Important For Wind Turbine Blade PDF
Important For Wind Turbine Blade PDF
Important For Wind Turbine Blade PDF
DETAILS
CONTRIBUTORS
GET THIS BOOK Committee on Assessment of Research Needs for Wind Turbine Rotor Materials
Technology, National Research Council
SUGGESTED CITATION
National Research Council 1991. Assessment of Research Needs for Wind Turbine
Rotor Materials Technology. Washington, DC: The National Academies Press.
https://doi.org/10.17226/1824.
Visit the National Academies Press at NAP.edu and login or register to get:
Distribution, posting, or copying of this PDF is strictly prohibited without written permission of the National Academies Press.
(Request Permission) Unless otherwise indicated, all materials in this PDF are copyrighted by the National Academy of Sciences.
ASSESSMENT OF RESEARCH
NEEDS FOR WIND TURBINE
ROTOR MATERIALS
TECHNOLOGY
Committee on Assessment of Research Needs for Wind Turbine Rotor Materials Technology
Energy Engineering Board
Commission on Engineering and Technical Systems
National Research Council
ii
NOTICE: The project that is the subject of this report was approved by the Governing Board of the National Research Council, whose mem-
bers are drawn from the councils of the National Academy of Sciences, the National Academy of Engineering, and the Institute of Medicine.
The members of the committee responsible for the report were chosen for their special competences and with regard for appropriate balance.
This report has been reviewed by a group other than the authors according to procedures approved by a Report Review Committee con-
sisting of members of the National Academy of Sciences, the National Academy of Engineering, and the Institute of Medicine.
The National Academy of Sciences is a private, nonprofit, self-perpetuating society of distinguished scholars engaged in scientific and
engineering research, dedicated to the furtherance of science and technology and to their use for the general welfare. Upon the authority of
the charter granted to it by the Congress in 1863, the Academy has a mandate that requires it to advise the federal government on scientific
and technical matters. Dr. Frank Press is president of the National Academy of Sciences.
The National Academy of Engineering was established in 1964, under the charter of the National Academy of Sciences, as a parallel
organization of outstanding engineers. It is autonomous in its administration and in the selection of its members, sharing with the National
Academy of Sciences the responsibility for advising the federal government. The National Academy of Engineering also sponsors engineer-
ing programs aimed at meeting national needs, encourages education and research, and recognizes the superior achievements of engineers.
Dr. Robert M. White is president of the National Academy of Engineering.
The Institute of Medicine was established in 1970 by the National Academy of Sciences to secure the services of eminent members of
appropriate professions in the examination of policy matters pertaining to the health of the public. The Institute acts under the responsibility
given to the National Academy of Sciences by its congressional charter to be an adviser to the federal government and, upon its own initia-
tive, to identify issues of medical care, research, and education. Dr. Samuel O. Thier is the president of the Institute of Medicine.
The National Research Council was organized by the National Academy of Sciences in 1916 to associate the broad community of sci-
ence and technology with the Academy’s purposes of furthering knowledge and advising the federal government. Functioning in accordance
with general policies determined by the Academy, the Council has become the principal operating agency of both the National Academy of
Sciences and the National Academy of Engineering in providing services to the government, the public, and the scientific and engineering
communities. The Council is administered jointly by both Academies and the Institute of Medicine. Dr. Frank Press and Dr. Robert M. White
are chairman and vice chairman, respectively, of the National Research Council.
This report and the study on which it is based were supported by Contract No. DE-AC04-89AL58181 between the U.S. Department of
Energy and the National Academy of Sciences-National Research Council.
Library of Congress Catalog Card No. 91-60990
International Standard Book Number 0-309-04479-0
NAT S-324
Additional copies of this report are available from:
National Academy Press
2101 Constitution Avenue, N.W.
Washington, D.C. 20418
Printed in the United States of America
iii
GEORGE E. DIETER (Chairman), Dean, College of Engineering, University of Maryland, College Park,
Maryland
JAMIE CHAPMAN, Power Systems Consultant, Boston, Massachusetts
H. THOMAS HAHN, Department of Engineering Science and Mechanics, Pennsylvania State University,
University Park, Pennsylvania
DEWEY H. HODGES, School of Aerospace Engineering, Georgia Institute of Technology, Atlanta, Georgia
CHARLES W. ROGERS, Bell Helicopter Textron, Inc., Fort Worth, Texas
LENA VALAVANI, Department of Aeronautics and Astronautics, Massachusetts Institute of Technology,
Cambridge, Massachusetts
MICHAEL D. ZUTECK, Consultant, Kemah, Texas
iv
Staff
ARCHIE L. WOOD, Executive Director, Commission on Engineering and Technical Systems, and Director,
Energy Engineering Board (to January 1991)
MAHADEVAN (DEV) MANI, Director, Energy Engineering Board
KAMAL J. ARAJ, Senior Program Officer
ROBERT COHEN, Senior Program Officer (retired)
GEORGE LALOS, Senior Program Officer
JAMES J. ZUCCHETTO, Senior Program Officer
JUDITH A. AMRI, Administrative/Financial Assistant
THERESA M. FISHER, Administrative Secretary
JAN C. KRONENBURG, Administrative Secretary
PHILOMINA MAMMEN, Administrative Secretary
NANCY WHITNEY, Administrative Secretary
PREFACE v
PREFACE
Wind-driven power systems represent a renewable energy technology that is still in the early stages of
development. These wind power plants installed in the early 1980s suffered structural failures chiefly because of
incomplete understanding of the wind forces (especially the turbulence component) acting on these large
structures and in some cases because of poor quality in manufacture. Failure of the rotor blades was one of the
principal and most serious structural failures. Failures from these causes are now somewhat better understood.
Another limitation to economical achievement of the potential of wind energy is uncertainty about the long-term
response of wind turbine rotor materials to the turbulent stochastic loadings to which they are subjected. These
structures can be subjected to as many as a billion stress cycles.
In accordance with its assessment of its long-term research responsibilities, the Department of Energy
requested the National Research Council to assess the research needs for wind turbine rotor technology. Such a
study would assist in organizing the information about the current status of wind turbine rotor materials, their
manufacture into blades, and their operation and life performance in service. Of special importance was an
assessment of current materials technology and design methodologies to provide perspective for future
investigations.
The Committee on Assessment of Research Needs for Wind Turbine Rotor Materials Technology was formed
by the Energy Engineering Board to specifically evaluate the following issues (see Appendix A for Statement of
Task):
In addition, the committee took as its responsibility the development, in broad outline, of a research and
development program that would place U.S. wind power technology in a preeminent world position.
In carrying out its assignment the committee examined an extensive literature on wind machines and wind
turbine rotor materials. The emphasis of the study was on wind machines suitable for utility applications. Many
U.S. experts on these subjects briefed the committee in two meetings of two-day duration (see Appendix B). The
committee wishes to acknowledge with gratitude the assistance of the following individuals for their time and
knowledge:
PREFACE vi
Holt Ashley, Stanford University; Charles Carlson and Marilyn W. Wardle, E.I. duPont de Nemours & Co.; John
C. Doyle, California Institute of Technology; Brant Goldsworthy, ALCOA/Goldsworthy Engineering; Richard H.
Hilt, Gas Research Institute; William Holley, U.S. Windpower; Donald Hunston, National Institute for Standards
and Technology; Edwin T.C. Ing; John Mandell, Montana State University; Robert C. Monroe, Hudson Products;
Robert H. Monroe, Gougeon Brothers; Peter Ogle, Dow United Technologies Composite Products; Donald
Pederson and Fred J. Policelli, Hercules, Inc.; Lawrence Rehfield, University of California-Davis; Forrest S.
Stoddard, Alternative Energy Institute; Robin Whitehead, Northrop; Daniel F. Ancona III, Leonard J. Rogers and
Jeffrey H. Rumbaugh, U.S. Department of Energy; James Tangler and Robert Thresher, Solar Energy Research
Institute; and Herbert Sutherland, Sandia National Laboratories.
CONTENTS vii
CONTENTS
LIST OF FIGURES ix
LIST OF TABLES x
EXECUTIVE SUMMARY 1
1 INTRODUCTION 5
Scope and Content 5
Wind-Driven Power Plants 5
Why Materials Knowledge Is Critical 8
The Evolution of Wind-Driven Power Plants 9
The Principal Component: Wind Turbines 12
Power Conversion Equations 16
The Wind Environment 19
Fatigue Cycle Accumulation 20
References and Bibliography 23
2 STRUCTURAL LOADING CHARACTERISTICS 25
Load Characterization 26
Blade Failure Experience 27
Lessons from Helicopter Experience 27
Recommendations 32
References and Bibliography 33
3 MATERIALS PROPERTIES AND LIFE PREDICTION 35
Fibers 35
Matrix Materials 39
E-Glass/Plastic Composites 43
Fatigue Life Prediction 48
Toughness Considerations 49
Wood/Epoxy Composites 50
Recommendations 60
References and Bibliography 61
4 WIND TURBINE ROTOR DESIGN 67
Airfoil Evolution 67
Aerodynamic Tip Brakes 69
Blade Root Retention 71
Glass-Reinforced Plastic (GRP) Blade Roots 71
Wood/Epoxy Blade Roots 75
Blade Joining 75
Blade Design Considerations 76
Fiberglass Blades 76
Wood/Epoxy Blades 78
Recommendations 79
GRP 79
Wood/Epoxy 79
Generic 79
References and Bibliography 80
CONTENTS viii
LIST OF FIGURES ix
LIST OF FIGURES
LIST OF TABLES x
LIST OF TABLES
EXECUTIVE SUMMARY 1
Executive Summary
The committee's charge was to define a research and development agenda for the Department of Energy's
wind energy program focusing on materials aspects of the wind turbine rotor technology. In particular, the
committee interpreted its mandate to include an assessment of the potential for new design methods, better
manufacturing processes, and advanced control methods in addition to advanced materials for improvement in
turbine rotor performance and lifetime.
In addition, the committee took as its responsibility the development, in broad outline, of a research and
development program that would place U.S. technology in rotors for wind power in a preeminent world position.
In carrying out its charge the committee examined an extensive literature on wind machines and wind turbine
rotor materials and was briefed by many experts on these subjects, both in the wind industry and in related
industries.
Wind-driven power systems represent a renewable energy technology that is still in the early stages of
development. Arrays of interconnected wind turbines convert the power carried by the wind into electricity for
users of the utility power grid. Major concentrations of this technology exist in California, Denmark, and Hawaii.
At the end of 1989, the wind power plants in California comprised a power-generating capacity of 1335 MW,
equivalent to a medium-sized utility power plant. Not only do these wind power plants produce no gaseous
emissions, particulates, or radioactive by-products, but they can be installed rather quickly as modular units, each
providing capacity from a few tens of kilowatts to hundreds of megawatts. They can be readily integrated into
existing utility generation-transmission-distribution systems. While economic utilization of wind energy depends
critically on location and siting, many available sites exist in the United States. While land intensive, wind power
plants can coexist with other uses of the land on which they are situated.
Wind power plants installed in the early 1980s suffered structural failures chiefly because of incomplete
understanding of the wind forces (especially the turbulence component) acting on these large structures and in
some cases because of poor quality in manufacture. Failure of the rotor blades was one of the principal and most
serious structural failures. Failures from these causes are now somewhat better understood. An associated
additional limitation to achievement of the full economic potential of wind energy is uncertainty about the long-
term response of wind turbine rotor materials to the turbulent stochastic loadings to which they are subjected.
Over their projected operational lifetimes (typically 20 to 30 years), these structures are subjected to as many as a
billion stress cycles.
Since a number of U.S. electric power utilities are continuing to add capacity, there will be an opportunity to
introduce new, longer-lasting designs. Moreover, renewed public interest in environmental issues associated with
power generation gives a renewed impetus to wind power. A new wind turbine system will probably take
advantage of advances in semiconductor power electronics to produce changes in the system configuration that
will make wind-generated electric power more amenable for use by electric utilities. New speed control schemes
will be introduced, but a major advance must come
EXECUTIVE SUMMARY 2
through the design of less expensive, longer-lived, and higher-efficiency rotor blades. A guiding principle in
creating this design should be that knowledge of aerodynamic forces must be carefully integrated with the
structural response of the material, all balanced by the practicalities of field experience and tempered by the need
to manufacture a consistently high-quality product at reasonable cost.
This committee has examined the experience base accumulated by wind turbines and the accompanying R&D
programs sponsored by the Department of Energy. We have concluded that a wind energy system such as
described above is within the capability of engineering practice. But certain gaps in knowledge exist, so achieving
the goal without costly and inefficient trial and error requires certain critical research and development. Because
of the fragile nature of the wind power equipment producers in the United States, this will require an R&D
investment from the Department of Energy.
The committee cannot conclude without commenting on the status of the wind power equipment industry.
Because of the decrease in the rate of installation of machines in the last 5 years (since the tax incentives expired),
there currently is only one major integrated manufacturer in the United States; only a few companies are actively
producing blades. In recent years a major Japanese manufacturer has entered the world market, joining the
European manufacturers who have been participants for some time. As a result, the U.S. industry is not in a
financial position to engage in the R&D necessary to gain worldwide technological leadership for what the
committee sees as a future growing worldwide market for wind power. The committee believes that the United
States is facing a future reduction in most fossil fuel sources of energy. When this is coupled with a resurgence of
public concern over environmental issues in energy production, the need to develop wind power energy to the
fullest extent possible seems compelling.
RESEARCH RECOMMENDATIONS
The committee has identified four goals to guide the research needs in wind turbine rotor technology:
Goal 1 To improve the material properties and design capability so that the structure will either withstand higher
stresses or the same level of stress for a much longer period of time.
Goal 2: To lower the operating stress levels by altering the structural/configuration design.
Gual 3: To improve the blade manufacturing process so that quality variations and cost are minimized.
Goal 4: To reduce the cost of blades enough so that periodic replacement becomes cost-effective.
The details of specific research recommendations are given in Chapter 7 (Conclusions and
Recommendations) and, in greater depth, at the ends of Chapters 2, 3, 4, 5, and 6. The main research tasks are
summarized below.
MATERIALS
Initiate a program to generate long-term (10-year-equivalent), high-cycle fatigue (108-109) data for candidate
structural materials: glass-reinforced plastic, wood/epoxy, and advanced high-modulus composite materials under
appropriate environmental conditions. The program would contribute an element in a needed databank for wind
turbine blade materials.
DESIGN
The wind turbine industry needs design tools that are beyond the capability for development by the private
sector. For example, it is necessary to be able to compute laminate stresses in three dimensions and to
EXECUTIVE SUMMARY 3
relate these stresses by means of a suitable failure theory. It is also necessary to rigorously relate blade material
and geometry to beam-like stiffnesses. In turn, the stiffness and mass must be combined with suitable aerodynamic
models to determine structural dynamic response in the rotating field of the rotor. This type of analysis will help to
assess the benefits of active control as well as passive (such as by elastic tailoring) load relief systems. Most of
these tools either exist or are being developed in the aerospace industry. They are largely developed under
government systems procurements and should be extracted and adapted to the special needs of the wind turbine
rotor.
Emerging technologies in active and passive control of both the rotor and the generator need to be studied on
an overall system basis considering the probable gains in structural efficiency and reduction in blade life-cycle
cost.
MANUFACTURING
Two emerging manufacturing processes, resin transfer molding, and pultrusion, offer significant
opportunities for cost reduction but with attendant limitations on blade design freedom. A feasibility study should
be conducted to evaluate each of these processes in a real application and at a reasonable scale, allowing realistic
design and cost trade studies to be accomplished. For example, the cost of a geometrically simple pultruded blade
can be little more than the cost of the materials, thus making it appreciably less costly than the curved and shaped
wind turbine blades currently employed. However, while less costly, such blades are also less efficient
aerodynamically. Thus, from a life-cycle economics perspective, it is not clear that the attendant reduction in the
initial cost of the wind turbine can compensate for the associated decrease in energy production over the wind
turbine lifetime.
EXECUTIVE SUMMARY 4
INTRODUCTION 5
1
INTRODUCTION
INTRODUCTION 6
Figure 1-1
Wind power plant in Altamont Pass, California. (Courtesy, U.S. Windpower, Inc.)
INTRODUCTION 7
wind power plants in California represented a power-generating capacity of 1335 MW, equivalent to a
medium-size nuclear or coal-fired power plant. During 1989 the wind power systems of California delivered
slightly more than 2 billion kWh of electricity. This amount of electrical energy is equal to the annual residential
needs of a city the size of San Francisco or Washington, D.C. In Europe, Denmark has been a pioneer in the use of
modern wind turbines for electricity generation. The approximately 2500 wind turbines installed throughout
Denmark currently supply about 1 percent of the country's annual electrical needs. We note here that
approximately half of the wind turbines installed in California during the 1980s were of Danish manufacture. This
is relevant to the longer-term impact of this study as it relates to the future viability of U.S. manufacturers.
The progress of the last decade was not achieved without difficulty and, at times, controversy. Early
structural failures, together with under-capitalization of some wind power plant operators, led to the perception,
from some quarters, that the California installations were ''tax farms.'' However, as the systems improved in
performance and began to produce large amounts of energy, these accusations were replaced by a growing
appreciation of the environmental benefits and advantages offered by this renewable power generation
technology. The advantages and benefits include the following:
Environmentally Benign: Wind power systems are environmentally benign. There are no gaseous emissions,
no particulates, and no radioactive by-products.
Rapid Modular Addition of Capacity: Wind power systems can be installed quickly, thus reducing financing
costs and providing flexibility in meeting demand growth.
Wide Range of Capacities: Wind power systems can range in size from very small systems to power plants of
utility scale (i.e., from a few kilowatts to hundreds of megawatts in power-generating capacity).
Ease of Integration: Wind power systems are readily integrated into existing utility generation-transmission-
distribution systems, using standard utility components and practices. The wind turbine is the only new
component.
Couse of Land: While land intensive, wind-driven power plants coexist with other uses of the land on which
they are situated. The wind turbine and towers, service roads, and electrical equipment typically occupy only 10
percent of the land on which the wind power plant is installed. Previous uses, typically agricultural, can continue
with little useful area removed.
Use of Indigenous Resources: Wind-driven power plants use indigenous resources for their fuel. Since the
fuel is without cost, the user country achieves a degree of energy independence and preserves hard currency
assets. This is important not only for developed countries with adequate wind resources but perhaps more so for
developing countries.
Economically Competitive: The cost of energy (CoE) from wind power plants is competitive now with some
conventional energy generation sources. Large-scale systems now being installed deliver energy at costs in the
range 7 to 9c/kWh. Under comparable wind resource conditions, third-generation wind turbines (under design
now) are expected to deliver energy at costs in the range of 4 to 6c/kWh.
These advantages are applicable not only to this country but globally as well. While not a complete solution
to environmental concerns, wind power systems do represent an attractive part of the total generation mix.
As with most technologies, the advantages of wind power systems must be weighed against the disadvantages
and limitations. The disadvantages include their visual impact, noise, potential interference with the reception of
television signals, and potential hazard to birds. Visual impact is inescapable, particularly at close distances.
The principal sources of noise in wind turbines have been the blades and gearboxes. Early machines were
relatively noisy compared with more recent designs. Advances in blade airfoil shape and manufacture have
significantly reduced the noise from wind turbine blades. Similarly, attention to the sources of noise in wind
turbine gearboxes has resulted in significant reductions. However, with both sources a certain amount of noise is
INTRODUCTION 8
inevitable. At reasonable distances (several rotor diameters), wind turbine noise can be difficult to discern from
ambient background and cultural noise sources (e.g., wind noise, automobile traffic).
Television interference has not been a problem with recent large installations. This may be due in part to the
use of composite materials in the blades and the fact that people typically do not reside in close proximity to wind
turbines.
There also are limitations to the broader application of the technology. The relative newness of the technology
has justifiably required the accumulation of significant operating experience. This limitation is being removed by
the favorable technical and economic performance accumulated over the past half decade or more by many
existing installations.
A fundamental limitation is that the economic generation of electricity is governed by the quality of the wind
resource—no wind, no energy; poor wind, some energy but unattractive economics. Wind systems thus have a
geographical limitation based on availability of the required resource—wind.
Finally, a limitation of current systems is that the electricity generated is time variable, reflecting the
variability of the wind. This variability occurs on several time scales. Seasonal and diurnal variations are
statistically predictable, such that the generated energy can be factored into the planning of the overall utility
generation mix. However, time variations on the order of minutes and tens of minutes (faster than the ramp rates
of conventional generation sources) can pose system voltage and frequency stability problems. The utility rule of
thumb is that wind systems of current design are limited to the range of 5 to 15 percent of the total generating
capacity. For a given utility system, the specific wind penetration limit depends on such technical details as
transmission line characteristics, the ratio of real and reactive power, and allowable excursions of voltage and
frequency. Thus far, with the large installations in California, system stability has not been a major problem due to
the very large utility systems to which the energy is supplied. The fraction of wind generation capacity is less than
10 percent of the total generation on-line. Depending on the time of day and the load profile, the wind systems on
the island of Hawaii do approach limitations imposed by system stability arguments. At night, when the load is
minimal, the installed wind systems can exceed the range cited above.
The system stability limitation represents an upper bound on the amount of current design wind capacity that
can be integrated with the balance of the utility generation mix. As such, this limitation is a market limitation
having technical origins. There are three approaches, now undergoing research and development, that will mitigate
and eventually remove this limitation. New wind turbine system designs form the first of these approaches, one
example of which is the variable-speed architecture with power electronic control and interface to the grid. Other
less complex examples are provided by innovative, passive, load-relieving rotor designs and by advances in self-
regulating airfoils. A second approach that will provide relief from this limitation is improved control of the wind
system and integration of control with the conventional generation sources. Finally, the third technique for
removing the penetration limitation is the incorporation of energy storage, which will compensate for the time
variability of the wind.
We are beginning to see all of these techniques employed in smaller integrated power systems. These are
systems used in regions where no large regional grid system is available, where the wind power capacity is
comparable to the conventional power generation capacity (often provided by diesels), and where the conventional
power sources can be turned off, with the load being supplied by the wind system.
INTRODUCTION 9
more particularly the turbulence component. Research, field measurements, and improved modeling, augmented
by operational experience, have greatly increased our understanding of these stochastic forcing functions and the
response of the wind turbine. Nevertheless, it is still true that the turbulence structure of the wind contributes the
most uncertainty in the design and sizing of the major structural components of a wind turbine.
From the perspective of a designer, this uncertainty continues to be due in part to incomplete knowledge of
the turbulence and its description and in part to the difficulty of modeling the structural vibrational responses.
However, a major contributor is uncertainty about the long-term responses of the wind turbine materials to the
turbulent stochastic loadings. The specific long-term responses of interest here are fatigue failures-that is, failures
due to the cumulative effects of many millions of flexural cycles of a structural component. Fatigue failures
represent the greatest uncertainty with regard to the long-term service lifetime (typically projected to be 20 to 30
years) of the major structural components of a wind turbine. Thus, fatigue failures represent a major uncertainty in
the life-cycle CoE.
Uncertainty about the properties of materials causes the wind turbine designer either to add more weight (and
cost) than is required or to misjudge and inadequately size a component so that failure occurs (usually more
costly). To further improve the economics of wind turbine systems and increase their range of use, improved
materials properties are required. This is particularly true with regard to the long-term fatigue properties of
composite materials. As used thus far in wind turbines, composite materials are combinations of glass, other
synthetic fibers, or wood in a resin matrix. With the anticipated increasing use of these and other composite
materials, improved knowledge about both their static strength and their fatigue properties becomes critical in
order to assure both short-term performance and the long-term life required of these power systems. This
knowledge base is particularly critical for composites because of the wide variation in their geometries,
constituents, and manufacture.
In addition to the need for improved materials knowledge, there is also a need for increased awareness in the
wind turbine community (within both the federal and the private sectors) of advanced manufacturing processes.
These processes, which may be used in other industries, are required to support not only the more extensive use of
composites but also their large-volume, economical manufacture.
INTRODUCTION 10
Figure 1-2
World energy generation by wind power plants during 1989 (units of millions of kWh). (Courtesy, Paul Gipe &
Associates.)
Figure 1-3
Growth of generating capacity for the California wind power plants. Sources: California Energy Commission (CEC);
CEC Performance Reporting Systems (PRS). (Courtesy, Paul Gipe & Associates)
INTRODUCTION 11
through the projection for 1990. The line graph of Figure 1-3 depicts almost the same information but in a
cumulative format. This graph shows the cumulative total of the net installed capacity at the end of each year
during this 10-year period.
There are four features in the data of Figure 1-3 that merit attention: (1) the current total rated capacity, (2)
the shape of the year-to-year growth in capacity, (3) a comparison of the capacity installed before and after the end
of 1985, and (4) the 1988 decrease in the cumulative total capacity. Discussion of these features will illuminate the
reasons behind the rapid evolution and growth of these large-scale renewable power systems. The first thing to
notice about the information in Figure 1-3 is the current rated capacity of the wind power plants. For a power
generation technology that has evolved over the recent decade, the nearly 1500 MW of installed capacity is
substantial. The second aspect of Figure 1-3 that merits comment is the rapid growth of the year-to-year added
capacity from 1981 through 1985 and the generally reduced rate of new installations in the succeeding years. This
is a manifestation of the mechanisms used to initiate the early, large-scale commercialization of a relatively new
and untried power generation technology. These mechanisms included tax incentives at the federal level and at
some state levels (including California) and federal legislation requiring regulated utilities to purchase the energy
produced by nonregulated or independent power producers (IPPs).
The development occurred first in California not only because of the three excellent wind regions, but also
because of a favorable legislative and regulatory climate at the state level. The legislative climate resulted in tax
credits to California residents who invested in renewable energy generation facilities. The favorable regulatory
climate, as represented through the actions of the California Public Utilities Commission and the California Energy
Commission, influenced the terms of power purchase contracts negotiated between the regulated utilities and the
IPPs.
These policies, factors, and influences permitted the attraction of the large amounts of private capital needed
to launch a new, capital-intensive industry. The wind power systems of California represent a private capital
investment of about $2.5 billion. The incentives and the attendant attractive returns compensated the investors for
participation in a risky and relatively untried power generation technology.
The tax incentives at the federal level expired at the end of 1985. The California tax incentives were phased
out over the 2-year period of 1986 and 1987. While there were some carryovers and grandfathering, for all
practical purposes the tax credit era that fostered the early development of wind power technology expired at the
end of 1985. The effects can be seen in the post-1985 decrease in the yearly installation rate of new capacity,
depicted in the bar graph of Figure 1-3.
At this point, we are equipped to appreciate the third feature in the data of Figure 1-3. If one sums in
Figure 1-3 the yearly capacity added from 1986 through the projections for 1990, this post-tax-incentive added
capacity is about equal to that installed from 1981 through 1985. Aside from any carryovers or phaseouts of
special tax incentives, two factors supported continued installations in the years following 1985: (1) the
continually improving energy production and overall economic performance of the wind systems and (2) the
carry-forward provisions of some of the pre-1985 power purchase contracts.
Finally, in Figure 1-3, the 1988 dip in the cumulative capacity on-line (the line graph) reflects the fact that a
number of older, first-generation wind turbines were retired. They were replaced, to some extent but not entirely,
by the new capacity (59 Mw) installed during 1988.
The growth in annual energy production of the California wind power systems is depicted in Figure 1-4. The
data in this figure show that energy production has steadily improved from 1981 through the present.
Nevertheless, averaged over the fleet, the current average capacity factor is still not up to the realistically
achievable potential in the wind regimes of the three California regions. For good sites in any of these regions, a
capacity factor of 0.25 is a reasonable performance value. The best of the current machines do achieve this value
or higher.
INTRODUCTION 12
Figure 1-4
Growth of annual energy production for California wind power plants.
Sources: CEC, CEC PRS. (Courtesy, Paul Gipe & Associates.)
The high energy prices in the early part of the 1980s were accompanied by the expectation of still higher
prices. This led to the development and approval of negotiated power purchase contracts wherein the energy prices
paid were consistent with the prevailing expectations of high, and higher, energy prices. By means of a long-term
fixed-price payment provision in certain contracts, the operator of wind power systems surrendered the possibility
of receiving near-future higher energy sales prices in return for defined energy prices that were fixed for a long-
term period (typically 10 years).
This benefited the utility rate payers by putting a cap on energy prices from these sources; it benefited the
IPPs by providing revenue certainty over the period of fixed prices. This period of revenue certainty (often
coincident with the term of third-party financing) supported the continued (post-1985) attraction of investment
capital from the private sector. Now, however, these investments, made in the absence of special tax incentives,
took on a different complexion.
These factors allowed wind systems to be financed on the basis of their economic performance. Thus,
beginning in 1986, large-scale wind power plants began to be financed in the same ways that other industrial
capital equipment is financed. These included sale-lease transactions and straight debt, secured by the equipment,
with repayment supported by the revenue streams generated by the equipment.
INTRODUCTION 13
renewable energy source will not be realized until costs can be reduced further.
Both the installed capital cost and the continuing maintenance costs need to be reduced. Two paths by which
this might occur are improved knowledge of the behavior of materials and the use of advanced manufacturing
techniques. Through utilization of these and other improvements, it is believed that the CoE from wind systems
can be brought to the range of 4 to 6c/kWh. This range will make wind power systems competitive not only with
nuclear but also with new coal-fired installations. At the same time, the performance characteristics will be
improved to the extent that wind turbine installations can be economically productive over a broader class of wind
regimes and geographical regions.
There are two generic types of wind turbines: (1) the horizontal axis wind turbine (HAWT) and (2) the
vertical axis wind turbine (VAWT). Although many of the findings and recommendations of this study are
applicable to both architectures, the focus here is on the HAWT. This is the type predominantly used in the
approximately 20,000 wind turbines now in service throughout the world.
Illustrated in Figure 1-5 are the principal elements of a wind turbine. These include (1) the rotor, consisting
of the blades and the supporting hub; (2) the drive train, consisting of the low-speed shaft (LSS), the gearbox or
transmission, the high-speed shaft (HSS), and the generator; (3) the machine bedplate or supporting frame; (4) the
yaw bearing and yaw orientation system; and (5) the tower. In the interest of clarity, the bedplate or frame has
been omitted from the drawing of Figure 1-5, as have the tower foundation and most of the tower.
Figure 1-5
Wind turbine subsystems.
INTRODUCTION 14
The wind turbine blades are rotating airfoil surfaces that transform power in the wind to mechanical shaft
power. As will be discussed later, the blades also provide one means for the control of power flow through the
drive train.
The wind turbine gearbox is a step-up type, used to convert the low-speed power and torque produced by the
rotor to their high-speed equivalents required for operation of the generator. The gearbox matches the different
operating speeds of the generator and the rotor.
The electrical machine typically used as a generator in wind turbines is the four-pole, squirrel cage induction
machine or, more briefly, an induction machine. An induction machine operates within a narrow range of speeds
centered about the synchronous speed of the machine. For a four-pole induction machine designed for use on a
60-Hz electrical system, the synchronous speed is 1800 rpm. At speeds below synchronous, the induction machine
functions as a motor. At speeds above synchronous, it operates as a generator.
The range of operating speeds is defined by the slip range. Typically, the slip range for the four-pole
induction machines used in wind turbines is 2 percent of synchronous speed, or about 35 rpm. Used as a
generator, this machine begins to develop power at about 1800 rpm and reaches full rated power at about 1835
rpm.
The aerodynamic characteristics of the blade airfoils, the rotor diameter, and the range of operating wind
speeds determine the optimum rotor speed. For wind turbines having power ratings in the range of 100 to 600 kW,
the range of current interest and emphasis, the corresponding rotor speed range would be approximately 35 to 70
rpm. Thus, for an 1800-rpm generator and this range of rotor speeds, the corresponding gearbox ratios would be in
the range of 26:1 to 51:1. Typically, such ratios would be implemented by two- or three-stage gearboxes.
The wind speeds over which turbines can generate power typically lie in the range of 10 to 50 mph. For the
California wind regions, the number of hours per year that the wind speeds lie within this operating range varies
from 3000 to 4500. The actual number of operating hours accumulated in any given year depends not only on
interannual wind statistics and local siting effects but also on the characteristics of a given wind turbine. For use in a
later illustration, we choose the intermediate value of 4000 hours per year. This means that the wind turbine
operates often enough to generate power 4000 hours out of the 8760 hours per year, or about 46 percent of the
time. Typically, there is a pronounced seasonal variation. In California, for example, about 70 percent of the
annual energy is produced during the five summer months, May through September.
In most regions there are many hours per year when the wind speeds are less than 10 mph. For these speeds
the wind does not carry enough power, with current technology, to overcome the effects of conversion
inefficiencies and losses. At the other end of this range, the number of hours during which the wind speeds are
greater than 50 mph usually is quite small. Thus, for wind speeds greater than 50 mph, the added revenues
generated do not compensate for the costs of the necessary increased structural strength. Finally, it is important to
note that the wind turbine structure is required to survive very high (nonoperating) wind speeds. A typical design
survival wind speed is 125 mph (56 m/s).
As a function of wind speed, the power output of wind turbines may be characterized by four numbers. In this
simplified but useful description, three of the numbers specify characteristic wind speeds and one defines the
power rating of the wind turbine:
vci the cut-in wind speed, defined as the wind speed at which the wind turbine begins to generate useable amounts of
electrical power. Typical values are vci = 10 mph = 4.5 m/s.
vr the rated wind speed, defined as the lowest wind speed at which the wind turbine reaches the maximum value of its
electrical power output. Typical values are vr = 27 mph = 12 m/s.
INTRODUCTION 15
Pr the rated power, defined as the maximum value of the electrical power output.
Vco the cut-out wind speed, defined as the wind speed at which the wind turbine reduces its electrical output from the
maximum rated power level to zero. Although this usually occurs over a narrow range of wind speeds, for simplicity
it is often assumed to occur at the single value vco. Typical values are vco = 50 mph = 22 m/s.
These values define the wind turbine power curve, a graph of the electrical power output as a function of
wind speed. The general features of a wind turbine power curve are depicted in Figure 1-6. The wind turbine
produces no power from zero wind speed through the cut-in value vci. As the wind speed increases further, the
power output rises, reaching the rated value Pr at the wind speed vr. As the wind speed increases beyond the value vr,
the wind turbine control system (which may be active or passive) endeavors to hold the output power constant at
the rated value Pr. This output value is maintained until reaching the cut-out wind speed vco, at which speed the
power output is reduced to zero.
The power curve of Figure 1-6 is an idealization. It neglects differences in wind turbine control strategies. It
accounts for rapid variations in wind speed only in an average sense. However, this representation does indicate
the average performance characteristics of a
Figure 1-6
Wind turbine power curve.
INTRODUCTION 16
given wind turbine and is useful in predicting its annual energy production. We shall use Figure 1-6 to illustrate
simplified equations that describe the conversion wind power to electrical power delivered by the wind turbine.
Equation (1) shows the dependence of the power in the wind incident upon a rotor of swept area A as a
function of the wind speed v. Equation (2) describes the electrical power output of the wind turbine resulting from
this flow. These equations are defined in terms of the following parameters:
Figure 1-7
Wind flow field and turbine loads.
INTRODUCTION 17
Figure 1-8
Comparison of logarithmic power in the wind with a wind turbine power curve (Pr = 100 kW).
The last portion of Equation (2) highlights the important dependencies. This relationship indicates that the
output power as well as the power flowing through the drive train is proportional to the cube of the wind speed, to
the area of the rotor, and to the power conversion efficiency of the rotor. Under the idealized assumptions of this
model, the maximum value achievable by the rotor power conversion coefficient is 16/27, the Betz limit:
INTRODUCTION 18
For comparison, the better-performing wind turbine rotors achieve maximum power coefficient values in the
vicinity of Cp actual= 0.45. This disparity indicates the deficiencies (e.g., neglect of angular momentum) in the
model used to derive the Betz limit and, to a limited extent, the potential for improvement in blade performance.
These equations illustrate the major problem in wind turbine design: the control or modulation of the power
flow through the wind turbine drive train. This power flow has both a steady-state, quasi-static component and a
fluctuating component. For the time being, we continue with the assumption of an incident wind flow field that is
constant in both time and space. This allows us to focus on the power control problem.
The problem is that the power in the wind flow field increases as v3 since the rotor area is fixed in practically
all designs. Thus, over the 10 to 50 mph operating range cited earlier, the incident wind power increases by a
factor of 125. Since the wind turbine is designed to deliver electrical power at the value Pr, a more realistic
measure of the dynamic range that a wind turbine must accommodate is given by the ratio of the wind speed at cut
out to the rated wind speed:
Thus, if the wind turbine's rated power is Pr= 500 kW, the cubic dependence of Pw means that the turbine
must be able to accommodate at least 3.2 MW of power flow in the wind.
To accommodate these power levels directly (in brute force fashion with no control actions) over the typical
range of operating wind speeds would place a severe economic penalty on the structural design margins of the
rotor, drive train and tower, and foundations. The remedy is to recognize that Cp is not a constant but may be
actively or passively controlled. To show this, we rewrite Equation (2) to show the explicit dependence of Cp on
the wind speed v:
Thus, an upper bound on the power and torque experienced by the drive train is achieved by modulating Cp
(v). This is done typically by either active blade pitch control or passive stall control techniques. The relationship
between the power in the incident flow field and the wind turbine power output (which is functionally the same as
the drive train power levels) is shown in the two graphs of Figure 1-8. These are simply plots of Equations (1) and
(5) wherein a function Cp(v) for a nominal 100-kW wind turbine has been used. The curves in this figure illustrate
the need for control of power through the drive train. For the example shown in Figure 1-8, the cut-out wind speed
is about 50 mph. At this wind speed, the wind turbine must accommodate the power in the wind whose value is
more than six times either the shaft power or the electrical power output.
As discussed above, the wind turbine blades provide one means for control of power flow through the drive
train. Several variations are used to effect this control. The two techniques currently used most widely are (1) pitch
control, wherein the entire blade is rotated about the pitch axis so as to change the power extraction characteristics
of the blade, and (2) fixed pitch or stall control, wherein the blade is kept at a fixed pitch angle chosen so that the
blade becomes increasingly less efficient as the wind speed increases. These are illustrated in Figure 1-9 along
with other variations, including partial-span pitch control, wherein only an outboard portion of the blade is rotated
about the pitch axis; the use of ailerons; and the use of deployable tip sections to limit overspeeds. Also illustrated
is a control technique seldom used thus far in wind turbine airfoils—pumped spoiling or boundary layer control.
INTRODUCTION 19
Figure 1-9
Wind turbine blade control methods.
The discussion thus far has dealt with steady state or nonfluctuating loads experienced by the wind turbine
structure and components. There is, in addition, a fluctuating component to the wind turbine loads that arises from
several sources. These include vertical wind shear, temporal and spatial turbulence, gravity, and interaction of the
rotor with the tower. These sources are discussed next in the context of a more realistic description of the wind
flow field.
INTRODUCTION 20
There are fundamental disparities between these wind measurement parameters, the resulting descriptors of
the wind flow field, and the actual wind flow field incident upon the wind turbines. Thus, it is not surprising, in
retrospect, that some aspects of early designs were marginal. First, the relatively efficient rotors of modern wind
turbines respond quickly to wind changes, inducing correspondingly rapid loadings. Thus, the rotor and drive train
experience torque and other transients that would not be expected from the 10-minute or longer averages of
standard meteorological measurements.
Second, the center of the wind turbine rotor, the hub, sits not at a 10 m height but more typically at two to
three times that height. Because wind speeds increase with height in the earth's boundary layer where wind
turbines operate, wind speeds are typically higher at the hub height relative to a 10-m height. This increase of wind
speed with height is termed vertical shear. The function often used to model vertical wind shear is given by
Equation (6):
This function connects the wind speed v(h) at the height h with the wind speed v(h1) at the reference height h1.
The power law coefficient usually chosen as representative has the value = 1/7. From this relationship we see
that the wind speed at height h = 20 m can be 1.10 times the value at the standard meteorological sensor height of
10 m. Similarly, at the height h = 30 m, the wind speed may be 1.17 times greater in value.
To appreciate the impact of vertical wind shear, consider a rotor of diameter 20 m situated at a hub height of
25 m. Due to vertical wind shear alone, each blade during each revolution will experience a +7 percent (top) and a
-8 percent (bottom) variation in wind speed. This maps into a 45 percent variation in the power density (watts/m2)
of the incident wind flow field.
Thus, we see that even though the vertical wind shear may be time independent, the vertical spatial
dependence, when sampled by a rotating blade, induces cyclic loads. Cyclic loads also occur when the blades pass
through any stationary wakes or bow waves associated with the tower. These cyclic load components are
superimposed on the quasi-static loads associated with the average wind speed and power density. Gravity is of
course also time independent, but again the rotating blades experience cyclic bending moments due to gravity. The
influence of gravity in fact becomes a major limiting factor as wind turbine rotors become larger.
The remainder of the contributors to the cyclic load components are associated with fluctuations in the wind
flow field incident upon the wind turbine rotor. These flow fields always have both a time and a spatial
dependence over the rotor swept area. While we are unable to treat these fluctuating components adequately in the
context of this chapter, the cyclic loads produced by these components are in practice responsible for a significant
fraction of the fatigue damage experienced by wind turbine structural components.
INTRODUCTION 21
In this estimate we thus focus on the effects of gravity, vertical wind shear, and blade-tower interactions.
Ignored are the stochastic loadings associated with the actual turbulence structure of the wind flow field, including
the passage of the blade through cells of spatial turbulence and the effects of yaw and off-axis operation. Ignored
also are the fatigue damage effects of wind turbine stop-start events. Although fewer in number, their loading
amplitudes and fatigue life extraction may be quite large.
The lower-bound fatigue cycle estimate is thus based only on the deterministic sources of cyclic loads. In fact
only one source need be considered, as all are associated with and driven by the rotational speed of the wind
turbine rotor. To obtain a simple result, no information about the spectrum of loadings is included. All effects are
considered to have the same unspecified loading amplitudes. Not included are the balance of the deterministic
sources; the effects induced by the temporal, spatial turbulence of the wind; and any stop-start effects. All of these
neglected effects act to increase the rate of accumulation of fatigue cycles. Thus, the estimate is a lower bound on
the number of fatigue cycles accumulated over time.
With these simplifications, the number of fatigue cycles accumulated as a function of calendar time is given
by the function where
t is the elapsed calendar time (years) over which cyclic events are accumulated;
Hop is the wind turbine operating hours per year (hours/year);
is the rotational speed (rpm) of the low-speed shaft;
k is the number of cyclic events per revolution (cycles/rev); and
N is the number of cyclic events accumulated (pure number) during the elapsed calendar time t as a function of the
operating hours per year Hop, the low-speed shaft
rotational speed ω, and the number of cyclic events per revolution k, where
The number of operating hours depends on the wind turbine, its operating strategy, and the wind regime, but
typically the value lies in the range of 3000 hr/year to 4500 hr/year. The value Hop = 4000 hr/year is used in
Figure 1-10.
The range of low-speed shaft rotational speeds considered is 35 to 70 rpm. Wind turbines at the smaller end
of the size spectrum typically will operate near the upper end of this range, while larger machines will operate
near the lower end.
The parameter k, the number of cyclic events per revolution, typically can take on the integer values k = 1, 2,
or 3. Given the present simplifications, from the perspective of a blade, this parameter has the value k = 1.
Similarly, for the cyclic events experienced by the low-speed shaft, this parameter is equal in value to the number
of blades. The value k = 1 is used in Figure 1-10.
For the parameters and their units as defined above, the cumulative number of fatigue cycles is given by the
numerical expression of Equation (7):
This function is illustrated in Figure 1-10 as a function of the elapsed calendar time t. The parameters k and
Hop were chosen to have the values k = I and Hop = 4000 hr/year. Curves are given for each of the low-speed
shaft rotational speeds of 35 rpm and 70 rpm. This estimate shows that the number of fatigue cycles accumulated
reaches 107 very quickly, in about a year's time. The number of fatigue cycles reaches 108 after about 10 years of
INTRODUCTION 22
operation. By 20 to 30 years of operation (typically the expected lifetime of wind turbines), the cumulative
number of fatigue cycles lies in the range of 108 to 109.1
Figure 1-10
Accumulation of fatigue cycles.
This extreme accumulation of fatigue cycles with such a large amplitude range is unequaled by practically
any other mechanical or structural system. Further, economics are a severe constraint for wind turbine systems.
Wind turbine systems must operate competitively under an economic ceiling associated with other means of
generating power. As a result, the three cost components of wind turbine economics—capital cost, maintenance
cost, and overhaul cost—are subject to stringent and demanding requirements. Increased knowledge about the
properties of materials can contribute in a major way to the advancement and further utilization of this large-scale,
renewable energy source.
1 This analysis shows that fatigue data at 109 cycles are needed. However, the magnitude of the task of performing fatigue
tests to 109 cycles should not be underestimated. Test frequencies are limited to around 20 Hz for fiberglass materials, which
means a single test to 108 cycles would require about 100 days and 109 cycles would require about 2 years. When difficulties
with test equipment failure and power interruption were considered, fatigue testing to 109 cycles is a major undertaking.
Although we have called for fatigue test data at 109 cycles in a number of places in this report, this is an idealized
requirement. Much guidance on design and materials selection would be provided by test data in the 107 to 108 cycle range.
INTRODUCTION 23
INTRODUCTION 24
Rashkin, S. 1988. Results from the Wind Project Performance Reporting System: 1987 Annual Report. California Energy Commission,
Sacramento, California, August. (One of a series of quarterly and annual reports that track the energy production performance of the
California wind power stations.)
Ringer, M. 1984. Relative Cost of Electricity Production. California Energy Commission, Sacramento, California, July.
Solar Energy Research Institute. 1987. Wind Energy Resource Atlas of the United States. Solar Technical Information Program, Report DoE/
CH 10093-4, SERI, Golden, Colorado, March. (Excellent treatise on the locations and quality of wind resources in the United States
suitable for wind power installations.)
Solar Energy Research Institute. 1988. Wind Energy Technical Reading List. Solar Technical Information Program, Report SERI/SP-320-3400
DE88001193, SERI, Golden, Colorado, May. (Bibliography containing references to technical reports from the federal wind research
program.)
Stoddard, F. S. 1990. Wind Turbine Blade Technology: A Decade of Lessons Learned. Presented at the World Renewable Energy Congress,
Reading, United Kingdom, September.
Sutherland, R. J., and R. H. Drake. 1984. The Future Market for Electric Generating Capacity: A Summary of Findings. Los Alamos National
Laboratory, Los Alamos, New Mexico, December. (Among other topics, portrays the component of capital cost attributable to the
cost of funds used during construction of a nuclear power station, prior to its becoming a performing, operational asset.)
Utility Data Institute. 1988. 1987 Production Costs: Operating Steam-Electric Plants. Utility Data Institute, Washington, D.C., October.
2
STRUCTURAL LOADING CHARACTERISTICS
Among wind turbine failures, those associated with structural dynamics are most common. Stresses in the
blades oscillate, particularly at connections of the blade with other components of the wind turbine, leading to
fatigue failures. These failures are not adequately predictable based on the current state of the art. The reasons for
this are (1) the inability of current methods to accurately characterize the loading, (2) inadequate response
prediction methodology, and (3) a paucity of material test data at the high-cycle numbers.
In general, there are at least four ways to deal with fatigue-related problems in a structure:
1. improve the material characteristics so that the structure will sustain higher cyclic stresses,
2. lower the operating stress levels by altering the structural/ configuration design,
3. improve the manufacturing process so that quality variations are minimized, and
4. make the structure so inexpensive to manufacture and maintain that to periodically replace certain
components is cost-effective.
This chapter concentrates on the second approach—a type of aeroelastic tailoring. Shirk et al. (1986)
presented a definition for aeroelastic tailoring and suggested that it become the standard: ''Aeroelastic tailoring is
the embodiment of directional stiffness into an aircraft structural design to control aeroelastic deformation, static
or dynamic, in such a fashion as to affect the aerodynamic and structural performance of that aircraft in a
beneficial way.'' They also pointed out that parallels exist between aeroelastic tailoring and active control
methodology. Although active control focuses primarily on the control law, there is nothing to prevent the
structure from being modified to assist the active controller in its task. Similarly, even in the absence of an
external energy source, aeroelastic tailoring itself uses a form of control law to modify the behavior of a structure.
In addition to being a form of passive control, structural tailoring provides a method to optimize the
performance of actively controlled structures. Shirk et al. (1986) also mentioned that the effective integration of
active controls and structural stiffness is another area of potential reward. The aeroelastic benefits derived from
deformation control due to structural tailoring and the movement of actively controlled surfaces each have limits.
They suggested that the synergistic effect derived by the optimum interaction of each be explored. Issues of
controllability and observability of aircraft dynamics are strongly influenced by the flexibility of the structure.
Also, simultaneous optimization of control gains and structural properties often becomes prohibitively expensive
for large degrees of freedom, and physical insight is often lost in the resulting control law.
Belvin and Park (1988) undertook a pioneering study along this direction. They presented a method for
optimization of closed-loop structural systems, using such structural tailoring schemes as tailoring the thickness to
maximize the model stiffness of selected modes. Results were obtained for some simple structures, including a
beam and a truss beam, showing the
importance of structural tailoring to increase dynamic performance and to reduce the control effort.
Armanios et al. (1990) explored the benefits of tailoring the macro- as well as the micro-structure; that is,
they examined altering ply stacking sequence, fiber orientation, and blend of material plies, to contain and resist
damage in flexible structures. One of their purposes was to demonstrate that damage tolerance can be designed
into a structure. It was shown that for a generic damaged ply model, which includes microcracking, delamination,
and fiber fracture and their interaction, damage modes alter the stillnesses of the structural component at the
damage site. Therefore, redistribution of local stiffnesses could be used to enhance toughness.
Rehfield and Atilgan (1987) showed that in closed-cell blades additional couplings arise, other than those
designed in, which must be accounted for. These "parasitic" couplings are extension shear (accompanies bending
twist) and bending shear (accompanies extension twist). With these additional couplings the analysis will predict
structural stillnesses smaller (more flexible) than without them.
In light of the above, structural, elastic, and aeroelastic tailoring concepts are promising for active/passive
control of flexible structures. Therefore, development of an interdisciplinary analysis methodology for design,
optimization, and control of structures would be useful. Haftka and Kamat (1989) developed an approach for
simultaneous nonlinear analysis and optimization of structures. This starting analysis shows how integrated
interdisciplinary approaches can lead to more understanding as well as computational benefits.
Various wind turbine configurations have been studied and tried, such as horizontal axis wind turbines
(HAWTs) versus vertical axis wind turbines (VAWTs), stall- versus pitch-controlled turbines, fixed- versus free-
yaw turbines, up- versus down-wind rotors, constant versus variable rpm, high-solidity slow running versus low-
solidity fast running, stiff versus flexible, and different manners of overspeed protection. No doubt there are other
configurations not yet tried (Watson, 1989) or even conceived. A few attempts have been made to study the gains
that could be made by elastic tailoring (e.g., Karaolis et al., 1987). These studies did not lead, however, to
significant changes in the way wind turbines are designed.
There are many questions to be answered concerning the design of wind turbines. Cost is one of the primary
considerations. Some very inexpensive methods for building blades exist, but will they produce practical blades as
far as performance is concerned? For example, it is known that pultrusion can produce blades at a fraction of the
cost of present blades. One restriction of such blades, however, is that they must be spanwise uniform. Given the
present structural design of HAWTs (i.e., without elastic coupling), this is known to be an energy-inefficient
design for aerodynamic reasons. No research has been done, however, to indicate whether such losses could be
compensated for by elastic tailoring, thus making pultrusion a practical manufacturing technique. The position of
the committee is that the impact of future research on modeling and the design process will be negligible unless
configuration, aerodynamic structural design, and materials properties issues are considered together.
LOAD CHARACTERIZATION
To successfully design a wind turbine blade structure, it is necessary that a representative model of the system
loads be developed. While not necessarily being able to accurately predict these loads, this model should predict
realistic peak-to-peak stress levels and oscillatory frequencies. The mechanisms for these stress reversals are
reasonably well understood. Blades are exposed to relatively high levels of turbulence and random gusts in the air
and, in some configurations, tower shadow effects. These, along with steady and quasi-periodic aerodynamic loads
due to steady components of wind and vertical wind shear, coupled with steady centrifugal forces and periodic
gravitational forces (in HAWTs) from the rotation create distributed loads along the blades. These distributed
loads, in turn, have steady, periodic, and random components.
The steady, periodic, random loads lead to static and dynamic deformations of the blades; these deformations
also consist of steady, periodic, and random components. All these distributed spanwise loads lead to oscillating
stresses in the blades, which vary spanwise and over each cross section. Because of the strong influence of
aerodynamics and structural dynamics on the internal loads, this problem is very much interdisciplinary. One
cannot choose a material for these machines "out of context" (i.e., without considering the aeroelastic aspects of
the problem). For additional discussion of the load characterization, see de Vries (1979).
refer to the inboard section of the blade, the blade/hub attachment and pitch-control hardware, and the hub itself).
Conventional articulated helicopter rotor configurations have flap and lead-lag hinges and pitch bearings. They
also require auxiliary mechanical dampers to dissipate energy from ground resonance mechanical instability.
Serious talk of so-called hingeless and bearingless rotors began about 25 years ago, along with hopes of
eliminating the mechanical dampers. The actual development and use of such rotors, however, has been slow in
coming. Nevertheless, the relatively clean bearingless rotor designs that are now being adopted are very different
from the conventional articulated rotor designs of 10 to 20 years ago. The weight, parts count, and drag of the
rotor/hub have been significantly reduced. On many systems the mechanical dampers have been replaced by
elastomeric restraints. The number of components that have to be replaced periodically has been reduced, and the
performance and flying qualities of the newer rotorcraft are better (Ormiston et al., 1988). The new designs did
not come without cost, however, and any significant change in the way wind turbines are designed will likely
require years of research.
Helicopter rotor blades spin at a relatively high angular speed and, although flexible, are effectively stiffened
by centrifugal forces. Because of the free rotation in flap and lead-lag directions at the root in articulated rotors, a
large component of their motion is as a rigid body. Loads in these configurations are transferred through the hub
into the fuselage primarily as shear forces at the roots of the blades. Blade aeroelastic instabilities can be avoided
by properly placing the axis of mass centers of the blade. The analyses of 20 years ago used in designing these
blades were, for the most part, linear.
When hingeless and bearingless helicopter rotors began to be designed, many new problems surfaced. First,
the loads and vibration problems multiplied because of the nonnegligible hub moments. Second, a number of
newly observed and sometimes catastrophic instabilities surfaced that had to be avoided. Because of the possibility
of these instabilities, the process of designing the blades became more complicated. In fact, to accurately predict
the behavior of such a rotor blade, highly sophisticated structural analyses are required. This is due to a number of
factors:
1. Effectively cantilevering the blade root into the hub brings root stresses into the boundary value
problem that governs the blade deformation. When combined with already strong axial and lateral
loads, these stresses make the problem geometrically nonlinear. Conventional rotor analyses were
linear. Early hingeless (not necessarily bearingless) rotor analyses were nonlinear, although they
allowed only for moderate rotations due to deformation. The nonlinearity implies that even the
linearized structural dynamics or aeroelasticity analysis (typically carried out in terms of assumed
modes) must be conducted relative to a deformed state. This state can only be determined from a
nonlinear analysis.
2. In order to control the blade's pitch and reduce the root stresses, a portion of the blade near the root
must be very flexible. Because of the wide range of pitch angles through which the outboard portion
of the blade must be rotated, rotations of the inboard portion due to deformation may be large. That
is, a moderate rotation analysis may not be adequate. This also implies that even the types of assumed
modes needed for rapid convergence may be functions of the deformed state of the blade--greatly
complicating the analysis.
3. The hardware necessary to control the blade pitch introduces additional load paths (see Figure 5-3).
These load paths render the structure statically indeterminate and disallow certain simplifications that
can be made in a determinate structure.
In addition to these complications, one must also account for the use of composite materials in modern
blades. This entails several more aspects of the analysis that must be upgraded. Most blade models are beams of
the classical type in which transverse shear deformation is assumed to be
negligible. It is well known that shear deformation must be accounted for in composite blades.
Also, with blades made of isotropic materials, simple engineering beam theories will usually suffice in the
determination of beam stiffness. Except for torsional stiffness, the stiffness constants are simply integrals of
modulus-weighted geometric quantities over the cross section. The reader is reminded that these integrals fall out
from St. -Venant (interior) elasticity solutions for the beam that implicitly include the effects of sectional
deformation (warping) in and out of the cross-section plane. When shear deformation is included, solution of the
flexure problem of St. -Venant is required. This problem is of the same order of difficulty as the St. -Venant
torsional problem, and both require some sort of computer solution for irregular-section geometries.
When we generalize this to beams made from anisotropic materials, all of the stiffness constants depend on
the in- and out-of-plane warping in complicated ways; and, in order to determine the equivalent-beam elastic
constants, it is necessary to solve the St. -Venant (interior) problem for in-and out-of-plane warping with the
recognition that these warping displacements are coupled. Only for single-celled, thin-walled sections can this be
done with sufficient accuracy without a computer. For further discussion of this, see Giavotto et al. (1983) and
Atilgan and Hodges (1990). This implies that in order to analyze composite blades, it is necessary to undertake two
separate analyses: one to get the section properties and one to use them in a response analysis. In addition, if
stresses over the blade section are desired, additional information concerning the stress distribution for the
applicable types of loading may need to be derived. See Figure 2-1.
There are a variety of computer-based blade analysis methods in existence that are applicable to the wind
turbine problem (Table 2-1). (See, for example, Rehfield, 1985 [analytical]; Bauchau, 1985; Kosmatka, 1986;
Borri and Merlini, 1986; Wörndle, 1982, and Giavotto et al., 1983 [two-dimensional or quasi-three-dimensional];
and Lee and Stemple, 1987 [three-dimensional]). The quasi-three-dimensional methods are solved by
discretization of the cross
Figure 2-1
Process necessary to analyze composite blades.
section plane by finite elements in order to obtain the properties. (The term quasi-three-dimensional indicates that
the axial coordinate is handled analytically.) They are far more efficient than a three-dimensional finite element
model would be. Unfortunately, however, most of these methods do not treat all aspects of the problem, and some
of these methods are too complex for a personal computer (PC).
TABLE 2-1 Blade Sectional Analysis Codes
Principal Investigator(s) Name of Code Restrictions Country of Origin Analysis Type
Bauchau - no in-plane warping; thin- U.S. 2-D finite element
walled (quasi-3-D)
Giavotto, Borri, et al. ANBA no restrained warping Italy 2-D finite element
(quasi-3-D)
Kosmatka - uniaxial stress field; no U.S. 2-D finite element
restrained warping
Lee - no in-plane warping; U.S. 3-D finite element
simple-section geometry
Rehfield/Nixon TAIL no in-plane warping; thin- U.S. analytical
walled; uniaxial stress
Wörndle - uniaxial stress field; no Germany 2-D finite element
restrained warping
Rehfield's method, the simplest of these methods, was programmed for a PC by Mark Nixon (NASA,
Langley). Although Rehfield's analysis takes restrained warping into account, Nixon's code does not. Explicit
treatment of in-plane warping (Poisson contraction and anticlastic deformation) is circumvented by the uniaxial
stress assumption. However, Rehfield does not consider initial twist and curvature, which are important for wind
turbines. Also, the shear stiffnesses obtained by this code are not sufficiently accurate because of the neglect of
out-of-plane St. -Venant flexural warping.
Bauchau, Kosmatka, Lee, and Wörndle have codes that are more general than Nixon's, but they are also
considerably more complex. The committee knows of no industrial users of Kosmatka's and Lee's codes in the
United States. Wörndle's code was developed in Germany and, although it is used by the German helicopter
company Deutsche Aerospace (previously MBB), the committee knows of no users in the United States.
Bauchau's code is used in the U.S. helicopter industry. It accounts for restrained warping and initial twist and
curvature. It is restricted, however, to the thin-walled case and yields results that are comparable to those of
Rehfield's analysis (Bauchau et al., 1987).
The code of Giavotto et al. (1983) (called ANBA) was developed in Italy. Although it is the most powerful of
all these codes, it does not account for
restrained warping (probably not an important limitation for wind turbine blade sections). It is also considerably
more computationally intensive than the others, requiring 20 minutes of CPU time (for large problems) on a MAC
II running A/UX. It is commercially available in the United States. While it is evident that there are PC-based
tools for determining blade elastic constants, to date there are no codes that incorporate the beam elastic constants
into a fully coupled aeroelastic response code for wind turbine blades. The distinction here is very important, and
both the determination and the incorporation must be compatible in kinematical and geometrical assumptions
(e.g., if pretwist is to be taken into account, then both the sectional and response analyses must include its effects).
For additional references on the subject of composite blade modeling, see Hodges (1990).
These couplings are very important in the response analysis. Modern fixed-wing aircraft designs have made
use of the properties of composites to tailor the structure so that certain performance or stability criteria are met or
enhanced. An example of this is the X-29 forward swept wing aircraft. Without composites, design of this airplane
would not have been possible (see Shirk et al., 1986, for a review of tailoring). With both helicopter and tilt-rotor
blades, various investigators have proposed these types of passive tailoring schemes to increase stability margins,
decrease weight, avoid high stresses in some localities, and increase efficiency.
The chief mechanisms for tailoring in these contexts are couplings between bending and twist deformation
modes and between extension and twist deformation modes. For example, Ormiston et al. (1976) incorporated two
coupling parameters in a rotor blade, one of which is equivalent to a bending-twist coupling and the other to an
elastic coupling between the flap and lead-lag directions. They showed that one could then passively extract some
of the damping coming from the air that is naturally present in the heavily damped flap mode and put it where it is
needed--in the otherwise weakly damped lead-lag mode. Theoretical and experimental work indicated that the
lead-lag damping could be increased by an order of magnitude without noticeably affecting the flap damping.
Current wind turbine rotor blades are quite stiff and heavy relative to helicopter blades. Part of the reason for
this is to prevent tower strikes. Given that wind turbine rotors must carry large edgewise gravity bending
moments, which argues for using their large planform dimensions to actively carry those moments, it may be that
hinges are the most practical way to get large pitch effects with such large and stiff rotor structures. Flap bending/
twist coupling could provide small outboard blade angle changes to help a constant-rpm machine retain optimum
efficiency over the low wind range, but variable rpm provides a much more effective and versatile way of doing
this, and it is on the near horizon. Using extension-twist coupling to provide overspeed limitation for fixed-pitch
rotors may be the most potent near-term use of elastic tailoring, as it could simplify rotor construction while
removing a lot of rotor cost and also be much more reliable than current mechanical latch mechanisms.
Determining whether wind turbine blades can be made lighter, less expensive, with lower loads (i.e., with
longer life), with performance equal to or better than present blades, and free from instabilities should be the goal
of any basic and applied research directed toward the wind turbine problem. Whatever elastic tailoring schemes
are proposed, however, must not compromise on total (manufacturing, installation, and maintenance) cost per year
of life. In order to speed up the incorporation of these technologies (along with active control as appropriate), the
designer must have access to computational tools. If active control is to be considered for achieving these gains
(see Chapter 6), the control design should be undertaken in combination with the passive measures.
RECOMMENDATIONS
The following areas are recommended for future research:
1. Simple cross-sectional analyses and codes need to be developed for determination of the sectional
elastic constants. To accomplish this in an accurate and computationally efficient manner will require
additional research. The resulting tools may likely be in the spirit of Rehfield/Nixon but with
additional wind turbine parameters included, such as initial curvature, initial twist, and taper and with
additional attention given to obtaining shear stiffnesses more accurately and accounting for restrained
warping. Although ANBA, for example, already deals with all these effects, a Rehfield/Nixon-type
code has the advantage of requiring very little CPU time on a PC. Bauchau's code may already meet
most of these requirements; what is unknown to the committee is the level of complexity and PC CPU
time that the wind turbine industry will deem simple enough.
2. Aeroelastic response analyses need to be developed for the wind turbine problem that incorporate the
appropriate types of elastic coupling derived in the analyses mentioned above for the first
recommendation. These could be simple modal analyses making use of a few natural modes and a few
perturbation modes as described by Bauchau and Liu (1989). (Note that modal analyses and finite
element analyses are not mutually exclusive. Modal analyses can be based on finite element models.)
The code need only contain a representative model of aerodynamics used to characterize the load
spectrum for design purposes. Obtaining this type of aerodynamic model is vital. This itself would
also require additional research.
3. Formal optimization procedures need to be employed in the systematic investigation of the effects of
certain aerodynamic, configuration, structural, and control parameters on the overall design. This will
require the development of complete simulation tools that are compatible with formal optimization.
Karaolis, N. M., P. J. Musgrove, and G. Jeronimidis. 1987. Passive Aerodynamic Control Using Composite Blades. Proceedings of Workshop
on the Use of Composite Materials for Wind Turbines. ETSU-N-109, Harwell, England, November 4, pp. 71-101.
Kosmatka, J. B. 1986. Structural Dynamic Modeling of Advanced Composite Propellors by the Finite Element Method. Ph.D. dissertation,
University of California, Los Angeles.
Lee, S. W., and A. D. Stemple. 1987. A Finite Element Model for Composite Beams with Arbitrary Cross-Sectional Warping. Proceedings of
the 28th Structures, Structural Dynamics and Materials Conference. AIAA Paper No. 87-0773, Monterey, California, April 6-8, pp.
304-313.
Ormiston, R. A., W. G. Bousman, D. H. Hodges, and D. A. Peters. 1976. Hingeless Helicopter Rotor with Improved Stability. United States
Patent 3,999,886. December 28.
Ormiston, R. A., W. G. Warmbrodt, D. H. Hodges, and D. A. Peters. 1988. Rotorcraft Aeroelastic Stability (Army/NASA Research
1967-1987). NASA Conference Publication 2495, Vol. l, pp. 353-529.
Rehfield, L. W. 1985. Design Analysis Methodology for Composite Rotor Blades. Presented at the 7th DoD/NASA Conference on Fibrous
Composites in Structural Design. AFWAL-TR-85-3094, Denver, Colorado, June 17-20, pp. (V(a)-1)-(V(a)-15).
Rehfield, L. W., and A. R. Atilgan. 1987. Analysis, Design and Elastic Tailoring of Composite Rotor Blades. Final Report, Grant No. NAG-
l-638. U.S. Army Aerostructures Directorate, Langley Research Center.
Rehfield, L. W., A. R. Atilgan, and D. H. Hodges. 1990. Nonclassical Behavior of Thin-Walled Composite Beams with Closed Cross
Sections. American Helicopter Society, Vol. 35, No. 2, pp. 42-50.
Shirk, M. H., T. J. Hertz, and T. A. Weisshaar. 1986. Aeroelastic Tailoring--Theory, Practice, and Promise. Journal of Aircraft, Vol. 23, No. 1,
pp. 6-18.
Stoddard, F., V. Nelson, K. Starcher, and B. Andrews. Horizontal Axis Wind Turbine (HAWT) Elastic Twist Determination. Final Report.
SERI Contract RL-6-06013.
Watson, R. 1989. Space Frame Wind Turbine. Ninth ASME Wind Energy Symposium. New Orleans, January, pp. 93-99.
Wörndle, R. 1982. Calculation of the Cross Section Properties and the Shear Stresses of Composite Rotor Blades. Vertica, Vol. 6, pp. 111-129.
Wörndle, R., and H. Mang, 1984. Zur Schubspannungs verteilung und Schubsteifigkeit bei querkraftbeanspruchten, inhomogenen Querschitten
beliebiger Form aus orthotropen Werkstoffen. Ingenieur-Archiv, Vol. 54, pp. 25-42.
Wright, A. D., B. L. Buhl, and R. W. Thresher. 1988. FLAP Code Development and Validation, SERI/TR-217-3125, Solar Energy Research
Institute Report, Golden, Colorado.
3
MATERIALS PROPERTIES AND LIFE PREDICTION
The characteristics that make composites, especially glass fiber-reinforced and wood/epoxy composites,
suitable for wind turbine blades are low density, good mechanical properties, excellent corrosion resistance,
tailorability of material properties, and versatility of fabrication methods. Although glass/vinylester and glass/
polyvinyl composites based on hand lay-up have been the most widely used materials so far, many more types of
fibers and resins have become available recently. The new carbon fibers are stronger and stiffer, while the new
resins provide higher toughness and shorter process cycle time. A number of handbooks are now available to the
designers of composites (Lubin, 1982; Engineered Materials Handbook, 1987; Composites & Laminates, 1987).
FIBERS
The most commonly used and lowest-priced fiber is E-glass fiber. Over the past several years, however, many
new fibers have become available. The commercially available fibers and their typical properties are listed in
Table 3-1.
While E-glass fiber is most widely used in wind turbine rotor blades mainly because of its low cost, carbon
fibers are the fibers of choice in many aerospace applications. Although more expensive, they provide higher
specific modulus and specific strength than glass fibers. The advantage of carbon fibers is further enhanced in
fatigue. However, carbon fibers are electrical conductors, and their contact with metals may lead to corrosion of
the latter. Polymeric fibers such as aramid and high-density polyethylene are the toughest of all the available fibers
and hence can be used where high-impact resistance and toughness are required. These polymeric fibers, however,
are weak in compression because of the fibrillar nature of their microstructure. Recently, a variety of ceramic
fibers such as alumina and silicon carbide have emerged mainly as reinforcements for metal and ceramic matrices.
These ceramic fibers have better oxidation resistance in high-temperature applications than carbon fibers.
However, they are still more expensive than most carbon fibers. Mechanical properties of epoxy matrix
composites made with the four most widely used fibers--aramid, carbon, E-glass and S-glass--are shown in
Table 3-2. Tensile fatigue behaviors of the first three composites are compared in Figure 3-1.
Since one type of fiber does not have all the desired properties, different fibers can be mixed to make a hybrid
composite. For example, in a glass/carbon hybrid composite the glass fiber can improve the impact resistance
while keeping the cost down, and the carbon fiber can provide the required strength and stiffness with less weight.
The weight savings resulting from the use of a hybrid composite reduces the load on the blade and hence will lead
to a longer lifetime. Furthermore, the material savings realized can also compensate partially for the higher cost of
the carbon fiber.
To further illustrate the cost differences based on fibers, consider a hybrid composite with a carbon/glass
volume ratio vc/g. The ratio of the cost of the hybrid composite to an all-glass composite to provide the same
structural stiffness is given by
The parameters in Equation (1) are defined as follows: $, price per unit mass; s, specific gravity; E, Young's
modulus; g, glass; and c, carbon. On the other hand, the weight ratio, defined as the weight of the hybrid
composite divided by that of the all-glass composite, is given by
E-glass fiber costs around $2/lb; a high-strength carbon fiber costs about $30/lb. Thus, an all-carbon fiber
composite costs 3.5 times more than an all-glass fiber composite to provide the same structural stiffness at a
weight savings of 76 percent. When a 50/50 glass/carbon hybrid composite is used, however, the calculated cost
ratio is reduced to 2.9 with a weight savings of 58 percent.
Figure 3-1
Trends of longitudinal tensile fatigue S-N data for unidirectional composites with various fibers.
Source: Mandell (1990).
As can be inferred from Table 3-2, the in situ tensile failure strain of E-glass fiber is as high as 2.5 percent,
whereas it is only 1 percent for carbon fiber. However, E-glass fiber has a much lower fatigue ratio than carbon
fiber, that is, 0.3 versus 0.75 at 10 million cycles (Figure 3-1). Therefore, both fibers have the same fatigue strain
of 0.75% at 10 million cycles. Beyond 10 million cycles, however, the carbon fiber is expected to outlast the glass
fiber. The cost ratio to obtain the same long-term fatigue strength is then at least the same as that needed to obtain
the same stiffness. An additional benefit is that a hybrid composite blade will have a longer lifetime because of the
reduced fatigue load resulting from the weight savings. Thus, a careful study is needed to explore the full benefits
of a hybrid composite blade from a life-cycle point of view.
Fibers are used in various forms. The commonly used fiber preforms include unidirectional tow, woven
cloth, knitted fabric, continuous strand mat, chopped strand mat, and braid as well as chopped fibers in sheet and
bulk molding compounds. Depending on the application and manufacturing process used, one fiber preform may
be preferred to others. The various fiber preforms are shown schematically in Figure 3-2.
Where high strength is required, unidirectional bundles of fibers known as tows should be used. Woven
cloths and knitted fabrics are easier to use, especially over complex contours. Woven cloths have disadvantages in
that
Figure 3-2
Schematics of various fiber preforms.
Source: Chou et al. (1986).
high fiber volume content cannot be attained and the inherent fiber cross-overs are susceptible to premature
fatigue damage. The fiber cross-overs also result in low compressive strength. While continuous strand mats
provide good strength, chopped strand mats are weaker because of the fiber discontinuity. Where delamination is a
prime concern and strengthening is required in the thickness direction, braids can work well.
MATRIX MATERIALS
The matrix resins commonly used in wind turbine blades are divided into three major classes of thermosetting
polymers: unsaturated polyesters, epoxies, and vinyl esters (Lee and Neville, 1967; Launikitis, 1978; Strong,
1989). The unsaturated polyester resins, typically based on the orthophthalic or isophthalic acid, are most widely
used because of short cure time and low cost. Despite the long cure time required and higher cost, epoxy resins are
gaining greater acceptance due to their superior chemical resistance, good adhesion, low cure shrinkage, good
electrical properties, and high mechanical strength. A reasonably good compromise between cost, cure time, and
the above-mentioned physical properties is achieved in epoxy-based vinyl ester resins, which have especially
shown very rapid growth over the past few years by filling the gap between polyesters and epoxies. The use of all
these resins as composite matrices is generally subjected to a temperature limit around 150° to 200°C. The upper
temperature limit is lower than the glass transition temperature Tg because above Tg the resin becomes soft and
rubbery, thus losing most of the load-carrying capability.
Polyesters, as their name implies, are resins in which recurring ester linkages are an integral part of polymer
chain backbone. The most commonly used system--isophthalic polyester resins--is more expensive but is
chemically more stable and a little less brittle than orthophthalic resins. The term unsaturated comes from the
presence of carbon-carbon double bonds in the polyester chain backbone that provide the location for cross-
linking, thereby eventually leading to a tight network. The most common type of cross-linking agent is styrene.
The styrene also lowers the initial viscosity of the polyester resin to improve processing of the composite by
facilitating impregnation and wetting of fibers. The cross-linking reaction, which is addition polymerization, is
triggered by an initiator (sometimes erroneously called a catalyst), such as organic peroxide, that produces free
radicals.
Many polyester resins are cured at room temperature with a rise in temperature due to the exothermic nature
of the reaction. In this case, the typical cure time is several hours or overnight. However, with a proper selection
of initiator, the cure reaction can be carried out at an elevated temperature in an extremely short time. Experience
in automotive composites has shown that the curing of an unsaturated polyester can be completed in 2 to 3 minutes
at 150°C with the use of a tert-butyl perbenzoate initiator.
Epoxy resins derive their name from the epoxide ring structure that serves as the principal cross-linking site.
Although various types of epoxy resins are produced commercially, so-called diglycidyl ether of bisphenol-A
(DGEBPA)-type resins (e.g., Shell EPON828), have achieved the widest market acceptance and demonstrated
versatility. For applications requiring special properties such as high modulus or higher use temperature, other
types of epoxy resins with higher functionality are used in place of difunctional DGEBPA-type resins. For
instance, tetrafunctional epoxy resins of tetraglycidyl methylene dianiline (TGMDA) (e.g., Ciba-Geigy MY720)
exhibit improved properties at elevated temperature. Greater strength and improved properties at elevated
temperature are also achieved in the resins of epoxidized phenolic novolacs (e.g., Dow DEN438) or tetraglycidyl
ether of tetrakis hydroxyphenyl ether (e.g., Shell EPON103). All these resins tend to have lower fracture
toughness than DGEBPA-type resins.
For polyfunctional epoxy resins, cross-linking or cure is effected by use of polyfunctional curing agents or
hardeners. The use of a hardener with higher functionality (e.g., tertiary amine versus primary amine) allows a
greater cross-link density of cured resins, which generally improves their physical properties. Lewis acids, such as
BF3, are catalytic curing agents
that promote the cure reaction but do not themselves serve as direct cross-linking agents. These catalytic curing
agents have the advantage of very long shelf life and require significant heat to initiate the reaction. The resultant
structure of Lewis acid-cured resin is very tightly cross-linked. Curing of most epoxy resins and composites is
done at 120° or 177°C with a typical cycle time of 2 to 3 hours under the molding pressure of up to 1.4 MPa.
The term vinyl ester can be applied to any number of chemical compounds comprising an ester linkage and
terminal unsaturation. There are several types of vinyl esters based on epoxy resins as well as nonepoxy resins.
However, to the composites industry, the vinyl ester resins usually mean methacrylate esters of epoxy resins
(Launkitis, 1978). Unlike polyesters, vinyl esters do not possess internal unsaturation. However, vinyl esters and
polyesters are similar in that they both utilize a coreactant or cross-linking agent, such as styrene, and free
radical-producing initiators, such as peroxides, to effect cure. As a result, vinyl ester resins can be cured in a very
short time like polyesters, but their static strength and modulus properties are similar or comparable to those of
epoxy resins.
Typical mechanical properties of polyesters, epoxies, and vinyl esters are shown in Table 3-3. While epoxy
resins are at the top of the scale as far as mechanical properties are concerned, they require the longest cure time
and are the most expensive. For example, epoxy resins cost about $1.80/lb, whereas unsaturated polyesters cost
only $1.00/lb. Vinyl ester resins fall in between these two resins, costing about $1.60/lb. Thus, the final selection
of a resin should consider the material cost as well.
The fatigue crack propagation rates of these resins vary with the stress intensity factor range K in the case
of other plastics for structural use (Hertzberg and Manson, 1980). In amine-cured DGEBA epoxy resins, the
values for exponent m in the equation for the crack growth per fatigue cycle, da/dN = const.( K)m, range from 7.7
to 20, which are higher than those for other plastics. In general, a lower fatigue crack growth rate and a higher
fracture toughness are observed with increasing molecular weight between cross-links.
As discussed previously, all of these resins suffer the problem of brittleness. A brittle resin results in
premature matrix cracking in the composite, which in turn facilitates moisture ingress. Although fracture
toughness of the resin systems can be raised by increasing Mc or adding diluents, these approaches result in
lowering of modulus and temperature resistance (Hertzberg and Manson, 1980; Owen, 1974; Christensen and
Rinde, 1979). At present, many commercially available resins utilize toughening agents in the form of discrete
particles of elastomers or ductile
TABLE 3-3 Properties of Cast Resins
Polyester Vinyl Ester Epoxy
Specific gravity 1.10-1.46 1.1-1.2 1.2-1.3
Flexural strength, MPa 60-160 120-140 110-215
Tensile strength, MPa 40-90 70-90 50-130
Compressive strength, MPa 90-200 - 110-210
Tensile elongation, % <5 <6 <9
Modulus, GPa 2-4 3-4 3-4.5
thermoplastics. In this case the increase of fracture toughness is achieved with a relatively small change of
modulus and temperature resistance.
Toughening of epoxy resins by the inclusion of elastomer particles is well established. Although several
different types of liquid or solid elastomers or their hybrids can be used, the most effective way of increasing resin
toughness is by using a liquid rubber with specific terminal functional groups (Lee, Riew, and Moulton, 1980;
Bucknall, 1977; Riew, Rowe, and Siebert, 1976; Scott and Phillips, 1975; McGarry and Sultan, 1969; Bascom,
Bitner, Moulton, and Siebert, 1980; Moulton and Ting, 1981; Lee, 1986). Examples are butadiene acrylonitrile
copolymers with carboxyl, amine, or vinyl groups at both ends of the chain which are selected depending on the
type of curing agent and curing mechanisms.
When carboxyl-terminated butadiene acrylonitrile copolymer (CTBN) is mixed in liquid state with an
amine-cured DGEBPA epoxy resin, their initial compatibility in liquid state followed by phase separation in the
solidification process leads to the formation of submicron-sized elastomer particles that are well bonded to the
surrounding resin. As a result, adding as small as 5 wt% of CTBN to Epon 828 resin increases the fracture
toughness G1c from 350 to 5260 J/m2. Yet the accompanying decrease of tensile modulus is only 2.5 to 2.8 GPa
(Riew, Rowe, and Siebert, 1976). A similar type of elastomer toughening in extremely brittle TGMDA resins (G1c =
80 J/m2) is less effective: a maximum twofold increase with considerable reduction of modulus (Lee, Riew, and
Moulton, 1980).
The use of elastomer-toughened epoxy resins as matrix materials has been shown to improve the interlaminar
fracture toughness of fiber-reinforced composites, particularly those with woven fabric reinforcement with resin-
rich areas between the plies (Bascom, Bitner, Moulton, and Siebert, 1980). Despite a high level of constraint on
thin films of resin matrix by densely packed surrounding fibers (Scott and Phillips, 1975), toughening of epoxy
matrices also increases the resistance of composites against local damage initiation and accumulation (McGarry
and Sultan, 1969; Bascom, Bitner, Moulton, and Siebert, 1980; Moulton and Ting, 1981; Lee, 1986). However,
their effects on fatigue lifetime or fatigue endurance limit of composites have not been fully confirmed.
In the case of toughening of unsaturated polyester resins, both liquid elastomers with terminal functional
groups and ductile thermoplastics are utilized as a secondary phase (McGarry, Rowe, and Riew, 1978; Lee,
Howard, and Rowe, 1983). Examples are vinyl- or epoxy-terminated butadience acrylonitrile elastomers,
hydroxyl- or vinyl-terminated epichlorohydrin elastomers, and polyvinylacetate. Compared with elastomer-
toughened DGEBPA epoxy resins, the size of elastomer particles formed in situ during cure is much bigger
(micron level) because of lower compatibility of the reactive mixture. As a result, the effectiveness of elastomer
toughening of polyester resins is much lower than that of DGEBPA epoxy resins. For instance, by adding 8 wt%
of vinyl-terminated polyepichlorhydrin elastomer, the fracture toughness G1c of isophthalic polyester resin is
increased from 60 to 110 J/m2 with a 23 percent reduction of tensile modulus. However, the use of elastomer-
toughened unsaturated polyester resins in short fiber-reinforced composites increases the local damage resistance
under impact and, in certain cases, that of tensile strength as well (McGarry, Rowe, and Riew, 1978; Lee,
Howard, and Rowe, 1983).
In addition to thermosetting resins, many engineering thermoplastics are available now as matrix resins.
Table 3-4 lists those new high-performance thermoplastic resins developed mainly for aerospace applications
(Witzler, 1988). Compared with thermosetting resins, the thermoplastic resins offer advantages in prepreg (pre-
impregnated laminate) stability and short processing cycle. However, they tend to be weak in solvent resistance
and fiber impregnation. Trade-offs between thermosets and thermoplastics as matrices are shown in Table 3-5.
Unidirectional prepregs also are available with commodity thermoplastics such as nylon and PET. Although
they are cheaper than the high-performance thermoplastic prepregs, their full potential has not been realized
because of the lack of low-cost manufacturing techniques.
aTg is the glass transition temperature, the temperature at which a polymer changes from a rigid glassy solid to a soft rubbery solid.
bPrepreg.
Source: Johnston and Hergenrother (1987).
The difficulty associated with fiber impregnation has led to the development of commingled yarns and fabrics
as well as powder-coated yarns and fabrics. A commingled yarn consists of both reinforcing fibers and
thermoplastic fibers. During processing the thermoplastic fibers melt and impregnate the reinforcing fibers
(Lynch, 1989). Thermoplastic powder is smaller in diameter than the corresponding thermoplastic fiber and hence
is more evenly distributed in the yarn, which facilitates a more uniform fiber impregnation (Hartnes, 1988).
E-GLASS/PLASTIC COMPOSITES
E-glass/plastic composites, commonly called glass-reinforced plastics (GRPs), have been widely used in the
manufacture of blades of various sizes. Typical resins used in GRPs are polyester, vinyl ester, and epoxy.
Although polyester resins can be cured in the shortest time, they show large shrinkage and hence may not be
appropriate for use in a hot, dry environment such as California, where moisture-induced swelling would be
minimal (Windpower Monthly, 1987). Vinyl ester resins have good environmental stability and are widely used in
marine applications. Epoxy resins have good mechanical properties and dimensional stability. Their drawback is
longer cure time and higher cost; however, new epoxy resins are now available for pultrusion and resin transfer
molding, which require fast curing.
GRP is the main structural material for a number of blades of large machines and is also used as cladding
over steel load-bearing frames in others (Phillips et al., 1987). GRP is the most popular blade material used in
medium-size machines in Denmark and The Netherlands. As of January 1987, approximately 81 percent of 15,059
wind turbines in the California wind farms had fiberglass rotor blades (Stoddard, 1989; Modern Power Systems,
1986).
Typical properties of a unidirectional E-glass/epoxy composite are compared with those of other
unidirectional composites in Table 3-2. For all the composites in the table, the stress-strain behavior is usually
quite linear except in transverse compression and in-plane shear. Another exception is thearamid/epoxy
composite, which is quite nonlinear in longitudinal compression also. Elastic properties of multidirectional
laminates can be calculated from unidirectional properties using laminated plate theory (Tsai, 1989; Jones, 1975).
Tensile stress-strain relationships of several representative E-glass/epoxy laminates are shown in Figure 3-3.
When a multidirectional
laminate, such as the quasi-isotropic laminate in the figure, is subjected to tension, cracks appear first in the matrix
and fiber/matrix interfaces of off-axis plies (Tsai and Hahn, 1975; Reifsnider, 1980). Most of these cracks, called
ply cracks, are entirely through the thickness of individual off-axis plies. The crack density increases with
increasing load until some of the cracks grow as delamination between plies with different fiber orientations.
Further increase of load breaks the fibers in the on-axis plies, leading to failure of the laminate as a whole. When
the load is along the 0° direction, 90° plies may fail at a strain as low as 0.3 percent, although the laminate does
not fail until 2.3 percent strain is reached. Thus, failure of a multidirectional laminate is preceded by matrix/
interface cracking in the off-axis plies and also by delamination between plies with different fiber orientations.
The ply cracking in general does not lead to immediate failure of the laminate. However, it facilitates
moisture diffusion through the cracks, thereby reducing the durability of the laminate. Even in the absence of
cracks, epoxy resins can absorb as much as 7 percent moisture by weight. The absorbed moisture plasticizes the
resin, thereby reducing its glass transition temperature. When fully saturated, the glass transition temperature of
epoxy resins can be reduced by as much as 100°C (Vinson, 1977).
At room temperature, absorbed moisture has a rather minimal effect on mechanical properties. When
combined with elevated temperature, however, it can seriously degrade mechanical properties through interfacial
debonding and matrix plasticization. The worst combination is thermal spiking to a temperature above the glass
transition temperature in the presence of moisture. The thermally spiked specimens show increased moisture
absorption and lower strengths (Springer, 1981).
Figure 3-3 Stress-strain relationships of glass/epoxy laminates under uniaxial tension (50 vol%). (LONG: strain
parallel to loading direction; and TRANS: strain [magnitude] normal to loading direction)
To provide a smooth surface and also to prevent moisture diffusion, a gel coat is frequently applied on the
surface of the blade. However, moisture still can diffuse through the gel coat and collect at the gel coat/composite
interface, resulting in blistering. Thus, the benefits of a gel coat as a means of preventing moisture infusion are not
clear (Phillips et al., 1987). Also, it has been found that E-glass/epoxy blades are prone to crazing in the surface
finish. In some blades, several small delaminations have been found under the outer covering of Nexus cloth, and
voids were present under the outer layer of fiberglass roving. Although these damages are not critical
immediately, they may facilitate water ingression, accelerating deterioration of the composite.
The tensile fatigue stress-lifetime (S-N) data of various types of composite laminates can be represented by an
equation of the form (Mandell, 1990),
where S is the maximum fatigue stress, UTS is the ultimate tensile strength, and N is the number of cycles to
failure. This is the form of the curve in Figure 3-1. The parameter B represents the fraction of UTS lost per decade
of fatigue cycles. For E-glass/epoxy laminates, B is about 0.1 regardless of whether the laminate is unidirectional
or multidirectional when the fatigue stress ratio R = 0.1 (Mandell, 1990). Since very little fatigue data exist beyond I
million cycles, fatigue strength at 100 million cycles, which is a minimum requirement for wind turbine blade
design, is estimated to be 20 percent of the ultimate strength. In terms of strain the fatigue strain at 100 million
cycles is only 0.5 percent although the static ultimate strain is as high as 2.5 percent. In comparison, aramid fiber/
epoxy composites lose about 6 percent of UTS per decade, while carbon fiber composites lose only 3 to 4 percent
of UTS. Thus, a carbon fiber composite enjoys a higher fatigue strain of 0.7 percent, although it has a much lower
ultimate tensile strain than glass/epoxy composites.
The failure sequence of multidirectional laminates under fatigue is quite similar to that under static tension
when the load is replaced by the number of fatigue cycles. As schematically shown in Figure 3-4, the damage is
initiated in the form of cracking of off-axis plies across their thickness. As fatigue proceeds, the crack density
grows rapidly, reaching a plateau. Thereafter, the ply cracks grow along the ply interfaces, leading to
delamination. Much of the fatigue life is spent in the delamination stage between the ply cracking and fiber failure
stages (Reifsnider et al., 1983). The ply cracking stage is especially short in glass composites because the
transverse failure strain is much lower than the longitudinal failure strain.
The ply cracking and delamination degrade compressive strength more than tensile strength when the fibers
used in the composite are not so fatigue sensitive. Delamination facilitates buckling and hence reduces
compressive strength substantially. This is the reason that composite laminates fail in compression under fully
reversed tension-compression fatigue. In fact, the maximum fatigue stress at R = -1 is only about half the maximum
fatigue stress at R = 0.1 (Ven Delft et al., 1987; Bach, 1988). Also, composite laminates are weaker in
compression-compression fatigue than in tension-tension fatigue, especially in the low-cycle region.
Another consequence of ply cracking and delamination in multidirectional laminates is reduction in stiffness.
For example, the fatigue failure of a quasi-isotropic glass/epoxy laminate at 10 million cycles (R = 0.1) can be
accompanied by a stiffness reduction as large as 35 percent (Zweben et al., 1989). Thus, design for fatigue may
have to be based on stiffness reduction rather than strength reduction if maintaining sufficient stiffness is more
critical.
Figure 3-4
Modes of damage growth in composite laminate under fatigue.
Source: Reifsnider et al. (1983).
may be determined using a rainflow counting algorithm, with each cycle being characterized by a mean
fatigue stress and an alternating stress (Murtha-Smith, 1985; Downing and Socie, 1982; Morgan et al., 1989).
Once the fatigue cycles are defined, the lifetime is predicted by using a combination of Miner's rule and constant-
amplitude fatigue data or a crack growth law (Sutherland, 1989).
It is difficult, however, to predict lifetimes of composite wind turbine blades because their failure modes are
complex and no sufficient database exists. Fatigue failure of a composite laminate depends on the specific lay-up
and style of material used. Therefore, the existing data are not likely to be directly applicable to the chosen
laminate. Furthermore, although the blade is subjected to multiaxial, variable-amplitude loading that involves
several different fatigue stress ratios, most available data are for uniaxial, constant-amplitude tensile fatigue.
Therefore, the current fatigue design practice is to generate constant-amplitude fatigue data for a laminate
simulating the actual blade lay-up and then use the metals methodology to predict the lifetime under operational
loading (Murtha-Smith, 1985).
Several unique characteristics of composite laminates should be recognized in applying the metals life
prediction methodology. First, composite laminates exhibit different hysteresis loops than metals. While the
unloading modulus does not change much with fatigue cycles in metals, it decreases with increasing fatigue cycles
in composite laminates as a result of ply cracking and delamination. Also, since composite laminates are weaker in
compression, especially under fatigue, the maximum alternating stress on a Goodman curve tends to be located in
the positive mean-stress quadrant. Implications of these differences in the application of the metals life prediction
methodology have not been examined in detail.
The first appearance of ply cracking in a laminate can be predicted from the fatigue behavior of constituent
plies using the laminated plate theory together with a failure criterion (Zweben et al., 1989). After a ply fails, the
state of stress changes. This stress redistribution is taken into account when the next failure event is predicted
using another failure criterion. This process continues until the laminate fails as a whole.
Although this life prediction procedure is straightforward conceptually, its implementation is rather difficult
because of the complex stress analysis involved and the lack of sui criteria for various failure events. An alternate
approach is to use residual strength to describe the overall degradation of the composite. Basically, the rate of
change of residual strength is postulated as a function of fatigue stresses, similar to a crack growth law. Fatigue
failure is then assumed to occur when the residual tensile/compressive strength is reduced to the maximum/
minimum fatigue stress. The parameters involved are determined by fitting the model prediction to the
experimental S-N data. The scatter in fatigue life is related to the scatter in static strength through the assumption
of similitude that a specimen with a higher static strength also has a longer fatigue lifetime. Variations of this
residual strength approach have been used by several investigators to analyze laboratory fatigue data (Hahn, 1979;
Sendeckyj, 1981; Yang and Shany, 1983; Reifsnider, 1980; Hwang and Han, 1989). These residual strength
models can predict lifetimes under spectrum loading on the basis of constant-amplitude fatigue data. Thus, they
can be used instead of Miner's rule, which has been found to be inaccurate for some composites (Bowen et al.,
1984).
While the prediction of final failure is mostly phenomenological, ply cracking and delamination, which are
two of the major subcritical failure modes, have been analyzed mechanistically. A damage growth law similar to
the crack growth laws for homogeneous materials can be used to describe the multiplication of ply cracks (Han
and Hahn, 1989) and the delamination growth (Johnson, 1985) under controlled conditions of loading and
specimen geometry. The subcritical damages may degrade the strength of the main load-bearing plies, thereby
accelerating the final failure and reducing the stiffness (Reifsnider and Stinchcomb, 1986). Reduced stiffness may
change the loading on the blade and eventually cause premature failure. Thus, successful design of blades requires
the ability to predict not only the final failure but also the growth of subcritical damages and the resulting stiffness
reduction.
Composite structures are sensitive to out-of-plane loads, which can cause delamination (Whitehead, 1990).
These loads are present especially at joints and impact-induced damage areas. While in-plane loads are effectively
carried by fibers, out-of-plane loads place an unexpectedly heavy burden on the matrix-controlled strength.
Unfortunately, the matrix-controlled strength is sensitive to the presence of defects such as porosity and
delamination. Since geometrically complex regions tend to contain a high level of porosity while being subjected
to out-of-plane loads, these regions of the blade must be identified and tested for long-term fatigue performance
under realistic loading conditions and environments. At the same time, realistic impact energy levels should be
established, and their effect on fatigue lifetime should be investigated.
As mentioned earlier, carbon fibers, unlike glass fibers, are almost insensitive to fatigue degradation.
Experience has shown that fatigue is not a problem in aircraft wings and fuselage structures, where carbon fiber
composites are used, if the maximum strain is kept below 0.6 percent.
Therefore, the same static knockdown factor can be used as well for fatigue strength to account for the effect
of defects (Whitehead, 1990). It is suggested that the current unfavorable economics should not preclude the use
of carbon fibers together with glass fibers in wind turbine blades to avoid fatigue problems.
TOUGHNESS CONSIDERATIONS
Wind turbine blades are expected to benefit from the use of tougher composites in several ways. One obvious
benefit is the reduction in matrix cracking and hence a better protection of fibers from moisture degradation.
Another benefit is the reduction in stress concentration and hence more efficient mechanical joining. In fact, one
of the main causes for the hub failure in Aerostar blades is believed to be the crushing of the brittle polyester resin
used. Last but not least, benefit is found in higher-impact resistance.
During their lifetime, wind turbine blades are subjected to impact by foreign objects such as rain drops, flying
debris, and other objects. Since most impact occurs on the leading edge, proper protection of that leading edge can
minimize the impact problems.
Composites are more susceptible to rain erosion than metals. Upon localized impact by rain drops, the matrix
resin can be chipped away, exposing fibers. Impact by larger foreign objects can break the fibers, crack the matrix
and interface, and induce delamination between plies. Impact damage may also occur during shipping and
handling. Unfortunately, even an invisible damage caused by impact can substantially reduce compressive strength
(Williams and Rhodes, 1981). Therefore, a low compressive strength after impact has been one of the main
obstacles to wider application of composites.
Reduction of impact damage requires the use of tougher resins; yet, too tough a resin may inadvertently
transfer too much load to the fibers during impact, resulting in unwanted fiber fracture. A compromise should be
made between matrix/interface cracking and fiber break through proper selection of resin toughness. Nevertheless, a
tough resin usually yields a higher compressive strength after impact than a brittle resin does.
As discussed earlier, much progress has been made lately in the development of toughened thermosetting
resins for composites. As a result, there are a variety of toughened resins available at present. In addition, the
high-performance thermoplastic resins in Table 3-4 all exhibit high toughness.
In general, a tougher resin yields a tougher composite. The relationship between the neat resin fracture energy
and the composite interlaminar fracture energy appears to be bilinear (Hunston, 1987). When the resin is brittle,
the composite has a higher interlaminar fracture energy than the resin fracture energy, mainly because of the fiber
bridging. As the resin becomes tougher, however, not all the resin toughness is translated to the composite
toughness. In fact, the composite fracture energy is only close to the rule-of-mixtures estimate.
There are other ways of improving the impact resistance of composite laminates. One method is to place a
tough resin interleaf between plies to arrest ply cracks and to accommodate large, local shear stains induced by
impact. The other is to provide reinforcements in the thickness direction by stitching (Hunston, 1990). These
methods can be implemented without incurring too high a cost.
Glass fibers are more resistant to impact damage than carbon fibers. Thus, any scheme of composite
toughening is expected to work better with glass fibers than with carbon fibers.
WOOD/EPOXY COMPOSITES
When the first wood/epoxy blades were designed for the 38-m (125-foot) diameter NASA MOD 0A wind
turbine, the available material properties data were limited. A survey of the compression, shear, and bending
strength of many
wood species was available in the U.S. Forest Products Wood Handbook (1974), along with equations to correct
the base values to different moisture levels. Values for along-the-grain modulus of elasticity and across-the-grain
tension and compression strength were also available in these s. These data were valuable in assessing which
species would likely be best suited to wind turbine use and in estimating the weight, stiffness, and static strength
of a candidate design. However, only minimal guidance concerning the effects of fatigue, load duration, and
component size on long-term allowable strength was given.
Some research on wood fatigue was available from long-term, constant-deflection, bending tests performed in
the 1940s by W. J. Kommers (1943a, 1943b). These tests did not cover a wide range of R ratios or allow a clean
separation of tension, compression, and shear strength, but they did provide lower bounds for those properties and
showed that the base wood material had potentially attractive high-cycle fatigue capability. The effect on strength
of the epoxy, or of the butt joints that would be needed to assemble a blade from individual veneer sheets, was
simply unavailable, for either static or fatigue conditions. Questions of how moisture and temperature variations
would interact with the above in high-cycle fatigue to determine long-term allowable strength were also
unanswered. Engineering judgment to extrapolate from a few sources of available test data plus conservative
design to cover areas of material property ignorance were, therefore, a requirement for the initial design work.
When the initial design studies began to reveal considerable promise for wood/epoxy as a wind turbine blade
material, testing of a few key strength features quickly followed. For the base material, this meant some strength
testing of fir/epoxy laminate, both with and without typical butt joints. For the critical blade-to-hub connection, it
meant both static and fatigue testing of the bonded steel stud load takeoff system, as the earliest versions of these
studs did not have the strength required for the MOD 0A application, and rapid design evolution was needed to
gain a demonstrated margin against the design loads (Faddoul, 1981). Static and fatigue performance was further
demonstrated by a 20-foot test article that was in essence the inboard one-third of the full size blade. While a great
deal was still unknown about the fundamental material and joint properties at this point, factors critical to this
particular design were well enough known that the resulting blades served successfully in the field for the duration
of the DOE/NASA MOD 0A program (Lark et al., 1983; Faddoul, 1983).
Smaller blades for the Enertech 13.4-m (44-foot) diameter and ESI 16.5-m (54-foot) diameter machines were
designed with about the same level of basic knowledge and also served successfully in the field, with one such
machine now having over 25,000 operating hours (Clark et al., 1985). Of the thousands of such blades built and
used, none has been retired because of fatigue. Only failures from lightning strikes, storms, machine runaways, or
similar incidents have been reported.
While the economics of small blades and early wind machines could tolerate some design margin to cover
incomplete materials data, larger machines aiming for more cost-effective energy production did not possess as
much weight tolerance. This became most evident in the DOE MOD 5A program, which was a teetered, two-
bladed upwind design that eventually grew to 122 m (400 feet) in rotor diameter and over 7 MW in rated power.
Steel, GRP, and wood/epoxy were all assessed in the initial rotor trade-off studies, and it became quite clear that
the gravity-induced stresses from rotor weight were a significant design driver and that high-cycle material
allowables had a big impact on rotor weight and cost. Wood/epoxy was eventually selected for this project, as it
showed the lightest weight and least cost. But it was known that major gaps in the material and joint properties
data would have to be filled to use the material efficiently and confidently, and so a major test program was
initiated to support this design effort. The eventual testing included tensile tests of specimens 9.1 m (30 feet) long,
with a 61-cm (24-inch) by 15-cm (6-inch) cross-section, that took over I million pounds to fail in static tension.
Both static testing and fatigue testing of planks up to 5 cm (2 inches) by 20 cm (8 inches) by 9.1 m (30 feet) were
performed. In addition, a great many lab-size specimens were tested. Materials with butt
joints, scarf joints, and without joints were compared to quantify the effect of joints. Shear testing at joints, both
with and without bonding voids, was performed both in fatigue and static loading. While the MOD 5A was never
built, the legacy of that program was a vast improvement in the depth and breadth of wood/epoxy material
property data, only part of which has been mentioned here.
In 1986, Gougeon Brothers, Inc., under a DOE/SBIR (Small Business Innovative Research) award, carried
out a study to advance the performance of jointed veneer laminate. This was followed by a Phase II grant to extend
the initial joint investigations with more fatigue tests (including micromechanical study of crack propagation at
joint sites) and to study a range of other material property topics, such as cross-grain tension ( y, z) and rolling
shear ( yz) strength (see Figure 3-5) in both static and fatigue loading as a function of size and the mechanical and
cost performance of graphite fiber augmentation between the veneer layers for high-stress applications.
The work covering the MOD 5A experience has been prepared by NASA and was recently released (Spera et
al., 1990). The experimental data from these extensive research efforts, along with in-house work at Gougeon
Brothers, and material research published as part of the British wood/epoxy wind turbine blade efforts (Ansell et
al., 1987; Bonfield and Ansell), have tremendously advanced current knowledge of these material properties, even
though some gaps in the data still remain. A discussion of some of the main features of these data follows to
outline a number of the major lessons learned, some of which may also be applicable to GRP or other candidate
blade materials.
The static tension and compression strengths of wood/epoxy laminate are distinctly different and respond
differently to various factors. For instance, loss of along-the-grain strength with increasing size is much more
severe in tension than in compression.
This reduction of strength with increasing size is called size effect. It is a stochastic process due to the
accumulation of material defects as item size/volume increases. The worst such defect in the stressed volume
limits strength in a statistical fashion. A tiny specimen of fir may show static tensile strength nearly twice its static
compression strength, but for laminated members the size of large wind turbine blades, the static tension and
compression strength will be about equal. This is due to a much more rapid loss of strength with size for tension
than for compression. Early tests of size effect relied on beams in bending (Bohannan, 1966), but due to the
nonlinearity of stress through the thickness resulting from crushing on the compression side, it was difficult to
cleanly separate the tension and compression stresses at failure. This nonlinear effect is why a stick snaps on the
tension side when bent, despite its greater tension strength capability. For adequate stress design of large wind
turbine blades, this ''size effect'' must be separated, understood, and accounted for, as the strength correction
relative to small laboratory samples can easily reach a factor of 2 in tension. Fortunately, size effect is now
reasonably well quantified for static tension in wood/epoxy, although the less severe static compression case is
still not well quantified.
While size effects along the grain ( x) are design drivers because they directly affect the allowables in the
primary structure of the wind turbine blade, across the grain strength ( y and z) (analogous to through the ply
strength z for GRP) (Figure 3-5) shows far bigger static strength differences and size effect (Barrett, 1974;
Barrett et al., 1975; Weibull, 1939). For typical wood/epoxy laminates, across the grain tensile strength may be
only one-fourth of the across the grain compression strength, and perhaps one-twentieth of the along the grain
tensile strength (Table 3-6). Making matters worse is the fact that size effect in cross-grain tension is much more
severe than along the grain, and differences between strength levels measured with small samples in the laboratory
and those for large structures in the field could exceed a factor of 5. The upshot of this is that cross-grain tension
stresses that may seem so modest as to be of little concern to the unwary designer could in fact lead to fracture
initiation and eventual failure of large, expensive blades in the field.
Figure 3-5
Laminate directional properties (top); shear direction nomenclature.
NOTES
a These typical properties represent a composite of laboratory test data from specimens having various moisture contents and uniformly
stressed volumes ranging from 19.5 in3 to 194.0 in3 (319.5 to 3179 cc). Properties are representative of structures similar to small wind turbine
blades. Larger structures are characterized by lower mechanical properties. Adjustments in properties due to size effects and other conditions
are considered by Gougeon Brothers in engineering each application.
b See Figure 3-5 for illustration of stress directions.
c See Figure 3-5 for laminate shear orientations.
d Assumed wood moisture content.
In contrast to the potentially dangerous cross-grain tension case, across the grain "rolling" shear has been
found to exhibit a very modest size effect, and lab data can be safely applied to large static structures with very
little correction. This also appears to be the case for cross-grain compression, although the data are rather
incomplete for that case.
It has been known for some time that both temperature and moisture have a direct effect on the static strength
of timber, and this holds true for the derivative wood/epoxy laminate as well, with some adjustments. Temperature
effects are essentially those of the wood species used since it carries the bulk of the stress, and, given that
temperature is relatively constant throughout a thick laminate, laboratory samples can be used to predict laminate
effects (Figure 3-6). On balance, temperature effects are modest and fairly well understood, and, like GRP, it is the
upper temperature extreme that must be accounted for in design. Ten to 30 percent strength reductions would not
be unusual in a computation of worst-case static strength compared to room temperature.
Moisture effects are modified for wood/epoxy laminate to the extent that the epoxy seals out liquid water
entirely and is a fairly effective barrier to the passage of moisture in the form of water vapor. As a result, a large
structure will not respond to the short-term moisture fluctuations in its environment, but will instead come to
equilibrium with the average humidity over a period of months to years (Figure 3-7). Static tension strength shows a
much more modest effect from moisture in the normal outdoor range than does compression strength (Figure 3-8),
but accurate design may require taking both into account depending on which property limits a specific design. If
typical 12 percent wood moisture content reference data are used in design, most use environments are already
covered, but a 10 or 20 percent strength credit may be available for dry environments, such as the California wind
farms.
It has long been known that duration of load has a direct effect on the static strength of timber beams1
(Gerhards, 1977; Barrett and Foshi, 1978). For a 10-year steady load duration, the maximum load would be about
60 percent of that which could be carried in a test whose duration was only minutes. Limited test data confirm that
duration and associated rate of loading effects exist in wood/epoxy laminates as well, but at present these data
cannot be used to make accurate predictions for long field exposure. Separate data for tension and compression
over a sufficiently long time span simply are not available. Also, there is reason to suspect that stabilization of the
moisture level within the laminate due to the epoxy will retard the creep rupture mechanism compared to the
unprotected timbers of earlier tests. However, no theory for estimating this, or reliable data for engineering use, is
currently available, so the rather large traditional strength reduction must still be used until better information
becomes available.
All of the above effects continue to play a role in determining the safe fatigue strength of wood/epoxy
laminate, but their description, interaction, and quantification become more complicated for fatigue compared to
static strength. To provide a logical separation of tension from compression fatigue effects, Department of Energy
(DOE) laminate testing has concentrated on three primary loading cases: R = 0.1 tension (Figure 3-9), where the
load is always tension and the least tensile stress is one-tenth of the largest; R = 0.1 compression (Figure 3-10),
where the load is always compressive and the least compressive stress is one-tenth of the largest; and R = -1
reversed axial (Figure 3-11), where the stress alternates between equal levels of tension and compression on each
fatigue cycle. As is seen, the first two cases are pure tension and pure compression, and the last involves both
tension and compression to reveal interaction effects. These cases correspond reasonably well to major design-
driving stresses in many wind turbine blade designs. The nonreversing R = 0.1 tension and compression loads are
similar to those found in the upwind and downwind blade structures due to aerodynamic lift in the
1 This is due to viscoelastic relaxation, which leads to creep rupture failure under large steady loads.
Figure 3-6
Wood strength change due to temperature at two moisture conditions (M.C. = moisture content).
Figure 3-7
Wood moisture content versus atmospheric relative humidity.
Figure 3-8
Effect of moisture content on laminate mechanical properties.
Figure 3-9
Typical tensile fatigue strength of wood/epoxy laminate.
Figure 3-10
Typical compression fatigue strength of wood/epoxy laminate.
Figure 3-11
Typical reversed stress fatigue strength of wood/epoxy laminate.
blade, while loads that are nearly fully reversing will be found near the blade's leading and trailing edges due
to the deadweight gravity bending moment.
Testing (both with and without joints in the laminate) shows the lowest strength levels and most rapid loss of
strength with increasing cycles for the reversed load R = -1 case. Strength levels for R = 0.1 tension fatigue can be
50 to 100 percent higher than for reverse axial, but the rate of strength loss with cycles is similar. Depending on
size, compression strength will generally be somewhat lower than tension at low cycles, but its strength loss rate is
only about two-thirds of that for tension, so it can cross over and exceed tensile strength at higher cycles,
particularly for large blades. These strength loss rates tend to be s for each type of loading over a wide range of
cycles. Consequently, the data plot as a straight line on a log-log stress versus number of cycles (S-N) plot. This
linear trend certainly holds valid up to 107 cycles; there are not enough data beyond that to determine where or if a
fatigue endurance limit exists. That knowledge could be valuable for better assessment of very high cycle strength
and life, but the appropriate data are expensive and time consuming to obtain. Current life and strength
assessments are forced to assume that the loss of strength shown at lower-cycle levels continues at the same rate to
arbitrarily high cycles.
The effect of size on strength varies dramatically with the type of loading. While size effect for static tensile
strength along the grain shows strength loss that can be a factor of 2 between laboratory samples and large blades,
it becomes worse for tension fatigue where it can reach roughly a factor of 3. For fully reversed R = -1 fatigue, the
size effect drops back about to the 2:1 level of static tension. Size effect data for compression fatigue and static
compression strength are minimal, but size effect appears to drop further to perhaps 1.5:1 for R = 0.1 compression
fatigue and to some lesser but ill-defined value for static compression (Spera et al., 1990).
Recent research work in size effect in the secondary material properties of cross-grain tension and rolling
shear was not able to cover the range of R ratios needed to draw conclusions similar to the above, but it revealed a
new phenomenon due to carefully controlled sample design and test structure, namely that size effect is more
severe for higher-cycle levels. While rolling shear showed very little loss of strength with size in static tests, that
was no longer true at 106 cycles; and while cross-grain tension did show substantial size effect in static tests, it
again became considerably greater at 10 6 cycles. An associated effect is that the fatigue curve slope, the rate of
strength loss with cycles, is larger for the larger specimens. So not only do larger specimens start out with a lower
initial static strength, that strength is lost faster on a fractional basis in fatigue than it is for smaller samples
(Bertelsen and Zuteck, 1991).
These lessons in size effect are important and will be increasingly so as wind turbines grow larger for
economic reasons. There is good reason to suspect that size effect will occur in all materials currently
contemplated for future wind turbine rotors, including GRP and metals, particularly for high-cycle fatigue. We now
have experimental proof that low levels of size effect in static tests are no assurance of its absence in high-cycle
fatigue, and designs undertaken in ignorance of these effects could lead to expensive or even dangerous failures in
future large-scale machines.
The interaction of long-term environmental effects with the fatigue process has been partially addressed in
wood/epoxy design by running fatigue tests at a range of temperature and moisture levels. However, it was found
that the cycle rates typical for long-duration tests drove substantial moisture from the test specimens, thereby
elevating their strength and fatigue performance. This may not occur in the same way at the lower-cycle rates and
intermittent use that occurs in the field, so caution must be exercised lest the elevated performance of laboratory
samples lead to excessive high-cycle fatigue expectations in the field.
The effect of veneer splice joints on strength is much more pronounced in fatigue than in static strength. Early
work quickly highlighted this and showed that 10 to 20 percent strength loss could easily occur owing to these tiny
joints distributed throughout the laminate. The effects were particularly apparent in tension and in high-cycle
fatigue. The first
attempts to combat this by using beveled scarf joints between the veneers met with only limited and frustratingly
intermittent success. In the large-scale fatigue specimens, performance was decidedly poor, and zipper-like
failures through large numbers of joints often occurred. This puzzling state of affairs has been investigated at
length in the recent Small Business Innovative Research (SBIR) research work and was traced to expanding gas
driving resin from the joints during the vacuum molding process (Bertelsen and Zuteck, 1991). The path to
higher-performance joints through mitigating this effect now seems clear in principle, but a practical
manufacturing method has yet to be proven. The important general lesson is that defect-driven failure modes that
are hard to spot in small specimens tested in static or low-cycle loading may emerge as major issues in bulk
specimens tested to moderately high cycles. It is difficult to see how the high-cycle fatigue performance of large
structures can be assured without at least a few tests of large-volume specimens to elevated cycles. Fortunately,
considerable data of this sort now exist to support the most critical design areas of large wood/epoxy wind turbine
blades. However, those data are by no means complete, and comparable data for design in GRP are unavailable.
Practical limits to the amount of materials data available will always exist, and perfect knowledge is not
necessary to execute a successful design. The benefit of improved materials knowledge is that a more appropriate
design that sidesteps pitfalls without undue cost can be provided and can be put into service with a much lower
uncertainty as to its operating envelope and expected life. The record of wood/epoxy blades in the field has been
excellent despite early materials data limitations due to confining the design window to stay within available
knowledge. A wider design window is now available due to knowledge accumulated over the past decade. This
will be needed in addressing the larger, more cost-effective designs that are needed in the future. A still broader
design window can be opened by addressing the known shortfalls in the existing data and any additional ones that
may be revealed as further advanced blade design occurs.
RECOMMENDATIONS
When designed and manufactured properly, glass/polymer and wood/epoxy rotor blades can provide tens of
thousands of hours of operating time. The recent advances in composites technology, however, may provide an
excellent opportunity to further improve the blade cost/performance. The knowledge base gained over the past two
decades in the aerospace industry can now be used to assess these more efficient blade designs using advanced
composites. What is needed is a focused research to refine the design along with improved materials knowledge in
certain key areas. Specific recommendations for further research are described below, in the order of priority:
1. Long-term fatigue data should be generated to at least 100 million cycles for the most useful
composite laminates and critical elements containing manufacturing defects, under appropriate
environmental conditions. Separate tension and compression fatigue data should be generated so that
independent lifetime projections for upwind and downwind portions of blade skin can be made. Fully
reversed fatigue data should also be obtained for the projection of life against edgewise gravity loads.
Moderate-cycle R ratio data may be needed to fully define material fatigue response. The data should
include documentation on the modes and growth of damage as well as the effect of damage on
properties.
2. At least a few large specimens of the favored composite laminates should be tested in moderate- to
high-cycle fatigue to quantify the fatigue size effect allowance that must be made for large rotors in
the primary loading regimes. All loading cases are needed for GRP. For wood/epoxy, size effect in
static compression and compression fatigue is not yet quantified. Sui tests comparing small and large
specimens of matched parent material are needed to fill this gap in the data.
3. Spectrum loading data to assess the available cumulative damage models are not yet available but are
needed to fully quantify high-cycle life predictions. A life prediction methodology should be
developed to predict
blade lifetimes based on constant-amplitude fatigue data. It should also be able to predict the modes
and growth of damage.
4. A database should be established for wind turbine blade materials. It should include mechanical
properties on fibers, matrix resins, and composites, including wood/epoxy. An extensive search of all
fatigue data on both GRP and wood/epoxy should be conducted and the results included in the
database.
5. Benefits of hybrid composites should be explored through design and limited testing. Design studies
should investigate savings on blade weight and life-cycle cost. The best hybrid reinforcement appears
to be carbon for either GRP or wood/epoxy.
6. Duration of load and creep effects are incompletely understood for wood/epoxy laminate. It appears
that epoxy stabilization of moisture levels reduces creep rate and magnitude, but data to verify and
quantify this for design are needed. Furthermore, moisture correction for laminate in the 5 to 12
percent operating range is still based primarily on old literature data. A better database would improve
life predictions.
7. Promising fiber preforms should be identified and examined under fatigue. Low-cost processability
should be taken into consideration.
8. Environmental effects, including ultraviolet exposure, moisture absorption, and temperature
fluctuations, should be delineated for the candidate composites and surface coatings.
9. Tough new resins should be examined for their applicability to rotor blades.
Launikitis, M. B. 1978. Chemically Resistant FRP Resins. Technical Bulletin of Shell Chemical Company.
Lee H., and K. Neville. 1967. Handbook of Epoxy Resins. McGraw-Hill, New York.
Lee, B. L., C. K. Riew, and R. J. Moulton. 1980. Rubber Toughening of Tetrafunctional Epoxy Resin. 12th National SAMPE Technical
Conference.
Lee, B. L., F. H. Howard, and E. H. Rowe. 1983. Effect of Matrix Toughening on the Crack Resistance of SMC Under Static Loading. 38th
Annual Conference of SPI, Session 9-A.
Lee, S. M. 1986. Compression After Impact of Composites with Toughened Matrices. SAMPE Journal, March/April.
Lubin, G., Ed. 1982. Handbook of Composites. Van Nostrand Reinhold, New York.
Lynch, T. 1989. Thermoplastic/Graphite Fiber Hybrid Fabrics. SAMPE Journal, Vol. 25, p. 17.
Mandell, J. F. 1990. Fatigue Behavior of Glass Fiber Composites. Presented at the NRC Workshop on Assessment of Research Needs for Wind
Turbine Rotor Materials Technology, Washington, D.C., January 22-23.
McGarry, F. J., and J. N. Sultan. 1969. Crack Phenomena in Cross-Linked Glassy Polymers. ASTM STP 460, ASTM.
McGarry, F. J., E. H. Rowe, and C. K. Riew. 1978. Improving the crack resistance of BMC and SMC. Polymer Engineering Science, Vol. 18,
p. 78.
Modern Power Systems, July, 1986, p. 19.
Morgan, C. A., A. D. Garrad, and U. Hassan. 1989. Measured and Predicted Wind Turbine Loading and Fatigue. Presented at the EWEC
Conference in Glasgow, 1989.
Moulton, R. J., and R. Y. Ting. 1981. Effects of Elastomeric Additives on the Mechanical Properties of Epoxy Resins and Composite Systems.
International Conference on Composite Structure.
Murtha-Smith, S. 1985. Loads and Fatigue Evaluation of Hamilton Standard WTS-4 Wind Turbine. Proceedings Wind Power '85, San
Francisco, August 27-30, p. 136.
Owen, M. J. 1974. Fatigue Damage in Glass Fiber-Reinforced Plastics. In Composite Materials, Vol. 5, ed. by L. J. Broutman, Academic
Press, New York.
Phillips, D. C., J. McCarthy, A. G. Davis, and J. E. P. Stott. 1987. The Use of Glass Reinforced Plastics and Advanced Composites for Wind
Turbine Blades. A Review of Experience in Their Use and of Relevant Design Procedures and Codes. Department of Energy, United
Kingdom, June.
Properties of Epoxy Resins, Hardeners, and Modifiers. 1982. Adhesive Age, April.
Reifsnider, K. L., Ed. 1980. Damage in Composite Materials. ASTM STP 775.
Reifsnider, K. L., E. G. Henneke, W. W. Stinchcomb, and J. C. Duke. 1983. Damage Mechanics and NDE of Composite Laminates.
Mechanics of Composite Materials. Recent Advances. Pergamon, New York, p. 399.
Reifsnider, K. L., and W. W. Stinchcomb. 1986. A Critical Model of the Residual Strength and Life of Fatigue-Loaded Composite Coupons.
Composite Materials: Fatigue and Fracture, ASTM STP 907, ed. by H. T. Hahn, p. 298.
Riew, C. K., E. H. Rowe, and A. R. Siebert. 1976. Rubber Toughened Thermosets. In Toughness and Brittleness of Plastics, ed. by R. D.
Deanin and A. Crugnola, American Chemical Society, Washington D.C.
Rosato, D. V., and C. S. Grove, Jr. 1964. Filament Winding. John Wiley & Sons, New York.
Scott, J. W. and D. C. Phillips. 1975. Carbon Fiber Composites with Rubber Toughened Matrices. Journal of Materials Science. Vol. 10, p.
551.
Sendeckyj, G. P. 1981. Fitting Models to Composite Materials Fatigue Data. Test Methods and Design Allowables for Fibrous Composites,
ASTM STP 734, ed. by C. C. Chamis, p. 245.
Spera, D. A., J. B. Esgar, M. Gougeon, and M. D. Zuteck. 1990. Structural Properties of Laminated Douglas Fir/Epoxy Composite Material.
NASA Reference Publication 1235, Report No. DOE/NASA/20320-76, May.
Springer, G. S., Ed. 1981. Environmental Effects on Composite Materials. Technomic Publishing Co.
Stoddard, F. S. 1989. Field Problems with Wind Turbine Rotors. Presented at the First Meeting of the NRC Committee on Assessment of
Research Needs for Wind Turbine Rotor Materials Technology, Washington, D.C., November.
Strong, A. B. 1989. Fundamentals of Composites Manufacturing. Society of Manufacturing Engineers.
Sutherland, H. J. 1989. Analytical Framework for the LIFE2 Computer Code. Sandia Report SAND 89-1937 UC-905.
Tsai, S. W. 1989. Composites Design, Think Composites.
Tsai, S. W., and H. T. Hahn. 1975. Failure Analysis of Composite Materials. Inelastic Behavior of Composite Materials, ed. by C. T.
Herakovich, AMD-Vol. 13, ASME, p. 73.
Van Delft, D. R. V., F. Hagg, and P. A. Joosse. 1987. The Influence of Fatigue Design Line Criteria on the Rotor Blade Design. Wind Energy
Conversion 1987, ed. by J. M. Galt, Mechanical Engineering Publications Ltd., London.
Vinson, J. R., Ed. 1977. Advanced Composite Materials--Environmental Effects. ASTM STP 658.
Weibull, W. 1939. A Statistical Theory of the Strength of Material, Royal Swedish Academy of Engineering Sciences, Proc. 151.
Whitehead, R. S. 1990. Damage Tolerance of Composites. Presented at the NRC Workshop on Assessment of Research Needs for Wind
Turbine Rotor Materials Technology, Washington, D.C., January 22-23.
Wilkins, N. J. M. 1984. Materials Aspects of Large Aerogenerator Blades . Proceedings of the European Wind Energy Conference. Hamburg,
West Germany, October 22-26, p. 281.
Williams, J. G., and M. D. Rhodes. 1981. The Effect of Resin on the Impact Damage Tolerance of Graphite/Epoxy Laminates. NASA TM
83213, October.
Windpower Monthly. 1987. June, p. 10.
Witzler, S. 1988. High-Temperature Thermoplastics: A Progress Report. Advanced Composites, March/April, p. 56.
Wood Handbook. 1974. Wood as an Engineering Material. U.S. Forest Products Laboratory, Agriculture Handbook No. 72.
Yang, J. N., and D. Shanyi. 1983. An Exploratory Study into the Fatigue of Composites Under Spectrum Loading. Journal of Composite
Materials, Vol. 17, p. 511.
Zweben, C., H. T. Hahn, and T.-W. Chou. 1989. Mechanical Behavior and Properties of Composite Materials, Vol. 1. Technolmic Publishing
Co.
4
WIND TURBINE ROTOR DESIGN ISSUES
Wind turbine rotor blades are a high-technology product that must be produced at moderate cost for the
resulting energy to be competitive in price. This means that the basic materials must provide a lot of long-term
mechanical performance per unit cost and that they must be efficiently manufactured into their final form,
including the cost of sufficient quality control. Unless a material choice and fabrication system can satisfy both of
these requirements, it will not be appropriate for advancing the state of the art for economical production of power
from the wind.
Both fiberglass-reinforced and wood/epoxy composites have been shown to have the combination of strength
and low material and fabrication costs required for competitive blade manufacture. Their fabrication requirements
and constraints, the current state of their materials database, and the areas where further research could advance
their efficient use are, however, often quite different. The gradual evolution of wind turbine configuration and
rotor designs also places a continuing demand on both systems to meet new requirements. In turn, the strengths
and limitations of each material system also feed back into the process of system configuration selection and
detailed design, as any design that cannot be efficiently produced will be at a competitive cost disadvantage. A
really good design will already include the compromises necessary so that it can be effectively manufactured with
acceptable quality in the material system chosen. A number of these considerations are discussed below.
AIRFOIL EVOLUTION
Early wind turbine blades typically used airfoil shapes borrowed from helicopter or low-speed aircraft use,
such as the NACA 23xxx or 44xx series (Figure 4-1). These airfoils have shapes that are everywhere convex and
in particular do not have a concave or reflex aft portion. While these were a natural starting point, they were found
to be sensitive to the buildup of insects on the leading edge, which caused substantial loss of power output for
many wind turbine designs.
Further evolution of airfoil choice in view of field experience has led to the use of the more modern
Wortmann or NACA laminar designs, as well as the aft loaded LS-1 type. These airfoil designs generally employ a
reflex or concave aft portion, which feature is also evident on the special-purpose SERI wind turbine airfoils
(Figure 4-2). That concavity effectively eliminates certain manufacturing processes, such as filament winding, from
consideration for producing the outer airfoil shell. While it also forces some modification to the female mold
fabrication technique, both fiberglass and wood/epoxy composites have successfully been produced in reflex
shapes without significant cost penalty.
Precise control of airfoil geometry is quite important in providing blades with consistent aerodynamic
properties. Small variations in outboard airfoil camber (±1/4 percent of chord) or twist (±1/4°) can lead to
substantial aerodynamic imbalance and rotor and turbine life reduction. For stall-controlled machines, off design
peak power can further reduce cost-effectiveness by either overstressing driveline and generator components, and
thereby increasing replacement costs, or by reducing total power output and revenue. This need for aerodynamic
consistency and accuracy has led to the adoption of molding as the fabrication method of choice for both
fiberglass
Figure 4-1
Commonly used early HAWT airfoils. Source: Tangler (1990).
Figure 4-2
SERI advanced wind turbine airfoils. Source: Miley (1982).
and wood/epoxy composites, as it provides control right at the outer aerodynamic surface, which determines
ultimate performance. Both material systems are able to provide the complete range of outboard airfoil shapes
currently of interest.
While the outboard portion of the rotor changes little with material choice, the inboard region is a different
matter. Fiberglass rotor blades often incorporate a large amount of inboard planform area and twist, and may carry
the maximum chord quite far in toward the root (Figure 4-3) (Stoddard, 1989). For a given rotor diameter, this
will produce the most power, albeit at the cost of a significant increase in total blade surface area and materials.
The flat sheet nature of the veneers used in wood/epoxy construction does not lend itself well to large inboard
planform and the twist and the double curvature surfaces that result. Instead, a gradual transition from the inboard
airfoil shape to an oval root is performed over the inner third of the blade (Figure 4-4). To regain the minor power
loss due to lessened inboard planform area, a slight increase in blade length is provided. Since the turbine rotor
designer is free to sweep energy out of the flow at whatever radius provides the least rotor cost, this is an effective
solution because the cost of the slight extra length is small compared to the large reduction in inboard planform
area. Low wind start-up torque is reduced, which could be a limitation for some turbine designs, but the reduction
of planform area also reduces storm wind loading on the turbine as a whole, so the cost trade-off at the system
level may or may not be favorable depending on start-up requirements.
Figure 4-3
Nine-meter GRP wind turbine blade. Source: Stoddard (1989).
Figure 4-4
Eleven-meter wood/epoxy blade airfoil thickness distribution and planform. Source: Musial et al. (1985).
savings and corrosion resistance that it provides may make it attractive for this application.
The wood/epoxy composite blades produced by Howden also use the pivoting tip arrangement. While
operating hours and machine numbers are not nearly as large as for the Danish designs, application with the
wood/epoxy material system has not shown any unusual problems to date.
The Enertech and ESI designs of U.S. origin both used pivoting tip plates for overspeed protection. These
were retained in place using threaded rod epoxy bonded into a veneer buildup at the blade tip. The load takeoff
principle is identical to the well-tested root stud method and has proven highly reliable in service, except for cases
of massive overspeed due to failure of the latch mechanism to deploy. This is not a material knowledge or fatigue
problem.
Future wind turbines may employ techniques such as outboard blade ailerons or boundary layer control for
overspeed protection. Such methods are not only potentially preferable from an aerodynamic standpoint, but also
require much less disruption of the primary load path within the blade and could thereby reduce the possibility of
long-term fatigue failure as well.
Figure 4-5
Flanged GRP blade root design. Source: Stoddard (1989).
mechanically captured both via the bolts/bushings and the flange clamping plates. Polyester-based filler material
was used to fill the gaps between the hub, metal clamping plates, and the GRP flange to render the assembly as
stiff and free of movement as possible.
The other major variation of the flanged root used root roving material that was molded uniformly outward to
form the flange, after which the flange drilled and metal bushings were added. While the extra security of a wrap
around the bushings was no longer present, mechanical clamping between the hub and flange plates provided
positive mechanical capture. While filler material was again used to resist movement and render the final assembly
as solid as possible, like the Hutter root, this design again did not depend a metal-to-GRP bond. Given the
modulus and strain values of these designs, some motion between metal and GRP appears virtually unavoidable in
the long run.
Both of these flanged root designs encountered considerable trouble in service (Stoddard, 1989). Partly due to
quality control and manufacturing problems, and partly due to the inevitable shrinkage of polyester upon cure, the
GRP flange was often not clamped well between the hub and flange plates. Once some gap occurred, the blade
root was free to move somewhat and could then begin to reduce the filler material to powder. At the same time, the
GRP bundles would begin to move relative to the steel flange plates (Figure 4-6) and would often floss deep
grooves into their surfaces due to the repetitive motion. Loss of root stiffness and the appearance of fine powder at
the root were the symptoms of this degeneration. It was estimated by Stoddard (1989) that this problem was
observed on 5 to 10 percent of the blades in service and was expected on 30 to 50 percent of the blades, given
operation and management attention. This amounts to over 20,000 blades.
A procedure called a root pack was developed to deal with these root problems. In essence, the flanges were
removed, and fresh filler material was added to take up the gaps and restore a tight fit. In many instances the
bushings were shortened or recesses were machined into the flange plates so that the bolt pretension could clamp
directly against the GRP flange. Bushings with a carefully selected crushing stress were also employed in some of
the later designs, so that the GRP and steel bushing could each share some of the bolt preload, which was the
original design intent of all the flange designs. It should be pointed out that some known flanged blades have over 5
years of operation with no problems and no root pack.
There is no doubt that the root pack procedure extended the useful life of many blades in the field. However,
the increased clamping stresses within the GRP may now be taking a toll, as long-term tensile failure right through
the fiber bundles has now begun to appear (Poore and Patterson, 1990). While a well-designed, well-manufactured
flanged root may offer life of several years, it does not yet appear able to offer the desired decades of life. This is
true even though the blades in question are generally less than 10 m in length. Life expectancy is only likely to get
worse as larger blades with increased gravitational bending moments are designed.
Another class of fiberglass root design is based on a tapered metal tube that is bonded to the root region of the
blade (Figure 4-7a). The tubes may be interior, exterior, or both and may occur with or without through bolts. For
its long-term fatigue performance, this class of roots generally depends on a metal-to-GRP bond to transfer load
via shear from one material to the other. These steel root tubes are provided with a steel flange so that they can be
bolted to the hub assembly. The increased amount of metal required for this class of retention may be more costly
compared to the types where the flange is formed directly from the GRP. However, roots of this sort appear to be
reliable when properly designed and manufactured and have served well up to the present time.
Figure 4-6
Flanged root design limitation due to steel/GRP strain incompatibility.
Figure 4-7a
Bonded steel root tube GRP hub design.
Figure 4-7b
Wood/epoxy blade root.
BLADE JOINING
Efficient field joining of blade sections has not yet emerged as a major issue for blade design and
manufacture, because most of today's machines are small enough that shipping the blades in one piece is not much
of a problem. However, conceptual studies have indicated potential cost and weight advantages for two bladed
rotors that are essentially one continuous structural piece due to the use of a composite ''flow through'' hub
(Figure 4-8). This system entirely eliminates the normal root retention and, in
Figure 4-8
All-wood/epoxy composite Hub. Source: Stroebel et al. (1984).
conjunction with a teeter system, needs to carry only net thrust and torque to the wind turbine. The resulting
design is structurally efficient and should have excellent long-life fatigue potential, but it will require some form
of field assembly to allow rotors to be shipped in pieces of convenient size. A cost-effective, fatigue-proven
joining system for both fiberglass and wood/epoxy surely seems within reasonable technical reach but has not been
developed and verified at this time. Such a field joining technique would also be of great value for vertical axis
machines, whose curving blade shape makes them particularly difficult to ship in large pieces. Many other ways to
use such a system to reduce machine-installed cost may be discovered once it is available and optimized.
Fiberglass Blades
The least-expensive structural material for fiberglass blade construction is unidirectional roving. This is a
high value-added material
for the manufacturer and, due to its straight fiber path and low resin content, is also one of the best-performing
fiberglass forms on the basis of strength and stiffness to weight. It is therefore of economic and structural
advantage to employ this material for as much of the primary structure as possible, consistent with efficient
fabrication. Its limitations are that it does not have the thickness uniformity of fabrics and does not distribute well
onto vertical or near-vertical mold surfaces. This can lead to fabrication difficulties near the blade root.
While conventional filament winding and transverse filament tape (TFT) are not applicable to the most recent
classes of airfoils, with their generally reflex trailing edge regions, these fabrication techniques could still be
considered for the primary structural spar in a spar/skin blade construction. It would be necessary for spar
fabrication by winding to be very low cost to compete effectively with the unidirectional roving reinforcement
mentioned above, but there may be designs or production levels for which that is possible. The method would gain
considerably if some of the cost or complexity of root formation could be included efficiently in spar construction,
as might be the case for winding directly over the root fitting.
While spar/shell construction has many attractions, it must overcome at least two significant structural
handicaps. The first is that it is difficult for the spar to be stiff enough edgewise to prevent large strains in the
blade trailing edge, which would lead to early shell cracking and failure. This could be handled by putting more
active material in the blade shells so they carry more of the edgewise loads themselves or by the addition of a
unidirectional trailing edge spline. The second handicap is that material concentrated in the spar rather than spread
through the blade shell results in thinner skin laminates, which then need more buckling support and are easier to
damage in the field. Simply put, it is difficult to beat the structural efficiency and ruggedness of a stressed skin
design, and significant fabrication economies must exist for the spar/shell construction to be competitive with it.
This is fundamental in assessing likely blade fabrication or material improvements.
The foregoing notwithstanding, it should be pointed out that several Danish blades, such as LM, MAT, and
Vestas, are currently produced with spar/shell construction and are performing successfully. WEG is producing a
wood/epoxy spar/GRP tail panel hybrid that also is performing well. Only time and further economic competition
can show if such designs will prove to be the most cost-effective in the long run.
Even stressed-skin fiberglass blades often need panel stiffening in the relatively flat aft blade area, especially
inboard where the span from shear web to trailing edge can be particularly large. Expanded foams are often used to
provide the required panel support, due to their light weight, moderate cost, and easy formability. However, foams
can be impact damaged by rough handling, and their long-term fatigue properties are not well characterized for the
numbers of cycles typical of wind turbine blades, particularly under conditions of variable outdoor environment or
in the presence of damage. Honeycomb is another skin-support possibility, but it has received little attention due to
its cost and greater formability problems compared with other choices.
A panel-stiffening material used in a great many fiberglass blades to date is a felt-like material called core
mat. Core mat is inexpensive and easy to conform to complex blade shapes. It is very light in the resin-free
condition and, while it does absorb considerable resin, still has acceptably low density once bonded in place. Since
this material is somewhat fibrous in nature, it may have rather good long-term fatigue properties, but at present
this useful component of blade fabrication is not characterized in long-term fatigue. However, it can be said that
shell-cracking problems seen in the field do not appear to be caused by core mat limitations for the kinds of
lifetimes and cycle counts currently achieved. This situation could change for larger blades with larger panel sizes
and deflections, but data for a quantitative assessment of these possible problems do not exist.
Since most blades are formed in female half-shell molds, joining of the blade shell halves and the interior
shear webs is a major part of the overall manufacturing process. A mixture of polyester resin and talc called
plasterite has been extensively used as a filler in the joining operations by certain fiberglass blade manufacturers.
This material is stiff enough to attract considerable load when used in large quantity, as has often been the case,
but is not possessed of fibrous reinforcement for strength and crack stopping in long-term fatigue. Another problem
is shrinkage after fabrication, which builds in residual tensile stresses that the material is not well suited to handle.
The only real attraction of this material appears to be its low cost, since the unfavorable combination of stiffness
without strength appears to lead to high-cycle shell cracking, after crack origination in the bulk plasterite bonding
material. Adequate fatigue characterization of this material does not exist, and it may be an impending problem for
many blades now in use in the field. Other, better-suited, fillers with greater fatigue capability and suitable cost,
handling, and health properties may be an important component in providing long-term blade life at a competitive
cost. Test data to identify, select, and predict the performance of better fillers are not available to the blade
industry at this time.
Another area of concern to the blade designer is the selection of blade skin-reinforcing materials. It is well
known that conventional fabrics with a lot of fiber crimp due to the weaving process suffer in high-cycle fatigue
due to resin breakdown and fiber stress concentrations at the intersections in the weave. So materials with less
fiber crimp will be preferred in high-cycle fatigue wind turbine applications. However, this still leaves a wide
selection of reinforcement styles across a broad range of costs in terms of the basic fiberglass chosen. A question
of the role and durability of chopped mat and double bias constructions in the high-cycle environment also exists.
Test data to allow a rational choice of the most cost-effective reinforcement are scant.
Another basic question exists in the choice of resin system. While epoxies are generally conceded to have the
best long-term fatigue strength, they are considered by some to be more difficult to work with and too expensive
for wind turbine blade fabrication. Current production has, therefore, concentrated primarily on the low-cost
polyester system, with some designers opting for the somewhat higher-cost vinylester systems. While there is no
doubt that vinylester is stronger at low- and moderate-cycle levels, it also appears to have a much larger fatigue
effect than polyester. Some current data appear to suggest (Burrell et al., 1986) that, at 10 million cycles and
above, polyester may, in fact, outperform the more expensive vinylester. An appropriate resolution of this question
would obviously benefit cost-effective blade manufacture.
Wood/Epoxy Blades
The fundamental fatigue database for wood/epoxy composite is in good condition compared to that for GRP,
as has been pointed out elsewhere. However, a number of topics to further improve the cost-effective use of this
material system do exist. One of these, the cost-effective field joining of fabricated pieces for advanced rotor
types, has already been mentioned. Additionally, research work by Gougeon Brothers completed under an SBIR
contract to the Department of Energy (Bertelsen and Zuteck, 1989) has shown that advanced laminate splice joints
can allow bulk laminate mechanical performance to approach that of unjointed material, but proving out practical
manufacturing methods to realize this promise has not yet been accomplished. Another area of possible interest is
that of alternative species to the Douglas fir used for the bulk of the previous work.
All of the above topics could improve the effectiveness of the wood/epoxy material system within the current
design envelope. However, substantial gains may also be made by expanding the design envelope for the material.
As was mentioned previously, the veneer sheets used with this construction technique constrain the acceptable
blade geometries because they can accept only a very small level of double curvature without splitting or creating
excessive resistance to the vacuum-molding forces. In this regard the problem is much like that for molding
thermoplastic sheets, which are also relatively stiff and lack the convenient "drape" of conventional fiberglass
materials. There may be many ways to move beyond current restrictions. For instance, high-temperature forming
processes may be able to push the wood veneer into its plastic range and provide forming possibilities that room-
temperature molding cannot. Or it may be possible to preprocess the veneer into a form that is much more
conformable than the current full veneer sheets but that is still easy and efficient to handle. The fact that the
database for wood/epoxy is in good condition for wind turbine blade fatigue design does not mean that further
significant advances in this material system are not available, but rather that they may lie in somewhat different
areas than for fiberglass.
Finally, both material systems can benefit from explicit feedback into the aerodynamic design process, so
that the shapes specified exploit the strengths and cost efficiencies of each to the best overall effect.
Whichever material system is used, long-term fatigue performance will benefit from using simple external
shapes with primary load paths that are as straight as possible, particularly in the highly loaded root and root
transition areas. Minor power losses due to simplification of the inboard rotor geometry can be offset by a slight
increase in rotor length, resulting in a potentially lighter and lower-cost blade of equal energy capture performance
and superior long-term fatigue life. New blade designs should explicitly consider this structurally beneficial
approach at the aerodynamic design phase.
RECOMMENDATIONS1
Grp
1. Fatigue data that can resolve whether polyester or vinylester is the appropriate resin for high-cycle
wind blade use should be obtained. Common E-glass reinforcing, such as unidirectional fabric and
roving, double bias/mat, knit triaxial, and chopped strand mat should be investigated to determine
which perform best with these low-cost resins.
2. Filled bonding material (such as plasterite) appears to have caused fatigue failures of blade shells.
Testing to identify and fatigue qualify a low-cost bonding material would be valuable.
3. Static and fatigue data (including shear) to define the mechanical performance of core mat or other
low-cost core material are needed to assure that fatigue design goals will be met.
Wood/Epoxy
1. Provide proof of the static and fatigue performance of a low-cost field joining technique for use with
flow-through rotor concepts and future large or complex shape rotors (such as MOD 5 or Darrieus).
2. Investigate the potential for low-cost techniques to mold wood/epoxy laminate into double curvature
shapes that cannot be made with current methods. Confirm fatigue performance of the resulting
laminate.
Generic
Both GRP and wood/epoxy rotors could achieve cost and fatigue benefits from an all-composite one-piece
rotor design. The potential life-cycle benefits of such designs should be studied and appropriate composite hub
designs identified.
1 The recommendations made here are those the committee deemed to be of high priority for research.
5
MANUFACTURING PROCESSES FOR ROTOR BLADES
were chosen for start-up or developmental phases by wind turbine blade producers. However, there is a practical
size limit imposed by either the risk of human error in a lengthy lay-up operation or just the risk of dealing with
thickness and warpage variations that increase with scale. The greater tooling and facility costs of the mechanized
winding process are justified when the processing risk is sufficiently great.
Wood/epoxy blades are made by laminating 0.10-inch-thick wood veneers together using an epoxy resin
system. Sheets of length 2.4 m (8 feet) are coated with resin and placed in mold halves end on end from the root to
tip. The outside plies are scarf jointed, but all inner plies are simply butt jointed in a staggered pattern. A fiberglass
surface ply is used on both the inner and outer surfaces of the skin.
Once lay-up is completed, a vacuum bag system is installed and the laminate is formed to the mold by vacuum
pressure. Cure is completed at room temperature in 2 to 8 hours depending on the size of the component. The
blade skin is made in two halves, which are trimmed at the mating plane after the skins are cured. A root to tip
shear web is made from quality plywood, and grooved blocks are bonded to both surfaces of the blade to receive
the web on assembly. The two halves and the spar are bonded together with a filled epoxy.
The root end attachment is accomplished by bonding custom-designed studs into spanwise holes drilled in a
thickened section of the spar. Bolts thread into these inserts for attachment of the blade to the hub. This process
yields a lightweight, reliable quality blade at costs under $50/lb (Stoddard, 1989).
low applied strain or even cure shrinkage stress. These cracks can propagate as delaminations in fatigue (Rogers et
al., 1990).
Movement of the laminate prior to cure, such as occurs during compaction of thick sections, can change the
fiber length required by the final geometry. Since the fiber will not stretch or compress elastically, because the
actual fiber stress is low, the fiber will either bridge, causing a void, or buckle, yielding local fiber curvature.
Either way, significant resin stresses are induced as the fiber tries to straighten out under tension loading or buckle
under compression loading. The matrix will crack, and the fracture will propagate under fatigue loading as a
delamination in tension or as fiber crimping in compression. The latter failure mode is catastrophic (Morse and
Piggon, 1990).
Heat is produced by the chemical reaction of cross-linking for all plastic matrix systems. This phenomenon is
called exotherm. The matrix gells and becomes rigid at a given location as a function of time at temperature. The
exotherm can raise the internal temperature of a laminate well above the surface temperature, which causes gelling
to occur at different locations and different times. The fiber and matrix generally have different thermal
coefficients of expansion, which results in residual stresses on cool down after cure. Filament winding processes
introduce residual fiber stresses in the fiber matrix assembly that are retained through cure. Each of these
phenomena can result in local residual stresses of significance in the matrix. Procedures for computing these
stresses were developed under U.S. government contract and reported by Rai and Brockman (1988); thus, the code
is available on request.
The design phase is not complete until the effects of these conditions have been determined and their stress
effects included in fatigue analysis.
Figure 5-1
UH-1H composite main rotor blade. Source: Covington (1980).
The RTM process is utilized in the production of cooling tower fan blades up to 12.2 m (40 feet) in diameter.
These blades look quite different from current wind turbine blades because of constraints of the process
(Figure 5-2). The rigid inner mandrel must be removed from the cured blade; therefore, the root end must be the
largest section of the blade. To live with this constraint, tower blades are made in two parts, an outer section
extending from quarter span out to the tip and an inner portion containing a root end joint. These sections are
joined by bonding, and both sections are at their largest cross section at the splice joint. Local fiber curvature can
occur due to nonuniform flow of the resin through the preform. Cooling tower blades produced by this process
sell for under $10/lb. (Monroe, 1990).
Figure 5-2
RTM cooling tower blade. (Courtesy, Hudson Products Corp.)
Pultrusion
Pultrusion is a process in which a continuous dry preform is pulled through a matched die while resin is
injected under high pressure. It is first heated for cure and then cooled in the die to a temperature where it
possesses sufficient strength to support the pulling clamp pressure. Once in operation, the process is almost labor
free, yielding a product at just above the material cost (Goldsworthy, 1990). Die cost is reasonable if any volume
is required. The most obvious restriction is that the part must be constant in cross section and laminate design. The
quality of the laminate is excellent.
Fiber Placement
Fiber placement is similar to filament winding in the sense that the raw material is tow rather than tape. It is
also similar to tape laying machinery, where a roller is used to stick the ply to the tool or prior ply. The merger of
these two ideas yields a machine that can place fiber on any tool surface in any direction and produce a quality
laminate. This process offers all the design freedom of manual lay-up together with the uniform quality of machine
fabrication. Although quality is high and design freedom great, production costs are not comparable to either RTM
or pultrusion processing. Lay-down rates of 10 pounds per man-hour have been quoted (Policelli, 1990).
Thermoplastics are touted as the low-cost composite material of the future largely because the requirement
for cure is eliminated. Simply melt the resin while the preimpregnated plies are held together and then let it
solidify. Currently, high processing temperatures and pressures keep tooling costs high. If the raw form of the
material is a precompacted sheet, design freedom in fiber orientation is restricted. If prepreg tow could be located
on the tool by a fiber placement machine that had heated rollers and could fuse and compact the added ply to the
laminate already laid, perhaps thermoplastics could then replace thermosets. Efforts in this direction have not been
successful to date.
MANUFACTURING RECOMMENDATIONS
To meet economic realities, the fatigue life and reliability of the blades must be improved. Fatigue life can be
significantly improved by advancements in design. Reliability will be improved through carefully controlled
mechanized manufacturing processes. This all must be accomplished with lower labor content.
Pultrusion offers an immediate solution to the economic issues but at reduced aerodynamic and structural
efficiencies. RTM offers slightly longer-term opportunities while retaining aerodynamic efficiency. Fiber
placement offers optimum aerodynamic and structural efficiency but will require further development to meet the
necessary cost objectives. Fiber placement may be the best match for the emerging family of thermoplastic
matrices. It is not clear which of these processes is most suitable for wind turbine blade production in this decade,
but it is clear that the next generation of blades can meet their performance goals, and it is likely that one or more
of these processes will achieve the cost objectives.
Figure 5-3
4BW hingeless bearingless rotor. Source: Friedman (1988).
The technology to design and manufacture wind turbine blades with the desired life and reliability is
emerging. Loads must be determined. Matrix strains must be computed and fatigue allowables not exceeded.
Quality in fiber alignment, resin content and weight, balance, and contour must be maintained and costs kept low.
To achieve these requirements, the design must be developed with state-of-the-art and emerging analytical
procedures. It is far more economical to examine design alternatives on the computer than in the field.
It is clear that aggressive use of composites to reduce the weight of the rotor is of benefit to the entire wind
turbine system.
Rogers, C. W., E. W. Lee, and D. A. Crane. 1990. The Effect of the Fracture Process on the Strength of Composite Laminates, 46th American
Helicopter Society Forum, May 22.
Stoddard, F. S. 1989. Field Problems with Wind Turbine Rotors. Presented to Committee on Assessment of Research Needs for Wind Turbine
Rotor Materials Technology, Washington, D.C., November 7.
6
ACTIVE CONTROL IN WIND TURBINES
1. setting upper bounds on and limiting the torque and power experienced by the drive train, principally
the low-speed shaft;
2. minimizing the fatigue life extraction from the rotor drive train and other structural components due to
changes in wind direction, speed (including gusts), and turbulence, as well as start-stop cycles of the
wind turbine; and
3. maximizing the energy production.
TABLE 6-1 Control Effectors and Sensors for a Pitch-Controlled Wind Turbine
Control Effectors and COMMANDS Sensors and MEASURED PARAMETERS
Generator contactor Contactor status
ON ON
OFF OFF
Nacelle yaw drive Nacelle orientation
ON ANGLE
OFF
DIRECTION Wind direction
RATE ANGLE
END POINT
Wind speed
Blade pitch actuator SPEED
ON
OFF Blade pitch angle
DIRECTION ANGLE
RATE
END POINT Generator power
REAL POWER
Generator torque REACTIVE POWER
TORQUE
ANGLE
Reactive power SPEED
REACTIVE POWER
Figure 6-1
Wind turbine drive train.
load is the input switch matrix of the power electronics. For a constant-speed wind turbine—one without
power electronics—the generator load is the utility grid.
The nacelle yaw drive may be an electric or hydraulic motor acting through gearing to rotate the nacelle and
rotor. The function of the yaw drive is to orient the wind turbine relative to the prevailing wind direction.
The blade pitch actuator effects the rotation of the rotor blades about their pitch axis. With the blade pitch
angle set at the full-power angle, maximum power is extracted from the incident wind flow field. As the blade
pitch angle is rotated toward the full feather position, the blades become less efficient at converting the power in
the wind flow field to shaft power.
Also listed in the first column of Table 6-1 are two controls that are effected by the power electronics. The
generator torque may be controlled by the power electronics acting on the electrical characteristics of the
generator. The power electronics, acting as the load for the generator, can electronically vary the load and thus the
drive train torque. The reactive power delivered to or received from the utility grid may also be controlled by the
power electronics. This is important for maintenance of one of the desired electrical characteristics (power factor)
of the power delivered to the utility system.
The second column of Table 6-1 lists the typical sensor complement associated with a wind turbine. Given
also is the physical information measured by these sensors. The contactor status sensor indicates whether the
contactor is open or closed, that is, whether the generator is connected to the load. The nacelle orientation sensor
measures the angular position of the nacelle, either relative to a fixed reference or to the wind direction. The wind
direction sensor measures the angular direction from which the wind blows. The wind speed sensor measures the
wind speed. This sensor may be shared by a number of wind turbines or may be absent altogether. If absent, wind
speed may be derived from knowledge of the blade airfoil characteristics, the power level, and the drive train
rotational speed. The blade pitch angle sensor measures the pitch angle of the blades. The generator power sensor
measures the real power flow into, or delivered by, the generator. Typically, the reactive power flow is also
measured. The generator speed sensor measures the generator rotational speed and, through knowledge of the
gearbox ratio, the speed of the low-speed shaft.
Of importance to control are two quantities that may be estimated from the sensor measurements. These are
the wind speed and the torque in the low-speed shaft. Estimation of the wind speed value was discussed above.
Estimation of the torque value proceeds from knowledge of the power, the rotational speed, and the gearbox ratio.
minimization, which may be important for some systems/operating conditions. Certainly, in the wind turbine
environment, a reasonable designer's objective would be to minimize the cost of energy. Furthermore, if
performance specifications are expressed in the time domain (as, for example, in terms of an upper bound on the
system impulse response), the absolute integral norm is the appropriate measure (Dahleh and Pearson, 1987). Each
design method is geared toward a particular aspect of system performance as expressed by the corresponding
norm, and it is conceivable that simultaneous suboptimization of all three norms could be achieved, by
appropriately modifying/combining the solutions to each individual norm optimization. Results along these lines
are presently becoming available (Bernstein and Haddad, 1989; Doyle et al., 1989).
However, in processes where parametric variation drastically affects their nominal characteristics, serious
performance degradation or even instability can occur. The need for design under parametric uncertainty/ (slow)
time variation gave rise to the area of adaptive control, which allows for real-time, on-line ''learning'' and
subsequent compensator parameter adjustment for performance, while maintaining overall system stability. This
class of algorithms is very distinct from "gain-scheduled" control algorithms, which are quite often perceived and
referred to as adaptive. Gain-scheduled controllers change the compensator characteristics as the operating
conditions change, in an open-loop fashion, from table look-ups, based on the scheduling parameter(s). On the
other hand, truly adaptive algorithms are real-time learning to update compensator parameters in a feedback
fashion.
Impressive theoretical advances were achieved, by the late 1970s, for adaptive algorithms designed to operate
with the assumption of no modeling (unstructured) errors. However, in the early 1980s such algorithms were
shown to be nonrobust in the presence of unmodeled dynamics and persistent disturbances (Rohrs et al., 1985).
Subsequently, various methods were developed for robustification, with a newly evolved notion for robust
adaptive control, which encompasses robust identification with robust control (Middleton et al., 1988). A
systematic study of the former has resulted in algorithms that identify the processes with guaranteed frequency
domain error bounds but at the expense of excessive—but not prohibitive, nowadays—computational burden that
requires real-time spectral monitoring of pertinent signals in the feedback loop (LaMaire et al., 1991). The
combination of robust identification with a robust controller structure as discussed above, with only infrequent
compensator parameter update, as indicated by the quality of the identification process information, has resulted in
the study of performance/stability robustness of time-varying systems, in particular as they arise in this context. A
number of results only recently became available (Dahleh et al., 1990; Meyer, 1988).
At the same time, analysis and design methods have been developed for nonlinear systems via the use of
appropriate transformations for "nonlinear inversion" and feedback (input-output) linearization.
In conclusion, the current state of the art in control theory provides the designer with a rich environment of
design modules/algorithms that can be ingeniously combined to meet large classes of design challenges. The wind
turbine rotor certainly belongs to the class of problems that can be readily addressed by the existing state of the
art, without the need for new theoretical developments. What is needed in the present problem is a good modeling
effort of the wind turbine/tower combination as well as the environment within which it operates and by which it
is affected.
But with the cost of power electronics decreasing, the variable-speed mode may be an attractive mode of
operation for such machines in the future.
In order for the LCCOE to be kept sufficiently low for wind power to be a competitive energy alternative, the
lifetime of a machine would extend over a period of up to 30 years and should be operating at a maximum of
"economic" performance. This presupposes that fatigue during "normal" operation is minimized while appropriate
measures are taken to avoid accidents (e.g., blade breakage, tower hits) during abrupt environmental changes.
Assuming that the structural, material, mechanical, and manufacturing aspects of the process are sufficient,
additional benefits can be gained by appropriately controlling the blade for minimization of wind-loading effects
and gust alleviation as well as maximizing efficiency of power production under variable wind speed profiles.
Aerodynamic and mechanical solutions in the form of stall-controlled blades that limit CLmas, teetered hub
versus rigid hub, and optimum choices of materials have been successful as one method of addressing problems
cited above. In addition, other less complex, potentially less costly control schemes are under development,
including passive control schemes. However, the full achievable potential offered by active control for
improvement in the overall performance of wind turbines has not yet been fully exploited.
The existing active (feedback) control strategies known to the committee have for the most part addressed the
issue of energy production (McNerney, 1981; Vachon, 1987; Ralph, 1989), although Vachon (1987) also
addresses fatigue issues primarily for vertical axis machines. Others are deemed proprietary and were not available
to the committee. The designed controllers referred to above are of the proportional integral (P-I) type and are
rather ad hoc. Typically, P-I action is taken on the basis of wind speed information—averaged over sampling
windows of various lengths—in conjunction with the turbine power curve. The latter provides the equivalent of a
command input to the overall system for desired performance, usually in terms of kilowatt-hours. At Sandia
National Laboratories, different algorithms have been tested for evaluation in a controlled simulation
environment. These are discrete wind speed; moving wind speed; moving power; and discrete, double-power.
These four algorithms have two adjustable parameters for optimum operation: cut-in/cut-off threshold (of wind
speed) and test window (for averaging). As they stand, no real safeguards for robustness are built in to the
algorithms, particularly in response to severe wind gusts. This is so because the system is set to respond to
averaged time information of wind speed profiles, in which high-frequency information is suppressed; certainly
this is true for the fast wind gust characteristic times. Thus, although statistical predictive capability exists,
application of a modern control design can further help anticipate or, at a minimum, initiate a fast system response
on an individual system basis so as to alleviate a severe wind profile that may be building up. Thus, alleviating
wind-loading effects under nominal operation can be more systematically addressed in real time for each specific
wind turbine.
performance requirements for control and have, as a by-product, a system operational fault tolerance resulting from
simpler control functions and implementation; moreover, it will be cost-effective in the long run.
Consequently, active and passive control strategies are not mutually exclusive, but, when both are
synergistically employed, they are capable of enhancing the overall system capabilities by alleviating the
requirements on either or both component strategies. Thus, actuator loads in an active setting could be greatly
alleviated if operating in a passively tailored structure. Suffice it to say that neither approach alone is capable of
delivering performance equivalent to that obtainable by the synergy of both.
A passive approach, on the one hand, is effected through a judicious hardware/material design and thus may
not be adaptable to changing requirements or dramatically changing operating conditions; "passive tailoring" of
the system cannot be done on-line, if the thresholds of a passive design have been exceeded. Moreover, passively
designed structures are not easily modifiable to changing requirements/forcing functions with any reliable
guarantees for consistent performance. Furthermore, the design margins of a passive control structure in response
to dynamic exogenous inputs can be severely limited in comparison with an active control computer-implemented
strategy that is easily modifiable as needed. On the other hand, an active controller may be considerably limited by
state-of-the-art actuator/sensor bandwidths, which may constitute an "unbreakable barrier" to system performance.
Certainly, weight, cost, and compatibility with an existing system also factor in but to a more manageable extent
(i.e., provided the system setup/architecture is modifiable for tolerance of such factors) once satisfactory
performance can be assured by the components in question. Ultimately, this is an area where trade-offs have to be
systematically worked out and weighed.
Before any detailed study of relative benefits can proceed, however, developing a good model of the overall
process intended for control is of utmost importance. This seems to be the crux of a successful engineering
endeavor, particularly in the wind turbine environment, which involves interaction of fluids (possibly unsteady
aerodynamics) with electromechanical structures. For the purposes of alleviating wind gust effects and/or
undesirable vibrations due to the interaction of the blade and unsteady aerodynamics (turbulence increases with
deeper location in the wind farm), mass flow or pressure sensors need to be placed at judiciously chosen locations
(perhaps even in the blade), in order to measure (precursors to) turbulent upcoming events (i.e., wind gusts). Such
sensor bandwidth should be extremely high given the nature of the latter. The sensor signal can be used to generate
an appropriate actuator command, via a microprocessor-implemented control algorithm. Actuation can be effected
through incorporation of aileron-like surfaces on the main blade structure. Thus, fast response can be achieved
locally, with very limited control actions, which will be effective in modifying the boundary layer of the unsteady
fluid process, thus alleviating excessive loading on the blade. On a more global actuation scale, control of
variable-speed machines through their power electronics effected via the rotor does appear to have the potential of
tailoring the blade structure to operating in close to optimal configuration in adverse wind profiles.
Based on experience with higher harmonic control (HHC) for helicopter vibrations (Goldenthal et al., 1987)
and, more recently, on active control of rotating stall and surge in compressors (Paduano et al., 1990), such models
of the fluid processes are readily obtainable. Sensor data of mass flow perturbations (velocity, pressure) can be
used while the perturbations are small (i.e., the nonlinear [unsteady] process is still at its inception and it evolves in a
close to linear fashion for a short period of time). Control designs based on models thusly obtained have proven
successful for HHC and for active surge and rotating stall control. Experience gained there can be brought to bear
in the wind turbine environment, given the basic similarities of the processes involved.
In fact, the dynamic models obtained for the fluid processes in the areas mentioned above are as simple as
low-order, linear, time-invariant (even scalar) systems that admit to the simplest control algorithm synthesis: fixed
parameter compensation that can be derived via the simplest theory, even classical approaches. Certainly, an
ingeniously designed lead/lag network will work effectively, whether derived via classical considerations or, more
formally, via a modern methodology such as the linear quadratic Gaussian with loop transfer recovery.
Computational and implementation/hardware requirements are within existing state-of-the-art capabilities,
and, as already stated, no new control theory is necessary for a successful outcome. The emphasis here is primarily
on a good modeling effort. However, the literature is severely lacking in complete methodologies for systematic
modeling of the effect on the machine of such fluid processes. Given its importance, this is an area that needs to be
mathematically formulated and formalized.
The foregoing suggestions are immediately implementable as a technology transfer from other related
industries (Paduano et al., 1990; Goldenthal et al., 1987). To be precise, the proof of concept in a realistic engine
(laboratory) setting was so recent that industrial production can only be in the very near future and is certainly not
in existence at the present time. However, implementation details in an industrial setting need to be worked out
specifically within the particular application industry at hand.
Further down the line, for future development, smart materials (e.g., piezoelectric sensors and actuators
embedded in them) may result in similar performance/operation improvements for wind turbines (Spangler and
Hall, 1990). In such a case, an essential continuum (or distributed network) of sensors and actuators results in
appropriate blade twisting so as to alleviate the effect of the aforementioned undesirable wind disturbances. This
technology is under development, actively being researched at this time in large space structures groups and
materials science laboratories. Preliminary results are encouraging (deLuis et al., 1989), although the problem of
forcing such a distributed structure into a desirable coordinated configuration via local sensor feedback entails
complexities and performance robustness risks. This is an area that requires further investigation. One should also
note that the existing preliminary results in large space structures rely on the overall passivity of a flexible
structure. However, the wind turbine environment is considerably more severe.
In conclusion, a systematic application of readily available control technology can offer substantial
advantages in enabling the wind power industry to realize its long-term objectives and become a competitive
energy alternative. More specifically, the following are recommended control technology areas for research and
development in the immediate (near) term:
1. Continue investigating viable models for the fluid structure interaction, together with signal-
processing algorithms for valid signal extraction.
2. Develop other active control techniques and incorporate the best response means (e.g., aileron
surfaces and boundary layer control).
3. Design control algorithms for wind loading/gust alleviation.
4. Develop models and algorithms for control of variable-speed machines to optimize energy production
versus fatigue life.
Stein, G. 1985. Beyond Singular Values and Loop Shapes. Proceedings of the American Control Conference, Boston, Massachusetts, June.
Stein G., and M. Athans. 1987. The LQG/LTR Procedure for Multivariable Feedback Control Design. IEEE Transactions on Automatic
Control, Vol. AC-32, No. 2, February, pp. 105-114.
Vachon, W. A. 1987. The Effects of Control Algorithms on Fatigue Life and Energy Production of Vertical Axis Wind Turbines. Vachon &
Associates Report .
7
CONCLUSIONS AND RECOMMENDATIONS
CONCLUSIONS
Wind turbine technology has demonstrated the potential for contributing to the energy needs of the United
States. If the sites with acceptable wind characteristics were fully utilized, they could contribute up to about 10
percent of the nation's electrical energy needs. The limitation is based on utility system stability issues rather than
available site locations. As in all energy investment decisions, the ultimate penetration level will be driven by the
cost of energy that is produced. In turn, this is decided by the initial cost of the wind energy plant and the annual
cost for maintenance and operation.
Since a number of U.S. electric power utilities are continuing to add capacity, there will be an opportunity to
introduce a new, longer-lasting design for a wind turbine system. Moreover, renewed interest by the public in
environmental issues associated with power generation gives a special advantage to wind power. A new wind
turbine system probably will take advantage of advances in semiconductor power electronics to improve energy
production as well as provide reactive power control, which will make wind-generated electric power more
amenable for use by the electric utilities. New speed control schemes will be introduced, but the major advance
must come through the design of less expensive, longer-lived, and higher-efficiency rotors. A guiding principle in
creating this design should be that knowledge of aerodynamic forces must be carefully integrated with the
structural response of the material, all balanced by the practicalities of field experience and tempered by the need
to manufacture a consistently high-quality product at reasonable cost.
This committee has examined the experience base accumulated by wind turbines, and the accompanying R&D
programs sponsored by the Department of Energy and has concluded that a wind energy system such as described
above is within the capability of engineering practice. However, certain gaps in knowledge exist. The achievement
of this goal without costly and inefficient trial and error requires certain critical research and development.
Because of the fragile nature of the wind power equipment producers in the United States, this will require an R&D
investment by the Department of Energy.
The committee cannot conclude without commenting on the status of the wind power equipment industry.
Because of the decrease in the number of machines installed in the past 5 years, since the tax incentives expired,
there currently is only one major integrated manufacturer in the United States. In addition, only a few companies
are actively producing blades. Moreover, in recent years, a major Japanese manufacturer has entered the world
market to join the European manufacturers who have been participants for some time. As a result, the U.S.
industry is not in a financial position to engage in the R&D necessary to gain worldwide technological leadership
for what the committee sees as a future growing worldwide market for wind power. The committee believes that
the United States is facing a future major reduction in fossil fuel sources of energy. When this is coupled with a
resurgence of public concern for environmental issues in energy production, the need to develop wind power
energy to the fullest extent possible seems compelling.
In the recommendations below, specific research tasks are listed that need to be carried out. Within each
category of research—materials, manufacturing, structural response, etc.—these research tasks are listed in
approximate order of priority. However, we wishes to emphasize that the overall goal of an R&D program should
be to develop a system to produce longer-lived, less expensive, and more efficient wind turbine rotors. Increased
knowledge of the fatigue properties and fatigue failure mechanisms of blades should take precedence, but this
cannot be separated from the search for better manufacturing processes or from design innovations that will either
minimize the likelihood of failure or ease the aerodynamic constraints of blade shape that impede process
innovation. The committee wish to emphasize that the four factors of fatigue, manufacturing, advanced materials,
and design are closely interrelated in the quest to produce a more cost-effective blade.
RESEARCH RECOMMENDATIONS
Recommendations are classified with respect to four goals. More detailed recommendations will be found at
the ends of Chapters 2, 3, 4, 5, and 6.
Goal 1: To improve the material properties and design capability so that the structure will withstand higher
stresses, or the same level of stress for a much longer period of time.
1. Long-term fatigue data should be developed for the most common glass-reinforced plastics (GRP)
laminates and critical elements under appropriate environmental conditions. The data should be
carried to 108 to 109 cycles if possible, at stress ratios of R = 0.1 (tension and compression) and R = -1
(fully reversed). An extensive search of all fatigue data on GRP composites should be conducted and
published in a source convenient to blade designers. This should evolve into a databank of wind
turbine blade materials.
2. The extensive compendium of mechanical data, including fatigue data, on wood/epoxy laminates
should be published and made available to domestic blade designers. Very high cycle (108 to 109)
fatigue tests should be conducted on this material to compare its fatigue response with that of GRP
materials.
3. The potential benefits for significant weight reduction in blades (about 50 to 70 percent) while
maintaining required stiffness through the use of hybrid composites (in which carbon or aramid fibers
are placed in critical blade locations) should be explored through design studies and limited blade
testing. Critical use of cost models must be a requirement in this work because of the strong industry
reluctance to utilize materials that are more expensive than E-glass/vinyl ester or wood/epoxy.
1. Simple cross-sectional analyses and computer codes need to be developed for determination of
sectional elastic constants in composite blades with elastic couplings. These design tools should
consider blade parameters such as curvature, twist, taper, and, above all, completely general material/
geometry.
2. An aeroelastic design code should be developed for wind turbine blades. This would permit the
investigation of aeroelastic tailoring as a passive control mechanism.
3. An investigation of new active control techniques for wind turbine blades should be initiated. This
should be aimed at a new generation design in which gust loads are essentially reduced, thereby
minimizing over-design of blades.
Goal 3: To improve the blade manufacturing process so that quality variations and cost are minimized.
1. The resin transfer molding process has demonstrated the capability of producing quality fan blades up
to 40 feet in diameter. Prototype studies to make GRP blades by this process should be undertaken.
The study must include trade-off studies of manufacturing cost and quality versus losses in
aerodynamic efficiency to enhance producibility.
2. The pultrusion process has a demonstrated capability to produce low-cost GRP blades but with a
relatively inefficient constant cross section. Aeroelastic tailoring may partially compensate for the lack
of twist and tapered planform compared to usual wind turbine blade geometries. A feasibility study
should be conducted.
3. The introduction of new manufacturing processes must be accompanied by fatigue testing of full-size
blades. Only in this way can the design details and the material quality be validated for the
manufacturing process. Baseline studies on blades produced by current practice are needed.
Goal 4: To reduce the cost of blades enough so that periodic replacementbecomes cost-effective.
1. Because fatigue crack propagation is slow in GRP composites, and is believed to be readily visible on
annual tower-top inspections, the strategy described by goal 4 may be feasible. This would eliminate
the need for extensive and expensive high-cycle fatigue testing. Eliminating the uncertainties in
accounting for long-life service in design calculations would result in considerable weight savings. A
detailed feasibility study with a realistic cost model should be undertaken if the manufacturing studies
show promise of significant reduction in blade life-cycle cost.
APPENDIX A 105
APPENDIX A
STATEMENT OF TASK
OBJECTIVES
The final technical report shall provide state-of-the-art information via a current survey of wind turbine
materials, fabrication, operation, and life performance. Research areas shall be identified and given priority. Also,
through an evaluation of research related to stress/deformation response, direction shall be given to Department of
Energy (DOE) for the development of models that will aid in the understanding of the stress problem. By
considering research related to failure of joints, fasteners, and critical sections, applied research in material
sciences shall be brought to the attention of the DOE program manager.
TASKS TO BE PERFORMED
The committee will assess:
a. The adequacy of existing models to predict dynamic stress patterns. A current preliminary
understanding of wind environment forcing functions shall be obtained from DOE.
b. The understanding of the structural stress/deformation response of various wind turbine materials
subject to representative forcing functions, with particular reference to dynamic and fatigue failure.
c. The understanding of the performance of joints, fasteners, and critical sections in relation to failure
modes and fractures and how they originate and propagate.
d. The effects of environmental degradation, including such factors as wear and erosion, corrosion, and
ultraviolet degradation over time in wind turbine materials.
e. The research tools needed to study these phenomena, such as sensors, computer hardware, software,
and databases.
f. The opportunities for new materials and coatings to improve wind turbine life performance.
g. The need for special laboratory facilities, models, and prototypes to improve the design and operation
of wind energy systems.
h. The prospects for improving the ability to predict wind turbine life and improved economics as a
function of material choice, design fabrication, and life-cycle maintenance factors.
ASSESSMENT OUTCOME
This work will influence and improve DOE's plan for wind energy research described in the latest version of
"Federal Wind Energy Program: Five-Year Research Plan." Policymakers will be provided with an improved
understanding of the role of wind technology research in relation to national policies on energy science and
technology.
APPENDIX A 106
APPENDIX B 107
APPENDIX B
COMMITTEE MEETINGS AND ACTIVITIES
APPENDIX B 108