptive management process is particularly valuable in the early stages of
technology development, when many of the potential environmental effects are unknown for
individual units, let alone the eventual build out of large numbers of units. Basing the
environmental monitoring programs on adaptive management principles, as advocated by many
resource and regulatory agencies, will take advantage of ongoing research and monitoring to help
refine technology designs and to improve environmental acceptability of future installations.

The second major report [24], sponsored by DOE, sought to develop a framework for identifying
key environmental concerns associated with marine renewable energy projects, including TISEC
projects. The third report [26] focused on environmental effects of tidal energy development. All
three of these reports constitute scoping documents that identify the range of issues potentially
faced by the industry as a whole and by developers of individual projects. Much work remains to
be done to narrow the scope of the data-gathering requirements placed on project developers
[15].

9-52

Ocean Tidal Energy

9.8 Installed Capacity and Estimated Growth


To date, worldwide installed TISEC capacity is about 6 MW. The devices are developmental and
commercial prototypes. Many commercial projects are in the planning and permitting stage. As
of November 2011, the installed TISEC capacity in the United States was 60 kW (the testing of
two 34-kW Verdant turbines in the East River in New York is complete and the turbines have
been removed). Table 9-9 presents EPRIs estimate of tidal in-stream capacity (in MW) that will
come on line (i.e., be commissioned and selling electricity into the grid to end users) in the
United States from 2012 through 2015. It assumes that licensing and permitting hurdles are
overcome, and governmental incentives and support are similar to those available in the recent
past.
Table 9-9
Installed and planned U.S. tidal energy capacity
Developer

Capacity (MW)

Project Name-Site

2012

2013

2014

2015

Verdant

RITE; East River, New York

0.110

ORPC

Cobscook Bay and Western


Passage; Eastport, Maine

0.150

Snohomish PUD

Admiralty Inlet, Washington

ORPC

Cook Inlet, Alaska

U.S. Navy

0.1

TOTAL NEW

YEARLY CAPACITY

0.260

4.1

5.5

0.5

4.6

5.6

11.1

CUMULATIVE

In general, EPRI expects that tidal energy will experience a growth rate that is constrained by
regulatory barriers and funding. EPRI estimates that there could be 500 MW of tidal energy plant
capacity by 2025. If Alaska is able to harness its tidal energy resource in this time period, this
estimate would be much larger (Alaska is believed to have more than 90% of the total U.S. tidal
energy resource). If small (tens of kilowatts) hydrokinetic turbines rated at lower speeds (less
than 2 m/s) become economical, the number and distribution of viable sites could increase.

9.9 Research Focus


Currently, DOE national laboratories are focusing on six critical research areas:
1. Physical Interactions with Devices
a. Fish/marine mammal attraction and avoidance
b. Strike risk to fish/marine mammals
2. Electromagnetic Fields
a. Effects of EMF on fish and marine mammals
9-53

Ocean Tidal Energy

3. Acoustics
a. Effects of MHK noise in riverine environments
b. Noise measurement and net-pen studies of select species
4. Toxicity
a. Effects of antifouling coatings on aquatic organisms
5. Benthic Habitat Alteration
a. Development of measurement methodology to evaluate effects of MHK devices
on benthic habitats
6. Data Aggregation and Risk Modeling
a. Development of a publicly available information database for MHK
environmental research
These national laboratories include the following:

Argonne National Laboratory

Idaho National Laboratory

National Renewable Energy Laboratory

Oak Ridge National Laboratory

Pacific Northwest National Laboratory

Sandia National Laboratories

In addition to the research being conducted by the DOE national laboratories, the department has
also funded the development of National Marine Renewable Energy Centers. As of 2012 the
department has helped establish three National Marine Renewable Energy Centers: Northwest
National Marine Renewable Energy Center (NNMREC), Hawaii National Marine Renewable
Energy Center (HINMREC), and Southeast National Marine Renewable Energy Center
(SNMREC).
The Northwest National Marine Renewable Energy Center is a DOE-funded partnership between
Oregon State University and the University of Washington. Oregon State University is
responsible for wave energy R&D, whereas the University of Washington is responsible for tidal
energy R&D. The Hawaii National Marine Renewable Energy Center is a partnership between
the DOE and the University of Hawaii, and focuses on wave energy R&D and Ocean Thermal
Energy Conversion (OTEC) R&D. The Southeast National Marine Renewable Energy Center is
a partnership between the DOE and Florida Atlantic University, and focuses on OTEC R&D and
ocean current energy R&D.

9-54

Ocean Tidal Energy

9.10 Conclusion
Considerable potential exists for generating electrical power from tidal energy in the United
States and many other places in the world. However, most of the U.S. tidal energy potential
exists in the state of Alaska, where the demand is low and where there is no transmission
infrastructure to bring the energy into the lower 48 states. There are only a few excellent tidal
sites in the lower 48 states. Those known to EPRI are the following:

Golden Gate, California: Excellent tidal resource (both volume and power density), adjacent
to a large load center with grid infrastructure, a nearby port, and high electricity prices.

Puget Sound, Washington: Excellent tidal energy resource (both volume and power density)
at a few sites, adjacent to a large load center with grid infrastructure, nearby port, but low
electricity prices.

Western Passage, Maine: Excellent tidal energy resource (both volume and power density),
adjacent to transmission infrastructure for delivering the power to large load centers, nearby
port, and high electricity prices.

Verdant Power believes that the East River in New York is a good tidal site, as do other firms. It
remains to be seen how FERC will treat multiple licenses on a single river, given that the
projects may have impacts on each other.

9.11 Internet Resources


EPRI tidal power reports are available at www.epri.com/oceanenergy.
Department of Trade and Industry (UK) reports are available at www.dti.gov.uk\\renewable.

9.12 References
1. Methodology for Estimating Tidal Current Energy Resources and Power Production by Tidal
In-Stream Energy Conversion (TISEC) Devices. EPRI, Palo Alto, CA: 2006. EPRI-TP-001NA (Revision 3).
2. Maine Tidal In-Stream Energy Conversion (TISEC): Survey and Characterization of
Potential Project Sites. EPRI, Palo Alto, CA: 2006. EPRI-TP-003 ME Rev 1.
3. Massachusetts Tidal In-Stream Energy Conversion (TISEC): Survey and Characterization of
Potential Project Sites. EPRI, Palo Alto, CA: 2006. EPRI-TP-003-MA (Revision 1).
4. Nova Scotia Tidal In-Stream Energy Conversion (TISEC): Survey and Characterization of
Potential Project Sites. EPRI, Palo Alto, CA: 2006. EPRI-TP-003-NS (Revision 2).
5. System Level Design, Performance, Cost and Economic AssessmentKnik Arm Alaska Tidal
In-Stream Power Plant. EPRI, Palo Alto, CA: 2006. EPRI-TP-006-AK.
6. System Level Design, Performance, Cost and Economic AssessmentTacoma Narrows
Washington Tidal In-Stream Power Plant. EPRI, Palo Alto, CA: 2006. EPRI-TP-006-WA.
7. System Level Design, Performance, Cost and Economic AssessmentSan Francisco Tidal
In-Stream Power Plant. EPRI, Palo Alto, CA: 2006. EPRI-TP-006-SF-CA.
9-55

Ocean Tidal Energy

8. Tidal In-Stream Energy Resource Assessment for Southeast Alaska. EPRI, Palo Alto, CA:
2006. EPRI-TP-003-AK.
9. Bryden, I. and S.J. Couch. 2006. ME1marine energy extraction: tidal resource analysis.
Renewable Energy 31(2): 133-139.
10. Black and Veatch. 2004. UK, Europe, and Global Tidal Energy Resource Assessment. The
Carbon Trust, London. Marine Energy Challenge Report No. 107799/D/2100/05/1.
11. Bryden, I.G. and S.J. Couch. 2007. How much energy can be extracted from moving water
with a free surface: a question of importance in the field of tidal current energy? Journal of
Renewable Energy 32(11): 1961-1966.
12. Polagye, B. 2009. Hydrodynamic Effects of Kinetic Power Extraction by In-Stream Tidal
Turbines. Doctoral dissertation. University of Washington, Seattle, WA.
13. Polagye, B., P.C. Malte, M. Kawase, and D. Durran. 2008. Effect of large-scale kinetic
power extraction on time-dependent estuaries. Proceedings of the Institution of Mechanical
Engineers, Part A: Journal of Power and Energy 222(5): 471484.
14. Survey and Characterization. Tidal In Stream Energy Conversion (TISEC) Devices. EPRI,
Palo Alto, CA: 2005. EPRI-TP-004-NA.
15. Jacobson, P.T. 2011. Challenges and opportunities in tidal and wave power. In: K.R. Rao
(ed.). Energy and Power Generation Handbook. American Society of Mechanical Engineers,
New York.
16. Fish Passage Through Turbines: Application of Conventional Hydropower Data to
Hydrokinetic Technologies. EPRI, Palo Alto, CA: 2011. 1024638.
17. Evaluation of Fish Injury and Mortality Associated with Hydrokinetic Turbines. EPRI, Palo
Alto, CA: 2011. 1024569.
18. Instream Tidal Power in North America: Environmental and Permitting Issues. Prepared by
Devine Tarbell & Associates, Inc., EPRI, Palo Alto, CA: 2006. EPRI-TP-007-NA.
19. New Brunswick Tidal In-Stream Energy Conversion (TISEC): Survey and Characterization
of Potential Project Sites. EPRI, Palo Alto, CA: 2006. EPRI-TP-003-NB (Revision 1).
20. System Level Design, Performance, Cost and Economic AssessmentMassachusetts
Muskeget Channel Tidal In-Stream Power Plant. EPRI, Palo Alto, CA: 2006. EPRI-TP-006MA. June 10, 2006.
21. System Level Design, Performance, Cost and Economic AssessmentMaine Western
Passage Tidal In-Stream Power Plant. EPRI, Palo Alto, CA: 2006. EPRI-TP-006-ME.
22. System Level Design, Performance, Cost and Economic AssessmentNew Brunswick Head
Harbour Passage Tidal In-Stream Power Plant. EPRI, Palo Alto, CA: 2006. EPRI-TP-006NB..
23. System Level Design, Performance, Cost and Economic AssessmentMinas Passage Nova
Scotia Tidal In-Stream Power Plant. EPRI, Palo Alto, CA: 2006. EPRI-TP-006-NS.

9-56

Ocean Tidal Energy

24. Kramer, S., M. Previsic, P. Nelson, and S. Woo. 2010. Deployment Effects of Marine
Renewable Energy Technologies: Framework for Identifying Key Environmental Concerns in
Marine Renewable Energy Projects. Prepared by RE Vision Consulting, LLC for DOE. RE
Vision DE-003.
25. USDOE (U.S. Department of Energy). 2009. Report to Congress on the Potential
Environmental Effects of Marine and Hydrokinetic Energy Technologies. Energy Efficiency
and Renewable Energy, Wind and Hydropower Technologies Program, Washington, DC.
December 2009.
26. Polagye, B., B. Van Cleve, A.E. Copping, and K. Kirkendall. 2011. Environmental Effects of
Tidal Energy Development. Seattle, WA.
27. Prioritized Research, Development, Deployment and Demonstration (RDD&D) Needs:
Marine and Other Hydrokinetic Renewable Energy. EPRI, Palo Alto, CA: 2008. 1014762.
28. UKERC (UK Energy Research Centre). 2008. UKERC Marine (Wave and Tidal Current)
Renewable Energy Technology Roadmap. Summary Report.
29. Previsic, M., Cost Profiles of Marine Hydrokinetic (MHK) Technologies, September 2011.

9-57

10

OCEAN WAVE ENERGY

10.1 Introduction
Ocean waves are generated by the influence of wind on the ocean surface (Figure 10-1). Ripples
on the surface create a steep slope against which the wind can push and cause waves to grow. As
the wind continues to blow, the ripples become chop, fully developed seas and, finally, swells. In
deep water, the energy in waves can travel for thousands of miles until their energy is dissipated
on distant shores.
Individual waves represent an integration of all winds encountered as they travel over the ocean
surface. Sea states (comprising wave height, period, phase and direction) can be accurately
predicted more than 48 hours in advance. This predictability, along with the relatively small shortterm variability of the available resource, will allow time for grid operators to dispatch other
generation resources to balance demand with supply. This is a major advantage over wind and solar
generation. Other characteristics of wave energy that make it especially attractive for electricity
generation are its high power density (kW/meter of wave crest length) and the potential for being
relatively environmentally benign if properly sited, sized, deployed, operated and maintained. See
Table 10-1 for an overview of ocean wave energy.
Table 10-1
Ocean wave energy overview
Installed Offshore
Wave Capacity
(as of November
2011)

Approximately 4 MW worldwide, and less than 0.1 MW in the United States.


Estimated annual incremental U.S. capacity additions:
2012: 0.150 MW
201315: 13.30 MW
EPRI estimates a U.S. cumulative capacity of approximately 10,000 MW by
2025
According to EPRIs 2011 assessment of the U.S. ocean wave energy
resource, the present outer shelf estimate of the total available wave energy
resource in the U.S. is approximately 2,640 TWh/yr. The majority of the
resource is located in Alaska.

Wave Energy
Conversion (WEC)
Technology
Readiness

WEC is an emerging technology. About a dozen full-scale prototype WEC


devices have been demonstrated at sea over the past five years, as well as
another dozen sub-scale prototypes.
Several commercial, grid-connected demonstration projects are currently
operating worldwide.

From 2009 to 2011 Ocean Power Technologies (OPT) tested its 40 kW


PowerBuoy in 30 m depth, mile offshore at the Kaneohe Bay Marine
Corps Base on Oahu, becoming the first grid-connected wave energy
device in the United States.

10-1

Ocean Wave Energy


Table 10-1 (continued)
Ocean wave energy overview
Economic Status

The DOE is funding a large multi-organizational project from May 2010 to


September 2013 to develop Reference Models. According to a DOE
presentation, the goal of the project is to develop a representative set of
Reference Models (RM) for the MHK industry to develop baseline costs of
energy (COE) and evaluate key cost component/system reduction
pathways. The Point Absorber Wave Energy Converter (WEC) report was
due September 30, 2011. The Oscillating Water Column WEC and Surge
Type WEC reports are due September 30, 2012.

Environmental Impact

Proper care in siting, installation, operation, and decommissioning may


enable ocean wave energy technology to be one of the more
environmentally benign electricity generation technologies.
Pilot demonstration testing is needed to understand the interactions
between the devices and their environment. Adaptive management will be
used to acquire and incorporate new information into decision-making
processes that will address project build out and cumulative effects.
The DOE provides funding to six national laboratories to conduct
environmental research in six critical areas: Physical Interactions
with Devices, Electromagnetic Fields, Acoustics, Toxicity, Benthic
Habitat Alteration, Data Aggregation, and Risk Modeling.
Additionally, the DOE funded the creation of three National Marine
Renewable Energy Centers in partnership with local universities,
including Northwest National Marine Renewable Energy Center
(NNMREC), Hawaii National Marine Renewable Energy Center
(HINMREC), and Southeast National Marine Renewable Energy
Center (SNMREC).
The U.S. Department of Energy and the Pacific Northwest National
Laboratory developed TETHYS, a database and knowledge management
system that provides access to information and research pertaining to the
potential environmental effects of marine and hydrokinetic (MHK) and
offshore wind development. TETHYS also hosts data from Annex IV, an
international collaboration to gather information on MHK environmental
research worldwide.

10-2

Ocean Wave Energy


Table 10-1 (continued)
Ocean wave energy overview
Regulatory Status

Numerous wave power plant projects have been permitted in Europe and
Australia.
On August 13, 2012 FERC granted Ocean Power Technology a commercial
license for the full build-out of the 1.5 MW, grid-connected wave power
station. This is the first commercial license issued for a wave power project
in the U.S. The license provides approval for the deployment of up to ten
grid-connected OPT devices for 35-years. OPT is working through the
permitting and licensing process to expand this site to 50 MW capacity. This
expansion project is called Phase III, and received a preliminary permit on
March 15, 2011 (P-13666).
The time, cost and complexity of the U.S. regulatory process can be difficult
for wave power project developers.
The Federal Energy Regulatory Commission (FERC) has primary
jurisdiction for licensing ocean wave energy under the Federal Power Act
(FPA), both in state waters and on the Outer Continental Shelf (OCS) ,
which lies 3 to 200 nautical miles off shore except in the Gulf of Mexico,
where it is 12 to 200 nautical miles..
FERC has developed a six-month license application process for pilot
demonstration plants.
The 2005 Energy Bill gave the Mineral Management Service (MMS) (now
BOEM) jurisdiction to lease lands on the Outer Continental Shelf (OCS 3
to 200 nautical miles offshore, except for the Gulf of Mexico, where it is 12
to 200 nautical miles).
FERC will no longer issue preliminary permits for wave plants on the OCS
but will continue to do so for wave plants in state waters.
BOEM lease rules for alternative energy on the OCS were issued in April
2009. There currently are no leases for wave plants on the OCS nor are
there any applications for leases for wave plants on the OCS.
On July 29, 2012 the Bureau of Ocean Energy Management (BOEM)
(formerly BOEMRE) and the Federal Energy Regulatory Commission
(FERC) announced revised guidelines for MHK developers pursuing
permitting and licensing on the OCS. The revisions further clarify the
regulatory process and help streamline the process for authorizing research
and testing of MHK devices.

10-3

Ocean Wave Energy


Table 10-1 (continued)
Ocean wave energy overview
Government Support
of Wave Energy
Technology

European governments (particularly in the UK, Ireland, Portugal and


Denmark) as well as those in Japan, New Zealand, and Australia, support
the development and deployment of WEC technology and are now
providing subsidies to stimulate a commercial market.
DOE initiated a Waterpower R&D Program in FY 2008 with a
congressionally mandated $10 million. FY 2012 received a budget increase
totaling $59 million ($34 million for MHK and $25 million for conventional
hydro). The FY 2013 budget proposal for the water program is only $20
million, which represents a 66% reduction from the FY 2012 budget.
So far in 2012 five Funding Opportunity Announcements have been issued
for water power technologies. The In-Water Wave Energy Conversion
(WEC) Device Testing Support FOA is particularly significant for the wave
energy industry because it offers $500,000 to one project to deploy and test
a buoy (point-absorber) WEC for one year at the Navy's Wave Energy Test
Site (WETS) off of Marine Corps Base Hawaii in Kaneohe Bay, Oahu. The
award was announced on May 18, 2012 and the submission deadline was
June 18, 2012.

Trends to Watch

Economical power from ocean waves requires the very best engineering
skills. In addition, resolution of sea space conflicts will require
communication outreach and skillful negotiation.
Demonstration projects and early commercialization projects, including
multi-megawatt wave farms, are expected to be deployed over the next
decade in Europe, South America, and Australia (and the United States if
regulatory obstacles are overcome).
Final reports for DOE Reference Models are expected by September 30,
2013.

Ocean waves are composed of orbiting particles of water, as illustrated in Figure 10-2. At the sea
surface, the diameter of water particle orbits is equal to wave height. Orbital motion decays
exponentially with depth, and its amplitude is only 4% of its surface value at half a wavelength
below the surface. The wave orbital motions are not significantly affected by the bottom in water
deeper than half a wavelength. The vector field of particle motion is illustrated in Figure 10-3.

Figure 10-1
Wind blowing over fetch of water,
producing waves

10-4

Figure 10-2
Particle motion in different water depths

Ocean Wave Energy

Figure 10-3
Vector field for particle motion in waves [1]

In deep water, wavelength is directly proportional to wave period squared. Therefore, the
wavelength of a 10-second wave is four times longer than that of a 5-second wave, and it will
begin to interact with the seabed in water that is four times as deep. Since the rate at which a
wave travels (its phase velocity) is equal to wavelength divided by period, a 10-second period
wave travels twice as fast as a 5-second wave. The combined potential and kinetic energy of the
waves travels at the velocity of the wave group, which in deep water is equal to half the phase
velocity.
Wave-power linear density (kilowatts per meter [kW/m] of wave crest width) is defined as the
flux of energy across a vertical plane intersecting the sea surface and extending to the depth of
no subsurface orbital motion (which is half the wavelength of the longest harmonic component).
For a 16-second wave, this depth is 200 m, which is the approximate depth of the continental
shelf edge.
The power of ocean waves is expressed in kilowatts per meter of wave crest front. Figure 10-4
depicts the power flux as the energy that crosses through a vertical plane 1 meter in width.
Annual averages range from 10 kW/m to 100 kW/m, depending on site location.

Figure 10-4
Wave power flux

10-5

Ocean Wave Energy

10.2 U.S. Wave Energy Highlights: Late 2011 to Mid-2012


Interest in ocean wave renewable energy continues to grow in the United States. This section
presents a brief accounting of notable federal and state developments as well as device
manufacturer activities within the sector from late 2011 through mid-2012.
10.2.1 U.S. Federal Highlights
The following sections present the latest marine and hydrokinetic (MHK) federal developments
with regard to congressional appropriations, awards issued, and regulatory updates.
10.2.1.1 Federal Appropriations
In FY 2012, the U.S. Congress appropriated $59 million to the Department of Energy Water
Power Program to conduct research and development (R&D) on advanced water power energy
generation technologies, including both MHK technologies ($34 million for wave, tidal, ocean
current, in-stream hydrokinetic and ocean thermal) and conventional hydropower ($25 million
for any hydropower technology that uses a dam or diversionary structure). This marks a
significant increase in funding from the $10 million appropriated in 2008, which was the first
year of the Water Power Program. However, the FY 2013 budget proposal announced in
February 2012 allotted only $20 million to the Water Power Program, representing a 66%
reduction from the previous year. Fortunately, in April 2012 the House and Senate
Appropriations Committee passed the FY 2013 Energy and Water Development Funding Bill.
The Senate funding measure provides $59 million to the program ($44 million for MHK and $15
million for conventional hydropower) and the House measure provides $45 million to the
program ($25 million for MHK and $20 million for conventional hydropower).
The DOEs MHK energy research is focused on assessing the potential recoverable energy from
MHK resources in the United States, and facilitating the development and deployment of
technologies to fully realize this potential. MHK technologies represent an opportunity for the
United States to engage directly in an emerging area of science and discovery, while developing
an entirely new suite of renewable energy technologies available to reduce emissions and help
states meet renewable energy portfolio standard (RPS) targets.
The DOEs priorities for marine power include the following:

Facilitating the deployment of prototypes, and collecting data on the energy conversion
performance and environmental impacts of the devices

Determining the available, extractable, and cost-effective resources in the United States

Characterizing and comparing the wide variety of existing MHK technologies

Improving technology performance and reliability, and reducing technology development


costs

Minimizing the cost, time, and negative impacts associated with siting projects

10-6

Ocean Wave Energy

10.2.1.2 Funding Awards


In 2008, the majority of DOE funding for wave power was awarded to specific technology and
project development efforts, selected through a competitive process. These awarded efforts
included the following:

Preparation of detailed design, manufacturing, and installation drawings of a bi-directional


air turbine for application in a floating oscillating water column

Engineering design, baseline environmental studies, and license construction and operation
applications to help Pacific Gas and Electric (PG&E), the largest investor-owned utility in
California, develop a hub to deploy wave energy converters and connect them to the grid

In addition, DOE selected and funded two National Marine Renewable Energy Centers: one at
the University of Hawaii, and a second run jointly by Oregon State University and the University
of Washington. Further, the DOE funded a market acceleration program, consisting of
nationwide wave and tidal hydrokinetic resource assessments, and a collaborative project to
address navigation and environmental issues as well as clarify the permitting process. EPRI was
selected by the DOE to conduct the national wave energy resource assessment. EPRI completed
the national wave energy resource assessment in late 2011. A report of the assessment [2] is
publicly available for download from EPRIs website. Related maps and data can be accessed at
the National Renewable Energy Laboratorys (NRELs) online Renewable Energy Atlas.
On August 3, 2010 the U.S. DOE announced the creation of the Southeast National renewable
Energy Center (SNMREC) at Florida Atlantic University. As the third National Marine
Renewable Energy Center, SNMREC will focus on ocean thermal energy conversion (OTEC)
and ocean current energy R&D.
Two FY09 solicitations were issued by the Wind and Water Power Program of the DOE in April
2009: (1) a Funding Opportunity Announcement (FOA) directed at industry partners and
industry-led teams; and (2) a Program Announcement (PA) directed at DOE national laboratories
to address technical challenges in water power development as well as market acceptance
barriers. The industry FOA consisted of six parts:
1. Marine and Hydrokinetic Energy Conversion Device or Component Design and
Development
2. Marine and Hydrokinetic Site-specific Environmental Studies/Information
3. Advanced Water Power Market Acceleration Projects/Analysis and Assessments
4. Hydropower Grid Services
5. Environmental Mitigation Effectiveness
6. University Hydropower Research Program

10-7

Ocean Wave Energy

Topic 3, Market Acceleration Projects/Analysis and Assessments consisted of six subparts:

An assessment of offshore ocean current energy resources along the U.S. coastline, excluding
tidal currents, to determine maximum practicably extractable energy

An assessment of in-stream hydrokinetic energy resourcesdefined as energy that can be


extracted from free-flowing water in rivers, lakes, streams, or man-made channels without
the use of a dam or diversionary structurein the United States to determine maximum
practicably extractable energy.

An assessment of projected life-cycle costs for ocean thermal energy conversion in the
United States over time

An assessment of global and domestic U.S. ocean thermal energy resources to determine
maximum practicably extractable energy

An assessment of projected life-cycle costs for wave, tidal, ocean current, and in-stream
hydrokinetic power in the United States over time

An assessment of the energy resources available from installing power stations on nonpowered dams and in constructed waterways and the construction of new pumped storage
facilities in the United States to determine maximum practicably extractable energy

On September 9, 2010, DOE announced the results of its FY 2010 solicitation from MHK energy
technology development projects. Substantial match with non-federal dollars is required of
grantees under the DOE funding opportunity. Wave energy proposals selected for funding
included the following:

$2.4 million grant to Ocean Power Technologies (OPT) to deploy a full-scale 150-kW
PowerBuoy system at its Reedsport Oregon site

$1.8 million grant to Northwest Energy Innovations to verify ocean wavelength performance
for a new point absorber wave energy device

$2.4 million grant to Wavebob, LLC, to develop and test an advance power take-off device
for their point absorber wave energy device

Several smaller grants for projects related to wave energy device and device component
development and monitoring

On October 26, 2010, DOE, BOEMRE, and the National Oceanic and Atmospheric
Administration (NOAA) announced awards of nearly $5 million for environmental research
projects to advance ocean renewable energy. Funded projects included the following:

Baysian Integration for Marine Spatial Planning and Renewable Energy Siting (Parametrix,
$499,000)

Characterization and Potential Impacts of Noise Producing Construction & Operation


Activities on the OCS (Cornell Lab of Ornithology, $499,000)

Developing Environmental Protocols and Modeling Tools to Support Ocean Renewable


Energy and Stewardship (University of Rhode Island, $745,000)

10-8

Ocean Wave Energy

Evaluating Acoustic Technologies to Monitor Aquatic Organisms at Renewable Sites


(University of Washington, $746,000)

Protocols for Baseline Studies and Monitoring for Ocean Renewable Energy (Pacific Energy
Ventures, $499,000)

Renewable Energy Visual Evaluations (University of Arkansas, $497,000)

Subseabed Geologic Carbon Dioxide Sequestration Best Management Practices (University


of Texas-Austin, $497,000)

Technology Roadmap for Cost Effective, Spatial Resource Assessments for Offshore
Renewable Energy (University of MassachusettsMarine Renewable Energy Center,
$748,000)

Throughout 2011 and 2012, the DOE offered multiple financial opportunities to MHK industry
businesses, including special loan guarantees and grants. The main types of awards offered to
MHK developers included Small Business Innovation Research (SBIR) programs and Small
Business Technology Transfer (STTR) programs. In FY 2010 the program selected 27 projects
for funding, with individual awards ranging from $160,000 to $10 million. In April 2012 the
DOE announced that there is $9 million available this year to fund approximately 50 small
businesses. Selected projects will receive 1-year awards of up to $150,000. Awardees with
successful projects will also have the opportunity to compete for follow-on funding in excess of
$1 million. These funding programs are intended to help emerging MHK technologies advance
along the sequence of DOE Technology Readiness Levels (TRLs). There are 10 TRLs, which are
organized as follows:

TRL 13 Concept Definition

TRL 4

Proof of Concept

TRL 5/6

System Integration and Laboratory Demonstration

TRL 7/8

Open Water System Testing, Demonstration, and Operation

TRL 9

Array Testing

TRL 10

Commercialization

Currently, the DOE is funding a large multi-organizational project to develop MHK reference
models. The project began in May 2010 and, according to a DOE presentation, the goal of the
project is to develop a representative set of Reference Models (RM) for the MHK industry to
develop baseline costs of energy (COE) and evaluate key cost component/system reduction
pathways. The presentation notes that there is a, need [to identify] COE targets with regard to
technology type and future innovation opportunities to prioritize research and cost reduction
pathways. The reference model completion date is slated for September 2013. The Point
Absorber Wave Energy Converter (WEC) report was due September 30, 2011. The Oscillating
Water Column WEC and Surge Type WEC reports are due September 30, 2012.

10-9

Ocean Wave Energy

So far in 2012 five funding opportunity announcements (FOAs) have been issued for water
power technologies:
1. Open Funding Opportunity Announcement
Advanced Research Projects Agency-Energy (ARPA-E)
2. Hydropower Advancement Project (HAP)Standard Assessments to Increase Generation
and Value
U.S. Department of Energy Office of Energy Efficiency and Renewable Energy
3. In-Water Wave Energy Conversion (WEC) Device Testing Support
U.S. Department of Energy Office of Energy Efficiency and Renewable Energy Water Power
Program
4. FY12 SBIR/STTR Funding Opportunity Announcement
Small Business Funding Opportunity Announcement (FOA)
5. DE-FOA-0000747: RFIImproving Marine and Hydrokinetic and Offshore Wind Energy
Resource Data
U.S. Department of Energy Office of Energy Efficiency and Renewable Energy
FOA #3 is particularly significant for the wave energy industry because it offers $500,000 to one
project to deploy and test a buoy (point-absorber) WEC for one year at the Navys Wave Energy
Test Site (WETS) off of Marine Corps Base Hawaii in Kaneohe Bay, Oahu. The opportunity was
announced on May 18, 2012 and the submission deadline was June 18, 2012.
10.2.1.3 Federal Regulations
On August 3, 2009, Minerals Management Service (renamed the Bureau of Ocean Energy
Management, Regulation, and Enforcement [BOEMRE]) and the Federal Energy Regulatory
Commission (FERC) issued guidance on the regulation of hydrokinetic projects on the Outer
Continental Shelf (OCS).
On July 29, 2012 the newly renamed Bureau of Ocean Energy Management (BOEM) (formerly
BOEMRE) and FERC announced revised guidelines for MHK developers pursuing permitting
and licensing on the OCS. The revisions further clarify the regulatory process and help
streamline the process for authorizing research and testing of MHK devices.
10.2.2 State Highlights
10.2.2.1 Projects in Hawaii
In October 2008, the University of Hawaii (UH) was selected as one of two national sites for
development of a Marine Renewable Energy Center (MREC). The Hawaii National Marine
Renewable Energy Center (HINMREC) is managed by the Hawaii Natural Energy Institute
(HNEI) at UH and funded by DOE at approximately $1 million per year with equivalent cost
share from UH and industrial partners. The center is an international partnership between
academia, industry, local and federal government agencies, and non-government organizations
(NGOs). The objectives of the Hawaii project are to facilitate the development and

10-10

Ocean Wave Energy

implementation of commercial wave energy systems with one or more systems deployed and
supplying power to the local grid at greater than 50% availability within five years, and to assist
the private sector to move ocean thermal energy conversion (OTEC) systems beyond proof-ofconcept to pre-commercialization.
The MREC will work closely with energy developers to conduct supporting research on system
performance and survivability, grid integration and environmental impacts, including completing
necessary environmental studies and assisting industrial partners to acquire required permits.
Partners include local engineering firms familiar with the permitting process.
MREC proposes to build on current and proposed marine energy projects in Hawaii to accelerate
the establishment of up to three field test facilities for hydrokinetic systems and one for OTEC
component testing. Proposed wave energy test sites include Pauwela Point on the northeast coast
of Maui, in cooperation with Maui Electric Company and Oceanlinx; the Kaneohe Marine Corps
Base on Oahu, where Ocean Power Technologies maintains an ongoing program; and off the
Makai Research Pier located west of Makapuu Point on the eastern tip of Oahu. The latter site is
proposed for obtaining long-term data series on wave energy resources, research on corrosion
and innovative materials, and an easily accessible site for deploying and testing small wave
energy conversion devices and components. The pier is already permitted for a range of marine
research activities. The HINMREC seeks to expand the existing facilities at Kaneohe Bay to
accommodate up to four wave energy conversion devices with 300 to 500 kW capacity each.
U.S. NavyOPT Project at Kaneohe Bay. From 2009 to 2011 Ocean Power Technologies (OPT)
tested its 40 kW PowerBuoy in 30 m depth, 0.75 miles offshore at the Kaneohe Bay Marine
Corps Base on Oahu, becoming the first grid-connected wave energy device in the United States.
Project accomplishments include the following:

Subsea and utility interconnection cables deployed and onshore station installed.

PowerBuoy interface with the electrical utility power grid certified as compliant with
international standards.

Validated hydrodynamic modeling using actual measurements through wave tank and ocean
testing.

Refined and ocean-tested PowerBuoy tuning software (tunes buoy performance to changing
wave conditions).

Environmental assessment (EA) resulted in a finding of no significant impact (FONSI) for up


to six buoys.

Exemption from FERC permitting and licensing requirements.

In 2011 OPT completed testing of its 40 kW PowerBuoy. On April 26, 2012 Philip Vitale, the
Deputy Chief Technology Officer and Director of Ocean Engineering for the Naval Facilities
Engineering Command, presented a project status update at the Global Marine Renewable
Energy Conference. He presented the following project updates for the Wave Energy Test Site
(WETS):

30 m test site is ready to be repopulated in FY12/13.

NEPA process (deep site) Environmental Assessment to be completed by end of FY12.


10-11

Ocean Wave Energy

Deep water system design FY12.

Deep water system installation FY13:

Shore station upgrade

Cabling form shore to two mooring locations

Moorings

Deep water system to be commissioned FY14.

Oceanlinx Maui Wave Energy Project (P-13521). FERC issued a preliminary permit to
OceanLinx Hawaii, LLC, on November 25, 2009, for it wave energy project off the coast of
Maui, about 0.6 miles north of Pauwela Point. The proposed project would deploy oscillating
water column wave energy conversion technology with a combined nameplate capacity of
2.7 MW. During the first year of the project, Oceanlinx worked with Maui Electric Company
(MECO) to refine the design and routing of the undersea cable to connect the project to MECOs
grid. MECO reported a preliminary cost estimate of $11 to $15 million. This estimate
significantly exceeds the initial estimates of $5 million; consequently, the project partners are
exploring cost-saving alternatives, including deploying a smaller device (up to 1.0 MW) in
shallower water closer to shore. Commencement of environmental studies awaits selection of a
lower-cost deployment site closer to shore. On January 13, 2012 FERC canceled Oceanlinks
preliminary permit following failure to submit a notice of intent and draft license application
(DLA).
10.2.2.2 Projects in Oregon
The Oregon Innovation Council granted $4.2 million to create the non-profit Oregon Wave
Energy Trust (OWET) in 2007. It is OWETs mission to establish the state as the preeminent
developer of wave energy in the United States, with the goal of producing 500 MW of clean
powerabout 3% to 5% of the states energy demandfrom the ocean by 2025. To achieve this
mission, OWETs strategy is to maintain technology neutrality and focus its resources on
reducing the barriers hindering the emerging wave energy industrys progress toward
commercial development. OWETs activities are grouped within four major program areas:

Stakeholder education and engagement: Specific activities include (a) Coastal Community
Open Houses, (b) creating a statewide network of coastal economic and community
development advisors, (c) facilitating development of organized fishermen groups to
participate in wave energy planning, (d) showcasing at consumer events, and (e) producing a
wave energy conference.

Regulatory and policy: Specific activities include development of Regulatory Roadmaps to


help wave energy developers navigate the complex network of state and federal permit and
license requirements. OWET actively monitors state legislative activities and provides wave
energy industry information to legislative representatives and committees.

Market development: OWET recently completed a Utility Market Initiative. This extensive
project sought to produce an effective market strategy to integrate wave energy projects into
the electric utility system and establish technical requirements to connect into the grid [3].
OWET hopes to create a utility pull and wave energy push to help meet the target

10-12

Ocean Wave Energy

production goals. On March 25, 2010, OWET released its Wave Energy Infrastructure
Assessment in Oregon report [4]. In July 2011, OWET published a report whose purpose was
to generate policy recommendations and market approaches to address the above-market
price of ocean energy-produced electricity [5].
In 2010 OWET also created the Oregon Wave Energy Commercialization Program (OWEC).
According to the OWET Oregon waveNEWS publication, the program is designed to
create market development activities by offering strategic business development services to
ocean industry companies and organizations. The two main elements of the program are (1)
commercialization grants and (2) commercialization acceleration support services. The
OWEC Grant Program helps the wave energy industry by bridging the gap between
traditional R&D funding sources and the availability of private investment. The following
section details recent OWEC Grant Program awards, quoted from OWETs Oregon
waveNEWS publication:

June 2012
o Neptune Wave Power was selected to receive an OWET grant for the fabrication,
deployment and monitoring of its WEC buoy at NNMRECs Newport test facility.

July 2012
o Atmocean, Inc. will undertake a full-size mini-array deployment near Coos Bay,
Oregon this summer to accelerate development of their Wave Energy/Sequestration
Technology (WEST) system. Upon completion, WEST will achieve TRL 7. OWET
funding is currently being utilized to aid in fabrication, deployment, and eventual
retrieval of the WEST system.
o Columbia Power Technologies will conduct experimental validation of recent, highimpact design modifications that offer a cost of energy decrease of well over 30%.
This project will include optimization of the new design, fabrication and wave tank
testing at Oregon State University of a 33rd scale WEC and reduced-impact mooring,
and post-processing and data analysis.
o M3 Wave Energy will enlist local industry partners to develop a submerged wave
energy device manufacturing plan with a focus on key Department of Energy TRL 4
criteria, which include a pilot project fabrication plan and a commercial-scale array
manufacturing plan.
o Northwest Energy Innovations (NWEI), in partnership with other industry leaders
from New Zealand, will verify the ocean wavelength functionality of the Wave
Energy Technology-New Zealand (WET-NZ) device through wave tank testing and
controlled open sea deployment of a 1:2 scale device. NWEI expects the WET-NZ
technology will be ready for commercial development in the United States upon
completion of this program.

10-13

Ocean Wave Energy

Research: OWET directs and funds environmental and applied research projects to answer
key questions about wave energy development. It has completed baseline assessment studies
on seabirds and whale migration at the proposed wave energy project sites. During 2009 and
2010, OWET supported additional research on seabirds, crab distribution, EMF, sediment
transport, and create a planning tool to model cumulative effects of wave energy projects.
OWET published a report in 2011 that makes recommendations regarding policies and
market approaches to address the above-market price of electricity produced by ocean energy
projects [5].

Douglas County Wave and Tidal Energy (P-12743). FERC issued a preliminary permit for the
Douglas County Wave and Tidal Energy Project on October 6, 2010. This project would be
located at the mouth of the Umpqua River near existing jetties. The preliminary permit
authorizes capacity up to 3 MW. The project developer considers an oscillating water column
system the most feasible for the site. As of May 11, 2012 the project continues to submit
progress reports to FERC.
Coos Bay OPT Wave Park (P-12749). On August 10, 2010, FERC issued a preliminary permit
for the Coos Bay OPT Wave Park Project. The project would be located in the Pacific Ocean
between 2.5 and 3.0 miles off the coast near Coos Bay, Oregon. The project would consist of 200
PowerBuoys having a combined, installed capacity of 100 MW, as well as an approximately 3.4mile subsea transmission cable and other facilities. As of August 16, 2012 the project continues
to submit progress reports to FERC.
Reedsport OPT Wave Park (P-12713). On February 1, 2010, OPT submitted a License
Application (LA) for the 1.5-MW Reedsport OPT Wave Park project. It is the first commercial
scale wave park on the West Coast of the United States. On July 28, 2010, OPT and several
intervenors to the FERC licensing process entered into a settlement agreement resolving issues
associated with issuance of an original license for the project, and requesting a license term of 35
years. On December 3, 2010, FERC published an environmental assessment (EA) that analyzed
the potential environmental effects of licensing the project and concluded that licensing the
project, with appropriate environmental protective measures, would not constitute a major
federal action that would significantly affect the quality of the human environment. On August
24, 2011, the Department of Energy issued a notice adopting FERCs EA and announcing a
FONSI for the single buoy project. The company had delayed plans to deploy a single
PowerBuoy until summer of 2012 and begin deployment of the additional nine PowerBuoys
sometime thereafter.
In June of 2012 OPT announced that it had successfully completed extensive factory acceptance
testing of its utility-scale PowerBuoy (PB150).85 On August 13, 2012 FERC granted OPT a
commercial license for the full build-out of the 1.5 MW, grid-connected wave power station.
This is the first commercial license issued for a wave power project in the United States. The
license provides approval for the deployment of up to 10 grid-connected OPT devices for 35
years.

85

PowerBuoy is a registered trademark of Ocean Power Technologies.

10-14

Ocean Wave Energy

Reedsport OPT Wave Park, Phase III (P-13666). Ocean Power Technologies (OPT) received a
preliminary permit on March 15, 2011, for Phase III of its Reedsport commercial wave energy
park. Phase III is intended to expand the project from the initial 1.5-MW project to 50 MW,
comprising 100 500-kW PowerBuoys. The Phase III project is intended to expand the size of the
project site, potentially utilizing the transmission cable and corridor associated with the 1.5-MW
project. OPT also intends to use the results of studies conducted at the 1.5-MW project to inform
the evaluation of the License Application for the 50-MW project. As of July 31, 2012 the project
continues to submit progress reports to FERC and collaborate with various regulatory agencies
and stakeholders.
Northwest National Marine Renewable Energy Center. Oregon State University (OSU) and the
University of Washington (UW) were awarded $6.25 million by DOE to develop the Northwest
National Marine Renewable Energy Center (NNMREC). OSU plans to deploy various devices
and subsequent environmental measurement devices at a location near Newport to study new
technologies and evaluate potential environmental effects. During the summer of 2012 the DOE
prepared and released a draft environmental assessment (EA) to analyze and describe the
potential environmental impacts associated with NNMRECs Newport facility. Public comments
were accepted until July 18, 2012.
NNMREC plans to develop multiple ocean test berths in two phases. Phase 1 consists of a
mobile ocean test berth (MOTB) to allow testing and analysis of WEC devices without
connecting to the electric grid. Phase 2 consists of a grid connected ocean test berth (GCOTB).
NNMREC is currently in Phase 1, with scheduled completion in late 2012. The NEPA process
has been initiated and NNMREC is moving forward with environmental reviews and permitting.
During the summer of 2012 OSU and NNMREC will test its first device, called WET-NZ, at the
Newport ocean test berth. The center will also deploy the Ocean Sentinel, a floating
instrumentation buoy developed at NNMREC. The buoy is designed for testing wave energy
converters (WECs) and provides power analysis and data acquisition, environmental monitoring,
and an active converter interface to control power dissipation to an on-board electrical load.
Currently NNMREC/OSU facilities are capable of modeling and testing small-scale WEC
prototypes at the Corvallis, Oregon campus, and a non-grid connected test berth for in-ocean
testing is available in Newport, Oregon. In recognition of the need for a facility capable of
accommodating commercial scale devices, NNMREC is working to develop the Pacific Marine
Energy Center (PMEC). PMEC will be a world class, full-scale, grid-connected ocean energy
test facility at NNMREC capable of accommodating commercial scale devices (TRL 5-9). It will
be the first full scale, grid connected test center in the United States. With funding from OWET,
NNMREC will partner with the widely successful European Marine Energy Center (EMEC) to
complete the development plan for PMEC.
10.2.2.3 Projects in California
Pacific Gas and Electric Humboldt WaveConnect Project (P-12779). On October 28, 2010,
PG&E announced that it was suspending the Humboldt WaveConnect project, citing permitting
challenges and unexpectedly high costs.

10-15

Ocean Wave Energy

PG&Es service territory borders 960 km of Pacific coastline, with wave power densities of 20
40kW/m, making wave power a renewable energy resource of strategic value to the utility. The
WaveConnect projects goal was to assess wave energys potential and examine the regulatory
and environmental issues associated with such a facilitys development. WaveConnect was
intended to provide the infrastructure to test small arrays of commercial wave energy conversion
devices in California, and therefore allow the emerging wave power industry and PG&E to gain
an understanding of the full life cycle of deployed technologies (Figure 10-5).

Figure 10-5
Schematic of PG&E WaveConnect system

In late 2008, DOE selected PG&E to receive a cost-sharing grant of $1.2 million. In early 2009,
PG&E received a decision of approval from the California Public Utility Commission to spend
$4.8 million of ratepayer-based funds toward the estimated design and licensing costs for the
pilot WaveConnect project.
PG&E received preliminary permits from FERC for the Humboldt site and a second potential
site in Mendocino County in early 2008. A preliminary permit for a third site near Vandenberg
Air Force Base referred to as the Central Coast site was granted in May 2010. In early May 2009,
PG&E announced that it was dropping the Mendocino County study because Port Noyo Harbor
in Fort Bragg, California, was found to be unsuitable to support a pilot wave energy project.
After considerable effort pursuing licensing, permitting, data collection, planning, and
communication with WEC manufacturers, local stakeholders, and the several public agencies
involved, PG&E suspended work on the Humboldt project in November 2010 and the Central
Coast project in May 2011. PG&E intended WaveConnect to take advantage of FERCs
expedited Pilot Project Licensing Process (PPLP). However, other permitting, regulatory,
environmental, and stakeholder engagement processes involved in siting a wave power project
largely negated any potential advantages offered by the PPLP, leading PG&E to conclude that
WaveConnect was untenable at both sites given the state of wave power technologies and the
available funding and resources.

10-16

Ocean Wave Energy

Green Wave Mendocino Project (P-14291) and San Luis Obispo Project (P-14292). FERC
took final action on October 26, 2010 canceling Green Wave Energy Solutions two projects off
the cities of Mendocino (P-13053), and Morro Bay (P-13052), California. These preliminary
permits were canceled for failure to file a notice of intent (NOI) to file a license application and a
pre-application document (PAD) for each project. Green Wave Energy Solutions filed new
applications for preliminary permits (100 MW each) off Mendocino and San Luis Obispo on
September 26, 2011.
On March 23, 2012 FERC rejected the preliminary permit application (P-14292) for the San Luis
Obispo Wave Park Project because Green Wave Energy Solutions failed to respond to FERCs
February 8, 2012 comments by March 9, 2012.
On July 19, 2012 FERC denied the preliminary permit application for the Green Wave
Mendocino Wave Park Project (P-14291) because it failed to demonstrate good faith and due
diligence during the term of its prior permit.
Sonoma Coast Hydrokinetic Energy Project Fort Ross (South) (P-13377). FERC issued
preliminary permits to the Sonoma County Water Agency for three projects on July 9, 2009, one of
which is still active. The project would deploy oscillating water column devices and/or buoy-type
wave energy conversion devices with a combined capacity of up to 5 MW. The preliminary permit
application noted possible future expansion of the project up to 40 to 200 MW.
On August 4, 2011 FERC canceled the preliminary permit because the project chose to suspend
further developments until funding was secured.
SWAVE Catalina Green Wave Energy Project (P-13498). FERC issued a preliminary permit
for the SWAVE project to Scientific Applications & Research Associates, Inc. (SARA) on
September 15, 2009, for an array of 10 to 40 SWAVE buoys approximately 0.75 miles off the
west coast of Santa Catalina Island. SARA voluntarily surrendered the preliminary permit for the
project on October 13, 2010, after determining that the project was not feasible at that time.
San Onofre OWEG Electricity Farm (P-13679). FERC issued JD Products, LLC, a preliminary
permit on October 29, 2010, for the proposed San Onofre OWEG Electricity Farm Project. The
project would consist of up to 11,443 Ocean Wave Electricity Generation (OWEG) units, an
experimental technology, with an estimated total installed capacity of 3,186 MW. The landbased portions of the project would be located on a site adjacent to the San Onofre Nuclear
Generating Station (SONGS). The application notes that SONGS Station 1 is inoperative, and
Stations 2 and 3 will be decommissioned by 2022. The application proposes to use the
distribution infrastructure currently associated with SONGS. The initial plan is to install only 50
of the experimental OWEG units for an estimated capacity of 8.9 MW.
On October 27, 2011 the project filed a notice of intent (NOI) and pre-application document
(PAD) with FERC. Later, it was determined that the project was unable to afford the costs
associated with conducting environmental studies prior to the construction and operation of the
proposed project. The project requested that the NOI/PAD process be put on hold until it could
obtain additional funding. On March 9, 2012 FERC terminated the integrated licensing process
for the project.

10-17

Ocean Wave Energy

10.2.3 U.S. Developer and Project Deployment Highlights


10.2.3.1 Ocean Power Technologies
OPT is pacing domestic wave energy development. In addition to OPT, the United States has
roughly a half dozen developers that are at an early stage of technology development. In late
2011, OPT announced it will begin work on a wave prediction model that will assess the
characteristics of approaching waves before they reach the wave energy converter, thereby
improving the tuning of the device to wave characteristics and increasing the capacity factor of
the device. Recent highlights from OPT projects include the following:

Invergordon, Scotland: OPT deployed its PB150 PowerBuoy on April 15, 2011 for ocean
trials approximately 33 nautical miles from Invergordon, Scotland. Peak electrical power
exceeded 400 kW, indicating the PB150 could produce its rated capacity of 150 kW on
average under favorable wave conditions. The six-month trials have concluded successfully.

Reedsport, Oregon: On July 28, 2010, OPT signed a settlement agreement with 11 federal
and state agencies and three non-governmental stakeholders for its 10-buoy, 1.5-MW
commercial power project off the coast at Reedsport, Oregon. This agreement followed
submission of the Final License Application for the project on February 1, 2010. On August
13, 2012 FERC granted OPT a commercial license for the full build-out of the 1.5 MW, gridconnected wave power station. This is the first commercial license issued for a wave power
project in the U.S. The license provides approval for the deployment of up to ten gridconnected OPT devices for 35 years. OPT is working through the permitting and licensing
process to expand this site to 50 MW capacity. This expansion project is called Phase III, and
received a preliminary permit on March 15, 2011 (P-13666). As of July 31, 2012 the project
continues to submit progress reports to FERC and collaborate with various regulatory
agencies and stakeholders.

Kaneohe Bay, Hawaii: On September 27, 2010, OPT announced that it had completed the
first-ever grid connection of a wave energy device in the United States at the Marine Corps
Base Hawaii (MCBH), Kaneohe Bay, Oahu, Hawaii. The 40 kW device was deployed on
December 14, 2009 about 0.75 miles offshore in approximately 100 feet of water. By August
2011, the PB40 PowerBuoy system had completed 5.6 million cycles of operation. In 2011
OPT completed testing of its 40 kW Power Buoy. Project accomplishments include the
following:

Subsea and utility interconnection cables deployed and onshore station installed.

PowerBuoy interface with the electrical utility power grid certified as compliant with
international standards.

Validated hydrodynamic modeling using actual measurements through wave tank and
ocean testing.

Refined and ocean-tested PowerBuoy tuning software (tunes buoy performance to


changing wave conditions).

Environmental assessment (EA) resulted in a finding of no significant impact (FONSI)


for up to six buoys.

Exemption from FERC permitting and licensing requirements.

10-18

Ocean Wave Energy

Cornwall, UK: OPT is poised to be the first customer at the South West of England Regional
Development Agencys (SWRDAs) Wave Hub project. The Wave Hub system underwent
its first full test in early November 2010.

Victoria, Australia: In July 2012 Victorian Wave Partners (owned by OPT) and Lockheed
Martin entered into an agreement to advance a 19-MW wave energy project in Victoria,
Australia. The project is one of the largest wave energy projects in the world and will utilize
a $66.5 million grant from Australias Department of Resources, Energy and Tourism.
Lockheed Martin will assist with the design of OPTs PowerBuoy technology, lead the
production and system integration of the wave energy converters, and support overall
program management. Construction will begin in 2013.

10.3 Worldwide Wave Energy Highlights: Late 2011 to Mid-2012


Interest in ocean wave renewable energy is strong worldwide. This section presents a brief
accounting of notable activities within the sector outside of the United States from late 2011
through mid- 2012.
10.3.1 United Kingdom and Ireland
The UK is maintaining its stature as the global leader in wave energy technology development.
In an effort to solidify its leadership position, the UK has established a goal of 2 GW installed
marine energy capacity by 2020 and adopted an energy policy designed to attract and support
WEC developers and equipment testing. The UK contains three wave testing facilities and is
home to many wave energy developers.
10.3.1.1 Wave Energy Developers
WEC developers in the UK with significant developments include Aquamarine Power and
Pelamis WavePower.
Aquamarine Power

Aquamarine Power's Oyster system (Figure 10-6) was officially launched at the European
Marine Energy Centre (EMEC) in Orkney on November 20, 2009. In July 2011, Aquamarine
Power unveiled its next-generation device, the 800-kW rated capacity Oyster 800. The device
commenced operational testing at EMECs Billia Croo test site in June of 2012. During that
same month, the device produced its first electrical power to the grid. Aquamarine Power has
also been granted consent to install up to two more Oyster devices at EMEC.
The Oyster is a hydroelectric wave energy converter that consists of an oscillator fitted with
pistons and fixed to the near-shore seabed. Each passing wave activates the Oscillator, pumping
high-pressure water through a subsea pipeline to the shore. Onshore, conventional hydroelectric
generators convert this high-pressure water into electrical power. Together, the three Oyster 800
devices will form a 2.4-MW array. The Oyster has been under development by Aquamarine Power
since 2005, in partnership with the Marine Energy Research Group at Queens University, Belfast.

10-19

Ocean Wave Energy

Figure 10-6
Aquamarine Power Oyster

Aquamarine Power received a $100,000 grant (requiring a $100,000 company match) from the
Oregon Wave Energy Trust to assess the wave energy potential along Oregons coast, but
forfeited the award, citing regulatory uncertainty surrounding Oregons ongoing Territorial Seas
Plan development. The plan was intended to be complete in 2010, but likely will not be complete
before mid-2012. Until the plan is complete, the state is precluded from offering seabed leases
for wave energy development.
In May, 2011, Aquamarine Power secured leases totaling 40 MW off Scotlands Western Isles
for deployment of its Oyster technology. The company secured $11 million in new funding in
September, 2011, following announcement of a $5.5 million loan from Barclays earlier the same
month.
Pelamis WavePower

The Pelamis wave energy converter (WEC) is rated at 750 kW with a target capacity factor of
25% to 40%. The device consists of five tube sections linked by flexible, bi-directional joints.
The power take-off system is driven by hydraulic cylinders at the joints. The company has
produced six full-scale devices to-date.
In July 2010, E.ON deployed a second-generation Pelamis WavePower device at EMEC.
Following four days of successful testing, the device was removed for inspection and prepared
for redeployment. The machine has been grid-connected since October 2010. A second device
owned by Scottish Power Renewables arrived in early November 2011, and in 2012 will operate
in tandem alongside the previously deployed device.

10-20

Ocean Wave Energy

In May 2011 Aegir Wave Power, a joint venture between Pelamis WavePower and Vattenfall,
was awarded an agreement for lease from the Crown Estate for a 10 MW wave farm at the
Shetland Islands, off the coast of Scotland. The project will consist of 10 to 14 Pelamis wave
machines and will begin construction in 2015.
On March 15, 2012 Vattenfall announced that it had reserved the last remaining berth at EMEC
to test a 750-kW Pelamis wave energy converter. Vattenfall plans to start testing a Pelamis
device in 2014 and, following successful tests, will purchase a single Pelamis device.
Pelamis WavePower is a partner in the Hebridean Marine Energy Futures Project in Scotland.
Through this project, Pelamis is planning to develop a 10 MW wave farm in the Bernera area off
the west coast of Lewis in the Outer Herbrides. The proposed wave farm will consist of up to 14
Pelamis wave energy devices. In October 2011 the Bernera project was successful in securing an
agreement for lease from the Crown Estate.
10.3.1.2 Wave Energy Test Centers
The UKs wave testing facilities include the New and Renewable Energy Center (NaREC), with
sub-scale prototype testing in a wave tank; the European Marine Energy Center (EMEC), with
single full-scale prototype testing in natural waters; and Wave Hub, for arrays of full-scale
prototypes tested in natural waters. Each facility is described here.
NaREC

NaREC is a leading research and development platform for new, sustainable, and renewable
energy technologies located in Blythe, England. Its range of development, testing and consulting
services work to support the evolving energy industry and transform innovative new
technologies into commercial successes. NaREC provides the emerging marine renewables
industry the support it needs to transform winning concepts into commercial successes. NaREC
services include the following:

Complete in-house prototype development facilities for wave technology, including a wave
tank

Mechanical and electrical design engineering and procurement

Electrical engineering consultancy and support for power conversion and drive train
development

Complete system testing from marine environment to grid connection

Resource and feasibility assessment and consultancy

Market analysis and research

Project management, funding and investment coordination

10-21

Ocean Wave Energy

European Marine Energy Centre (EMEC)

In operation since 2003, EMECs current wave energy clients include Aquamarine Power, E.ON,
ScottishPower Renewables, Seatricity, Vattenfall, and Wello Oy. The first developer to test its
wave energy device at EMEC was Pelamis WavePower. Between 2004 and 2007 Pelamis
WavePower its their first-generation P1 device at EMEC. The second wave energy developer to
use EMECs facilities was the Finnish company AW-Energy. In 2005 the company tested its
fully submerged WaveRoller device.
Successful testing of the Pelamis P1 device led to the development of Pelamiss second
generation P2 device, which was deployed at EMEC by the utility company E.ON in July 2010.
Another utility company, ScottishPower Renewables, deployed a Pelamis P2 device adjacent to
the E.ON device at EMEC in May 2012. A third utility company, Vattenfall, plans to deploy a
Pelamis P2 device at EMEC in 2014.
Edinburg-based Aquamarine Power is the developer of the Oyster wave energy device. To date,
Aquamarine Power has deployed and tested two full-scale Oyster devices at EMEC: the 315 kW
Oyster 1, and the 800 kW Oyster 800. The Oyster 800 was grid-connected in June 2012 and is
currently undergoing testing at EMEC. Aquamarine Power has also been granted consent to
install up to two more Oyster devices alongside the Oyster 800 device at EMEC.
The British company, Seatricity, plans to deploy its 1-MW device at EMEC during the summer
of 2012. The Seatricity concept involves multiple floats traveling up and down with the waves,
operating pumps to pressurize sea water, which is piped ashore to drive a standard hydroelectric
turbine.
The Finnish company Wello Oy successfully installed its full-scale Penguin device at EMEC on
June 30, 2012.
Figure 10-7 illustrates Ocean Power Technologies 150-kW PowerBuoy (potential future EMEC
client), Aquamarine Powers Oyster, and a second-generation P2 Pelamis WavePower device.

10-22

Ocean Wave Energy

Figure 10-7
EMEC testing configuration in 2011

The EMEC test site at Billia Coo in mainland Orkney provides the worlds only multi-berth,
open-sea test facility for wave energy converters. The EMEC offices and data facilities are in
Stromness. Orkney was chosen because of its natural and man-made resources. The wave test
facility site receives uninterrupted Atlantic waves of up to 15 meters. Orkney is also the most
northerly community connected to the UK national grid, has excellent harbor facilities, and a
significant professional community experienced in working with renewable energy.
On July 30, 2012 the Marine Energy Park was launched at EMEC to heighten the international
profile of the region and its reputation as a world leader in marine energy.
The UK Wave Hub

Wave Hub is a 42 million test facility for wave energy technology. In early September 2010,
Wave Hub deployed a 12-ton, grid-connected socket on the ocean floor 10 miles offshore of
Cornwall, South West England in 55 m of water. On November 2, 2010, the complete system
was connected to the grid and tested. Four 300-m tails extended from the hub, each
constituting a test berth for deploying a wave energy conversion device with a capacity up to 4 to
5 MW. The project funders are the South West Regional Development Agency (12.5 million),
the European Regional Development Convergence Fund Programme (20 million), and the UK
government (9.5 million). In April 2011 Wave Hub welcomed the first wave energy device for
testing, Fred Olsens BOLT Lifesaver wave energy converter. The device is being tested at the
Falmouth Bay test site, part of the South West Marine Energy Park.

10-23

Ocean Wave Energy

Ireland Test Centers and Deployments

Irelands target installed capacity for ocean energy is 500 MW by 2020. According to the 2012
GMREC presentation by The Sustainable Energy Authority of Ireland (SEAI), Ireland is
developing two open sea test sites and three options for commercial wave energy deployments
(WestWave project). The full scale Atlantic Marine Energy Test Site (AMETS) is located in
Belmullet. According to the GMREC presentation, the foreshore lease application was submitted
in December of 2011 and is currently in the public consultation phase. The substation planning
application is to be submitted by the summer of 2012. In addition to the Belmullet test site,
Ireland is also developing a -scale Wave Energy Test Site in Galway Bay.
The WestWave project is part of the Irish Governments Ocean Energy Strategy, and seeks to
become the first wave energy project in Ireland by 2015. According to the projects website
(http://www.westwave.ie) the strategy is divided into four phases: (1) prototype and R&D
development, (2) development of a wave energy test site to facilitate pre-commercial single
device demonstration, (3) development of pre-commercial array projects between 5 and 10 MW,
and (4) development of commercial array projects. Phases 1 and 2 are already under way. The
purpose of the WestWave project is to fulfill Phase 3 by 2015. Phase 4 seeks to have commercial
array projects being developed toward 2020 and beyond. Currently, three locations are being
considered for the WestWave commercial deployments: Belmullet, Achill, and Clare. The
project consortium includes four of the leading wave energy developers: Wavebob, Ocean
Energy, Aquamarine Power, and Pelamis.
10.3.2 Portugal and Spain
Portugal and Spain are currently involved in a variety of research projects and multi-national
collaboration programs aimed at advancing the status of the MHK energy industry. The
following information presents a brief overview of these activities.
10.3.2.1 Developments in Portugal
Portugals Wave Energy Centre (WavEC) is a non-profit organization, founded in 2003, and
dedicated to the development of ocean wave energy and other offshore renewables. WavEC is
currently funding a large variety of projects to advance the status of the ocean energy industry.
The following information presents a brief overview of select projects.
The Pico Power Plant is a 400 kW wave energy demonstration power plant on Pico Island in the
Azores. The plant has been in operation since 2008 and utilizes oscillating water column (OWC)
technology to generate electricity.
The WaveRoller prototype of a bottom-mounted flat-plate oscillating device developed by the
Finnish company AW-Energy was deployed in April 2007 at Peniche, Portugal, 100 km north of
Lisbon. In 2008, AW-Energy announced plans to construct a 1-MW plant. The company now
has all the necessary permits to install the project and connect it to the grid. The company
expected to ship the first 3 100-kW WaveRoller by the end of December 2011, followed by
deployment in Peniche as soon as the weather allowed.

10-24

Ocean Wave Energy

10.3.2.1 Developments in Spain


Commissioned in July 2011, Spains Mutriku oscillating water column (OWC) plant was the
worlds first commercial scale wave energy power plant. The plants total capacity is 296 kW.
The Wave Energy Lift Converter Multiple Espaa (WELCOME) project was installed in April
of 2011, and includes a 1:5 scale (150 kW) WEC prototype device developed by PIPO Systems.
The project is located approximately four nautical miles from Las Palmas Harbor in the Spanish
Canary Islands.
One of the most notable developments in Spain is the Biscay Marine Energy Platform (BIMEP).
BIMEP is an open ocean infrastructure for research, demonstration, and operation of offshore
wave energy converters (WEC). Construction was expected to begin during the second quarter of
2012, with operation during the fourth quarter of 2013.
10.3.3 Denmark, Norway and Sweden
A grid-connected wave energy test site at Nissum Bredning in the northwestern corner of
Denmark was built to enable various technology developers to test and demonstrate their
technologies at different scales. A 24-m long, 1:10 scale model of one such WEC device, the
Wave Star, was tested at Nissum Bredning until August 2008. Figure 10-8 shows the Wave Star
at Nissum Bredning with the buoys raised in the survival position. On September 18, 2009,
Wave Star deployed its next-generation device at Hanstolm, Denmark. The 40-m long device is a
test section of the companys commercial 500-kW machine (Figure 10-9). The device has been
connected to the grid since February 2010.

Figure 10-8
Wave Star 1:10 machine with buoys raised

Figure 10-9
Wave Star 1:2 machine with buoys raised

Floating Power Plant A/S has constructed a 37-m model for a full off-shore test at Vindeby
offshore wind turbine park, located off the coast of Lolland in Denmark (Figure 10-10). The test
system, named Poseidon 37, is 37 meters wide, 25 meters long, 6 meters high (to deck), and
weighs approximately 300 tons. It is a combined wind-wave energy converter. The test plant was
launched in Nakskov Harbour in May 2008, and towed to the test site and installed in August
2008. On June 14, 2010, Poseidon was redeployed at a test site in Onsevig, Denmark, to begin
Test Phase 2. The platform has demonstrated its stability even when the wind turbines are
operating at peak efficiency. The Poseidon 37 was brought ashore on October 24, 2011, for
repair and refitting with a new power take-off system and a more advanced system for
conditioning monitoring.
10-25

Ocean Wave Energy

In Sweden, the Seabased AB system is being tested. It uses a three-phase, permanent-magnet,


linear generator and is especially developed to be used in ocean bed arrays and directly driven by
point absorbers (buoys) on the surface. Figure 10-11 shows the preparation of the generator for
launch off the coast of Lysekil, Sweden. The WEC unit consists of a buoy coupled directly to the
rotor of a linear generator by a rope. The tension of the rope is maintained with a spring pulling the
rotor downwards. The rotor moves up and down at approximately the same speed as the wave. The
linear generator has a uniquely low pole height and generates electricity at low wave amplitudes
and slow wave speeds. Directly driven linear wave energy converters are deployed and coupled in
arrays at intervals of 25 to 50 meters. The Swedish Energy Authority awarded Seabased AB 139
million Swedish kroner (approximately $21 million US) on February 11, 2010, toward the 250
million Swedish kroner ($37 million US) needed to build a 400- to 500-unit, 10-MW project off
the coast of Smgen, Sotens, Sweden. On November 9, 2011 the European Commission
authorized the plans for the Swedish award.

Figure 10-10
Floating power plant AS Poseidon

Figure 10-11
Seabased AB linear generator

In March, 2011, Wave Dragon announced it had started development of a 1.5-MW North Sea
demonstration of its Wave Dragon device. The company plans to deploy the device offshore of
Hanstholm at the DanWEC test center in Denmark. A 1:4.5 scale prototype of a larger device
was launched in 2003 and was the worlds first offshore grid-connected wave energy conversion
device. It was deployed off the cost of Denmark at Nissum Bredning, and as of August 2009 had
accumulated over 20,000 hours of experience supplying electricity to the grid.
10.3.4 Australia, New Zealand, and Tasmania
Australia, New Zealand, and Tasmania are also pursuing wave energy projects. Companies
involved in developing wave energy in Oceana include BioPower Systems Pty, Oceanlinx, and a
partnership among Industrial Research Limited, the National Institute of Water and Atmospheric
Research (NIWA), and energy industry consultants Power Projects Limited.

10-26

Ocean Wave Energy

10.3.4.1 BioPower Systems Pty Ltd.

BioPowerSystems, a Sydney-based company, announced on March 1, 2010 that it had


secured land access and onshore development rights and project intellectual property for a
commercial-scale project at Port Fairy, Australia. The project plan calls for initial
deployment of a 250-kW bioWAVE system, to be followed by a commercial array of 1-MW
bioWAVE units. Preliminary assessments indicate the site could support up to 100 MW of
installed capacity. Figure 10-12 shows a subscale model of the bioWAVE. On August 8,
2011, BioPower Systems announced that it had completed extensive tests of its full-scale,
250-kW O-Drive power conversion module, using a test rig. The device is designed to plug
into both the bioWAVE and bioSTREAM devices and convert oscillating mechanical motion
to grid-ready AC power. On December 1, 2011, the company announced that it had received
$5 million in funding from the Victorian (Australia) government. BioPower Systems must
raise another $3.6 million toward the $14 million pilot demonstration project to be gridconnected near Port Fairy, Victoria. The pilot project will demonstrate a single 250-kW
bioWAVE module, four of which, arranged in parallel, will constitute a 1-MW commercialscale device. On July 3, 2012 the project was awarded $5.6 million in additional funding
from the Federal Minister for Resources and Energy, Martin Ferguson. The project is
currently in the final design phase and fabrication is scheduled to begin towards the end of
2012. Installation is expected in 2013 and the project will operate for 21 months, and then be
decommissioned.

Figure 10-12
BioPower BioWave

Figure 10-13
Oceanlinx oscillating water column

10.3.4.2 Oceanlinx Limited


Oceanlinx, which manufactures an oscillating water column device (Figure 10-13), has made
steady advances toward a commercial device:

Port Kembla (New South Wales, Australia): Oceanlinx deployed its third-generation, 1:3
scale model, pre-commercial device, known as the Mk3PC, on February 26, 2010, and
connected it to the grid on March 19. On May 14, 2010, the device broke free of its mooring.
Otherwise, the Mk3PC performed as designed, thereby validating the design for Oceanlinxs
full-scale 2.5-MW commercial device.
10-27

Ocean Wave Energy

Oceanlinx is planning to deploy the worlds first 1-MW wave energy converter,
greenWAVE, in Port MacDonnell, South Australia. The device utilizes oscillating water
column (OWC) technology. On July 4, 2012 the Australian government awarded
approximately $4 million to Oceanlinx for their project. The funding includes one year of
operating costs after grid connection.

Hawaii: A memorandum of understanding was signed with Maui Electric Co. (MECO) in
Hawaii for up to 2.7 MW. On November 9, 2011, FERC notified Oceanlinx of probable
cancellation of the preliminary permit issued on November 25, 2009, for failure to file the
required notice of intent and draft license application within two years of preliminary permit
issuance.

10.3.4.3 Aotearoa Wave and Tidal Energy Association (AWATEA) of New Zealand
According to the 2012 GMREC presentation by AWATEA, there are several active ocean energy
projects underway in New Zealand, including Crest Energy, Wave Energy Technology-New
Zealand (WET-NZ), Chatham Islands Marine Energy (CHIME), Neptune Power, Tangaroa
Energy, and Energy Pacifica. The following information presents a brief update of select
projects:

WET-NZ: The WET-NZ project is an R&D project to develop a wave energy converter
(WEC). After multiple device design concepts, the third device (20 kW) was deployed for
three months and is currently undergoing refurbishment. WET-NZ and Northwest Energy
Innovations (NWEI) were recently awarded $2 million to further testing of the WET-NZ
device at Oregon State Universitys Wave Energy Test Project beginning in the summer of
2012.

Chatham Islands: The Chatham Islands project plans to deploy a Wavegen LIMPET device
on the main Chatham Island, totaling 220 kW. In July 2010 the project received $2.16
million in funding. Currently the project is facing delays due to issues over ownership of the
islands electricity system.

10.4 Wave Power and Energy Resources


A number of sources provide wave power and energy resource data for assessing potential sites.
These sources include in situ measurements, satellite measurements, and wind-wave models.
In late 2011 EPRI completed a DOE-funded assessment of the United States ocean wave energy
resources [2]. EPRIs estimate of the available U.S. offshore wave energy resource along the
Outer Continental Shelf (notional 200 m depth contour) is 2,640 TWh/yr. State-by-state
estimates (including estimates for the inner shelf (notional 50 m depth contour) are provided in
Tables 10-2 to 10-6. For all tables, estimates along the Outer and Inner Continental Shelf are for
the notional 200 m depth contour and 50 m depth contour, respectively.

10-28

Ocean Wave Energy


Table 10-2
Alaska available wave energy resource breakdown (TWh per year)
Alaska

Outer Shelf

Inner Shelf

194

79

Pacific Ocean

1,356

724

Total Alaska

1,550

803

Bering Sea

Table 10-3
West Coast available wave energy resources (TWh per year)
State

Outer Shelf

Inner Shelf

Washington

116

72

Oregon

179

143

Northern California

65

45

Central California

185

148

Southern California

43

12

587

419

California

Total West Coast

Table 10-4
Hawaii available wave energy resources by major island (TWh per year)
Hawaii

Total Hawaii

Outer Shelf

Inner Shelf

Kauai

21

19

Oahu

22

14

Molokai

22

16

Maui

16

17

Hawaii

35

33

130

110

10-29

Ocean Wave Energy


Table 10-5
East Coast available wave energy resources by state (TWh per year)
State

Outer Shelf

Inner Shelf

Maine

19

13

New Hampshire

n/a

Massachusetts

45

36

Rhode Island

New York

16

12

New Jersey

14

Delaware

Maryland

Virginia

North Carolina

57

30

South Carolina

24

12

Georgia

Florida

41

36

Total East Coast

237

172

Table 10-6
Gulf of Mexico available wave energy resources by state (TWh per year)
State

Outer Shelf

Inner Shelf

Florida

23

15

Alabama

Mississippi

n/a

Louisiana

29

19

Texas

27

23

Total Gulf of Mexico

83

60

The percent of the total available wave energy that is technically recoverable was estimated for
three assumed array capacity densities, and assuming a 100-fold difference in wave power
density between the threshold operating condition and the maximum operating condition. The
three assumed capacity densities were 10 MW/km and 15 MW/km, which bracket the current
state of technology, and 20 MW/km, which represents an achievable improvement. Percent
recoverable wave energy in Tables 10-7 through 10-9 should be multiplied by the corresponding
values in Tables 10-2 through 10-6 to obtain estimates of technically recoverable wave energy in
units of TWh per year.
10-30

Ocean Wave Energy


Table 10-7
Percent technically recoverable wave energy by region for capacity packing density of
10 MW/km under assumptions described in text
Outer Shelf
Recoverable
Resource

Inner Shelf
Recoverable
Resource

TOC
(MW/km)

MOC
(MW/km)

West Coast
(WA, OR, CA)

31%

37%

300

East Coast
(ME through NC)

57%

70%

200

East Coast
(SC through FL)

67%

78%

100

Gulf of Mexico

68%

71%

100

Alaska
(Pacific Ocean)

29%

46%

300

Alaska
(Bering Sea)

40%

50%

300

Hawaii

54%

56%

200

Coastal Region at
10 MW/km Packing Density

Table 10-8
Percent technically recoverable wave energy by region for capacity packing density of
15 MW/km under assumptions described in text
Outer Shelf
Recoverable
Resource

Inner Shelf
Recoverable
Resource

TOC
(MW/km)

MOC
(MW/km)

West Coast
(WA, OR, CA)

42%

48%

300

East Coast
(ME through NC)

65%

81%

200

East Coast
(SC through FL)

76%

87%

100

Gulf of Mexico

77%

79%

100

Alaska
(Pacific Ocean)

39%

52%

300

Alaska
(Bering Sea)

49%

59%

300

Hawaii

64%

56%

200

Coastal Region at
15 MW/km Packing Density

10-31

Ocean Wave Energy


Table 10-9
Percent technically recoverable wave energy by region for capacity packing density of
20 MW/km under assumptions described in text
Outer Shelf
Recoverable
Resource

Inner Shelf
Recoverable
Resource

TOC
(MW/km)

MOC
(MW/km)

West Coast
(WA, OR, CA)

50%

55%

300

East Coast
(ME through NC)

73%

88%

200

East Coast
(SC through FL)

82%

93%

100

Gulf of Mexico

84%

85%

100

Alaska
(Pacific Ocean)

46%

59%

300

Alaska
(Bering Sea)

56%

65%

300

Hawaii

72%

73%

200

Coastal Region at
20 MW/km Packing Density

Additional results and methodology of the resource assessment are available in an EPRI report
[2]. Results are also displayed via Geographic Information System (GIS) as part of the NRELs
Renewable Energy Atlas, which includes wave power densities for specific coordinates along
with user-selectable annual and monthly statistical products. The database also includes the
probability distributions of sea state parameters, which developers need to predict the annual and
monthly energy yield of their devices and projects. The wave power density, significant wave
height (Hmo), energy wave period (Te), the directions of the primary wind-wave and primary and
secondary swell waves, and bathymetry (iso-baths) are displayed on a GIS map from a depth of
50 meters out to either 200 meters depth or 50 nautical miles from shore (whichever is closer to
the shoreline). The GIS map includes the entire coastline of the United States, including Alaska,
Hawaii, and Puerto Rico.
The influence of the ocean floor reduces wave power levels in shallow waters (<50 m).
Submerged features such as canyons can also focus energy, leading to hot spots in close
proximity to shore. A number of shallow-water wave transformation models take into account
the bathymetry to calculate near-shore wave data. High-resolution bathymetry is required. The
input boundary condition for these shallow water models is the output from the NOAA
WAVEWATCH III model at the edge of the OCS. Experience with shore-based devices shows a
need for extensive modeling in such locations.

10-32

Ocean Wave Energy

10.4.1 Measurement Data Sources


The two largest inventories of long-term measured wave data in the United States are maintained
by the National Data Buoy Center (NDBC) of NOAA (www.ndbc.noaa.gov), and the Coastal
Data Information Program (CDIP) of Scripps Institution of Oceanography (http://cdip.ucsd.edu/).
NDBC data buoys are equipped with strapped-down accelerometers for measuring wave conditions
derived from buoy heave response. Wave spectra are computed from 20-minute time-series
measurements of sea surface elevation changes, and these records are archived at one-hour intervals.
West Coast and Hawaii reference stations are shown in Figures 10-14 and 10-15, respectively.

Figure 10-14
West Coast reference stations

Figure 10-15
Hawaii reference stations
(Point Makapuu is CDIP 0098)

10.4.2 Wind-Wave Model Data Sources


The operational ocean wave predictions of NOAA/NWS/NCEP are performed using the windwave model NOAA WAVEWATCH III using operational products of NCEP as input. The windwave model is implemented on global and regional scales, as shown in Figure 10-16. NOAA
WAVEWATCH III [6, 7] is a third-generation wave model developed at NOAA/NCEP. It is a
further development of the model WAVEWATCH I, developed at Delft University of
Technology, and WAVEWATCH II, developed at NASA, Goddard Space Flight Center. NOAA
WAVEWATCH III differs from its predecessors in many important ways such as its governing
equations, model structure, numerical methods, and physical parameterizations.
NOAA WAVEWATCH III solves the spectral action density balance equation for wave numberdirection spectra. The implicit assumption of this equation is that water depth and current, as
well as the wave field itself, vary on time and space scales that are much larger than that of a
single wave. A further constraint is that the parameterizations of physical processes included in
the model do not address conditions where the waves are strongly depth-limited. These two basic
characteristics imply that the model can generally be applied on spatial scales (grid increments)
larger than 1 to 10 km and outside the surf zone.
10-33

Ocean Wave Energy

Figure 10-16
NOAA Wave Watch III global coverage

All regional models obtain hourly boundary data from the global model. All models are run on
the 0000 GMT, 0600 GMT, 1200 GMT and 1800 GMT model cycles, and start with a 6-hour
hindcast to ensure continuity of swell. The models provide 126-hour forecasts, with the
exception of the North Atlantic Hurricane model, which provides a 72-hour forecast.
Graphical products are maps and spectra; binary and text products are gridded binary (GRIB)
files, spectral data, and spectral bulletins. Detailed parameter definitions can be found in the
NOAA WAVEWATCH III Users Manual [7].
10.4.3 Wave Power Forecasting
Reliable electric power system operation requires precise balancing of supply and demand. Grid
operators manage supply-demand balance on a minute-to-minute basis considering load forecasts
and using current resources, rules, and procedures. Accurate wave power forecasts will help
system operators meet the challenge of integrating wave-generated power with the electric grid.

10-34

Ocean Wave Energy

In 2007, EPRI investigated the accuracy of forecasting wave power as a function of forecasting
time horizon [8]. In the original wave energy study co-funded by the Bonneville Power
Administration (BPA), WAVEWATCH III forecasts were available only at NDBC buoy
locations in very deep water 100 nautical miles or more from the coastline. The EPRI Project
Team accomplished the forecast accuracy study in two steps:
1. Virginia Tech compared the WAVEWATCH III forecast accuracy with far offshore buoys.
2. SAIC correlated time-lagged data from the far offshore buoys to a few buoys at about 50 m
depth, the depth currently favored for offshore wave power plants.
NOAA WAVEWATCH III now forecasts wave sea states shoreward to the 50-m depth contour.
EPRI conducted a study for Pacific Energy Ventures and Oregon Wave Energy Trust in 2009
making a one-to-one comparison of WAVEWATCH III forecasts with co-located measurements
at a NDBC buoy on the 50-m depth contour [9]. Time-lagged correlations between far offshore
and 50-m depth buoys are no longer required.

10.5 Wave Energy Conversion (WEC) Technology Description


Wave power research programs in industry, government, and at universities have established an
important foundation for the emerging wave power industry over the past decade. In the late
1970s and early 1980s, the UK regarded wave power as an alternative to nuclear generation and
had the most aggressive R&D program in the world. Although the program contributed to
important basic research on optimal control and tuning of wave power conversion devices, it
ultimately stalled as oil prices dropped and government funding ceased. In the past decade,
wave-powered generation has advanced, and in the last five years a half dozen full-scale
prototypes have been tested in natural waters.
10.5.1 Harnessing Wave Energy
Wave energy extraction is complex and many device designs have been proposed. Four of
the best known device concepts and their principles of operation are depicted in Figure 10-17:

Point absorber: A bottom-mounted or floating structure that absorbs energy in all


directions. The power take-off system may take a number of forms, depending on the
configuration of displacers/reactors. The illustration shows a floating buoy; however, a point
absorber could be a bottom-standing device with an upper floater.

Oscillating water column (OWC): At the shoreline, this could be a cave with a
blow-hole and an air turbine/generator in the blow hole. Near shore or offshore, an OWC
device is a partially submerged chamber with air trapped above a column of water. As waves
enter and exit the chamber, the water column moves up and down and acts like a piston. A
column of air, contained above the water level, is compressed and decompressed by this
motion to generate an alternating stream of high-velocity air in an exit blowhole. The air is
channeled through an air turbine/generator to produce electricity.

Overtopping terminator: A floating reservoir structure with reflecting arms and a ramp.
Arriving waves overtop the ramp and are contained in the reservoir. The collected water
turns turbines connected to generators as it flows out of the device.

10-35

Ocean Wave Energy

Attenuator or linear absorber: An example of the attenuator principle is a long floating


structure that is orientated perpendicular to the traveling wave front. The structure is
composed of multiple sections that move relative to each other, and this motion activates
hydraulic power converters that drive electrical generators.

Overtopping

Reservoir

Figure 10-17
Wave energy device principles

10-36

Waves
overtopping
the ramp

Ocean Wave Energy

Figure 10-18 shows examples of the four machine types summarized earlier. There are many
other design concepts that are related to these machine types.

OPT Power Buoy

Oceanlinx

Pelamis

Wave Dragon

Figure 10-18
Wave energy device concept

10.5.2 WEC System Developers


Today, a number of small companies are leading the commercialization of technologies to
generate electricity from ocean waves. In 2004, EPRI requested information from all known
WEC device developers [10] and updated the survey in 2006 [11] and 2008, excluding those
with only concepts or patents. Table 10-10 is a list of wave energy conversion device developers
as of November 2011. Survey details are now maintained by the DOE.

10-37

Ocean Wave Energy


Table 10-10
Wave energy conversion device developers as of November 2011
Device Developer(1) & Website

Device Name

Type(2)

Able Technologies
abletechnologiesLLC. com
Advanced Wave Power
www.advancedwavepower.com/
AeroVironment
www.avinc.com/engineering/marine_e
nergy
AlbaTERN
www.albatern.co.uk/index.html
Applied Technologies Company
www.atecom.ru/wave-energy/
Artificial Muscles
www.artificialmuscles.com
Aquamarine Power
www.aquamarinepower.com
Aqua-Magnetics
www.amioceanpower.com/home
Atargis Energy
www.atargis.com/index.html
Atmocean
www.atmocean.com/
AW Energy
www.aw-energy.com
AWS Energy
www.waveswing.com
Balkee Tide and Wave Electricity
Generator
BioPower Systems
www.biopowersystems.com
Bourne Energy
www.bourneenergy.com/future.html

Wave Pipe

Point Absorber

Nautilus

Point Absorber

Eel Grass

Point Absorber

Squid

Attenuator

Float Wave Electric


Power Station
Unknown

Point Absorber

Development
Status(3)
Laboratory Proof of
Concept
Experimental

Point Absorber

Experimental

Oscillating Wave
Surge Converter
Point Absorber

Technology
Demonstration

Cycloidal Wave
Energy Converter
Atmocean

Attenuator

Experimental

WaveRoller

Oscillating Wave
Surge Converter
Point Absorber

Oyster
Electric Buoy

Archimedes Wave
Swing
TWPEG

Point Absorber
Technology
Demonstration
Commercial
Demonstration

Point Absorber

bioWave

Oscillatory

OceanStar

Attenuator

Brandl Motor
brandlmotor.de/index_eng.htm

Brandl Generator

Point Absorber

Experimental

Caley Ocean Systems

Wave Plane

Other

Concept

Carnegie Wave Energy


www.carnegiewave.com/

CETO

Point Absorber

Commercial
Demonstration

C-Wave Limited
www.cwavepower.com

C-Wave

Attenuator

Experimental

Checkmate Sea Energy


www.checkmateuk.com

Anaconda

Articulating linear
absorber

Laboratory Proof of
Concept

College of the North Atlantic


www.cna.nl.ca/news/default.asp?Mess
ageID=789

Wave Pump

Point Absorber

Experimental

Columbia Power Technologies


www.columbiapwr.com/

Multiple devices

Point Absorber

Commercial
Demonstration

10-38

Technology
Demonstration

Ocean Wave Energy


Table 10-10 (continued)
Wave energy conversion device developers as of November 2011
Device Developer(1) & Website

Device Name

Type(2)

Development
Status(3)

ConWEC AS

Controlled WaveEnergy Converter


(ConWEC)

Point Absorber

Experimental

Daedalus Informatics
www.daedalus.gr/weca.html

WECA

Oscillating Wave
Surge Converter

Lab Testing

Dartmouth Wave Energy


www.dartmouthwaveenergy.com/

SeaRaser Buoy

Point Absorber

Experimental

DEXAWAVE Energy
www.dexawave.com/

DEXAWAVE
Converter

Attenuator

Experimental

Dresser-Rand

HydroAir

Oscillating Water
Column

Lab Testing

Ecofys
www.ecofys.co.uk

Waverotor

Hydrodynamic Lift

Experimental

Ecole Centrale de Nantes


www.bulletinselectroniques.com/actualites/52074.ht
m

SEAREV

Oscillating Water
Column

Lab Testing

Ecomerit Technologies
www.ecomerittech.com/centipod.php

Centipod

Attenuator

Experimental

Edinburgh University
www.mech.ed.ac.uk/research/wavepo
wer/sloped%20IPS/Sloped%20IPS%2
0intro.htm

Sloped IPS Buoy

Attenuator

Lab Testing

Edinburgh University, Wave Power


Group

Salter Duck

Attenuator

Experimental

ELGEN Wave

Horizon Platform

Point Absorber

Lab Testing

Sperboy

Experimental

www.sperboy.com

Oscillating Water
Column

Euro Wave Energy

Point Absorber

Experimental

www.elgenwave.com/
Embley Energy

www.eurowaveenergy.com/
Float Incorporated
www.floatinc.org/

Rho-Cee

Oscillating Water
Column

Lab Testing

Floating Power Plant

Poseiden

Attenuator

Technology
Demonstration

Fred Olsen Ltd

Buldra

Point Absorber

Technology
Demonstration

Green Wave Energy


www.gedwardcook.com/

Syphon Wave
Generator

Submerged
Pressure
Differential

Experimental

10-39

Ocean Wave Energy


Table 10-10 (continued)
Wave energy conversion device developers as of November 2011
Device Developer(1) & Website

Device Name

Type(2)

Development
Status(3)

Green Wave Energy


www.gedwardcook.com/

Floating Wave
Generator

Attenuator

Experimental

Green Ocean Energy

Wave Treader

Attenuator

Experimental

Ocean Treader
Green Ocean Wave Energy
greenoceanwaveenergy.com

Ocean Wave Air


Piston

Point Absorber

Experimental

Green Cat Renewables


www.greencatrenewables.co.uk/

Wave Turbine

Attenuator

Experimental

GyroWaveGen

GyroWaveGen

Other

Concept

Hann-Ocean

Drakoo

Point Absorber

Experimental

Hidroflot s.L.
www.hidroflot.com

Ocean Converter

Multiple Pt.
Absorbers

Laboratory Testing

Hydam Technology

McCabe Wave
Pump

Attenuator

Technology
Demonstration

Independent Natural Resources


www.inri.us

SEADOG water
pump

Point Absorber

Early Commercial

Intentium
www.intentium.com/

Offshore Wave
Energy Converter

Other

Concept

Interproject Service
http://www.ips-ab.com/

IPS OWEC Buoy

Point Absorber

Experimental

IWAVE

IWAVE

Point Absorber

Concept

JAMSTEC
www.jamstec.go.jp/jamstece/30th/part6/page2.html

Mighty Whale

Oscillating Water
Column

Technology
Demonstration

Jospa
http://www.jospa.ie/

Irish Tube
Compressor (ITC)

Overtopping/

Lab Testing

hann-ocean.com/

Waveenergy.nualgi.com

Terminator/
Oscilating Water
Column

Joules Energy Efficiency Services

Tetron

Point Absorber

Laboratory Testing

Kinetic Wave Power


www.kineticwavepower.com

PowerGin

Overtopping/

Laboratory Proof of
Concept

Lancaster University Renewable


Energy Group
www.engineering.lancs.ac.uk/lureg/

PS Frog

Point Absorber

WRASPA

Oscillating Wave
Surge Converter

10-40

Other

Laboratory Proof of
Concept

Ocean Wave Energy


Table 10-10 (continued)
Wave energy conversion device developers as of November 2011
Development
Status(3)

Device Developer(1) & Website

Device Name

Type(2)

Langlee Wave Power


www.langlee.no/

Langlee System

Oscillating Wave
Surge Converter

Experimental

Leancon Wave Energy


http://www.leancon.com/

Multi Absorbing
Wave Energy
Converter
(MAWEC)

Oscillating Water
Column

Laboratory Testing

Manchester, Univ of
www.reuk.co.uk/Manchester-BobberWave-Power.htm

Bobber

Point Absorber

Laboratory Proof of
Concept

Motor Wave
www.motorwavegroup.com

Motor Wave

Point Absorber

Experimental

Muroran Institute of Technology

Pendulor

Oscillating Wave
Surge Converter

Experimental

Navatek
www.navatekltd.com

Navatek WEC

Attenuator

Laboratory Proof of
Concept

Neptune Renewable Energy


www.neptunerenewableenergy.com/w
ave_technology.php

Triton

Point Absorber

Technology
Demonstration

Neptune Systems

Magneto Hydro
Dynamics

Other

Concept

Ocean Energy Industries


www.oceanenergyindustries.com/

WaveSurfer

Point Absorber

Commercial
Demonstration

Ocean Energy Ltd


www.oceanenergy.ie

Ocean Energy
Buoy (OEBuoy)

Oscillating Water
Column

Technology
Demonstration

Ocean Harvesting Technologies


www.oceanharvesting.com/

Ocean Harvester

Point Absorber

Laboratory Proof of
Concept

Ocean Navitas
www.oceannavitas.com/index.html

Aegir Dynamo

Point Absorber

Experimental

Oceanlinx (formerly Energetech)


www.oceanlinx.com

Uiscebeatha

Oscillating Water
Column

Commercial
Demonstration

Ocean Power Technologies


www.oceanpowertechnologies.com

PowerBuoy

Point Absorber

Early Commercial

Ocean Wave and Wind Energy


www.owwe.net/

Wave Pump Rig

Point Absorber

Experimental

Ocean Wave and Wind Energy


www.owwe.net/

OWWE Rig

Hybrid Wind-Wave:
Point Absorber and
Overtopping/

Concept

Terminator

10-41

Ocean Wave Energy


Table 10-10 (continued)
Wave energy conversion device developers as of November 2011
Development
Status(3)

Device Developer(1) & Website

Device Name

Type(2)

Ocean Wave Energy Company


www.owec.com

OWEC

Point Absorber

Ocean Wavemaster

Wave Master

Other

Oceantec Energias Marinas

Oceantec Wave
Energy Converter

Attenuator

Offshore Islands
www.offshoreislandslimited.com/

Wave Catcher

Point Absorber

Offshore Wave Energy, Ltd


www.owel.co.uk/

OWEL WEC

Other

Laboratory Proof of
Concept

OreCON Ltd
www.orecon.com

MRC1000

Floating Point
Absorber &
Oscillating Water
Column

Technology
Demonstration

Oregon State Univ


www.eecs.orst.edu/msrf

Various direct drive


buoys

Point Absorber

Technology
Demonstration

Pelagic Power
www.pelagicpower.no/about.html

W2Power

Point Absorber

Experimental

Pelamis Wave Power


www.pelamiswavepower.com

Pelamis

Attenuator

Early Commercial

Perpetuwave Power
www.perpetuwavepower.com/index.ph
p

Wave Harvester

Attenuator

Laboratory Testing

Pontoon Power
www.pontoon.no/

Pontoon Power
Converter (PPC)

Attenuator

Laboratory Testing

Protean Energy Ltd


www.proteanenergy.com/

Energy Conversion
Platform (ECP)

Point Absorber

Experimental

Purenco AS
www.straumekraft.no

Fisherman

Point Absorber

Experimental

Renewable Energy Holdings


www.reh-plc.com

CETO

Point Absorber

Technology
Demonstration

Renewable Energy Pumps


www.renewableenergypump.com

Wave Water Pump

Point Absorber

Experimental

Resen Waves
www.resenwaves.com

Lever Operated
Pivoting Float

Point Absorber

Experimental

Resolute Marine Energy


www.resolutemarine.com/

Point Absorber

Laboratory Proof of
Concept

Resolute Marine Energy


www.resolutemarine.com/

Oscillating Wave
Surge Converter

Laboratory Testing

10-42

Laboratory Proof of
Concept

Experimental

Ocean Wave Energy


Table 10-10 (continued)
Wave energy conversion device developers as of November 2011
Device Developer(1) & Website

Device Name

Type(2)

Development
Status(3)

Ryokuseisha
www.ryokuseisha.com/eng/product/po
wer_supply/index.html

Wave Activated
Generator (WAG)
Buoy

Oscillating Water
Column

Technology
Demonstration

Scientific Applications & Research


Associates
www.sara.com/RAE/ocean_wave.html

MHD Generator

Other

Laboratory Proof of
Concept

Oscillating Wave
Surge Converter

Commercial
Demonstration

SDE
www.sde.co.il/
Seabased AB
www.seabased.com

Direct Driven
Linear Gen

Point Absorber

Laboratory Proof of
Concept

Sea Power
www.seapower.com

Floating Wave
Pres. Vessel

Point Absorber

Experimental

SeaNergy

Submerged
Pressure
Differential

Technology
Demonstration

Seatricity
www.seatricity.net/

Point Absorber

Experimental

SeaVolt Technologies

Wave Rider

Point Absorber

Seawood Designs
www.surfpower.ca

SurfPower

Point Absorber

Experimental

SEEWEC Consortium
www.seewec.org

FO3

Point Absorber

Technology
Demonstration

Sieber Energy
www.wave-energy-accumulator.com

SIE-CAT Wave
Energy
Accumulator

SRI International
http://sri.com

Electroactive
Polymer Artificial
Muscle (EPAM)
Generagor

Point Absorber

Experimental

SyncWave Energy
www.syncwaveenergy.com

SyncWave

Point Absorber

Technology
Demonstration

Surf Buoy
www.cosmotheist.com/Surfbuoy.htm

SurfBuoy

Point Absorber

Experimental

Trement Electric
www.npowerpeg.com/index.php/blog/7
7-npower-wave-energy-converter

nPower WEC

Point Absorber

Proof of Concept

Trident Energy
www.tridentenergy.co.uk/index.php

Direct Energy
Conversion

Point Absorber

Laboratory Proof of
Concept

Versabuoy Intl
www.vbuoy.com

VersaBuoy

Point Absorber

Experimental

Laboratory Proof of
Concept

10-43

Ocean Wave Energy


Table 10-10 (continued)
Wave energy conversion device developers as of November 2011
Device Developer(1) & Website

Device Name

Type(2)

Development
Status(3)

Vigor Wave Energy


www.vigorwaveenergy.com/

Vigor WEC

Attenuator

Laboratory Testing

Voith Hydro Wave Gen


www.wavegen.com

Limpet,

Oscillating Water
Column

Commercial demo
w/Limpet and
breakwater systems;
floating system status
unknown

Nearshore OWC,
Breakwater Turbine

Wave Dragon ApS


www.wavedragon.net

Wave Dragon

Overtopping

Commercial
Demonstration

Wave Energy AS
www.waveenergy.no

Seawave Slot Cone


Generator

Overtopping

Technology
Demonstration

Wave Energy Technologies


www.waveenergytech.com/

WET EnGen

Point Absorber

Experimental

Wave Energy Technology


www.wavenergy.co.nz

WET-NZ

Point Absorber

Experimental

Wave Star Energy


www.wavestarenergy.com

Wave Star

Point Absorber

Technology
Demonstration

Waveberg Development
www.waveberg.com

Waveberg Pump

Attenuator

Experimental

Wavebob Ltd
www.wavebob.com

Wavebob

Point Absorber

Technology
Demonstration

Wavemill Energy

Wavemill

Wavepiston
www.wavepiston.dk

Wavepiston

Oscillating Wave
Surge Converter

Experimental

WavePlane
www.waveplane.com

WavePlane

Overtopping/

Experimental

Wello OY
www.wello.fi/

Penguin

Terminator
Attenuator

Experimental

Notes:
1. This list excludes most individual inventors with conceptual-level-only technology.
2. The principle of operation: Point Absorber, Attenuator, Overtopping/Terminator, Oscillating Water Column, Oscillating
Wave Surge Converter, Submerged Pressure Differential, and Other.
3. The following definition of development status was used:
Laboratory testing stage
Experimental Subscale at sea testing
Technology Demonstration Large size engineering prototype at sea testing whose purpose is to test for function and
performance
Commercial Demonstration Large size manufacturing prototype at sea testing whose purpose is to test for
commercial viability
Early Commercial Offering many units of large size for purposes of generating and selling the electricity produced

10-44

Ocean Wave Energy

10.5.3 Survival in Storms and Hostile Marine Environments


Modern wave energy conversion technologies are designed to survive a 100-year wave and are
the result of years of testing, modeling, and development by many developer organizations. Fullscale prototypes have been deployed, although not continuously, in natural waters since 2004.
The harsh marine environment presents challenges for device survival. Device failures are to be
expected in the prototype stages. Successful operation in the ocean will require significant
investment in ocean engineering. It should be noted, however, that oil and gas platforms are
surviving 50 years or more in similarly hostile marine environments.
10.5.4 Effect of Wave Power Plants on the Environment
Given proper care in siting, deployment, and operations, offshore wave power can be one of the
more environmentally benign electricity generation technologies. Cumulative effects can be
controlled by appropriately limiting deployment at project and regional scales. Early
demonstration and commercial offshore wave power plant projects should include rigorous
monitoring of the environmental effects of the plant and similarly rigorous monitoring of nearby
undeveloped sites in their natural state so that natural variation can be distinguished from longterm effects of wave energy projects. Monitoring and adaptive management plans will likely be
required for most significant wave power projects.
10.5.4.1 Report to Congress: Potential Environmental Effects of Marine and
Hydrokinetic Energy Technologies [12]
Section 633(b) of the Energy Independence and Security Act of 2007 (EISA) called for a report
to Congress that addresses the potential environmental impacts of marine and hydrokinetic
energy technologies, options to prevent adverse environmental impacts, the role of monitoring
and adaptive management, and the necessary components of an adaptive management program.
The EISA Report to Congress was prepared based on a review of peer-reviewed literature,
project documents, and U.S. and international environmental assessments of these new
technologies. The information was supplemented by contacts with technology developers;
experts in state resource, regulatory agencies, and non-governmental organizations; and input
and reviews by federal agencies (NOAA Fisheries, Minerals Management Service [now the
Bureau of Ocean Energy Management, Regulation and Enforcement], U.S. Fish and Wildlife
Service, National Park Service, Bureau of Indian Affairs, FERC).
There are numerous conceptual designs for converting the energy of waves, river and tidal
currents, and ocean temperature differences into electricity. Most of these technologies remain at
the conceptual stage. They have not yet been tested in the field or as prototype, full-scale
devices. Consequently, there have been few studies of their environmental impacts. Most
considerations of the environmental impacts have been in the form of predictive studies and
speculative environmental assessments that have not yet been verified.
The assessments have identified common elements among these technologies that may pose a
risk of adverse environmental effects. Potential impacts include the alteration of currents and
waves; alteration of substrates and sediment transport and deposition; alteration of habitats for
benthic organisms; noise during construction and operation; emission of electromagnetic fields;
toxicity of paints, lubricants, and antifouling coatings; and interference with animal movements
10-45

Ocean Wave Energy

and migrations. Project installation and operation will change the physical environment. Possible
effects on biological resources could include alteration of the behavior of animals, damage and
mortality to individual plants and animals, and potentially larger and longer-term changes to
plant and animal populations and communities. Some effects are expected to be minor, but the
potential significance of many of the environmental issues cannot yet be determined owing to a
lack of experience with operating projects.
Although there have been few environmental studies of these new concepts, a preliminary
indication of the importance of each of these issues can be gained from published literature
related to other technologies (e.g., noises generated by similar marine construction activities,
EMF emissions from existing submarine cables, and environmental monitoring of active offshore
wind farms). Experience with other, similar activities in freshwater and marine systems can also
suggest effective impact minimization and mitigation measures that can be applied to these new
renewable energy technologies. However, some aspects of the environmental impacts are unique
to the technologies, and will require operational monitoring to determine the magnitude and
significance of the effects. This is particularly true for the cumulative effects of large numbers of
ocean energy devices that commercial-scale projects will comprise.
Impacts to bottom habitats, hydrographic conditions, or animal movements that are
inconsequential given a few units may be significant if large, multi-unit projects exploit large
areas in a near-shore ocean area. For some environmental effects it will be difficult to extrapolate
the effects from small to large numbers of units because of complicated, non-linear interactions
between the placement of the machines and the distribution and movements of aquatic
organisms. Assessment of these cumulative effects will require careful environmental monitoring
as the projects are deployed.
Evaluation of monitoring results might be usefully conducted in an adaptive management
framework. There are numerous state and federal agencies and environmental laws and
regulations that will influence the development of marine and hydrokinetic technologies. Federal
licensing of these renewable energy projects is the responsibility of FERC and BOEMRE. Their
licensing decisions include input from other federal and state agencies, tribes, environmental
groups, and other stakeholders. After a licensing decision has been made and operation of the
energy project has begun, the identification (and correction) of environmental impacts will
depend on appropriate monitoring.
The ability to modify a project to mitigate unacceptable environmental impacts identified by
operational monitoring might be based on applying adaptive management principles reflected in
the project license conditions. In the context of marine and hydrokinetic energy technologies,
adaptive management is a systematic process by which the potential environmental impacts of
installation and operation could be evaluated against quantified environmental performance goals
during project monitoring. Early information about undesirable outcomes could lead to the
implementation of additional minimization or mitigation actions which are subsequently reevaluated. An adaptive management process is particularly valuable in the early stages of
technology development, when many of the potential environmental effects are unknown for
individual units, let alone the eventual build out of large numbers of units. Basing the
environmental monitoring programs on adaptive management principles, as advocated by many
resource and regulatory agencies, will take advantage of ongoing research and monitoring to help
refine technology designs and to improve environmental acceptability of future installations.
10-46

Ocean Wave Energy

10.5.5 Permits for Offshore Wave Power Plants


As of August 2, 2012 there were five active FERC-issued preliminary permits, and no pending
preliminary permits (Table 10-11). A preliminary permit gives the permit holder the first right of
refusal to a site for a three-year period to study the site and file a construction license application.
Further information can be obtained from the FERC website:
http://www.ferc.gov/industries/hydropower/indus-act/hydrokinetics/permits.asp.
Table 10-11
FERC active preliminary permits
Docket
No.

Project/Permit Name

Capacity
(MW)

State

P-12743

Douglas Co. Wave & Tidal


Energy

P-12749

Coos Bay OPT Wave Park

P-13666

Filing
Date

Issue Date

Issued/
Pending

OR

10/6/2010

Issued

100

OR

8/10/2010

Issued

Reedsport OPT Wave Park,


Phase III

48.5

OR

3/15/2011

Issued

P-13679

San Onofre OWEC Electricity


Farm

3,186

CA

10/29/2010

Issued

P-13823

Killisnoo Tidal Energy

0.25

AK

1/21/2011

Issued

2/1/2010

8/5/2010

10.5.6 Overview of Regulatory Status for Offshore Wave Power Plants


Agreements between FERC and the Mineral Management Service (now BOEM) in early 2009
have produced the permitting, licensing, and leasing framework depicted in Table 10-12. On July
29, 2012 the Bureau of Ocean Energy Management (BOEM) and FERC announced revised
guidelines for MHK developers pursuing permitting and licensing on the Outer Continental Shelf
(OCS). The revisions further clarify the regulatory process and help streamline the process for
authorizing research and testing of MHK devices.
Table 10-12
FERC, MMS, and state lands permitting, licensing and leasing framework
State Seabed Lands (1)

Federal Seabed Lands (2)

Preliminary Permits

FERC

None

Pilot Licenses

FERC

FERC

Construction and
Operation Licenses

FERC

FERC

State Agency

BOEMRE

Leases

1. To 3 nautical miles except in Texas and Gulf of Mexico states where state lands extend to 12 nautical miles.
2. On the Outer Continental Shelf (OCS).

The regulatory permitting, licensing, and leasing processes associated with U.S. ocean wave
projects can be lengthy, costly, and complex. Indeed, regulatory issues represent a significant
barrier to ocean wave energy development in the United States.
10-47

Ocean Wave Energy

Since 1920, construction and operation of a non-federal hydroelectric project in the United States
has required a license issued by FERC in accordance with the Federal Power Act. In 2004, as a
result of legal interpretation from FERC, wave energy hydroelectric projects were placed under
FERC licensing jurisdiction.
Meanwhile, the 2005 Energy Policy Act (EPAct05) gave jurisdiction for leasing on the Outer
Continental Shelf to the Department of Interiors Mineral Management Service (now BOEM). In
addition to FERC and BOEM, approvals to install and operate a pilot or commercial project are
still required from many other federal, state, and local regulatory agencies (upward of 20
different agencies).
10.5.7 WEC Power Plant Footprints
Use of sea space by wave energy power plants is of critical concern to many environmental
advocacy and ocean user groups. The oceans are held in trust by the federal and state
governments for the good of society as a whole, and these areas are currently used for multiple
purposes (commercial fishing, recreation, commercial shipping, dump sites, military training,
etc.).
The footprints of some of the existing technologies for a 10-MW and a 100-MW wave power
plant are shown in Table 10-13. The Orecon machine, the only tension-moored device in the
table, will have the smallest footprint. Slackly moored devices such as the Pelamis and the
PowerBuoy will require larger footprints. The Oyster is a near-shore device that operates only in
the surge zone.
Table 10-13
WEC device areal footprints
Absorber Dimensions

Footprint Single Unit

Length (m)

Width (m)

Length (m)

Width (m)

Pelamis P1 (1)

123

4.6

300

150

OPT 500 PowerBuoy (2)

18

18

100

100

Aquamarine Power
Oyster (3)

12

18

30

30

Orecon MRC (4)

30

45

245

130

Farm Arrangement Footprint 10 MW


# Devices

# Rows

Length (km)

Width (km)

Pelamis P1 (1)

19

1.0

0.90

OPT 500 PowerBuoy (2)

20

0.70

0.30

Aquamarine Power
Oyster (3)

33

1.0

0.03

Orecon MRC (4)

1.47

0.13

10-48

Ocean Wave Energy


Table 10-13 (continued)
WEC device areal footprints
Absorber Dimensions
Length (m)

Width (m)

Footprint Single Unit


Length (m)

Width (m)

Farm Arrangement Footprint 100 MW


# Devices

# Rows

Length (km)

Width (km)

Pelamis P1 (1)

210

10

1.80

OPT PowerBuoy 500 (2)

200

0.30

Aquamarine Power
Oyster (3)

333

10

0.03

Orecon MRC (4)

68

4.16

0.52

Notes:
1. Based on preliminary design performed by EPRI for Oregon Pelamis Wave Power Plant [13].
2. Based on March 2008 Coos Bay Preliminary Application Document filed with FERC. Project size estimates
based on 100-m lateral spacing and 100 m between rows for PB500 PowerBuoys. Overall 100-MW project size
includes three transit lanes. Each lane is 400 m in length.
3. Based on Aquamarine Power website, 300-kW machine located at 15-m depth and 12 18-m size (EPRI assumed
12-m lateral spacing)
4. Based on input provided by Orecon MRC rated at 1.5 MW.

10.6 Design, Performance, Cost, and Economic Feasibility Issues


10.6.1 WEC Sites
EPRI has investigated the attributes required for a good wave energy site. EPRI site assessments
are documented for Hawaii [14], Washington [15], Oregon [16], and Maine [17]. These and
other reports are available from the EPRIs Ocean Energy website at
http://oceanenergy.epri.com/waveenergy.html#reports.
There are many factors to consider when evaluating potential sites for a wave energy plant. The
primary factors are the following:

Favorable wave energy climate

A nearby harbor with sufficient depth, size, and port infrastructure to support the assembly
and deployment of the plant and maintenance operations

A transmission and distribution system that can flow the power from the wave plant into the
grid and a substation interconnection point close to shore

An easement for the submerged cable from the wave power plant to shore and possibly under
the beach to the substation, similar to the easement required for an outflow pipe

Bathymetry that provides a depth of about 50 m within two to three miles of shore
10-49

Ocean Wave Energy

A sandy seabed (for anchor placement) and a sandy route to shore for trenching the
submerged cable

Minimum conflicts of sea space use (e.g., fishing, crabbing, whale migration)

Absence of environmentally sensitive habitats

Local labor to be trained for employment in this new industry

10.6.2 WEC Design, Performance, and Cost


While tidal and river hydrokinetic technologies can draw from decades of experience with windturbine rotor design, the design drivers for the cost-effective development of wave energy
absorber systems remain to be explored and optimized. This leads to a wide-range of very
different approaches being developed. These substantial differences in design approaches lead to
a wide range of outcomes in terms of performance, cost, and economics. This section does not
attempt to fully quantify this full array of outcomes, but instead captures a range of values
around some of the leading designs under development.
The wave energy sector can be subdivided into onshore, near-shore, and deep-water
technologies. EPRIs cost assessment focuses on devices that are installed in deep water. This
sector has the most significant deployment potential because of limited siting constraints and
generally higher power densities.
In 2004, EPRI performed an Offshore Wave Power Feasibility Definition Study examining five
locations and two WEC technologies. Offshore Wave Power Plant Feasibility Reports have been
published for sites in Hawaii [18], Oregon [13], San Francisco [19, 20], Massachusetts [21], and
Maine [22]. Performance estimates were developed using local wave data obtained from
measurement buoys and performance data supplied by manufacturers.
Two manufacturers provided sufficient information such that EPRI could perform a design and
performance analysis: Pelamis WavePower and Oceanlinx. Cost estimates were developed by
creating a detailed breakdown of the various cost centers and outlines of installation and
operation procedures, and by cross-checking them with a variety of sources, including local
operators, the design team, local manufacturers, and similar offshore projects in the oil-and-gas
and offshore-wind industries.
For the purpose of this assessment, a number of cost centers were defined that would allow
sufficient detail to provide a solid understanding of the impacts of the major cost drivers. The
following are cost centers that contribute to capital cost estimates in this analysis:

Installation: Includes all activities associated with the plant construction, including:
transport, underwater cable installation, mooring installation, device deployment and
commissioning. The plant is assumed to terminate on shore. No cost is included for the
interconnection of the plant with the electrical grid.

Permitting and environmental compliance: Includes all cost components required to permit a
plant. Uncertainties in this cost category are high because there is limited experience to draw
from. These costs include site baseline assessments, environmental baseline studies, and
consultant fees.

10-50

Ocean Wave Energy

Infrastructure: All costs except for those of the devices and moorings themselves, including
underwater cables, cable landing to shore, dockside improvements, and specialized vessels
used to operate the project.

Mooring: Includes all components required to keep the device on-station. These may involve
mooring chains, anchors, piled foundations, shackles, and other mooring components.

Structural: Includes the main structural components, typically built from structural steel,
concrete or fiberglass.

Power take-off: Includes all components and subsystems required to convert the primary
mechanical energy into electricity that can be fed into the electric grid. Subsystems include
items such as a generator, gearbox, hydraulic system, frequency converter, step-up
transformer, and electrical riser cable.

Operating and maintenance (O&M) costs are divided into the following cost centers. To
normalize costs from operational activities, such costs are expressed as present value, divided by
the plants operational life and rated capacity to yield a unit of dollars per kilowatt capacity per
year ($/kW-year). The following are cost centers that contribute to O&M cost estimates in this
analysis:

Insurance: Insurance costs for one-off offshore installations are typically on the order of 2%
of capital cost. As technology matures and technology-related risks are reduced, such
insurance costs are expected to decline to a level similar to that of wind energy today.

Marine monitoring: It is expected that many early-adopter projects will require on-going
monitoring of the plants environmental effects to satisfy regulatory agencies. This may
include measures such as active and passive acoustic monitoring, fish studies, and sediment
transport studies.

Operations: This cost center contains all expenses associated with the operation, maintenance
and repair of the plant. This category includes labor cost, fuel cost, management expenses,
and facility leases.

Replacement parts: Includes the cost of all parts that require replacement of the life of a
project.

The first analysis summarized in the text that follows assesses the cost and performance of a
specific wave energy conversion device in a variety of applications and environments. That
analysis is followed by another completed in 2011 for a generic wave energy conversion device
located off the coast of northern California. When possible, the cost, performance, and economic
assessments presented are based on actual testing and pilot-plant data to inform and drive technoeconomic models.
10.6.2.1 Assessment for Linear Absorber Devices in Massachusetts and Oregon
Table 10-14 summarizes EPRIs cost and performance estimates for a Pelamis-type linear
absorber wave power plant deployed in two size rangesa single-unit pilot plant and a 300,000MWh/yr commercial-scale plantin two locations: on the East Coast at Wellfleet,
Massachusetts, and on the West Coast at Reedsport, Oregon. All cost figures are expressed in
2009 dollars.
10-51

Ocean Wave Energy


Table 10-14
Cost and performance estimates for linear absorber (Pelamis) wave power plants
(December 2009 dollars2)
0.75 MW
Wellfleet MA

103 MW
Wellfleet MA

0.75 MW
Reedsport OR

90 MW
Reedsport OR

1 0.75

206 0.5

1 0.75

180 0.5

Annual Electrical Energy at Busbar (MWeh)

964

300,000

1,000

300,000

Plant Nameplate Rating (MW)

0.75

103

0.75

90

Physical Plant
Seabed, km2
Unit Life, Years

<0.5
20

18.5
20

<0.5
20

16.2
20

Note 1
12

Note 1
24

Note 1
12

Note 1
24

$1,050
$1,528
$367
$2,316
$1,282
$0
$954
$755
Note 3
($91)
$8,159
$693
$8,852

$66
$53
$264
$1,412
$555
$132
$133
$125
Note 3
($13)
$2,728
$330
$3,061

$874
$452
$367
$2,316
$1,282
$0
$10,653
$633
Note 3
($1,663)
$5,318
$452
$5,770

$32
$23
$264
$1,412
$555
$150
$143
$123
Note 3
($126)
$2,577
$254
$2,830

$408

$136

$266

$129

$9,260

$3,197

$6,035

$2,959

N/A7

$0.53

N/A7

$0.48

N/A

$1.07

N/A

$1.03

85

95

85

95

PreCommercial

PreCommercial

PreCommercial

PreCommercial

Simplified

Simplified

Simplified

Simplified

Rated Capacity and Site


Plant Size (number of units unit size, MW)

Scheduling
Development time, Months
Construction time, Months
Capital Cost ($/kW)
On-shore transmission and grid interconnection
Subsea cables
Mooring
Power Conversion Modules
Structural sections
Facilities
Installation
Construction Mgmt and Commissioning
Contingencies
Less State Renewable Inv Tax Credit (4 & 5)
Total Plant Cost (TPC)
AFUDC (interest during construction)
Total Plant Investment
Owner Costs
Due diligence, engineering, permitting, legal,
financial fees, etc.
Total Capital Requirements
Yearly O&M Costs (% of TPC per year)
10 Year One-Time Retrofit Costs (% of TPC)
Unit Availability (%)

Confidence and Accuracy Rating


Technology Development Rating
Design & Cost Estimate Rating

Notes:
1. Development time for permitting is an unknown at this early point with emerging ocean energy technology.
2. The costs are in November 2004 dollars in References 15 through 19 and were adjusted with a 2.5% inflation
rate to Dec 2009 dollars.
3. Contingency costs are built into each of the subsystems.
4. The Oregon credit is 25% of the project cost up to a maximum of $10 million.
5. The Massachusetts credit is 9.5% of the installation cost.
6. Permitting cost is assumed to be 5% of total plant cost; Permitting cost is an unknown at this early point with
emerging ocean energy technology.
7. O&M costs for a pilot plant cannot be estimated.

10-52

Ocean Wave Energy

10.6.2.2 Assessment for Typical Wave Power Device Installed in Northern California
The following cost and economic assessments are based on a hypothetical installation at a
reference site and the following table shows the relevant parameters of the reference deployment
site. In contrast to the assessment above, this analysis assumes generic wave power arrays of
5 MW, 20 MW, and 50 MW operating in an ocean wave environment with the following
characteristics:

Deployment area: Northern California, United States

Average significant wave height (Hs): 2.38 m

Average energy period (Te): 9.3 s

Average power density (P): 33.5 kW/m

Water depth: 70 m

Distance to shore: 5 km

Table 10-15 summarizes the results of the assessment, with costs for each size plant broken
down by capital expenses (CAPEX in $/kW), O&M expenses (OPEX in $/kW-year), plant
performance, and the resulting cost of electricity (cents/kWh). CAPEX includes the costs of
permitting and environment, installation, infrastructure, mooring, structural components, and
power take-off equipment. OPEX includes the costs of insurance, marine monitoring,
replacement parts, and operations.
Figure 10-19 graphically summarizes the CAPEX results for the three sizes of wave energy
plants, and Figure 10-20 does the same for O&M.

10-53

Ocean Wave Energy


Table 10-15
Typical cost, performance and economic profiles for wave energy plants (2011$)
Plant Capacity
5 MW

20 MW

50 MW

Power Take-off

$915

$821

$764

Structural

$3662

$3134

$2828

Mooring

$375

$360

$351

Infrastructure

$249

$133

$103

Installation

$708

$318

$194

Permitting & Environment

$1003

$269

$107

$6913

$5035

$4347

Operations

$90

$49

$34

Replacement Parts

$81

$73

$68

Marine Monitoring

$154

$39

$15

Insurance

$113

$78

$45

$439

$239

$163

6.3%

4.7%

3.7%

Capacity Factor

30%

30%

30%

Availability

95%

95%

95%

CAPEX Contribution

30

22

19

O&M Contribution

22

12

52

34

27

CAPEX ($/kW)

TOTAL
O&M ($/kW-year)

TOTAL
O&M Percent of CAPEX
Performance

Cost of Electricity (cents/kWh)

TOTAL

10-54

Ocean Wave Energy

Figure 10-19
Capital cost breakdown for a typical wave energy technology

$500
$450

OPEX($/kWyear)

$400
$350
$300

Insurance

$250

MarineMonitoring
ReplacementParts

$200

Operations

$150
$100
$50
$0
5MW

20MW

50MW

Figure 10-20
Operational cost breakdown for a typical wave energy technology (in $/kW-year)

10-55

Ocean Wave Energy

Average power density at the deployment site has a significant impact on device performance,
cost and the resulting cost of electricity (COE). Figure 10-21 shows the COE sensitivity to power
density at an installed capacity of 50 MW. This sensitivity was evaluated by developing
performance cost and economic assessments at different potential deployment sites in the United
States and Europe.
35

CoE(cents/kWh)

30
25
20
15
10
5
0

20

25

30

35
40
PowerDensity(kW/m)

45

50

55

Figure 10-21
Cost of electricity as a function of the power density at the deployment site

10.6.3 WEC Economic Feasibility


EPRIs models assume that initial early-adopter issues have been resolved and reliability similar
to that of comparable technologies has been obtained. Because real-world experience is limited,
typical cost and performance estimates must also be developed from technical specifications
supplied by leading device manufacturers. It is important to recognize that there is significant
diversity in technical approaches in this emerging field. It is also important to note that system
performance is heavily influenced by site conditions; resource power density tends to be the
dominant driver of the COE.
Manufacturers typically underestimate cost in the early stages of development. As the
technologies approach commercial maturity, such cost projections increase. The actual
construction and operational costs of a pilot device reveal a more complete economic picture and
provide a solid starting point for further cost studies. Once a technology reaches commercial
maturity, economy of scale applies and volume production begins to drive down costs. This
typical projected cost-trajectory (Figure 10-22) can make it difficult to compare early-stage

10-56

Ocean Wave Energy

technologies with more developed options. Based on its extensive experience with emerging
energy technologies, EPRI has developed a cost estimate rating methodology which assesses the
likely range of uncertainty based on a technologys design maturity and the amount of detail
available for the cost estimate. To reduce the uncertainty in this assessment, only data from
devices that are at a fairly advanced stage of development are used. As a result, all of the cost
data given in this report have an estimated uncertainty range of 30%.
Lab/Idea

Prototype

Commercial

Volume
Production

Cost

Stage of Development
Figure 10-22
Cost projection as a function of development status

It is an established fact that learning through production experience reduces costs. Cost reduction
follows a logarithmic relationship such that for every doubling of the cumulative production of
new wave power units, a specific percentage reduction in production costs would be expected.
The specific percentage used in this study was 82%, which is consistent with documented
experience in the wind energy, photovoltaic, shipbuilding, and offshore oil and gas industries.
As occurred with computers, flat-screen televisions, and wind turbines and PV panels, the costs
of WEC devices will decline as the industry moves toward larger-scale manufacturing and higher
cumulative production. The top line in Figure 10-23 is the industry-documented wind energy
learning curve. This curve was developed by EPRI based on data from a multitude of sources.
Figure 10-23 also shows the lower and higher bounds of the wave energy cost estimates as well.
The 82% learning curve is applied to the wave power plant installed cost but not to the O&M
component of the cost of electricity, which is why the three curves are not parallel.

10-57

Ocean Wave Energy

Wave Low Bound

Wave Upper Bound

Wind

COE (cents/kWh)

100.00

Actual
Wind
COE
History

10.00
Projected Wave
Upper and Lower
COE

1.00
100

1000

10000

100000

Installed Capacity (MW)


Figure 10-23
Levelized COE comparison to wind: Oregon example with federal and state financial
incentives

The adoption of ocean power technologies will be based upon the value of electricity these
devices generate and supply to the grid as well as their ease of integration with the grid. In
addition to the capital cost of installation, O&M will make a significant contribution to the total
cost. O&M costs related to unplanned maintenance is a major factor in the overall cost of
electricity.
EPRIs economic assessments have been based on a wave power plant book lifetime of 20 years.
The cost flow profile for wave energy, much like many renewable energy technologies, is
heavily front-loaded; 90% or more of the COE from wave power plants will derive from initial
capital costs and installation costs. Conversely, typical fossil fuel power plants experience fuel
and ongoing operations that are about 80% of the plant's cost of electricity.
Levelized COE takes into account all fixed and recurring costs of a wave power plant as a
function of the electrical energy it generates. The costs, annual energy produced, and financial
assumptions on which the estimates are based are documented by EPRI [13].
Although there are some federal and state tax incentives to build renewable power systems, these
incentives are insufficient to finance early-adopter, small-scale projects. As a result, developers
will be put in the position of having to push for large commercial installations to drive cost
down, and in the process may be forced to assume the significant technical, economic, and
environmental risks of deploying unproven technologies at large scale.

10-58

Ocean Wave Energy

Dozens of institutional investors in U.S. renewable energy projects pulled out of the market
when the nations liquidity dried up in 2008. Some found more lucrative investments elsewhere
while others found themselves unable to take advantage of tax credits because they lacked the
profits to do so. The American Recovery and Reinvestment Act (ARRA) of 2009 changed the
investor ground rules; however, construction delays engendered by permitting requirements
precluded ocean energy projects from exploiting those funds.
Large insurance companies and investment banks that engage project developers provide the
majority of the renewable energy project financing.

10.7 Installed Capacity and Estimated Growth


Installed offshore wave power capacity as of November 2011 was less than 0.1 MW in the
United States and approximately 4 MW worldwide.
Provided regulatory barriers are overcome and government support and incentives are provided,
it is anticipated that wave energy in the United States could grow rapidly and reach 10,000 MW
rated capacity by 2025.

10.8 Research Focus


Currently DOE National Laboratories are focusing on six critical research areas:
1. Physical Interactions with Devices
a. Fish/marine mammal attraction and avoidance
b. Strike risk to fish/marine mammals
2. Electromagnetic Fields
a. Effects of EMF on fish and marine mammals
3. Acoustics
a. Effects of MHK noise in riverine environments
b. Noise measurement and net-pen studies of select species
4. Toxicity
a. Effects of antifouling coatings on aquatic organisms
5. Benthic Habitat Alteration
a. Development of measurement methodology to evaluate effects of MHK devices
on benthic habitats
6. Data Aggregation and Risk Modeling
a. Development of a publicly available information database for MHK
environmental research

10-59

Ocean Wave Energy

These national laboratories include the following:

Argonne National Laboratory

Idaho National Laboratory

National Renewable Energy Laboratory

Oak Ridge National Laboratory

Pacific Northwest National Laboratory

Sandia National Laboratories

In addition to the research being conducted by the DOE National Laboratories, the department
has also funded the development of National Marine Renewable Energy Centers. As of 2012 the
department has helped establish three National Marine Renewable Energy Centers: Northwest
National Marine Renewable Energy Center (NNMREC), Hawaii National Marine Renewable
Energy Center (HINMREC), and Southeast National Marine Renewable Energy Center
(SNMREC).
The Northwest National Marine Renewable Energy Center is a DOE-funded partnership between
Oregon State University and the University of Washington. Oregon State University is
responsible for wave energy R&D, while the University of Washington is responsible for tidal
energy R&D. The Hawaii National Marine Renewable Energy Center is a partnership between
the DOE and the University of Hawaii, and focuses on wave energy R&D and Ocean Thermal
Energy Conversion (OTEC) R&D. The Southeast National Marine Renewable Energy Center is
a partnership between the DOE and Florida Atlantic University, and focuses on OTEC R&D and
ocean current energy R&D.

10.9 Conclusions
Considerable potential exists for generating electrical power from wave energy off the coast of
the United States and many other places in the world. Wave energy climates are the most
favorable for west-facing coasts between 35 and 55 degrees latitude in the Northern and
Southern Hemispheres. In the United States, the prime locations (i.e., those with a good wave
climate, port infrastructure, and/or coastal grid infrastructure) are the following:

Northern California and Hawaii, which have excellent wave energy climates, good coastal
grid infrastructures, good ports, and high electricity prices.

Oregon, which has excellent wave energy climate, good coastal grid infrastructure, good
ports, but low electricity prices.

Washington, which has excellent wave energy climate, poor coastal grid infrastructure, good
ports, but low electricity prices. The system load is in the Seattle area and there is no
transmission infrastructure to transmit power across the Olympic peninsula.

Alaska, which has excellent wave energy climate, poor coastal grid infrastructure, good
ports, and high electricity prices. The system load is relatively small in the Anchorage,
Fairbanks, and Juneau areas and there is no transmission infrastructure to move the power to
those areas.

10-60

Ocean Wave Energy

The potential recoverable electricity from wave energy resources is estimated by EPRI to be
about 6.5% of todays electricity consumption in the United States. Initial studies suggest that
given sufficient deployment scale, these technologies will be commercially competitive with
other forms of renewable power generation. However, significant technical, economic,
operational, environmental and regulatory barriers remain to be addressed in order to allow this
emerging industry to move progress to commercial development.
The experience related to ocean energy is limited to prototype installations that provide a limited
understanding of economic, operational, environmental, and regulatory issues. It will be
critically important to gain a full understanding of all life-cycle-related issues over the coming
years to pave the way for larger-scale commercial deployments and a successful wave energy
industry.
Such understanding can only be gained from demonstration and commercial systems deployed
by early adopters. First-generation commercial systems will not only address technology-related
issues, but will also provide confidence to regulators, the general public and investors. Both
market push (R&D) and market pull (economic incentives) mechanisms will be required to
successfully move this technology sector forward and develop the capacity to harness wave
energy from the ocean.

10.10 Internet Resources


EPRI: http://oceanenergy.epri.com
European Wave Energy Thematic Network: www.wave-energy.com
Department of Transportation and Industry (UK): www.dti.gov.uk/renewable
Australian Renewables including Wave Energy: www.greenhouse.gov.au/renewable/index.html
Danish Wave Energy: www.waveenergy.dk
European Wave Energy Research Network (EWERN):www.ucc.ie/ucc/research/hmrc/ewern.htm
European Wave Energy Thematic Network: www.wave-energy.net
World Wave Atlas: www.oceanor.no/projects/wave_energy
World Energy: www.worldenergy.org/wec-geis/publications/reports/ser/wave/wave.asp

10.11 References
1.

Zurkinden, A.S., F. Campanile, and L. Martinelli. 2007. Wave energy converter through
piezoelectric polymers. COMSOL Users Conference. Genoble, France.

2.

Mapping and Assessment of the United States Ocean Wave Energy Resource. EPRI, Palo
Alto, CA: 2011. 1024637.

3.

OWET (Oregon Wave Energy Trust). 2009. Utility Market Initiative: Integrating Oregon
Wave Energy into the Northwest Power Grid. Prepared by Pacific Energy Ventures on
behalf of the Oregon Wave Energy Trust. December 2009.

4.

OWET (Oregon Wave Energy Trust). 2009. Wave Energy Infrastructure Assessment in
Oregon. Prepared by Advanced Research Corporation. December 1, 2009.
10-61

Ocean Wave Energy

5.

OWET (Oregon Wave Energy Trust). 2011. Ocean Energy Blue Tag Study: Bridging the
Gap Between the Cost of Wave Energy and its Market Value. Prepared by Robert K.
Harmon & Company, LLC, July 2011.

6.

Tolman, H.L. 1991. A third-generation model for wind waves in slowly varying, unsteady
and inhomogeneous depths and currents. Journal of Physical Oceanography 21: 782-797.

7.

Tolman, H.L. 2009. User manual and system documentation of WAVEWATCH III version
3.14. Environmental Modeling Center, Marine Modeling and Analysis Branch, National
Oceanic and Atmospheric Administration. MMAB Contribution No. 276. May 2009.

8.

Feasibility of Using Wavewatch III for Days-Ahead Output Forecasting for Grid Connected
Wave Energy Projects in Washington and Oregon. EPRI, Palo Alto, CA: 2008. EPRI-WP012.

9.

Wave Energy Forecasting Accuracy as a Function of Forecast Time Horizon. EPRI, Palo
Alto, CA: 2009. EPRI-WP-013.

10. E2I EPRI Assessment: Offshore Wave Energy Conversion Devices. EPRI, Palo Alto, CA:
2044. E2I-EPRI-WP-004-US (Revision 1).
11. California Wave Power Demonstration Project: Bridging the Gap Between the Completed
Phase 1 Project Definition Study and the Next PhasePhase 2 Detailed Design and
Permitting. EPRI, Palo Alto, CA: 2007. EPRI-WP-011-CA.
12. USDOE (U.S. Department of Energy). 2009. Report to Congress on the Potential
Environmental Effects of Marine and Hydrokinetic Energy Technologies. Energy Efficiency
and Renewable Energy, Wind and Hydropower Technologies Program, Washington, DC,
December 2009.
13. System Level Design, Performance and CostsOregon State Offshore Wave Power Plant.
EPRI, Palo Alto, CA: 2004. E2I-EPRI Global-WP-006-OR (Revision 1).
14. E2I EPRI Survey and Characterization of Potential Offshore Wave Energy Sites in Hawaii.
EPRI, Palo Alto, CA: 2004. E2I-EPRI-WP-003-HI (Revision 1).
15. E2I EPRI Survey and Characterization of Potential Offshore Wave Energy Sites in
Washington. EPRI, Palo Alto, CA: 2004. E2I-EPRI-WP-WA-003.
16. E2I EPRI Survey and Characterization of Potential Offshore Wave Energy Sites in Oregon.
EPRI, Palo Alto, CA: 2004. E2I-EPRI-WP-OR-003.
17. E2I EPRI Survey and Characterization of Potential Offshore Wave Energy Sites in Maine.
EPRI, Palo Alto, CA: 2004. E2I-EPRI-WP-003-ME.
18. System Level Design, Performance and CostsHawaii State Offshore Wave Power Plant.
EPRI, Palo Alto, CA: 2005. E2I-EPRI-WP-006-HI.
19. System Level Design, Performance and CostsSan Francisco California Energetech
Offshore Wave Power Plant. EPRI, Palo Alto, CA: 2004. E2I-EPRI-006B-SF.
20. System Level Design, Performance and Costs for San Francisco California Pelamis
Offshore Wave Power Plant. EPRI, Palo Alto, CA: 2004. E2I-EPRI Global-006A-SF.
10-62

Ocean Wave Energy

21. System Level Design, Performance and CostsMassachusetts State Offshore Wave Power
Plant. EPRI, Palo Alto, CA: 2004. E2I Global-EPRI-WP-006-MA.
22. System Level Design, Performance and CostsMaine State Offshore Wave Power Plant.
EPRI, Palo Alto, CA: 2004. E2I-EPRI Global-WP-006-ME.
23. Prioritized Research, Development, Deployment and Demonstration (RDD&D) Needs:
Marine and Other Hydrokinetic Renewable Energy. EPRI, Palo Alto, CA: 2008. 1014762.
24. UKERC (UK Energy Research Centre). 2008. UKERC Marine (Wave and Tidal Current)
Renewable Energy Technology Roadmap. Summary Report.

10-63

11

RIVER IN-STREAM ENERGY

11.1 Introduction
The overall hydrokinetic power potential of U.S. rivers with discharge rates (Q) greater than
113 m3/s and velocities greater than 1.3 m/s was estimated by New York University (NYU) in
1986 to be 12,500 MW, using conservative assumptions of turbine array deployment [1]. The
best river in-stream hydrokinetic resources in the United States are in Alaska and the Pacific
Northwest; however, all states have some hydrokinetic resources, assuming conversion
technology is economical at lower power densities than was assumed in the NYU study. Table
11-1 provides an overview of river in-stream energy and Figure 11-1 shows the major North
America rivers and their annual discharges in km3/year.
Table 11-1
Overview of river in-stream energy

About 300 kW worldwide; 130 kW in the United States.

Estimated annual incremental U.S. capacity additions:

2012: Hundreds of megawatts are in the FERC preliminary permit


process pipeline; the number of projects which will actually be
completed is unknown.

Estimated cumulative capacity by 2025: 500 MW.

Technology Readiness
(as of November 2011)

River in-stream energy conversion (RISEC) is an emerging technology.


Ocean Renewable Power Corporation (USA), HydroGreen (USA), and
New Energy Corporation (Canada) demonstrated their technologies in
small tidal or river applications in 2010.

Economic Status

Results from the first EPRI economic feasibility studies evaluating


remote village applications in Alaska show a simple payback period of 4
to 5 years for two remote villages (avoided cost of electricity is
$0.65/kWh) and 9 to 11 years for one grid-connected village (avoided
cost of electricity is $0.18/kWh). A 2011 assessment of potential
megawatt-scale RISEC applications in the Mississippi River yielded a
cost of electricity ranging from $0.22/kWh to $0.57/kWh. Due to a lack
of real-world experience with RISEC technology, cost estimates are
subject to significant uncertainty.

Installed Capacity
(as of November 2011)

11-1

River in-Stream Energy


Table 11-1 (continued)
Overview of river in-stream energy

In-stream river power projects are likely to be controversial due to


uncertainty regarding environmental impacts. However, small projects
serving isolated communities in remote locations tend not to be
controversial.

Given proper care in siting, installation, operation, and


decommissioning, RISEC technology may be one of the more
environmentally benign generation technologies.

Pilot demonstration testing is needed to understand the interactions


between the devices and their environment. Adaptive management will
be used to incorporate new information into decision-making processes
that will address project build out and cumulative effects

The Federal Energy Regulatory Commission (FERC) has primary


jurisdiction for licensing river in-stream energy under the Federal Power
Act (FPA).

FERC has developed a six-month license application process for pilot


demonstration plants. Alaska Power and Telephone applied for the first
pilot license for its Yukon River at Eagle project. A license at an existing
hydroelectric project (Mississippi Lock and Dam No. 2, Hastings,
Minnesota) has been modified to allow installation of a hydrokinetic
array in the existing projects tailrace.

As of August 2, 2012, there were 81 active FERC-issued preliminary


permits and three pending preliminary permits. Of the issued permits,
75 were for projects on the Mississippi River and 58 were issued to
Free Flow Power Corporation. Furthermore, 43 of the 81 Preliminary
Permits were issued in 2012. Of the pending permits, one is for the St.
Lawrence River (New York) and two for the Kilarc Canal and Creek
(California).

Government Support of
RISEC Technology

DOE initiated a Waterpower R&D Program in FY 2008 with a


congressionally mandated $10 million. FY 2012 received a budget
increase totaling $59million ($34million for MHK and $25million for
conventional hydro). The FY 2013 budget proposal for the water
program is only $20million, which represents a 66% reduction from the
FY 2012 budget.

Trends to Watch

Preliminary permits give project developers an opportunity to evaluate


the permitted site and prepare a license application without
encroachment by other developers for a period of three years. An
important indicator of industry development will be the number and
progress of applications for FERC-issued pilot licenses and commercial
licenses. There are 70 such projects, totaling 7,883 MW, in the pre-filing
phase of the commercial license application process.

Environmental Impact

Regulatory Status

11-2

River in-Stream Energy

Figure 11-1
Major North American rivers and their yearly discharges in km3/year

In North America, the discharge of fresh water is dominated by the Mississippi (Q = 600 km3/yr)
and the St. Lawrence (Q = 440 km3/yr), both of which receive large inputs from the highprecipitation region east of the mountain ranges that dominate the western third of the continent.
Other large rivers in North America include the Mackenzie and the Yukon (which discharge into
Arctic marine waters), the Western Hudson Bay Group, the Columbia, the Sacramento, the
Colorado, and the Hudson rivers.
11.1.1 Main Components of River Hydrokinetic Resource
There are four major components of river in-stream or current flow at a given transect of a river:
discharge, velocity, cross-section area and stage. Discharge is the volume of water moving past a
given cross section of river per unit of time. The units generally used to express this
measurement are cubic meters per second (m3/s). Velocity is the speed of water movement in
meters per second (m/s). Cross-section area is the water area (in m2) in a plane perpendicular to
the flow at a given point along the river. Stage is the height (in meters) of the water level above
an arbitrary zero point. All of these components are interrelated: as water levels increase, so do
discharge and speed; when the stage increases, so does cross-sectional area.
Flow velocity is dependent on the amount of friction between the water and the stream channel.
Smoother channels have less friction and, therefore, faster flow. Channel roughness contributes
to turbulence, which dissipates energy and reduces flow velocity. The Manning equation is used
to calculate average flow velocities in open-channel systems:

k R 2 / 3 S 1/ 2
V
n

11-3

River in-Stream Energy

In this equation, V is the average flow velocity in feet per second or m/s; R is the hydraulic radius
in feet or m, and is the ratio of the cross-sectional area of the stream divided by the length (feet
or m) of channel bottom along that cross section (wetted perimeter), S is the slope of the water
surface, k is a conversion constant that is 1.49 for English units and 1 for metric units, and n is
the Manning roughness coefficient. A larger value of the Manning coefficient yields a smaller
value for flow velocity. Typical values of n are shown in Table 11-2.
Table 11-2
Manning coefficients for representative river substrates
Substrate Type

Manning Coefficient (n)

Mountain streams with rocky beds

0.040.05

Winding natural streams with weeds

0.035

Natural streams with little vegetation

0.025

Straight, unlined earthen channels

0.020

Smoothed concrete

0.012

Current speed varies spatially within any river or channel. Current speed is influenced by
proximity to obstacles such as rocks and vegetation found along the riverbed, which reduce the
flow rate. Therefore, the current is usually fastest in the middle and slowest along the edges and
bottom of the river. In addition, the current is usually faster at the outside of a river bend than on
the inside of the bend. Fast current may also scour out the outer channel if the substrate is
erodable, thereby deepening the channel there.
Erosion and water velocity have a direct relationship, as erosion and the maximum particle size
transported increases with flow velocity. Thus, substrate particle size is sorted by local water
velocity. Just as speed affects erosion, erosion affects speed. The speed of a river affects the
shape of the river itself. At a high gradient, the speed of a river is at its maximum and the shape
of the river is very narrow, stretching out as the slope gradient decreases.
11.1.2 The Conversion of River In-Stream Hydrokinetic Energy to Electricity
A river in-stream energy conversion (RISEC) device harnesses the unidirectional flow of a river
or man-made canal. The power of a river current, in watts, is given by
Pwater = A V3
where A is the cross-sectional area of flow intercepted by the device (in square meters), is the
water density (in kilograms per cubic meter), and V is the water velocity (meters per second).
The energy recovery efficiency and performance of an in-stream energy conversion device can
be estimated using the simplified model of a generic RISEC device as described later. The
calculation addresses the energy collection and conversion efficiency of each step in the process,
beginning with the kinetic energy of the flowing water stream and proceeding through the
turbine, drive train, generator, and power conditioning steps as illustrated in Figure 11-2.

11-4

River in-Stream Energy

Water
Kinetic
Energy

Turbine
Efficiency

Drive
Train
Eff.

Gen
Eff.

Power
Cond.
Efficienc

Grid
Ready
Electric
Energy

Figure 11-2
Steps affecting hydrokinetic turbine efficiency

The individual efficiencies (f) of the RISEC components determine the net usable power that can
be delivered to the grid as described in following equation:
Pdelivered = Pwater fRISEC,
where
fRISEC = fTurbine fDriveTrain fGenerator fPowerConditioning
Because all RISEC devices rely on moving water, and the water must retain enough kinetic
energy to exit the turbine, there is a limit to the portion of kinetic energy that can be removed by
a turbine. The maximum energy that can be extracted by a turbine immersed in a free (unducted,
unconstrained) stream is known as the Betz limit. The Betz limit of 59% is sometimes used as a
theoretical maximum efficiency of a turbine. Typical values for component efficiencies are fTurbine
= 40%, fDriveTrain = 95%, fGenerator = 95%, and fPowerConditioning = 97%. The resulting overall efficiency in
this example is 35%.
Figure 11-3 is an illustration of a typical open-rotor, horizontal-axis RISEC device. The
hydrokinetic energy available to such a water turbine per unit cross-sectional area (in this case,
the area of the rotor blades) is equal to one-half the density of water multiplied by the cube of the
water speed (Figure 11-4).

(P/A)flow = 0.5 x density of water x V3

Figure 11-3
Example of a hydrokinetic turbine

Figure 11-4
Water power density

11-5

River in-Stream Energy

11.2 U.S. In-Stream River Energy Highlights: Late 2011 to Mid-2012


Interest in river in-stream renewable energy continues to grow in the United States. This section
presents a brief accounting of notable federal and state developments and device developer
activities within the sector from late 2011 through mid-2012.
11.2.1 U.S. Federal Highlights
11.2.1.1 Federal Appropriations
In FY 2012, the U.S. Congress appropriated $59 million to the Department of Energy Water
Power Program to conduct research and development (R&D) on advanced water power energy
generation technologies, including both marine and hydrokinetic (MHK) technologies ($34
million for wave, tidal, ocean current, in-stream hydrokinetic and ocean thermal), as well as
conventional hydropower ($25 million for any hydropower technology that uses a dam or
diversionary structure). This marks a significant increase in funding from the $10 million
appropriated in 2008, which was the first year of the Water Power Program. However, the FY
2013 budget proposal announced in February 2012 allotted only $20 million to the Water Power
Program, representing a 66% reduction from the previous year. Fortunately, in April 2012 the
House and Senate Appropriations Committee passed the FY 2013 Energy and Water
Development Funding Bill. The Senate funding measure provides $59 million to the program
($44 million for MHK and $15 million for conventional hydropower) and the House measure
provides $45 million to the program ($25 million for MHK and $20 million for conventional
hydropower).
The DOEs marine and hydrokinetic energy research is focused on assessing the potential
recoverable energy from MHK resources in the United States, and facilitating the development
and deployment of technologies to fully realize this potential. MHK technologies represent an
opportunity for the United States to engage directly in an emerging area of science and
discovery, while developing an entirely new suite of renewable energy technologies available to
reduce emissions and help states meet renewable energy portfolio standard (RPS) targets.
The DOEs priorities for marine and hydrokinetic energy include the following:

Facilitating the deployment of prototypes, and collecting data on the energy conversion
performance and environmental impacts of the devices

Determining the available, extractable, and cost effective resources in the United States

Characterizing and comparing the wide variety of existing marine hydrokinetic technologies

Improving technology performance and reliability, and reducing technology development


costs

Minimizing the cost, time, and negative impacts associated with siting projects

11-6

River in-Stream Energy

11.2.1.2 Funding Awards


The DOE issued a solicitation in April 2009 that resulted in the selection of EPRI in late 2009 to
conduct a national river in-stream hydrokinetic energy resource assessment. That assessment is
ongoing and will be concluded in early 2012. The DOE issued another solicitation in April 2010
that focused on advancing marine and hydrokinetic technology readiness. Awards totaling $37
million were announced in September 2010, $22 million of which were related to RISEC
technology development. Matching funds from non-federal sources will double that amount.
Throughout 2011 and 2012 the DOE offered multiple financial opportunities to MHK industry
businesses, including special loan guarantees and grants. The main types of awards offered to
MHK developers included Small Business Innovation Research (SBIR) programs and Small
Business Technology Transfer (STTR) programs. In FY 2010 the program selected 27 projects
for funding, with individual awards ranging from $160,000 to $10 million. In April 2012 the
DOE announced that there is $9 million available this year to fund approximately 50 small
businesses. Selected projects will receive 1-year awards of up to $150,000. Awardees with
successful projects will also have the opportunity to compete for follow-on funding in excess of
$1 million. These funding programs are intended to help emerging MHK technologies advance
along the DOE Technology Readiness Level (TRL) chain. There are 10 TRL levels, which are
organized as follows:

TRL 13

Concept Definition

TRL 4

Proof of Concept

TRL 5/6

System Integration and Laboratory Demonstration

TRL 7/8

Open Water System Testing, Demonstration, and Operation

TRL 9

Array Testing

TRL 10

Commercialization

Currently, the DOE is funding a large multi-organizational project to develop MHK reference
models. The project began in May 2010 and according to a DOE presentation, the goal of the
project is to, Develop a representative set of Reference Models (RM) for the MHK industry to
develop baseline costs of energy (COE) and evaluate key cost component/system reduction
pathways. The presentation notes that there is a, Need [to identify] COE targets with regard to
technology typeand future innovation opportunities to prioritize research and cost reduction
pathways. The reference model completion date is slated for September 2013. The Point
Absorber Wave Energy Converter (WEC) report was due September 30, 2011. The Oscillating
Water Column WEC and Surge Type WEC reports are due September 30, 2012.
So far in 2012 five Funding Opportunity Announcements have been issued for water power
technologies:
1. Open Funding Opportunity Announcement
Advanced Research Projects Agency-Energy (ARPA-E)
2. Hydropower Advancement Project (HAP) Standard Assessments to Increase Generation
and Value
U.S. Department of Energy Office of Energy Efficiency and Renewable Energy
11-7

River in-Stream Energy

3. In-Water Wave Energy Conversion (WEC) Device Testing Support


U.S. Department of Energy Office of Energy Efficiency and Renewable Energy Water Power
Program
4. FY12 SBIR/STTR Funding Opportunity Announcement
Small Business Funding Opportunity Announcement (FOA)
5. DE-FOA-0000747: RFI Improving Marine and Hydrokinetic and Offshore Wind Energy
Resource Data
U.S. Department of Energy Office of Energy Efficiency and Renewable Energy
11.2.1.3 Federal Regulations
In 2011 FERC attempted to revise permitting and licensing for RISEC projects, specifically
projects on the lower Mississippi River. In an April 1, 2011 letter Jeff Wright, FERC Director of
the Office of Energy Projects, announced plans to deny additional hydrokinetic permits to Free
Flow Power Corporation and Northland Power Mississippi River LLC. Between these two
developers there are 141 issued and pending FERC preliminary permits. FERC stated that:
Given the number and scope of these permit applications, and the fact that it appears unlikely, as a
pragmatic matter, that either applicant will be able to develop and file license applications for more
than a small percentage of these sites during the preliminary permit term, commission staff is
concerned that the Federal Power Act's goal of promoting competition in the comprehensive
development of the nation's waterways would not be furthered by issuing to two applicants such a
large number of permits covering such an expansive portion of a single river.

Despite objecting to FERCs decision, Free Flow Power immediately withdrew 58 of 60


applications and chose not to renew five other permits. Northlands attorney, Carolyn Elefant,
stated:
Here, the commission's proposal to decline to issue any preliminary permits for the Mississippi River
sites is inconsistent with the commission's (1) past precedent on issuance of successive permits and
(2) policy of promoting competition and limiting site banking articulated in the Notice of Inquiry and
Policy Statement Preliminary Permits for Wave, Current and Instream New Technology Hydropower
(February 15, 2007).

In response to the two companies enormous opposition to the decision, FERC reversed its
decision in a June 9, 2011 letter, stating, After reviewing all of the resulting filings, staff has
determined that it is appropriate to continue processing permit applications on the lower
Mississippi River at this time. Additionally, 43 of Free Flow Powers permits were accepted for
filing, whereas 40 of Northland Powers permits were listed as deficient, requiring additional
information.

11-8

River in-Stream Energy

11.2.2 State and Regional Highlights


11.2.2.1 Developments in Alaska
The first RISEC project in the United States was installed in August 2008 in the Yukon River at
Ruby, Alaska, where a 5-kW EnCurrent turbine from New Energy Corporation of Canada
delivers electricity to a remote village grid (Figure 11-5). The project was coordinated by the
Yukon River Intertidal Watershed Council, and the design and installation were performed by
ABS Alaskan. This project is ongoing, with annual deployment of the turbine during the openwater season.

Figure 11-5
Hydrokinetic turbine deployment from barge at Ruby, Alaska, on the Yukon River

Another Alaska project was the first to seek a FERC pilot hydrokinetic license. Alaska Power &
Telephone (AP&T) submitted a draft pilot license application along with a notice of intent to test
a 100-kW hydrokinetic system on the Yukon River at the small community of Eagle. In 2007,
the Denali Commission of Alaska awarded $1.6 million in grant funds to assist in the
development and deployment of the companys river hydrokinetic project. The project was
awarded an additional $1.5 million grant in 2010. A 25-kW EnCurrent turbine, adaptable to a
variety of locations where sufficient tidal or river current flow is available, is being evaluated
over a five-year period. In 2012, the project will relocate to the Alaska Center for Energy and
Power test facility at Nenana, Alaska, on the Tanana River for further studies.
In early May 2009, rapidly rising temperatures in excess of 70F caused a rapid melt-out of
ice and snow along the Yukon River, resulting in unprecedented flooding that nearly wiped out
Eagle. An aerial view (Figure 11-6) shows at least two dozen buildings submerged in a sea of
car-sized ice chunks and 30 feet of muddy floodwater. The Yukon rose 30 feet over its normal
level when 4- to 7-foot-thick ice pans surged downstream, choking the river and bulldozing
islands and shorelines. Despite this setback, the project continued with seasonal deployments of
the 25-kW turbine. Debris management during the open water season also presents a challenge,
although the EnCurrent turbine itself has proven to be robust to debris in deployments through
2011. However, in June 2011 the company surrendered its FERC preliminary permit.

11-9

River in-Stream Energy

Figure 11-6
Village of Eagle during spring breakup of 2009

EPRI System Definition and Feasibility Study at Three Sites in Alaska. This study was
conducted under the sponsorship of the Alaska Energy Authority (AEA), Chugach Electric, and
Anchorage Municipal Light & Power. EPRI first conducted a site survey for six locations in
close proximity to small villages that typically have power needs of a few hundred kilowatts. The
attributes needed for a good RISEC site were characterized. Of particular importance is the river
current velocity profile [2].
Based on the survey results, the projects Alaskan sponsors selected three sites for preliminary
design of a RISEC power plant: Kvichak River at Igiugig, Eagle on the Yukon River, and
Whitestone on the Tanana River. EPRI then prepared the design and estimated the cost to
construct, deploy, operate and maintain this plant. Last, EPRI assessed the technical and
economic feasibility of RISEC technology applied at these sites. The simple payback periods
estimated for the remote villages of Iguigig and Eagle, both of which have an isolated grid, were
three to four years and four to five years, respectively. The simple payback period estimated for
Whitestone, a remote but grid-connected village, was eight to nine years [3]. Detailed studies are
being conducted by the Alaska Energy Authority on the Kvichak River at Igiugig in support of a
project deployment there. Studies include acoustic current Doppler profiler (ACDP), and are
expected to lead to site selection and deployment of a device in 2012.
Alaska Energy Authority Contract Awards. The AEA awarded two contracts to local
universities to characterize the river in-stream resource potential in Alaskan rivers:

University of Alaska Anchorage (UAA): UAA is conducting resource assessments of


hydrokinetic energy potential in rural Alaska. This project is assessing the potential
hydrokinetic energy resources for many sites in rural Alaska by collecting velocity,
bathymetric, and water surface elevation/slope data. UAA will begin by creating a list of
about 24 sites/communities which appear to have the greatest potential. Alaskas larger rivers
(e.g., the Yukon, Koyukuk, Kuskokwim, and Susitna rivers) are obvious places to evaluate
for in-stream energy. UAA is examining available data to develop a list of river stretches and
community partners. The UAA Native Science and Engineering Program (ANSEP) has a
well-established relationship with many communities throughout Alaska and has agreed to
assist the faculty in establishing working relationships with the communities. Working in
cooperation with the communities, UAA is surveying the selected river stretches to obtain

11-10

River in-Stream Energy

bathymetric, current distribution, and water surface elevation/slope data. UAA has installed
water-level gauges at the various study sites to obtain flow/velocity data through the openwater period. Using data from U.S. Geological Survey (USGS) gauging stations, UAA will
estimate the long-term hydrologic (i.e., velocity and depth) conditions at the selected rural
sites.

University of Alaska Fairbanks (UAF): UAF will deploy measurement devices and
characterize the long-term velocity profile at Nenana on the Tanana River. UAF will perform
environmental effects analysis, analyze the effects of cold temperatures and ice on the water
turbine, and plan a water turbine pilot demonstration project.

Project Permitting and Licensing:

Igiugig RISEC Water Power Project (P-13511). On October 29, 2009 the Igiugig Village
Council in Alaska was awarded a preliminary permit for its project, the Igiugig RISEC Water
Power Project (P-13511), located on the Kvichak River in the Lake and Peninsula Borough,
Alaska. The project would utilize 12 RISEC devices, totaling 40 kW. On September 27, 2011
FERC granted the project an 11-month time extension for filing the Notice of Intent and
Draft License Application. As of March 30, 2012 the project continues to submit progress
reports.

Nenana RivGen Power Project (P-13883). On March 2, 2011 ORPC Alaska, LLC. was
granted a preliminary permit from FERC for the Nenana RivGen Power Project (P-13883),
located on the Tanana River in the vicinity of Nenana, Alaska, in the unorganized borough of
Yukon-Koyukuk, Alaska. The project would utilize up to six 50-kilowatt RivGen devices
with a combined capacity of 300 kilowatts. As of February 27, 2012 the project continues to
submit progress reports.

Whitestone Poncelet River-In-Stream-Energy-Conversion Project (Microturbine


Hydrokinetic River-In-Stream-Energy-Conversion Project) (P-13305). On April 17, 2012
Whitestone Power and Communications filed an application with FERC for a pilot project
license, effectively superseding their application for a preliminary permit. The proposed
project would be located on the Tanana River near Delta Junction, Alaska. The project is
anticipated to operate from April through October 2012, and generate approximately 200
MWh/year.

11.2.2.3 Mississippi River Projects


First FERC-Licensed Commercial Hydrokinetic Plant Installed in the Continental United
States. Hydro Green Energy installed a hydrokinetic turbine in the tailwater of the U.S. Army
Corps of Engineers Lock and Dam #2 on the Mississippi River at Hastings, Minnesota, in
January 2009. The 3.6-m diameter turbine, suspended under a barge in the tailwater (Figure
11-7), had a rated power output of 38 kW at 2 m/s. Full commercial power operations
commenced on August 20, 2009 with a 100-kW nameplate-capacity hydrokinetic turbine. Hydro
Green Energy is currently focusing its activities on un-powered, low-head dams.

11-11

River in-Stream Energy

Figure 11-7
Hydro Greens Hydrokinetic Turbine (left) and the deployment site at lock and dam #2 on
the Mississippi River at Hastings, Minnesota (right)

Free Flow Power Mississippi River Proposal. Free Flow Power, a Massachusetts-based
company, proposes to place 180,000 underwater turbines along 500 miles of the Mississippi
River. As of August 2, 2012 the company held 58 preliminary permits from FERC for projects
totaling 7,344 MW authorized capacity. Figure 11-8 shows project locations in the Memphis,
Vicksburg, and New Orleans districts of the U.S. Army Corps of Engineers.
In 2010, Free Flow Power tested interactions between its 3-meter turbine and fish in a flume at
the USGS Conte Anadromous Fish Laboratory in Turners Falls, Massachusetts (Figure 11-9).
Beginning on June 20, 2011, the company began testing its first full-scale hydrokinetic turbine
generator in the Mississippi River at Plaquemine, Louisiana. The turbine is installed on a
research platform that is instrumented to monitor water velocity and turbine performance;
however, the turbine is not grid-connected.

11-12

River in-Stream Energy

Figure 11-8
Preliminary permit sites held by Free Flow Power on the Mississippi River; Memphis
District (upper left), Vicksburg District (upper right), and New Orleans District (lower left)
of the U.S. Army Corps of Engineers

11-13

River in-Stream Energy

Figure 11-9
Pre-test inspection of Free Flow Power 3-meter turbine in flume at the USGS Conte
Anadromous Fish Laboratory

FERC Attempts to Limit Permitting on the Mississippi River. In April 2011 FERC issued notice
to Free Flow Power Corp. and Northland Power Mississippi River LLC stating that the two
companies already hold or have applied for 141 preliminary permits on the Mississippi River and
will not be issued any more. FERC reasoned that it appears unlikely, as a pragmatic matter, that
either applicant will be able to develop and file license applications for more than a small
percentage of these sites during the preliminary permit term, and that holding so many permits
limits fair competition among other developers. After receiving enormous opposition from the
two companies, FERC reversed its decision in June 2011. In 2012, 20 preliminary permits were
issued to Free Flow Power and one was issued to Northland Power Mississippi River LLC.
11.2.2.4 Detroit River Projects
RISEC Project Feasibility Study on the Detroit River. The Detroit/Wayne County Port
Authority (DWCPA), NextEnergy, the Detroit Riverfront Conservancy, and the University of
Michigan School of Naval Architecture and Marine Engineering are engaged in a collaborative
effort to develop and deploy a RISEC system on the Detroit River. Vortex Hydro Energy LLC, a
company set up by the University of Michigan, is developing the patented RISEC technology,
intended to be used at the DWCPAs new Public Dock and Terminal project. On August 2, 2010,
Vortex Hydro Energy LLC conducted an open water test of its latest Vortex Induced Vibration
(VIVACE) converter. The device was tested in the St. Clair River at Port Huron, MI.

11-14

River in-Stream Energy

11.3 Canadian River In-Stream Energy Highlights: Late 2011 toMid-2012


In Canada, there is strong interest in both river in-stream renewable energy and tidal in-stream
generation. This section presents a brief account of notable Canadian activities within the sector
between late 2011 and mid-2012.
RISEC Project Deployment in the St. Lawrence River at Cornwall, Ontario. Verdant Power
Canada is developing the Cornwall Ontario River Energy (CORE) Project on the St. Lawrence
River (Figures 11-10 and 11-11). The CORE project has received grant support through the
Ontario Ministry of Research and Innovation (MRI) for a three-year kinetic hydropower
demonstration project in the fast-flowing waters of the St. Lawrence, 3.5 km downstream of the
bi-national New York Power Authority-Ontario Power Generation Moses-Saunders hydroelectric
dam. The project, to be deployed in three stages, will primarily test the foundation and
mountings of Verdants proprietary Kinetic Hydro Power System (KHPS) turbine for fastflowing riverine systems. In addition, the demonstration project will continue to expand the
science and evaluation of RISEC technology impacts on aquatic resources through a testing
program undertaken with the St. Lawrence River Environmental Institute of Science. If
successful, the project could be expanded in phases to 5 MW and potentially to as much as
15 MW in the general vicinity (see www.verdantpower.com for more information). The Phase I
deployment has been delayed due to permitting issues.

Figure 11-10
Location of Verdants Cornwall Ontario Renewable Energy (CORE) Project in the St.
Lawrence River

11-15

River in-Stream Energy

Figure 11-11
Site of the CORE Project on the St. Lawrence River (left), and Verdant Turbine (right)

Deployment on a Microgrid at Fort Simpson. A 25-kW EnCurrent hydrokinetic turbine was


deployed at Fort Simpson on the Mackenzie River, Northwest Territories, Canada, during the
summer of 2010. The microgrid-connected turbine was removed for the winter season and was
reinstalled following ice-out in the spring of 2011. Floating debris caused shutdown of the
system three times each in 2010 and 2011. Equipment limitations in this remote location posed
challenges to system deployment and retrieval (Figure 11-12).

Figure 11-12
Positioning the anchor for the EnCurrent turbine deployment at Fort Simpson, NT, Canada

11-16

River in-Stream Energy

Upcoming and In-Progress Canadian Projects. New Energy Corporation is currently working
on several projects to deploy their EnCurrent hydrokinetic turbines, including the following:

30 kW Pointe du Bois, Manitoba, Canada

10 kW Coffee Creek, Yukon Territory, Canada

500 kW Canoe Pass Tidal Energy Demonstration Project, Campbell River, British Columbia

25 kW Cardinal, Ontario, Canada

25 kW Fort Simpson, Northwest Territories, Canada

5 kW NDulee Ferry Crossing, Northwest Territories, Canada

11.4 River In-Stream Power and Energy Resources


EPRI has performed in-stream river resource power and energy assessments for six sites in the
state of Alaska. For the Alaska site survey, only USGS gauging station transects were used.
EPRI obtained the hand-written calibration sheets from the USGS Office in Anchorage, Alaska.
Following is a general description of how the river speed profile was obtained for the Yukon
River site at Eagle, Alaska, which could represent an analytical approach that would prove useful
at other sites.
EPRI researched available data from USGS gauging stations and developed a methodology for
estimating in-stream hydrokinetic resources and predicting its performance. USGS calibrates
most of its 420 measurement stations in Alaska six times per year. To do this, the USGS
establishes a cross-sectional profile at the calibration cross section, measures velocities at various
locations throughout the cross section, calculates the water discharge volume, and correlates the
discharge or volumetric flow rate with the water level in the river.
The USGS maintains a stream gauging station on the Yukon River at Eagle and has recorded
daily discharge data for 57 years. To develop the river speed profile, a relationship was first
established between discharge rate and velocity. The relationship was then applied to the full
data set to determine the statistical parameters shown below in Figures 11-13 through 11-18.
Note that the velocity profiles and associated power densities are valid only for the transect that
the USGS used to calibrate the flow data and which EPRI used to calibrate the velocity data.

11-17

River in-Stream Energy

3.00

Velocity (m/s)

2.50
2.00
1.50
1.00
0.50
0.00
0

50

100

150

200

250

300

350

400

450

500

Distance from shore (m)

Figure 11-13
Stream velocity profile across the Yukon River at the USGS Gauging Station at Eagle,
Alaska for a discharge rate of 183,000 ft3/s

0.0

Water Depth (m)

-2.0
-4.0
-6.0
-8.0
-10.0
-12.0
0

50

100

150

200

250

300

350

400

450

500

Distance from shore (m)

Figure 11-14
Channel cross section at the USGS Gauging Station at Eagle, Alaska, at a discharge rate
of 183,000 ft3/s

11-18

River in-Stream Energy


8
y = 1.9989Ln(x) - 17.952
7

Velocity (ft/s)

6
5
4
3
2
1
0
0

50000

100000

150000

200000

250000

300000

Discharge (ft^3/s)
Velocity

Fitted

Log. (Velocity)

Figure 11-15
Velocity versus discharge at the USGS Gauging Station at Eagle, Alaska

35%
30%

Frequency

25%
20%
15%
10%
5%

2.
85

2.
65

2.
45

2.
25

2.
05

1.
85

1.
65

1.
45

1.
25

1.
05

0.
85

0.
65

0.
45

0.
25

0.
05

0%

Velocity (m/s)

Figure 11-16
Velocity distribution at the USGS Gauging Station at Eagle, Alaska

11-19

River in-Stream Energy

250,000

200,000

ft^3/s

150,000

100,000

50,000

0
Jan

Feb

Mar

Apr

May

Jun

Jul

Aug

Sep

Oct

Nov

Dec

Month

Figure 11-17
Average discharge by month at the USGS Gauging Station at Eagle, Alaska

9.00
8.00
7.00
Velocity (ft/s)

6.00
5.00

Average Vel ft/s

4.00

Max Vel ft/s

3.00

Min Vel ft/s

2.00
1.00
0.00
-1.00

Jan

Feb Mar

Apr

May Jun

Jul

Aug Sep

Oct

Nov Dec

Month

Figure 11-18
Average velocity by month at the USGS Gauging Station at Eagle, Alaska

11.5 River In-Stream Energy Conversion Development


A RISEC device differs from a tidal wave energy conversion device in that it does not need to
accommodate bi-directional flow, nor is it exposed to a salt water environment. However, large
floating debris must be coped with in the riverine environment.

11-20

River in-Stream Energy

11.5.1 Harnessing In-Stream River Energy


In-stream river energy extraction is complex and many device designs have been proposed. It is
helpful to describe their physical arrangements and energy conversion mechanisms, which may
be classified as one of the following three types:

Axial flow: the axis of rotation is parallel to the direction of water flow.

Cross flow: the axis of rotation is perpendicular to the water stream and may be oriented at
any angle, from horizontal to vertical, with respect to the water surface.

Non-turbine: oscillatory hydrofoil, vortex induced motion, or hydro Venturi device.

Figure 11-19 illustrates axial and cross-flow types of turbines.


Axial and Cross-flow Turbines

Marine Current Turbines Axial Turbine


(Courtesy Marine Current Turbines)

Lucid Energy Cross-flow Vertical Turbine


(Courtesy Lucid Energy)

ORPC Cross-flow Horizontal Turbine


(Courtesy ORPC)

Figure 11-19
River in-stream energy conversion devices

11-21

River in-Stream Energy

The subsystems for a hydrokinetic turbine typically include a blade or rotor, which converts the
kinetic energy in the water to rotational shaft energy; a drive train, which usually comprises a
gearbox and a generator; a support structure for the rotor and drive train; and other equipment,
including controls, electrical cables, and interconnection equipment.
Additional ways of classifying these devices include the following:

Support structures: These may be either gravity-foundation-based, attached to a monopile


foundation, or anchored and moored and allowed to fly in the stream using buoyancy
and/or dynamic pressure forces. RISEC devices may also be hung from a floating platform
that is anchored and moored to maintain its position in the flow.

Open vs. ducted: Turbine rotors may be either open, much like wind turbines, or enclosed in
a duct or shroud. Since the energy extracted from the current is a function of the crosssectional area of the flow directed through the swept area of the turbine, using a duct of a
given cross-sectional area is the equivalent of using an open rotor with the same crosssectional area. The wind energy industry has found that adding length to a rotor blade is more
economical than adding a duct to increase the cross-sectional area and power of a wind
machine. However, the economics may differ for water turbines.

Fixed vs. variable-pitch blades: Pitch control is used to limit power, maximize the efficiency
of the turbine in variable flows, and enable bi-directional operation. There are other ways to
accomplish these three functions with fixed-blade turbines and many different design
concepts for implementing them.

Closed-center vs. open-center hubs: Instead of a fixed hub and rotating blades, a design
variation uses an outer fixed rim and an inner rotating bladed disc. One potential benefit of
an open-center design is eliminating the need for a gearbox by encapsulating the stator of a
generator on the rim of the machine.

Savonius vs. Darrieus vertical-axis turbines: First invented in Finland, the Savonius turbine
is S-shaped when viewed from above. This drag-type turbine turns relatively slowly, but
yields a high torque. The Darrieus turbine was invented in France in the 1920s. Often
described as looking like an eggbeater, this turbine has vertical blades that rotate into and out
of the flow. Using hydrodynamic lift, these turbines can capture more energy than drag
devices.

Helical vs. cycloidal: Aerodynamic lift-type vertical-axis turbine blade configuration. A


helical blade traces a three-dimensional curve that lies on a cylinder or cone, such that its
angle to a plane perpendicular to the axis is constant. A cycloidal blade has the shape of a
curve traced by a point on the circumference of a circle that rolls on a straight line.

Non-turbine types are also being investigated, such as hydro Venturi and hydrofoil or vortex
induced oscillation. A few examples are described in the following subsection.
11.5.2 RISEC Technology Developers
Today, a number of entrepreneurial companies are leading the commercialization of in-stream
river energy conversion technologies. Any tidal in-stream energy conversion (TISEC)
technology can be applied to river in-stream energy conversion (RISEC). Table 11-3 lists known
RISEC/TISEC developers that are developing in-stream river applications as of August 2, 2012.
11-22

River in-Stream Energy


Table 11-3
In-stream river energy conversion device developers
Device Developer 1
Website

Device Name

Type

Development
Status 2

Unknown

Oscillatory

Laboratory

FFP Turbine
Generator

Axial Horizontal
Axis

Technology
Demonstration

Lucid Energy
www.licidenergy.com

Gorlov Helical
Turbine (GHT)

Cross Flow
Vertical Axis

Technology
Demonstration

Hydro Green Energy


www.hgenergy.com

Krouse Turbine

Axial Horizontal
Axis

Commercial
Demonstration

EnCurrent Turbine

Cross Flow
Vertical Axis

Commercial
Demonstration

OCGen

Cross Flow
Horizontal Axis

Commercial
Demonstration

Underwater Electric
Kite

Axial Horizontal
Axis

Commercial
Demonstration

Verdant Power
www.verdantpower.com

Free Flow Turbine

Axial Horizontal
Axis

Commercial
Demonstration

Vortex Hydro
www.vortexhhydro.com

VIVACI

Oscillatory

Laboratory

Microturbine River
In Stream

Axial Horizontal
Axis

Laboratory

AeroHydro Research and Technology


www.ahrta.com
Free Flow Power
www.freeflowpower.com

New Energy Corporation


www.newenergycorp.ca
Ocean Renewable Power Corp
www.oceanrenewablepower.com
UEK
www.uekus.com

Whitestone Power & Communications


Notes:

1. This list excludes individual inventors with conceptual-level technology.


2. The following definitions of development status were used:
Laboratory testing stage
Experimental: Sub-scale at sea testing
Technology Demonstration: Large engineering prototype field (in water) testing whose purpose is to test for function
and performance
Commercial Demonstration: Large manufacturing prototype field (in water) testing whose purpose is to test for
commercial viability
Early Commercial: Offering many large units for purposes of generating and selling the electricity produced.

11.5.2.1 Aero Hydro Research and Technology Associates


Aero Hydro Research and Technology Associates (AHRTA) is developing a hydropower
generator that uses oscillating wings to convert the flow energy of rivers and tidal streams into
electrical energy. AHRTAs concept manipulates a hydrofoil to oscillate in both plunge (pure
translation) and pitch (rotation about some axis on the hydrofoil chord line) to extract energy
from water flow. The phase angle between the pitch and plunge oscillations must be close to 90
degrees. AHRTA has constructed an experimental model which enforces the plunge and pitch
oscillation with the proper phasing between the two motions, as shown in Figure 11-20. The

11-23

River in-Stream Energy

model has two hydrofoils arranged in a tandem configuration so that the two hydrofoils also
operate with a 90-degree phasing. Thus far, the model system has operated satisfactorily.
AHRTA is now in the process of developing a new model with a simpler mechanism to enforce
the phasing between the pitch and plunge motion.

Figure 11-20
Experimental configuration of the AHRTA oscillating turbine

11.5.2.2 Free Flow Power


Free Flow Power (FFP) is developing a RISEC turbine system, called SmarTurbine, that uses a
rim-mounted, permanent-magnet, direct-drive generator to maximize efficiency, with front and
rear diffusers and one moving part, the rotor. The generator uses a start-up bearing and a
combination of magnetic levitation and hydrodynamic bearings. At a flow of 9 feet per second
(2.7 m/s), the turbine can produce 20 kW.
Magnetic arrays, using rare-earth neodymium magnets, provide high field strength for greater
efficiency and lower harmonic content. This arrangement facilitates easier grid synchronization
than traditional bipolar magnet arrays. Meanwhile, the generators rotor is designed to operate
over a wider range of flow speeds (2 to 5 m/s). Figure 11-21 illustrates both the cross section and
component parts of the FFP turbine generator and an experimental rotor.

Figure 11-21
Free Flow Power SmarTurbine generator

11-24

River in-Stream Energy

FFP plans to place six to 12 turbines in arrays on pilings, 25 feet off the bottom of a river and at
least 40 feet below the surface to stay clear of ships and boats. The company expects to
manufacture two versions of the Free Flow Turbine Generator: a 2-meter turbine designed to
generate 10 kW at a flow velocity of 2 m/s, and a 1-meter turbine designed to generate 10 kW at
a flow velocity of 3 m/s.
Currently, all of Free Flow Powers hydrokinetic projects are located on the Mississippi River.
The company is investigating the feasibility of RISEC projects at several locations along the
river and holds 58 FERC-issued preliminary permits. In April 2011 FERC issued notice to Free
Flow Power Corp. and Northland Power Mississippi River LLC stating that the two companies
already hold or have applied for 141 preliminary permits on the Mississippi River and will not be
issued any more in order to promote fair competition among other developers. After receiving
enormous opposition from the two companies, FERC reversed the decision in June 2011. In 2012
20 preliminary permits were issued to Free Flow Power and only one was issued to Northland
Power Mississippi River LLC.
11.5.2.3 Hydro Green
Houston-based Hydro Green Energy, LLC has developed and patented a hydrokinetic turbine
array (HTA) system. The company intends to operate as an independent power producer (IPP),
selling the power generated from its HTAs via long-term, wholesale power purchase agreements
(PPAs), although the company has recently focused its attention on low-head hydropower
applications. Figure 11-22 illustrates a 2-by-2 hydrokinetic HTA configuration, and Figure 11-23
is an underwater illustration of the patented hydrokinetic in-stream river current device array
configuration.

Figure 11-22
Hydro Green turbines

11-25

River in-Stream Energy

Figure 11-23
Hydro Green turbine array configuration

11.5.2.4 Lucid Energy Technologies


Formed in March 2007, Lucid Energy Technologies is a joint venture between GCK
Technology, Inc. and Vigor Clean Tech, Inc. The company is focusing on designing and
commercializing complete hydrokinetic electricity generation systems based on the Gorlov
Helical Turbine (GHT). Figures 11-24 and 11-25 show a Lucid Energy turbine prototype and an
array configuration, respectively.

Figure 11-24
Lucid Energy Gorlov Helical Turbine

11-26

River in-Stream Energy

Figure 11-25
Lucid Energy array configuration

11.5.2.5 New Energy Corp.


New Energy Corp. is the Canadian manufacturer of the proprietary EnCurrent turbine, whose
design is based on the Darrieus wind turbine. When the turbine rotor is placed within a water
current, the hydrofoils generate a lift vector in the forward direction and the energy is captured as
shaft rotation. The hydrofoils experience their maximum forward torque at the top and bottom of
their rotation, when the water moving past them is tangential. The turbine rotates in the same
direction regardless of the direction of the water current and captures between 35% and 40% of
the energy in moving water. It rotates at a very low speed, with the blade speed between 2 and
2.5 times the speed of the water passing through the device.
One of the unique properties of the Darrieus turbine design is that it is able to capture the energy
from the water regardless of the direction of the current. This feature enables the EnCurrent
turbine to harness the energy of both flood and ebb tides in tidal applications. A permanent
magnet generator is mounted on the turbine shaft to convert the torque generated by the rotor
into electricity. The output from the permanent magnet generator is a variable-frequency AC
current which is rectified to DC and then converted to constant-frequency AC by an inverter.
New Energy currently manufactures 5-, 10-, and 25-kW models of the EnCurrent Power
Generation System, and is developing 125- and 250-kW models. New Energy also provides a set
of ancillary products that support the installation of the EnCurrent Power Generation System.
Figure 11-26 illustrates such a device mounted on a double-hull pontoon boat, and Figure 11-27
illustrates the turbine design. The generator is mounted on the shaft above the water surface.

11-27

River in-Stream Energy

Figure 11-26
New Energy system concept

Figure 11-27
New Energy EnCurrent turbine

11.5.2.6 Ocean Renewable Power Corp.


Founded in 2004, Ocean Renewable Power Co. is developing a proprietary RISEC turbine called
the River Generation Turbine Generating Unit (RivGen TGU), based on the companys Ocean
Current Generation TGU. The turbine rotates in one direction regardless of the stream flow
direction. Two cross-flow turbines drive a permanent magnet generator on a single shaft. Figure
11-28 shows the OCGen TGU.
RivGen modules are held in the current on a tripod deployed on the river bottom. Power output
up to 50 kW is reportedly achievable in a 10 ft/sec current.

Figure 11-28
Ocean Renewable Power Co. OCGen module

In mid-May 2007, ORPC initiated an OCGen TGU demonstration project in tidal currents
in the Western Passage (Passamaquoddy Bay) near Eastport, Maine. The demonstration was
completed in early 2008, and it successfully proved the basic design and technical feasibility of
the OCGen TGU. Data were also collected for use in the subsequent TGU commercial designs.
ORPC has continued to test improved designs at Eastport and was the recipient of a $10 million
grant from DOE to facilitate commercialization of its technology. ORPC holds a preliminary
permit for a site on the Tanana River, Alaska, where the RivGen TGU may be deployed.
11-28

River in-Stream Energy

On March 2, 2011 ORPC Alaska, LLC. was granted a preliminary permit from FERC for the
Nenana RivGen Power Project (P-13883), located on the Tanana River in the vicinity of Nenana,
Alaska, in the unorganized borough of Yukon-Koyukuk, Alaska. The project would utilize up to
six 50 kW RivGen devices with a combined capacity of 300 kW.
11.5.2.7 UEK Corp.
UEK Corp. was founded in 1981 by Philippe Vauthier, an early visionary in TISEC technology.
Since his death in 2008, the company has been led by his widow Denise Vauthier. UEK stands
for Underwater Electric Kite, Vauthiers description of a system that consists of two joined,
free-flow turbines tethered to the river or sea bed. Figure 11-29 illustrates deployment of a
prototype in the tailwater of a conventional hydropower project.

Figure 11-29
UEK prototype demonstrated in 2000
Source: UEK

11.5.2.8 Verdant Power


Founded in 2000, New York-based Verdant Power has fabricated, tested, and deployed four
working marine energy system prototypes. Dubbed the Free Flow, the system consists of arrays
of three-bladed, horizontal-axis turbines that resemble and operate like wind turbines (Figures
11-30 and 11-31). The turbine rotor spins at approximately 32 rpm, driving a gearbox to increase
rotational speed and power the grid-connected generator. The gearbox and generator are encased
in a waterproof, streamlined nacelle mounted on a streamlined pylon.

11-29

River in-Stream Energy

Figure 11-30
Verdant Power Free Flow turbine

Figure 11-31
Verdant Power turbine lowered into the East River prior to mounting on a monopile

Verdant Powers Free Flow turbines can operate in both tidal and river settings. Turbines
deployed in tidal settings are assembled with internal yaw bearings, which allow the turbines to
pivot with the changing tide. Turbines designed for rivers are fixed. Depending on the site,
various types of devices can be used to anchor the turbines underwater. Verdant is currently
pursuing the permits required for a phased commercial project in Canadian waters of the St.
Lawrence River near Cornwall, Ontario (CORE project).
Following the successful testing of Verdants fourth- generation Free Flow device at the RITE
project from 2006 to 2009, Verdants fifth-generation device is planned for installation at that
site, as well as the CORE project site. As of January 23, 2012, Verdant Power is currently the
first developer to hold a commercial pilot project license from FERC for a tidal power project,
which utilizes Verdants TISEC/RISEC Free Flow device.

11-30

River in-Stream Energy

11.5.2.9 Vortex Hydro


Founded in 2004, Vortex Hydro Energy LLC is a Michigan company that has developed a
technology it calls Vortex Induced Vibrations Aquatic Clean Energy (VIVACE). As its name
suggests, VIVACE uses vortex-induced vibrations (VIV) to extract energy from ocean, river,
tidal, and other water currents.
For decades, engineers have tried to prevent VIV from damaging offshore equipment and
structures. As shown in Figure 11-32, VIV results from vortices forming and shedding on the
downstream side of a bluff body in a current. Vortex shedding alternates from one side to the
other, thereby creating a vibration or oscillation. The VIV phenomenon is nonlinear and selfregulated, which means it can produce useful energy at high efficiency over a wide range of
current speeds. VIVACE works by maximizing and exploiting VIV rather than preventing it.

Figure 11-32
Vortex-induced vibrations oscillating objects in fluid currents

VIVACE devices can be positioned below the surface, thereby avoiding interference with other
river uses such as fishing, shipping, and tourism. In addition, VIVACE utilizes vortex formation
and shedding, the same mechanism fish use to propel themselves through the water, which
makes it compatible with marine life.
A VIVACE prototype was tested in the St. Clair River, Michigan in late 2010, and Vortex Hydro
expects to begin commercializing the technology by 2013.
11.5.2.10 Whitestone Power & Communications
Whitestone Power & Communications is a publicly owned company in Delta Junction, Alaska.
Whitestones axial flow device, Microturbine River In Stream, is currently listed at DOE
Technology Readiness Level (TRL) 5/6: System Integration, and Technology Laboratory
Demonstration. A diagram of the device is shown in Figure 11-33, and the DOE MHK
Technology Database lists the following description of the device:
HDPE blades are the only moving parts in the water. This gives the turbine high resistance to silty
or salty water. Blades designed to survive impact of 1500 lb object. HDPE provides flexibility
and strength. Blades penetrate water 24 inches allowing for deep and shallow operation.
Mounting design allows for variable depth operation for varying river conditions. All submerged
prime-mover parts constructed from HDPE. No underwater gearboxes, generators or electrical
cables. Velocity of blades 50% of velocity of river current.

11-31

River in-Stream Energy

Figure 11-33
Whitestone Power & Communications Microturbine River-In-Stream device

The project, Whitestone Poncelet River-In-Stream-Energy-Conversion Project (aka Microturbine


Hydrokinetic River-In-Stream-Energy-Conversion Project) (P-13305) filed an application with
FERC for a pilot project license, effectively superseding the application for a preliminary permit,
on April 17, 2012. The proposed project would be located on the Tanana River near Delta
Junction, Alaska. The project is anticipated to operate from April through October 2012, and
generate approximately 200 MWh/year.
11.5.3 Survival in Hostile River Environments
In-stream river energy conversion systems that are submerged below the surface do not bear
the full brunt of ice and floating debris. In Alaska, rivers normally begin to freeze in October,
and freezing to a thickness of 4 to 8 feet is not uncommon. Ice breakup is potentially destructive,
with large pieces of ice scouring the river bottom and edges. Figure 11-34 depicts an ice jam
during a spring breakup in the Yukon River at Eagle, Alaska. Ice accumulation is a potential
problem even in areas of open water during winter in northern climates. Liquid water splashed
above the waterline freezes and accumulates layer by layer potentially reaching large volume and
mass. Free-floating ice crystals, called frazil ice, can stick to underwater structures, increasing
drag on hydrokinetic equipment.

11-32

River in-Stream Energy

Figure 11-34
Ice jam during spring breakup: upstream view of the Yukon River at Eagle, Alaska

11.5.4 Environmental Impacts of In-Stream River Power Plants


Given proper care in siting and scaling projects, river in-stream power promises to be one of the
most environmentally benign electrical generation technologies. Early demonstration and
commercial river in-stream power plants should include rigorous monitoring to support
assessment of impacts on the aquatic environment.
EPRI recently completed desktop and laboratory flumes studies of interactions between
hydrokinetic turbines and fish [4, 5].
11.5.5 Permits for In-Stream River Power Plants
The novelty of RISEC technology has to date triggered conservative evaluations and an
extensive environmental review process at the federal, state, and local levels. As of August 2,
2012, there were 81 active FERC-issued preliminary permits and three pending preliminary
permits. Of the issued permits, 75 were for projects on the Mississippi River and 58 were issued
to Free Flow Power Corporation. Furthermore, 43 of the 81 preliminary permits were issued in
2012. Of the pending permits, one is for the St. Lawrence River (New York) and two for the
Kilarc Canal and Creek (California). There were no preliminary permits for inland hydrokinetic
projects issued before 2008. A list of hydrokinetic preliminary permits and licenses is posted on
the FERC website at www.ferc.gov/industries/hydropower/indus-act/hydrokinetics/permits.asp.
11.5.6 Overview of Regulatory Status for River In-Stream Energy Conversion
Projects
The regulatory permitting and licensing processes associated with U.S. river in-stream projects
can be lengthy, costly, and complex. Indeed, regulatory issues represent a significant impediment
to capital acquisition and in-stream river energy development in the United States.
11-33

River in-Stream Energy

Since 1920, constructing and operating a non-federal hydroelectric project in the United States
has required a license issued by FERC in accordance with the Federal Power Act. In 2004, as a
result of legal interpretation from FERC, unconventional tidal and wave energy hydroelectric
projects were placed under FERC jurisdiction. In October, 2007, FERC announced a new
streamlined permitting approach for marine and hydrokinetic energy projects intended to reduce
the duration of the approval process from five years to six months for five-year licenses for pilot
projects under 5 MW. FERC issued a white paper in April 2008 explaining the new pilot project
licensing process, posted online at http://www.ferc.gov/industries/hydropower/indusact/hydrokinetics/pdf/white_paper.pdf.
In April 2011 FERC attempted to revise permitting and licensing for RISEC projects,
specifically projects on the lower Mississippi River, by denying additional preliminary permits to
Free Flow Power Corporation and Northland Power Mississippi River LLC. The reason, FERC
claimed, was that these two companies hold too many permits on the Mississippi River, which
prevents fair competition among other developers. Between these two developers there are 141
issued and pending FERC preliminary permits. Despite objecting to FERCs decision, Free Flow
Power immediately withdrew 58 of 60 applications and chose not to renew five other permits. In
response to the two companies enormous opposition to the decision, FERC reversed its decision
in June 2011. Subsequently, 43 of Free Flow Powers permits were accepted for filing by FERC,
whereas 40 of Northland Powers permits were listed as deficient, requiring additional
information.

11.6 Design, Performance, Cost, and Economic Feasibility


11.6.1 RISEC Sites
There are many factors to consider when evaluating potential sites for an in-stream river energy
plant, including the following:

First and foremost, a high annual current flow, in terms of both volume and speed

A transmission and distribution system that can transport the power from the river plant into
the grid and a substation interconnection point close to shore

Sufficient depth for the device to be deployed off the river bottom and completely
submerged, and which provides navigation clearance if required

Minimum conflicts with other river uses

11.6.2 RISEC Devices Studied


In 2008, EPRI completed a site characterization study of the six sites in Alaska shown in Figure
11-35. After reviewing the data for those six sites, three sites were selected for conceptual
feasibility design studies: Tanana River at Whitestone, Yukon River at Eagle, and Kvichak River
at Igiugig. The Igiugig and Eagle sites are connected to small isolated village grids, whereas the
Whitestone site is located near a 26-kV transmission line.

11-34

River in-Stream Energy

Tanana at Whitestone

Tanana at Manley Hot


Springs

Yukon at Eagle

Taku at Juneau
Yukon at Pilot Station
Kvichak at Igiugig

Figure 11-35
Site location overview, with chosen sites indicated in yellow

For each of the three sites, point designs for both a single-unit demonstration and a commercial
tidal power plant were used to estimate cost and performance. Performance estimates were
developed using river current data predictions obtained from USGS. Figure 11-36 shows the
cross-sectional water depth profiles at the USGS measurement stations of the three sites of
interest.
0

Water Depth (m)

-2

50

100

150

200

250

300

350

400

450

500

-4
-6
-8
-10

Kvichack @ Igiugig

-12

Yukon @ Eagle
Tanana @ Big Delta

-14
Distance from Shore (m)
Figure 11-36
Water depth profiles at three Alaska sites during typical river discharges

Figures 11-37 and 11-38 show the monthly average velocities and monthly average crosssectional power densities at the three sites. Rotor power output is directly related to the
hydrokinetic power density at a river site and significantly affects the project economics. As
noted earlier, the hydrokinetic power density of a free-flowing stream is proportional to the cube
11-35

River in-Stream Energy

of fluid velocity. Although the power densities for all three sites are much higher during summer
than winter, the Iguigig site exhibits much less summer/winter variability than the other two
sites. This is a direct result of the storage provided by Lake Llama upstream of the Iguigig site.
The higher river discharge rates during summer and associated higher velocities are a direct
result of the annual snow melt.
2.50

Yukon River @ Eagle

Monthly Average Velocity (m/s)

Tanana River @ BigDelta


Kvichak River @ Igiugig

2.00

1.50

1.00

0.50

0.00
Jan

Feb

Mar

Apr

May

Jun

Jul

Aug

Sep

Oct

Nov

Dec

Month

Monthly A verage Power D ensity (kW/m ^2)

Figure 11-37
Average water velocities by month at three sites in Alaska
4.50

Yukon River @ Eagle

4.00

Tanana River @ BigDelta

3.50

Kvichak River @ Igiugig

3.00
2.50
2.00
1.50
1.00
0.50
0.00
Jan

Feb

Mar

Apr

May

Jun

Jul

Aug

Month

Figure 11-38
Average power densities by month at three sites in Alaska

11-36

Sep

Oct

Nov

Dec

River in-Stream Energy

11.6.3 RISEC Design, Performance, and Cost


The unit capacity of RISEC devices is limited by the rotor-size constraints imposed by the water
depths in rivers. Scalability imposes significant constraints on the cost-effectiveness of the
technology. However, device performance and cost are well understood for these rotor devices.
Most device manufacturers now target niche-market applications, powering remote villages
where power requirements are small and cost of electricity is high. Device capacity is typically 5
to 20 kW per device, with some manufacturers claiming capacities as high as 100 kW. Because
of the small rated capacity of such devices, cost profiles were established for plant sizes between
500 kW and 5 MW. Even though some manufacturers have proposed larger-scale plants, there
are significant technical, economic and siting issues that remain to be addressed before such
large-scale deployment can proceed.
Cost and performance assessments of six representative RISEC projects are presented in the
following sections. The first set of three assessments refers to the three Alaska sites extensively
studied by EPRI in 2008 and described in Section 11.6.2. The nameplate ratings of these sites
range from 41 kW in the Kvichak River to 593 kW in the Tanana River. The second set of three
assessments was completed in 2011 for hypothetical RISEC plants deployed in the Mississippi
River near Baton Rouge, Louisiana with capacities of 1 MW, 5 MW, and 10 MW [6]. Between
the two sets of data, a wide range of project capacities and characteristics is considered.
11.6.3.1 EPRI Methodology
When possible, the following cost, performance, and economic assessments are based on actual
testing and pilot-plant data to inform and drive techno-economic models. These models assume
that initial early-adopter issues have been resolved and reliability similar to that of comparable
technologies has been obtained. Because real-world experience is limited, typical cost and
performance estimates must also be developed from technical specifications supplied by leading
device manufacturers. It is important to recognize that there is significant diversity in technical
approaches in this emerging field. It is also important to note that system performance is heavily
influenced by site conditions; resource power density tends to be the dominant driver of the cost
of electricity (COE).
Manufacturers typically underestimate cost in the early stages of development. As the
technologies approach commercial maturity, such cost projections increase. The actual
construction and operational costs of a pilot device reveal a more complete economic picture and
provide a solid starting point for further cost studies. Once a technology reaches commercial
maturity, economy of scale applies and volume production begins to drive down costs. This
typical projected cost trajectory (Figure 11-39) can make it difficult to compare early-stage
technologies with more developed options. Based on its extensive experience with emerging
energy technologies, EPRI has developed a cost estimate rating methodology that assesses the
likely range of uncertainty based on a technologys design maturity and the amount of detail
available for the cost estimate. To reduce the uncertainty in this assessment, only data from
devices that are at a fairly advanced stage of development are used. As a result, all the cost data
given in this report have an estimated uncertainty range of 30%.

11-37

River in-Stream Energy

Lab/Idea

Prototype

Commercial

Volume
Production

Cost

Stage of Development
Figure 11-39
Cost projection as a function of development status

For the purpose of this assessment, a number of cost centers were defined that would allow
sufficient detail to provide a solid understanding of the impacts of the major cost drivers. The
following are cost centers that contribute to capital cost estimates in this analysis:

Installation: Includes all activities associated with the plant construction, including:
transport, underwater cable installation, mooring installation, device deployment and
commissioning. The plant is assumed to terminate on shore. No cost is included for the
interconnection of the plant with the electrical grid.

Permitting and environmental compliance: Includes all cost components required to permit a
plant. Uncertainties in this cost category are high because there is limited experience to draw
from. These costs include site baseline assessments, environmental baseline studies, and
consultant fees.

Infrastructure: All costs except for those of the devices and moorings themselves, including
underwater cables, cable landing to shore, dockside improvements, and specialized vessels
used to operate the project.

Mooring: Includes all components required to keep the device on-station. These may involve
mooring chains, anchors, piled foundations, shackles, and other mooring components.

Structural: Includes the main structural components, typically built from structural steel,
concrete or fiberglass.

Power take-off: Includes all components and subsystems required to convert the primary
mechanical energy into electricity that can be fed into the electric grid. Subsystems include
items such as a generator, gearbox, hydraulic system, frequency converter, step-up
transformer, and electrical riser cable.

11-38

River in-Stream Energy

Operating and maintenance (O&M) costs are divided into the following cost centers. To
normalize costs from operational activities, such costs are expressed as present value, divided by
the plants operational life and rated capacity to yield a unit of dollars per kilowatt capacity per
year ($/kW-year). The following are cost centers that contribute to O&M cost estimates in this
analysis:

Insurance: Insurance costs for one-off offshore installations are typically on the order of 2%
of capital cost. As technology matures and technology-related risks are reduced, such
insurance costs are expected to decline to a level similar to that of wind energy today.

Marine monitoring: It is expected that many early-adopter projects will require ongoing
monitoring of the plants environmental effects to satisfy regulatory agencies. This may
includes measures such as active and passive acoustic monitoring, fish studies, and sediment
transport studies.

Operations: This cost center contains all expenses associated with the operation, maintenance
and repair of the plant. This category includes labor cost, fuel cost, management expenses
and facility leases.

Replacement parts: Includes the cost of all parts that require replacement of the life of a
project.

11.6.3.2 Kilowatt-Scale Assessment: Alaska


For each of the three Alaskan sites described in Section 11.6.2, EPRI prepared a point design,
performance, and cost estimate for both a single-unit demonstration and a commercial RISEC
power plant. The performance and cost estimates were used to assess the economic feasibility in
terms of a simple payback period. Table 11-4 presents the cost estimates for the three
commercial plants. These cost estimates are in 2008 dollars and do not include permitting and
engineering costs.

11-39

River in-Stream Energy


Table 11-4
Cost and performance estimates for Alaska in-stream river power plants
Kvichak at Iguigig

Yukon at Eagle

Tanana at Whitestone

30

Annual Electrical Energy


at Busbar (MWh/yr)

207

107

1,325

Nameplate Rating (kW)

41

47

593

December 20092

December 20092

December 20092

Onshore Transmission
and Grid Interconnection

$2,575

$2,246

$178

Subsea Cables

$1,025

$894

$71

Powertrain Mounting

$1,100

$589

$627

Power Conversion
Modules

$1,325

$1,505

$1,168

Structural Sections

$925

$327

$565

Assembly

$225

$131

$150

Transport to Site

$150

$65

$105

Onsite Assembly

$150

$44

$105

Device Deployment

$38

$11

$26

Contingencies

$350

$196

$232

Total Plant Cost (TPC)

$7,863

$6,008

$3,229

AFUDC (interest during


construction)

$786

$601

$326

Total Plant Investment

$8,649

$6,609

$3,555

Due Diligence,
Permitting, Fees and
Taxes

$393

$300

$161

Total Capital
Requirement

$9,042

$6,909

$3,717

Yearly O&M Costs


$/kW-yr

$302

$140

$224

90

38

90

+30% Capital Cost

+30% Capital Cost

+30% Capital Cost

30 to +80% O&M
Costs

30 to +80% O&M
Costs

30 to +80% O&M
Costs

Number of Units

Capital Cost $/kW


Month/Year Dollars

Unit Availability (%)


Cost Estimate Accuracy
Rating

11-40

River in-Stream Energy

11.6.3.3 Megawatt-Scale Assessment: Mississippi River, Louisiana


This analysis will consider 1-MW, 5-MW, and 10-MW RISEC plants in the Mississippi River
near Baton Rouge, Louisiana with the following riverine parameters:

Mean bulk velocity: 1.07 m/s

Mean surface current speed: 1.25 m/s

Vertical power law profile: 1/6.

Table 11-5 summarizes the results of the assessment, with costs for each size plant broken down
by capital expenses (CAPEX in $/kW), O&M expenses (OPEX in $/kW-year), plant
performance, and the resulting cost of electricity (cents/kWh). CAPEX includes the costs of
permitting and environment, installation, infrastructure, mooring, structural components, and
power take-off equipment. OPEX includes the costs of insurance, marine monitoring,
replacement parts, and operations.
Table 11-5
RISEC energy cost, performance and economic profiles (2011$)
Plant Capacity
1 MW

5 MW

10 MW

Power Take-off

$1790

$1099

$898

Structural

$2043

$1746

$1632

Mooring

$204

$181

$172

Infrastructure

$493

$306

$250

Installation

$864

$294

$188

Permitting & Environment

$1556

$347

$174

$6950

$3974

$3313

Operations

$171

$100

$79

Replacement Parts

$26

$19

$17

Marine Monitoring

$174

$35

$17

Insurance

$100

$42

$16

$470

$196

$129

6.8%

4.9%

3.9%

Capacity Factor

30%

30%

30%

Availability

95%

95%

95%

CAPEX Contribution

31

18

15

O&M Contribution

26

11

57

29

22

CAPEX ($/kW)

TOTAL
O&M ($/kW-year)

TOTAL
O&M Percent of CAPEX
Performance

Cost of Electricity (cents/kWh)

TOTAL

11-41

River in-Stream Energy

Figure 11-40 graphically summarizes the CAPEX results for the three sizes of RISEC plants,
whereas Figure 11-41 does the same for O&M.
$8,000

CAPEX($/kW)

$7,000
$6,000

Permitting&Environment

$5,000

Installation

$4,000

Infrastructure
Mooring

$3,000

Structural

$2,000

PowerTakeOff

$1,000
$0
1MW

5MW

10MW

Figure 11-40
RISEC capital expense as a function of plant scale

Figure 11-41
RISEC O&M expense as a function of plant scale

11-42

River in-Stream Energy

River current velocity and corresponding power density are highly localized and can vary
significantly over a few hundred meters. To illustrate the sensitivity of a plants levelized cost of
electricity (LCOE) to the power density at the site, the velocity distribution during the
performance assessment was simply scaled to evaluate the technologies sensitivity to the
resource strength (Figure 11-42). This assessment does not consider regional differences in river
discharge and other site-specific resource differences, but can serve as an indicator to evaluate
the sensitivity of LCOE to resource strength.
35

CoE(cents/kWH)

30
25
20
15
10
5
0
0

MeanPowerFlux(kW/m2)

Figure 11-42
Cost of electricity as a function of mean power flux

11.6.4 RISEC Economic Feasibility


Table 11-6 summarizes the economic analyses for the three potential commercial-scale river
power plants in Alaska. The cost estimates are in 2008 constant dollars and do not include
permitting and engineering costs. Iguigig and Eagle are remote villages and their RISEC plants
were sized to meet the summer low daily load (40 kW for Iguigig and 70 kW for Eagle).

11-43

River in-Stream Energy


Table 11-6
Cost estimates for three feasibility evaluation sites
Site Parameters

Iguigig

Eagle

Whitestone

No

Yes

No

Annual Average Power Density

1.48 kW/m2

1.5 kW/m2

0.67 kW/m2

Mid-channel Average Power density

3.24 kW/m2

3.2 kW/m2

1.48 kW/m2

719 kW

4,601 kW

762 kW

1:4

1:20

1:10

60 m

150 m

50 m

40 kW

70 kW

> 5 MW

# of RISEC Devices

30

# Rotors per Machine

Rotor Diameter

1.5 m

2m

2m

Rated Capacity

42 kW

61 kW

593 kW

220 MWh/year

113 MWh/year

1356 MWh/year

Capacity Factor

65%

57%

29%

Availability

90%

38%

90%

$269,000

$308,000

$1,821,000

Installed Cost per kW

$5,800/kW

$7,500/kW

$3,100/kW

Avoided Cost (selling price)

0.65 $/kWh

0.65 $/kWh

0.18 $/kWh

3 Years

4 Years

8 Years

Ice Freeze-over?

Average Total Kinetic Power


Summer/Winter Power Density Variability
Site Distance from Shore
Grid Feed-In Limit
RISEC Plant Parameters

Plant Annual Generation

Cost and Economic Parameters


Installed Cost

Simple Payback Period


Notes:
1.

The value of electricity revenues is the avoided cost. For a rural Alaskan utility running on diesel, the avoided
cost is essentially the fuel cost. With fuel costs of $8/gallon delivered and efficiencies of 13 kWh/gallon, the
avoided cost is typically $0.65/kWh. The O&M cost of a diesel genset is $0.02-$0.05/kWh, but we
conservatively assume that a genset idling in the background has no O&M savings. For the grid-connected
Whitestone case, a value of $0.18/kWh was used.

2.

The simple payback period (SPP) is the year when the yearly revenues (calculated as the product of the annual
electricity produced multiplied by the avoided electricity cost) equals the sum of the capital cost plus O&M
cost. Escalation of non-fuel cost was assumed to be 3% per year and escalation of fuel costs was assumed to be
8% per year.

11-44

River in-Stream Energy

11.7 Environmental Issues


Table 11-7 summarizes the key environmental issues, possible impacts, and potential mitigation
approaches surrounding RISEC project implementation.
Table 11-7
Major environmental issues and mitigation recommendations
River Issue

Impact(s)

Mitigation

Withdrawal of
River Energy

Retardation of river discharge rates.

Limit the withdrawal of river energy


to a level that does not result in any
noticeable ecological effects.

Interactions with
Aquatic Life,
Seabirds and
Benthic
Ecosystems

Fish mortality seems to be the most


significant issue. Impacts on the benthic
ecosystems occur primarily during
installation; impacts on sediment
transport processes are also of concern.

Limit the rotation speed of blades to


a level which precludes fish
mortality.

Emissions

Applies to devices with closed-circuit


hydraulic systems where working fluid
(which may be biodegradable) may leak
or spill during transfers, and use of fluids
for installation (i.e., drilling sockets in
hard rock seabed).

Use only non-toxic fluids.

Visual
Appearance and
Noise

The aesthetic effect of visually impacting


a pristine river scene may be
unacceptable in some situations.

Fully submerge the energy


conversion devices, land the cable
under the shore.

Conflicts with
Other Uses of
River Space

Potential conflicts with recreational uses,


commercial shipping, commercial fishing,
dredging, and other activities.

Hold siting, design and installation,


operation, and procedure
discussions with all local
stakeholders prior to making final
plant detail design decisions.

Debris and Ice


(including ice
breakup)

Potential damage to turbine rotor and the


entire system.

Install trash racks to reduce debris


impacts and remove devices during
spring ice breakup to avoid damage.

The U.S. Department of Energy and the Pacific Northwest National Laboratory developed
TETHYS (after the mythical Greek titaness of the seas), which is a database and knowledge
management system that provides access to information and research pertaining to the potential
environmental effects of marine and hydrokinetic (MHK) and offshore wind development.
Tethys also hosts data from Annex IV, an international collaboration to gather information on
MHK environmental research worldwide. As of August 2012, there were 16 research reports in
the TETHYS database pertaining to RISEC devices.

11-45

River in-Stream Energy

11.8 Installed Capacity and Estimated Growth


As of November 2011, there are three installed RISEC devices in the United States: a 100-kW
Hydro Green turbine at Hastings, Minnesota; a 25-kW EnCurrent system on the Yukon River at
Eagle, Alaska; and a 5-kW EnCurrent system on the Yukon River at Ruby, Alaska. Table 11-8
presents the EPRI estimate of river in-stream capacity (in megawatts) that is expected to come on
line in the United Statesmeaning commissioned and delivering electricity to the gridduring
2011 through 2013.
Table 11-8
Installed and planned U.S. river in-stream power capacity (MW)
Developer

Project Name/Site

2011

2012

2013

ABS Alaskan

Yukon River at Ruby, Alaska

0.005

Hydro Green

Hastings, Minnesota

0.100

Alaska Power &


Telephone

Yukon River at Eagle, Alaska

0.025

Free Flow Power


Projects 1

Many on the Mississippi River

100s of
MW

100s of
MW

Other Hydro Green


Projects 1

Many in Alaska, Mississippi River

100s of
MW

100s of
MW

Other 2008-Issued
FERC Prel. Permits

Various

100s of
MW

100s of
MW

TOTAL NEW

YEARLY CAPACITY

0.130

0.2

CUMULATIVE
Note:
1.

If Free Flow Power and Hydro Green complete their license applications within the three-year time frame of the
FERC preliminary permit, they should have their construction and operation licenses in 2011 and construct and
begin operation in 2012 or shortly thereafter.

In general, EPRI expects that river in-stream energy will experience a growth rate that is limited
by regulatory barriers and government support for the emerging river energy industry. EPRI
estimates that there could be 500 MW of RISEC plant capacity by 2025. This assumes that
FERC and other regulatory agencies change their practices to enable river in-stream plants to be
permitted at about the same cost as onshore wind plants, and that Congress enacts the same
financial incentives for river in-stream energy as are available for wind energy. There is a
potential for river in-stream capacity to significantly increase if Free Flow Power and Hydro
Green build their plants under the existing preliminary permits issued by FERC. Of course, it is
unknown at this time whether and how many of the RISEC projects in the FERC queue will
actually be built.

11-46

River in-Stream Energy

11.9 Research Focus


Currently DOE National Laboratories are focusing on six critical research areas:
1. Physical Interactions with Devices
a. Fish/marine mammal attraction and avoidance
b. Strike risk to fish/marine mammals
2. Electromagnetic Fields
a. Effects of EMF on fish and marine mammals
3. Acoustics
a. Effects of MHK noise in riverine environments
b. Noise measurement and net-pen studies of select species
4. Toxicity
a. Effects of antifouling coatings on aquatic organisms
5. Benthic Habitat Alteration
a. Development of measurement methodology to evaluate effects of MHK devices
on benthic habitats
6. Data Aggregation and Risk Modeling
a. Development of a publicly available information database for MHK
environmental research
These national laboratories include the following:

Argonne National Laboratory

Idaho National Laboratory

National Renewable Energy Laboratory

Oak Ridge National Laboratory

Pacific Northwest National Laboratory

Sandia National Laboratories

In addition to the research being conducted by the DOE National Laboratories, the department
has also funded the development of National Marine Renewable Energy Centers. As of 2012, the
department has helped establish three National Marine Renewable Energy Centers: Northwest
National Marine Renewable Energy Center (NNMREC), Hawaii National Marine Renewable
Energy Center (HINMREC), and Southeast National Marine Renewable Energy Center
(SNMREC).

11-47

River in-Stream Energy

The Northwest National Marine Renewable Energy Center is a DOE-funded partnership between
Oregon State University and the University of Washington. Oregon State University is
responsible for wave energy R&D, while the University of Washington is responsible for tidal
energy R&D. The Hawaii National Marine Renewable Energy Center is a partnership between
the DOE and the University of Hawaii, and focuses on wave energy R&D and Ocean Thermal
Energy Conversion (OTEC) R&D. The Southeast National Marine Renewable Energy Center is
a partnership between the DOE and Florida Atlantic University, and focuses on OTEC R&D and
ocean current energy R&D.

11.10 Conclusions
Considerable potential exists for generating electric power from river energy in the United States
and many other places in the world. Most of the high-power-density U.S. river hydrokinetic
energy potential exists in the lower Mississippi River, the Pacific Northwest, and particularly in
the state of Alaska. Many preliminary permits have been issued for the Mississippi, Missouri,
and Ohio rivers, due in part to the proximity to electrical load.

11.11 Internet Resources


EPRI River Power (RP) Reports are available from the River page at
http://oceanenergy.epri.com/.

11.12 References
1. Miller, G., J. Franceschi, W. Lese and J. Rico. 1986. The Allocation of Kinetic Hydro Energy
Conversion SYSTEMS(KHECS) in USA Drainage Basins: Regional Resource and Potential
Power. NYUDAS 86-151.
2. River In-Stream Energy Conversion (RISEC) Characterization of Alaska Sites. EPRI, Palo
Alto, CA: 2008. EPRI-RP-003-Alaska.
3. System Level Design, Performance, Cost and Economic AssessmentAlaska River In-Stream
Power Plants. EPRI, Palo Alto, CA: 2008. EPRI-RP-006-Alaska.
4. Fish Passage Through Turbines: Application of Conventional Hydropower Data to
Hydrokinetic Technologies. EPRI, Palo Alto, CA: 2011. 1024638.
5. Evaluation of Fish Injury and Mortality Associated with Hydrokinetic Turbines. EPRI, Palo
Alto, CA: 2011. 1024569.
6. Previsic, M., Cost Profiles of Marine Hydrokinetic (MHK) Technologies, September 2011.
7. Prioritized Research, Development, Deployment and Demonstration (RDD&D) Needs:
Marine and Other Hydrokinetic Renewable Energy. EPRI, Palo Alto, CA: 2008.

11-48

12

GRID INTEGRATION CHALLENGES AND


TECHNOLOGIES
The integration of variable renewable generation, particularly variable wind and solar PV
generation, affects the power delivery system at both the transmission and distribution levels. At
the transmission level, variability and uncertainty of generation is causing system operators to
redefine ancillary requirements, such as reserve requirements and ramp rates, in order to
maintain grid reliability. Other issues, such as frequency response of the system, conventional
generator cycling and increasing transmission requirements, are also being driven, at least in
part, by increased shares of variable renewable generation. Operational changes may affect many
tools and processes currently in use, and will increase in importance for operators working to
meet reliability standards. Systems integrating large amounts of variable generation will require
new methods to plan the system, both in terms of the transmission infrastructure and the
generation portfolio.
At the distribution level, variability of renewable generation will also require changes. The issues
are different but equally important for enabling penetration of distributed generation into existing
and future systems. These include evaluating interface devices, analytics, studies, applications
with end-use resources, and assessment of new technology for effective interconnection and
integration of renewable and other distributed generation. Recognizing the need for change and
adapting the electric grid can enable higher penetration levels without reducing safety, reliability,
or asset utilization effectiveness.
This chapter first outlines in brief how the grid is operated and planned. This is followed by a
description of the key characteristics of variable renewables, and the key challenges for
integration of these sources based on experience to date and studies. Finally, it addresses
technologies and strategies that will ease the integration of renewable energy into the electricity
grid. Energy storage, power electronics, resource forecasting, grid planning and operating tools,
increased flexibility from conventional plant, demand response, and increased transmission
capability can all be used to mitigate the impacts of adding large amounts or wind, solar, and
other variable generation to the electricity grid.

12.1 Basics of Power System Operation and Planning


This section outlines the planning operation of the electrical grid, particularly highlighting areas,
to be discussed later, that will be most affected by variable generation integration.
12.1.1 Functions of System Generation
The main function of generation in an electric power system is to produce electric energy to
serve system demand. This function may include regeneration from stored energy. The
Authoritative Dictionary of IEEE Standards Terms, ANSI/IEEE Std. 100, defines generation as
12-1

Grid Integration Challenges and Technologies

producing or storing electric energy with the intent of enabling practical use or commercial sale
of the available energy. A generating station is also defined as a plant wherein electric energy is
produced from some other form of energy by means of suitable apparatus (for example,
chemical, mechanical or hydraulic).
The operation of electric power systems is fundamentally different from services provided by
other types of utilities. Electric systems have two unique physical operating constraints:

Electric energy is not commercially stored86 like natural gas and water. Production and
consumption (generation and load) must be balanced in near real-time. This requires
continuous monitoring of load, generation, and the voltages and energy flows throughout the
power system, as well as adjusting generation output to match consumption.

The transmission and distribution network is primarily passive, with few control valves or
booster pumps to regulate electrical energy flows on individual lines. Flow-control actions
are limited primarily to adjusting generation output and to opening and closing switches to
add, remove, or reroute transmission and distribution lines and equipment from service.

These two operating constraints lead to numerous reliability consequences with practical
implications that dominate power system design and operation, including the following:

Every action can potentially affect all other activities on the power system. Therefore, the
operations of all bulk-power participants must be coordinated.

Cascading problems that quickly escalate in severity are a real threat. Failure of a single
element can, if not managed properly, cause the subsequent rapid failure of many additional
elements, potentially disrupting the entire power system.

Contingency readiness may limit current operations. For example, likely power flows that
occur if another element fails could limit the allowable power transfers.

Maintaining system stability and reliability often requires that actions be taken
instantaneously (within fractions of a second), requiring automatic computations,
communications, and controls.

These consequences bear directly on how wind and solar power interacts with the power system.
12.1.2 Expectations Based on Traditional Generator Performance
Reliable and low-cost operation of the interconnected electric utility is not a natural state but
relies on operational scheduling, planning, and coordination of a multitude of electric utility
control systems. A traditional electric utility generator is not just a source of energy. Operators
rely on an array of performance capabilities and specific services from traditional generators.
Some of these services are simply provided by generator operators as part of their participation
in an energy market. Others are obtained from bidding into separate ancillary services markets
that have evolved with deregulation of generation services. The exact definition of these services
varies in different markets but, in general, they fall into the categories described in Table 12-1.
86

Electricity is not stored directly. Rather, it is converted to another form of energy and re-converted later.
Pumped storage hydro converts electricity to mechanical potential energy by lifting water. Batteries convert electric
energy to chemical potential energy. Their conversion to electricity uses conventional generators or inverters.

12-2

Grid Integration Challenges and Technologies


Table 12-1
Functions and services provided by generation
Functions and Services
Baseload Units
(non-regulating)

Description
Energy (firm) scheduled well in advance, based on
availability, price, and long-term contracts.

Time Frame
Long-term
commitments

Committed Units (usually Energy (firm) scheduled based on availability and


with regulation capacity) price to meet block load, with LOLE1 and load
forecasts considered.

Day before plan,


hourly resolution

Load-Following or
Energy-Balancing Units

Energy ramping to follow the load, met by adjusting


generation schedules and the imbalance energy
market.

Hourly plan with 5- to


10-minute resolution

Frequency Regulation
(regulating reserves)

Every few minutes,


Service provides capacity based on a signal from
dispatcher, with AGC2 to meet CPS1 and CPS23 and minute-to-minute
resolution with a
no net energy.4
faster signal sent out
by ISOs/Balancing
Authorities

Reactive Supply and


Voltage Control

Service of injecting or absorbing of reactive power


to control local transmission voltages (usually
provided with energy).

Continuous with
response in seconds

Spinning Operating
Reserves

Service to provide energy in response to


contingencies and frequency deviations.

Begin within 10 sec


full power in 10 min

Non-Spinning Operating
Reserves

Service to provide load/generation balance in


response to contingencies, not frequency response.

Respond within 10
minutes

Replacement Reserves

Service to restore contingency capacity to prepare Respond within 60


for the next generation or transmission contingency. minutes, run up to 2
hours

System Black Start

Service to restore all or a major portion of the power As required


system without outside energy after a total collapse.

Notes:
1.

Loss of load expectation (LOLE) is the probably measure that a load cannot be served with available generation.

2.

Automatic generation control (AGC) is a method for adjusting generation to minimize frequency deviations and
regulate tie-line flows.

3.

Control performance standards (CPS1 and CPS2) are minute-to-minute and 10-minute average criteria for load
frequency control in each control area. These criteria require control areas to maintain their area control errors (ACE)
within tight limits. ACE is measured in MW and is defined as the instantaneous difference between the actual and
scheduled interchanges plus frequency bias (imbalances that bias the system toward maintaining 60 Hz).

4.

Frequency regulation service is usually provided by generation that is on-line and delivering some level of base energy
power on a full-time basis or scheduled. The service increases and decreases power output so that the average output
over the scheduled period does not changethat is, there is no net change in delivered energy attributed to the
frequency regulation service. Consequently, energy-storage devices could provide this service. in some ISOs,
particularly California, this service may be split into two, with separate up and down regulation.

12-3

Grid Integration Challenges and Technologies

Response time is one of the critical factors in defining generation capabilities and services. The
reaction time is critical because of the nature of the electric utility system, which operates with
very little electrical storage, as described earlier. Furthermore, the degree to which variable wind
and solar power can be absorbed without significant integration issues, as well as the value
attributed to wind power, is determined by how well the traditional generation functions are
fulfilled, as well as the standard to which they need to be fulfilled.
12.1.3 Typical Power System Operations and Planning
This section outlines power system operations and planning, so that the reader can have a better
understanding of how variable renewable generation may have an impact on power system
operations and planning. From a planning perspective, the main aim is to ensure there are
sufficient resources available in periods from months to decades ahead to meet demand at the
desired level of reliability, while ensuring that cost, environmental, and other regulatory
constraints are met. With increased levels of variable generation, their particular characteristics,
described in the next section, will need to be accommodated. Reliability metrics and economic
planning are important aspects of long-term power system planning.
With or without variable renewable generation, the system operating strategy will likely include
energy balancing from both committed and reserve units, unit commitment block scheduling, and
an economic dispatch procedure. This strategy starts with a forecast of system load with timevarying characteristics and expected minimums and maximums. Inventories of available
generation and the estimated cost of energy as well as other services are maintained, or bid into a
market by market participants if appropriate. Desired reserve margins are set. Base loaded units
and firm energy blocks can be scheduled so that, in the short term, the operator is only dealing
with small differences between the predicted and actual load and/or available generation. Thus,
the system operator is effectively managing supply and demand over three time scales: (1) daysand hours-ahead scheduling, (2) intra-hour load following and balancing, and (3) fast regulation
to maintain system frequency and voltage, as shown in Figure 12-1. The impact that variable
renewable generation may have on this strategy is discussed later. All three time frames are also
important to effectively integrating any generation resource, including wind. This section first
describes day-to-day system operation, followed by a short description of some of the key
planning issues.

12-4

Grid Integration Challenges and Technologies

Figure 12-1
System operation over minutes/hours (top) to days (bottom)

12.1.3.1 Resource Adequacy and Transmission Planning


Resource adequacy is the process by which an electricity system control area (ISO/RTO, VIU,
etc.) projects future load growth and generator decommissioning to ensure that sufficient
generation exists to meet expected peak instantaneous demand on the system. Increasingly,
demand response resources and energy efficiency initiatives are entering to share the role of
traditional generating resources. Periodically, load growth scenarios are simulated, and a
resource adequacy is examined, in which each balancing authority (BA) is required to ensure it
has sufficient capacity to meet future peak load. In addition, transmission constraints may require
that certain peak generation capacity is located in specific areas to serve load pockets, or to
mitigate the worst potential failure on the bulk grid system.
Resource adequacy is ensured through metrics such as loss of load expectation (LOLE) or
expected unserved energy (EUE). These metrics are based on probabilistic methods that examine
generator characteristics and project the sufficiency of resource on the system to meet projected
demandfor example, a system may have an LOLE of 1 day in 10 years. If there is insufficient
generation to meet this target, then additional generation needs to be procured. Resource
adequacy provision may be accomplished through a market auction process and is utilized to
12-5

Grid Integration Challenges and Technologies

guarantee a certain available supply of generation for a period in the future. This could also have
the side effect of reducing the volatility of energy prices in the day-ahead and real-time markets.
Resource adequacy generation resources are typically required to bid into the day-ahead energy
market under a bid cap. In non-market regions, the resources may receive a payment based on a
long-term agreement. Resources will each have a capacity credit, which is the amount of firm
capacity they can contribute to resource adequacy. Generation with low capacity credits tends to
have low output at peak, which is often the case with variable renewable generation.
In addition to ensuring that there is sufficient generation years ahead of time, the transmission
system must also be designed to ensure that generation can be delivered to load in a reliable
fashion. Therefore, transmission planning aims to plan for build-outs of new transmission
resources, while ensuring that reliability is maintained. In particular, the system must be able to
operate in a secure fashion, which means that it must be able to handle the loss of any
transmission line. Transmission planning often aims to balance reliability with economicsthat
is, the system should be built to ensure it is robust to contingencies (it can maintain stability with
the loss of any critical element), but overbuilding the system is expensive and inefficient. As
described later, the dispersed nature and low capacity factor of variable renewable generation
may increase transmission planning challenges.
12.1.3.2 Unit Commitment
Day-to-day operation starts with the unit commitment (UC) process. UC is part of a set of
programs for scheduling generation on an hourly to weekly basis. The primary goal is
determining which generation should be online at various parts of the day, while considering
generator characteristics as in Table 12-3. Vertically integrated utilities typically run their UC
optimization computer programs the day before the operation is needed. These programs accept
as inputs detailed information on the characteristics of the individual generating units that are
available to produce electricity on the following day, including unit status, minimum and
maximum output levels, ramp-rate limits, startup and shutdown costs, durations and lead times,
minimum runtimes, and unit fuel costs at various output levels. They then schedule generation
based on the objective of lowering overall system costs while meeting forecasted energy and
reserve requirements. Unit commitment is run for both markets and non-market systems.
Generally, markets will schedule a day ahead based on generator and load bids received, with a
reliability unit commitment run by the system operator to ensure security on the system once the
market has cleared. In many markets, unit commitment processes will continue to be run at
various intervals prior to the time of delivery to ensure sufficient energy and ancillary services to
meet demand at the desired reliability level. For example, the California Independent System
Operator (CAISO) runs day-ahead, hour-ahead, and real-time markets.
12.1.3.3 Economic Dispatch
Once generators are committed (turned on and synchronized to the grid), they are available to
deliver power to meet customer loads and reliability requirements. Utilities will typically run
their least-cost dispatch model every five minutes or so. This model forecasts load for the next
five-minute interval and decides how much more or less generation is needed during the next
interval for regulation to meet the system load. The model then selects the least-cost combination
of units that meet the need during the next intra-hour interval.
12-6

Grid Integration Challenges and Technologies

Load following resources are dispatched to follow within-the-hour load changes in the load (or
net load when including variable generation) consistent with the economic dispatch cycle (5 to
10 minutes per cycle). These are therefore used to follow variability in the load over periods of 5
to 20 minutes; sufficient load-following is procured in the economic dispatch stage to meet
variations in load. Load-following requirements tend to be somewhat correlated; most area
profiles rise in the morning and drop off in the evening. Still, because load-following
requirements are not perfectly correlated, the total system load-following requirement is less than
the sum of the load-following requirements of individual end users. This aggregation of loads
has a powerful effect on system planning and operation because the system must respond to total
variations, not the sum of individual variations
12.1.3.4 Role of Ancillary Services
In addition to committing and dispatching units, system operators contract for ancillary services
to support operation of the grid. FERC has defined ancillary services as those necessary to
support the transmission of electric power from seller to purchaser, given the obligations of
control areas and transmitting utilities within those control areas to maintain reliable operations
of the interconnected transmission system. This statement recognizes the importance of ancillary
services for bulk-power reliability and to support commercial transactions. Although the
resources necessary to create these services are generators, the services themselves must be
deployed and controlled by the same system operator that controls the transmission system.
Ancillary services are conceptually well defined. They have existed throughout the history of the
power system. However, the details of obtaining them from markets are still evolving. Ancillary
services normally provided by generation include regulation, contingency reserves, voltage
control, and black start capability. In many market regions, the resources to provide ancillary
services are determined based on a co-optimization with energy resources in the day-ahead
market. All resources bid in both energy and regulating capability and costs, and the market
optimizer (or vertically integrated utilities UC tool) determines the best mix of resources to
provide energy and ancillary services to meet scheduled demand at lowest cost. Energyboth in
day-ahead and from the perspective of interval-to-interval economic dispatch and load
followingwas already described. A brief description of different reserve types follows. Note
that these will vary from region to region and are constantly evolving, but the following serves as
an approximate guide.
Regulating Reserves/Regulation
System operators run regulation markets to ensure that there is adequate on-line generation
capacity for ramping up or down to follow the load and regulate the faster and more random
changes in load, especially in time periods shorter than the dispatch interval (i.e., within 5
minutes in many markets). Regulating units must be on-line and provide fast response
(regulation reserve, or simply regulation) to meet minute-by-minute fluctuation in the system
energy balance. The North American Electric Reliability Council (NERC) has set guidelines on
response-time performance for plants providing frequency regulation.
Both regulation and load following are dealt with on a balancing or control area basis through
NERC standards. It is not necessary to compensate for each and every load variation directly, but
the aggregate change in the area must be balanced; this includes any deviations from scheduled
flows on tie-line, thus the balancing authority or independent system operator (ISO) must ensure
12-7

Grid Integration Challenges and Technologies

that the net balance (of generation and supply on the system and tie line flows versus scheduled)
are maintained. The fast, random fluctuations associated with regulation are typically
uncorrelated. Consequently, the total regulation requirement is not the sum of all the regulation
requirements of the individual loads and uncontrolled generators, but is instead the sum of the
correlated components. Various plants in the generation mix will bid in or be scheduled to
provide balancing (energy regulation) to the system. It should be noted that controlling loads is
another way to balance supply and demand. Load control or load as a resource is being
considered by most ISOs. Demand response can be relatively fast and may be the least-cost
option. With the advent of the smart grid and associated technologies, controllable or responsive
loads may well increase significantly in the next decade, and such resources may prove very
useful in providing load following and regulating reserves.
Recently, FERC issued an order that would pay regulation revenues based on performance rather
than the capacity alone. This gives resources a chance to correct the higher frequency ACE
signals instead of respond to AGC and possibly receive higher payments for fast and accurate
response to these signals. FERC 755 requires regional transmission officers (RTOs) to base the
regulation market clearing price on the total cost to supply regulating reserve, including the
opportunity cost of the marginal resource. This may make the supply of regulation more
profitable for certain resources, while improving system-wide provision of regulation. The
largest impact will likely be for storage and demand response, which have high capital costs but
low opportunity costs to provide regulation; it may also be important for wind generation.
Table 12-2 shows the typical controllability characteristics of different types of generation that
may be in a fleet. Note that variable generation is shown here; whereas for the other technologies
what is shown is the controllability, however, for these resources, what is shown is how likely
the technology is to change in that period of time.

12-8

Grid Integration Challenges and Technologies


Table 12-2
Comparison of output controllability for various generation technologies
(EPRI, T. Key, November 2012)
Controlled
Ramp Rate
(Up or Down)

UnControlled
Variability

Minimum
Output % of
Rated**

Typical
Unit Size
[MW]

Minutes

100% in 30 sec
to 2 min

Seasonal

10%

1200

Hours-days

Once/day

1 to 2% per
minute

Rarely

40 to 50%

10750

SteamCoal
Critical Pressure

Hours-days

Once/day

0.2 to 2% per
minute

Rarely

40 to 50%

10750

Combined Cycle
Natural Gas

Hours

Twice/day

2 to 3% per min

Rarely

45 to 55%

40400

Combustion
Turbine Simple
Cycle

Minutes

More than
twice/day

3 to 5% per
minute

Rarely

55 to 65%

20250

Diesel Internal
Combustion

Minutes

More than
twice/day

100% in 15 to
30 seconds

Rarely

10%

0.120

Hours-days

Days

0.2 to 2% per
minute

Rarely

90 to 100%

7501500

Wind*

Seconds*

Minutes*

Seconds*

10%/min

2 to 5%

0.14.5

Solar PV*

Seconds*

Minutes*

Seconds*

10%/sec

2 to 5%

<0.11

Solar Thermal
6-hr Storage*

6090
minutes

More than
twice/day

~2% per min


(faster down)

Daily sitedependent

10%

50150

Compressed Air
Energy Storage

Minutes

More than
twice/day

30 to 40% per
minute

Rarely

25 to 30%

20250

Seconds to
minutes

Minutes

Seconds

Rarely

2 to 5%

<0.11

Startup
Time

Minimum
Cycling Time

Conventional
Hydro*

Minutes

SteamCoal

Generation Source

SteamNuclear

Battery Energy
Storage

* Controlled output changes depend on available resource (water, wind, sun), whereas uncontrolled output
variability, such as output ramp rates, vareis by technology and weather. Wind and solar PV are only partially
controllable and as such should be treated differently than other resources described here. In addition, it should be
noted that most wind, and much of solar PV, will be in the form of larger plants, from 10 to 100s of MWs in siz
** Minimum output is dictated by economics as well as technical limits, and market incentives could act to lower
minimum outputs.

12-9

Grid Integration Challenges and Technologies

Contingency generation reserves are needed in case of sudden loss of generation, off-system
purchases, unexpected load fluctuations, and/or unexpected transmission line outages. These
reserves may be spinning or non-spinning reserves, or simply serve as replacements. Reserve
generation units may be categorized as follows:

Contingency reserves. The power system must always be prepared to survive the
unexpected loss of a generator or a transmission line. To accomplish this, the power system
operator is required to maintain contingency reserves consisting of spinning reserve and nonspinning reserve sufficient to meet the largest credible contingency.

Spinning operating reserve. At least half the contingency reserve is typically spinning
reserve. Spinning (sometimes called synchronous) reserve comes from generation that is online, not fully loaded, capable of responding fully within 10 minutes, and able to maintain
that output for at least two hours. Spinning reserve units are also required to be responsive to
frequency deviations. In some regions, demand response is now allowed to provide spinning
reservefor instance, it can contribute to PJMs synchronous reserve market.

Non-spinning reserve. The definition of non-spinning reserve is similar to that of spinning


reserve except that the reserve is not required to be on-line or frequency responsive. In
addition, non-spinning reserve does not have to be provided by generation; instead, it can be
provided by dispatchable demand, interruptible exports, certified off-line generation, or
external imports.

Replacement reserve. Contingency reserves must be restored so that the system is prepared
for a subsequent unexpected outage. Some regions specifically identify replacement reserves,
which are capable of responding within one hour and sustaining that response for an
additional two hours. They can be generators, loads, or resources from outside the ISOs
control area.

Together, spinning reserve, non-spinning reserve, and replacement reserve provide resources that
begin responding immediately to an unexpected event, are fully deployed within 10 minutes, are
capable of responding to a second event within one hour, and can sustain the total response for
three hours. This coordinated set of resources is designed to provide sufficient time for markets
to begin functioning again and return the system to normal operations. All of these services and
regulation are typically procured through day-ahead and hour-ahead markets run by the ISO or
balancing authority.
Figure 12-2 provides a summary of deployment times for various ancillary services. Reserves are
deployed only during contingency operations, whereas regulation and voltage control are
required during both normal and contingency operations.

12-10

Grid Integration Challenges and Technologies

Replacement Reserve

C ontinge ncy Oper atio ns


Non-Spinning Operating Reserve

Spinning Operating Reserve

N orma l & Co ntin gen cy Op era tions


Regulation

Voltage Control

0.1

98041

10

100

1000

TI ME (MINUTE S)

Figure 12-2
Ancillary services distinguished by their deployment times and durations

12.1.3.5 Voltage Control


Voltage control refers to maintaining voltages on the power system within pre-defined limits.
Reactive power is a critical component of voltage control; as it is a local phenomenon, ensuring
sufficient sources of reactive power in different regions of the system is important. Therefore, it
is not normally procured through markets. Instead, all generators are required to provide limited
reactive support and voltage control without monetary compensation. Reactive support provided
by generation or absorption of reactive power (MVAR) is used to control voltage in the vicinity
of the generator. All generators are required to be capable of following a system-operatorsupplied voltage schedule under automatic voltage control within a power factor range of 0.90
lagging and 0.95 leading. This is usually required in grid code or interconnection requirements,
for both conventional and, lately, variable renewable generation, over a certain MW size.
Generators need to be able to provide reactive support during both normal operation, and also
during faults. The system operator will typically compensate generators if they are required to
provide additional reactive support and voltage control.
12.1.3.6 Frequency Response
The final aspect of system operation relevant to renewable integration is frequency response.
This refers to the behavior of the system after a fault on the network. After a fault, frequency
starts to decay quickly. This is initially arrested by the spinning inertia on the system. Then,
governor response (primary frequency response) on conventional generators starts to bring the
frequency back to its normal operating point. The system must have sufficient inertia and
primary frequency response to ensure that load shedding does not occur when frequency drops
too low after a fault. Once primary frequency response has acted, secondary frequency response
(or spinning reserve) acts to bring frequency back to its nominal value over the next few minutes.
In recent years, frequency response has become more significant an issue, with FERC, among
12-11

Grid Integration Challenges and Technologies

other entities, identifying it as an area that requires further study and possibly new standards or
regulations. Changing plant mix has been identified as a key reason for this, with coal being
replaced by gas (which may provide less inertia) and, more recently wind, and solar PV. The
nature of variable renewables means they do not provide the natural inertia and governor
response of conventional plant, and thus may affect frequency response.
12.1.3.7 Black Start
Because of relatively strict geographic requirements for black-start capability, it is also not
normally procured through markets. System operators determine the systems black-start
requirements to ensure that the system could be restored to service expeditiously if it should ever
fail completely. Long-term contracts are established with selected black-start units. Each blackstart unit must be capable of starting without external assistance within 10 minutes. It must be
capable of supplying the reactive power requirements and controlling the voltage of the
energized transmission system and operating for a minimum of 12 hours.

12.2 Characteristics of Variable Generation and Experience to Date


The grid integration challenges related to higher penetration of renewable resources are
dependent on the location, technology, and penetration level. For example, rapid expansion of
wind generation capacity may create more challenges to the grid than additions of solar, wave,
tidal, biomass and geothermal plants, owing to its greater variability and unpredictability. All
have unique generation characteristics that will affect their integration into the electric grid. The
primary challenges for the non-thermal renewable generation technologies are variability and
predictability of generation output, coupled with the fact that many of these technologies do not
behave in a similar way to conventional electrical generators (Table 12-3).
Table 12-3
Comparison of non-thermal renewable generation technologies
Technology

Wind

Solar Thermal

Solar PV

Wave

Tidal Current

Development
Status

Commercial

Emerging
Commercial

Commercial

PreCommercial

PreCommercial

Energy
Source

Uneven solar
heating

Sun

Sun

Wind over
water

Gravity of
moon and sun

Power Density
(annual 8760
average per
unit area)

0.250.35
kW/m2
(US Great
Plains)

0.180.3
kW/m2
(Southwest
US)

0.150.24
kW/m2
(Southern &
Western US)

2025 kW/m2
(US West
Coast)

210 kW/m2
(Northern US)

Variability

Weather
patterns

Day-night;
clouds, haze,
and humidity

Day-night;
clouds and
haze

Sea states

Diurnal and
semi-diurnal

Predictability

Hours

Minutes

Minutes

Days

Years

12-12

Grid Integration Challenges and Technologies

Thermal renewable generation resources, such as biomass, geothermal, and biogas, will behave
much as thermal plants currently do. Therefore, the issues likely to be seen on the distribution
and transmission systems will not be different than those seen when installing conventional
sources of electricity. This chapter does not examine these in any great detail. When certain
characteristics of these plants are relevant they are mentioned, but such plants are not the focus
of this chapter.
More than 50% of currently deployed renewable generation cannot be dispatched and available
power output depends on the weather. This number is likely to grow and increase variable
generation in the grid. Thus, grid-operating flexibility is reduced and reserve requirements
increase to accommodate these less controllable resources.
Wind in particular, but also run-of-river and geothermal, are typically remote and not near
primary transmission rights-of-way. The economics of adding transmission or distribution for
renewable energy can be challenging if capacity factors are low or output is variable. Tidal and
wave energy devices are currently not near commercial deployment, so they are not focused on
in this discussion. In the case of run-of-river hydropower, not shown in the table, output may be
more predicable with changes related to wet and dry seasons and longer-term weather patterns.
Power delivery contracts, as well as environment constraints, can also affect renewable
generation flexibility. For example, many hydro facilities must meet river flow requirements
regardless of grid power needs, geothermal is often purchased as must take energy, and
municipal solid waste facilities are driven to dispose of their accumulating fuel. In contrast, solar
thermal electric technology includes about 30 minutes of natural thermal inertia and is very
conducive to the addition of additional thermal storage without significantly increasing cost. In
many areas of the West, two to six hours of thermal storage can provide solar plant flexibility
and carry afternoon solar peaks into early evening load peaks.
When wind, solar, and other variable generation resources are part of the generation mix,
operators must consider their unique characteristics when planning and operating the power
system. Conventional power plant performance metrics are designed for dispatchable
generation. These can be difficult to apply to variable generators such as wind and solar power.
Currently IEEE 762-2006 is the standard for reporting individual power plant performance. The
Standard assumes fuel is available, and indicates to what extent the rest of the plant performs
relative to its rating and availability. In contrast, performance metrics used for renewable plants
normally assume that fuel is variable and the main determining factor of plant availability. This
contrast, between traditional controllable generation and emerging variable generation, has led to
some confusion and hesitation to incorporate solar plants into a conventional generation fleet.
The following characteristics of variable generation are likely to require changes in system
operation and planning:

Variability

Uncertainty

Non-synchronous nature of the resource

Distributed generation

Remote location of some renewable resources


12-13

Grid Integration Challenges and Technologies

12.2.1 Variability
The output of most traditional generators can be planned, as well as controlled, even minute-tominute to meet the expected daily load and to balance or regulate load variations. Variable
generation output depends on the weather and is determined by characteristics such as timevarying wind speeds, solar insolation, or tidal current. Its output varies day to day, as well as
from hour to hour and even minute to minute. Seasonal changes are also likely. Figure 12-3
shows the annual wind output at a site in the Midwestern United States with hourly resolution,
and Figure 12-4 shows the hourly output of a solar PV plant over one week.
Although two variable generation facilities may have the same capacity rating and annual energy
production, different wind regimes or solar insolation patterns may result in very different
hourly, daily, and seasonal operating schedules. Consequently, aggregating output over a large
area can reduce variability because the output of different widespread wind farms or solar plants
is not perfectly correlated. Additionally, variability is reduced when examining wind and solar
PV output together.
Hourly Wind Production
250

200

MW

150

100

50

0
1

10

Month

Figure 12-3
Wind output varying diurnally, seasonally, and with weather changes

12-14

11

12

Grid Integration Challenges and Technologies

Figure 12-4
Daily variation of solar PV system output over a month (recorded for 1-MW site near
Knoxville, TN, 2012)

Figure 12-5 compares typical variations in system load (electricity demand) with variations in
wind power generation. The difference is the amount of required generation scheduling and
energy regulation for non-wind power plants. Experience to date has shown that wind output is
not likely to coincide with peak load requirements (and thus has a low contribution to resource
adequacy) and that there are interesting successes and significant challenges to absorbing higher
levels of deployment. As the footprint covered by the installed variable generation increases, the
total variability decreases because variability is smoothed as outputs from wind plants are not
fully correlated. Therefore, when considering the effect that wind plants have on the system, it
should be noted that each does not need to be balanced individually, but rather the impact on the
system net load as a whole should be considered.

12-15

Grid Integration Challenges and Technologies

Figure 12-5
Hourly load shapes with and without wind generation (from DOE 20% Wind Energy by
2030, Report DOE/GO-102008-2567, May 2008)

In addition to balancing supply and demand on an hourly basis, maintaining grid stability usually
involves moment-to-moment balancing of the load and generation by holding a constant
frequency, as described previously. A substantial load change may cause sudden changes in
frequency to which speed governors on individual generator units must respond. The number of
plants and their capacity to provide load balancing and regulation are limited in any power
system. The challenge is to have sufficient energy balancing and regulating generation with ramp
rates to meet the worst-case net load fluctuations. Introducing variable generation usually means
that additional flexibility will be required on the part of other conventional generation to
accommodate sub-hourly ramps. Figure 12-6, from NREL, shows the distribution of 5-minute
changes in generator loading requirements with and without wind generation.
Wind power outputs have significant spatial variations. Outputs from nearby wind plants are
correlated, but outputs from distant wind plants are not. This spatial variation can smooth output
in show time frames. As shown in Figure 12-7, also from NREL, short-time fluctuations of wind
power are independent regardless of distance between wind plants. However, longer-time wind
power fluctuations are correlated for nearby wind power plants.

Figure 12-6
Number of P observation changes in average power for 4 ramp rate intervals: 10 seconds,
1 minute, 10 minutes, and 1 hour (The period shown is summer of 2012; from NREL.)

12-16

Grid Integration Challenges and Technologies

Figure 12-7
Example of wind power plant output correlation with distance (from NREL)

12.2.2 Uncertainty
As well as being variable, most variable generation (with the exception of tidal power) is
unpredictable. The degree of predictability varies with technology; wind power can be well
forecast for the next few hours, but beyond that is difficult to predict. Solar PV can be difficult to
forecast over one hour for smaller arrays. This uncertainty of output increases the need for
regulating reserves, and should also be considered when making unit commitment decisions. The
likely forecast error, often measured as mean absolute error (MAE) or root mean square error,
needs to be accounted for by including additional flexibility in the commitment, so that forecast
errors can be covered by plants in an economical and reliable way.
The prediction error for variable generation increases with timeforecasts are less accurate a day
ahead than an hour ahead. As with variability, uncertainty is decreased when variable generation
is spread out over larger areas. The main prediction errors tend to be for large ramps, which are
decreased when a larger area is considered since there is not perfect correlation between the
output of plants in different areas.
12.2.3 Non-Synchronous Nature of the Resource
Wind and solar PV (and other inverter connected generation) does not connect synchronously to
the grid, which is different from other conventional resources, which spin at the same frequency
as the grid. This means that these resources do not naturally provide inertia or primary frequency
response. Increased levels of inverter-based generation therefore cause concerns about the
frequency response capabilities of the system. As well as not providing the frequency response
characteristics, they will also displace other generators which can provide these critical services,
and thus the ability of the system to respond to faults may be degraded. Recently, various
vendors have begun to implement active power control on wind turbines which would allow this
service to be provided from wind, by emulating inertia and primary frequency response. Solar
PV may also be able to provide primary frequency response in the future.
12-17

Grid Integration Challenges and Technologies

In addition to frequency issues, wind and PV may displace traditional reactive power resources
on the system, both for normal operation and following a fault. Without these resources
available, there may not be sufficient reactive resources on the system to maintain voltages at
required levels. In addition, wind and/or PV plants may actually consume reactive power if not
controlled to avoid this, thus making the situation worse. However, in the past decade, much
work has been done to provide interconnection standards such that now most wind plants and
larger PV plants will provide reactive power support to the system, either through their own
power electronics or through associated reactive power equipment.
12.2.4 Distributed Nature of the Generation
Wind, solar, and other renewable generation are deployed both as large projects interconnected
to the transmission system and as small distributed facilities connected to the distribution system.
Even though wind is considered primarily as a bulk-generation resource, a growing number of
installations are being deployed on distribution systems at or near smaller load centers. Because
distributed systems are designed for one-way power flow, distributed generators require special
arrangements for connection and protection.
Two types of distributed wind and solar applications are common. Some facilities use local
distribution to collect energy from individual units and deliver it to the distribution system. Other
facilities are more dispersed and are individually connected to the system. Integration issues that
may surface in these distribution scenarios include voltage regulation, protection coordination,
and interferences or harmonic flows along rights-of-way. In general, these are the same
integration concerns that occur when generation is added into the distribution system.
12.2.5 Remote Location of Some Renewable Resources
Although good quality renewable energy sites exist throughout the United States, the best wind
and solar resources are located relatively far from large load centers. In addition, sites selected
for building large wind farms are likely to be away from population centers. This is not unique to
wind. For example, other generation types, such as hydro and coal-fired stations, may also be
located at significant distances from load centers. However, wind developments have tended to
not consider the available transmission capacity to these sites or include the transmission-related
costs of taking the wind to market.
12.2.6 Experience of Integration of Wind Power
In some regions, such as Northern Europe and the upper Midwest and California in the United
States, relatively high penetration levels are managed via a strong existing transmission
backbone and by relying on other controllable generation, such as gas turbines and hydropower.
Other regions with less flexibility have experienced forced curtailment of wind generation, such
as transmission congestion in west Texas, transmission and distribution line voltage control
issues along the Buffalo Ridge in Minnesota and northwest Iowa, and challenges for grid
operators to follow load and manage ramp rates in territories such as the California ISO, New
Mexico and the Big Island of Hawaii (EPRI Wind Integration Technology Assessment and Case
Studies, 1004806, March 2004). Curtailment will generally happen for one of two reasons: lack
of flexibility in the generation fleet (plant cannot be ramped down or turned off quick enough to
accommodate wind) or congestion in the transmission network.
12-18

Grid Integration Challenges and Technologies

As penetration levels of variable generation increase, so does the challenge of managing this
resource. Wind penetration levels are already 10% to 15% of capacity in some areas of New
Mexico and California. Utilities such as Public Service of Colorado have seen instantaneous
penetrations of over 50% of demand being met by wind energy during low load periods. Even
higher penetrations have occurred on islands and in European grids. Although in the main,
integration of this level of wind generation has not caused any particular issues, with higher
penetration, the response of wind turbines to weather and power system transients have created
operating concerns. It can also have an impact on the frequency performance of the system as
shown in Figure 12-8.

Figure 12-8
Example of wind plant output variation on area control error (ACE)
Courtesy of Public Service of New Mexico

12.2.7 Experience of Integration of Solar Power


Compared to wind, solar energy may be easier to integrate into the electric grid. Good solar
resources are more often close to load centers and are more likely to be available during daily
load peaks. One exception is the rapid rate of change in output occurring as clouds pass over
solar arrays, which may cause increased variability.
The energy resource potential for photovoltaic electric generation is very well distributed.
Although some areas, such as the southwestern United States, Spain, northern Africa, and
Australia, have excellent solar resources, the range of the best to the worst locations for solar is
much narrower than that of most of the other renewables. In the United States, the best locations
are less than twice as sunny as the worst. This is illustrated in Figure 12-9, which shows the
relative cost of solar electric energy for a number of U.S. cities based on resource availability.

12-19

Grid Integration Challenges and Technologies

By its nature, solar energy is more available during the summer than in the winter, as shown in
Figure 12-10. For most areas with summer peaking of demand, this profile is a good match with
load requirement. Even so, seasonal and daily variations of available output affect capacity factor
and overall system economics.

Figure 12-9
Relative cost of photovoltaic electricity due only to resource variability

Figure 12-10
Monthly solar energy variation over one year

12-20

Grid Integration Challenges and Technologies

Widespread deployment of solar would certainly challenge both retail and wholesale grid
operations. Some studies estimate that such challenges could manifest as early as 2020 in some
parts of the country. (See the EPRI white paper, Distributed Photovoltaics: Utility Integration
Issues and Opportunities, August 2008, 1018096, as well the prior Western Governors
Associations Clean and Diversified Energy Initiative Solar Task Force Report, completed
January 2006.)

12.3 Variable Generation Integration Impacts


The technical issues and opportunities for integrating variable renewable generation differ
depending on its point of connection to the power system. The key question is: What grid
changes are needed to enable large-scale deployment of distributed renewable generation and to
integrate with other load and generation resources? These changes will vary depending on low-,
medium-, and high-voltage points of connection to the power system. The appropriate strategy
also depends on the penetration level relative to the power system capacity at the point of
connection. The rules, concerns, and potential paybacks all vary at different system levels and
have been treated separately in the past. This point is illustrated in Table 12-4. From this, it can
be seen that wind and solar PV connected to the bulk system will results in very different
challenges than those connected at the distribution system. Therefore, these are described
separately in the remainder of the chapter.
Table 12-4
Distributed power system performance expectations at various connection points in the
electric system
Distributed
Generation (DG)
Expectations and
Connection
Connection at End
Use (Low Voltage)
Connection at
Distribution
(Medium Voltage)
Connection at
Transmission
(High Voltage)

Interconnection
Rules

System Integration
Concerns

Local connection
requirements, e.g.,
IEEE 1547 and
derivatives.

Feeder-level issues such


as power flows, protection,
and voltage impacts, e.g.,
at higher penetration
levels (IEEE P1547.8)

Special grid rules for


generators <20 MW

Understanding system
response for planning and
analysis of scenarios

Local and System


Values or Payoff
Power, heat, load
control, quality and
reliability
Ancillary service support
to utility T&D, e.g.,
reserve capacity,
demand response,
expansion deferral, etc.

12.3.1 Transmission Connected Generation Impacts


This section addresses renewable generation integration issues and costs with regard to power
system planning and operation. It is based on information presented in Impacts of Wind
Generation Integration (EPRI 21023166, April 2011), as well as on other industry reports and
studies. Previous EPRI reports in the area also include Wind Power Integration Assessment and
Case Studies (1004806, March 2004), Wind Power Integration: Smoothing Short-Term Power
Fluctuations (1008389, March 2005), Wind Power Integration: Energy Storage for Firming and
Shaping (1008388, March 2005) and Survey of Wind Integration Study Results (1011883, March
2006).
12-21

Grid Integration Challenges and Technologies

Depending on the location and the relative size of renewable output, integrating variable
generation ultimately leads to increased variability and uncertainty, remote locations requiring
increased transmission, with impacts on system dispatch, network stability, load following, and
load balancing. The impacts of wind generation on both technical operating limits and ancillary
costs are summarized in the EPRI report, Survey of Wind Integration Study Results (1011883,
March 2006), covering North American sites. A related report also covers European sites (EPRI
1012734, March 2007). A number of key integration impacts have been identified.
Increased Need for System Flexibility and Transmission
The need for additional system flexibilitythe ability of the system to respond to variability and
uncertaintyis principally determined by wind-related features, with wind penetration being the
single greatest factor; studies and experience to date have shown that most power systems can,
on an energy basis, reliably accommodate up to 10% penetration, with only minor cost and
operating impacts. The cost to provide this system flexibility (as opposed to the costs to actually
connect wind to transmission and distribution networks) is best described as the additional costs
induced by balancing the variability and uncertainty of wind, and is a function of the existing
power system, plus any new investment required. A 2009 IEA study that compared the wind
integration costs from studies across Europe and North America, cited a cost range of 0 to 4.0
/MWh, or about 0 to $5.63/MWh;87 see Figure 12-11. In addition to these balancing costs to
provide flexibility during operations, another key aspect is to ensure that flexibility is provided in
the planning process.

Figure 12-11
Results from estimates for the increase in balancing and operating costs due to wind
power [1]88

87

March 29, 2011 exchange rate: 1 = $1.41.


The currency conversion used here is 1 = 0.7 and 1 = 1.3 US$. For the UK 2007 study the average cost is
presented here; the range in the last point for 20% penetration level is from 2.6 to 4.7 /MWh.

88

12-22

Grid Integration Challenges and Technologies

With regard to moving wind energy to load, transmission issues are particularly challenging.
From planning to permitting, from capital cost estimation to cost allocation, there is no clear
indication how and whether sufficient transmission will be constructed to tap into our vast,
remote, domestic wind resources. However, it remains that of all the flexible resource
possibilities, it seems that increasing the transmission capabilities of an area, both in improved
transmission within the area and improved transmission to other regions in order to access
flexibility, is key to integrating variable generation.
Minimum load considerations present a technical limit on the amount of wind that a power
system can accommodate. Proposed federal environmental regulations directed at thermoelectric
power plants will inadvertently decrease operational flexibility from mainly baseload plants
namely by increasing higher minimum loads and slowing ramp rates. At periods of low net load,
thermoelectric power plants may be shut down to accommodate windfor example, over a night
or a weekend. Many combined cycle natural gas-fired units and some coal units already operate
in this two-shift mode, meaning they shut down overnight; this type of operation may be needed
for more units, which will be particularly difficult for those which have operated as baseloaded
units for most of their lifetime. This additional generator cycling is expected to be an
increasingly challenging aspect of variable generation integration. With increased requirements
for flexibility, plants operated in a more flexible manner may see increases in variable O&M
costs, increases in forced and scheduled outages, reductions in lifetime, and reductions in
efficiency, all in the face of reduced operating hours and possible reduced revenue. Therefore,
planning and operating the power system will have to incorporate these issues to ensure that the
flexibility required is available. System planning will likely have to account for increased
flexibility requirements in the future as well as traditional resource adequacy requirements.
Situational Awareness Impacts
With increased levels of variable generation, operators will be required to deal with larger, more
frequent ramps in system net load (load minus variable generation) as well as more uncertainty
in when they occur. This will require greater situational awareness. With increasing levels of
variable generation, the system will move from a mainly deterministic system to a more
stochastic system, with increased uncertainty causing significant challenges for system operators.
In particular, bulk system operators faced with significant amounts of distributed energy
resources, such as PV, may not have direct visibility of the minute-to-minute production of that
resource, and thus will not know what the conventional plant on the system needs to be doing.
This is already being seen in places such as the Big Island of Hawaii and, to a lesser extent, parts
of Germany and Southern California. Figure 12-12 gives an example of the type of large ramps
that may be seen by an operator with increased levels of variable generation. This shows how
reserves available to the operator dropped off as wind on the system decreased at the same time
as a load rise. It can be seen that this was not forecasted wellforecasting, though improved
significantly in the past decade and now in use in all of the North American ISOs, is still never
going to be perfect, increasing the need for situational awareness.

12-23

Grid Integration Challenges and Technologies

Figure 12-12
Large wind ramp in Texas showing need for increased situational awareness

Impacts on System Stability


With increased levels of inverter-based generation, there are concerns about system stability. In
particular, frequency response and voltage, as mentioned earlier, will both be of concern. These
are starting to be addressed by improved interconnection standards, which require wind and PV
plants above a certain size to offer support to the system. In particular, voltage concerns have
been reduced significantly with standards being applied for inverter-based generation to connect
to the bulk system. Although in the United States these are not part of a grid code, most utilities
do ensure that new resources can provide a certain range of reactive power control. An example
is given in Figure 12-13, which also shows how losses on collector system may reduce the
reactive power available.
From the point of view of frequency response, increased levels of wind and/or PV may reduce
the primary frequency response. A recent study for the California ISO [2] has examined the cases
with high levels of wind and solar generation. It was found that reduction in system inertia due to
higher levels of renewable generation will not have a significant impact on frequency response
when compared to the reduction of primary frequency control action. However, fast transient
frequency support using controlled inertial response from wind power will help increase the
under-frequency load shedding (UFLS) margin and reduce the risk of load shedding. It was
shown in [2] that benefit of these responses can be several times greater, per MW, than was
observed for primary frequency control in the conventional generation fleet. Currently, only
ERCOT (in Texas) and Hydro Quebec of the North American system operators require all wind
plants to provide active power control such that they can respond to faults on the system. It is
expected that as vendors continue to develop this capability, more utilities or ISOs will include it
in interconnection agreements. In many areas of the worldfor example, in Irelandinertial
12-24

Grid Integration Challenges and Technologies

concerns due to displaced conventional plant are limiting the amount of energy that can be met
instantaneously by non-synchronous sources (in Ireland, currently 50% can be met before wind
is curtailed; this is expected to be raised in the near future).
0.5

Reactive Power (per-unit)

0.3
0.1
-0.1 0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

-0.3
-0.5
-0.7
-0.9

Wind Plant (w ith collection system)


Wind Plant (no collection system)
Active Pow er (per-unit)

Figure 12-13
Power factor curves for example wind plant

A final impact on transmission system stability may come from significant penetration of
distributed energy resources, such as small (rooftop) PV. Here, issues such as a lack of visbility
of output and the impact on load are important and will make it more difficult to operate the
system In addition, these generators are currently connected through the IEEE 1547 standard for
distributed generation, which has a requirement to disconnect the generation if a low voltage or
low frequency event is seen. Although the primary purpose of that is to maintain safety and
ensure that local reliability is maintained, it has the side effect of possibly causing a fault on the
bulk system to be exasperated as the PV disconnects. There is currently a proposal to change
IEEE 1547 to accommodate bulk system as well as distribution system needs to account for this.
As well as the impacts listed earlier, other impacts may also arise from the integration of variable
renewable resources in the bulk system. However, those listedincreased need for flexibility
and transmission, possible impacts on system stability, and increased need for situational
awarenessare those which have been identified as most crucial based in studies and experience
to date.
12.3.2 Distribution Connected Generation Impacts
Todays electric distribution systems have evolved over many years, in response to load growth
and changes in technology. The largest single investment of the electric utility industry is in the
distribution system. Most common are radial circuits fed from distribution substations designed
to supply load based on customer demand requirements while maintaining an adequate level of
power quality and reliability, as shown in Figure 12-14.

12-25

Grid Integration Challenges and Technologies

Figure 12-14
Todays typical distribution feeder topology [1]

Figure 12-14 shows how the system is designed to be fed from a single source. Protection is
based on time-overcurrent relays and fuses that use nested time delays to clear faults by opening
the closest protective device to a fault and minimizing interruptions. It is designed to safely clear
faults and get customers back in service as quickly as possible. In areas of high load density,
network systems are common. These systems are fed by multiple transmission sources, and
thereby provide high reliability. Both radial and network distribution systems have been
designed to serve load, with little planning for generation connected at these levels.
Circuit sectionalizing switches are manually controlled to restore load in unfaulted sections
downstream from a failure. The system voltage is maintained in compliance with American
National Standard Institute (ANSI) Std. C84-1, which specifies that service voltage be delivered
within 5% of the system rated voltage. These systems are generally considered to be ready to
support small photovoltaic installations without change, when the PV inverters meet appropriate
IEEE, UL, and FCC standards, and the overall penetration levels are very low.

12-26

Grid Integration Challenges and Technologies

The designs and technologies associated with todays distribution systems impose important
limits on the ability to accommodate rooftop solar and other distributed generation, end-user load
management, distributed system controls, automation, and future technologies such as plug-in
hybrid electric vehicles (PHEV). The system characteristics that lead to these limitations include
the following:

Voltage control is achieved with devices (voltage regulators and capacitor banks) that have
localized controls. These schemes work well for todays radial circuits but they do not handle
circuit reconfigurations and voltage impacts of local generation well. This results in limits on
the ways circuits can be configured and important limits on the penetration of distributed
resources. It also limits the ability to control the voltage on distribution circuits for
optimizing customer equipment energy efficiency.

There is little communication and metering infrastructure to aid in restoration following


faults on the system.

There is no communication infrastructure to facilitate control and management of distributed


resources that could include renewables, other distributed generation, and storage. Without
communication and control, the penetration of distributed generation on most circuits will be
limited. In addition, the distributed generation must disconnect in the event of any circuit
problem, limiting reliability benefits that can be achieved with the distributed generators.

There is no communication to customer facilities to allow customers and customer loads to


react to electricity price changes and/or emergency conditions. Customer-owned and
distributed resources cannot participate in electricity markets, limiting their economic
payback in many cases. Communications to the customer would also provide feedback on
energy use, which has been shown to help customers improve energy efficiency.

The infrastructure is limited in its capacity to support new electrical demand such as home
electronics and PHEVs. These new loads have the potential to seriously affect distribution
system energy delivery profiles. Communication and coordinated control will be needed to
effectively serve this new demand.

At the same time, the distribution system infrastructure is aging, resulting in concerns for
ongoing reliability. Utilities are struggling to find the required investment just to maintain the
existing reliability, much less achieve higher levels of performance and reliability. New
automation schemes are being implemented that can reconfigure circuits to improve reliability,
but these schemes do not achieve the coordinated control needed to improve energy efficiency,
manage demand, and reduce circuit losses.
The bottom line is that todays power distribution system has not been designed for distributionconnected PV or other high penetration of distributed generation. In the past this was not an
issue, but today, with larger amounts of PV connecting to the electric system, we can expect new
challenges in how distributed and variable generation can be safely and reliably interconnected.

12-27

Grid Integration Challenges and Technologies

The following impacts will be seen when large amounts of PV are integrated at the distribution
level:

Steady-state voltage regulation: Because distributed energy resources (DER) raise voltage
levels when they inject power into the grid, they may cause high voltage conditions at high
penetration levels. This will have an impact on the stability of the grid, and needs to be
mitigated.

Voltage flicker: Voltage flicker is a sudden change in voltage that occurs in seconds or
fractions of a second that can cause objectionable changes in the visible output of lighting
systems. The existing standards are likely adequate for high penetration of distributionsystem-connected PV.

Harmonics: Harmonics are distortions in the regular 60-Hz sine wave in North American
power systems. Too much harmonic distortion can cause adverse operation of customer and
utility equipment. Improvements over the past 10 years in the quality of inverter output have
drastically reduced issues with PV system harmonics. The existing standards are likely
adequate for high penetration of distribution system connected PV.

Unintentional islanding: Utilities are regularly required to isolate a section of the power
system by disconnecting the section with network protectors or switches. Unintentional
islands can be established when a section or the grid is isolated from the substation supply
while the load continues to be maintained by an energy source within the isolated section that
continues to provide power. Unintentional islands pose a threat to proper utility system
operation due to their unsynchronized nature, hazard to public safety, and inadequacy to meet
customer loads. The IEEE 1547 standard and utility interconnection guidelines require PV
systems connecting to the network to have anti-islanding protection set to disconnect in the
event that the network voltage or frequency goes outside of predefined limits. As a result,
unintentional islands are not currently considered a significant concern. However, current
anti-islanding techniques require PV systems to drop offline rather than ride-through
temporary faults, contributing to voltage drop or frequency problems. Large amounts of PV
could be prone to tripping during severe transmission system disturbances that typically
affect a wide geographical area.

In order to directly address the issues related to connecting large amounts of PV in the
distribution system, four key areas will need to be addressed:
1. Voltage regulation practices
2. Overcurrent protection practices
3. Grounding practices
4. Switching and service restoration practices
Fortunately, due to the robustness of the existing design practices, the distribution system can
handle some level of distributed renewable generation without modification. The basis for
simplified interconnect rules, such as California Rule 21, is that some level of robustness in
existing distribution design allows connection without detailed engineering studies. Standards
such as IEEE 1547-2003 and U.L. Inverter Test Standard 1741 evolved to enable connection
without major design changes to the electric system.
12-28

Grid Integration Challenges and Technologies

In most areas today, distributed renewable generation is treated as negative load and the usual
functions of generation are not expected or required. However, as PV deployments grow, feeder
cases that cross the threshold are expected, requiring changes in operation rules such that
distributed generation will need to provide voltage support and eventually energy balancing. This
is anticipated to occur initially on individual feeders with high penetration and later on at the
substation and sub-transmission level. Table 12-5 shows how the role and the rules for behavior
of distributed generation need to need to change with increasing penetration levels.
Table 12-5
Grid penetration scenarios and changing role of distribution generation
% of Generation

2%

10%

Grid Penetration
Scenarios

I. Low-numbers
and level of PV
with relatively stiff
grid connection

II. Moderate-level
of PV with
relatively soft grid
connection

III. High-level of
PV with capacity
of grid less than
the load demand

IV. PV operates
part time as an
island or microgrid

PV Impact and
its role in the
Grid

Very low, not


significant to grid
operation

Non critical, can


affect distribution
voltage near PV

Critical to power
delivery and
meeting demand

Primary power
source for
standalone
operation

Interconnection
and Integration
Objectives

Non interference,
good citizen and
compatible

Manage any local


distribution
impacts

Engage PV for
system
operations and
control

Rely on PV for
stability and
regulation

Rules / Standard
Operating
Procedures

IEEE 1547-2003
current practice
radial feeders

Modified 1547,
add network and
penetration limits

New rules include


operation and
grid support
requirement

Standalone rules
that are system
dependent

Main Concerns
with Respect to
System Dynamic
Grid Impacts

-Voltage and
current trip limits

-Interfere with
regulation

-Availability

-Availability

-Recovery times

-Regulation
provided

-Load following

-Response to
faults
-Synchronization

-Coordination

-Ramping
response

-Normal and
reserve capacity

-Islanding

30%

100%

-Voltage control

-Interactions of
machine controls
Transitions On- and Off-Grid

12-29

Grid Integration Challenges and Technologies

12.4 Integration Technologies and Strategies for Transmission-Connected


Renewable Generation
Renewable generation integration technologies include generator interface and output controls,
new flexible resource including energy storage and demand response, and renewable generation
forecasting tools. Together with these technology advances that allow integration, multiple
strategies related to how the system is planned and operated can enable integration. As seen in the
previous section, the primary issues for integration on the transmission system relate to balancing
supply and demand, which becomes more difficult with the additional variability and uncertainty
of variable generation, as well as some impacts on transmission requirements and stability.
12.4.1 Variable Generator Interface to Maintain System Stability
The variable generation technologies themselves can be used mainly to meet system stability
challenges mentioned previously. In the case of wind, solutions including better wind turbines,
improved fault tolerances, application of power electronics for stabilization, compensation and
the like are already being applied in wind farm applications, whereas others are only in the
design and development stage. In the past decade, interconnection standards have advanced
significantly for variable generation connecting to the bulk system. Most of the wind plants
connected in the past five years have to perform to interconnection standards on low voltage ride
through, reactive power control for steady state and dynamic performance. This has been
achieved in different manners for different types of wind turbine technology. For example, Type
4 wind turbines, with full four-quadrant control, can provide reactive power support through
power electronics. On the other hand, induction-based machines, as were more common in the
past, may need separate equipment installed onsite for reactive compensation, including switched
capacitor banks, static series compensators, static voltage regulators, and so forth. This reactive
compensation equipment was described in detail in previous versions of this report; however, as
the technology has advanced significantly recently and interconnection standards are requiring
this from all new plants, it will not be focused on here. Similarly, low voltage ride through
capabilities on bulk connected wind and PV plants are now standard in all new build due to
interconnection standards and thus are not discussed in detail. From a stability standpoint,
although reactive compensation will always be relevant, it is not as crucial a limiting factor in
wind integration as it may have been in the past. In general, therefore, transmission connected
variable renewable generation technology has improved significantly in the past decade in terms
of how it can maintain system stability.
An area related to power system stability that does require further R&D and technology
development is in the area of frequency response. Several vendors now have active power
control schemes available for wind turbines, such as GEs WindINERTIA. These provide
emulated inertial response and primary frequency response to aid systems in maintaining
frequency response within desired levels after a fault on the system. These technologies are now
being demonstrated in various wind plants throughout the world. More details can be found at a
joint EPRI/National Renewable Energy Laboratory project (website:
http://www.nrel.gov/electricity/transmission/active_power.html).

12-30

Grid Integration Challenges and Technologies

12.4.2 Variable Energy Forecasting Tools


Variable energy forecasting is becoming increasingly important as installed capacity continues
to grow. Development of wind forecasting tools is focusing on improving forecast accuracy,
predicting high ramp rates of wind generation in response to rapid changes in wind speed, and
integrating the forecasts into grid operations. All the major ISOs in the U.S. system now use
centralized forecasts provided regularly by forecast providers, for both day-ahead and close to
real-time operations. Solar forecasting, which is further behind, is still being deployed, with
significant research ongoing, particularly in the area of the forecasting of large amounts of
distributed PV it is expected that as the levels of installed solar power increase, solar forecasts
become as important and widespread as wind forecasts. Solar forecast will have different
challenges and opportunities than wind forecasts. Whereas wind speed is the driving factor of
wind forecasts, insolation is the main factor for solar forecasts. Cloud formation and movement
can be extremely hard to predict, but on the other hand, for short term forecasts, sky imaging and
satellite techniques can be used. This may make solar variability more challenging to forecast
day ahead, but easier in the short term.
Forecasts are used by ISOs to ensure that sufficient resources are available to meet load, and to
ensure that the system can meet all expected ramps. In addition, ramp forecasts are now often
provided to ISOs, which report the likelihood of a large ramp occurring at various forecast
horizons. Centralized forecasts such as this are generally made available to wind and solar power
producers. These have various contractual provisions that were devised to handle deviations
between forecast and actual power delivered. Better forecasting tools will directly help with the
characteristics of wind power, with positive financial consequences. It is clear that more accurate
wind forecasts help both the buyer and seller of wind energy. From the point of view of the
seller, better forecasts minimize contractual penalties, and greater confidence maximizes the
value of the wind power. From the buyers perspective, accurate wind power forecasts reduce
risk and increase power system reliability.
One of the main results of the uncertainty of wind and solar PV is increasing reserve
requirements. This additional reserve is required to maintain the same level of system/load
balance in minute-by-minute regulation and minimize any additional operating cost due to the
balancing of intra-hour slow variation. Expected limits in control and variability for wind power
increase reserve requirements, as additional reserves are required for managing variability and
uncertainty. Better forecasts can reduce the need for reserves and increase the value of wind and
solar PV energy.
Many different components go into developing an accurate wind power forecast. They include
numerical weather predictions; real-time and recent wind speed, direction, and energy
generation; physical models of wind flow; wind plant power curves; and artificial neural
networks and other sophisticated statistical techniques [16]. EPRI has been monitoring and
evaluating the latest forecasting technologies for many years, and worked with DOE and the
California Energy Commission (CEC) to develop and test wind forecasting systems in Texas and
California [1725].
All forecasting systems are similar in some respects, but they use different techniques to forecast
wind speed, direction, and corresponding wind-energy generation. The model structure typically
blends physical and statistical models, with the blend dependent on the forecast window, such as
the forecast interval into the future. Short forecast time intervals up to about six hours typically
12-31

Grid Integration Challenges and Technologies

rely on real-time wind speed, direction, and power generation data; rapid-update numerical
physical model data; simple persistence, and adjustments using self-learning statistical methods
such as screening multiple linear regression (SMLR), artificial neural network (ANN), and
support vector machines (SVM) (Figure 12-15).
Longer forecast time intervals greater than four to six hours typically use numerical weather
prediction, physical wind flow models, and statistical techniques to predict weather conditions,
wind flow patterns near the wind turbines, and wind energy generation. The technology has
advanced to the point that the mean absolute error of same-day wind energy forecasts has
decreased to 6% to 8% and the error of next-day forecasts has decreased to about 15%.

Predictors

P1,P2,...
SMLR
ANN
SVM

Predictand

F
Training
Algorithm

F = f(P1,P2,...)
Figure 12-15
Wind predictions adjusted using self-learning statistical methods
Source: AWS Truepower [26]

Figure 12-15 is a schematic diagram of an example longer-term forecast model used by AWS
Truepower to generate 48-hour forecasts [26]. The longer-term forecast model typically involves
the following steps:
1. Download numerical weather forecast data from the National Center for Environmental
Prediction (NCEP) website.
2. Calculate wind flow around wind plant using a meso-scale or wind-flow model.
3. Apply model operating statistics (MOS) to estimate adjusted wind speed and direction at
wind plant.
4. Estimate hourly power generation at 100% availability using wind plant model.
5. Apply MOS to adjust the hourly power generation forecast.
6. Adjust forecast to account for curtailment and forced outages of individual wind turbines (if
available).
7. Issue the 48-hour forecast of hourly wind speed and energy generation.

12-32

Grid Integration Challenges and Technologies

Although forecast accuracy has improved and commercial forecasting systems are being applied
throughout the world, there are still several issues that stand in the way of further reductions of
wind forecast error. They include (1) the general unavailability of robust real-time wind speed,
direction, and generation data for individual wind plants and for regional wind generation
resources; (2) the unavailability of operational status data for individual wind turbines required
to adjust forecasts for machines that are shut down or curtailed; and (3) methodologies to
accurately forecast rapid changes of wind speed and generation and that result in high ramp rates
of regional wind generation.
A recent Department of Energy project, the Wind Forecast Improvement Project (WFIP), is
examining the benefits of increasing the number of sensors and the nature of the sensor
technology, being provided to the National Oceanic and Atmospheric Administration (NOAA),
which runs all the numerical weather prediction (NWP) forecasts for North American entities
(the ECMWF, the European body, and other bodies also run forecasts for North America as part
of their global models). This project is examining how these data can be utilized with new NWP
forecasting models developed at NOAA, such as the Rapid Refresh model and the High
Resolution Rapid Refresh model, which runs more frequently and at higher resolution than
previous models, thus allowing more up-to-date and more granular data, which can then be used
as input to the commercial providers forecasting tools. Similar work is expected for solar
forecasting.
As the installed wind capacity continues to grow, the value of more accurate wind energy
forecasts will increase. Thus, there will be a large incentive to further improve forecast accuracy
in future years. The need for more accurate forecasting of high ramp rates up and down has
become especially important because the combined output of large concentrations of wind power
plants can change quickly and affect operation of the grid. Because of the complexity of weather
impacts on wind plant output, ramp rate forecasting is particularly challenging. The factors
affecting ramp forecasting include passage of weather fronts and storms that cause sudden
changes in wind speed and direction, large-scale eddy currents that cant be predicted, and
variations of atmospheric stability that affects local wind patterns.
A final recent development in variable generation forecasting is in the area of probabilistic
forecasting. This method produces forecasts with uncertainty intervals around a mean forecast.
This allows operators greater knowledge of how confident the model is in the forecast ability, as
well as giving operators greater knowledge of whether they will have a wide range of forecast
error for any particular forecast. It can also be used with stochastic optimization based methods
as described later to better operate the system.
12.4.3 Increased Diversity through Increased Transmission and BA Cooperation
As mentioned earlier, many integration studies show the benefit of diversity when integrating
variable generation. Increased diversity in the area in which variable generation is installed
reduces variability and uncertainty, making variable generation easier to manage. For example,
while wind power can be very variable at one wind plant, the variability of 10 wind plants will
be significantly less. Considering a larger area increases the flexible resources that can be used to
deal with wind power variability, as well as provide inertia. Therefore, strengthening the
transmission network allows variable generation to be accessed and its total impact reduced. This
would also require increased balancing area (BA) cooperation to ensure the flexible resources on
12-33

Grid Integration Challenges and Technologies

the system as a whole are well utilized to minimize the effect of variability. Instead of each
balancing area needing to manage its own variability, increased cooperation would allow a
sharing of reserves needed to manage variability. The total reserves needed over all regions
would thus be decreased, and the flexible resources needed could be reduced. A good example of
this can be found in Europe, where Denmark, with very high wind penetration, is able to access
flexible resourcesparticularly hydroin neighboring Norway to mitigate variability effects.
Currently, there is much focus in the Western Interconnection of the United States on possible
costs and benefits of setting up an energy imbalance market there, which could be used to
share energy imbalance requirements throughout the region rather than each BA being required
to meet its own.
12.4.4 Flexibility from Conventional (Non-Variable Generation) Plant
Increased demand for flexibility from the conventional fleet can be met by building new
technology, such as energy storage, and can also be managed, to a certain extent, by existing
resources. Dispatchable plantswhich can respond to commands from the system operator to
change outputare the dominant source of flexibility today, and will remain of crucial
importance in the future. All types of dispatchable plants can change their output, but certain
types are technically better able to respond, both in the extent to which they can change their
output as well as the time it takes to do so. The critical characteristics of conventional plant
which can be used to maintain the balance of supply and demand are their ramp rates, startup
times, and minimum turn-down level.
Expected or unexpected increases in wind power output will force conventional generators to
ramp down or shut-down entirely, while reductions in wind power output, will force
conventional plant to ramp up their output, or if sufficient ramping capability is not available,
fast-starting units will need to come online. Some generation types (such as hydro or even open
cycle gas turbines) are more suited to frequent cycling, but for others, particularly units designed
for baseload operation, the large temperature and pressure gradients associated with cycling
operation impact the plants heat rate and can accrue large levels of damage within the plants
components. Baseload generators were designed to be run at, or close to, maximum output on a
continual basis. As such, the components of these generators were designed and specified for
creep conditions (i.e., constant stress), however in reality these generators are now operating
under fatigue conditions (i.e., fluctuating stress). Creep and fatigue can interact in a synergistic
manner in that creep will reduce fatigue life and, likewise, fatigue reduces creep life.
Therefore, when a baseload unit, which has been operating under creep conditions, switches to
cycling operation, the creepfatigue interaction renders the unit highly susceptible to component
failure leading to increased maintenance requirements and forced outage rates. Thermal shock,
metal fatigue, corrosion, erosion, and heat decay are common damage mechanisms that result
from cycling operation. However, the time it takes before this damage manifests itself is
variable. Analysis by Intertek Aptech suggests that following a switch to cycling operation it can
take in the region of five to seven years for the effects to be seen in a new plant or nine months to
two years in an older plant. Thus, plant operators and system planners, when considering the
issue of plant cycling, will need to take a longer-term perspective.

12-34

Grid Integration Challenges and Technologies

12.4.5 Flexibility from Storage and Demand Side Resources


With the increased levels of wind power being seen worldwide and the projected further
increases of both wind and solar PV over the coming decades, there has been significant
attention paid to emerging flexible resources such as storage and demand side options as a means
to meet many of the challenges associated with wind integration. In particular, these resources
are seen as a means to meet the increased flexibility requirements of systems with high wind due
to variability and uncertainty of wind across time scales from seconds to minutes to hours or
even days. Storage and demand side options are both characterized by the ability to respond
quickly and accurately, to increase load, as well as provide generation resources and by the fact
that they are energy limited in nature. Demand response can contribute to wind integration in
different ways, using either price- or incentive-based programs; the extent to which they can be
used to integrate wind depends on the nature of the program. Electricity storage comes in many
forms, with pumped hydro storage presently accounting for the majority share, along with recent
advances in battery and compressed air energy storage. The ability of storage to contribute to
integrating wind power is limited mainly by the cost of deployment and the efficiency losses.
Although these features may not be as significant for demand response, there are significant
regulatory and policy barriers in deploying demand side options. These emerging resources
provide support for wind integration in particular regarding provision of ancillary services and
reduction of wind curtailment.
Storage options include pumped hydro, compressed air, thermal storage, battery storage, and
flywheels. These have a wide range of operating characteristics; some are more suited for certain
types of operation than others. For example, flywheels have a very small capability for storing
energy, and thus may not be good for performing long-term load following or smoothing. On the
other hand, they can provide frequency regulation very accurately. Demand side options can be
categorized as either price based (e.g., time of use pricing) or incentive based (e.g., aggregation
of large number of water heaters providing frequency regulation). Technologies here include
large industrial loads such as aluminum smelters, but also aggregated residential and commercial
loads such as electric vehicles, space heating, and air conditioning. In the next decade, up to tens
of GWs of demand side resources could be available according to various projections, including
from NERC and FERC.
Storage and demand side options seem ideally placed to provide many of the services required
for increased penetrations of variable generation. In particular, they can help on the provision of
increased regulating and flexibility reserves. They can also aid in longer balancing requirements
(day-ahead and hourly changes in net load) and in some cases help improve system stability. In
many regions they may be used to improve transmission utilization, and possibly defer upgrade
of transmission or distribution. At the same time, there are significant challenges. Demand side
response is still relatively small in all but a few regions (such as PJM, where it can provide up to
25% of synchronous reserve), and it is not clear exactly how big the resource will become.
Similarly, many of the energy storage technologies are still in development phase, and are
currently too expensive to justify their costs except in relatively rare cases. This is expected to
change as markets adopt to increasing levels of variable generation and place more value on
flexibility and as the technologies themselves improve.

12-35

Grid Integration Challenges and Technologies

12.4.6 Grid Planning and Operating Tools


New tools to enable grid operators to better visualize the real-time status of the grid and
anticipate contingencies are needed to moderate additions of variable generation resources.
Several tools and methods are being developed in this area, including the following:

Modeling tools to forecast renewable energy output on a near-real-time frame and up to 24


hours, with confidence levels for grid operators to have advanced notice to commit adequate
capacity, plan and schedule to meet energy requirements. Benefits include minimizing
unscheduled power flows and real-time voltage and dynamic stability problems, and
preventing blackouts. Research scope should include the use of time-dependent optimal
power flow techniques to develop adaptive strategies for real-time operation.

Methods to evaluate the capacity value of an aggregation of variable generation, energy


storage, and demand response providers, which allows for coordinated dispatch and
scheduling. This would involve energy markets to help balance the economics and risks in
power system operations. The probabilistic nature of these resources will need to be taken
into account when considering reserve provision and dispatch to meet demand.

Planners will need to be able to ensure sufficient flexibility on the grid to deal with the
increased variability and uncertainty. This needs to recognize contributions from distributed
resources such as storage, electric vehicles and demand response. Additionally, planners
must consider more than just the traditional peak load or snapshot hours to ensure that hourto-hour variability can be managed.

Visualization for grid operation needs to be extended to cover intermittent generation in a


geographical display, together with visualization of load and demand response capabilities
and the effect they have on the transmission loadings, voltages and frequencies of the power
grid. This concept, developed by EPRI, will provide greater wide-area situational awareness
to operators. Control room visualization may be a key technology for increasing intermittent
generation without significant cost or loss of grid reliability. R&D is needed to develop
automation concepts on adaptive relay protection and voltage instability load shedding to
handle the integration of substantial amounts of intermittent generation.

Bulk system coordination of photovoltaic for market and bulk system control: Dispatch
center control of distributed renewable generation on the transmission system will be needed.
This will allow renewable generation to participate and be aggregated into energy markets as
well as allow for control to preserve system stability, power quality, and reliability at the
bulk level.

12-36

Grid Integration Challenges and Technologies

12.5 Integration Technologies and Strategies for Distribution-Connected


Renewable Generation
With large amounts of distributed connected generation, new strategies to manage power on the
distribution level, as opposed to transmission level, will be needed. The distribution system will
need to evolve to accommodate two-way power flows. In the future, a more automated
distribution system will interact with distributed energy resources, including photovoltaic
generation and battery systems. This will, in turn, enable better utilization of these resources and
higher penetration. As smart grid concepts are implemented on the distribution system, it will
become increasingly feasible to integrate distributed PV into grid operations. This will need large
amounts of time-synchronized data to be fed back to a control room. Better visibility of
distributed PV from the control room, coupled with PV inverter technology that can be
responsive to system needs via operator commands and price signals, would enable distributed
PV to participate in energy and capacity markets and help support system reliability. Emerging
communication and control technologies make it feasible for PV to be aggregated into virtual
power plants. Combining solar with demand response creates a compelling reliability product,
but this concept is in its infancy. All these changes will add complexity to the operations of the
distribution system. The changes in the distribution grid for high-penetration PV will likely
include the following:

Interactive voltage regulation and VAR management: Utility voltage regulator and
capacitor controls will be interactive with each other and photovoltaic sources. The central
controller shown in Figure 12-16 will help manage the interactivity to ensure optimized
voltage and reactive power conditions.

Protective relaying schemes designed for renewable generation: The distribution system
and sub-transmission will include more extensive use of directional relaying,
communication-based transfer trips, pilot signal relaying, and impedance-based faultprotection schemes (like those used in transmission). These can work more effectively with
multiple sources on the distribution system.

Advanced islanding control: Extensively automated switchgear and renewable generation


with enhanced islanding detection to improve capability for detecting unintentional islands.
Also, these systems should be able to reconfigure the grid/renewable generation into
reliability-enhancing intentional islands.

12-37

Grid Integration Challenges and Technologies

Figure 12-16
Distributed controller results aggregated to manage area power and system voltage
profiles

Interactive service restoration: Sectionalizing schemes for service restoration allow


distributed photovoltaic and other renewable generation to pick up load during the restoration
process, as shown in Fig 12-17. Once separated, these must deal effectively with overloads
from cold-load pickup and the current inrush required to recharge the system.

PV
Inverter
Source

SUBSTATION

Rotating
Machine
Source

Switch B
(3rd to close)

Switch C
(Last to close)

FEEDER

Circuit Breaker
(first to close)
Load
Area 1

Switch A
(2nd to close)
Load
Area 2

Rotating
Machine
Source

Load
Area 3

PV
Inverter
Source

Load
Area 4

Figure 12-17
Cascaded restoration of distributed generators

Improved grounding compatibility: Consider new devices and architectures in both


renewable generation and distribution that address grounding incompatibilities between
power system sensing, protection, and harmonic flows. Examples of such techniques are the
following:
- Control or limit ground fault overvoltage via relaying techniques or ancillary devices
instead of effectively grounded renewable generation requirements.

12-38

Grid Integration Challenges and Technologies

- Harden the power system and loads to be less susceptible to ground fault overvoltage
(increase voltage withstand ratings).
- Change protective relaying for ground faults so a high penetration of grounding sources
does not affect the ground fault relaying.
- Change feeder grounding scheme or load serving scheme back to a grounded three-wire
system.

Distributed energy storage: Energy storage of various forms will apply to correct temporary
load/generation mismatches, regulate frequency, mitigate flicker, and assist advance islanding
functions and service restoration. Table 12-6 shows the costs of distribution scale storage.
Table 12-6
Energy storage characteristics by application (kilowatt-scale)
Technology
Option

Maturity

Capacity
(kWh)

Power
(kW)

Duration
(hrs)

% Efficiency
(total cycles)

Total Cost
($/kW)

Cost
($/kW-h)

Energy Storage for Distributed (DESS) Applications


Advanced
Lead-Acid

Demo100250
Commercial

2550

25

8590
(4500)

16003725

400950

Zn/Br Flow

Demo

100

50

60
(>10000)

14503900

7251950

Li-Ion

Demo

2550

2550

14

8093
(5000)

28005600

9503600

8590
(1500-5000)

45205600

2260

Energy Storage for Residential Energy Management Applications*


Demo10
Commercial
20

Zn/Br Flow

Demo

930

315

24

6064
(>5000)

20006300

7851575

Li-Ion

Demo

740

110

17

7592
(5000)

1250
11,000

8002250

Lead-Acid

2
4

1400

Refer to the full EPRI report for important key assumptions and explanations behind these estimates. All
systems are modular and can be configured in both smaller and larger sized not represented. Figures are
estimated ranges for the total capital installed cost estimates of current systems based on 2010 inputs
from vendors and system integrators. Included are the costs of power electronics if applicable, all costs
for installation, step-up transformer, and grid interconnection to utility standards. Smart-grid
communication and controls are also assumed to be included. For batteries, values are reported at rated
conditions based on reported depth of discharge. Costs include process and project contingency
depending on technical maturity. The cost in $/kW-h is calculated by dividing the total cost by the hours of
storage duration.
For CAES and Pumped Hydro, larger and smaller systems are possible. For belowground CAES the heat
rate may range from ~3845-3860 Btu/kWh and the energy ratio is 0.68-0.78; for aboveground CAES the
heat rate is ~4000 Btu/ kWh and the energy ratio is ~1.0.
For C&I and residential applications lower CAPEX costs may be possible if the battery system is
integrated and installed with a photovoltaic system.
First-of-a-kind system costs will be higher than shown. Future system costs may be lower than shown
after early demonstrations are proven and products become standardized.

12-39

Grid Integration Challenges and Technologies

These system changes and technology upgrades represent an extensive investment on the part of
electric utilities, rate payers, and equipment manufacturers, and a huge change in the way the
power system is designed and operated. These changes will not come overnight and will require
many decades to implement, as well as considerable engineering planning and development to
determine the balance of features and capabilities needed against the cost and complexity of
implementation. Nonetheless, these are the approaches needed to move to high-penetration PV,
and the industry needs to begin work now on research and development so that the technologies,
tools, and approaches will be available in a timely manner.
Another configuration on the horizon is likely to coordinate multiple generators at dispersed
locations and with a variety of energy resources, including various combinations of solar, wind,
other generators and energy-storage devices. Microgrids may be applied in a broad range of sizes
and configurations. Figure 12-18 shows examples of possible microgrid subsets that could be
developed on a typical radial distribution system. These microgrid subsets include a single
customer, a group of customers, an entire feeder, or a complete substation with multiple feeders.
A very large substation could have up to 100 MW of capacity, eight or more feeders, and could
serve more than 10,000 customers.

Figure 12-18
Concept of distribution microgrids of various sizes and levels allowing reliability islands
and grid tie operation

Challenges with microgrids are many. Regardless of their size, they must take on key control
responsibilities while operating in the islanded state; otherwise, serious damage may result. At
the same time, these distributed generators must not adversely impact reliability, voltage
regulation, or power quality on the bulk power system while the microgrid is interconnected.

12-40

Grid Integration Challenges and Technologies

Advanced inverter/controllers and energy management will need to be sufficiently sophisticated


to interface with emerging smart grid technology. As such, they must be capable of supporting
communication protocols utilized by current energy management and utility distribution level
communication systems. This shift from todays central control system to a future intelligent
control is illustrated in Figure 12-19.

Figure 12-19
Distributed controller must be integrated with overall distribution control system to
maximize system value and reduce capacity requirements

The master controller is the key to providing a future sophisticated microgrid operation that
maximizes efficiency, quality, and reliability. Although some of the capabilities identified for an
intelligent microgrid master controller are currently being researched, other capabilities do not
yet exist. The Galvin Electricity Initiative documented the functional requirements for master
controller software in the report Master Controller Requirements Specification for Perfect Power
Systems (02162007 Rev. 3, 2007), which is available from the Galvin Electricity Initiatives
website at www.galvinpower.org.
In 2008, the German Association of Energy and Water Industries (BDEW) introduced new grid
codes for connecting power plants to the medium-voltage grid. Generating plants connected to
the medium-voltage networks have to support grid stability and must not disconnect from the
grid during a fault, as is common practice today in many countries. With the increasing
penetration of distributed generation (DG) sources in the medium-voltage grid, it is necessary
that these plants include both dynamic and static grid-support capabilities.

12-41

Grid Integration Challenges and Technologies

Due to high penetration of DG plants in the medium-voltage grid, new medium voltage grid code
in Germany requires that these plants provide voltage stability during voltage drop in the grid,
often referred to as fault ride-through (FRT). Figure 12-20 illustrates one ride-though concept
proposed in this grid code. PV systems connected to medium voltage grid need to do the
following:

Stay connected during a voltage drop down to 0% UC with duration up to 150 ms.

Offer stable operation above Limit 1.

Ride through the low voltages within Limits 1 and 2.

Support the grid by providing additional reactive current during a symmetrical fault.

In case of an unsymmetrical fault, the reactive current must not cause voltages increase
above 1.1 UC in the non-faulty phases.

Figure 12-20
FRT limiting curves proposed in the new German grid codes for connecting PV systems to
the medium-voltage power grid

According to this medium-voltage grid code, a plant must be able to curtail active power output
in steps of 10% (or smaller) of the agreed rated output power. Above a system frequency of 50.2
Hz, all generating units have to reduce their power output with a gradient of 40%/Hz of the
instantaneously available power. Above 51.5 Hz and below 47.5 Hz, the plant has to disconnect
from the grid. In response to the following cases, the network operator can temporarily limit the
feed-in power or disconnect the plant:

Risk of unsafe system operation

Risk of bottlenecks and congestion in the network

12-42

Grid Integration Challenges and Technologies

Risk of unwanted islanding

Risk of static or dynamic grid instability

Risk of system instability due to frequency increase

Carry out repairs or construction

In the context of production management, feed-in management, and network security


management

This new German grid code also requires DG plants connected to the medium-voltage grid to
provide static grid support by providing reactive power. The reactive power set point can be
either fixed or adjustable by a signal from the network operator, as follows:

A fixed displacement factor

A variable displacement factor depending on the active power

A fixed reactive power value in MVR

A variable reactive power depending on the voltage

12.6 References
12.6.1 In-Text Citations
1. Wind Report 2004: E.ON Netz GmbH, Bayreuth, Germany. 2004, updated by German
Renewable Energy Sector Shows Impressive Growth, Greenprices, January 24, 2007.
2. Study of Electric Transmission in Conjunction with Energy Storage Technology: Lower
Colorado River Authority, Austin, TX, Ridge Energy Storage & Grid Services, Austin, TX,
RnR Engineering, Austin, TX, and Walter J. Reid Consulting, Austin, TX, August 2003.
3. Giovanni, D. Wind Energy in the Big Island: Hawaii Electric Light Company, Hilo, HI.
Presented at USDOE/DBEDT Forum, April 2002.
4. Brooks, E.D.L., Smith, J., Pease, J., and McGree, M. Assessing the Impact of Wind
Generation on System Operations at Xcel Energy-North and Bonneville Power
Administration. Windpower 2002. AWEA Washington, D.C.
5. Brooks, D., Lo, E., Zavadil, R., Santoso, S., and Smith, J. Characterizing the Impact of
Significant Wind Generation Facilities on Bulk Power System Operations Planning.
UWIG, May 2003.
6. Dragoon, K. and Milligan, M. Assessing Wind Integration Costs with Dispatch Models: A
Case Study. Windpower 2003, Austin, TX.
7. Hirst, E. Integrating Wind Energy with the BPA Power System: Preliminary Study.
Prepared for BPA Power Business Line, September 2002.
8. Hirst, E. Integrating Wind Output with Bulk Power Operations and Wholesale Electricity
Markets. Wind Energy 5(1) 2002, pp. 1936.
9. Electrotek Concepts, Inc. Systems Operations Impacts of Wind Generation Integration
Study. Prepared for We Energies, June 2003.
12-43

Grid Integration Challenges and Technologies

10. United States Patent #5,083,039, Variable Speed Wind Turbine, Richardson et al., issued
January 21, 1992.
11. Asplund, G. et al. DC Transmission Based on Voltage Source Converters. CIGRE Session
1998 paper No 14-302.
12. Grainger, W. and Jenkins, N. Offshore Wind Farm Electrical Connection Options. British
Wind Energy Association, 2002.
13. Torbjrn Thiringer et al. Grid Disturbance Response of Wind Turbines Equipped With
Induction Generator and Doubly-Fed Induction Generator. Chalmers University of
Technology, IEEE T&D Dallas, September 2003.
14. IEEE Draft Trial-Use Guide for Application of Power Electronics for Power Quality
Improvement on Distribution Systems Rated 1 kV through 38 kV, P1409, D10,
September 5, 2003.
15. Sannino, A., Sversson, J., and Larsson, T. Power-Electronic Solutions to Power Quality
Problems. Electric Power Systems Research 66 (2003) 71-82, Department of Electric Power
Engineering, Chalmers Univ. of Technology, Gothenburg, Sweden.
16. Short-Term Wind Generation Forecasting Using Artificial Neural Networks. EPRI, Palo
Alto, CA: 2003. 1009219.
17. California Wind Energy Forecasting System Development and Testing Phase 1: Initial
Testing. California Energy Commission, Sacramento CA; EPRI, Palo Alto, CA: 2002.
1003778.
18. California Wind Energy Forecasting System Development and Testing Phase 2: 12-Month
Testing. California Energy Commission, Sacramento CA, EPRI, Palo Alto, CA: 2003.
1003779.
19. Texas Wind Energy Forecasting System Development and Testing Phase 1: Initial Testing.
EPRI, Palo Alto, CA; U.S. Department of Energy, Washington DC: 2003. 1008032.
20. Texas Wind Energy Forecasting System Development and Testing Phase 2: 12-Month
Testing. EPRI, Palo Alto, CA; U.S. Department of Energy, Washington DC: 2004. 1008033.
21. Wind Energy Forecasting Applications in Texas and California. EPRI, Palo Alto, CA: 2003.
1004038.
22. California Regional Wind Energy Forecasting System Development, Volume 1: Executive
Summary. EPRI, Palo Alto, CA; California Energy Commission, Sacramento, CA: 2006.
1013262.
23. California Regional Wind Energy Forecasting System Development, Volume 2: Wind Energy
Forecasting System Development and Testing and Numerical Modeling of Wind Flow over
Complex Terrain. EPRI, Palo Alto, CA; California Energy Commission, Sacramento, CA:
2006. 1013263.
24. California Regional Wind Energy Forecasting System Development, Volume 3: Wind Tunnel
Modeling of Wind Flow over Complex Terrain. EPRI, Palo Alto, CA; California Energy
Commission, Sacramento, CA: 2006. 1013264.

12-44

Grid Integration Challenges and Technologies

25. California Regional Wind Energy Forecasting System Development, Volume 4: California
Wind Generation Research Dataset (CARD). EPRI Palo Alto, CA, California Energy
Commission, Sacramento, CA: 2006. 1013265.
26. Zack, J.W. Wind Forecasting: State of the Art. CanWEA Wind and Power Systems
Seminar, May 20, 2009, Montreal, Quebec.
27. Cruden, A. and Dudgeon, G.J.W. Opportunities for Energy Storage Devices Operating with
Renewable Energy Systems. EESAT 2000, September 1820, Orlando, FL.
28. Strbac, G., Bathurst, G.N., and Jenkins, N. Integration of Renewable Generation into the UK
MarketOpportunities for Energy Storage. EESAT 2000, September 1820, Orlando, FL.
29. Taylor, R.E., and Hoagland, J.J. Using Energy Storage with Wind Energy for Arbitrage,
EESAT 2002, April 1517, 2002, San Francisco.
30. Taylor, R., Hoagland, J., and Bradshaw, D. Energy Storage for Ancillary Services, EESAT
2002, April 1517, 2002, San Francisco.
31. Key, T., and Kamath, H. Electric Energy Storage: Will it Help Enable DE? Primen
Perspective on Distributed Energy, Primen, DE-PP-23, August 2003.
12.6.2 General
Wind Power Integration Assessment and Case Studies. EPRI, Palo Alto, CA: 2004. 1004806.
EPRI-DOE Handbook of Energy Storage for Transmission and Distribution Applications).
EPRI, Palo Alto, CA; U.S. Department of Energy, Washington, D.C.: 2003. 1001834.
Norris, B., EPRI Energy Storage Handbook: Wind Energy Storage Applications, Gridwise
Engineering Co., Danville, CA: 2003.
Technical Assessment GuideStorage. EPRI, Palo Alto, CA: 2000. 1001203.
12.6.3 Renewable Generation Grid Impacts and Integration
Characterizing the Impacts of Significant Wind Generation Facilities on Bulk Power System
Operations Planning: Utility Wind Interest GroupXcel Energy-North Case Study, Utility Wind
Interest Group, Reston. VA, EPRI, Palo Alto, CA: 2003. 1004807.
Wind Report 2004: E.ON Netz GmbH, Bayreuth, Germany. 2004.
Eastern Wind Integration and Transmission Study, NREL, 2009.
Western Wind and Solar Integration Study, NREL, 2010ew.
IEA Wind Task 25, Design and Operation of Power Systems with Large Amounts of Wind
Power: Final Report, Phase One 20062008. VTT Research Notes 2493, 2009.

12-45

13

GREENHOUSE GAS EMISSIONS CONTROL

13.1 Greenhouse Gases


Several greenhouse gases are known to contribute to humanitys effect on the radiation balance
in the atmosphere and, hence, on global temperature and potential climate change. The
greenhouse gases of concern include carbon dioxide (CO2), methane (CH4), nitrous oxide (N2O),
and certain chlorofluorocarbons (CFC) that react with and deplete the ozone layer. The estimated
lifetimes of these gases in the atmosphere before they oxidize to carbon dioxide and their
infrared absorbing strengths vary. Table 13-1 compares those characteristics relative to carbon
dioxide as a function of time.
Table 13-1
Lifetimes in the atmosphere and relative infrared absorption strengths of the greenhouse
gases
Gas

Lifetime in the
Atmosphere

Infrared Absorbing Strength Per Unit Mass vs. CO2


20-Year

50-Year

100-Year

Carbon Dioxide (CO2)

Variable

Methane (CH4)

12 (3) years

56

21

Nitrous Oxide (N2O)

120 years

280

310

170

Chlorofluorocarbons
(CFCs)

Not given

4,900

3,800

Not given

Source: U.S. DOE EIA, Emissions of Greenhouse Gases in the U.S. 1996, October 1997.

Over a 50-year timeframe, methane absorbs 21 times as much infrared radiation as an equal mass
of carbon dioxide. Thus, we say that the infrared absorbing equivalent of a ton of methane
released by a landfill is 21 tons of carbon dioxide. Similarly, the infrared absorbing equivalents
of a ton of nitrous oxide and a ton of chlorofluorocarbons are 310 tons and 3,800 tons of carbon
dioxide, respectively.

13.2 Role of Renewable Energy Technology


Renewable generation technologies typically produce no direct emissions of fossil fuel carbon
dioxide, methane, or other greenhouse gases, and can be used to offset greenhouse gases emitted
by the fossil-fuel-fired component of the system generation mix. Consequently, renewable
energy can be considered to be a greenhouse gas emissions-reduction technology as well as a
renewable energy power technology. In addition, should greenhouse gas emission-reduction
mandates be enacted in the future, renewables would become an important component of a
greenhouse gas emissions-reduction strategy and would likely play a key role in carbon dioxide
emissions trading.
13-1

Greenhouse Gas Emissions Control

The carbon dioxide emissions-reduction potential of a renewable energy power plant is a


function of the fuel mix of the existing generation system. The effective carbon dioxide
emissions-reduction cost is a function the CO2 emission rate and the difference between the
average generation costs of the base system and those of the renewable energy power plant.

13.3 Fossil Carbon Intensity of Fuels


The carbon intensity of generation technologies is measured by the carbon emissions released by
burning a fuel to generate a megawatt-hour of electricity. It depends on the carbon and heat
contents of the fuel as well as the net thermal efficiency of the power generation technology.
Table 13-2 compares fuel heat and carbon content with fossil carbon intensity for coal, oil, and
natural gas [1, 2]. Dry wood is included in the table to highlight the zero carbon intensity of
sustainably grown biomass fuel. The data are presented in units of mass of carbon released per
unit of fuel in both mass and higher heating value (HHV) heat content units. Sub-bituminous
coal has the highest carbon intensity per unit energy content (27.77 kg carbon/GJ) and natural
gas has the lowest fossil carbon intensity (13.8 kg carbon/GJ).
Table 13-2
Fossil carbon intensity of coal, oil, natural gas, and wood fuels
Fuel Type

Heat ContentHHV

Carbon Content

Fossil Carbon Intensity

(Btu/lb)

(MJ/kg)

(lb-C/lb)

(kg-C/kg) (lb-C/MBtu)

(kg-C/GJ)

Coal: Bituminous

13,700

31.798

0.782

0.782

57.08

24.59

Coal: Sub-bituminous

9,000

20.889

0.580

0.580

64.44

27.77

Oil

20,000

46.420

0.92

0.92

46.0

19.8

Natural Gas

23,400

54.311

0.75

0.75

32.1

13.8

Wood (dry)

8,000

18.568

0.45

0.45

0*

0*

*Note: Fossil carbon intensity is the measure relevant to greenhouse gas. By this measure, wood from
renewably grown trees has zero carbon intensity. If the carbon content of the fuel were put into the same
formula used for fossil fuels, the carbon intensity of wood is 54.2 lb-C/MBtu (23.4 kg-C/GJ).

Table 13-3 compares the impacts of fuel carbon content and net heat rate to the fossil carbon
intensity of selected coal- and natural-gas-fired generation technologies. The data are presented
in units of mass of CO2 and carbon (C) released per megawatt-hour of electricity generated.
Existing coal-fired power plants operating at 34.1% thermal efficiency (HHV) exhibit the highest
generation carbon intensity per megawatt-hour generated (0.26 metric ton C/MWh) while
advanced, natural-gas fuel cell plants operating at 63.7% net thermal efficiency exhibit the
lowest generation carbon intensity (0.08 metric ton C/MWh) (note: one metric ton = one tonne).

13-2

Greenhouse Gas Emissions Control


Table 13-3
Carbon intensity of generation technologies
FuelTechnology
(Net HHV Efficiency)
English Units

Carbon Content
(lb/MBtu)

Net Heat Rate


(Btu/kWh)

Typical existing (0.341)

56.7

Pulverized, 95% scrubbed


(0.376)
Advanced, IGCC (0.467)

Fossil Carbon Emission


CO2
(ton/MWh)

C
(ton/MWh)

10,000

1.04

0.28

56.7

9,087

0.94

0.26

56.7

7,308

0.76

0.21

Existing Steam Plant


(0.331)

32.1

10,300

0.61

0.17

Advanced, CC (0.538)

32.1

6,350

0.37

0.10

Advanced, CT (0.427)

32.1

8,000

0.47

0.13

Advanced, fuel cell (0.637)

32.1

5,361

0.32

0.09

SI Units

Carbon Content
(kg/GJ)

Net Heat
Rate (kJ/kWh)

CO2
(tonne/MWh)

C
(tonne/MWh)

Typical existing (0.341)

24.43

10,550

0.95

0.26

Pulverized, 95% scrubbed


(0.376)

24.43

9,587

0.86

0.23

Advanced, IGCC (0.467)

24.43

7,710

0.69

0.19

Existing Steam Plant


(0.331)

13.83

10,867

0.55

0.15

Advanced, CC (0.538)

13.83

6,699

0.34

0.09

Advanced, CT (0.427)

13.83

8,440

0.43

0.12

Advanced, fuel cell (0.637)

13.83

5,656

0.29

0.08

Coal

Natural Gas

Coal

Natural Gas

13.4 CO2 Emissions Offsets


13.4.1 Factors Affecting CO2 Emissions Offset
Calculation of the CO2 emissions offset resulting from adding wind, solar, and other renewable
generation to a utility system at first appears to be straightforward and uncomplicated. However,
several indirect factors complicate the calculation and typically result in actual net CO2
emissions offsets that are lower than those estimated using the straightforward levelized method
presented below. These factors include the following:

Renewable generation penetration into the generation mix

Renewable generation duty cycle characteristics

Hydro, pumped storage, and peaking capacity in the existing generation mix
13-3

Greenhouse Gas Emissions Control

Impacts on the dispatch and annual capacity factor of each component of the generating mix

Possible increased cycling of baseload coal and gas units

Increased spinning reserves for load following and regulation

The impacts of these factors vary with the renewable generation type and grid characteristics and
tend to be greatest for wind, which is a variable generation resource and not dispatchable, and
lowest for biomass, geothermal, and hydro, which are typically dispatchable. The impacts for
solar photovoltaic and thermal generation are intermediate between those of wind and the other
renewables, as discussed further in the following sections.
13.4.1.1 Wind
In most cases, the daily operating cycle of wind generation does not follow the daily load
demand cycle. For example, peak wind generation may occur at night during the winter and in
the afternoon during the summer. In addition, wind generation can ramp up or down quickly as
weather fronts move by, creating the need to rapidly increase or decrease other generation
connected to the grid in order to follow load. When the penetration of wind generation in the mix
is a few percent, the impacts are usually minimal and the electricity grid can easily absorb the
wind generation. However, if wind penetration rises above 10% to 15%, especially during low
load periods at night, the impacts become more significant. In the latter case, addition of wind
generation usually requires an increase in spinning reserves for load following and regulation,
increased cycling of base load coal units, and possible curtailment of wind generation at night.
Each of these factors decreases the net CO2 emissions offset resulting from adding wind to the
generating mix. In fact, the net emissions offset can even be negative, e.g. the CO2 emissions can
actually increase. For more information, consult recent references that address wind integration
impacts (UWIG-EPRI [3], White [4], and E.ON Netz GMbH [5]).
13.4.1.2 Solar PV and Thermal
Solar PV and thermal generation are also variable resources, but their operating cycles are
predictable and more synchronized with the daily load demand cycle. In addition, solar thermal
generation is usually configured to include backup gas-firing heating or thermal energy storage
to allow the plant to operate during off-peak hours. Thus, with backup gas firing, their actual net
CO2 emissions offsets would be somewhat lower than the levelized emissions offsets presented
later.
13.4.1.3 Biomass, Geothermal, and Hydro
Biomass, geothermal, and hydro units are typically dispatchable and their CO2 emissions offsets
should be close to the levelized offsets presented later.
13.4.1.4 Planting Trees and Other Fast-Growing Crops
In addition to the CO2 emissions offsets of renewable generation technologies, plantations of
fast-growing trees and other biomass crops such as eucalyptus, hybrid poplar, and switchgrass
act as carbon sinks that absorb CO2 from the atmosphere via photosynthesis.

13-4

Greenhouse Gas Emissions Control

For example, as discussed in Reference 8, the annual growth of a hybrid poplar plantation in the
Pacific Northwest absorbs 5 to 8 tons of carbon per acre or 18.3 to 29.3 ton CO2/acre per year
(11.2 to 18.0 metric tons C/hectare per year; one hectare = 10,000 square meters).
As shown in Table 13-3, a coal-fired power plant generates up to 1.04 tons of CO2/MWh (0.95
metric ton carbon/MWh). Therefore, an acre of hybrid poplar would absorb the CO2 produced by
the generation of 17.6 to 28.2 MWh from coal.
A 500-MW coal-fired power plant operating at 85% capacity factor generates 3,723,000
MWh/year and 3,872,000 tons CO2/year. In order to absorb the entire CO2 volume generated by
the plant each year, the required hybrid poplar plantation would cover 132,000 to 211,000 acres,
or 206 to 330 square miles.
13.4.2 Levelized CO2 Emissions Offsets
This section presents charts and other information to estimate the CO2 emissions offsets resulting
from adding renewable generation to an electricity system. The estimates assume that (1) each
MWh of electricity generated by renewable resources replaces 1 MWh of electricity that would
otherwise be generated by the existing generation mix; and (2) the relative contribution of each
component of the mix to generation and CO2 emissions does not change.
Figure 13-1 plots the fossil CO2 emission offset (metric ton CO2/MWh) for coal, oil, and natural
gas generation as a function of the percentage of total kilowatt-hours of generation contributed
by the respective fossil fuel sources. The chart shows separate curves for 100% fossil fuel
generation by coal, oil, and natural gas. These curves provide representative data for illustration
only and can vary greatly depending on the specific generation technology involved (see Table
13-3). The levelized CO2 emission offset for a specific blend of coal, oil, and natural gas
generation is estimated by either interpolating a value between the three curves or by calculating
a weighted average of the values determined by the coal, oil, and gas curves on the chart.
Figure 13-2 shows the levelized CO2 emissions control cost ($/metric ton CO2) as a function of
the cost premium of renewable power relative to the existing system generation cost and the CO2
emissions offset. Figure 13-3 is similar except that it shows the CO2 emission control cost as a
function of the fossil fuel mix and assumes that 50% of the generation is contributed by coal, oil,
and natural-gas-fired technologies.

13-5

Greenhouse Gas Emissions Control

CO2 Emissions Offset vs.% Fossil Fuel Generation


1.0
0.9

Metric Ton CO2/MWh

0.8
0.7
100% Coal

0.6
0.5

100% Fuel Oil

0.4
100% Natural Gas

0.3
0.2
0.1
0.0
0%

10%

20%

30%

40%

50%

60%

70%

80%

90%

100%

Fossil Fuel Generation, % of Total MWh


Figure 13-1
Levelized CO2 emissions offset vs. % fossil fuel generation

CO2 Emissions Control Cost vs. Cost Premium of


Renewable Generation
300

CO2 Emissions Offset, Metric Ton/MWh

$/Metric Ton CO2

250

0.2

200

0.3
150

0.4

100

0.6

50

1.0

0
-50
-10

10

20

30

40

50

Cost Premium vs. System Generation Cost, $/MWh

Figure 13-2
Levelized CO2 emissions control cost vs. cost premium and CO2 emissions offset

13-6

60

Greenhouse Gas Emissions Control

CO2 Emissions Control Cost vs. Cost Premium


50% Fossil Fuel Generation
250

$/Metric Ton CO2

200
150

100% Natural Gas

100% Fuel Oil

100
100% Coal

50
0
-50
-100
-10

10

20

30

40

50

60

Cost Premium vs. System Generation Cost, $/MWh


Figure 13-3
Levelized CO2 emissions control cost vs. cost premium and fossil fuel mix at 50%
generation from fossil fuels

To illustrate the application of the charts, if a fossil generation mix is 70% coal and 30% natural
gas, and fossil fuels supply 90% of the kilowatt-hours generated, Figure 13-1 indicates that the
CO2 emissions factors for coal and natural gas (at 90%) are about 0.85 and 0.46 metric ton
CO2/MWh, respectively. Thus the weighted CO2 emissions offset factor is (0.85 70%) + (0.46
30%) = 0.74 metric ton CO2/MWh. If the cost premium is $10/MWh, then the CO2 emissions
control cost is $10/MWh / 0.74 metric ton/MWh = $13.60/metric ton CO2.

13.5 Factors Affecting Generation Cost of Renewable Technologies


Renewable energy generation costs are sensitive to various factors. The following section
discusses the factors that affect the generation cost and, in turn, the levelized carbon dioxide
reduction cost of using a given renewable energy technology. All levelized costs are in constant
end-of-year or third-quarter 2010 U.S. dollars (zero inflation rate). The federal production tax
credit and investment tax credit are not considered in this analysis.
13.5.1 Wind
Wind generation cost is most sensitive to the annual average wind speed at the site. As wind
speed increases, wind energy generation increases, and thus the unit cost of energy ($/MWh)
decreases. Table 13-4 presents wind plant performance and cost data vs. location. The source of
the data is the 2009 EPRI report, Engineering and Economic Evaluation of Utility-Scale Wind
Power Plants [6].

13-7

Greenhouse Gas Emissions Control


Table 13-4
Wind plant performance and cost vs. location (4th Quarter 2010$)
Source: EPRI [6]
Base
Case

California

Texas

Michigan

New York

New York
Offshore

100

100

200

150

50

202

4Q 2010

4Q 2010

4Q 2010

4Q 2010

4Q 2010

4Q 2010

Total Capital
Requirement $/kW)

$2,159

$2,663

$2,101

$2,345

$3,277

$4,240

Operation &
Maintenance Cost
($/kW-yr)

$31.8

$43.0

$31.6

$37.5

$40.0

$131.3

Annual Capacity
Factor

35%

33%

31%

28%

29%

45%

Levelized Cost of
Electricity ($/MWh)

$58.1

$77.9

$64.6

$81.0

$101.7

$110.2

Location
Rated Capacity (MW)
Reference Date
of Costs

The levelized wind energy cost ranges between about $58/MWh at the 100-MW base case plant
and about $110/MWh at the 202-MW New York offshore wind plant.
Figure 13-4 presents the resulting levelized CO2 emissions control cost vs. system generation
cost at the 100-MW California wind project for five CO2 emissions offset factors between 0.2
and 1.0 metric ton CO2/MWh. The analysis assumes the CO2 emissions offset factor is 0.755
metric ton CO2/MWh, which is consistent with the advanced IGCC coal plant example in Table
13-3 and a representative mix of coal and cleaner natural gas combustion. The levelized CO2
emissions control cost decreases as both the CO2 emissions offset factor and system generation
cost increase, and it becomes negative for system generation costs higher than the $78/MWh
wind generation cost. Negative values of the levelized CO2 emissions control cost correspond to
a net credit for reducing CO2 emissions.
Figure 13-5 presents the levelized CO2 emissions control cost vs. system generation costs for
each of the six wind plant cases described in Table 13-4. The CO2 control cost increases as the
wind generation cost increases and is highest for the 202-MW New York offshore wind plant
and lowest for the 100-MW base case plant.

13-8

Greenhouse Gas Emissions Control

CO2 Emissions Control Cost - Wind Power


100-MW California Project, 76 $/MWh
4th Quarter 2010 $

$/Metric Ton CO2

400
300
200
100

0.2

Emissions Offset (Metric Ton CO2/MWh)

0.3
0.4
0.6
1.0

0
-100
-200
-300
0

20

40

60

80

100

120

140

160

System Generation Cost, $/MWh


Figure 13-4
Levelized CO2 emissions control cost at 100-MW California wind plant vs. CO2 emissions
offset and system generation cost (4th Quarter 2010$)

CO2 Emissions Control Cost - Wind Power

$/Metric Ton CO2

Emissions Offset: 0.755 Metric Ton CO2/MWh

4th Quarter 2010 $

200
150
100
50
0
-50
-100
-150
-200
-250

NY Offshore 202 MW
NY Onshore 50 MW
MI Onshore 150 MW

CA Onshore 100 MW
TX Onshore 200 MW
Base Case Onshore 100 MW

25

50

75

100

125

150

175

200

System Generation Cost, $/MWh


Figure 13-5
Levelized CO2 emissions control cost vs. location and system generation cost (4th Quarter
2010$)

13-9

Greenhouse Gas Emissions Control

13.5.2 Biomass
Biomass generation cost is most sensitive to the delivered biomass fuel cost, the installed cost of
the biomass power system, and the net thermal efficiency of the power cycle. The installed cost
is expected to remain relatively stable over time, but the delivered biomass fuel cost may decline
as new crop planting, cultivation, harvesting, and handling technologies improve. The net
thermal efficiency of the power cycle has the potential to improve as well.
Table 13-5 summarizes the total capital requirement, fixed and variable O&M, plant
performance, and levelized cost of electricity for a 60-MW, 100% biomass repowered , a 250MW biomass-cofired plant (at 10% heat input from biomass), and a 50-MW biomass-fired
bubbling fluidized bed boiler power plant.
Figure 13-6 shows the levelized CO2 emissions control cost for the three cases as a function of
the system generation cost. The source of the data is the 2010 EPRI report, Engineering and
Economic Evaluation of Biomass Power Plants [7]. Assuming that the system generation cost is
$50/MWh, the levelized CO2 cost ranges between minus $25/metric ton CO2 for the 10%
biomass-cofired plant and plus $101/metric ton CO2 for the 100% biomass-fired bubbling
fluidized bed plant. The cofiring cost is based on the 25-MW contribution of the biomass fuel to
the 250-MW output of the coal-fired plant and incremental cost of the biomass fuel vs. coal. A
negative CO2 control cost corresponds to a net credit for reducing CO2 emissions.
Table 13-5
Levelized cost of electricity for 50-MW biomass-fired stoker and fluidized bed boiler power
plants ($/MWh, 3rd Quarter 2010$)
Source: EPRI [7]
100%
Biomass
Repowering

10% Biomass
Cofiring

100% Biomass
Bubbling FBC

1,970

1,690

5,590

Fixed OM

132

21.6

113

Variable O&M

3.5

4.9

5.8

Btu/kWh

12,600

9,760

13,964

GJ/kWh

13,284

10,290

14,723

Fuel $/GJ

3.37

0.24

3.37

Capacity Factor

85%

85%

85%

Capital

22

20

56

Fixed O&M

18

15

Variable O&M

Fuel

45

50

TOTAL

88

31

126

Total Capital Requirement

Levelized COE $/MWh

13-10

Greenhouse Gas Emissions Control

CO2 Emissions Control Cost - Biomass Power


Biomass Fuel Cost for Repowering and FBC: 3.74 $/GJ
Incremental Fuel Cost for Cofiring: + 0.26 $/GJ
CO2 Emissions Offset: 0.755 metric ton CO2/MWh

200
100% Biomass Bubbling FBC (50 MW)

$/Metric Ton CO2

150

100% Biomass Repowering (60 MW)

100
50
0
-50

10% Biomass Cofiring (250/25 MW)

-100
-150
0

20

40

60

80

100

120

140

System Generation Cost, $/MWh


Figure 13-6
Levelized CO2 emissions control cost vs. system generation cost for 100% biomassrepowered and bubbling fluidized bed plants and 10% biomass co-fired coal plant
(incremental cost vs. coal for co-fired plant, 3rd Quarter 2010$)
Source: EPRI [7]

13.5.3 Solar Photovoltaics


Photovoltaic (PV) generation cost is most sensitive to the annual insolation rate at the site and
the installed cost of the PV system. Installed cost is expected to continue to decline over time as
new technology is developed and commercialized. Table 13-6 presents the 20-year levelized cost
of electricity for 10-MW central station solar photovoltaic power plants at each combination of
four locations and three PV technologies. The locations include Las Vegas, Nevada; Alamosa,
Colorado; Jacksonville, Florida; and Columbus, Ohio. The PV technologies include fixed flatplate and single-axis tracking modules using crystalline silicon cells and two-axis tracking
concentrating PV using multijunction cells. The estimated levelized cost of electricity ranges
between $247/MWh for the fixed flat-plate CdTe system at Las Vegas and $421/MWh for the
single-axis tracking c-Si system at Columbus. The source of the data is the 2009 EPRI report,
Engineering and Economic Evaluation of Central-Station Photovoltaic Power Plants [7].
Figure 13-7 shows the expected sensitivities of the levelized CO2 emissions control costs for the
six technologies to the system generation cost at the Las Vegas site (top) and the Jacksonville
site (bottom). The CO2 control cost decreases with system generation cost. The analysis assumes
that the CO2 emissions offset factor is 0.755 tonne CO2/MWh.

13-11

Greenhouse Gas Emissions Control


Table 13-6
Performance and cost estimates for 10-MW solar photovoltaic power plants (4th Quarter 2010$)
Source: EPRI [8]
Fixed FlatPlate a-Si

Fixed FlatPlate CdTe

Fixed FlatPlate c-Si

Rated Capacity

10 MW

10 MW

10 MW

Total Capital Requirement ($/kW)


Operation and Maintenance Cost
($/kW-yr)

$3,834

$3,721

$4,027

$53

$53

$48

22.6%
19.1%
17.0%
14.6%

22.2%
20.7%
16.9%
14.6%

22.4%
21.1%
16.9%
14.4%

$249
$294
$330
$384

$247
$264
$324
$380

$258
$275
$342
$402

1-Axis
Tracking c-Si

Tilted 1-Axis
Tracking c-Si

2-Axis
Tracking CPV

Rated Capacity

10 MW

10 MW

10 MW

Total Capital Requirement ($/kW)


Operation and Maintenance Cost
($/kW-yr)

$4,557

$5,066

$4,937

$62

$62

$67

25.0%
25.3%
18.4%
15.8%

26.9%
26.8%

27.7%
27.9%

$267
$263
$362
$421

$258
$275

$260
$258

First-Year Capacity Factor


Las Vegas, Nevada
Alamosa, Colorado
Jacksonville, Florida
Columbus, Ohio
Levelized Cost of Electricity ($/MWh)
Las Vegas, Nevada
Alamosa, Colorado
Jacksonville, Florida
Columbus, Ohio

First-Year Capacity Factor


Las Vegas, Nevada
Alamosa, Colorado
Jacksonville, Florida
Columbus, Ohio
Levelized Cost of Electricity ($/MWh)
Las Vegas, Nevada
Alamosa, Colorado
Jacksonville, Florida
Columbus, Ohio

13-12

Greenhouse Gas Emissions Control

CO2 Emissions Control Cost - Solar PV


10-MW PV Plant, CO2 Emissions Offset: 0.755 metric ton/MWh
December 2010 $

$/Metric Ton CO2

600

Columbus, OH

500

1-Axis cSi

400
Fixed CdTe

300
200
100
0
0

25

50

75

100

125

150

175

200

System Generation Cost, $/MWh


Figure 13-7
Levelized CO2 emissions control cost vs. system generation cost for 50-MW solar
photovoltaic plants, Las Vegas, NV (top) and Columbus, OH (bottom) (December 2010$)
Source: EPRI [8]

13.5.4 Geothermal
Geothermal generation cost is most sensitive to the geothermal energy production and the
installed cost of the geothermal wells and power system.
Table 13-7 presents the 20-year levelized cost of electricity for 50-MW flash-steam and 50-MW
binary-cycle geothermal plants. The data are from the 2010 EPRI report, Engineering and
Economic Evaluation of Geothermal Power Plants [9].The levelized COE is $66/MWh and
$70/MWh for the flash-steam and binary-cycle geothermal plants.
Figure 13-8 shows the resulting CO2 emission control costs versus the base system generation
cost for the 50-MW flash-steam and 20-MW binary-cycle geothermal plants. The analysis
assumes the CO2 emissions offset factor is 0.717 tonne CO2/MWh, including the 95% adjustment
for the CO2 emissions released during the production of geothermal power.

13-13

Greenhouse Gas Emissions Control


Table 13-7
Levelized cost of electricity for 50-MW flash-steam and binary-cycle geothermal power
plants (December 2010$)
Source: EPRI [9]
Flash
Steam

Binary
Cycle

50

50

Capital Cost, $/kW

$4,989

$5,192

Fixed O&M, $/kW-yr

$76.4

$80.6

Variable O&M, $/kW-yr

$9.60

$10.50

Capacity Factor

96%

96%

Capital @ 8.3 %/yr

$47.7

$52.4

O&M

$9.1

$9.6

Fuel Cost $/GJ

$9.6

$7.6

TOTAL

$66.4

$69.6

Geothermal Technology
Rated Capacity, MW

Levelized COE, $/MWh

Figure 13-8 shows that, as the base system generation cost increases, the CO2 emissions control
cost decreases and becomes negative for generation cost greater than about $66/MWh for flashsteam and $70/MWh for binary-cycle geothermal plants.

13-14

Greenhouse Gas Emissions Control

CO2 Emissions Control Cost - Geothermal Power


Basis: MW Binary Cycle and Flash-Steam Plants
December 2010 $
CO2 Emissions Offset: 0.755 metric ton CO2/MW x 95%

120
100

$/Metric Ton CO2

80
60

50-MW Binary-Cycle

40
20

50-MW Flash-Steam

0
-20
-40
-60
0

20

40

60

80

100

Base System Cost, $/MW


Figure 13-8
Levelized CO2 emissions control cost for 50-MW flash-steam and binary-cycle power
plants vs. base system generation cost (December 2010$)
Source: EPRI [9]

13.5.5 Solar Thermal


Solar thermal generation cost is highly sensitive to the annual insolation rate at the site and the
installed cost of the solar thermal system. The installed cost is expected to continue to decline
over time as new technology is developed and commercialized.
Table 13-8 presents the levelized cost of electricity for four solar-thermal power technologies.
They include 250-MW parabolic trough plants with and without 6-hour thermal energy storage,
100-MW tower with 10-hour storage, and 100-MW dish/Stirling plants. The data assume 30-year
project life and constant December 2010 dollars, and do not include the federal production tax
credit or investment tax credit. The source of the data is the 2010 EPRI report, Solar Thermal
Technology Status and Performance and Cost Estimates2010 [10].

13-15

Greenhouse Gas Emissions Control


Table 13-8
Constant-dollar levelized cost of electricity for solar thermal technologies (December
2010$)
Parabolic
Trough:
No Thermal
Storage

Parabolic
Trough:
6-hr Thermal
Storage

Central
Receiver:
10-hr
Thermal
Storage

Dish/Stirling

250

250

100

100

4,070

6,160

6,510

4,540

Fixed O&M $/kW-yr

64

68

66

62

Variable O&M, $/MW-hr

Net Heat Rate, GJ/kWh

Fuel, $/GJ

25.4%

40.7%

49.1%

23.7%

Capital @ 6.68%/yr

210

199

174

250

O&M

45

30

24

47

Fuel

255

229

198

297

2 125 MW

2 125 MW

100 MW

4,000 25
kW (100 MW)

25.4%

40.7%

49.1%

23.7%

LCOE with ITC, $/MWh

116

100

86

133

FOM Portion of LCOE

35

23

19

36

Capital Charge Portion of LCOE

81

77

67

97

196

175

152

228

FOM Portion of LCOE

35

23

19

36

Capital Charge Portion of LCOE

161

152

133

191

Rated Capacity, MW
Total Capital Requirement, $/kW

Capacity Factor
Levelized COE $/MWh

TOTAL

Plant Size (no. of units unit size)


Capacity Factor

LCOE without ITC, $/MWh

Figure 13-9 presents the resulting levelized CO2 emissions control costs for the four solar
thermal power plant technologies. The analysis assumes that the CO2 emissions offset factor is
0.755 metric ton CO2/MWh. The CO2 emissions control cost decreases with increasing system
generation cost. The solar thermal power tower with 10-hour thermal energy storage has the
lowest CO2 emissions control cost and the dish/Stirling plant has the highest cost.

13-16

Greenhouse Gas Emissions Control

CO2 Emissions Control Cost - Solar Thermal Technologies


Net CO2 Emissions Offset: 0.755 metric ton CO2/MWh
Basis: December 2010 $

400

$/Metric Ton CO2

300

Dish Stirling

Parabolic Trough

200
100
Tower
10-hr Storage

Parabolic Trough
6-hr Storage

-100
0

25

50

75

100

125

150

175

200

System Generation Cost, $/MWh


Figure 13-9
Levelized CO2 emissions control cost for four solar thermal technologies (December
2010$)
Source: EPRI [10]

13.6 References
1. Greenhouse Gas Reduction with Renewables. EPRI, Palo Alto, CA: 2000. TR-113785.
2. Renewable Energy Technology Characterizations. EPRI, Palo Alto, CA: 1997. TR-109496.
3. Characterizing the Impacts of Significant Wind Generation Facilities on Bulk Power System
Operations Planning: Utility Wind Interest GroupXcel Energy-North Case Study.
EPRI, Palo Alto, CA: 2003. 1004807.
4. White, D. J., Danish Wind: Too Good to be True?, The Utilities Journal,
July 2004, pp, 37-39.
5. Wind Report 2004, E.ON Netz Gmbh, Bayreuth, Germany, 2003.
6. Engineering and Economic Evaluation of Utility-Scale Wind Power Plants. EPRI, Palo Alto,
CA: 2009. 1017599.
7. Engineering and Economic Evaluation of Biomass Power Plants, 100% Biomass
Repowering, Biomass Co-Firing, and Bubbling Fluidized Bed Biomass Combustion. EPRI,
Palo Alto, CA: 2010. 1019762.
8. Engineering and Economic Evaluation of Central-Station Solar Photovoltaic Plants. EPRI,
Palo Alto, CA: 2009. 1017600.
13-17

Greenhouse Gas Emissions Control

9. Engineering and Economic Evaluation of Geothermal Power Plants. EPRI, Palo Alto, CA:
2010. 1019761.
10. Solar Thermal Technology Status and Performance and Cost Estimates2010. EPRI, Palo
Alto, CA: 2010. 1022504.
11. Stanton, B., Eaton, J., Johnson, J., Rice, D., Schutte, B., and Moser, B., Hybrid Poplar in the
Pacific Northwest, the Effects of Market-Driven Management, Journal of Forestry, June
2002.

13-18

Export Control Restrictions

The Electric Power Research Institute, Inc. (EPRI, www.epri.com)

Access to and use of EPRI Intellectual Property is granted with the spe-

conducts research and development relating to the generation, delivery

cific understanding and requirement that responsibility for ensuring full

and use of electricity for the benefit of the public. An independent,

compliance with all applicable U.S. and foreign export laws and regu-

nonprofit organization, EPRI brings together its scientists and engineers

lations is being undertaken by you and your company. This includes

as well as experts from academia and industry to help address

an obligation to ensure that any individual receiving access hereunder

challenges in electricity, including reliability, efficiency, health, safety and

who is not a U.S. citizen or permanent U.S. resident is permitted access

the environment. EPRI also provides technology, policy and economic

under applicable U.S. and foreign export laws and regulations. In the

analyses to drive long-range research and development planning, and

event you are uncertain whether you or your company may lawfully

supports research in emerging technologies. EPRIs members represent

obtain access to this EPRI Intellectual Property, you acknowledge that it

approximately 90 percent of the electricity generated and delivered in

is your obligation to consult with your companys legal counsel to deter-

the United States, and international participation extends to more than

mine whether this access is lawful. Although EPRI may make available

30 countries. EPRIs principal offices and laboratories are located in

on a case-by-case basis an informal assessment of the applicable U.S.

Palo Alto, Calif.; Charlotte, N.C.; Knoxville, Tenn.; and Lenox, Mass.

export classification for specific EPRI Intellectual Property, you and your
company acknowledge that this assessment is solely for informational

Together...Shaping the Future of Electricity

purposes and not for reliance purposes. You and your company acknowledge that it is still the obligation of you and your company to make
your own assessment of the applicable U.S. export classification and
ensure compliance accordingly. You and your company understand and
acknowledge your obligations to make a prompt report to EPRI and the
appropriate authorities regarding any access to or use of EPRI Intellectual Property hereunder that may be in violation of applicable U.S. or
foreign export laws or regulations.

Program:
Renewable Generation

2012 Electric Power Research Institute (EPRI), Inc. All rights reserved. Electric Power
Research Institute, EPRI, and TOGETHER...SHAPING THE FUTURE OF ELECTRICITY are
registered service marks of the Electric Power Research Institute, Inc.

1023993

Electric Power Research Institute


3420 Hillview Avenue, Palo Alto, California 94304-1338PO Box 10412, Palo Alto, California 94303-0813 USA
800.313.3774 650.855.2121askepri@epri.comwww.epri.com

Você também pode gostar