You are on page 1of 272

DESIGN OF

COMPOSITE
STRUCTURES
AGAINST
FATIGUE
Applications t o Wind Turbine Blades
Edited by R M M a y e r

Warns

ISP*

Design of Composite Structures Against Fatigue

> is*

CU\ .

Design of Composite Structures Against


Fatigue
Applications to Wind Turbine Blades

Edited by
Rayner M Mayer
BSc, MSc, PhD, CEng, MIMechE

BUS

First published 1996


This publication is copyright under the Berne Convention and the International Copyright Convention. All rights reserved. Apart from any fair
dealing for the purpose of private study, research, criticism, or review, as
permitted under the Copyright Designs and Patents Act 1988, no part may be
reproduced, stored in a retrieval system, or transmitted in any form or by any
means, electronic, electrical, chemical, mechanical, photocopying, recording
or otherwise, without the prior permission of the copyright owners. Unlicensed
multiple copying of this publication is illegal. Inquiries should be addressed
to: The Managing Editor, Mechanical Engineering Publications Limited,
Northgate Avenue, Bury St Edmunds, Suffolk, IP32 6BW, UK
ISBN 0 85298 957 1
R M Mayer
A CIP catalogue record for this book is available from the British Library.
EUR 16687
The data here are provided in good faith, but neither the authors, the original
providers of the data, nor the sponsors, are able to accept responsibility for the
accuracy of any of the information included, or any of the consequences that
may arise from the use of the data or designs or constructions based on any of
the information supplied or materials described. The inclusion or omission of a
particular material in no way implies anything about its performance with
respect to other materials.
Neither the publishers, the European Commission nor anyone acting on their
behalf are responsible for any statement made in this publication. Data,
discussion and conclusions developed by the Author are for information only
and are not intended for use without independent substantiating investigation
on the part of the potential users. Opinions expressed are those of the Author
and are not necessarily those of the Institution of Mechanical Engineers, its
publishers, or sponsors.
Printed in Great Britain by
Antony Rowe Ltd, Chippenham, Wiltshire
The bulk of the funding for this work was provided under contract JOUR-007 1
with the European Commission (Joule programme).

Contents
About the author

ix

Scope of the book

How to use this book

Preface

xi

Acknowledgements

xii

Notation

xiii

Units

xiv

Chapter 1 Fatigue considerations


1.1 Introduction
1.2 Design and manufacture
1.3 Fatigue considerations
1.4 Structural design
1.5 Materials and evaluation
1.6 Data collection and analysis
1.7 Non-destructive evaluation (NDE)
1.8 Conclusions

1
1
2
4
6
10
11
12
13

Chapter 2 Properties of aligned


2.1 Introduction
2.2 Materials
2.3 Mechanical characterization
2.4 Significance for design
2.5 Conclusions
Chapter 3 Influence of matrix and fabric
3.1 Introduction
3.2 Materials and testing
3.3 Static properties
3.4 Fatigue properties
3.5 Effect of matrix
3.6 Effect of fabrics
3.7 Discussion
3.8 Conclusions

fibres

15
15
16
18
30
31
33
33
33
36
38
43
44
44
49

vi

Contents

Chapter 4 Influence of spectral loading


4.1 Introduction
4.2 The spectrum
4.3 Materials and testing
4.4 Constant amplitude data
4.5 Tests with WISPER/WISPERX
4.6 Discussion
4.7 Conclusions

51
51
51
52
55
59
61
63

Chapter 5 Effects of environment


5.1 Introduction
5.2 Materials and specimens
5.3 Exposure to humidity
5.4 Hailstone simulation
5.5 Testing procedure
5.6 Results
5.7 Statistical evaluation
5.8 Conclusions and recommendations

65
65
65
68
70
72
74
80
86

Chapter 6 Glass and hybridfibreperformance


6.1 Introduction
6.2 Materials and static properties
6.3 Fatigue of glass fibre laminates
6.4 Fatigue of hybrid fibre laminates
6.5 Fatigue of glass/carbon bolted joints
6.6 Conclusions

89
89
90
94
100
103
103

Chapter 7 Fatigue properties of wood composites


7.1 Introduction
7.2 Manufacture
7.3 Advantages
7.4 Constant amplitude data
7.5 Life predictions and the WISPERX spectrum
7.6 Fatigue properties of alternative species
7.7 Effect of joint configuration on fatigue performance
7.8 Infra red condition monitoring of joints
7.9 Conclusions

107
107
107
108
110
115
116
118
119
121

Chapter 8 Benchmark tests


8.1 Introduction
8.2 Material selection
8.3 Flexural testing
8.4 Tensile testing
8.5 Significance of static results
8.6 Fatigue testing

123
123
123
125
126
127
129

Contents

8.7 Analysis of fatigue data


8.8 Recommendations

vii

131
131

Chapter 9 Comparison of coupon and spar tests


9.1 Introduction
9.2 Materials and manufacture
9.3 Material characterization
9.4 Fatigue testing of coupons
9.5 Methods of monitoring damage
9.6 Component testing
9.7 Discussion
9.8 Conclusions

133
133
134
136
137
138
141
145
147

Chapter 10 Response of blade roots to high bending moments


10.1 Introduction
10.2 Experimental details
10.3 Pin-hole
flange
10.4 Trumpet
flange
10.5 T-bolt
flange
10.6 Discussion
10.7 Conclusions

149
149
150
151
156
165
167
170

Chapter 11 Influence of moisture on GFRP bolted joints


11.1 Introduction
11.2 Effect of moisture
11.3 Materials and methods
11.4 Results and discussion
11.5 Conclusions

171
171
171
173
176
180

Chapter 12 Influence of complex loading on blade-root joints


12.1 Introduction
12.2 Description of tests
12.3 Pin-hole
flange
12.4 Rectangular blade-root
12.5 Conclusions

181
181
182
188
189
193

Chapter 13 Evaluation of T-bolt root attachment


13.1 Introduction
13.2 Load attachment principle, FEA-simulation
13.3 Tests on ENERCON rotor blades
13.4 Discussion
13.5 Conclusions

195
195
195
198
203
207

Chapter 14 Comparison of fatigue curves for glass composite


laminates
14.1 Introduction
14.2 Linear regression analysis

209
209
210

viii

Contents

14.3
14.4
14.5
14.6
14.7
14.8
14.9

Reverse loading
Tensile loading
Compressive loading
Flexural loading
Other analysis methods
Comparing new materials against the standard fatigue curve
Conclusions

213
218
220
220
222
222
224

Chapter 15 Recommendations for good-working practices, norms


and standards
15.1 Introduction
15.2 Material selection and characterization
15.3 Design
15.4 Manufacture
15.5 Coupon testing
15.6 Structural testing
15.7 Type approval
15.8 Conclusions

227
227
227
229
229
231
235
236
237

Glossary

239

Index

243

Principal authors and addresses

247

About the author


Rayner M. Mayer is the principal consultant with Sciotech, a company
involved with product innovation and development. He obtained his PhD at
Cambridge University and is a member of the Institution of Mechanical
Engineers and a chartered engineer. He is widely published in technical and
scientific journals.
The consultancy specializes in the application of fibre reinforced plastics to
load bearing components and structures for energy and transport applications.
He is responsible for the scientific and co-ordination of collaborative European
energy research and development programmes, such as Joule and Thermie.

Scope of the Book


The first chapter outlines the process for designing with composite materials
and how fatigue impacts on design. This follows the engineering design
standard, BS 7000, A Guide to Managing Product Design.
The principal considerations are discussed in the next six chapters (Chapters
2-7), starting with the fibre orientation and considering, in turn, the influence
of: the fibres; matrix; fabrics; spectral loading; moisture; impact; hybrid
fabrics; and wood veneers.
The various coupon test methods are then compared by way of a benchmark
test using the same material. The effect of scale is considered in Chapter 9 when
results from coupons and components are contrasted.
Aspects of structural testing of full-size components are described in
Chapters 9-12. Both the methodology and the ability to monitor the damage is
discussed.
The final two chapters summarize the results from two viewpoints - that of
design and of good working practice.

How to use this Book


Information can be sought at various levels.
For those not acquainted with the mechanisms of the fatigue process,
Chapter 1 should be consulted. The design strategy is outlined, together with
how to assess the type of tests needed for materials and structures.
Specific design aspects are then considered in Chapters 2-9, starting with the
orientation of the reinforcement and type of matrix and ending with the
influence of the environment. In Chapters 913 aspects of structural testing and
verification of the design are discussed.
Whilst design information is derived throughout the text, the coupon data is
collated and analysed in Chapter 14. The standard curves can form the starting
point for design purposes. The recommendations in Chapter 15 form a design
check list; no doubt designers will add to this list as their experience grows. If
the text is to be used as a source of information then the index, list of contents,
or summary tables in Chapter 1 should be consulted.
A glossary is provided to define the technical terms used within the text and
the industry. It is consistent with that adopted by the American ASM International Handbook Committee being the most authoritative source available.

Preface
The use of composite materials for load-bearing structures is gradually gaining
acceptance as knowledge is gained about their performance and durability.
The idea is intrinsically simple - to utilize two (or more) constituents, which
together have more attractive properties than the individual constituents.
In this book, the use of two types of composite materials is considered;
namely fibre reinforced plastics and, to a lesser extent, wood-composites. The
question of how these materials can be successfully incorporated into primary
load-bearing structures is also addressed.
The study is applied to the inclusion of such materials into wind turbine
blades, which experience both static and dynamic (fatigue) loading. It is not
merely the loads that must be withstood, but also the effects of the environment
over prolonged periods of time.
This work describes a systematic approach by engineers from seven European countries over a period of five years to elucidate the material and
structural response to fatigue loading. Methods of detecting the accumulation
of damage over long periods of time are also discussed.
The information presented in this book will assist in the design of composite
structures against fatigue, though it will be necessary to characterize the
material combination and manufacturing process used at the detailed design
stage.
With its low environmental impact, the generation of electricity from
renewal energy sources will increasingly provide for the needs for the next
millenium. This study will allow, inter alia, the next generation of wind
turbines to be developed based on new designs of rotor blades.
Whilst this book is self-contained, the accompanying volumes by the editor
(Design with Reinforced Plastics and Design Data for Reinforced Plastics,
(Chapman & Hall) may be consulted if readers are not familiar with the design
process.
As this work is on-going, we would hope that our experience and those of our
readers could be incorporated in any future editions.
Ray ner M Mayer
Yateley 1995

Acknowledgements
Funding support was provided by:
European Commission Directorate General XII - Science, Research,
Development; contract number JOUR-0071-DK
Danish Energy Agency
Department of Trade and Industry, UK
Jotun Polymer A/S, Norway
Norges Vassdrags - og Energiverk, Norway
NOVEM, Netherlands
Research Council of Norway
Renewable Energy Centre, NEL, UK
Swedish National Board for Industrial and Technical Development
Vlaamse Gemeenschap, Dept. Economie, Belgium
Werkgelegenheid en Binnenlandse Aangelegenheden, Belgium
and the Institutes and Companies to whom the authors are affiliated.
We wish to acknowledge the support and assistance of the following:
Giancarlo Caratti and Komninos Diamantaras, EC.DG XII
Ragnar Arvesen, Jotun Polymer, Norway
Geoffrey Dutton, Rutherford Appleton Laboratory, UK
Mark Hancock, Wind Energy Group Limited, UK
Monica Jong, ECN, Netherlands
W. Kurz, Aerodynamik Consult Company, Germany
David Mayer, Sciotech, UK
Jens-Peter Molly, DEWI, Germany
Hans Reiter, DLR, Germany
David Richmond, Flemings Industrial Fabrics, UK
B.A.J. Schaap, ECN, Netherlands

Notation
Fabric and laminate notation
Process
SB - stitch bonding
FW - filament winding
W
- weaving
I
- inlaid
CSM - chopped strand mat
Notation numbers
(numbers)
letters
+
s
[p]

alignment angle or plies


mass in g/m2
method of fabrication
combination fabric
symmetrical
number of plies

Example 1 Combination fabric.


0/90
/
alignment

SB (800)
/
process

CSM (100)

\
\
mass combination

process

mass

Example 2 Biased fabric.

/ \ \ w
0

(567) /

90 (35) W

alignment mass alignment

mass process

Example 3 Plied laminate.


[45 W (250) /
/
\
\
alignment process

0 (150) [4p]]s

/
\
mass plies symmetrical

Units
Units

Properties
Modulus
(a)-(d)
Stress

Stress

tensile
flexural
compressive
torsional
tensile
flexural
compressive
torsional
interlaminar shear
minimum (e)
maximum
mean
amplitude
minimum/maximum (f)

Strain (g)
Poissons ratio
Barcol hardness
Toughness (impact strength)
Fibre fraction - by volume
- by weight
Water absorption
Glass transition temperature

GPa
GPa
GPa
GPa
MPa
MPa
MPa
MPa
MPa
MPa
MPa
MPa
MPa

J/m or

Designator

Ef

Ec
G
,
Of
Oc

Omin
Oman
Ornean
Oall

Bn
J/m2

%
%

Vf

%
C

w,
Tg

Notes
(a) If a stress-strain curve is non-linear then one can measure the initial slope
(tangent, Fig. 0.1) or the secant slope. For the latter, the strain range
should be defined if it is not taken between zero and the failure strain.
(b) Subscripts t, c, and f are used to designate tensile, compressive and flexural
loading, respectively.
(c) Subscript f may also be used to designate the fatigue value of the modulus
once the loading has been specified, e.g. Etf.
(d) Parameters may be normalized by dividing their current value by their
initial value. For example, moduli may be normalized by dividing Elf by ,.

Units

Tangent

S^T

<

>^,
<>
//'

y^

Secant

y
,

^^r"""
j

1
^Linear limit

Strain ()
Fig. 0.1 Determination of the modulus from the tangent or secant slope of a stress-strain curve

Time
Fig. 0.2 Definition of the parameters used to define an alternating stress of constant amplitude.
<Tmi is always the most negative stress value, even if the stresses are negative

Units

/Strain amplitude

/Strain range

Time
Fig. 0.3 Definition of parameters used to define strain. As with <r,. c,h, is always the most
negative strain

(e) The relationship between the various stress parameters (Fig. 0.2) are
"mean
*-*^ \Vmax ' ^rnin/
tfalt = 5 ("max ~ Omi)
**
"minimax
=

(Omcan ~ OaltV^mcan +

CT

alt)

(f) Common values of the R ratio are


R = 0.1 (tension/tension) or (flexure/flexure)
= 10 (compression/compression)
= 1 reverse loading; (tension/compression)
(g) the definition of the various strain parameters is illustrated in Fig. 0.3.

CHAPTER 1

Fatigue Considerations
ft M. Mayer*
Composite materials provide an optimum design solution to fabricating structures that move or
rotate. The design principles are discussed with reference to fatigue loading. The selection and
manufacture of such materials is outlined, together with the various aspects of fatigue that have
been investigated.

1.1 Introduction
The composite materials discussed in this volume refer to two principal classes
of materials, namely polymers reinforced by either fibres or wood veneers. In
wood, the reinforcement is the cellulose microfibrils in the cell walls. The use of
such reinforcements produces a range of properties which cannot be sustained
by the matrix itself, such as stiffness, strength, and fatigue resistance (1)(2).
The reinforcement provides the mechanical strength, whereas the matrix
provides the means of transferring the load into and out of the reinforcement
and protecting it against the environment. However, the choice of such
materials is restricted by design requirements and consideration of shape,
processing ability, and cost.
In general, the stiffness and strength of a composite increase with reinforcement fraction up to some limiting value. For glass reinforced plastic (GRP) the
reinforcement is typically 40-50 percent by volume, whilst for wood composites it is in excess of 80 percent as the glue line has negligible volume.
Nevertheless, the material properties of these two groups are much closer to
one another than to metals since they are both composite materials whose
properties only arise as a result of lay-up.
Common properties of wood andfibrecomposites include: high strength and
stiffness to density ratios; good environmental resistance; suitability for use in
the rapid manufacture of large structures. Moreover, moulding direct to final
shape facilitates assembly and minimizes wastage.
The materials and structures discussed in this publication are typical of
general applications in mechanical engineering in which composites are only
slowly penetrating. As these have not yet been investigated to any significant
extent, designers have hitherto been very conservative in their designs.
Sciotech, UK.

Design of Composite Structures Against Fatigue

Laminated wooden beams have been used for some time as structural
members for largespan trusses and roofs such as those in sport halls or
swimming pools. These comprise wooden planks, typically 1250 mm thick,
glued together to give the appropriately shaped beam. Their loading is
primarily static in nature. For complex shaped structures such as wind turbine
blades, wood veneers are used, typically 35 mm thick, and the loading is
primarily dynamic in character.
1.2 Design and manufacture
Structures can be designed on the basis of British Standard BS 7000, in which
the design is advanced in successive stages from conceptual through embodi
ment to detailing and issuing instructions for manufacture (3). The design
sequence for proceeding to the embodiment (or layout) stage is shown in Fig.
1.1 (4).
The nature of composite materials is such that the selection of manufacturing
processes, materials, and properties are interrelated; choice of any one may
define the other two! This reduces the design options by restricting the number
of possible material combinations.
The fundamental design rule is to lay the reinforcement in the direction of
the principal stresses and to check that the stresses in the other directions are
such that the matrix has sufficient strength to withstand the loading. If the
matrix alone cannot withstand the subsidiary stresses then the reinforcement
also needs to be laid in those directions.
As the design evolves, one must check forfitnessof purpose; if this cannot be
achieved, the structure, materials, or manufacturing process may need to be
reassessed. As the sequence is iterated in successively greater depth, it is
necessary to ensure that there is no fundamental difficulty which would prevent
the design from being released for manufacture.

Manufacture
The design of composite structures involves the interrelation of structural
considerations, material properties, and processing. The processes listed in
Table 1.1 are capable of producing large structures, though some of these have
restrictions on the geometrical shape (4).
The most common process, and one that is also the most labour intensive, is
that of contact moulding, in which the dry reinforcement (fabric or veneer) is
laid up in a mould and is then impregnated with resin and consolidated. In
prepreg moulding the reinforcement and resin are already intimately mixed
and prepared in sheet form by the material supplier.
Either fibres or fabric in the form of tape are used in the winding technique
and these are passed through a resin bath before being laid onto a mandrel at
the appropriate angle. They have a low labour content as automation is

Fatigue Considerations
Conceptual design

_-rzr

Materials, processes

Shape, form

Load path, joints

Resin, reinforcement type and lay-up

Strength, stiffness

Mass

Durability

Environmental impact

L
Fig. 1.1

Embodiment design
The steps in establishing the embodiment from the conceptual design; dotted
lines indicate stop and think before iterating the loop. The process needs to be
repeated for the detailed design stage (4)

Table 1.1

Principal manufacturing processes


for large structures
Contact moulding
Filament winding
Tape winding
Prepreg moulding
Resin transfer moulding

Design of Composite Structures Against Fatigue

possible; spars for wind turbines of up to 40 m in length have been manufactured by this method.
Resin transfer moulding involves laying up the reinforcement in the form of
fabrics inside a mould, closing the mould, and injecting the catalyzed resin
mixture. Once the resin has hardened, the component can be demoulded with
good surfaces all round.
It is essential that components and structures are evolved which can be
moulded or manufactured directly to final shape. In this way no machining is
required and the outer (moulded) surface can act as an environmental barrier.
Jointing
The principal methods of jointing are glueing and bolting, and careful attention
to design detail is required in order to ensure adequate load transfer between
components. A key rule is that joints between dissimilar materials (such as
composites and metals) should be located in areas of low stressing if at all
possible.
1.3 Fatigue considerations
The fundamental response of composite materials to fluctuating loading
(fatigue) is known (5) and has been characterized for some specific material
combinations like carbon fibres in an epoxy matrix for aerospace applications.
As with metals, the load bearing capability decreases as the number of fatigue
cycles increases.
However, the way in which the damage nucleates and grows is very different.
In metals, a crack (or cracks) is nucleated and damage increases by crack
growth. Owing to the anisotropic nature of composites, one tends tofindareas
of damage arising, which can grow and eventually lead to failure.
Following Reifsnider (5), it is believed that damage accumulates in three
stages of varying time length (Fig. 1.2).
During the initial load period (stage 1), there is generally a small drop in
stiffness associated with the formation of some damage. This is followed (stage
2) by a much longer time period in which the damage seems to increase linearly
with time and the stiffness falls very gradually. If the stress is sufficiently high, a
third stage (3) is observed which is characterized by an ever increasing amount
of damage which ultimately leads to failure.
Substantiation
Fatigue can be investigated in various ways:
(a) increasing the size of the test piece from coupon through to component;
(b) increasing the complexity of loading from constant amplitude through to
loading encountered in service;
(c) or a combination of (a) and (b) (Fig. 1.3).

Fatigue Considerations

' Residual strength

^
.Stiffness

/
-+-3

10'

10

10*

IV

IV

- Stages

IV

Cycles
Fig. 1.2

The three characteristic stages of damage of composites (5)

The ultimate goal is to apply the service loading to the complete structure
(6). Such tests are complex and time consuming, but provide information
which could only otherwise be obtained by evaluation in service.
The majority of tests are always performed on coupons for one or more of the
following reasons:
-

low cost;
ability to use standard test machines and fixtures;
ease of testing;
prospect of testing at high frequencies circa 5 Hz or higher because of the
lower loads;
rapid investigation of various materials and lay-ups;
response of a material to various types of loading;
effects of environment;
ability to obtain design allowable values for fatigue.

All this information is necessary in order to be able to design a structure against


fatigue.

Design of Composite Structures Against Fatigue


Increasing complexity
of specimen

Goal

Structures

Components

Specimens

Basic
fatigue
data

Coupons

Constant
amplitude

Fig. 1.3

Variable
amplitude

Service
loads

Increasing
complexity
of applied
loads

Possibilities of fatigue testing in terms of increasing complexity of loads and components


(6)

The inverse relationship between the number of tests, their complexity, and
cost (called the pyramid of substantiation) (Fig. 1.4) was originally developed
for aircraft components (7).
A similar strategy has been adopted in the current investigations, that is a
large number of tests on coupons and a much smaller number on components
and structures.
1.4 Structural design
The design sequence for a load-bearing structure such as a rotor blade is
illustratedin Fig. 1.5.
The aerodynamic, structural, and manufacturing considerations are intertwined. Aerodynamics will set the shape envelope from which the loads can be

Fatigue Considerations

40
Number of tests
Fig. 1.4

60

80

100

Relative cost (%)

Substantiation of the design indicating the number of tests and the related costs for each
stage of testing of a composite component for Airbus Industrie (7)

Concept
Regulation
Blade number

Location
Terrain
Wind conditions

Load Calculations
Simulations

Static strength
analysis

zz

Fatigue
spectrum

Fatigue life
prediction

Estimated life

Fig. 1.5

The steps in establishing the ability of a rotor blade to withstand the imposed loadings

Design of Composite Structures Against Fatigue

calculated for a specific wind regime. The blade structure has to withstand the
imposed loads and transfer these loads to the shaft. The manufacturing process
then has to ensure the production of the desired shape and the location of the
reinforcement in appropriate directions.
Material selection is dependent on the manufacturing process, and this in
turn enables both static and fatigue properties to be determined. From these
data, fatigue lives can be predicted from the knowledge of the load spectrum
(Chapter 4). The structure has to withstand these loads with an adequate
margin of safety (10). If the loading is too high, then the design has to be
iterated.
There are a number of codes governing the general principles of structural
design, such as ISO 2394 and the Eurocodes (9)-(ll). These are based on limit
state analysis in which the designer has to identify the ways in which a structure
fails to fulfil its function in terms of either ultimate loads or service loads.
The uncertainties in loads and materials are considered by using partial
safety coefficients. The value of these coefficients depends upon aspects such
as:
-

material variability.
whether the damage is progressive or catastrophic.
whether the design is verified by testing.
inspection in service.
repairability.
lifetime reliability target.

The methods by which some of these aspects can be determined is illustrated in


subsequent chapters.
For wind turbines, the safety requirements have been set out in a new
international design code IEC 1400-1 (12). It specifies inter alia the limits for
the statistical analysis of the fatigue strength, and also makes allowance for
whether a structure fails in a safe manner.
Rotor blade design
Owing to the diffuseness of air, wind turbine blades need to be large in order to
capture any appreciable amount of energy from wind. For example, a turbine
fitted with blades some 17 m in length would typically generate a maximum
power output of 500 kW.
For energy costs to be competitive with other sources, it is essential to design
the blade so that: (a) its mass is effectively used and (b) blades can be
manufactured in a cost-effective manner (8).
In the spar/shell design (Fig. 1.6), the spar is designed to be the prime load
bearing member whilst the shell provides the aerodynamic shape and torsional
stiffness (13). The spar is generally wound by a tape winding process with the
filaments transverse to the length of tape (Fig. 9.2 in Chapter 9), the shell being

Fatigue Considerations
1

Shell

Spar

Glass/polyester

Gelcoat

Fig. 1.6

Spar/shell design developed originally for the blades of the wind turbine erected at Nibe
(13)

Table 1.2
Component
Spar
Blade root

Evaluation of components

Flange type
Trumpet, pin-hole
Pin/hole
T-bolt joint
Flexhat
Trumpet

Purpose oftest

Description
(chapter)

Manufacturing evaluation
Fatigue response
Design verification
Fatigue response
Design verification
Fatigue response

9
10,12
10
13
12
10

made by contact moulding in a mould (4). The advantage of this design is that
each component can be separately optimized.
The stiffened shell is more commonly used with one or more stiffeners in the
form of webs (Fig. 7.2 in Chapter 7 and Fig. 13.3 in Chapter 13). As buckling
can be a problem with such large structures, structural foam is often used
between the skins of the shell to provide sufficient rigidity.
Blade root
The design of the blade root is complex as the torque from the blade has to be
transferred onto the shaft, usually via a flange. There is also a change of
material, from composite to steel and, less commonly, to aluminium. Types in
common use include the trumpet, pin-hole, pre-stressed T-bolt (IKEA type),
and stud (14).
The tests shown in Table 1.2 cover both existing and improved blade root

10

Design of Composite Structures Against Fatigue


Table 1.3

Materials and processes investigated

Reinforcement

Resin

Process

Glass
Aligned fibres at various
angles

Filament winding

Transverse filament tape


Combination fabrics

Tape winding
Contact moulding

Description
(chapter)

Ortho-Polyester

2, 4
9
3, 4
3
3
3
6
5

Balanced woven fabrics

Prepreg moulding

Epoxy
Iso-polyester
Iso-polyester
Iso-NPG polyester
Vinyl ester
Epoxy
Iso-polyester
Epoxy

Glass/carbon
Discrete layers of fabrics

Contact moulding-

Iso-polyester

Contact moulding

Epoxy

Woven and stitch


bonded fabrics

Wood
Khaya
Poplar
Birch
Beech
Baltic pine

Vacuum bagging

connections, and were evolved to validate designs as well as their structural


response to load. These tests require large rigs, great care in introducing the
load(s), and the imposition of loads greater than the maximum design load, in
order to accelerate the test.
1.5 Materials and evaluation
Materials have been evaluated in the form of test plates manufactured by the
appropriate process. They are typical of those in use in general mechanical
engineering applications. Fabric types have included woven, inlaid, and stitchbonded (4) whilst fibre orientations varied from unidirectional to balanced
fabrics in the warp and weft directions. These are summarized in Table 1.3.
The following design aspects outlined below have been considered.
- Fibre alignment through manufacturing laminate plates by the filament
winding process in a laboratory. Consequently these values will set an upper
limit to what could be achieved in a factory environment (Chapter 2).
- Matrix and fabric construction using various types of matrices with the same
type of reinforcement and various fabrics with the same resin. This provides
a link with other matrix systems such as vinyl ester and epoxy, whose
structural applications are more widely known (Chapter 3).
- Humidity and moisture as laminates can take up water under high humidity
and conversely give up water under very dry conditions. The amount of
water absorbed or desorbed has been measured and the effect on properties
determined. Since these measurements were made on thin coupons, they set

Fatigue Considerations

11

an upper limit to what could occur with structures where the laminates are
thicker and have fewer edges (Chapters 5, 11).
Impact damage of ice particles has been investigated using an 'ice gun'
as well as the combined effect of moisture penetration and damage
(Chapter 5).
Spectral loading on rotor blades is generally evaluated using a load set
averaged over a number of different types of turbines. It is designated
WISPER. The relationship of this spectrum to a truncated spectrum
(WISPERX) and constant amplitude loading has also been investigated
(Chapter 4).
Hybrid fabrics comprise mixtures of glass and carbon fibres, as carbon fibres
are appreciably stiffer than glassfibres.Thus a combination of thesefibresin
a resin matrix could provide advantages in terms of obtaining the best
properties of bothfibres,yet at lower cost than that of carbon alone (Chapter
6).
Bolted joints. Bolting introduces compressive and shear stresses, which are
superimposed on the other loads seen by the blade rotor. This has been
studied for both glass and hybrid fabrics (Chapter 6).
Wood composites. Various design aspects are considered including veneer
type, jointing, and spectral loading (Chapter 7).
Effect of scale on going from coupons to components has been investigated
for tape wound coupons and spars. Difficulties can arise with the quality of
manufacture, the degree of alignment of the fabric or veneer, the method of
load introduction, and the method of testing (Chapter 9).

1.6 Data collection and analysis


There is still incomplete agreement for the static testing of composites and even
less for fatigue testing. Consequently, each institute has tended to develop its
own test methods, which will be internally consistent and valid. For this reason,
a benchmark test has been undertaken to validate the consistency of tensile and
flexural test methods (Chapter 8).
Many engineers represent fatigue results by normalizing a property by its
initial value; this may be stress, modulus or the number of cycles to failure.
Normalized data need to be converted to engineering units in order to establish
permissible values for design purposes. This approach has been developed by
Sims (15) (amongst others) to cover various aspects of fatigue loading, testing,
and component geometry. In this work, strain is generally used. This approach
also facilitates comparison between data sets and helps to establish trends and
design principles.
The scatter in the data can be evaluated by statistical methods, and this is
discussed in Chapter 14. The confidence in the data and the probability of
survival are designated by limits, which are generally prescribed by design
codes (9)-(12). Following IEC 1400-1, these limits have been calculated for

12

Design of Composite Structures Against Fatigue

Table 1.4 Principal NDE techniques


Cause

Description
(chapter)

Multiple
Overall change in stiffness
Local change in stiffness
Change in load-rdeflection curve
Overall change in vibration of component
Local increase in temperature
Local areas which contain damage

2-13
3-9
9, 10, 12,13
2,3,7,9
9
9
7, 9, 10, 13

Technique
Visual
Stiffness change
Strain gauges
Hysteresis in area under damping
Resonant frequency
Temperature rise
Infra-red emission

95 percent survival probability and 95 percent confidence limits and are


given for relevant data sets. Other limits could be used depending upon the
design philosophy. This is discussed further in Chapter 15.
1.7 Non-destructive evaluation (NDE)
A variety of NDE techniques can be used to monitor the accumulation of
damage (Table 1.4). Of these, visual inspection is the most straightforward and
is also the most important. Cracks and damage areas are readily visible as most
of the coupons and components are translucent to transmitted light.
Local strains can be determined by strain gauges and this has been extensively used for all structural testing where damage generally occurs on a local
scale rather than a global scale (Chapter 10). Some institutes also use these for
coupon tests to determine extension rather than using extensometers or crosshead deflection of an actuator.
Stiffness has been monitored throughout the study. This enables the damage
stages to be identified (Fig. 1.2), and any deflection limit to be maintained. For
coupons, changes in the modulus can be determined from the stress-strain
curve, either during fatigue testing or by static loading.
The load-deflection curve which characterizes a material or structure can be
used to determine the hysteresis damping (Fig. 1.7). For composites the
internal damping is much higher than metals and is an indicator of damage in
the material.
Resonant frequency and natural frequency are also two methods of
measuring the damping. They provide information at a global level of the state
of the material or structure at micro-level.
Damage induced in composites will generate heat and as these materials
have poor conductivity it should be generally easy to detect. Infra-red
emissions can be detected with a suitable camera and will locate the areas of
damage and how these will grow. Thermocouples are simple to install and have
been placed in locations where high strains have been detected, possibly by a
strain survey (Chapter 9).

Fatigue Considerations

13

Average dynamic
modulus
Hysteresis
damping

Strain
Fig. 1.7

A typical load-deflection curve showing the average dynamic modulus and the hysteresis
damping

Such NDE techniques can also be used to decide a suitable failure criterion.
As composite structures are generally stiffness-limited compared with metals,
a substantial drop in stiffness may be sufficient for the structure to have
effectively 'failed'.
For wind turbine blades, the majority of which are generally positioned
upwind of the tower to avoid the effect of the tower shadow on the loading, a
sufficient loss of stiffness could, for example, result in insufficient clearance
between the blade and the tower.
1.8 Conclusions
Composite materials like wood-veneer laminates and fibre reinforced plastics
have similar manufacturing techniques and similar properties. The strategy for
identifying the influence of specific fatigue parameters is outlined and the
framework is sketched within which the testing and evaluation has been
undertaken.
Coupon data are analysed in individual chapters and are subsequently
gathered together in Chapter 14 to permit a more detailed statistical analysis.
The principal observations andfindingsare assessed in Chapter 15 to provide a
set of recommendations for good working practices.
References
(1)
(2)
(3)
(4)
(5)

RICHARDSON, T., 1987, Composites: a design guide (Industrial Press, New York).
D I N W O O p i E , J. M., 1981, Timber, its nature and behaviour (Van Nostrand, New York).
Guide to managing product design, BS 7000, 1991 (BSI, Milton Keynes).
MAYER, R. M., 1993, Design with reinforced plastics (Chapman & Hall, London).
REIFSNIDER, K. F. (Editor), 1991, Fatigue of composite materials (Elsevier, Amsterdam).

14

Design of Composite Structures Against Fatigue

(6) HAIBACH, E. 1981, Fatigue data for design applications in materials, experimentation and
design in fatigue, Fatigue Conference.
(7) SCHNEIDER, K. and LANG, R. W., 1990, Secondary source qualification of carbon fibre
prepregs for primary and secondary Airbus structures, 11th SAMPE Conference.
(8) PRETLOVE, A. J. and MAYER, R. M., 1994, Rotor size and mass - the dilemma for
designers of WECS, Wind Engineering, 18, 317-28.
(9) General principles on reliability for structures, ISO 2394,1986 (ISO, Geneva).
(10) Basis of design, Eurocode 1, part 1,1995 (CEN, Brussels).
(11) Design of wooden structures, Eurocode 5, 1995 (CEN, Brussels).
(12) Wind turbine generator systems. Part 1. Safety requirements, IEC 1400-1, 1994 (IEC,
Geneva).
(13) JOHANSEN, . S., LILHOLT, H, and LYSTRUP, Aa, 1980, Wingblades of glass fibre
reinforced polyester for a 630 kW windturbine, Third International Conference on Composite
Materials, (Elsevier, Amsterdam).
(14) SANDBERG, O., 1989, Blade root design, a state of the art survey, (FFA, Stockholm).
(15) SIMS, G. D. and GLADMAN, D. G., 1978, Effect oftest conditions on the fatigue strength
of a glass-fabric laminate. Plastics and Rubber, 1978, p. 122 et seq.

CHAPTER 2

Properties of Composites with Long


Fibres
S. /. Andersen*, H. Lilholt* and Aa. Lystrup*
The fatigue properties of glass/polyester composites are presented in a wide spectrum of fatigue
diagrams under various loading conditions. These diagrams can be used both for comparison of
material configurations and for establishing design allowables. Material parameters are
recorded, e.g., stiffness and hysteresis loops. These serve as monitoring parameters for the state
of damage of the composite materials during fatigue and they can thus be used in combination
with design allowables.

2.1 Introduction
The composite material of glass fibre reinforced polyester is a common
material for wingblades of windturbines. This material has been used and
studied in many configurations over the past 1015 years. Initially, the glass
fibres were used in the configuration of randomlyoriented fibres typically in
the form of chopped strand mats (CSM), in a matrix of polyester. This material
has moderate stiffness and strength, in particular during fatigue loading. It has
only been successfully used for relatively small wingblades (35 metres in
length).
For larger (heavier) and longer wingblades the need for increased stiffness,
to reduce deflections, and for increased fatigue strength, to increase life time, is
apparent. The improved fibre configuration involves using a relatively large
fraction of fibres oriented in the loading direction, i.e., along the length of the
wingblade. Such fibre arrangements provide composite materials of high
stiffness and good fatigue strength. This is important for large wingblades,
where fatigue by gravity loads during rotation becomes increasingly important,
in addition to the loads caused by the (fluctuating) wind.
Materials are investigated based on a selection of glass/polyester composites,
with various combinations of random fibres and oriented (straight) fibres.
These systematic and welldefined arrangements are composed of 0 degree
orientation, orientations, and 0 degree/0 orientations. Of these, the
0 degree/0 arrangement can approximately simulate the random fibre orien
tation (e.g. CSM).
Ris National Laboratory, Denmark.

16

Design of Composite Structures Against Fatigue

Table 2.1 Materials


Glass fibres
Type
Designation
Supplier
Diameter
Density
Stiffness
Strength
Failure strain
Bundle characteristics
Surface treatment/coating

E-glass
RPA38 20/21EC12-300
Skandinavisk Glasfiber
12/im

2.615 g/cm3
70-72 GPa
-2000 MPa
~3 percent
300 tex; 1000filamentsper tow
For polyester

Polyester
Type
Designation
Supplier
Density
Modulus
Ultimate tensile strength
Failure strain
Characteristics

Unsaturated polyester
UP333 and Alpolit UPS294V
Hoechst/Polyplex
-1.213 g/cm3
- 4 GPa
-100 MPa
3-5 percent
General polyester for glassfibres

These straight-fibre configurations of nominally continuous fibres can all be


fabricated by a winding technique (described below) ; this ensures a high degree
of similarity in fabrication of all composite materials, and comparisons can be
made more easily and with greater confidence.
The fatigue loadings on the wingblades are mainly caused by the wind
fluctuations, causing flap-wise bending of the wingblades, and by the gravity
loading, causing edge-wise bending of the wingblades during rotation. The
type of fatigue loading seen by the blades is tension-tension fatigue (R = 0.1)
and compression-compression fatigue (R = 10), both due to wind loads; in
addition tension-compression fatigue (R = 1) is caused by gravity.
During fatigue loading of materials in general, and composites in particular,
the microstructure, i.e.,fibre-configuration,may undergo changes. The effect
of such damage is recorded and studied through changes in the material
stiffness and in the development and change in the hysteresis-loops (stressstrain loops during cyclic loading).
2.2 Materials
Characteristics
The constituent materials used to fabricate the composites are listed in Table
2.1.
The composites for wingblades are represented by several series of glass/

Properties of Composites with Long Fibres

17

polyester composites; series 1 contains well-defined fibre configurations, and


these are fabricated in the laboratory at Ris; series 2 is commercially
manufactured with conventional fibre arrangements (Chapter 8) and delivered
by the company Jotun Polymer. Table 2.2 lists the specifications.
Fabrication
The glass/polyester composites of series 1 are fabricated as laminates by a
laboratory procedure, while the materials of series 2 are industrially made as
plates.
The laboratory fabrication is performed as a simple winding of fibre rovings
onto a flat steel plate; the first step of the winding is done as a 90 degreewinding leading to a nominal ply of straight, aligned fibres. The plate is then
remounted in the winding machine into a position corresponding to the
required fibre orientation of the next ply/plies; this next-step winding is a
90 degree-winding in itself, leading to a ply of the actual orientation. These
steps are repeated to obtain the required plies and stacking sequences. The
winding is performed in the 'wet state', i.e., the fibres are pre-soaked with
polyester before being led onto the plate during the winding.
The thickness of the laminate is adjusted/calibrated by steel spacers placed
on the winding plate, before cover-plates are screwed onto the winding plate;
these cover-plates ensure a flat surface on the laminates.
The curing of the polyester UP333 (for the first series of laminates) is made in
a small autoclave at a pressure of 1-3 bar and a temperature of 30C. The curing
of the polyester UPS294V (for the later series of laminates) is made in a heating
oven at normal pressure and at a temperature of 40C. The cured laminates are
cut to specimens for mechanical testing.
The glass/polyester composites of series 2 are made by hand lay-up according
to the prescribed fibre configurations and stacking sequences, (Tables 2.2 and
8.3).
Specimens (test coupons)
The same specimen size and shape is used for all types of mechanical
characterization.
The normal specimen is shown in Fig. 2.1(a). This specimen is well suited for
nearly all materials and for both loading types (static and fatigue) and for both
loading modes (tension and compression). For composites of high stiffness and
strength, i.e., composites with a large fraction of fibres parallel to the loading
direction, a narrower gauge section is necessary; this specimen is shown in Fig.
2.1(b).
All specimens have end tabs, which are made of glass/epoxy material (print
plate quality), pre-cut to shape, glued onto the composite specimens by epoxy
glue, and cured in a fixture to ensure perfect position and alignment.

18

Design of Composite Structures Against Fatigue

Table 2.2 Composites


Glassipolyester (Ris)
Fibres
Volume fraction
Number of plies
Stacking sequences

'

Straight
50%
8 or 10
[0]
[02, , 0]s

[ek.

Glassipolyester (benchmark)'
Fibres
Volume fraction
Number of plies
Stacking sequence
Resin

Combination fabric, warp biased woven roving plus chopped


glass layer
35 percent
Seven
[0(565), 90(37)W + CSM (113) [3p]/0(565), 90(37)W [l/2p]s
Norpol 2080 (Jotun Polymer)

*Note: see Chapter 8

2.3 Mechanical characterization


The mechanical testing is performed on servohydraulic testing machines
which allow tensile loading and compressive loading, both monotonically and
cyclically.
The specimens are mounted in speciallydesigned grips; a taper of 15 degrees
in the grips and specimens allows a hydrostaticallysupported gripping, which
reduces the risk of damage/splitting in the end sections of the specimens.
Compression loading is made on these 'large' dimension specimens, so that
the same amount of material is used in mechanical characterization for both
tension and compression. These 'large' specimens are not stable against
buckling during compression loading. A special antibuckling device has been
constructed to support the specimens in compression loading, see Fig. 2.2. The
frame supports the (flat) specimens against sideways bending and the opening
in the frame allowsextensometers to be mounted on the specimen. The guiding
rods of the antibuckling device ensure perfect alignment during fatigue
loading.
The basis for this arrangement for compressive loading is the following:
the same (large) specimens for tension and compression ensure characteriz
ation based on same amount of material;
a large specimen for compressive loading (rather than a small, fat specimen
which is inherently stable against buckling) ensures characterization based
on a large amount of material, i.e., this gives better materials statistics;
the antibuckling device itself is a disadvantage, because (a) the support to the
specimen may suppress other failure modes, and (b) there is an inherent
friction between the specimen and the support frame.

Properties of Composites with Long Fibres

Ca)
Fig. 2.1(a)

fcla'i

'M't'iVrMVi'ri'iiVrrrrtVi'iVi'i'

' ' 'it^jjjLiiiaiiiitliiaii

19

(b)

Specimen for static and fatigue testing of composites with moderate strength. Dimen
sions in mm. (b) Specimen for static and fatigue testing of composites with high strength
(large fraction of fibres in the 0 degree direction). Dimensions in mm

The load application during mechanical characterization is made by hydraulic


control, both for monotonic (static) and cyclic (fatigue) loading. The load is
monitored continuously during testing.
The strain is recorded via two extensometers mounted on either side,
respectively, of the specimen; they are electrically coupled to give the net axial
strain, and thus to eliminate any possible bending strain induced in the
specimen. The (net) strain is recorded continuously during testing.
The mechanical characterization of the composite materials is made accord
ing to the plan presented in Table 2.3.

20

Design of Composite Structures Against Fatigue

Fig. 2.2

Anti-buckling device for compression loading of specimens

Table 2.3

Material

Mechanical characterization

Fraction
of (f-plies

Static tests
tension
compression

Fatigue tests
R = 0./
R = /0

(%)
G/poly 0 degrees
10 degrees
45 degrees
60 degrees
07+10 degrees
0730 degrees
0745 degrees
G/poly (benchmark)

100
0
0
0
60
60
60
84

X
X
X
X
X
X
X
X

X
X

X
X
X
X
X
X
X
X

X
X
X
X

Properties of Composites with Long Fibres

21

Table 2.4 Monotonie (static) tests


Material
(volume fraction
of fibres 50 percent)

Stiffness
E
GPa

G/poly 0 degrees
10 degrees
45 degrees
60 degrees
07 10 degrees
030 degrees
045 degrees
G/poly (benchmark)

47.2
40.9
17.4
15.4
42.0
30.4
37.2
23.8

Tension
Strength

Strain

<

MPa

862
546

2.3
1.5

55
874
634
654
445

0.49
2.24
2.4
2.6
2.13

Stiffness
E
GPa

Compression
Strength
l,

MPa

Strain
^uc

16.2

132

>2.6

40.6

720

1.9

Monotonic (static) testing


Tensile and compressive testing is made in the following cases and with the
parameters:
-

temperature
20C
humidity
'natural'
displacement rate 0.5 mm/min
strain rate
~8 x 10 -4 sec~'
stress-strain curve is recorded
stiffness, E, is calculated as initial slope of the stress-strain curve
strength at failure, o u , is calculated
strain at failure, EU, is calculated

The results of mechanical characterization are given as stiffness E, strength ,


and strain eu in Table 2.4.
The stress-strain curves for some of the composites are shown in Fig. 2.3; in
general fibre orientation has a large effect, such that 0 degree orientation gives
high values of stiffness and strength, while the off-axis orientations (45 degrees,
60 degrees) are responsible for very low values. The combination of orien
tations, such as 0 degrees and 45 degrees, gives values dominated by the
0 degree orientation.
The data under monotonic (static) loading are of the expected level, and
serve as quality-control of the glass/polyester composites. The stress-strain
curves, both in tension and in compression, are of the 'normal' type and shape.
Cyclic (fatigue) testing
Tests are made in the following cases and with the parameters:
- tension-tension
R = omm/omux = 0.1
- compression-compression R = omm/om.M = 10
- temperature
20C

22

Design of Composite Structures Against Fatigue

800

0.5

1.0

1.5

Strain (%)

2.0

2.5

3.0

(b)
Fig. 2.3(a)

Stress-strain curve for glass/polyester of fibre orientation 45 degrees, tested in


compression, (b) Stress-strain curve for fibre orientation 0 degrees/ 45 degrees,
tested in tension

humidity
'natural'
frequency
~5 Hz
load control is used during fatigue testing
number of cycles 7 is recorded, directly by counting, and indirectly by
frequency x time
- stiffness, E, is recorded
- stress-strain loops, -, are recorded at intervals during fatigue testing
(hysteresis loops)
The identification of maximum and minimum loads during testing is made at
the start of the test by loading (slowly and monotonically) to the calculated

Properties of Composites with Long Fibres


Table 2.5

23

Cyclic (fatigue) tests

Glass/polyester, comparisons
Fatigue
ratio R

Composites

0.1; 10
0.1; 10
0.1
0.1
10
0.1
0.1

G/poly

**
"
"
"
"

G/poly (benchmark)

Fibre orientations

Figure

All
0/30
0;10; 45; 60
0;0/10;0/30;0/45
0/10;0/30;0/45
0; 10; 0/10
Warp biased

2.4(a)
2.4(b)
2.4(c)
2.4(d)
2.4(e)
2.4(f)
2.4(g)

maximum load; the minimum load is calculated via the prescribed R ratio.
These loads are used as load-control parameters during fatigue testing.
The initial maximum strain emax is used in the diagrams (S-N curves). The
results of the mechanical characterization are presented in similar fatigue
diagrams. These are plots of the initial maximum strain versus the logarithm of
number of cycles; for compressive loading the numerical, maximum strain is
used. The results are presented such that comparisons are made easy, i.e.,
diagrams are plotted on the same scale, and a reference line is used; this
reference is the fatigue curve for glass/polyester composite with a fibre
orientation of 0 degrees, tested at R = 0.1 (tension-tension fatigue). Several
series and comparisons are listed in Table 2.5.
The data for cyclic (fatigue) loading are presented and compared in a series
of diagrams (Fig. 2.4). The glass/polyester laminates form a large group of data
and allow several types of comparisons.
The overview of all data in Fig. 2.4(a) shows the general shape of the S-N
curves, with a possible fatigue limit at cycles beyond IO7108 cycles. The curves
are approximately straight lines with a slope which (numerically) decreases at
increasing values for the fibre orientations. This behaviour allows for simple
analytical expressions to be established for the S-N curves, and this leads to
potential design values for fatigue up to about 107 cycles.
The comparison (Fig. 2.4(b)) of tension-fatigue (R = 0.1) and compressionfatigue (R = 10) shows, generally, that there is little difference between tension
and compression. The tendency is that at fibre orientations close to zero
composites are stronger in tension than in compression, while at fibre orientations near 90 degrees, the compressive fatigue strength is higher than the
tensile fatigue strength.
The individual comparisons for glass/polyester are made by comparisons of
'related' composites under the same loading type and are illustrated in the Figs
2.4(c)-2.4(f).
For angle ply laminates (Fig. 2.4(c)) the fatigue strength reduction is small

24

Design of Composite Structures Against Fatigue

+ 0 degrees =0.1
10 degrees
o 45 degrees
x 60 degrees
0/45
0/45
R = 10

_L

Ulli

10

10

R = 0.1
R = 10

IV

IV
Cycles

IV

IV

10'

10

(b)
Fig. 2.4(a)

Fatigue diagrams of maximum nominal strain versus logarithm of number of cycles to


failure, for glass/polyester composites all composites al K = 0.1 and R = 10. (b)
Fatigue diagrams of maximum nominal strain versus logarithm of number of cycles to
failure, for glass/polyester composites0 degree/30 degree composites at R = 0.1 and
R = 10

Properties of Composites with Long Fibres

25

3.0

0 degrees
O 10 degrees
V 45 degrees
x 60 degrees

= 0.1

2.0

D 0 degrees 10 degrees
+ 0 degrees 30 degrees R = 0.1
0 degrees 45 degrees
0 degrees

0.5

(d)
Fig. 2.4(c)

10

10!

IV

IV
Cycles

IV

IV

10'

10

Fatigue diagrams of maximum nominal strain versus logarithm of number of cycles to


failure, for glass/polyester composites 0 degree, 10 degree, 45 degree, and 60
degree composites, at R = 0.1. (d) Fatigue diagrams of maximum nominal strain versus
logarithm of number of cycles to failure, for glass/polyester composites 0 degree,
0degree/10 degree, 0 degree/30 degree, and 0 degree/45 degree composites at
= 0.1

26

Design of Composite Structures Against Fatigue


2.5
0 degrees 10 degrees
ffl 0 degrees 30 degrees
A 0 degrees 45 degrees

degrees R = 0.1

oJ-

10

10'

10'

IV
Cycles

10s

H = 10

10'

10'

10'

(e)

o 0 10 degrees
O 10 degrees

= < u

0 degrees

os

10

10'

IO3

IO4
Cycles

10'

10'

10'

10

(f)
Fig. 2.4(e)

Fatigue diagrams of maximum nominal strain versus logarithm of number of cycles to


failure, for glass/polyester composites 0 degree, 0 degree/10 degree, 0 degree/30
degree, and 0 degree/45 degree composites at R = 10. (f) Fatigue diagrams of
maximum nominal strain versus logarithm of number of cycles to failure, for glass/
polyester composites 0 degree, 1 0 degree, and 0 degree/10 degree composites at
= 0.1

Properties of Composites with Long Fibres


2.5

-i

27

IV

IV

10'

0 degrees

0.5

Fig. 2.4(g)

' ' II ml

10

IV

IV

IV
Cycles

10

Fatigue diagrams of maximum nominal strain versus logarithm of number of cycles to


failure, for glass/polyester composites combination fabric (benchmark) at R = 0.1

for angles of about 10 degrees, while it is large for 45 degree and 60 degree fibre
orientations.
For combination laminates with a significant fraction of 0 degree oriented
fibres the fatigue strength for R = 0.1 is practically unaffected by the angje
oriented fibres, as seen in Fig. 2.4(d) for R = 0.1 (tensionfatigue) while a
clear reduction in fatigue strength is recorded in Fig. 2.4(e) for R = 10
(compressionfatigue), when the angle orientation is increased.
The comparison of laminates with 0 degreefibresand anglefibresfor R = 0.1
(tensionfatigue) is seen in Fig. 2.4(f) for angles of 10 degrees. The presence of
0 degree fibres ensures fatigue strength values close to those of 'pure' 0 degree
laminates.
The results for the glass/polyester (benchmark) at R 0.1 (Fig. 2.4(g)) are
displaced below the reference line, which is probably caused by the imperfect
manufacturing of this practical material, made under industrial conditions.
Stiffness reduction
In most tests the stiffness changes are recorded regularly during fatigue testing,
typically at every tenth of a decade of cycles. The results are displayed in
diagrams showing stiffness reduction, E/E0, versus normalized lifetime (num
ber of cycles), both with a linear parameter (Fig. 2.5(a)) and logarithmic
parameter (Fig. 2.5(b)). This figure shows the same data plotted in two
different ways. The realtime plot (Fig. 2.5(a)) also shows that the steep
reduction in modulus starts at about 85 percent of the (normalized) lifetime,
while the logarithmic plot (Fig. 2.5(b)) obscures this situation, although such
plots are often used for presentation of fatigue data.
The reduction in stiffness during fatigue loading is an indirect measure of the

28

Design of Composite Structures Against Fatigue

1.0

~i

0.9

1 0.7
o

0.6

0.5

(a)

(b)
Fig. 2.5(a)

0.1

0.2

0.3

0.1

0.2

0.4
0.5
0.6
0.7
Normalized lifetime (N/N()

0.3
0.4
0.5
0.6
0.7
Normalized lifetime (log AVlog Nt)

0.8

0.8

0.9

0.9

1.0

1.0

Stiffness reduction diagram, with normalized stiffness ElEB versus normalized lifetime
;V/,Vf for 0 degree/30 degree composites, fatigue tested at R = 10 with a maximum
strain of 0.8 percent, (b) Stiffness reduction diagram, with normalized stiffness /-."//',,
versus normalized logarithmic lifetime log '/log /Vr for 0 degree/30 degree compo
sites, fatigue tested at R = 10 with a maximum strain of 0.8 percent

damage (cracks, delaminations) produced in the material. The selected results


show that a slow reduction in stiffness occurs during most of the life-time, and
that a final, rather fast, reduction takes place before failure.
Hysteresis loop change
In selected tests, the hysteresis loops (- loops) are recorded at intervals
during fatigue testing. An example of the results is shown in Fig. 2.6, typically

Properties of Composites with Long Fibres

0.75
1.0
Strain (%)
Fig. 2.6

29

1.75

Hysteresis loop diagram, for composite under fatigue testing of maximum initial strain of
1.0 percent. For cycle 1, the modulus is 31.2 GPa; after 74,935 cycles, 24.9 GPa and after
290,925 cycles, 18.0 GPa

as the fir st hyster esis loop at the star t of fatigue testing and the loop after , a
specific number of cycles.
It is noted that the following phenomena occur as the number of cycles
increases:
(a) the loop area incr eases;
(b) the 'loop slope' (equivalent to the stiffness) decr eases;
(c) the minimum str ain and the maximum str ain incr ease.
Similar r esults ar e found with other composites (see Chapter 9). The last
phenomenon is a shift on the loop in the - diagr am, this is a result of the fact
that the fatigue testing is under load contr ol and that the mater ial stiffness is
reduced. For the numer ical values of str ess and str ain the following r elations
hold
R -

omJomaii

. = rl "min

'P min
'max

' "ma x

min

OmmIL =

max

omaxlb.

(PmmIA)IE
(rmax/A)IL

The last two relations show that both min and max increase when E is reduced,
and that the increase is largest for e max (because P m;ix > Pmin).

30

Design of Composite Structures Against Fatigue

Initial

^
^ ^ After N cycles

ss
Strain

Fig. 2.7

Schematic hysteresis loops at (/:" and V = 1 ) and at cycle ' (E and N); E is the elastic
modulus (approximately equal to 'slope' of loop), ' is cycle number; the R ratio is 0.1

The 'loopslope' is expected to extrapolate to , = 0,0. (Fig. 2.7): this is not


the case for several of the hysteresis loop results. This could be caused by
experimental error at the start of (fatigue) testing, by an initial state of internal
stress in the composites, and by changes in these internal stresses during fatigue
loading.
2.4 Significance for design
These results set upper limits for those data which can be obtained under
commercial manufacture, asfilamentwinding produces highly orientated, high
quality laminates. The best way of comparing fatigue data is nominal strain
versus the logarithm of cycles.
Orientation
The effect offibreorientation is clear the higher the angle to the load axis, the
lower the nominal strain for any given number of cycles.
Having somefibreslying along the load axis and somefibresat an angle is the
best design practice. The tensile fatigue data presented on the basis of strain
show the same fatigue curves for all laminates with a substantial proportion of
0 degree fibres (typically 50100 percent). For the designer the loadcarrying
capacity is of importance, and therefore the stiffness is also significant ; this calls
for a moderate proportion of offaxisfibresto avoid a low modulus and thus low
fatigue strength. Ideally, the offaxis fibres should be orientated to resist the
offaxis loads.

Properties of Composites with Long Fibres

31

Loading
Composites are better able to withstand tensile than compressive loading, even
if the laminate contains fibres lying along the loading axis. In general,
composites in service will be subjected to a mixture of tensile and compressive
loading. The fatigue properties of reverse loading (tension/compression)
would thus be less than that of pure tensile loading, as can be seen in Fig. 8.5.
The effect of spectral loading, which is a combination of compressive and
tensile loading, is discussed in Chapter 4.
Damage accumulation
Stiffness is an important property for designers, because it governs the
(maximum) deflection during loading. The reduction in stiffness is thus an
important consideration for long-time design. The curves of EIE0 versus log Nf
(Fig. 2.5) can lead to design-allowables for cases where a maximum reduction
in stiffness can be accepted, and the curves can be used to establish the
corresponding number of cycles. Alternatively, the point of the curve where
the fast reduction starts may be used as a design-allowable, if the actual
stiffness at that point can be accepted.
The hysteresis loops can also be correlated to temperature changes (increases) during fatigue loading, and such recordings can serve as damagemonitoring techniques. These may be developed into practical methods to be
used in service.
Both of these damage parameters are discussed further in subsequent
chapters (for example, Chapters 7 and 9).
2.5 Conclusions
The fatigue of well-orientated fibre composites shows a characteristic decrease
in properties as a function of time. Somefibresin the primary loading direction
are beneficial in terms of fatigue properties. Such composites are better able to
withstand tensile than compressive fatigue.

CHAPTER 3

Influence of Matrix and Fabric


A. T. Echtermeyer*, B. Engh, and L. Buene*
Fatigue properties of eleven different laminates are evaluated for various combi
nations of glass fabrics and polymeric matrices. Damage build up is monitored
continuously by recording stressstrain curves during testing.
The effect of the various resins on fatigue properties is found to be small in a
laboratory environment. Damage is found to develop even when specimens are
cycled below the static linear limit of stress with strain. A standard strain/cycle fatigue
lifetime curve can be developed for all laminates provided that some continuous
fibres are present in the load direction.
3.1

Introduction

This chapter investigates the properties of fibre reinforced plastics (FRP)


which are typically used today for wind turbine blades, marine vessels, and
other engineering structures. These composites have glass reinforcements like
woven rovings and chopped strand mat. Typical resins are polyesters and
vinylesters. More advanced reinforcements, like stitchbonded fabrics and
multiaxial fabrics having continuous straight parallel fibres are investigated
too.
For all materials damage development during fatigue was monitored. The
loss of stiffness (reduction of Young's modulus) was taken as a measure of the
amount of damage in the specimen. In particular matrix cracking has been
identified to cause stiffness reduction in static and fatigue tests (l)(4). The
relative effect of matrix and fibres on fatigue properties can be better under
stood by taking development characteristics into account.
3.2 Materials and testing
Eleven different laminates were tested, the layup and resin of which are listed
in Table 3.1. In addition the fabric types and fibre fractions are indicated as
both can influence the properties of the laminate. All laminates were manufac
tured by hand layup and postcured at 363 (90C) for 120 minutes.
The influence of resin type was evaluated for five laminates having the same
widelyused combination mat as reinforcement. These comprised a balanced
woven roving of 800 g/m2 and 100 g/m2 of chopped strand mat (CSM) stitch
*DNV Research AS, Norway.
tDNV Classification AS, Norway.

34

Design of Composite Structures Against Fatigue


Table 3.1

Description of laminates
1

Laminate
Code
Combi iso
Combi NPGiso
Combi vinyl
Combi rub-vinyl
Combi Ortho
CSM
SB-Combi
MA-O
MA-90
MA-0/90
MA-45

Fibre
weight
fraction

Fraction of
fibres in
load
direction

Layup*

(%)

(%)

[[0(400),90(400)WR],((100CSM)]SB].,
[[0(4O0),90(400)WR],[(100CSM)]SB]5
[[0(400),90(400)WR],[(100CSM)]SB].,
[[0(400),90(400)WR],[(100CSM)]SB],
[[0(400),90(400)WR],[(100CSM))SB]5
[(100CSM)],,
[0(400),90(400),(100CSM)]SB]5
[0(578),90(14)I].,S
[90(578),0(14)I],5
[0(400).90(40)I].
[+45(400),-45(400)K]., s

53
51
53
52
53
31
64
54
54
53
53

45
45
45
45
45
Random
45
98

Resin type
lso-polyester
NPG/iso-polyester
Vinyl ester
Rub. mod. vinyl ester
Ortho-polyester
NPG/iso-polyester
NPG/iso-polyester
NPG/iso-polyester
NPG/iso-polyester
NPG/iso-polyester
NPG/iso-polyester

50
(1

'Notes: For layup conventions refer to Table 0.1


SB fabric Rovimat 800/100 (Chomarat)
Multiaxial knitted fabrics (Devoid AMT)

Table 3.2

Resin properties from Jotun Polymer (5)

Ortho
polyester
Type
Tensile Modulus (GPa)
Tensile Strength (MPa)
Elongation at break (%)
Flexural modulus (GPa)
Flexural strength (MPa)

NP41-90
3.6
65
3.7
3.3
125

so

polyester

NP72-90
3.7
76
3.5
3.6
135

NPG/iso
polyester
NP20-80
3.6
73
6.5

124

Vinyl ester

Rubber
modified
vinyl ester

NP92-20
3.3
80
5
3.1
130

NP92-40
2.9
68
9
2.8
125

All resins were supplied by Jotun Polymer

bonded together. Five different resins were used - three polyester based and
two vinyl ester based; their properties are listed in Table 3.2 (5). The resins
differ from the greatest extent in their elongation to break (3-9 percent).
The other six laminates were made to study the influence of different fabric
layups with the same resin (NPG-iso polyester). These comprised a stitch
bonded combination mat (SB) and the chopped strand mat by itself and four
laminates of inlaid construction having straight glass fibres held together by
polyester yarns.
Laminates MA-0 have 98 percent of their fibres running parallel to and 2
percent normal to the load direction. MA-90 is the same laminate as MA-0,
but tested perpendicular to the main fibre direction. MA-0/90 is a cross plied
multi-axial laminate with the same amount of fibres running parallel and
normal to the load direction. Laminate MA-45 is laminate MA-0/90 rotated

Influence of Matrix and Fabric


Grips

Qtl

35

Anti-buckling
supports
Specimen

Specimen

V\

Support jig

Extensometer

Fig. 3.1

Grips

Experimental set-up for coupon fatigue tests

by 45 degrees. The MA laminates were also used to obtain basic ply properties
for laminate theory calculations.
All tests were performed on a servo-hydraulic MTS testing machine.
Specimens were cut as straight 25 x 5 x 270 mm coupons from the laminate,
according to ISO 3268. In order to prevent buckling in compression an
antibuckling device similar to the support jig of ASTM 695 was used. Dimensions of the device had to be slightly changed to a length of 140 mm and width of
25 mm to fit the specimens; there is no standard procedure for GRP tensioncompression fatigue tests (Fig. 3.1). The gap between the antibuckling device
and the grips was kept as small as possible (less than 2 mm). This set-up worked
well for most specimens, but thin specimens with highly oriented fibres had the
tendency to fail in the gap between the antibuckling device and grips. In those
cases an extra set of metal supports was inserted into the test fixture, as shown
in Fig. 3.1. These supports closed the gap near the grips. The slit in the middle
of the supports, with its orientation at 45 degrees to the specimen, did not seem
to cause any buckling of the specimen in compression.
Strains were measured with an MTS Model 632.12C-20 extensometer,
attached to the edge of the specimen. Stress-strain curves could be taken
continuously by a computerized data acquisition system. Testing rates of quasi
static tests were 2 mm/min for tensile and compressive tests. Frequencies of
fatigue tests were varied between 2 and 5 Hz to keep the average load rate
about constant, avoiding viscoelastic effects. Low frequencies were chosen to
prevent the specimens from heating up. All fatigue testing was done in the load
control mode.

36

Design of Composite Structures Against Fatigue


Table 3.3

Laminate code
Combi iso
Combi NPG/iso
Combi vinyl
Combi rub-vinyl
Combi ortho
CSM
SB-Combi
MA-0
MA-90
MA-0/90
MA-+45

Linear
Young's
modulus
(GPa)
17.1
16/1
16.8
16.9
17.3
8.4
16.0
28.3
10.6
20.4
10.4

0.4
0.4
0.5
0.7
0.7
0.4
0.7
0.7
0.4
1.0
0.3

Static properties of laminates


Strain at
linear
limit

Ultimate
tensile
strength
(MPa)

0.28
0.34
0.38
0.33
0.30

209 2 0
205 6
227 5
237 14
226 7
125 1
250 16
602 8
71 2
366 2 5
605

(%)

0.32

0.13
0.13

Strain at
tensile
failure

(%)

Ultimate
compr.
strength
(MPa)

1.40 0.20
209 9
1.62 0.12
258 14
1.72 0.03
288 25
259 14
1.68 0.07
1.77 0.10
294 23
1.61 0.03 211 6
2.25 0.06
248 10
2.34 0.30 >450
2.35 0.20
83 2
2.50 0.30
285 30
>12
908

*no linear limit

3.3 Static properties


Static properties of all laminates were measured in tension and compression.
Tensile static curves can be characterized by an initially linear slope, the linear
limit, and ultimate failure properties (Fig. 3.2). These parameters are listed for
all laminates in Table 3.3. Tensile elastic moduli , were taken as the slope of
the linear part of the stress-strain curve. The linear limit was defined as the
stress-strain level when data points drop significantly below the line defining
the linear modulus Et. This is taken as a 10 percent drop of the secant modulus
from the Et value. Note that the MA-0 laminate with 98 percent of its fibres
running in the load direction does not show a linear limit based on this
definition. Typical tensile stress-strain curves are shown in Fig. 3.3 for NPG/iso
laminates.
Different resins with the same combimat reinforcement showed very similar
static characteristics. Ultimate strength and strain have similar values within
the experimental scatter (Table 3.3), therefore, only the stress-strain curve of
the combi NPG/iso laminate is shown in Fig. 3.3. The main difference between
the five combi laminates is the onset of non-linearity. Although the exact
position of the onset is difficult to determine, a trend between the different
resins can be seen. The iso- and ortho-resins have a lower linear limit than the
vinylesters and the NPG/iso polyester. This reflects the relativeflexibilityof
the various resins (Table 3.2). Compressive tests show a fairly linear stressstrain curve up to the point of failure. Compressive properties are also listed in
Table 3.3.
Different reinforcements with the same resin have a large influence on the
static stress-strain behaviour (Fig. 3.3). As expected, modulus and strength
increase with the amount of fibres in the load direction (Table 3.1). However,
the combimat and the CSM fail already at a strain of about 1.5 percent. All

Influence of Matrix and Fabric

37

Failure

Fig. 3.2

Strain
Schematic of a stress-strain curve of a quasi static tensile test

other laminates fail at strains of more than 2.1 percent. It seems that stitchbonding and knitting allow better utilization of the fibres. Note that 90 degree
laminates (with nofibresin the load direction) generally fail at very low strains,
but the MA-90 laminate failed at a strain of 2.3 percent. The 2 percent of the
fibres running in the load direction could bridge matrix cracks and keep this
MA-90 laminate together until the few fibres in the load direction failed.
The MA45 laminates showed very non-linear behaviour and failed at
strains of more than 12 percent (6). It seems that the yarns running through the
thickness of the laminates (used in the knitting process) help to keep the plies
together for a long time. 45 epoxy-prepreg laminates fail at much lower
strains of about 3 to 4 percent (7).
Ply properties
The static data are important for design and quality control. Advanced
structural calculations, based on laminate theory (e.g. (8)(9)) require the full
set of orthotropic ply properties to characterize a laminate. To use laminate
theory calculations properly, plies should represent layers of fibres running

38

Design of Composite Structures Against Fatigue

600
500

400

300

SB-combi

Vi

200

1.0
Fig. 3.3

1.5
Strain (%)

2.0

2.5

3.0

Typical static tensile stress-strain curves for laminates with NPG/iso-polyester resin

continually and parallel in one direction (100 percent unidirectional plies).


These properties cannot be derived from complicated laminates, but the simple
multiaxial laminates tested here (MA-0, MA-90, and MA-45) can be used to
deduce such ply properties (10)(11). Ply properties for the multiaxial laminates
are listed in Table 3.4. Note that the axial properties (parallel and normal to
fibres) of a 100 percent unidirectional ply are slightly different from the
properties of the laminates MA-0 and MA-90 (Table 3.3). The ply properties
can be used to predict properties of more complicated laminates. Predictions
for the MA-0/90 laminate were found to be very good. Predictions for (0/90/45)
laminates were good in the linear range. Ultimate properties can be predicted
with progressive failure analysis, 'last ply failure' calculations determine the
first fibre failure (11)(12). In all these cases the ply properties have to be
adjusted to the fibre volume fractions of the laminates, using micromechanic
correction formulae (12).
3.4 Fatigue properties
Fatigue tests of all specimens were performed under reverse loading (7? = 1),
because it was found to cause the most severe fatigue conditions in other

Influence of Matrix and Fabric


Table 3.4

39

Ply properties of glass polyester plies, obtained from multiaxial laminates


with NPG/isophthalate resin

Property

Symbol

Value

Units

M
t

592
0.63
1.1
2.54
0.39
0.60
1.87

g/m2
mm
kg/m3
kg/m1

29.2
9.6
3.3
0.29

GPa
GPa
GPa

Fibre Weight
Thickness/ply
Density of matrix
Density of fibre
Fibre Volume Fract.
Fibre Weight Fract.
Ply density

Pm
Pi
Vf
Vw

ft

Young's Mod in fibre direction


Young's Mod normal to fibre direction
Inplane shear modulus
Major Poisson's ratio

Et
2

Gn
>;12

X
X'
Y
Y'
XY

Tensile Strength in fibre direction


Compr. Strength in fibre direction
Tensile Strength normal to fibre
Compr. Strength normal to fibre
In plane shear strength

300

V,

S 200

R = l
MAO/90
combi y"^

SB combi

150

^^,

\
'.

MA 90

Fig. 3 . 4

^ Df c U A * .
MA 45
0==.:

10'

^ ."

1
2

^^

100

50

^S^

..CSM

MPa
MPa
MPa
MPa
MPa

0.

MAO

250

kg/m3

678.9
nm
13.3
61.2
19

"^

".

1
,
IV
IV
Cycles to failure

^U
'

~-.^^

10 s

10'

10'

.S'-' curves for laminates with NPG/iso-polyester resin and various reinforcement fabrics

composites (13). Figure 3.4 shows the S-N curves for all the different laminates
made with NPG/iso resin. The differences in the curves are large, due to the
very different amounts of fibres in the load direction. Similar to the static
results, laminates with the highest amount of fibres in the load direction

40

Design of Composite Structures Against Fatigue

showed the highest fatigue strength. The S-N curves for the combimat
laminates with different resins were very similar to the one for NPG/iso (14).
Such curves do not reveal any information about the damage processes taking
place in the specimens. However, damage development studies can help to
assess the change of properties during the fatigue life, and they can lead to a
more simplified analysis of ultimate fatigue failure.
Damage development
Stresses and strains were monitored continuously for all specimens during
fatigue testing. A change in the stress-strain curve results from some damage
development. A reduction of the slope of the curve corresponds to a drop in
Young's modulus. Initially, such a drop corresponds to the onset and increase
of matrix cracking in the laminate. Later, other damage forms, like delamination and fibre failure, can contribute to a modulus drop too (14) (15).
Reverse loading fatigue shows more special characteristics in the change of
the stress-strain curve with cycle number. The slope of the tensile part of the
curve drops to lower values due to an increase in damage (matrix cracking).
But the compressive slope of the curve remains at the same value through most
of the fatigue life (Fig. 3.5). Compression seems to close the cracks in the
material. This means that the damage introduced to the specimen during the
tensile part of the fatigue cycle does not influence the compressive modulus of
the material. Once these cracks are closed under compression, the material
appears undamaged to a stress-strain measurement. Only shortly before
failure of the specimen do both the tensile and compressive modulus drop
rapidly. It is very likely that at this point severe damage, including fibre failure,
develops in the specimen, which causes rapid degradation of the properties.
Stress-strain characteristics during fatigue, including hysteresis effects, are
described in more detail in (14) (15).
Change of the tensile fatigue modulus is a good way to characterize damage
development during fatigue. Figure 3.6 shows the change of the modulus with
cycle number for different fatigue strains for the combi-NPG/iso laminate.
Cycling below the linear limit
Specimens have the same initial tensile fatigue modulus Etfi as the primary
static modulus Et. The tensile fatigue modulus El{ remains constant for a
number of cycles before it starts to drop. This indicates that the specimen is
initially free of damage, i.e., no matrix cracking. Cracking sets in after an
initiation period, when the modulus starts to drop. The initiation period
increases if a specimen is cycled at very low amplitudes, i.e., far below the
linear limit.
Figure 3.6 shows two specimens which were cycled at strains below the linear
limit. The modulus Et{ of the specimen cycled at 0.32 percent strain drops after
about 1500 cycles, while the modulus of the specimen cycled at 0.25 percent

Strain

Fig. 3.5

Schematic of stress-strain curves of a fatigue test

strain drops after approximately 3000 cycles. This onset of matrix cracking
happens at much lower numbers of cycles than ultimate failure (Fig. 3.7).
Cycling above the linear limit
Specimens develop matrix cracking already when loaded in thefirstcycle. The
initial modulus Et depends on the fatigue strain (Fig. 3.6). High strain levels
cause much initial damage, giving low initial fatigue moduli tfi. The moduli
drop immediately with increasing cycle number, reflecting an instant growth of
damage.
Critical fatigue modulus Ecr
Once matrix cracking occurs, the crack density increases, causing the tensile
modulus to drop. The modulus Elf drops slowly and fairly linearly with the
logarithm of the cycle number up to a critical value crit (Fig. 3.6). This

Design of Composite Structures Against Fatigue

42
1.1

No
damage

l.o
.25
s

I 0.9
s
3

Slight

^,0.32
'S

-^^59
0.8

damage
'

->

1.61"^'^

I 0.7

0.6

0.5

Fig. 3.6

to

io 3

:
io 4
Cycles to failure

Severe
damage

0.59 :

10'

10.96

105

10'

10'

Change of tensile fatigue modulus Er with cycle number for various initial strains (%) of
the combi mat NPG/iso-polyester laminate (loaded in weft direction). The upper two
curves show data for specimens cycled below the linear limit

modulus appears to be independent of the strain amplitude. Since a linear


relationship between modulus drop and matrix crack density was found in
other composites (3), matrix crack density seems to accumulate linearly with
the logarithm of the cycle number.
Slightly different behaviour was observed for specimens cycled below the
linear limit (0.35 percent). After the linear drop of stiffness Etf with cycle
number the reduction slowed down and eventually halted, indicating a satu
ration of the damage. The modulus Ef at saturation was slightly above the
critical modulus crit . Damage saturation was also found in crossplied (0/903)s
GRP (3) (16). This means that all matrix cracks had been formed, but the
laminate was able to sustain the damage.
Ecrit reflects the stiffness of the laminate at a state of damage, which is mostly
due to matrix cracking. At this level, thefibresseem to be still undamaged and
so able to sustain the applied load.
Failure
Once the modulus drops below Ecrit all properties of the laminate degrade very
rapidly. A rapid drop in tensile stiffness coincides with a rapid drop of
compressive stiffness Ec{, which stays constant at the previous static value Ec
(14). A different damage mechanism, probably damage of the load bearing
fibres, gets activated and rapidly causes ultimate failure. A further indication of

Influence of Matrix and Fabric

2.0

43

iso / combi
NPG iso
vinyl
rub-vinyl
ortho
NPG iso / SB combi

1.8
1.6

1-4
95/95
limit

h 1.2
(A

I 1.0

I 0.8

Some
damage

a
0.6

0.4

O.Q.

No

0.2

1 ml

10

IO2

S.

N.

IO3

IO"

10s

IO6

IO7

Cycles
Fig. 3.7

/V curves for combi laminates with various resins and one SB-combi reinforcement. The
stages of damage in the Combi-NPG/iso-poly ester laminates are indicated. Initial strain is
plotted because tests were performed under load control, which means strains increase
slightly during fatigue

severe damage formation is the sudden appearance of and increase in the


hysteresis effect in the fatigue stress-strain curve (between loading and
unloading curve) when the fatigue modulus drops below crjt (14) (15).
3.5 Effect of matrix
The influence of the matrix material on the fatigue properties of laminates
appears to be small. Combi-laminates show exactly th same characteristics
regarding modulus drop, as described above and shown in Fig. 3.6. Even
laminates with a phenolic matrix have a similar characteristic, except that
phenolic composites show a reduction of compressive modulus even at low
numbers of cycles (15).
Critical fatigue moduli cri , for all laminates are listed in Table 3.5. For
combi-laminates the modulus tf drops by 21-25 percent during the slow drop
period, until crit is reached. The fatigue modulus of SB-combi drops by 35

44

Design of Composite Structures Against Fatigue


Table 3.5

Laminate
code
Combi iso
Combi NP Giso
Combi vinyl
Combi rub-vinyl
Combi ortho
CSM
SB-Combi
MA-0
MA-90
MA-45

Critical fatigue moduli of laminates

Linear
Young's
Modulus
E,
(GPa)

Critical
fatigue
modulus

Tensile
modulus
drop
during Fat.

(GPa)

(%)

17.1
16.1
16.8
16.9
17.3
8.4
16.0
28.3
10.6
10.4

13.4
12.4
12.8
12.7
13.3
5.9
10.4
27.0
3.0

21.6
23.0 '
23.8
24.9
23.1
29.8
35.0
4.6
71.7

Secant modulus
at static
tensile fail.
ECJ
F

t
"Stt
L-sec
(GPa)
(GPa)
14.9
12.7
13.2
14.1
12.8
7.8
11.1
25.7
3.0
-

0.90
0.98
0.97
0.90
1.04
0.76
0.94
1.05
0.99

percent. These moduli drops are quite large and may constitute failure if
stiffness or resonance frequency is a critical design parameter.
Fatigue life for a maximum initial strain is shown in Fig. 3.7 for all laminates
with combi-mat reinforcement. These -ZV curves have the same characteristics
as S-N curves for these laminates. Laminates show some scatter in the data at
high fatigue strains, while data are all close together at low fatigue strains.
Fatigue far above the linear limit (short life times) (Fig. 3.2) seems to be slightly
influenced by the matrix properties (14). However, if the material is cycled at
fatigue amplitudes close to or below the linear limit, all laminates have about
the same lifetime. Matrix properties are not important anymore, because the
matrix is in a state of damage saturation. Fibre properties control fatigue
response at low strain amplitudes. Therefore, failure at low and high cycle
fatigue appears to be governed by different failure mechanisms.
All data were obtained under laboratory conditions. Results may be some
what different if specimens are tested under severe environmental conditions.
Specimens which show more matrix cracking early in their fatigue life may
provide an easier path for water or other media to reach the fibres. Since fibre
strengths can be reduced by the presence of water etc. (17), such laminates may
fail earlier.
3.6 Effect of fabrics
Fabrics have the largest influence on the fatigue properties of laminates.
However, even though the fatigue data are different, the basic mechanisms
governing damage accumulation do not change, as long as somefibreslie in the
load direction.
The change of modulus with cycle numbers follows the same principles as
described above, but the critical fatigue moduli are very dependent on the
fabric type (Table 3.5). Figure 3.8 shows typical fatigue modulus versus cycle

Influence of Matrix and Fabric

45

1.1

1.0

0.9

MAO

a
oE 0.8
U

25
3
3

\ i.

0.7

o
O

\
MA 9oV

ti 0.6
<U
Bl

MA 45

'
SB combi

0.5

l combi/NPG iso.
\

'1.J

0.4

I I I I Mili

0.3

IO2

ili'
Fig. 3. 8

1IO

10"
Cycles to failure

10s

10'

10'

Typical changes of tensile fatigue modulus Ef with cycle number for laminates with
different glass reinforcements. All curves are taken from specimens with a lifetime of
approximately 10 6 cycles

number curves for different reinforcements. The multiaxial laminate MA0


(98 percent of the fibres running parallel to the load S) has a fairly constant
fatigue modulus until rapid degradation sets in. Matrix cracking occurs in this
laminate in the same way as in the previouslydescribed laminates, but it does
not influence the stiffness of a laminate with such a high fraction of the fibres in
the load direction. The modulus in the fibre direction u d of a nearly
unidirectional laminate can be calculated from the moduli of the fibres fibrc
and the matrix matrix
ud = l^ffibre
o

(1 ~ Vt) ^matrix

'

where V , is the fibre volume fraction in the load direction. (V? is the fibre
volume fraction of the laminate multiplied by the fraction of fibres in load
direction). Matrix cracking causes a reduction of matrix> but this does not
influence u d to any large extent, if V? is large and Efibre ^matrix Only fibre
degradation has a measurable effect on u d . This explains why the modulus of
the MA0 laminate does not change with increasing cycle numbers. Other
composites dominated by 0 degree fibres do not lose much stiffness during
fatigue either. For example, UD laminates made from carbon (18) and aramid
(19) fibres show no drop in tensile stiffness before ultimate failure. On the

46

Design of Composite Structures Against Fatigue

other hand, a large drop in modulus is observed for the MA-90 laminate. Here
V? is very small and changes in the modulus of the matrix due to cracking have a
large influence on ud . The other laminates show characteristics which are
between these two extremes. The only exceptions are the 45 degree laminates. These laminates have no fibres running in the load direction. The
modulus drop curves do not show a critical fatigue modulus and they do not
show a linear drop with log cycle number (Fig. 3.8).
3.7 Discussion
A comparison between static tensile moduli critical fatigue moduli cri
and static secant moduli see at failure is given in Table 3.5. Critical fatigue
moduli are very close to the static secant modulus at failure. At the point of
failure in a static test the material develops much matrix cracking, which
reduces the static modulus before thefibresfail. A similar situation happens in
a fatigue test. Matrix cracking increases with cycle numbers (up to a saturation
level if the fatigue amplitude is low), before the fibres start to fail. Therefore,
crit and sec are very similar. So, a simple static tensile test can give a good
estimate of how much the modulus will drop during the fatigue life of a fibre
reinforced laminate.
Fibre-dominated properties can be well characterized by laminate strains.
Strains are approximately the same throughout the laminate (an assumption of
laminate theory if no bending is present), while stresses change from ply to ply
and between the matrix and fibres. At a given strain C|0ad in load direction, all
fibres in the load direction are under a similar stress CTfibre
"fibre

^fibre^load

where fibre is the Young's modulus of the fibre. The average stress of the
laminate can, on the other hand, be very different, depending on the stiffness
matrix of the total laminate. This matrix is influenced by the fibre orientations
in the laminate and also by the stage of damage.
Characterizing fatigue of composites by a strain-life diagram was suggested
by Boiler (20) and Talreja (21). This method provides a conceptual framework
for interpretation of fatigue results, based on matrix and fibre properties. The
method also allows comparison of fatigue data of different laminates directly
without the need to normalise data by the tensile strength.
Figure 3.9 shows the cycle numbers to failure versus maximum fatigue strain
for all laminates made with NPG/iso resin. The fatigue life curve is very similar
for all the different laminates, except for the MA45 laminate. As long as
some fibres run in the load direction, fatigue life is fibre dominated. Only the
chopped strand mat data points are consistently slightly below the group of the
other data points. This is probably due to the non-continuous fibres in CSM,
which cannot always bridge the damage areas as well as the continuous fibres.
All specimens have about the same lifetime at a given maximum fatigue strain.

Influence of Matrix and Fabric

IO5

47

IV
Cycles

Fig. 3.9

/V curves for laminates with different glass reinforcements but same resin (NPG/isopolyester). Initial strain plotted because tests were performed under load control, which
means strains increase slightly during fatigue. MA 45 degree data is shown for
comparison and is not used in deriving means and limits

This is even true for laminates which show a rather low strain to failure in static
tests. CSM and combi-laminates failed at 1.6 percent static strain (Table 3.3),
while all other laminates failed at more than 2.1 percent strain.
The 45 laminates have very different fatigue characteristics. Their fatigue
life is dominated by the ability of the matrix to keep the laminate together.
When the matrix cannot take the shear and normal stresses between the fibres,
the laminate will fail. Note that these results were obtained from narrow
specimens where fibres do not run continuously from one load introduction
point to the other. If 45 laminates are used in larger structures fibres could
connect load introduction points. Such structures should be less sensitive to
matrix cracking and the fatigue life may be longer than indicated by the small
specimens tested here.
Failure and design
Failure of FRP due to fatigue can be best described by the number of cycles to
failure for a maximum strain, i.e., -7 curves, as long as some fibres of the

48

Design of Composite Structures Against Fatigue


Table 3.6

Parameters to describe nV curves of laminates

Laminate
code

log(e) = log{e(l)} -alog(N)


Standard
log{e(l)}
deviation

Combi iso
Combi NPG/iso
Combi vinyl
Combi rub-vinyl
Combi Ortho
CSM
SB-Combi
MA-0
MA-90
MA-0/90
All laminates

0.28
0.44
0.44
0.44
0.47
0.34
0.51
0.41
0.55
0.50
0.45

0.02
0.02
0.03
0.02
0.44
0.03
0.04
0.01
0.03
0.04
0.06

0.10
0.12
0.13
0.12
0.14
0.11
0.13
0.11
0.15
0.12
0.13

laminate are oriented in the load direction. Data are obtained by a linear fit
equation and a logarithmic equation of the form
() = (1) a\og(N)
linear fit equation
log{(/V)} = log{e(l)} - log (N) logarithmic fit equation
where is the number of cycles to failure and is the maximum fatigue strain
(Figs 3.7, 3.9).
Both linear fit and logarithmic fit curves are shown in Figs 3.7 and 3.9. In
both cases, the 45 laminate was not included when obtaining the failure
curve, because it has no fibres running in the load direction. The lower 95
percent tolerance bound with 95 percent confidence, constructed as described
in Chapter 14, is also shown in these figures. Parameters for the best fit of all
materials are given in Table 3.6.
The linear fit seems best suited to describe the data for high strain-stress
amplitudes (short lifetimes below 103 cycles), while the logarithmicfitseems to
describe the low strain-stress amplitude (long lifetimes, more than 103 cycles)
better. A reason for the difference may be the influence of the matrix on the
high amplitude fatigue, while fibre properties dominate low amplitude fatigue.
However, both methods to fit the data are empirical and have no relation to
failure mechanisms. Since lifetimes of more than 103 cycles are usually of
interest, the logarithmic description should be used to predict fatigue life.
Since the modulus of the laminates will gradually drop during the fatigue life,
this drop should be taken into account for all design calculations. An estimate
of the loss of modulus can be obtained from the ratio of initial static modulus
and static secant modulus at failure. Modulus drop data for the laminates tested
here are given in Table 3.5.
These data are expected to describe the worst fatigue conditions for uniaxial
loading, because all tests were performed under reverse loading (R = -1).
They can, therefore, be used as conservative estimates for other loading ratios

Influence of Matrix and Fabric

49

(see Fig 8.6). However, to optimize fatigue designs, data at other R ratios
should be obtained. Fatigue data were measured up to a few million cycles. In
many applications more cycles are expected during the lifetime of the struc
ture. These data do not justify the assumption of a fatigue limit at low fatigue
strains; more tests at very low fatigue amplitudes would be needed to clarify
this point. Recent data on different laminates indicate that there is no fatigue
limit up to 109 cycles (22).
Since fatigue failure is basically fibre dominated for high cycle numbers, all
continuous glass fibre reinforced laminates have about the same fatigue
performance in strain space (the same -curve), as long as some fibres run in
the load direction. The 45 laminates showed much lower fatigue properties,
which means fabrics without any fibres in the load direction should be avoided
for fatigue applications. A [0,45] layup will perform well, because some
fibres run in the load direction (see Chapter 5). Laminates with higher moduli
in the load direction have the better fatigue properties in stress space (Fig. 3.4).
3.8

Conclusions

- The matrix affects the static linear limit of laminates and has some influence
on fatigue at high stress-strain amplitudes (short lifetime). An influence on
long time fatigue (low stress-strain amplitudes) is not apparent, because
laminates are saturated with matrix cracks before fibre failure. Matrix cracking
develops in laminates, even if the specimens are cycled at stress-strain
amplitudes below the linear limit.
- Fibre orientation and fabric type affect static strength and fatigue strength,
but the fatigue strain to failure is very similar for all glass laminates, as long as
some fibres run in the load direction. Laminates with no fibres in the load
direction fail much earlier than fabrics with some fibres in the load direction.
- All laminates show a drop in tensile modulus during fatigue. The drop is
gradual over most of the fatigue life, falling to a critical fatigue modulus. The
critical fatigue modulus is similar (within 10 percent) to the static secant
modulus at failure. Therefore, the higher the fibre-content in load direction,
the higher the critical fatigue modulus and the lower the drop of modulus
during fatigue.
- Fatigue failure for all laminates is best predicted by a strain based analysis
using - curves which describe the cycles to failure at a given maximum
fatigue strain. A standard curve can be used to predict failure of all laminates
with some fibres in the load direction.
References
(1) BROUTMAN. L. J. and SAHU. S. 1969. Progressive damage of glass reinforced plastic
during fatigue. Proceedings of the 24th Annual Technical Conference SPI. Section
l l - D , p . 1.
(2) HAHN, H. T. and KIM. R. Y.. 1976. Fatigue behavior of composite laminates. composite
Mater.. 10. 156-180.

50

Design of Composite Structures Against Fatigue

(3) HINGHSMITH. A. L. and REIFSNIDER. K. L... 1982. Stiffness reduction mechanisms in


composite laminates. Damage in composite materials, ASTM STP 775, (ASTM, Philadel
phia, USA), pp. 103117.
(4) OGIN, S. L., SMITH, P. A. and B EAUMONT. P. W. R., 1985, Matrix cracking and
stiffness reduction during fatigue of a (0/90),, GFRP laminate. Composites Sci. Technol., 22,
2331.
(5) ARVESEN, R., 1992, Jotun Polymer, Sandefjord, Norway, personal communication.
(6) ECHTERMEYER, A. T., 1992, Evaluation of the [45] s inplane shear test method for
composites reinforced by multiaxial fabrics, DNV Research Technical Report 922035.
(7) CARLSSON, L. and PIPES, R. B ., 1987, Experimental characterisation of advanced
composite materials, (Prentice Hall, Inc., N.J., USA).
(8) TSAI, S. T., 1988, Composites design. Fourth edition. Think composites, (Dayton, OH,
USA).
(9) HULL, D., 1981, An introduction to composite materials, (Cambridge University Press,
Cambridge, UK).
(10) ECHTERMEYER. A. T. and ENGH, B ., 1992, Obtaining ply properties from 0/90 fabric
reinforced laminates' DNV Research Technical Report 922025.
(11) ECHTERMEYER, A. T., MCGEORGE, D. and BUENE.L., 1993, 'Effect of various glass
and aramid reinforcements on static and fatigue properties of composites'. Proceedings of
Advanced Composites '93, Australia.
(12) ECHTERMEYER, A. T., MCGEORGE, D., ENGH, B . and HAYMAN, E., 1992.
Evaluation of multiaxial glass, kcvlar, and glasskevlar hybrid marine laminates, DNV
Research Technical Report 922062.
(13) SCHUETZ, D., GERHARZ, J. J. and ALSCHWE1G, E., 1981, Fatigue properties of
unnotched, notched, and jointed specimens of a graphite/epoxy composite, Fatigue of fibrous
composite materials, ASTM STP 723, (ASTM, Philadelphia, USA), pp. 31^17.
(14) ECHTERMEYER, A. T., B UENE, L., ENGH, B . and SUND, O. E., 1991, Significance of
damage caused by fatigue on mechanical properties of composite laminates, Proceedings of
the 8th International Conference on Composite Materials, Hawaii, USA, p. 38B l.
(15) ECHTERMEYER, A. T., B UENE, L. and ENGH, B ., 1995, Lifetime and Young's
Modulus changes of glassphenolic and glasspolyester composites under fatigue. Compo
sites, 26, 1016.
(16) B ADER, M. G. and B ONIFACE, L., 1985, Damage development during quasistatic and
cyclic loading in GRP and CRFP laminates containing 90 degree plies. Fifth International
Conference on Composite Materials ICCMV, San Diego, California, pp. 221232.
(17) B ULDER, B . H. and B ACH. P. W., 1991, A literature survey on the effects of moisture on
the mechanical properties of glass and carbon fibre plastic laminates. Netherlands Energy
Research Foundation, ECN, Report ECNC91033.
(18) CAMPONESCHI. E. T. and STINCHCOMB , W. W.. 1982, Composite materials: testing
and design. ASTM STP 787, (ASTM. Philadelphia, USA), pp. 225246.
(19) GANCZAKOWSKI. H. L.. ASHB Y. M. F., B EAUMONT, P. W. R. and SMITH, P. .,
1990, Composites Sci. Technol.. 37. 371392.
(20) B OLLER. . ., 1969, Fatigue fundamentals for composite materials, ASTM STP 460.
(ASTM. Philadelphia. USA), pp. 217235.
(21) TALREJA, R., 1982, Damage models for fatigue of composite materials. Fatigue and creep
of composite materials (edited by H. Lilholt and R. Talreja), Ris National Laboratory.
Denmark, pp. 137153.
(22) VAN DELFT, D. R., RINK. H. D. and JOOSE. P. .. 1994. Fatigue behaviour of fibreglass
wind turbine blade material in the very high cycle range. Proceedings of the European Wind
Energy Association Conference and Exhibition, Greece.

CHAPTER 4

Influence of Spectral Loading


M. Poppen* and P. Bachi

Fatigue is often critical in the design of wind turbine blades. Rotation of the blades in a turbulent
and sheared wind field causes a high number of alternating loads on the wind turbine rotor.
These special load conditions usually result in more than 10s number of cycles during a lifetime.
This means that it is important to be able to predict the fatigue life of the wind turbines with
sufficient accuracy and conservatism. Spectral loading and methods to predict fatigue damage in
fibre reinforced plastics are discussed in this chapter. It is shown that the Palmgrcn-Mincr's rule
for predicting fatigue damage can give unconservativc predictions.

4.1 Introduction
This work considers the different steps in the fatigue design after the loads have
been calculated. The materials can be tested under 'realistic' conditions with
the load spectrum, WISPER, which simulates the loads on wind turbine
blades. Palmgren-Miner's rule can then be used to predict the lifetime of the
spectrum loaded specimens, and that lifetime can be compared to the
measured lifetime of the specimens. This gives a measure for the reliability of
the fatigue life prediction models.
Three different materials have been tested. The tests have been performed
as a collaboration between three institutes, FFA, ECN, and DNV.
4.2 The spectrum
WISPER (Wind SPEctrum Reference) is a standardized load sequence for
horizontal axis wind turbine blades. It simulates the load conditions in the flap
direction at a point close to the blade root. The spectrum was developed in an
IEA working group in 1987 (1). It is based on measurements from nine
different wind turbines in Europe. The sequence represents approximately two
months' service of a generic wind turbine at a rotation speed of 45 rpm.
WISPER consists of a row of 265 423 loading reversal points (end levels). The
maximum load corresponds to level 64 and the rest of the load levels are scaled
accordingly, with zero load at level 25.
The WISPER load sequence is intended for comparative purposes only, i.e.,
*FFA, Sweden.
ECN, The Netherlands.

52

Design of Composite Structures Against Fatigue

to evaluate materials and structural details, dimensions, design alternatives,


and fatigue lifetime prediction procedures.
Simulating an operational period of ten years, WISPER has to be run for
more than fifty times. Experiments with composites cannot be accelerated as
much as tests with metals as the internal heating has to be kept low; this
requires lower test frequencies. Material testing will, therefore, be extremely
time consuming and to obtain statistically reliable results a large number of
specimens is required.
A new spectrum was then developed, WISPERX (1). This new spectrum
contains only one tenth of the cycles present in the original spectrum. The
difference between WISPER and WISPERX is that low load cycles (low in
range) have been excluded. After WISPER had been rainflow counted the
cycles which gave low ranges were omitted. These cycles are supposed to be
low enough not to have a large influence on the fatigue life. Compared to
WISPER's 265 423 end levels, WISPERX contains approximately one tenth or
25 662 end levels. A rainflow count result of the WISPER and WISPERX
sequences are shown in Fig. 4.1. Here the number of cycles of different ranges
is shown. In Fig. 4.2 and 4.3 the distribution of peaks and troughs in WISPER
and WISPERX are presented.
Some specimens are tested using WISPER and some specimens are tested
using WISPERX. The difference in lifetime between the two spectra is also
considered.
4.3 Materials and testing
Tests were performed on three different glass fibre-based composites, two of
which are considered in other chapters. The laminates differed in manufactur
ing method, lay-up and type of glass fibre, and resin (Table 4.1).
Specimens were cut to the shapes specified in Table 4.2. The tensile
properties of the three materials are summarized in Table 4.3.
Glass fibre I epoxy, 30 degrees (FFA)
The material was filament wound at 30 degrees. The coupons were taken
from a cylinder manufactured by Hamilton Standard. The cylinder diameter
was large enough to supply specimens that could be considered as flat. The
material combination and manufacturing method were the same as for the 80
meter rotor on the Maglarp wind turbine in the south of Sweden.
The fibres were -glass. The resin was an Epon 826/Jeffamine D-230 epoxy.
Test procedure
The constant and variable amplitude tests were performed at ECN on loadcontrolled servo-hydraulic testing machines equipped with the Harrier com-

Influence of Spectral Loading

o.i

io

53

WISPER

WISPERX

2 3
io

io

io5

Number of half cycles


Fig. 4.1

Rainflow-counted ranges in WISPER and WISPERX

54

Design of Composite Structures Against Fatigue

37
40
43

46
49
fc

52-

a
6 4 f

0.1

1 (

WISPER
WISPERX

H"

io

io 2

io 3

10"

105

Occurrences
Fig. 4.2

Peak distribution in WISPER and WISPERX

puter interface and a HewlettPackard 9816 microcomputer. In order to


control the fatigue machines in the WISPER sequence tests a special software
program was developed.
The tests at FFA were performed on a servohydraulic MTS machine, with a
MTS Test Processor Interface and a Digital Micro PDP 11/73 Control Com
puter. The tests were run in load control mode. The maximum test frequency
was 5 Hz.
Tests were performed at DNV on a servohydraulic testing machine. In
order to prevent buckling in compression an antibuckling device similar to the
support jig of ASTM 695 was used (Chapter 3). Frequencies of fatigue tests
were varied between 2 and 5 Hz to keep the average load rate constant,
avoiding viscoelastic effects. All tests were run in load control.
All WISPER/WISPERX tests were performed with a constant loading rate
and a sinusoidal (ECN) or saw teeth (FFA) wave shape. It has been shown that
for a similar material, the difference in loading type does not affect the fatigue
life (9). The failure criterion was complete failure of the specimen, i.e., when
the specimen is more or less in two parts.

Influence of Spectral Loading

55

WISPER
WISPERX

31
34-

0.1

10

IO2

IO3

^
IO4

10s

Occurrences
Fig. 4.3

Trough distribution in WISPER and WISPERX

Table 4.1
Reinforcement
type
Fabrics
Fibres
Combination fabric

Types of GRP materials

Orientation
0 degrees and 45 degrees
30 degrees
0/90 WR + CSM

Resin

Reference
(chapter)

Material
type

Isopolyestcr
Epoxy
Vinyl ester

6
This chapter
3

ECN
FFA
DNV

4.4 Constant amplitude data


The designer wishes to predict the lifetime of a particular type of wind turbine
under specific meteorological and conceptual conditions. After the load
spectrum has been evaluated based on the knowledge of this wind turbine, the
next step is to predict the lifetime for a structural component using this
material. This is done by using PalmgrenMiner's rule. It is based on the

Design of Composite Structures Against Fatigue

56

Table 4.2 Specimen shapes


Material
type

Geometry

Length
(mm)

Width
(mm)

Thickness
(mm)

ECN
FFA
DNV

Dogbonc
Straight
Straight

200
150
270

45 to 23*
50
25

S.5
3.5
5

*at waist

Table 4.3

Mechanical properties of the three materials

Material
Glass fibre/polyester, 0 dcgrees/45 degrees (ECN)
Glass fibrc/cpoxy, 30 degrees (FFA)
Glass fibre/vinyl ester, 0/90, CSM (DNV)

Tensile
modulus
GPa

Tensile
strength
MPa

Weight
% fibres

17.8
14
16.8

424
280
227

50
69
53

assumption that each of the blocks of load cycles is consuming a fraction of the
life proportional to the expected lifetime at constant amplitude loading at that
load level. This schematic explanation is valid when only the stress range is of
importance. For glass fibre composites the mean level has to be included. The
material has to be tested at different mean levels as well as different amplitudes
to establish a complete knowledge of the material.
The results can be plotted in a constant life diagram (Fig. 4.4). Each curve
represents a certain expected lifetime, and the combinations of mean level and
amplitude of the alternating stress that the specimen is being subjected to,
which will result in that lifetime. The intersection at the abscissa is the static
strength value as the stress amplitude is zero. Lines originating from the origin
represent a constant ratio between minimum and maximum load, i.e. an R
ratio. To establish a complete diagram, an infinite number of R ratios have to
be tested. Usually a simplification is used with two different R ratios being
tested. Typically R = 1 which means a symmetric tensile/compressive
loading around zero, and also R = 0.1, which results in pure tensile loading.
For applications where compressive loading is present, further R ratios have to
be measured.
The results from the constant amplitude fatigue tests can now be plotted as
Whler curves, i.e., maximum stress versus number of cycles, one for each R
ratio (as in Fig. 6.4). The data can be fitted to a straight line in either a linear-

Influence of Spectral Loading

57

Static tensile stress

Mean stress level


Fig. 4.4

An example of constant life diagram. Each line represents a constant lifetime for a
specimen which has been subjected to an alternating load with a specific combination of
mean stress and stress amplitude

log or a log-log representation. Tests (2) show for a glass/polyester material


that the log-log representation is the most correct. The linear-log algorithm is
too pessimistic for the ultra-high cycle fatigue (Chapter 3). These curves are
then transferred to the constant life diagram. To establish the complete
diagram an interpolation between the tested R ratios and the known ultimate
tensile strength has to be performed. This can be performed by using straight
lines, power curves, or some other type of equation (Chapter 7).
For our materials the following choices were made. The constant amplitude
data were fitted to straight lines in log-log diagrams using least square fit. The
interpolation between different R ratios in the constant life diagram was made
by using straight lines between the known set of data points. The constant
amplitude tests are described in references (3)-(5).
For the combination fabric material (DNV) the constant amplitude testing
was only performed at R = 1. A few specimens were tested at one load level
at R = 0. Based on the knowledge of the other two materials, a slope was
chosen which made it possible to fit a line to the data. The constant life
diagrams of the three materials are shown in Figs 4.5-4.7.
The different shapes of the curves give information about how detrimental
compressive loading (to the left of R = 0) is compared to tensile loading (right
of/? = 0).

58

Design of Composite Structures Against Fatigue


300

500

200
300
Mean stress (MPa)
Fig. 4.5 Constant life diagram, glassfibre0/45 /polyester material

ISO

R0
/

100
3

Q.

so;

^ ^ ^ _ ^ ^ ^ ^

^ s .

=aC^.

1
100

1
200

"' * *

Mean stress (MPa)


Fig. 4.6 Constant life diagram, glass fibre 30/epoxy material

300

Influence of Spectral Loading

59

200 t

*"

s
5 100

\102

io

100

200

300

Mean stress (MPa)


Fig. 4.7

Constant life diagram, Glass fibre/vinyl ester, 0/90, CSM

4.5 Tests with WISPER/WISPERX


Damage summation
For the calculation of the damage of a WISPER/WISPERX sequence, a
rainflow counting procedure resulting in a 64 x 64 fromto matrix is used. Each
element in the matrix can then be seen as the number of half cycles at constant
amplitude loading at a specific R ratio and maximum load (the 'AI'). A damage
64x64 matrix can be determined by dividing the number of cycles in each
element of the fromto matrix by the number of cycles to failure from the
constant life diagram (/V). The damage is determined by summation of the
elements in the matrix. This can be written as

D = Y (nJN
For each tested load level the real life of a specimen can be compared to the
predicted life using PalmgrenMiner's rule. The structural component is
supposed to have failed in fatigue, as the sum D reaches 1.
Instead of using constant life diagrams to obtain N an empirical relation for
the number of cycle to failure for different R ratios, can be used. Appel and
Olthoff (8) modified the Mandell (6) relation and the degradation factor D
after analysing data from literature
= US (1 D log 0 with US = UTS if \ > \
else US = UCS
D = 0.095 0.015 R*
with R* defined as: R* = sign tfmn|amn|/sign
'mnxl^rnax

60

Design of Composite Structures Against Fatigue


Table 4.4

Damage of WISPER and WISPERX according to


Palmgren-Miner's rule

Load level
(MPa)

WISPER

WISPERX

Difference

(%)

148.6
171.4

0.1191
0.5576

0.1184
0.5353

0.6
4.1

with |o"min| the smallest absolute stress or strain and |amax| the largest absolute
stress or strain.
This method has been used as a comparison for one of the materials.
Comparison of WISPER and WISPERX results
In order to compare the lifetime of specimens loaded with the WISPER
sequence with specimens tested using the much shorter WISPERX, some tests
were performed. As some cycles were omitted, the lifetimes of the WISPERX
tested specimens were expected to be longer. The question was whether the
difference was large enough to be taken into account. The tests are more
thoroughly described in (8).
The selected material was glass/epoxy. As several tests were performed on
'identical' material, a good comparison of the two spectra could be carried out.
For the experiments with WISPER, 42 specimens were used at seven different
stress levels. For the tests using WISPERX it was decided that two different
stress levels were enough. One low and one medium-high level were chosen.
Six specimens were tested at the low level and nine specimens were tested at the
medium-high level.
Palmgren-Miner's rule was used to calculate the difference in damage build
up between the two spectra. Calculations were performed to investigate how
important the omitted cycles are based on log-log representation of the
constant amplitude data and a constant life diagram to interpolate between the
tested R ratios. The results are shown in Table 4.4.
For the higher load level the difference in damage sum is larger than for the
lower level. This is expected since it is the small cycles that were omitted, and
for a high load level these small cycles can be large enough to influence the
lifetime of the specimen.
The calculated differences are too small to have a significant influence on the
lifetime of a specimen. According to Palmgren-Miner's rule the two spectra
are almost identical for the stress levels that the tests were run at.
The results from the experiments with WISPER and WISPERX are pre
sented in Fig. 4.8. The load levels on the y axis are the maximum load in the
spectrum. On the logarithmic axis the number of passages through the
spectrum is plotted, which for WISPER corresponds to 265 423 end levels and
25 662 end levels for WISPERX. One passage corresponds to a certain period

Influence of Spectral Loading


300

61

250

WISPER
WISPERX
WISPERX
WISPER

~C^O
% 200

r<BQ,o

1 |

150

100

^>..

1
10
Sequences

Fig. 4.8

*.

100

Comparison between lifetimes of specimens tested using WISPER and WISPERX. Lines
represent the best fit to the data

of service time for an imaginary wind turbine which is the same for both
spectra. The two sets of data points can, therefore, be directly compared. The
plotted lines are the least square fit of the data to a linearlog representation.
Surprisingly, the WISPERX specimens show a shorter lifetime than the
specimens tested with WISPER. However, the difference is small compared to
the scatter in the measured lifetimes for each spectrum.
The tests with WISPERX show that for this material, WISPERX can be used
instead of WISPER. The conclusions are only valid for the tested material in
this load range, but they can also be used for materials with similar fatigue
characteristics. WISPERX is most needed for low load levels, as using
WISPER for these tests would be too time consuming. PalmgrenMiner's rule
indicates that the difference between the spectra decreases as the load level
decreases.
4.6 Discussion
The three materials were tested using either WISPER or the WISPERX load
sequence. The lifetimes of the specimens were compared to the predicted life
times using PalmgrenMiner's rule.
In Fig. 4.9, the results for the glass fibre/polyester 0|45 is shown. The
predicted life is somewhat longer, which means that the method is optimistic
for this material. The alternative way of making the prediction by using a
simple algorithm (Appel/Olthoff) gives a more optimistic result. For glass
fibre/epoxy 30 degrees (Fig. 4.10) the predicted lifetime is almost always
shorter than the measured lifetime. This means that for this material the
method is conservative. However, for the third material (Fig. 4.11), a combi

Design of Composite Structures Against Fatigue

62
350 r.

^^>
Is 300

^^
^N.
"~ ^
O C O C t ^ " ^ *"* * ^,

WISPER
WISPERX
Prediction
AppelOlthofT

5 250

J
I 200
150

' '

1
10

1
102

103

Sequences
Fig. 4.9 Comparison between measured lifetimes and predicted lifetimes of glass fibre
0/45/polyester material

300

WISPER
WISPERX
Prediction

200
E
3

'i

100

-J

100
io
Sequences
Fig. 4.10 Comparison between measured lifetimes and predicted lifetimes of glassfibre30/epoxy
material

nation fabric/vinyl ester shows higher vulnerability to spectrum loading than


the PalmgrenMiner's rule predicts. The difference between the predicted
lifetimes and the achieved lifetimes is very large. The method is, therefore,
unsafe for this material. Earlier reported tests on glass fibre composites using
WISPER (10), (11) showed conservative results. This work indicates that, for
at least one material, calculating the damage with the PM rule can give non
conservative predictions.

Influence of Spectral Loading

63

200

WISPERX
Prediction

150

100

IO2

10

IO3

Sequences
Fig. 4.11 Comparison between measured lifetimes and predicted lifetimes of glass fibre/vinyl ester,
0/90, CSM material

It should be pointed out that all predictions are made using the same
calculation method. No safety factors on the constant amplitude tests are
applied.
It is also possible to find the most severe load cycles according to the
prediction model. Depending on the shape of the constant life diagram,
different cycles will be the most detrimental for a material. For the glass fibre/
epoxy material the most detrimental cycles are the cycles with the highest
range, which for WISPER/WISPERX will be the cycle from 1 to 64 levels. This
cycle does not have a large mean value, but due to the fact that this material is
more or less independent of mean level in this load range it will be the most
severe cycle. For the other two materials, cycles with an R ratio in the region of
0 (pure tension, no compressive part) have a large influence on damage,
according to the prediction model.
4.7 Conclusions
This work indicates that, at least fr one material, the Palmgren-Miner's rule
can give unconservative predictions. The WISPER tests on the other materials
can be predicted using a linear damage calculation based on a constant life
diagram.
References
(1) TEN HAVE, . ., 1991, WISPER and WISPERX final definition of two standardised
fatigue loading sequences for wind turbine blades, NLR report NLR TP 91476.
(2) JOOSSE, P. A. and VAN DELFT. D. R. V., 1993. Fatigue behaviour of fibre glass wind
turbine blade material. Proc. ECWEC '93. Travemunde.

64

Design of Composite Structures Against Fatigue

(3) BACH, P. W., 1992, Fatigue properties of glass- and glass/carbon-polycstcr composites for
wind turbines, ECN-C-92-072, 1992.
(4) BLOM, . F., 1982, Fatigue evaluation of WTS-3 glass fibre blade material, FFA TN 198226, The Aeronautical Research Institute of Sweden.
(5) ECHTERMEYER. A. T., ENGH, B., and BUENE, L., 1993, Fatigue of composite
structures-influence of matrix and fabric. Det Norske Veritas Report No. 92-2067.
(6) MANDELL, J. F., 1982, Development in reinforced plastics-2, (edited by Pritchard),
(Applied Science Publishers), p. 67.
(7) POPPEN, M.. 1991, Comparisons between the two spectra WISPER and WISPERX, FFAPH-l 128, The Aeronautical Research Institute of Sweden.
(8) APPEL, N. and OLTHOFF. J.. 1988, Voorontwerpstudic NEWECS-45, Polymarin rapport.
(9) BACH, P. W., 1991, High cycle fatigue of glass fibre reinforced polyester and welded
structural details, ECN-C-91-010.
(10) POPPEN, M. and BACH, P. W., 1991, Fatigue testing using the WISPER Spectrum, Proc.
EWEC'91, Amsterdam.
(11) POPPEN, M., 1989, Fatigue testing of glass fibre reinforced epoxy using the WISPER
sequence, FFA TN 1989-45.

CHAPTER 5

Effects of Environment
C. W. Kensche*
The exposure offibrecomposites to moisture and the impact of hailstones are studied. After such
environmental treatment, the accumulation of damage is monitored during both static and
fatigue testing. A design strain level of 0.6 percent is recommended for safe-life operation of
glass/epoxy composites.

5.1 Introduction
Rotor blades, like boat hulls, are exposed to a hostile environment throughout
their service life. Such effects include: temperature fluctuations; rain; formation of ice; thunderstorms; hailstones; erosion from sand particles; and
extreme dryness in desert environments. As with boats, a resin-rich layer (gel
coat) is always used on the outer surfaces to give protection against such
elements.
Three effects were studied in some depth - hailstones, high and low humidity
(1). The results were compared with data of identical but unaffected material.
Additional comparison was possible with the results of (2), obtained with the
same material. The work on moisture complements that described in Chapter
11 with a different material composition and specimen configuration. The
effect of hailstones was studied using an air-gun to launch the projectiles. By
investigating both static and fatigue properties, it is possible to propose a
design strain level for safe-life operation of glass/epoxy composites.
5.2 Materials and specimens
Laminates were manufactured by DLR in order to obtain specimens of
laboratory quality to avoid scatter in lifetime by poor quality materials. A
combination of two glass fabrics orientated at 45 degrees and 0 degrees was
used (Fig. 5.1, Table 5.1). The 0 degreefibrestook the principal loading, whilst
the 45 degree-angled fibres were incorporated to withstand the shear forces
inherent in structures such as rotor blades. The matrix was an epoxy resin type
L20/SL from Bakelite. Small voids were present, but the estimated porosity <1
percent was unlikely to influence the properties significantly.
Ten plates measuring 520 x 500 mm2 and with a mean thickness of 2.2 mm
*DLR, Germany.

66

Design of Composite Structures Against Fatigue

t X X > l 45degrees

UD

Fig. 5.1 Lay-up of the Gl-Ep specimens

Table 5.1 Materials and lay-up


Fabrics
Lay-up
Resin
Manufacture
Curing
Post-cpring

(a) ' glass, 2/2 twill balanced, 280 g/m2, type 92125 (Interglass)
(b) 'E' glass, 387 (0)/38(90), 425 g/m-, type 92146 (Interglass)
[45(a)[lp]/0(b)[2p]]5
Epoxy type L20/SL (Bakelite)
Pressed between plates
24 h, 20C
6 h, 60C; 15 h, 80C

were manufactured with a mean fibre volume fraction of about 40 percent.


They were supplied with a thin peel ply of Ulmia 7849 for ease of glueing tabs
without roughening the surface. The plates were cut into two parts; one of them
was taken for exposure to humidity (without peel ply), the other one was stored

Effects of Environment

67

Tabs

Fig. 5.2

Compression test specimen (Celanese type) with clamping device

Tabs
Fig. 5.3

Specimen

Tabs with notches for prevention of debonding

in a laboratory climate to avoid differences in properties due to the influence of


fabrication.
Three different geometries were used for the specimens, depending on the
purpose of the static tests: tension, ILSS (interlaminar shear stress), and
compression:
- Tension
- ILSS
- Compression

Straight, 250 x 20 x 2.2 mm3, without tabs;


20 X 10 x 2.2 mm3;
Celanese type, principle see Fig. 5.2.

The free length between the tabs of the compression test specimens was varied
over a range between 8 and 32 mm. The reference results were obtained with 12
mm free compression length.
The geometry for the fatigue specimens tested in reverse loading was similar
to that of the tension test specimens, but 20 mm shorter and supplied with tabs
(Fig. 5.3). These tabs were notched in a special manner to avoid the high shear
stresses at the beginning of the bonding area between specimen and tabs. Since

68

Design of Composite Structures Against Fatigue


5

^--Tab

- Sample

Clamp
plates
Spacer
plate
Fig. 5.4

Principle of antibuckling guide

the fatigue tests were performed in axial compression loading, the use of an
antibuckling guide was inevitable (Fig. 5.4).
5.3 Exposure to humidity
Polymeric materials (like epoxy and polyester resins) are able to absorb
moisture either from the humidity in the surrounding air or by diffusion of
water through the matrix. The water molecules deposit on the polymer chains
and lead to a softening of the resin and, thus, of the fibre compound. The
velocity of the water diffusion is influenced by the temperature; but is
independent of the surrounding humidity. However, the maximum moisture
content in the material is dependent on the humidity of the environment. An
objective indication of the softening effect is the glass transition point Tg, which
can be measured by the DSC method. The increase of moisture is stated by
measuring the weight gain which is schematically recorded against the square
root of time. The diffusion procedure is reversible when the material is

Effects of Environment

5
Fig. 5.5

69

10
Time (years)

Numerical simulation of moisture penetration in a rotor blade

exposed again in dry conditions. This procedure is used to determine the


absolute moisture content after the receipt of the material.
Moisture absorption can be calculated by using Fick's law as the coefficient
of diffusion into the matrix system L20/SL is known (3). Using the FEA
program system SMART, a simulation of thirty years' moisture absorption for
alO mm thick rotor blade was made under the assumption of cyclic cold-wet
and dry-warm weather conditions related to a one-year cycle measured at the
DLR wind turbine testfield (4). The calculation showed that twenty years'
natural weathering resulted in a constant moisture content of 0.4-0.5 percent
(Fig. 5.5). The reduction in the moisture penetration rate due to the gel coat
protection was not considered in the simulation.
In many tests moisture gain is caused by simply exposing the material in
warm water. But then the moisture content does not seem to be controllable.
Some airworthiness requirements demand an exposure to a relative humidity
of 85 percent for certification of structures to meet more realistically the
changes of surrounding humidity during the lifetime of an aircraft. In an FFA
study, the material was prepared for testing in 95 percent relative humidity (5).
In our investigation the exposure of the material took place in a tropical
chamber at 90 percent relative humidity. To accelerate the uptake of moisture,
the surrounding temperature was raised to 45C. A higher temperature might
have compensated the desired negative moisture effects by postcuring of the
matrix. The increase of the moisture content was measured on 'traveller
specimens'. Figure 5.6 shows that saturation was achieved after about 120 days
of exposure at a humidity level of 0.4-0.5 percent. This agrees with the
analytically simulated moisture absorption. The plates which were exposed to

70

Design of Composite Structures Against Fatigue

-0.3

Fig. 5.6

6
8
Time (days)"

I
10

L
12

Moisture absorption of traveller specimen at 45C/90 percent RH and desorption at RT/5


percent RH

humidity (later called wet) were stored in the humid climate, without further
weight gain, up to the time they were prepared for testing.
To determine the absolute moisture content, traveller specimens were also
desiccated at room temperature (RT). The loss of weight was between 0.1 and
0.2 percent and corresponds with the moisture content in the as-received status
(later called 'dry'). One half of each plate was exposed to humidity. After
saturation the fibre volume fraction, , and glass transition temperature, TQ,
were measured on both halves of the plate. The dry and wet plates were
impacted as described above. Then the test specimens were cut and prepared
for further investigation.
5.4 Hailstone simulation
Literature differentiates up to seven forms of hailstones, depending on the time
of growth and the conditions in thunderstorm clouds (6). The mean density is
about 0.9 g/cm3. Their frequently milky appearance is due to voids in the
volume which determines the density.
The terminal velocity of hailstones can be described by the empirical formula
V = 9 x D 0 8
with V*, in m/s and D in cm. To determine the impact velocity of a hailstone
onto a rotor blade, the terminal velocity of the hailstones was superimposed

71

Effects of Environment
Table 5.2

Velocity and energy of hailstones on the tip of a 25 m diameter wind turbine

Stone 0
(mm)

Mass

V
(m/s)

Vwinrf

(m/s)

(m/s)

total

(m/s)

Piotai

(g)

10
15
21
25
31
37

0.3
1.9
5
8
15
25

9
12.4
16.2
18.7
22.2
25.6

65.4
65.4
65.4
65.4
65.4
65.4

35
35
35
35
35
35

101.5
102.5
104
105.1
106.9
108.8

1.54
10
27.1
44.22
85.74
148

(J)

Specimen

Fixture
Foam

Time counter

Fig. 5.7

Impact testing rig

with the maximum velocity components of a rotorblade of the type DEBRA 25


with a diameter of 25 m and with 50 min -1 rpm. This resulted in a hailstone
velocity of more than 100 m/s for hailstone diameters between 10 and 37 mm,
see Table 5.2.
In the laboratory Gl-Ep plates were impacted during pre-investigations with
artificially-manufactured iceballs up to 37 mm in diameter to choose a suitable
size for the projectiles and to gain impacting experience. Figure 5.7 shows the
principle of the impact testing rig.
By controlling the air pressure, the projectile was accelerated in the tip of a
sabot up to the necessary velocity which was measured by two light barriers just
before striking the specimen. This triggered a camera with a time-adjustable
flash light to take a photograph either before or during impact, as shown in Figs
5.8 and 5.9 for the case of 37 mm iceballs. A hard foam disk was mounted
behind the circularly-clamped Gl-Ep plate to prevent it from failure in the rim.

72

Design of Composite Structures Against Fatigue

Fig. 5.8

Icehall of 037 mm before impact

Fig. 5.9

Impact of a 037 mm iceball

After impacting the whole plate was cut into suitably-sized pieces for the test
specimens.
The damage by the hailstone impact on the front side of the specimen was
barely visible, whereas the rear side showed a whitened area of a size
corresponding to the diameter of the iceball. The impact tests were continued
with a diameter of 20 mm. Thus, the 20 mm-wide test specimens were damaged
over the whole cross-section, and as many specimens as possible were cut out of
the plates. Considering the chosen velocity of about 105 m/s the hailstone
impact energy resulted in about 27 Joules.
The whitening on the rear side of the laminate plates after hailstone impact
was not due to delaminations, but to microcracks between matrix andfibresin
the outer 45 degree fabric (Fig. 5.10). Figure 5.11 shows a photomicrograph
of the whole cross-section with microcracks and voids. Both dry and wet plates
were impacted.
5.5 Testing procedure
The following combinations of material were now available.
- dry, undamaged;
- dry, damaged;
- wet, undamaged;
- wet, damaged.
The static properties tension, ILSS, and compression were tested in a screw
test machine (Zwick). An average of six specimens were measured for each

Effects of Environment

Fig. 5.10

Damage by hailstone
impact on rear side of
Gl-Ep plate

Fig. 5.11

73

Cross-section micrograph of an impact damaged


specimen

type of static test and condition status to obtain statistically well-determined


results.
The tensile specimens were clamped between hydraulic grips without tabs.
The strain was measured by strain gauges mounted in the gauge length.
The problem with compression tests is that the free compression length must
not be too large, in order to avoid instability failure. In our investigation the
Celanese-testing rig was used, which has a norm free length of 8 mm, see Fig.
5.2. This appeared too small for testing the effects of impact and humidity on
compression strength. So a parametric study was made of the free length (see
below).
For calculation of the 'E' moduli the initial modulus was taken, i.e., the
linear fit between 0 and the linear limit (Fig. 0.1). In our case the linear limit
was defined to be between 10 percent and 50 percent of the failure load.
Interlaminar shear stress in such structures occurs especially in bondings
between spar caps and shear webs as well as in areas of load introduction. Thus,
some ILSS tests were performed on short-beam specimens to obtain information about the conditioning effects on that loading. Since the crack pattern
on the front and rear side of the impacted specimens was different, they were
tested from both sides.

74

Design of Composite Structures Against Fatigue

The fatigue properties in the different modes of conditioning were of special


interest since they can give information about the lifetime of the rotor blades.
A basis for comparison with non-damaged material can be obtained from
investigations described in (2).
For the fatigue tests which were carried out at an R ratio of 1 (tensioncompression), the application of an antibuckling guide was necessary to avoid
instability of the specimens (dimensions 230 x 20 x 2.2 mm3). For observation
purposes, and also to admit failure mechanisms of the conditioned speci
mens probably differing from the dry/undamaged ones, the antibuckling guide
was supplied with a window of the size 16 x 16 mm2, see Fig. 5.4. The long
time tests with humid specimens (at strain levels lower than 0.8 percent)
were carried out in a humid atmosphere to prevent the material from drying
out.
5.6 Results
Fibre content and glass transition point
Three pieces were cut out from each manufactured plate to specify the fibre
content. The fibre volume content, , was between 38.6 and 42.4 percent for
the individual plates, the mean value for all plates being 40.551.65 percent
(Table 5.3). During the tests, no effect on the results induced by the scatter was
observed.
Table 5.3 also shows the glass transition temperature Tg for dry and wet
material which was measured by means of the DSC method. The Tg for the
former was about 100C, as measured previously (2). As expected, the Tg for
the wet material was only about 91C, indicating a softening effect due to the
moisture.
Static tests
The results for the tensile tested specimens are presented for stress, strain, and
the 'E' modulus in Fig. 5.12 and Table 5.3. Only small differences can be
observed, since the tensile properties are dominated more by thefibresthan by
the matrix. The strength for the wet material is about 3 percent lower than for
the dry material. The strength of the damaged specimens in the dry state is a bit
higher than for the undamaged ones in the dry and the wet stage. This could
be explained by the fact that the damage-induced cracks slow down the
maximum shear stresses between the layers of the laminate. The initial moduli
do not show significant differences, as the failure strains follow the stresses in
their variation.
The ILSS tests show a decrease of about 16 percent for the wet material
properties compared to the dry ones; whereas there is nearly no change
between damaged and undamaged material. The results are shown in Fig. 5.13

Table 5.3

Stiffness
Initial tensile modulus
Compressive modulus
Tensile failure data
Stress at failure
Strain at failure
Secant modulus at failure
Compressive, failure data
12 mm free length:
Stress at failure
Strain at failure
8 mm free length:
Strain at failure

Material properties of the tested GlEp (Type 2A)

Dry. undamaged

Dry, damaged

Wet, undamaged

Wet, damaged

Standard deviation

Standard deviation

Standard deviation

Standard deviation

25.29
25.10

0.76
2.49

25.48
25.51

0.94
2.83

9.03
0.04
0.442

556.43
2.263
24.586

8.52
0.035
0.694

557.64
2.258
24.693

17.9
0.02
0.59

CD
O

484
1.672

18
0.079

322
1.271

13
0.164

n
o
m
<
3'

2.277

0.155

E,, GPa

26.04
30.56

1.31
1.54.

27.57
28.97

MPa
%
Er. GPa

575
2.26
25.1

11
0.04
1.18

587.71
2.34
25.13

485
1.604

24
0.127

2.199

0.156

E, GPa

o.
f.

F,

MPa
%

ILS
Interlaminar Shear Stress

MPa

Other properties
Fibre volume fraction
Glass transition point

Tg, C

o,

Initial modulus defined as:


Compression modulus defined as:
Linear limit defined as:

1.68
0.99

29
0.13

337
1.337

m
3S

3
CD
3

52.25

1.44

51.63

0.93

44.78

40.55
100

1.65

40.55
100

1.65

40.55
91

3
1.65

44.57

0.65

40.55
91

1.65

linear fit between 0 and linear limit


linear fit between 0 and linear limit
offset between 10 and 50 percent of failure load
~4

76

Design of Composite Structures Against Fatigue


600 r

3" 400 -

200 -

eu
O
3

s
Dry

Dry

Wet

Wet

Undamaged Damaged Undamaged Damaged


Fig. 5.12

Static tensile properties of differently conditioned Gl-Ep; the scatter about the mean is
shown by horizontal shading

and in Table 5.3. Thefigureshows that there is no difference in the mean ILSS
value for the dry, damaged specimens whether they are loaded from the front
or the rear side of the impact. Only the standard deviation in the latter case is
higher.
The static compression properties were of special interest for the design of
the size of the window in the antibuckling guide. The results of the preinvestigation with specimens of a free length between 8 and 32 mm show that

Effects of Environment

Dry
Undamaged
Fig. 5.13

Dry
Damaged

77

Wet
Wet
Undamaged Damaged

Interlaminar shear strength of differently conditioned Gl-Ep

600

20
Free length (mm)
Fig. 5.14

32

Compression strength versus free length

instability begins at a free length between 10 and 12 mm, see Fig. 5.14. It was
decided therefore, to continue the tests with a 12 mm free length for the
compression tests.
Tests were carried out with six specimens each for the four different
combinations of conditioning. Figure 5.15 and Table 5.3 show the results for
the compression strength, strain, and initial modulus. There is a decrease of
about 30 percent between the strengths of the wet and dry material, but no
difference between the undamaged and the damaged material in the dry state.

78

Design of Composite Structures Against Fatigue

Dry
Undamaged
Fig. 5.15

Dry

Wet

Damaged

Undamaged

Wet

Damaged

Static compression properties of differently conditioned Gl-Ep, 12 mm free length

The wet, damaged specimens achieved a little higher strength than the wet,
undamaged ones. For the failure strain and the initial modulus the decrease is,
however, not as high but still about 21 percent for the strain and about 16
percent for the modulus. In strain the damaged materials again show higher
values than the undamaged ones. No significant differences in the stiffness can
be recognized due to the hailstone impact.
The relatively large degradation of the static compression strength and strain
must be related to the moisture-induced weakness of the polymer matrix which
supports the axial compressed glass fibres. In view of the fatigue behaviour it

Effects of Environment

79

2.5

S
a
2.0

-95/95 limit [1]

'I 1.5
I i-o

0.5

Undamaged
* Damaged
G Undamaged, no window

10
Fig. 5.16

102

13
io

JL4

10
Cycles

105

10'

107

10'

e- data of unconditioned GI-Kp at R = 1

must be considered that at maximum strains close to the failure strains, which
are at about 1.3 percent, a large decrease of the load cycle numbers may be
expected.
Fatigue tests
As a first step in the fatigue investigations comparison tests were performed
with :hp 1 dry undamaged.ehp specimens to compare with earlier results (2).
Figure 5.16 shows that the new data points are within the scatter of the previous
investigation and beyond the curve of 95 percent survivability from (2).
The tests were continued by fatigue testing of the other specimens. The
stiffness was measured by clip gauge extensometer in a first static tension test
immediately before starting the fatigue loading and at 103 and 104 load cycles.
There was some stiffness degradation from the condition dry undamaged:ehp
down to wet damaged, see Fig. 5.17. The largest decrease of stiffness was about
4 percent. Damage accumulated in the dry, undamaged samples by cracking in
the 45 degree layer after 103 cycles (Fig. 5.18) and after 10 cycles in the rim
zones (Fig. 5.19). Both sets of cracks caused a small decrease in stiffness. As
testing continued, delaminations increased in size, and, shortly before failure,
fibre cracks started as well. The dry, damaged material differed only from the
dry undamaged material by having the 45 degree microcracks present from the
beginning. The wet material developed microcracks between the 45 and 0
degree layers.

80

Design of Composite Structures Against Fatigue


101

looov^
^

. ^*~ ;~--~
*
___
^^
~~~

*^
'

9 9

~~
^^^
^ w^

...
x,
v

^v

'

>

"*

*** ^

x Dry, undamaged
* Dry, damaged

*"

ml
10

"" ~

**

^ ^.
^

v
*s

*^.
^ vX
.\ . *
*"* ^."""*

""ri
v.

O Wet, undamaged
o Wet, damaged

1
io 2
Cycles

1
10'

"^ 0

IO4

Fig. 5.17 Stiffness changes versus load cycles for GlEp with different states of conditioning,
measured with clip gauge

Figure 5.20 shows the data for the various conditioned materials. Table
5.4 contains the single static and fatigue data points. The fatigue data of the dry
and wet material do not show differences between the undamaged and the
hailstonedamaged specimens. B ut, as shown by the evaluation of the static
ILSS and compression data (Figs 5.13 and 5.15), a negative influence due to the
moisture must also be stated in fatigue. At strain levels of 0.8 percent and
more, the load cycle number was significantly reduced, i.e., by more than an
order of magnitude. However, at a strain level of 0.6 percent, load cycle
numbers were similar for the wet as for the dry material. Thus, another failure
mechanism seems to occur at this lowest tested strain level.
5.7 Statistical evaluation
For each data set a statistical evaluation was made using the twoparameter
Weibull distribution according to the Sendeckyj method (7) (Fig. 5.20).
Following the European certification rules, the curves represent the fatigue
lives of 95 percent survivability with 95 percent lower confidence limit. The
shape and scale parameters and and the model parameters C and 5 for this
data and those described in (2) are given in Table 5.5.
The resulting fatigue lives were plotted only one order of magnitude higher
than the highest fatigue results, and for the case in which no static tests were
available, they were plotted an order of magnitude lower than the lowest

Effects of Environment

Fig. 5.18

81

Damage propagation on dry specimen

fatigue test point. This was the case for the wet material since the static
compression tests were carried out with a free length of 12 mm, as mentioned
above, and failed too early due to instability effects. Thus, these tests cannot be
used for the statistical evaluation. For the fatigue curves for the dry material.

82

Design of Composite Structures Against Fatigue

Fig. 5.19

Damage propagation on wet specimen

the individual static tensile and compressive results were used. It is interesting
that the static compression properties (Table 5.3) and the fatigue properties of
the damaged Gl-Ep were slightly superior to the undamaged ones.
However, since the resulting fatigue curves of the dry and wet material were
close together a pooling of all dry and wet material data including those of (2)

Effects of Environment

83

x Dry, undamaged
* Dry, damaged
O Wet, undamaged
o Wet, damaged
Curves 95/95 limits

Fig. 5.20

e- data and curves of 95 percent survivability and 95 percent lower confidence limit of
differently conditioned Gl-Ep at R = 1. is static tensile and C static compressive
strain

was carried out. The new fatigue lives are plotted in Fig. 5.21. Since they are
presented as curves of 95 percent survival ability and 95 percent lower
confidence limit they represent certifiable values. Their statistical parameters
are given in Table 5.5. In the low-cycle area, the fatigue line of the wet
compared with the dry Gl-Ep shows a significant decrease, whereas in the
(extrapolated) high-cycle region, it crosses the fatigue line of the dry material.
To characterize the dry and the wet Gl-Ep, it is proposed that the slope 5 be
used in combination with the strain at a high load cycle number. In (8) the
reciprocal value k of 5 is used as a parameter. Therefore, in Table 5.5 the k
values and the certifiable strains at 108 load cycles are given for the curves of the
pooled data. For the dry material, k is 11.5 and the corresponding high-cycle
strain nearly 0.4 percent; for wet Gl-Ep, k is 15.7 and the strain is very close to
0.4 percent.
Design allowables for glass epoxy
On the basis of these results, the possibility of new possible design allowables
for rotor blades made of Gl-Ep material arises. The fatigue curves for both dry
and wet Gl-Ep showed a certifiable 0.4 percent level at 108 load cycles. These
were obtained by testing at a stress ratio of R = - 1 which gives the most
conservative values. The average loading ratio of rotor blades, however, is
either more tensile or compressive dominated (refer Fig. 4.1). This would lead
to fatigue lives superior to R = 1 curves for Gl-Ep and for glass polyester
(Gl-UP) (9).

84

Design of Composite Structures Against Fatigue

Table 5.4 Test results of dry and wet Gl-Ep


Gl-Ep. dry. damaged

Gl-Ep. dry. undamaged


Load cycles
1
1
1
1
1
1
1
1
1

Load cycles
2346760
358460
268190
199010
94740
82170
45860
40570
25450
14580
19930
15820
2820

Static tests
Strain (%)
2.339
2.272
2.183
2.148
2.461
2.017
2.284
2.104
2.179

Fatigue tests
Strain (%)

Observation
Tension
Tension
Tension
Compr. 8mm
Compr. 8mm
Compr. 8mm
Compr. 8mm
Compr. 8mm
Compr. 8mm

Observation

7638390
5500000
22530
9630
8880
5500
3960
3410
5120
4390
2900
2880

Fatigue tests
Strain (%)
0.600
0.600
0.800
0.800
0.800
0.8(H)
0.800
0.80(1
1.000
1.000
1.000
1.000

Static tests
Strain (%)
2.258
2.350
2.411
2.232
2.408
2.126
2.372
2.106
2.556
2.208
2.216

Load cycles
1404870
759000
718430
349530
53230
52640
49670
45020
20360
4010
3310
2270

0.800
0.800
0.800
0.800
0.800
1.000
1.000
1.000
1.000
1.000
1.200
1.200
1.200

Fatigue tests
Strain (%)

Observation
Tension
Tension
Tension
Compr. 8mm
Compr. 8mm
Compr. 8mm
Compr. 8mm
Compr. 8mm
Compr. 8mm
Compr. 8mm
Compr. 8mm
Observation

0.800
0.800
0.800
0.800
1.000
1.000
1.000
1.000
1.000
1.200
1.200
1.200

Gl-Ep. wet. damaged

Gl-Ep. wet. undamaged


Load cycles

Load cycles

Observation

Load cycles

Runout
Runout

6500000
29870
21720
15330
6380
1110
1030
730
670

Fatigue tests
Strain (%)
0.600
0.800
0.800
0.800
1.000
1.000
1.000
1.000
1.000

Observation
Runout

Effects of Environment

85

Table 5.5

Weibuli parameters for the fatigue curves

Fatigue
tests

Weibull parameters

Model parameters
C
S

13
12
12
9

13.756
24.494
13.366
17.562

2.243
2.398
1.617
1.563

0.4
4.84
1
1

Dry, undamaged (1)

14

20.109

2.236

0.01744

0.11756

8.5

Pooled data, dry


Pooled data, wet

39
21

13.988
15.066

2.230
1.600

0.22
1

0.0868
0.0635

11.5
15.7

Dry,
Dry,
Wet,
Wet,

undamaged
damaged
undamaged
damaged

Slope k
= l/S

0.0794
0.072
0.0646
0.061

Strain at
N=108

12.6
13.9
15.5
16.4
0.394
0.370

2.5

Dry
A Wet
2.0:

95/95 limit dry

95/95 limit wet

' '

10

Fig. 5.21

lll 2

_L

_L

J_

_L

_L

I fr'

IV

10'
Cycles

10*

10'

10"

IO-

10"

Fatigue curves of pooled data of differently conditioned GlEp at R = 1

The design allowable, however, refers to the highest load which is antici
pated to occur only once during the life of a rotor blade. Provided that this
'century gust' yields about 50 percent higher loads than the maximum loads
during normal operation, it can be argued that at a strain level of 0.6 percent,
very high load cycle numbers were achieved for the tested specimens (Fig.
5.20). Thus, without looking more deeply into the service life of a rotor blade,
the 0.6 percent level is proposed as the possible design strain level for GlEp.
However, the stress distribution in a rotor blade or a structural element is
more complex because of
(a) stresses in the shear web;

86

Design of Composite Structures Against Fatigue

(b)
(c)
(d)
(e)

interlaminar shear stresses between theshear web and the spar cap;
stress concentrations in the load introduction zone;
areas of steep decrease or increase of stiffness;
local instabilities etc.

All these effects may diminish the possible lifetime of a structure. Therefore, it
is recommended that the fatigue life of a whole structure be ascertained by
means of constant amplitude tests (service life tests at this low level would take
too much time) at a relatively high strain level, e.g., 0.6 percent or higher. If no
damage growth is observed after a specified number of cycles (e.g., 10000 or
100000) it may be assumed that during the service life of a wind turbine, no
fatigue failure will occur on the rotor blades. This should be proven in further
investigations.
5.8 Conclusions and recommendations
Environmental effects on the static and fatigue properties of GRP rotor blade
material were studied. Gl-Ep laminates were exposed to 90 percent relative
humidity up to saturation. As-received (dry) and moisture-affected (wet)
laminates were impacted by iceballs of 20 mm diameter and a velocity of about
105 m/s, i.e., with an energy of about 27 Joules. The moisture content after the
exposure was between 0.4 and 0.5 percent. The damage by the hailstones was
limited to microcracks in the 45 degree layers on the front and the rear side of
the specimens.
For the static and fatigue investigations pre-conditioned material was available in the state
-

dry, undamaged;
dry, damaged;
wet, undamaged;
wet, damaged.

It was shown that the hailstone-simulating iceball impact does not harm the
properties of the Gl-Ep investigated, whereas humidity caused a significant
decrease in the static properties and also in the fatigue properties at R 1 and
strain levels of 0.8 percent and more. At 0.6 percent, however, the results of
the wet and the dry Gl-Ep were similar.
The static and fatigue results were statistically evaluated according to the
Sendeckyj method (7) using the two-parameter Weibull distribution. The
fatigue lives were presented as curves of 95 percent survival ability with 95
percent lower confidence limit. The test results of the dry material including
those described in (2) and the wet material each were pooled together for the
statistical evaluation.
Whereas in the low-cycle area, the fatigue life of the wet Gl-Ep was much
lower than that of the dry one, the curves showed similar results in the highcycle space, e.g., at 108 load cycles, both curves were only just below 0.4

Effects of Environment

87

percent. The reciprocal k values of the slope are 11.5 for dry and 15.7 for wet
GlEp.
A proposal is made for a new possible design allowable on the basis of the
results obtained with the coupon specimens and additional tests carried out on
structural components or complete rotor blades.
References
(1) JGG, S, 1992, Einflu von Hagelschlag und Feuchte auf das Festigkeits und Steifigkeitsver
halten von G FKLaminaten, DLR, IB 43517/92.
(2) KENSCHE, Ch. W, 1992, High cycle fatigue of glass fibre reinforced materials for wind
turbines, DLR Forschungsbericht 9217.
(3) SPRINGER, G. S, 1981, Environmental effects on composite materials, (Technomic Publish
ing Company).
(4) ROHM, A, 1989, Berechnung der Feuchteaufnahme einer G FKStruktur und der daraus
resultierenden Spannungen mit der Methode der Finiten Elemente, DLR, IB 43509/89.
(5) B LOM, . F., 1984, Influence of moisture and elevated temperature on the fatigue properties
of WTS3 glassfibre blade material, FFA TN 58.
(6) LAUB E, HLLER, Numerical data and functional relationship in science and technology.
Vol. 4, Meteorology, (Springer Verlag).
(7) SENDECKYJ, G. P., 1981, Fitting models to composite materials fatigue data, Tesis methods
and design allowables for fibrous composites, ASTM STP 734, (Edited by C. C. Chassis), pp.
245260.
(8) Richtlinie fr die Zertifizierung von Windkraftanlagen, 1993, Chapter 5.2. (Germanischer
Lloyd).
(9) KENSCHE, Ch. W. (editor) et al., 1995, Fatigue of materials and components for wind turbine
rotor blades, (European Commission).

CHAPTER 6

Glass and Hybrid Fibre Performance


P. W. Bach*
Fatigue tests arc performed with coupon specimens and bolted joints from glass- and hybrid
glass/carbon-fibre reinforced polyester plates (GFRP and GCRP), produced under industrial
conditions.
The analysis of the constant amplitude tests on a 0/45 degree GFRP laminate shows that
there is no indication of a fatigue limit and that the data fit better to a line on a log-log than on a
linlog basis. It is also observed that both on a strain and on a normalized stress basis the fatigue
behaviour is dictated by the unidirectional layers.
A positive hybrid effect is not observed in the present fatigue investigation on laminates and
bolted joints where the results arc compared with equivalent plain glass specimens.

6.1 Introduction
The most fatigue critical part of a wind turbine is the rotor blade structure and
its connection to the hub. In a design life extending over more than twenty
years it has to ensure high cycle (>10 8 ) fatigue loading of variable amplitude
due to gusts, windshear, start and stop procedures, etc. Very few data are
available in this high cycle fatigue range for the materials and substructures
used in wind turbines, especially in combination with environmental effects.
One of the materials that is extensively used in wind turbine blades is glass
fibre reinforced polyester (GFRP). There is currently also some interest in
hybrid material (glass/carbon fibres, GCRP) in order to save weight in the
heavily loaded parts, e.g. in the joints.
The inexpensive glass fibres have a large elongation at fracture and a high
impact energy compared to carbon, but a lower modulus and compression
strength. Carbon fibres generally have a better fatigue resistance (rather flat
fatigue curve) than glass. Due to their low compression strength, aramid fibre
reinforcement is not considered.
A composite prepared by combining mixed fibres has mechanical properties
which can deviate from the linear summation of the contribution of the relative
volume fractions of the fibres (rule-of-mixtures). It is expected from a literature survey that a positive hybrid effect exists for the fatigue performance (1).
The possible benefit of incorporating carbon fibres in glass fibre reinforced
polyester is investigated by a comparative fatigue study of a plain glass laminate

*ECN, The Netherlands.

90

Design of Composite Structures Against Fatigue


Table 6.1

Fabrics

Mass
(g/m1)

Warp
(g/m2)

Weft

Glass UD

500

495

Glass 45

480
800

Glass + or - 4 5

700

Carbon UD

200

690
10
199

240
(-45)
100
(random)
10 (+45)
690 (-45)
1 (polyester)

Stitch bonded

Glass UD 700/CSM 100

240
(+45)
700

Carbon 45

430

215

-(+45)
215 ( - 4 5 )

Stitch bonded

Type

Manufacturing
process
Woven

Inlaid; powder
bonded
Woven
Woven

Fabric type
Syncoglass
NW5/500
Gevetex
Emax 430
Ahlstrm
700/100
Ahlstrom
R24-700-45
Enk
C-UD-200
Soficar C mux
430

Notes: (i) Resin: polyester Synolitc 593-E-l (DSM)


(ii) Postcuring 24 h at 60C

and a glass/30 percent carbon laminate, both in coupon form and as bolted
joints.
The design strength and stiffness of rotor blades have to be maintained
during the lifetime of a wind turbine. The stiffness is especially important
because stiffness degradation of the materials has an effect on the aerodynamics and the resonance frequency of the rotor blade. In this investigation
special attention is paid in monitoring the stiffness degradation of the specimens.

6.2 Materials and static properties


Glass/polyester coupons
For this investigation coupon specimens were cut on-axis (0 degrees) from a
0 degree and 45 degree reinforced glassfibrepolyester plate (8.5 mm thick).
The plate contained seven layers of 500 g/m2 UD and eight layers of 480 g/m2
45 degree glass fabric in a symmetrical lay-up (Tables 6.1 and 6.2). The resin
was an iso-polymer Synolite 593-E-l. The plate was produced under industrial
conditions by hand lay-up and with a glass/polyester ratio of 50 %wt (30
%vol).
The load introduction to fibre reinforced composites is very critical due to
the multi-axial stress distribution and stress concentration in the grips. For a
smooth load introduction the specimens were machined in a dogbone shape
(Fig. 6.1). Sims and Gladman (2) observed that there is only a slight difference

Glass and Hybrid Fibre Performance


Table 6.2
Material
polyester +
Glass coupon
8.5 mm thick
Glass/Carbon
coupon
8.5 mm thick
Glass bolted
15 mm thick
Glass/Carbon
bolted
15 mm thick

fibre

91

Material layup and mechanical properties

Layup, symmetrical
about midlayer/s
45/0/45/0/45/0/
45/^0/s
45/0/4545c/2p0c/0/

Test speed 'E'modulus


(mm/sec) G( Pa)
2
0.02
2

UTS/
(MPa)/%

UCS/
(MPa)/%

17.8/18.1*

424

310

26.1

370/2.6
367/1.8

286/2.0
251/1.0

454/?0/s
+45/45/3p0/+45/45/l/?0/s

0.25

490

2p45/45c/0/0c/0/45c/
45/0/2p0c/^p0/s

545

Notes: (i) * indicates compression


(ii) fibre volume fraction typically 32 percent
(iii) layers glass unless designated c(carbon)

Delamination

Fig. 6.1

Fibre reinforced polyester coupon specimen and schematical illustration of the damage
development. Dimensions in mm

92

Design of Composite Structures Against Fatigue

between the fatigue results of specimens with parallel and continuous waists,
with the parallel specimens being slightly stronger. On an ultimate strength
normalized basis the fatigue curves were essentially identical.
For laminates containing angle-plies it is often observed that damage is
initiated at the free edges. This will result in conservative fatigue data, because
damage may not occur as early in edge-free components or in components
where the edges are at a lower stress level.
The static mechanical properties of the laminates were measured with a
speed of 1 mm/min (3). The measurements show an Ultimate Tensile Strength
(UTS) of 365 MPa (emax = 2.38 percent) and an Ultimate Compression
Strength (UCS) of -395 MPa (emax = -2.64 percent). The 'E' modulus of the
material was calculated after measurement of the strain with strain gauges. The
initial 'E' modulus (Fig. 0.1) was 17.8 GPa (18.1 GPa in compression) which is
somewhat lower than the value of 18.5 GPa calculated from laminate theory.
The ultimate strengths values of the specimens were initially measured with a
speed of 0.02 mm/sec, which gave a loading rate of about 2.85 MPa/sec. The
UTS was 370 MPa (emax = 2.6 percent) and the UCS was -286 MPa (emax =
-2.0 percent). The tensile strength of glass fibre reinforced plastics, however,
showed a significant loading rate sensitivity (4)(5). The low loading rate used
gave a conservative value of the tensile strength.
For normalizing the fatigue data with the ultimate strength, UTS/UCS
values obtained with about the same rate of stress application were used. With
a loading rate of 285 MPa/sec (2 mm/sec), which is comparable to the loading
rate in the fatigue test, the measured UTS of the specimen was 424 MPa and the
UCS was -310 MPa. This was an increase in strength with about 7.5 percent
per order of magnitude of the loading rate (Table 6.2), which has also been
observed by others (5).
Glass/carbon-polyester coupons
For the investigation into the possible benefit of hybrid fibres, coupon specimens were cut on-axis from a 0 degree and 45 degree reinforced glass/30
percent carbonfibrepolyester plate. The plate containedfivelayers of 500 g/m2
UD, six layers of 480 g/m2 45 degree glass fibre, four layers of 200 g/m2 UD,
and two layers of 430 g/m2 45 degree carbon fibres in a symmetrical lay-up
(Tables 6.1 and 6.2). The plate was produced under industrial condition by
hand lay-up and with afibre/polyesterratio of 50% wt (~32% vol).
For a smooth load introduction the specimens were machined in a dogbone
shape. The specimens had the same dimensions and resin as the plain glass
specimens. The ultimate strengths values of the specimens were measured at a
speed of 2 mm/sec, which gave a loading rate of about 82 kN/sec which is
comparable to the rate in the fatigue tests. The UTS was 367 MPa (emax = 1.8
percent) and the UCS was -251 MPa (emax = -1.0 percent).
The 'E' modulus of the material was calculated after measurement of the

Glass and Hybrid Fibre Performance

93

strain with strain gauges. The initial 'E' modulus was 26.1 GPa which was
somewhat higher than the calculated value of 24.2 GPa.
Glassicarbon-polyester bolted joints
To investigate the influence of hybrid fibres on the fatigue performance of
structural details, the same type of joints as the plain glass/polyester bolted
specimens (see Chapter 11) were fabricated. Specimens were cut on-axis from
a 0 degree and 45 degree reinforced glass-/31 percent carbon fibre polyester
plate (15 mm thick). The plate contained a combination glass fabric of seven
layers of 700 g/m2 UD fibres plus 100 g/m2 CSM, six layers of 700 g/m2 + or
-45 glass fibres, eight layers of 200 g/m2 UD and four layers of 430 g/m2 45
carbon fibres in a symmetrical lay-up (Tables 6.1 and 6.2). The plate was
produced under industrial conditions by hand lay-up and with a glass-polyester
ratio of 50% wt ( - 3 2 v/o fibres).
The ultimate strength of the bolted plate was measured with a loading rate of
about 200 kN/sec. The UTS was 131 kN per bolt hole, which gave a bearing
stress of 545 MPa in the holes. The specimen geometry is shown in Fig. 11.1.
The bearing strength of bolted holes of glass/carbon hybrid composites
varies with the carbon fibre content, with a minimum at 20 percent carbon
fibres (6). The carbon fibre content of the investigated plate is above this
minimum and it was observed that the static bearing strength of the hybrid
laminate was better than of the plain glass reinforced plate (Table 6.2).
However, the loading rates differed significantly, resulting in a possible
conservative value for the plain glass plate.
Fatigue test method
The constant amplitude tests were performed on load-controlled servohydraulic testing machines (Instron 1342 and Instron 1343) with tensiontension or tension-compression loading (stress ratio R = omin/omux = 0 . 1 , - 1 ,
or 0.5). During fatigue tests the GFRP specimens experienced heating due to
hysteresis, which depends on the stress amplitude and the frequency. Due to
the duration of the tests, a frequency up to 15 Hz was used, the temperature of
the specimen being kept low by a cooled air flow of 15-18C.
In the steady dynamic state there was a temperature difference between the
centre and the surface of the specimen of about 7C. This implied that the
average temperature of the specimens was about 20C.
After measurement of the initial stiffness, the stiffness degradation was
determined on-line during the test at regular intervals (100-10 000 cycles). The
stiffness was calculated as the difference in load divided by the difference in
displacement of the specimen, as measured by the displacement transducer of
the actuator. The stiffness degradation of the specimens influenced the load
performance in the closed-loop servo-hydraulic system. In order to have a

Design of Composite Structures Against Fatigue

94

03
Fig. 6.2

0.4
0.5
0.6
Normalised lifetime (N/N,)

0.7

Typical stiffness degradation plot of GFRP coupon specimens, (a) R = 0.1 ; (h) R = 1.
Data has been normalized by the initial stiffness and the number of cycles to failure

constant load until failure the load was corrected on-line during the fatigue life
of a specimen.
Loading of composites with stress ratio R = - 1 is, relative to the magnitude
of the stress, the most severe constant amplitude test. In addition to tension
and compression, the specimens can also be subjected to buckling. The design
of the specimens are such that they are stable to Euler buckling. Anti-buckling
guides, however, were not applied as recommended by Schtz because these
can also restrain the development of micro-buckling damage and may cause
extra heating of the specimen (7). On a local scale, micro-buckling of the fibres
can occur due to misalignment of the 0 degree reinforcement or small voids in
the resin.
6.3 Fatigue of glassfibrelaminates
Glass-polyester coupons
In Fig. 6.2 the typical stiffness degradation S/S0 of a specimen loaded with a
stress ratio R = 0.1 is shown. The development of the fatigue damage observed
in the specimens proceeds gradually and is consistent with the measured
stiffness degradation. In an early stage of the fatigue life (0-10 percent) the
damage starts with cracks in the polyester perpendicular to the load, located on
the stitch bond lines of the glass fibres (Fig. 6.1). This causes a very small
reduction in the stiffness, because the contribution of the polyester to the initial
stiffness is only about 10 percent in that direction.
At a later stage (10-70 percent) interfacial shear damage occurs and the first
delaminations appear in the 45 degree layers. From a laminate calculation it is
expected that the first-ply failure will be in the 45 degree layers. The damage

Glass and Hybrid Fibre Performance

95
Fracture

IV
Fig. 6.3

\V

101
Cycles

IV

IO*

IO'

IO"

Initial strain versus cycles of GFRP coupons (0/45 degrees) tested at R = 0.1 and R = 1

initiates at the edge about 15 mm from the middle. The delaminations expand
in the last stage across the whole specimen until finai fracture occurs. In Fig.
6.1(c) the damage development is schematically illustrated in four stages.
A typical stiffness degradation plot of a test with a stress ratio R = 1 is also
depicted in Fig. 6.2. The stiffness degradation in the first 20 percent of the
fatigue life is quite strong and is probably due to microbuckling effects. After
this initial stage interfacial shear damage occurs and small delaminations start
to grow and expand in the last stage across the specimen (see Fig. 6.1).
In order to obtain fatigue endurance curves ( or ) tests up to 107108
cycles were conducted. In Fig. 6.3 the number of cycles until fracture of the
specimens versus the maximum initial strain level is shown for R = 0.1 and
R = l.
The results of these constant amplitude tests can be compared with an
empirical relation for the fatigue strength of glass fibre reinforced plastics.
Appel and Olthoff (8) modified the Mandell (9) relationship and the degra
dation factor D after analysing data from literature (Chapter 4). These
modified curves are shown in Fig. 6.4 on a maximum stress basis. The
Appel/Olthoff relationship predicts fatigue lifetimes which are close to the data
points for the R = 0.1 and R = 1 tests.
From these fatigue tests, constant life diagram (Goodman diagram) can be
constructed for lifetimes from 102 to 108 cycles by interpolation of the fatigue
data. The result is shown in Fig. 4.5. The Goodman lines are bending
downward for the high cycle range, which is unusual. The same behaviour is,
however, also observed in a similar 0 degree and 45 degree glass fibre

Design of Composite Structures Against Fatigue

96
500

Fig. 6.4

Fatigue results (stresscycles) of the GFRP coupons (0/45 degrees) on maximum stress
basis. The AppelOlthoff predictions AO are indicated with lines

10.0 _

--

^"5$<^^

*&>

s 1-0

_
-

"*^

****^">w

-"

v^* -

*****

O R = 0.1
= -l

0.1

Fig 6.5

unni

10'

10'

IV

IV
Cycles

IV

10'

10'

10'

Initial strain versus cycles of GFRP coupon (0/45 degrees) for 10 percent reduction in
stiffness

Glass and Hybrid Fibre Performance


Table 6.3

97

Failure strain in percent at 107 cycles for the GFRP


laminates

Loading

Fracture criterion

10 percent stiffness
reduction criterion

0.75
0.50

0.71
0.42

Tensile = 0.1
Reverse/?=-l
Note: from Figs 6.3 and 6.5

reinforced epoxy laminate (10). It implies that the tension-compression fatigue


behaviour of this material is better than normal.
These non-woven glass fabrics are, according to Mandell, in the higher stress
range more fatigue resistant than woven fabric (11). The fatigue of the latter
material is affected by local friction and micro-delamination near the cross
over point. Only at low stress levels near a possible fatigue limit will the fatigue
life of woven material approximate to the fatigue performance of laminates
with non-woven reinforcement.
Failure criteria
The choice of a failure criterion is, besides fracture, a rather arbitrary matter.
The fatigue damage develops rather gradually and can be represented by the
stiffness degradation.
In a survey of the long-term durability of fibre glass reinforced plastic
structures at NASA, it was found that the static strength could be retained for a
considerable period of time under real service conditions (12). Structures
which were continuously exposed to ultraviolet radiation and weathering
required appropriate surface protection. The stiffness modulus decreased by
about 10 percent in a period of about twenty years. Such a reduction in stiffness
of 10 percent can be used as a failure criterion for rotor blades with a design life
of approximately twenty-five years. In Fig. 6.5 the number of cycles until 10
percent stiffness reduction is shown. The results are compared in Table 6.3. At
the lowest applied strain levels (0.4 percent) there is still fatigue failure
observed at 3.107 cycles. The slope of the - or - curves is continuous on
log-log curves, which implies that there is still no indication of a fatigue limit in
this high cycle range.
The observed damage mechanism is consistent with the conceptual frame
work for the interpretation of fatigue damage by Talreja (13). In this model the
fatigue limit for unidirectional reinforced composites is correlated with the
fatigue limit of the matrix material. For intermediatefibreorientations (such as
45 degrees) a mixed-mode matrix and interfacial damage occurs.
For metallic materials many cumulative damage theories have been devel
oped which, except for Miner's linear damage theory (Chapter 4), are based on

98

Design of Composite Structures Against Fatigue


3.0
O O 4*
OOOO
fltt

1.0

*"

*.*4___ = . I 0 V
^~^

~ - ^ - / f = 0.1 ~
-R = -l

O 0 45 degrees X = 0.1
0 45 degrees = 1
0 degrees
45 degrees

0.1

11 mul
10

102

IV

IV

IV

IV

10'

10

Cycles
Fig. 6.6

Initial strain versus cycles of GFRP coupon (0, 4 5 degrees) at fracture. Unidirectional (0)
and + 4 5 data from (16) and (17)

a crystalline multigrain microstructure and thus are not applicable to the


fatigue behaviour of glass fibre reinforced plastics.
The fatigue behaviour of fibre reinforced plastics is different from that of
metallic materials where most of the time is spent on developing the first crack.
In the case of fibre reinforced plastic laminates a first crack can initiate in the
first load cycle. The crack, however, is confined to the ply in which it is initiated
for the most of the fatigue life. A reduction in stiffness and in strength will only
be measured after the formation of many, usually dispersed, small cracks
(Chapter 3).
Because of the statistical nature of the fatigue behaviour of composite
materials, the research on the cumulative damage models is directed to a
statistical approach of the Miner theory (14)(15).
Discussion
Fatigue lifetime data of fibre reinforced plastics can follow a variety of trends.
The most frequently observed are linlog or loglog relations, which show
large differences for low amplitudes (Chapter 3). In Fig. 6.6 the constant
amplitude data for ? = 0.1 and R = 1 are presented in a log-log diagram. The
data fit better to a line on a log-log than on a lin-log basis (Fig. 6.4), so it is
sensible to represent these data with a power law type of equation.
The best way to compare the properties of fibre composites is on a strain
basis, because different laminates can be compared without knowing the
dimensions and modulus data, and the strain in a laminate is, contrary to the
stress, the same in the diverse layers. In Fig. 6.6 the data of previous fatigue
tests on unidirectional and on 45 degree glass-polyester laminates are also
indicated (16)(17). This shows that the data of the present laminate fits on the

Glass and Hybrid Fibre Performance


1.0
0.8

6
.

.. __

_,

04
c

oco?.

0.4 &

0.2

'S,

O
0/ 45 degrees R =0.1

0/ 4 5 degrees R=-l
0 degrees
45 degrees

1
10

10

IV

(a)

IV

10'

10'

10'

10"

Cycles
1.0

0.8

"

0.6

o <
ooo.
..

^.

*~

O 0/ 45 degrees R =0.1
0 degrees
45 degrees

0.2

(b)
Fig. 6.7

ml
10

10'

IV

IV
Cycles

I
IV

..

.
o

I
10'

10'

10"

Stress normalized comparison of 0 degrees, 45 degrees and 0 degrees/45 degrees GFRP


coupon results as a function of cycles, (a) Fracture; (b) 10 percent reduction in stiffness

curves of the unidirectional material. From some preliminary results it was


observed that the data of several materials fit on a master curve if there are
enough unidirectional fibres in the layup (18). This is considered further in
Chapter 14. Carbon and glass/carbonhybrid laminates show an analogous
response to fatigue. In terms of strainlife comparisons all results fall within a
single scatter band (19).
The overall stress normalized by the UTS or UCS also gives a relative basis
for comparison, if ultimate strength values obtained with about the same
loading rate as in the fatigue tests are available. As is shown on a normalized
stress basis in Fig. 6.7, the fatigue results of the tested laminate are also
close to the data of UD and of 45 degree glass fibre reinforced polyester
(16)(17)(20).
It can be noted that the fatigue lifetime for fracture of the laminate is

100

Design of Composite Structures Against Fatigue

predicted by the UD data, but also by the diagonal data. This is, however,
probably a coincidence. The fatigue data of the diagonally plied material were
obtained in load controlled tests which showed a steep decrease in stiffness,
while the fatigue loading in the layers in a 0/45 degree laminate is governed by
the overall strain and the reduction in stiffness of the most fatigue resistant
layers, i.e., the UD lamellae (refer similar conclusion in Chapter 2).
This effect becomes clear when the data for 10 percent stiffness reduction are
plotted (Fig. 6.7(b)), showing a large reduction for the diagonal data. The 45
degrees data are also flattered by normalizing, because only a conservative
(low stress rate) UTS was tested. So both on a strain and on a normalized stress
basis the fatigue behaviour of the 0/45 degree fibre reinforced coupons is
dictated by the unidirectional layers.
6.4 Fatigue of hybridfibrelaminates
Glassi carbon!polyester coupons
A hybrid composite of two types of fibres makes it possible to adapt the
laminate performance for special material requirements. In the case of multi
directional laminates, damage will start in the off-axis plies (1). GFRP
composites seem to be more susceptible to this damage than CFRP and hybrid
laminates. The strain-life curves reflect this behaviour, indicating a positive
hybrid effect of the hybrid composite. This effect is also present when the
fatigue ratio (fatigue stress for a certain life divided by the tensile strength) is
considered.
Damage accumulation
The development of tensile (R = 0.1) fatigue damage observed in the
specimens proceeds gradually, and is consistent with the measured stiffness
degradation. As with the plain glass material, the damage starts with cracks in
the matrix perpendicular to the load located on the stitch bond lines of the glass
fibres. In a later stage of the fatigue life (10-70 percent) interfacial shear
damage occurs and thefirstdelaminations appear. As in Fig. 6.1(c) the damage
initiates at the edge about 15 mm from the middle and the delaminations
expand in the last stage across the whole specimen until final fracture occurs.
Also the stiffness degradation of the R = 1 tests is comparable to the plain
glass material (Fig. 6.2(b)) and is quite large in thefirst20 percent of the fatigue
life (due to micro-buckling effects). After this initial stage, interfacial shear
damage occurs in the 45 degree carbon layers and small delaminations start to
grow and expand in the last stage across the specimen. In the final stage the
specimen splits within the carbon layers and fails by buckling.
Fatigue resistance
In order to compare the fatigue behaviour of the hybrid with the glass coupons,
tests up to 107 cycles were conducted. Due to the large difference in '

101

Glass and Hybrid Fibre Performance

1.0
0.8

a 0.6

E
3

0.4
O = 0.1
= l Glass
fi = 0.1
Glass/carbon
+ K = l
0.2

Fig. 6.8

10

IV

IV

IV
Cycles

10s

10'

10'

10"

Comparison of glass G and hybrid GC coupons on a normalized stress basis as a function of


cycles for R = 0.1 and R = 1

modulus and straintofailure of carbon and glass fibres, the results of the
constant amplitude fatigue tests can only be compared with the plain glass
material on a normalized stress basis or with load carrying capacity of
equivalent layups.
In Fig. 6.8 the fatigue data are plotted on a normalized stress diagram. As
can be expected for reinforcement with carbon fibres, the fatigue behaviour is
improved for pure tensile loading (R = 0A). This is also true for the lower stress
range at R = 1.
The stress is obviously below the fatigue critical stress of the carbon layers. A
longer life can then be expected from a hybrid than from a GRFP carrying the
same load, because the stress in the glass will be lower. The slope of the fatigue
curve is lower for the hybrid due to the incorporation of the carbonfibresin the
laminate, which carry a high fraction of the load.
The results for reverse loading (R = 1) are, however, rather disappointing
in the higher stress range. Probably the stress is beyond the fatiguecritical
stress of the carbon fibres, leading to an early failure of the carbon and hence
the composite. This is also indicated by the low compression strength of this
material (Table 6.2).
One possible explanation is that the bonding of the carbon fibres to the
polyester was poor. An inquiry into the effects of fibre sizing on the bonding
characteristics of the carbonpolyester system was made (21). From this work
one could conclude that the ultimate stresses as found in the coupon tests were
correct. No effect of sizing of the carbon with epoxy is known to the contacted
manufacturers. (Glass fibres can also be sized with epoxy and there are no
bonding problems.) Laminate calculations showed that thefirstplyfailure is in
the 45 degree carbon layers.
Comparing the load carrying capacity of the laminates with equivalent lay

102

Design of Composite Structures Against Fatigue

Fig. 6.9

Comparison of glass G and hybrid GC coupons on a maximum stress basis for R = 0.1
and = - 1

400

a.
S
r 3001

+_

E 200

O Glass
+ Glass/carbon
100

101

10
WISPER sequences

Fig. 6.10

10J

Comparison of glass G and hybrid GC coupons on a stress basis using the WISPER
spectrum

up, fibre fraction, and thickness gives a quite different view on the results, as is
shown in Fig. 6.9. The reverse stress fatigue, in particular, is worse on this basis
for the hybrid material. This is due to the rather low ultimate strength values of
the hybrid (Table 6.2) resulting in more optimistic fatigue data on a normalized
basis. One should, therefore be very careful when comparing data normaled on
a stress basis.
Some specimens were tested with the WISPER spectrum (Chapter 4). The
results are compared with the plain glass coupons on a stress basis (Fig. 6.10).
There is only some improvement with the substitution of carbon for glass as can
be expected from the constant amplitude data (Fig. 6.9) for a spectrum loading
with a majority of tension-tension cycles.
In designing turbine blades the load carrying capacity in fatigue will be

Glass and Hybrid Fibre Performance

103

400

300

B.

200
Glass

Glass/carbon
= 0.1 Glass
R = 0.5 ' a S S

'

100

x = 0.1 Glass/carbon
+ = 0.5

10
Fig. 6.11

10:

10'

IV
Cycles

10 !

10'

10'

10'

Comparison of glass G and hybrid GC bolted joints on a stress basis. Criterion fracture

decisive. In this way one can conclude that there is hardly any benefit in
incorporating a modest fraction of carbon fibres in glass fibre reinforced
polyester.
6.5 Fatigue of glass/carbon bolted joints
Fatigue tests with R = 0.1, R = 0.5, and R = 0.615(WISPER) were
conducted up to the high cycle range with bearing stresses sometimes equival
ent to the plain glass joints. The R ratio 0.5 was chosen because it was closer
to the real stressfluctuationsin a blade root than R = 1 . The fatigue damage
was similar to that observed in the plain glass bolted joints (see Chapter 11).
In designing bolted joints for turbine blades the stress bearing capacity of the
holes in the laminate is decisive (22). In Fig. 6.11 the number of cycles until
fracture of the specimens versus the maximum initial bearing stress level is
shown for the R = 0.1 and R = 0.5 tests, and compared to the plain glass
results.
The data show some, but not significantly improved fatigue behaviour in the
R = 0.1 and R = 0.5 tests. Although only a few data are available the relative
insensitivity to the stress level in the R = 0.5 tests is remarkable. The data for
the fatigue life at 10 percent stiffness reduction show the same trend as for
fracture.
Also for the WISPER or WISPERX tests only some improved fatigue life is
observed (Fig. 6.12). So a positive hybrid effect for fatigue may not exist for
this glass/carbon material in the bolted joint, depending upon the loading that
the joint is subjected to.
6.6 Conclusions
The analysis of the constant amplitude tests on a 0/45 degree GFRP laminate
showed that there was no indication of a fatigue limit and that the data fitted

Design of Composite Structures Against Fatigue

104
500
a.

+ ..

' 400

E
I 300

'I

^""^o

"""

O Glass
+ Glass/carbon

S
200

,,,

IO1

10'

WISPER sequences

Fig. 6.12

Comparison of glass G and hybrid GC bolted joints on a stress basis using the WISPER
spectrum

better to a line on a loglog than on a linlog basis. It was also observed that
both on a strain and on a normalized stress basis the fatigue behaviour was
dictated by the unidirectional layers. For normalizing the fatigue data on a
stress basis, ultimate strength values obtained with a similar rate of stress
application as in the fatigue tests should be used.
A hybrid composite prepared by combining glass and carbonfibrescan have
mechanical properties which deviate from the linear summation of the contri
bution of the relative volume fractions of the fibres (ruleofmixtures). A real
positive hybrid effect was not observed in the fatigue investigation on laminates
and bolted joints where the results were compared with equivalent plain glass
specimens. Especially, the fatigue performance with the stress ratio R = 1
was rather disappointing. In designing turbine blades the load carrying capacity
in fatigue will be decisive. In this way one can conclude that there is hardly any
benefit in incorporating a modest fraction of carbon fibres in glass fibre
reinforced polyester.
References
(1) OLDERSMA, ., 1991, A literature survey on the fatigue behaviour of hybrid composites,
NLR TP 91345 L.
(2) SIMS, G. D. and GLADMAN, D. G., 1980, Effect oftest conditions on the fatigue strength
of a glassfibre laminate; Part B Specimen condition, Plastics and rubber, Materials and
applications, p. 122.
(3) LAMERIS, J., 1991, Statische eigenschappen van het standaard laminaat gebruikt in het
nationaal onderzoekprogramma 'Vermoeiingsonderzoek vezelversterkte kunststoffen voor
windturbinerotoren' NLR CR 91479 L.
(4) JONES, C. J. tal., 1983, Environmental fatigue of reinforced plastic, Composites, 14,288.
(5) B ACH, P. W., 1992, Fatigue properties of glass andglass/carbonpolyester composites for
wind turbines, ECNC92072.
(6) MATTHEWS, F. L. et al., 1982, The bolt bearing strength of glasscarbon hybrid composites,
Composites, 13, 225.
(7) METONDANG, T. H. and SCHTZ, D., 1981, The influence of antibuckling guides on the
compression fatigue behaviour of carbon fibre reinforced laminates, LFB report.

Glass and Hybrid Fibre Performance

105

(8) APPEL, N. and OLTHOFF, J, 1988, Voorontwerpstudie NEWECS-45, Polymarin rapport.


(9) MANDELL, J. F., 1982, Development in reinforced plastics-2, (edited by) O. Pritchard,
(Appi. Science Publishers) p. 67.
10) KENSCHE, Ch. W., 1992, High cycle fatigue of glass fibre reinforced epoxy materials for
wind turbines, DLR Final Report EC-contract EN3W-0041-D.
11) MANDELL, J. F., 1975, Composite reliability, ASTM STP 580, (ASTM, Philadelphia,
USA) p. 515.
12) LIEBLEIN, S., 1981, Survey of long term durability of fiberglass reinforced plastic structures, Report DOE/NAS-A/90599-1, NASA CR-165320 TRS 106.
13) TALREJA, R., 1987, Fatigue of composite materials, (Technomic Publishers).
14) SENDECKYJ, G. P., 1991, Fatigue of composite materials, (edited by K. L. Reifsnider)
(Elsevier) p. 431.
15) YANG, J. N. and JONES, D. L., 1981, Load sequence effects on the fatigue of unnotched
composite materials Fatigue of fibrous composite materials, ASTM-STP 723, ASTM, Philadelphia, USA, p. 213.
16) BACH, P. W., 1988, High cycle fatigue investigation into wind turbine materials, Proc.
EWEC, (edited by W. Palz), (Stevens & Ass., England), p. 337.
17) BACH, P. W., 1989, High cycle fatigue of glass fibre reinforced polyester, Proceedings of the
IEA Workshop on Fatigue in Wind Turbines, (edited by K. F. McAnulty), ESTU-N-113.
18) MAYER, R. M., 1992, Fatigue properties and design of wing blades, Proceedings of the
Wind Energy Contractors Meeting, Alghero, EUR 14607, p. 241.
19) DICKSON, R. F. etal., 1989, Fatigue behaviour of hybrid composites; Part 2 - Carbon-glass
hybrids, J. Mater. Sci., 24,227.
20) BACH, P. W., 1991, High cycle fatigue testing of glass fibre reinforced polyester and welded
structural details, ECN-C-91-010.
21) WAGENDORP, M. H. P., 1991, Hybrid fabric interlaminar bonding, Polymarin, rapport
N/21342-002.
22) GODWIN, E. W. and MATTHEWS, F. L., 1980, A review of the strength of joints in fibre
reinforced plastics, Composites, 11, 155.

CHAPTER 7

Fatigue Properties of Wood Composites


M. Ansel!, I. Bond, P. Bonfield and C. Hacker*
Wood composites are an excellent design solution for the construction of wind turbine blades,
since they are light, notch-insensitive, and inexpensive. S-N data and constant life lines have
been generated for Khaya and a successful fatigue life prediction analysis has been developed
and applied to the complex reference wind spectrum WISPERX. Khaya shows no size effect in
axial tension-compression fatigue. Several alternative wood species, including beech, poplar,
Baltic pine, and birch have been assessed in fatigue, and beech and poplar are good alternatives
to Khaya. Joints and joint quality substantially influence the fatigue life of wood composites.
Infra-red thermography is an effective method for following damage accumulation around joints
throughout the course of a fatigue test.

7.1 Introduction
In the last ten years there has been intense interest in the development of
commercial wind turbines triggered by the construction of wind farms in
California (1)(2). Various types of wind turbine have been built (1)(3)
including horizontal and vertical axis designs. Most commercial machines are
now of a standardized horizontal axis format, Fig. 7.1, with two or three blades
feeding power to a generator set mounted on top of a concrete or steel tower.
The majority of wind turbine blades are constructed from glass-reinforced
plastic (GRP), steel, and wood composites. Commercial wind turbines are
generally powered by either GRP blades or wood composite blades. The Wind
Energy Group (WEG) construct the wood composite blades for their 33 m
diameter MS-3 turbines (4) at UK and Californian windfarms from laminated
Khaya ivorensis (African mahogany) and Westinghouse use Douglas fir wood
composite for their 142 ft diameter WWG-0640 turbines at Kahuku, Hawaii
(5).
7.2 Manufacture
An exploded diagram of the structure of WEG's MS-3 turbine blade is
presented in Fig. 7.2. The wood composite leading edge, the main structural
element of the blade, is fabricated from wood veneer (6) with a moisture
content of approximately 8 percent. Any growth defects are spread out over
the blade volume, although major defects are selected out at the preproduction stage. The trailing edge is a foam sandwich and the complete
'University of Bath, UK.

Design of Composite Structures Against Fatigue

108

Nacelle

Hub

Blades

Tower

Fig. 7.1

Horizontal axis wind turbine

structure is protected by a GRP skin. The blade is moulded in two halves which
are subsequently bonded together in conjunction with a GRP shear web.
Each half of the blade is laid up by coating the pre-cut sheets of veneer with
room temperature cure epoxy resin and laying them into the mould. When the
final lay-up has been achieved the veneer is covered by a plastic membrane and
air is evacuated from the mould. The resulting vacuum is sufficient to press the
veneer into the mould and the resin is allowed to cure.
Blades are attached to the hub of the wind turbine via conical steel studs
which are bonded directly into the blade root using a carbonfibre-filledepoxy
grout. The conical form of the studs alleviates the inevitable stress concentrations at the interface between the high modulus steel and the low modulus
epoxy. The wood veneer is 4 mm thick rotary-cut Khaya ivorensis, but a
number of alternative species, including European softwoods and hardwoods,
are being evaluated to diversify the sources of supply.
7.3 Advantages
The specific mechanical properties of laminated Khaya are compared with
those of other engineering materials in Table 7.1.
At first sight the specific properties of wood appear to offer no obvious

109

Fatigue Properties of Wood Composites

GRP skin
Foam

'Epoxy grout
Fig. 7.2

Table 7.1

Material
High tensile steel
High strength Al alloy
Khaya/epoxy
Glass/polyester
Carbon/epoxy

The structure of a WEG wood composite blade

Mechanical properties of materials for wind turbine blades


Specific
gravity
gravity

Tensile
strength
(GPa)

Strength/
s-g
(MPa)

Tensile
modulus
(GPa)

Modulus/
s-g
(GPa)

7.8
2.8
0.5
2.0
1.5

1.55
0.40
0.082
1.1
1.4

199
143
164
550
940

210
70
12
36
130

26.9
25.0
24.0
18.0
86.6

Note: Glass/polyester and carbon/epoxy are typical laminates with 50% vol fraction of UD fibres.

advantages over the other candidate materials for 'turbine blade manufacture.
In fact tensile strength is not the main requirement. The main requirements are
fatigue strength, particularly of the joints, and buckling resistance in compression. The welded construction of steel blades results in a very heavy structure.
This is because welded joints have very low stress allowables in fatigue, and
fatigue loads are generated by the self weight of the blades. For example, a
steel blade for a 20 m diameter turbine weighs 1300 kg, compared with 450 kg
for a 25 m blade in laminated wood. Half the weight is in the laminated wood

110

Design of Composite Structures Against Fatigue

Table 7.2 Test methodology


Shape
Size
Loading
Control
Frequency
Environmental conditioning

Profiled dumbell
30 mm wide x 12 mm thick
36 mm wide x 40 mm thick
Axial, constant amplitude, and
complex
Load
Constant rate of stress application
150 and 200 MPa/s
65% RH at room temperature

and half in the resin, studs, and GRP components. Wood veneer and foam can
be placed in the mould for one side of a 16 m blade in one hour prior to vacuum
bagging.
Wood composite blades are two thirds of the weight of the best GRP (glass/
polyester) blades. Furthermore the range offibre-reinforcedcomposites available is vast, varying in price from 1.50/kg to more than 50/kg, compared with
2.75/kg for wood/epoxy laminates. Each fibre-reinforced composite system
needs to be evaluated individually in fatigue, whereas the fatigue life of
different species of wood is less variable and will depend mainly on the
cellulose content reflected by the density.
The mechanical properties of carbon fibre-reinforced plastics (CFRP) in
Table 7.1 appear to offer an attractive alternative to steel, GRP, and wood
composites. However, the high specific stiffness is not an advantage in a large
aerofoil structure. The ratio of the blade wall thickness to the chord length
(leading edge to trailing edge distance) is between 2 and 4 percent for a wooden
wind turbine blade. Below 2 percent Euler buckling is likely to occur so there is
little incentive to increase the stiffness, and hence reduce the thickness, of the
blade wall by using CFRP even if cost is of little concern.
Further detailed information on materials selection for wind turbine blades
is provided by Bonfield (7) and the virtues of American wood/epoxy blades are
discussed by Davidson (8).
7.4 Constant amplitude data
Wood composite turbine blades are large hollow structures and during rotation
each side of the blade is subjected to predominantly tensile or compressive
loads. High-capacity servohydraulic fatigue machines (20 tonnes) are used to
apply axial tension/compression loads to large profiled specimens which are
sufficiently thick to resist buckling in compression (Table 7.2). Tests are
sinusoidal and load-controlled with a constant rate of stress application of
200MPa.s _I .
Specimens are waisted with a central section of constant cross-sectional area.
Two specimen geometries have been used to evaluate size effect in fatigue. The

Fatigue Properties of Wood Composites


60

= 10

10
-0.1

Fig. 7.3

10

IV

IV
IV
Cycles

IV

IV

IV

10'

S-N characteristics for Khaya ivorensis/epoxy laminate loaded in axial tension/


compression at R = 1, 2, 10, and 10 at a moisture content of 10 percent

first (standard) has three veneer layers plus a 0.5 mm glass layer on either face
and is 12 mm deep by 30 mm wide; the second (large) is 40mm deep (10 veneers
plus a glass layer on either face) by 36 mm wide. Tests are carried out within an
enclosure containing a saturated sodium nitrite solution to maintain a relative
humidity of approximately 65 percent.
Constant stress amplitude tests have been performed at various R ratios (9)
(refer Units discussion in the introductory chapter).
S-N curves (maximum stress versus number of cycles to failure) at R ratios of
1 (reversed loading), 2, 10 (compression-tension) and 10 (compressioncompression) are presented in Fig. 7.3. Wood is seen to have a highly variable
fatigue life in common with other materials. Log-linear regression lines have
been fitted to results plotted as log N-S data because N is the dependent
variable and 5 is the independent variable. The plot is then inverted into the
conventional S-log N format. The fatigue data passes through the mean static
compressive strength (mean 49.4MPa) at one quarter of a cycle (log10 N =
0.6). The fatigue and static points exhibit considerable scatter and there is a
possible sigmoidal form to some of the curves which suggests the existence of a
fatigue limit.
At R = 1 the fatigue life is lowest at any given maximum stress level. The
life increases as the tensile component of the reversed load becomes less
(passing from R = 2 to -10) until the load is all compressive (R = 10) where
there is noticeably less scatter, and damage rates are low. In tension-tension at
R = 0.1 (Fig. 7.4) the straight line fit to the fatigue points passes through the
mean static tensile strength (80.0 MPa) and a 5 percent prediction boundary is
included, below which there is a less than 5 percent probability of failure.

112

Design of Composite Structures Against Fatigue

90
^ 5 0 % probability

70

1^
****>.

7
50 _

~~J3

"*^

95% confidence

Cl_

^\s

~" *

D
>s

^^^*s^^
D

"S.

30

Ss
s*

10 . .
0.1

Fig. 7.4

10

IO1

l
IV
IV
Cycles

10s

10'

10'

Ss

S*

10

S-N characteristic for Khaya ivorensis/epoxy laminate loaded in axial tension/


compression at R = 0.1 at a moisture content of 10 percent

Constant life lines


SN fatigue data from Figs 7.3 and 7.4 are used to construct a constant life
diagram (Fig. 7.5) for mean fatigue lives of 105,106, and 107 cycles. Alternating
stress o"a|t is plotted versus mean stress amean and, for each life, the constant life
line is constructed from seven data points obtained from thefiveSN curves at
R = 10, 10, 2 , 1 and 0.1 and the static tensile and compressive strengths
(o"ait = 0) Th e constant life diagram can be split into four regions which,
passing clockwise around the diagram, represent compressioncompression (R
= +1 to o), compressiontension (R = to 1), tensioncompression (R
= 1 to 0) and tensiontension (R = 0 to +1).
As the R ratio passes from 10 to +10 the failure mode changes from mixed
mode to compressive and this transition is accompanied by a region of inflexion
in the constant life diagram. At positive R ratios in compressioncompression
(Fig. 7.3) SN curves become almost horizontal and fatigue life is stress
independent. Hence the constant life lines in this region of the constant life
diagram become superimposed. Between R = o and R = 10 the constant life
line is driven upwards temporarily before the decline in the alternating
component of the stress draws the constant life line down to the R = +1
ultimate compressive strength limit.
From the point of view of simple fatigue design, the complex form of the
constant life lines is unfortunate. However, a safe simplification of the lines can
be made by constructing straight Goodman lines between the two R = +1
points and the R = 1 for each life. These straight lines fall below the envelope
of the constant life lines. The form of the constant life lines for wood in tension,

Fatigue Properties of Wood Composites

10 O 10

40

30

0.1

"N.

-m

^v

:^^

^V/\

\SX

rado

10' cycles
O 10'

'

= 20 -

113

10 ,

if

-50

-30

y/
10

10

30

50

l \
70

1
90

Mean stress (MPa)


Fig. 7.5

Constant life diagram for axially fatigued Khaya ivorensis/epoxy laminate for lives of 10s,
106, and 107 cycles

compression, and mixed loading modes are similar to those for fibre compo
sites where failure modes in tension and compression are also, in general,
different (Chapter 4).
Size effect
Theflexuralstrength of statically loaded wood (as well as most other materials)
is influenced by size, so it might be expected that increasing the size of wood
specimens would also have a detrimental effect on the axial fatigue life because
of the increased likelihood of a large surface flaw occurring.
Size effect tests have been carried out on the standard and large wooden
laminate samples to establish whether such a size effect is observed in axial
loading. Large samples are four times the crosssectional.area of the standard
samples. Seventeen large samples have been fatigued to failure at R = 1.
These results are plotted with the R = 1 data for standard specimens in Fig.
7.6. The data points for the large sample are close to those for the standard
samples and, if anything, they lie above rather than below the line fitted to the
standard data points.
The fatigue results reported above suggest that in axial tensioncompression
tests a size effect is not observed. Wood is resistant to propagation of lateral
cracks because of its orthotropic structure, so there is little incentive to perform

Design of Composite Structures Against Fatigue

114
60

^ ^ ^ ^ ^ ^

a 30

a Standard
Large

a. 50

40

OD

"

^ j a s o

*s^^^S

20

10
0.1

1
1
1
l
I
10s
IV
IV
10'
10'
Cycles
Fatigue data for standard and large samples at R = 1

101

10

Fig. 7.6

10'

0.957

I
o.i
Fig. 7.7

L
10

J
10'

IV
IV
Cycles

10'

10'

10'

10'

Interpolated S-N curves for Khaya at many R ratios. These lines have been constructed
from data points at 10\ 10s, 10'', and 107 cycles

fracture mechanicsrelated fatigue tests on wood laminates. The absence of a


size effect in tensioncompression suggests that there is no limit to the size of
wood composites blades for commercial wind turbines.
Deriving SN curves at any R ratio by interpolation
SN curves at many R ratios (Fig. 7.7) for all loading configurations have been
interpolated from the set of constant life lines (Fig. 7.5). The modulus of the

Fatigue Properties of Wood Composites

115

peak stress at each R ratio is plotted versus log (cycles) to failure. At some R
ratios between 0.1 and 1 thefittedlinear curve to the S-N data does not pass
through the origin. At R ratios between +1 (compression) and 1 the curves
intercept the y axis close to or at the compressive strength, whereas between R
0 and +1 (tension) the curves intercept the y axis close to or at the tensile
strength. Mathematical relationships have been developed which predict the
slope and intercept of S-N curves for Khaya at any R ratio and these have been
incorporated into software. By characterizing S-N curves for all R ratios in this
way it becomes unnecessary to resort to over-simplified mathematical relationships to predict the form of S-N curves, based on limited fatigue data.
7.5 Life prediction and the WISPERX spectrum
Turbine blades are subjected to deterministic loads (e.g., gravity, aerodynamic, and centrifugal) and highly variable stochastic loads (e.g., turbulence) which impose cyclic fatigue stresses on the blade material. Wood
composites have been tested to failure when subjected to complex loads
incorporated in standardized wind spectra.
The WISPER spectrum developed by Ten Have as a reference spectrum for
wind turbines has 2.654.105 maximum and minimum turning points and has
been constructed by summing complex flap-wise loads from nine wind turbines. Low amplitude cycles which cause negligible fatigue damage have been
eliminated resulting in the WISPERX spectrum, which is approximately one
tenth of the duration of WISPER (Chapter 4). Large wood composite specimens are subjected to repeated exposure to the WISPERX spectrum, scaled to
a range of peak loads, until failure occurs.
The expected number of cycles to failure is obtained by using a rainflow
counting algorithm to sort the components of the complex spectrum into bins
containing cycles of constant peak stress and R ratio (or constant stress range
and mean stress). The rainflow-counted spectrum is plotted three dimensionally in Fig. 7.8. The tall columns represent the many lower stress range cycles
and it can be seen that the damaging high stress range cycles are in a minority.
A Miner's summation has been performed to predict (10)(11) life based on S-N
data from Fig. 7.7 applied to the rainflow-counted data of Fig. 7.8 as a function
of peak static strength level. (The ratio of the peak tensile stress to the peak
compressive stress in the WISPERX spectrum is by coincidence almost equal
to the ratio of the tensile and compressive strengths of wood.)
Predicted lives, in terms of the number of passes through WISPERX, versus
percentage static strength are plotted in Fig. 7.9 and actual lives from fatigued
Khaya are superimposed. It can be seen that the predictions are conservative
but that measured lives are pleasingly close to the prediction boundary. It can
be concluded that Miner's rule adequately sums the damage caused by each
cycle in the complex spectrum and that accurate characterization of S-N curves
at all R ratios enables realistic life predictions to be made.

Design of Composite Structures Against Fatigue

116

3000

Fig. 7.8

Results of a rainflow analysis of WISPERX

100

90
+

Predicted
Measured

80
E
s

H-

70
60
50
40
30

0.1
Fig. 7.9

10

IV

L
IV
IV
IV
WISPERX sequences

J
IV

L
IV

10

Predicted versus actual lives as a function of passes through WISPERX

7.6 Fatigue properties of alternative species


Most of the wood composite wingblades manufactured in the last ten years
have been manufactured from Khaya ivorensis, an African mahogany. Khaya
is still considered to be an excellent material for blade fabrication but commer
cial and environmental pressures have prompted the evaluation of other wood
species (12). Poplar, B altic pine, birch, and beech veneers are available in

117

Fatigue Properties of Wood Composites

Table 7.3

Species
Khaya
Poplar
Baltic pine
Birch
Beech

Wood species and properties

Density
(kg/mS)

Tensile
strength
(MPa)

Compressive
strength
(MPa)

550
450
550
670
720

80
63
105
117
103

50
37
40
55
69

Notes: (i) VeneersReliance veneers and other suppliers


(ii) Resin epoxy SP 110/210 (SP Systems)
(iii) Manufacture lamination/vacuum bagging

70

'

I
S 40

Khaya
Baltic pine
Birch
o Poplar
Beech

'"~^

o^
a 30
S

OB C H P ^ *

S 20

5"

*"

10
0.1
Fig. 7.10

1
10

1
IV

I
1
IV
IV
Cycles

I
IV

l
10'

1
10'

10'

.S'.'V curves at R 1 for five wood species containing outer scarf joints and inner butt
joints

commercial quantities at competitive prices and offer adequate mechanical


properties (Table 7.3)
Laminated panels were prepared by WEG for fatigue evaluation and
samples cut from the panels contained either butt and scarf joints or all scarf
joints. All samples had an external glassfibre/epoxyouter skin to simulate the
structure of a blade.
Fatigue lives measured at R = 1 for allfivespecies are presented in an SN
curve in Fig. 7.10. All species contained scarf joints in the outer veneers and
butt joints in all the inner veneers. Similar lives are obtained for the beech,
Baltic pine, Khaya, and birch, although the B altic pine lives are more
scattered. Poplar is the least fatigueresistant material at any given peak stress
level, although it has the lowest density.

118

Design of Composite Structures Against Fatigue


80
Butt scarf
Butts
Over scarfs
Scarfs
Under scarfs
Unjolnted

70
60
s 50

I
I

* 40

30

*+~^
*

nW- + -

+^

DD

20
10
0
0.1

10

10'

IV

IV

10'

Kr-

10'

10

Cycles
Fig. 7.11

S-N curves at R = 1 for Khaya with six joint configurations butt/scarf, all butts,
all scarfs, verstarts, underscarfs and unjointed

Birch and Baltic pine failed by crack initiation from an inner butt joint and
cracks ran through the wood from butt joint to butt joint emerging at the
surface away from the scarf joints. Khaya and poplar failed by crack initiation
from an inner butt joint followed by a more lateral crack propagation through
adjacent veneers, again without interaction with surface butt joints. There is,
therefore, no correlation between the failure mode and the softwood or
hardwood microstructure.
Beech failed by crack initiation from an inner veneer joint followed by
delamination along a glue line and horizontal cracking via a veneer or at a joint.
Despite the poorer glue line quality the fatigue lives were longer. Overall,
beech is a low cost alternative to Khaya, offering good static and fatigue
properties, and it is available in a 6 mm veneer thickness (4 mm for Khaya)
which substantially reduces the number of glue lines.
7.7 Effect of joint configuration on fatigue performance
Early wood composite wingblades contained butt joints only, but more recent
designs have incorporated surface scarf joints which are known to transfer
stresses between the ends of veneers quite effectively. Fiveveneerthick
Khaya samples were manufactured with six joint configurations, namely all
butt joints, inner butts and outer scarfs, all scarfs, all underscarfs, all over
scarfs and unjointed. R = 1 SN results are plotted in Fig. 7.11.

Fatigue Properties of Wood Composites

119

Jointing significantly reduces the unjointed static tensile strength (average 80


MPa) for most configurations except for perfectly scarfed samples where no
sample failed below 75 MPa. Unjointed fatigue life is superior by a factor of 10
to all the jointed configurations. Of the jointed samples the overscarfed
samples perform marginally better in fatigue. Fatigue tests of jointed poplar
demonstrate that all scarfed poplar performs as well as butt-scarfed Khaya.
Hence carefully jointed, low density species can perform as well as less
accurately jointed high density species.
7.8 Infra red condition monitoring of joints
The study of damage mechanisms in materials is often a difficult and complicated one. This is particularly so in the case of wood with its three-dimensional
cellular structure and its characteristic failure modes. There is a need for a
method to monitor and understand the various physical processes which occur
within blades during service, with special reference to the performance of
joints. When a structure is fatigued, heat is generated due to bulk hysteresis
and internal friction effects, especially at joints. Heat generation will be related
to the degree of damage in the wood composite. The surface manifestation of
this heat can be detected using an infra-red camera (13) and analysed to give
information about the underlying flaw.
The infra-red camera used was an AGEMA 880 with a liquid nitrogen cooled
mercury cadmium telluride detector. This instrument operates in the 8-13/im
waveband. The minimum resolvable temperature difference at the sample
surface is 0.1C. A system of rotating prisms directs the infra-red radiation onto
the detector and an image of 140 x 140 pixels is produced. Raw data are stored
both on video (25 frames/second) and as freeze frames on computer disk at
rates of up to 25 frames/second. Data stored on video loses its calibration
constants, so frames have to be periodically stored on disk to retain this
information for subsequent use with the video recording.
Khaya samples were tested with various joint geometries, representative of
those likely to be found in blades, e.g., scarfs, butts, and imperfect joints.
Testing was carried out under reversed loading (R = 1) conditions at various
stress levels and at a stress application rate of 400 MPa/s and samples were
viewed from the side, allowing the inspection of damage accumulation at
joints.
A schematic of a four scarf jointed sample, indicating the path along which
failure occurred, is viewed in Fig. 7.12. An infra-red image of this sample at a
point during fatigue testing when the peak stress had just been raised to 35 MPa
is presented in Fig. 7.13. A hot spot with a temperature of approximately 34C
is clearly visible at the left hand side, lower, outer scarf joint. Another image of
the same sample, Fig. 7.14, captures the situation at the point when failure
occurred. This time the highest temperature zone is over the left-hand side,
upper, inner scarf joint which showed a rapid increase in temperature after the

120

Design of Composite Structures Against Fatigue


_,, ,
if joint

Initiation point

GRP skin

for fanure

fa.

'

>

. . ,
I

I * ' *

^^AA,
I

l\

I '

_rack path
Fig. 7.12

Fig. 7.13

Schematic of four scarf jointed sample indicating failure path

Infra-red thermographic image of four scarf sample during testing

peak stress level was raised to 35 MPa. This joint region was the initiation site
for failure.
Localized heat build-up in the form of hot spots occurred at joints in all the
samples during fatigue testing with temperatures of 50C or more being
observed at the onset of failure. The location of the highest temperature region
indicated the joint at which failure would initiate. However, sites of maximum
damage changed during a single test. This was manifested by the movement of
the hottest spot from one joint to another as peak stress levels were altered
during testing.
It will be useful to attempt to establish a quantitative link between tempera
ture increase and the corresponding damage seen by the wood. The mapping of
damaged zones and crack development will be a helpful aid for verifying and

Fatigue Properties of Wood Composites

Fig. 7.14

121

Infra-red thermographic image of four scarf sample at a point just before failure

developing life prediction methods as well as being a very effective NDT


method.
7.9 Conclusions
(a) Experience gained from operating horizontal axis wind turbines with wood
composite blades in Great Britain and the USA has demonstrated the cost
effectiveness and fatigue resistance of laminated wood/epoxy aerofoil
structures.
(b) Wood composites are notch-insensitive and do not show a size effect when
subjected to axial fatigue loads.
(c) A life prediction analysis has been developed which is able to predict the
life of wood composite turbine blades in different, site-dependent, wind
environments.
(d) The choice of alternative species is dictated by the availability of thick (4-6
mm) veneer from sustainable sources and the fatigue performance
depends on density.
(e) Joints have a critical influence on fatigue life and careful jointing of low
density species such as poplar can result in fatigue lives greater than poorly
jointed higher density wood.
(f) Infra-red thermography (IRT) has clearly demonstrated that internal
damage initiates close to joints and that significant temperature rises of the
order of 50C occur very close to the point of failure.
References
(1) LIPMAN, N. H., 1983, California's wind farm progress Windirections, 3, 16-18.
(2) BWEA, 1987, A second UK windfarm now completed, Windirections, 6, 13.

122

Design of Composite Structures Against Fatigue

(3) PRETLOVE, A. J. and WORTHINGTON, P. J., 1983, A review of aero-generator fatigue


problems, Int. J. Fatigue, 15-22.
(4) SPRUCE, C , 1988/89, Wind Energy Group MS-3 prototype wind turbine. Windirections, 8,
20-22.
(5) QUARTON, D., 1988, The Westinghouse WWG-0640 wind turbine, Windirections, 7, 1617.
(6) SCURLOCK, J., 1991, Series production of wood blades started. Windpower Monthly, 7,
18-19.
(7) BONFIELD, P. W., 1989, Materials selection for wind turbine blades, Windirections, 8, 2829.
(8) DAVIDSON, R., 1990, Wood still beats every aerospace material, says Gougeon, Windpower Monthly, 6, 19-20.
(9) BONFIELD, P. W. and ANSELL, M. P., 1991, The fatigue properties of wood in tension,
compression and shear, J. Mater. Sci., 26,4765-4773.
(10) BONFIELD, P. W. and ANSELL, M. P., 1990, Fatigue testing of wood composites for
aerogenerator rotor blades. Part V - Life prediction analysis and hysteresis. Wind energy
conversion 1990, (edited by T. D. Davies, J. A. Halliday and J. P. Palutikof, (Mechanical
Engineering Publications Limited, London, UK) pp. 19-24.
(11) BONFIELD, P. W. and ANSELL, M. P., 1991, Fatigue testing of wood composites for
aerogenerator rotor blades. Part VI - Spectrum fatigue loading, life prediction, and damage
rates. Wind energy conversion 1991, (edited by D. C. Quarton and V. C. Fcnton),
(Mechanical Engineering Publications Limited, London), pp. 311-316.
(12) BONFIELD, P. W., BOND, I. P., HACKER, C. L. and ANSELL, M. P., 1992, Fatigue
testing of wood composites for aerogenerator rotor blades. Part VII-Altcrnative wood
species and joints. Wind energy conversion 1992, (edited by B. R. Clayton, Mechanical
Engineering Publications Limited, London, UK), pp. 243-250.
(13) DUTTON, A. G., IRVING, A. D., LIPMAN, N. H., CLAYTON, . R., AFTAB, ., and
BOND.L. J. (1992) Infra-red condition monitoring of wind turbine blade fatigue tests, Wind
energy conversion 1992, (edited by B. R. Clayton) (Mechanical Engineering Publications
Limited, London, UK), pp. 221-228.

CHAPTER 8

Benchmark Tests
R. M. Mayer*
There is a lack of agreed methods for fatigue testing of coupons. A benchmark test is undertaken
using a common GRP laminate with more than 80 percent of the fibres in the primary load
direction. The methods are reported and discussed, and the data are analysed.

8.1 Introduction
The available GRP test methods are listed in Table 8.1 for both tensile and
flexural testing of coupons.
Principal parameters which can vary include: the shape of the specimen; the
method of introducing the load; and, in the case of flexural fatigue whether a
roller or hemispherical loading face is used.
Various laboratories have evolved methods that are self-consistent in themselves, although a recent survey of the available data (1) shows that most data
are not recorded in a manner which would allow it to be reproduced.
For these reasons, a benchmark test was conducted amongst seven institutes
using a common material.
8.2 Material selection
The tensile testing of coupons becomes more problematic as the proportion of
glass rovings in the prime loading direction increases. At the same time,
Table 8.1 Relevant coupon test methods
Mode

Property

Static

Tensile
Flexure

Fatigue

Sciotech, UK.

Tensile
Flexure

Number
ISO 527
ISO 3268
EN 61
ISO 178
EN 2562
EN 63
ASTM D 3479
AFNORT51-120

Scope
Plastics
GRP
FRP
GRP
FRP
Plastics

124

Design of Composite Structures Against Fatigue


Table 8.2 Resin properties

Type

Norpol 20-80 (Jotun Polymer)

Tensile Strength
Modulus
Strain to failure
Flexural strength
Modulus
HDT

Table 8.3

73 MPa
3.1 GPa
6.5 percent
135 MPa
3.0 GPa
81C

Fabric and fabrication

Type

LM 600/150*

Warp
Weft
Process
Chop layer
Binder
Number of plies
Lay-up

3.1 ends/cm, 1800 tex, RP 251t


0.9 ends/cm, 400 tex, RP 251t
Woven, 602 g/m2
113 g/m2, 2400 tex, P7t
Powder
Seven
[0(565), 90(37)W + CSM (113)
[3p]/0(565),90(37)W[l/2p]]
Contact moulding
3 h at 80C

Fabrication method
Post cure
'Flemings Fabrics.
Rovings supplied by Vetrotex

reinforcement for rotor blades is mainly required along the length of the blade.
For this reason, a pseudo-unidirectional fabric was selected, which is typical of
many industrial applications. It is a combination fabric, comprising primarily of
UD woven fabric, to which a chop strand layer is bonded; 84 percent of the
fibres are in the zero direction. The resin is a medium reactivity iso-NPG
polyester from Jotun Polymer. Details of the resin and fabric are given in
Tables 8.2 and 8.3.
A single plate 2400 x 1200 x 5 mm was manufactured by Jotun Polymer as
follows:
- three layers of combination fabric;
- one layer of the woven fabric only;
- three layers of combination fabric.
Since the woven side of the fabric was always to the outside of the laminate, the
laminate was balanced. Following postcuring, a set of plates of size 400 x 600
mm were then cut and supplied to each institute. The fibre volume fraction was
35 percent.

Benchmark Tests
Table 8.4

DLR
DNVR
ECN
FFA
UWE
Scio

125

Flexural testing details

Coupon width
(mm)

Roller loading
(mm)

Diameter outer
(mm)

15
15
25
30
25
25

10
10
20
30
25
20

4
10
10
50
25
10

(EN 63)

Note: A span to depth of 20:1 was used

3
Fig. 8.1

4.5
6
Deflection (mm)

7.5

10.5

Two typical load-deflection curves for three-point flexural bending (DLR)

8.3 Flexural testing


A variety of coupon widths measuring 15-30 mm were used whilst the span was
kept constant at 100 mm. Three rollers were used for loading, with the radius of
the support rollers varying from 4 to 50 mm and that of the loading roller from
10 to 30 mm (Table 8.4).
Two typical load-deflection cures are shown in Fig. 8.1. It can be seen that
the curve is linear over most of its length except at high strains when some
compressive damage started to occur - thus the slope can be clearly defined,
and so can the failure stress.
Samples failed either by compression or shear or both.

Design of Composite Structures Against Fatigue

126

Table 8.5 Flexural properties of benchmark laminates


E/
(GPa)
DLR
ECN
FFA
UWE
Scio

25.0
23.1
23.1
21.5
21.7

1.5
1.0
1.5
0.7
2.3

(MPa)

(%)

Number
samples

616 3 8
644 3 0
nm*
601 2 6
612 3 1

2.6 0 . 3
2.9 0.2
nm
2.8 0 . 1
2.9 0 . 3

8
17
4
.8
6

'Note: shear failure; deemed invalid

Table 8.6 Tensile sample details and testing

DLR
DNVR
ECN
FFA
UWE
Scio
Ris

Length
(mm)

Width
(mm)

100
250
80
120
80
UK)
85

25
25
21
35
25
25
25

Tabs
(mm)

Chapter
reference

5.2
3.2
6.2
4.3

50x1.5

100 x 2.0
50 x 5.0
60x2.0

2.2

Note: Strains measured using an extensometer, except for Sciotech, who


used a strain gauge

The results were evaluated using standard formulae with no allowance for
the shear component due to the span/depth ratio on the modulus (Table 8.5).
The greatest variation is found in the modulus, where a difference of 3.5 GPa
separates the highest and lowest means. There is a much narrower spread in the
failure stress and strain, which is surprising as it is generally more difficult to
determine the failure point than the slope of the curve.
8.4 Tensile testing
There are significant variations in specimen width, gauge length, specimen
shape, and method of gripping. The test methods were those described in other
chapters, and the details are tabulated in Table 8.6.
All samples were straight sections, with the exception of ECN, which was
curved (Fig. 8.2). The type of tabs varies widely, but their use is generally
deemed essential to ensure good load transfer between the hardened steel grips
and the softer GRP coupons.
A typical load-deflection curve is shown in Fig. 8.3 and it has similar
characteristics to theflexuralcurve, except that it has a slight curvature at high
strains. Therefore, when measuring the modulus it is important to specify
whether the slope is the initial tangent or the secant at some specified stress or
strain (Fig. 0.1).

Benchmark Tests

127
,i.s

-IM

l.l.l.i.t.l.l.l.l.l.l.l.

21
- Specimen
r400

Tabs

25
Fig. 8.2

Sample shape for tensile test used by ECN. Dimensions in mm

The secant modulus will decrease as the strain range over which it is
measured increases. This is illustrated for three coupons in Fig. 8.4.
To obtain a valid value of the failure strength and strain, checks were made
to ensure that failure occurred within the gauge length and not in the grip area.
The mean data are tabulated in Table 8.7.
The values of the moduli are in close agreement, with little scatter. The
scatter in stress is somewhat larger than the modulus and the corresponding
values forflexure.This reflects the difficulty in measuring the tensile properties
of highly orientated laminates.
8.5 Significance of static results
The average of the data in Tables 8.5 and 8.7 is given in Table 8.8 with the
deviations covering all the values.
The most striking feature is that the tensile modulus is higher and the tensile
strength lower than the corresponding values for flexure. This result is identical

128

Design of Composite Structures Against Fatigue


1

701

r-,

60 -

2
3
Elongation (mm)
Fig. 8.3

261

Two typical load-deflection curves for samples loaded in tension (ECN)

-|

23

22

Fig. 8.4

_l

l_

0.5

1.0

1.5
Strain range (%)
Variation in secant modulus with strain as measured from zero strain for 3 coupons
(DNVR)

Benchmark Tests

129

Table 8.7 Tensile results from benchmark test

DLR
DNVR
ECN
FFA
UWE
Scio
Ris

E, (initial)
(GPa)

E, (secant)
(GPa)

o,
(MPa)

(%)

25.5
25.0
27.5
23.9
27.5

24.3 1 . 0
23.5 0.5
25.0 1 . 0
nm
24.7 1 . 1
25.2 1 . 0
21.6 0 . 4

572 27
483 20
538 9
488 23
534 14
468 3 2
447 1 7

2.4 0 . 1
2.0
nm
nm
2.1 0 . 1
1.9 0 . 2
2.1 0 . 1

0.9
0.5
1.1
0.5
1.1

23.8 0.4

e,

Table 8.8 Averages of the mean values of the tensile and


flexural properties as measured by various institutes

Flexure
Tension

E
(GPa)

(MPa)

(%)

22.9 2.1
24.5 1.0

618 26
514 58

2.8 0.2
2.1 0.2

to that reported for all other fibre-reinforced plastics, as a recent survey has
shown (1).
The lower tensile strength is attributed to the larger volume of material being
loaded; this increases the probability of finding a flaw. The lower flexural
stiffness can be attributed to the effects of shear and the proximity of the
reinforcement to the surface of the laminate.
The scatter in data is regarded as satisfactory as no attempt was made to
standardize the test methods. As these methods were subsequently used to
obtain fatigue data, the associated change in property will have less scatter than
the absolute values, as they will be independent of the test method.
8.6 Fatigue testing
The same coupons and test methods were used to undertake fatigue testing.
Three types of loading were applied:
(a) tensile loading with an R ratio of 0.1;
(b) flexural loading with R = 0.1;
(c) reverse loading with R = 1.
All tests were conducted in load control and the reduction in strain (and,
therefore, modulus) was monitored.
The results are plotted in Figs 8.5 and 8.6 against the normalized stress and
strain as a function of the number of fatigue cycles. Curves have been drawn
through the points to illustrate the trends.

Design of Composite Structures Against Fatigue

130
1.0

+ UWE
X Riso
O ECN
D DNV

0.9
0.8
0.7

Flexure ( = 0.1)
0.6
0.5
Reverse loading*
( = D

0.4

"""boo

0.3
0.2

J_

J.

10'
10'
10'
Cycles
Fig. 8.5 Benchmark test of a pseudounidirectional GRP composite as tested by various institutes.
Data plotted as a function of normalized stress versus number of cycles for three types of
loading
10=

10'

+
X
O
O

2.0

UWE
Ris
ECN
DNV

Flexure

o^oo

1.2

Tension
x

Reverse loading ^

Ox

*o

0.8

0.4

102

I I I I III

IV
Cycles

IV

10"

Fig. 8.6 B enchmark test of a pseudounidirectional GRP composite. Data plotted as a function of
strain versus number of cycles for three types of loading

Benchmark Tests

131

8.7 Analysis of the fatigue data


The most striking observation is that the three fatigue curves have a similar
slope, especially when plotted on a normalized stress basis. Though different
loading regimes, sample shapes, and methods of gripping are involved, this
implies that these methods are detecting a similar rate of structural degradation.
Evidence from previous work on other GRP material combinations supports
this observation (1)(2). This is discussed further in Chapter 14, in the development of standard fatigue curves. One can, therefore, conclude that the test
methods used in this benchmark test for measuring fatigue give self-consistent
results.
There is more scatter amongst the data points for the tensile loading than
flexural or reverse loading. The reason for this is not known, but it does imply
the need to measure sufficient data points to obtain a valid trend with fatigue
curves.
After longer time periods (i.e., higher cycles) it is unlikely that the fatigue
curves will maintain their linearity when plotted on a linear-log basis. The
tendency for the curves to flatten is illustrated by the strain curve for reverse
loading in Fig. 8.6. Thus extrapolation from short to long times could be
inaccurate, particularly if a test method that has not been benchmarked is used.
8.8 Recommendations
The following recommendations can be made on the basis of this benchmark
test:
(a)
(b)
(c)
(d)

any non-standard fatigue test method should be benchmarked;


failure of the coupon should occur within the gauge length;
the method of static loading should be validated;
sufficient data points should be collected in order to derive a valid trend for
a fatigue curve (see Chapter 14).
(e) care should be taken in extrapolating data from short to long times (i.e.,
high cycles).
(f) this methodology could be used to benchmark other test methods (2).

References
(1) HANCOX, N. L. and MAYER, R. M., 1993, Design data for reinforced plastics, (Chapman &
Hall, London).
(2) Research Programme for Fatigue-Critical Structures of Small and Medium Sized Wind
Turbines, 1995, (edited by C. Kensche) (DLR).

CHAPTER 9

Comparison of Coupon and Spar Tests


G. A. Lowe andN. Satterly*
Fatigue testing of glass-reinforced polyester coupons is undertaken as well as D-spars manufactured using the transverse filament winding process. The effect of changing the manufacturing
technique is investigated in order to improve the design and production process.
The relative effect of size is considered by comparing the results obtained from spars with
those of coupons extracted from spars and cut from flat plate.
The fatigue life of spars is considerably lower than that predicted from testing small coupons
used to characterize the material. The production quality should be similar in order to compile
accurate data for design purposes.

9.1 Introduction
The rotor blades of wind turbine generators can expect to be subjected to over
100 million fatigue cycles within their lifetime. Thus the choice of materials and
an adequate knowledge of their static and fatigue properties are essential
elements of rotor design.
The method of manufacturing spars by tape winding has been used since the
first rotor blades were made for the turbines at Nibe in 1978 (1). It allows
automated manufacture of the 'D'-shaped spar (Fig. 1.6), thus leading to a
reliable and reproducible component.
This investigation consisted of two distinct aspects:
(a) base material properties obtained from tests plates produced under laboratory conditions;
(b) blade spars and coupons produced from the same materials using bulk
manufacturing techniques, characteristic of volume production. (The spar
is the primary load-carrying element of the chosen blade design and hence
the aerodynamic shells did not form part of this investigation.)
The objectives of the work were to establish the following:
(a) how long such material and components would last under a given loading
cycle;
(b) how damage accumulated in such materials and structures. Various
damage parameters were monitored throughout the tests, to establish
"University of the West of England, UK.

134

Design of Composite Structures Against Fatigue

whether their rate of change would provide an estimate of the severity of


the damage and hence the structural integrity at any point in the life of the
component.
These topics are of direct interest to designers and operators who wish to
produce and operate more cost effective machines more efficiently.
9.2 Materials and manufacture
Materials
Glass-reinforced isophathallic polyester resin has been used for the production
of flat plates and spar sections. This material is typical of that used for the
production of wind turbines.
The current authors used two types of glass tape reinforcement: the tradi
tional woven and the more recent knitted or inlaid construction. Both materials
were in the form of a 210mm wide tape, with polyester yarns in the longitudinal
(warp) direction and glass rovings in the transverse (weft) direction (Fig. 9.1).
The polyester yarns in the woven construction pass over and under success
ive glass rovings, causing them to kink. This is believed to give rise to stress
concentrations, with a consequential reduction in fatigue performance. When
winding takes place and tension is applied to the polyester yarns the above
problem is extenuated. However, the inlaid construction keeps the fibres
straight, eliminating the kinks and associated stresses.
Additionally this construction allows more stretch and hence is more capable
of complying to the tapered shape of the mandrel used in the manufacture of
the spar.
The materials are listed in Table 9.1 and the results of investigating the
different types of reinforcement and tape tension are given in (2).
Manufacture
Flat plates were manufactured using the leaky mould technique, with a
nominal pressure of 0.1 bar applied to consolidate the laminate. Post curing for
3 h at 80C was carried out. The plates were made to provide a benchmark for
comparison with the tape wound spars.
Spars were produced by two manufacturers using their own versions of the
tape winding process. Glass tape is wound onto a tapered rotating mandrel,
with the resin applied automatically or manually by a pumped roller. An
overlap of approximately 40 percent between successive layers, in which the
fibres lay along the mandrel axis, is maintained by careful control of the angle
to account for the taper and change in geometry of the mandrel (Fig. 9.2). This
0 degree laminate is essential for maximum strength and stiffness (3). Follow
ing winding the spar is allowed to cure for 24 h before the component is pressed
off.

Comparison of Coupon and Spar Tests


.Glass rovings (weft)

135

.Polyester yarns (warp)

~21

e polvesi
yarns at edge

(a)

iHHtttHHHu

Glass rovings (weft)

p = p Polyester yarns (warp)


Chain stitching

(b)
Fig. 9.1 Construction of inlaid transverse filament tape showing: (a) location of polyester yarns and
glass rovings; and (b) method of joining

Table 9.1

Reinforcements and resins

Material

Type

Transverse filament tape

Resins

Woven, mass 278 g/m


Inlaid, mass 410 g/m2
'E' glass yarn (1200 tex)
Polyester yarn (30 tex)
Isopolyester UP334
Isopolyester Crystic 272
Catalyst MEKP solution
Accelerator copper cobalt styrene solution

Manufacturer
Ahlstrm Glassfibre
Flemings Fabrics
Hoechst
Scott B ader

136

Design of Composite Structures Against Fatigue


Overlap between
successive layen
Fibres aligned with
mandrel axis

Mandrel

Glass tape

-Tlipe carriage

Fig. 9.2 Tape being wound onto a tapered mandrel

Fig. 9.3 Arrangement of three-point flexural test jig

9.3 Material characterization


Flat plates and spar sections have been characterized by a series of tests.
Flexural testing (as shown in Fig. 9.3) has established the flexural modulus,
strength, and failure strain using standard three-point bend tests (4). This type

Comparison of Coupon and Spar Tests

137

of testing has been chosen because the tape wound material is unidirectional
and of varying thickness. Some tensile testing has been carried out, but valid
failures have been impossible to obtain.
Fibre volume has been obtained from resin burn off tests (5) (6), whilst
interlaminar shear strength has been measured by short beam shear tests (6)
(7). This test is a good quality check and the type of failure is independent of
features, such as joints, that exist in the tape wound sections.
Static test results
The results show that the flat plates are of better quality than the spar sections
and that the inlaid material (I) is better than the woven (W) (Fig. 9.4).
The plate results show similar strengths and stiffness, the mode of failure
being tensile. It is, therefore, believed that the small differences in properties
are due to the kinks in the woven construction.
The greater difference in properties between the two types of spar material
can be explained by the failure mode. Whilst the inlaid samples always failed in
tension the woven samples exhibited significant signs of compression damage.
The latter is almost certainly due to the kinks permitting buckling on a
macroscopic scale.
The large difference between the properties of the plate and spar coupons
can be explained by a general reduction in quality arising from the production
method, including high voidage, and the effect of the tape joints. This was
confirmed by measurements of the interlaminar shear stress (Table 9.2).
Table 9.2

Interlaminar shear strength and fibre volume (Vf)-test coupons

Sample form

Material

Wound spar section

Woven
Inlaid
Woven
Inlaid
Inlaid

Hand laminated flat plate


Wound hoop (6) hand tension
1.75 kg
3.0 kg
4.5 kg

Interlaminar shear
strength (MPa)

V,(%)

29.6
32.5
45.5
41.7
34.7
35.3
35.8
35.8

38.3
45.0
35.0
40.6
34.1
37.7
39.7
38.4

9.4 Fatigue testing of coupons


Fatigue testing was carried out in flexure using a similar arrangement as for
static testing with some modifications (8). The test machine was a Schenck
Pulsermat operating at approximately 8 Hz under load control. An R ratio of
0.25 and a span to depth ratio of 16:1 was used.

138

Design of Composite Structures Against Fatigue

S 30

9.0

V
20

2
s

. 10

/ /
/
//
/
/

'

" n0
'n

\s

S oo
oo ;

..'
"
oo

\\

oo "

'; oo'

/
'/ / /
/
/_/

oo

\
\
\\
\
\
\ \
\v

S oo
OO " :

o"o
S 0 0

"

OO U t

S oo

1000
800
8 600
g 400
9

200

^d

W
spar

spar

W
plate

I
plate

Fig. 9.4 Flexural modulus and stress for four diffrent types of components. W = woven fabric;
I = inlaid fabric

The stiffness of the samples was used as the failure criterion, a 10 percent
reduction of stiffness from the original value meaning failure and equivalent to
100 percent life (8) (9). The results are displayed in Figs 9.5 and 9.6.
Figure 9.5 shows normalized stress data, enabling a comparison of the
material's fatigue characteristics, regardless of their ultimate strength (10).
The plate data indicate that there is some feature of the woven material that
significantly affects the fatigue performance. This is again associated with the
kink stresses which are more serious in fatigue than in the static situation. The
effect is not evident in the spar coupon results, indicating that there is a more
dominant feature affecting the material. It is suggested that the poor quality

Comparison of Coupon and Spar Tests

139

1.0
* Wspar
I spar
W plate
-I-1 plate

0.8

0.6

0.4

0.2

10"

I I I III!

KT

10"

10*

10'

10-

Cycles

Fig. 9.5 Flexural fatigue of four components displayed as the normalized stress versus logarithm of
number of cycles to failure, R = 0.25

and the overlap joints cause the lower values, emphasizing the need to optimize
the production process to ensure that the material is used to its full capacity.
Figure 9.6, relating test strain to fatigue life, clearly shows that the tape
winding technique produces a material inferior in quality to the plates. This
highlights the need for manufacturers to ensure that fatigue data used for
design purposes originate from representative coupon tests.
9.5 Methods of monitoring damage
Throughout the fatigue testing programme various parameters were moni
tored on both coupons and spars, using a computercontrolled data logging
system developed at the University of the West of England. The parameters
recorded were: flexural stiffness, hysteresis area, surface temperature, and
local strains (spar only). Components were visually inspected at regular
intervals.
The time history of these parameters was used to correlate the rate of change
of the parameter with amount of damage with regard to the life of the sample.
Figures 9.7 and 9.8 show the stiffness, hysteresis area, and surface tempera
ture change during the normalized life (9) of typical coupons under compara
tively high and low stress (75 percent and 45 percent UFS), respectively.
The reduction of stiffness with time is similar, indicating little dependence on
stress. However, the other two parameters are very different; the high stress

Design of Composite Structures Against Fatigue

140
2.5

*
x

-I-

2.0

Wspar
Ispar
W plaie
I plate

I 1.5
TS

3
E
1.0

0.5

10V

IO1

10*

10'

10'

Cycles

Fig. 9.6 Flexural fatigue data plotted as a function of maximum strain versus cycles to failure

Fig. 9.7 Changes in modulus, temperature, and hysteresis areas as a function of the life of the
coupon; 100 percent life = 10 percent reduction in modulus. Inlaid flat plate tested at 75
percent UFS; 10 500 cycles to failure

Comparison of Coupon and Spar Tests

141

EIE,

3 0.8

.'..
* if

V,' .'<!.-A
* '

20

'

40

I tJ

60
Life (%)

/'V
"
80

100

' 1
2 g.

0
120

Fig. 9.8 Property changes as a function of coupon life. Inlaid flat plate tested at 45 percent UFS;
15 000 000 cycles to failure

producing large changes and, conversely, the low stress producing small
changes. There is a good correlation between the changes in the hysteresis area
and temperature, as would be expected from a viscoelastic material such as
GRP, both showing an accelerating trend towards the end of the sample's life.
The dependence of these parameters on high stress and hence high damage rate
is obvious.
Reifsnider (9) observed the damage accumulating in composites with cross
plies and described the damage in three stages (as shown in Fig. 1.2). Stage 1 is
where damage initiates, matrix cracks occur, and weak fibres break; this stage
is characterized by a fairly rapid but short accumulation of damage (up to 20
percent of life). In stage 2 other damage mechanisms take over with a slow,
steady and relatively long period of damage accumulation, up to 80 percent of
life. Stage 3 sees an accumulation of the damage until fracture.
Comparing the shape of Reifsmider's damage curve with those in Figs 9.7
and 9.8 for stiffness change, we can see stages 2 and 3 are quite well defined, but
there is little evidence of stage 1. It can only be assumed that the unidirectional
material tested in this work is not as susceptible to weakening caused by the
initial degradation of fibre breakage and matrix cracking as was the material
with cross plies.
9.6 Component testing
Three spars of two different designs were tested at high strain conditions in
order to accelerate damage. The testing was carried out in the structural test
rig, as shown in Fig. 9.9, and components were fatigue tested with a frequency
between 0.3 and 1 Hz. Typical local instrumentation used for monitoring
damage is shown in Fig. 9.10, which shows strain gauge devices and thermo
couples.

Design of Composite Structures Against Fatigue

142

5200

Portal frame

Gauges

Spar

Fig. 9.9 General arrangement of rig for spar testing; dimensions in mm; Ivdt is a displacement
transducer

Extensometer

Strain gauge

Thermocouple

4P . .

Demec stud
Fig. 9.10 L ayout of local instrumentation in an area of high strain - strain gauges, thermocouples,
and Demec gauges

Spar 1 (O L Boats, Denmark)


For the first spar component it was predicted, from coupon data, that there
would be a 10 percent drop in stiffness after 1 million cycles at a strain of 0.6 or
35 percent of the material's ultimate strength. This was a much higher strain
than wind turbines in service would ever see (where a typical maximum stress
would be 10 percent of ultimate), but to test at a low strain would have been too

Comparison of Coupon and Spar Tests

143

/'

High damage

Some damage

40

60

100

Llfe(%)
Fig. 9.11 Increase in temperature during fatigue of three areas with varying amounts of damage

time consuming. Cycling commenced with periodic monitoring of damage


parameters and visual inspection made on a regular basis.
The visual damage proved to be the best indicator of damage. At 55 000
cycles the global drop in stiffness measured with a displacement transducer was
only 8 percent, but the spar was on the verge of failure. Such a drop in stiffness
on a coupon would only have produced minor visible damage.
Clearly common criteria, such as a 10 percent loss of global stiffness, cannot
necessarily be applied to both coupons and spars. Any loss of stiffness shown by
a large component must be treated seriously and global damage monitoring
must be very accurate to detect local damage. This very much affects the
structural integrity of the component, but can sometimes be masked by the
global measurements.
Some of the other damage monitoring parameters used in the coupon
testing, such as change of temperature and hysteresis area, did not provide
significant information when used with the components. The strain gauges
which were mounted on the surface gave disappointing results since they failed
early in the test due to cracks in the surface of the spar to which they were
attached.
The test on this spar was continued, however, with cracks clearly visible. It
was found that the monitoring of temperature in thisfinalphase and applied to
areas containing varying degrees of damage, was effective in proving the
potential of the technique. Figure 9.11 shows the increases in the temperature

144

Design of Composite Structures Against Fatigue

Table 9.3 History of damage for spar 3, strain 0.3%


Cycles

Area mm2

<30 000
<110 000
<1 840 000

17
286

<2 500 000

325

Description
Small crack transverse to glass filament, 10 mm in length
Crack is now area of delamination 3 mm by 11 mm
Area has increased steadily and has linked up with other areas of
damage
Slow growth of damage: an additional small crack (90 degrees)
has now linked up with damage area

measured at three different areas, and the trends are clear. The same success
was not noted in the measurement of hysteresis area which was subject to much
scatter.
Spar 2 (Polymarin, Holland)
It was decided to test this spar at a high strain of 0.6 percent, in order to give a
reasonable amount of time in which to gather data. It was estimated that this
spar would last 106 cycles, but in fact it failed at less than 105 cycles due mainly
to cracks and delaminations at the tape overlaps. Because of the unexpected
rapid development of damage some of the measuring techniques, such as strain
and temperature, did not capture enough data in the programmed intervalsof
inspection and visual inspection usingreflectedand transmitted light was found
to be the best indicator.
Spar 3 (Polymarin)
Following the early failure of previous spars it was decided to lower the test
strain to 0.3 percent, which was intended to give a life of at least 106 cycles
based on the coupon data for the same material. In the event this spar did not
fail, but the most prominent damage in the form of a small crack started to
appear at 33 000 cycles and was an area of delamination. It spread steadily until
it joined up with other damage later in the test. Table 9.3 summarizes the
damage development.
The growth of this damage is shown in Fig. 9.12 where the area of damage
was traced at inspections and quantified by measurement. The curve can be
divided into three definite phases and relates in some respects to the three
stages discussed previously. There is a damage initiation phase during which
the damage started and progressed at an even rate for thefirst25 percent of the
test (Fig. 9.12). The rate of damage then accelerated to another steady state
where growth continued for some time. At approximately 75 percent of the test
the damage area joined up with other local damage which was running
transverse to its direction of progression. This slowed the rate of damage
accumulation considerably, because it was the most active damage front that
had been blocked.

Comparison of Coupon and Spar Tests


J3U

145
1

300

250

200

g 150

100
50
01^

20

40

60

80

100

Life(%)

Fig. 9.12 Growth of visual damage as a function of spar life; maximum strain 0.3 percent

For this test the sensitivity of the surface temperature measurements was
improved by a factor of 2, to resolve to approximately 0.1C. Even so, there
was very little indication of temperature change. Similarly, there was no
change of hysteresis area for this lower strain test.
Strain gauges were placed at the positions of high strain, so the measurement
of local strain was far more successful than with the previous spar, all gauges
remaining operational throughout the test. The results, not surprisingly when
considering the lack of significant damage, showed no real change in strain, but
this was comparable with the results of the other methods used.
More testing was carried out on this spar by progressively increasing the
strain level (Table 9.4). Areas of damage were studied and a typical plot of one
significant area is shown in Fig. 9.13, which indicates that the rate of accumu
lation of visual damage is proportional to the applied strain and shows the
various stages of damage as displayed by the coupons.
9.7 Discussion
The testing showed that the inlaid construction is significantly better in fatigue
than the woven material, although static properties are similar. This is believed
to be due to the kink stresses induced in the woven construction when loaded
both statically and dynamically.
It was found that the fatigue lives of components are significantly shorter

146

Design of Composite Structures Against Fatigue

Table 9.4 Strain history of spar 3


Cycles

Description of damage

0.3

<35 000

0.3

<111000

0.3
0.3

<330000
<840000

Small crack at edge of tape overlap on outer surface @ 2.62m


from flange face (FFF)-damage no. 1.
Crack growing and additional small cracks just outboard @
2.64m FFF- damage nos 2 and 3.
Initial damage is now an area, other cracks lengthening.
Area of damage growing and has now linked up with cracks
outboard at 2.64m FFF.
Strain increased, additional cracks forming in areas outboard of
original area (2.62m FFF) which has now linked up with another
crack.
Strain increased, steady growth of existing damage, lots of
additional damage initiating from the edge of tape joints on outer
surface. Several areas of delamination appearing at tape joints on
inner surface of spar.
Strain increased, damage development speeding up, inner
surface delaminations developing fairly rapidly.
Strain increased, all damage developing quickly. Large inner
surface delamination advancing particularly rapidly.

Maximum
strain (%)

0.375

<2 500 000

0.45

<4000000

0.6

<5 300 000

0.6

<5 500 000

40

60

80

100

Llfe(%)

Fig. 9.13 Cumulative increase in visual damage throughout various strain increments

than those predicted from coupon data, typically by a factor of 10. However, it
should be noted that only a small number of components were tested and the
data are subjected to scatter.
Whilst differences in quality have an effect on fatigue performance for
typical wind turbine applications, i.e., high cycle, low stress conditions, these
effects are less significant. This indicates that GRP is tolerant to manufacturing
errors and is a good material for turbine blades. The results also show that

Comparison of Coupon and Spar Tests

147

quality control of the production method is important in achieving the best


performance.
Coupons should be representative of the lay-up of the final component to
achieve true data. The coupon testing programme, with the associated damage
accumulation measurements, established that there were several methods of
monitoring that were reliable in assessing the amount of damage in a coupon.
Failure of the spar components generally initiated from the overlap joints in
the tape. The spars manufactured from the inlaid tape of greater mass with
fewer layers clearly failed in this way, whereas the woven tape of lesser mass
and more layers suffered more fibre breakage.
Reducing the amount of load transfer per layer by decreasing the thickness
of the tape may help to strengthen the joint.
Transferring the damage methods developed with coupons to spars had
limited success and great accuracy was required, as small changes could be
overwhelmed by scatter and environment factors. The large area of the spar to
which the instrumentation was attached made correct positioning of the local
measurement methods rather unpredictable since the damage, when it did
occur, remained confined to a small area; nevertheless these should be located
after a strain survey has been made. Even when the instrumentation was
correctly positioned, sensing the damage was very difficult with the methods
used. Any changes in the global parameters that were recorded, i.e., stiffness
and hysteresis area, were found to be small. After losing only 5-8 percent of
global stiffness, spars failed catastrophically, whereas coupons would hardly
have shown any signs of damage.
All the spar measurement techniques were subject to scatter; this could mask
real effects when changes are small. The scatter can be induced by various
factors such as experimental error, change of ambient temperature, and nonpermanent creep. Care has to be taken in assessing the test results.
The best indicator of the spar condition proved to be visual inspection;
monitoring the growth rate of an area of damage was found to illustrate the
various phases of damage accumulation and displayed the same characteristics
as that shown by the coupons.
9.8 Conclusions
The effect of scale on testing does exist and, therefore, an iterative design
strategy is required in order to take this factor into account. It is important to
strive to obtain as good a quality component as the test plate. Damage
development will vary as the size of the component increases. Some monitoring
methods are better with coupons and some with components. The transverse
filament tape method is an effective way of orienting fibres along the axis of a
component.
A typical rotor blade is only likely to see 1 cycle of 0.6 percent strain during
its lifetime (Chapter 5). Thus only testing at such high strains may not be a

148

Design of Composite Structures Against Fatigue

sensible accelerated test, as failure mechanisms may be initiated which are not
typical of lower strains. This is shown by comparing the lifetime of the three
spars tested here. An accelerated test in which the strain is increased sequen
tially (spar 3) appears to be a better way of testing. A similar method has been
used for testing other types of spar (Chapter 10).
References
(1) L O W E . G . A. andSATTERLY.N. D. 1993. Fatigue properties of wingblade materials and
components manufactured by the transverse filament winding process. Proceedings of the
1993 European Community Wind Energy Conference (H. S. Stephens, Bedford), pp. 191194.
(2) LOWE, G. ., SATTERLY, N. D, MAYER, R. M., and PRETLOVE, A. J., 1994, Fatigue
properties and design of wing blades for wind turbines, ETSU Contract W/44/00230, DTI
Report.
(3) RICHARDSON, T., 1987, Composites- design guide, (Industrial Press, New York).
(4) BS2782: Part 10: Method 1005,1977, Determination of flexural properties-three point method.
(5) BS2782: Part 1 : Method 107 K: 1970, Determination of fibre content of resin matrix composites
by resin burn off.
(6) SATTERLY, N. D., 1994, Fatigue studies of uni-directional glass reinforced polyester for
wind turbine blades, MPhil Thesis, University of the West of England, UK.
(7) BS2782: Part 3: Method 341 A: 1977, Determination of interlaminar shear strength of
reinforced plastics.
(8) SIMS, G. D., 1989, A Vamas round robin on fatigue test methods for polymer matrix
composites. Part 1 - tensile and flexural tests of unidirectional material, (NPL Report DMA
(A) 180).
(9) REIFSNIDER, K. L., 1991, Fatigue of composite materials. Composite Materials Series,
Volume 4, (Elsevier, Amsterdam).
(10) ANSELL, M. P. et al., 1988, Testing of wood in fatigue. Proceedings of IEA workshop on
fatigue in wind turbines.

CHAPTER 10

Response of Blade Roots to High Bending


Moments
A. Vinckier and W. Sys*
The blade root is an area of complex load transfer from the blade to the hub which also involves a
change of material (GRP to steel). Design is, therefore, difficult and verification by accelerated
laboratory testing is necessary. Tests on five full-scale blade roots are described and the way in
which damage accumulation is monitored.

10.1 Introduction
It has been well known for many years that to get specific and practical results,
for example accurate life estimates, realistic fatigue tests have to be performed.
This requires an accurate simulation of the actual structure and actual loading
conditions. One can easily fulfil the first condition by performing tests on fullscale components which are fabricated in the same way as the actual structure.
The second condition is more difficult to reproduce. Service loads may have
very complicated load sequences. They can act as distributed loads, such as
wind loads, or as quasi point loads. Further complications arise when high
numbers of cycles, for instance 109 cycles, and varying environmental conditions are involved. Generally, the test loads might be more or less crude
approximations of the service loads.
The most likely places for fatigue weakness are at joints. Here all of the
factors which unfavourably influence fatigue strength are met: stress concentration, bolt holes, built-in stresses, offsets, changes of section, and application
of different materials. Further, the calculated stresses in the joints are frequently of questionable value. Hence, the test component must at least
comprise a complete full-scale root section of the blade, including all attachments. Ideally, the component is an entire blade which might allow for quasi
perfect load application. This is almost impossible when mechanical resonance
systems are used. The main disadvantage of resonance systems is the fact that
variable loading is quasi unfeasible. The servohydraulic fatigue systems with
closed-loop control are well-suited for complex load-time sequences. However, these systems are more impractical when large deflections are involved.
Instead of complete blades, root sections varying from about 2 m to 4 m in
length are commonly used with servohydraulic fatigue systems. The internal
'University of Gent, Belgium.

150

Design of Composite Structures Against Fatigue

loads in these short sections will, however, differ from practice. For example,
the bending moments are reproduced accurately while the shear forces are
exaggerated.
The fatigue test may be performed under constant-amplitude loading or
under varying-amplitude loading. The load levels may be equal to the service
levels, or they may be higher (Fig. 1.3). Generally, it is important to bear in
mind that full-scale testing is carried out because it reduces, or even eliminates,
all questions and uncertainties about points such as: dimensions of the
structure, stress raisers, and production techniques. Something similar is
achieved by varying-amplitude or some other type of programme loading.
There are, however, uncertainties about using Miner's rule to calculate fatigue
lives (Chapter 4). When the real life of a structure is 106 to 107 cycles, it is
possible to perform the full-scale test under programme loading, which
simulates, more or less, the real service loads. The test will last a reasonable
length of time, about 3-4 months.
If cracks occur in the tested structure, its performance must certainly be
increased by modifications in the fatigue critical zones, which will very likely be
revealed by the test. If no cracks occur, the confidence in its fatigue performance is undoubtedly augmented. However, for rotor blades the expected real
life lies between 108 and 109 cycles (a service life of thirty years and about 70 000
cycles per day). So real service load levels for testing are clearly not appropriate. Higher loads are needed to end the test within a few months.
Programme loading is certainly recommended; accurate life predictions will
be probable. However, such tests are always very specific. Constant-amplitude
loading may be considered to produce fatigue data of full-scale structures.
These data, combined with a good cumulative damage rule, are the bases for
life estimates when various load sequences are examined. It is not possible to
state whether the program loading or the constant-amplitude loading will
provide the best answers to the general problem of estimating fatigue lives.
Fatigue damage is a process occurring under fatigue loading. From the start
of the test, damage accumulates until failure occurs and a certain critical
amount of damage is obtained. Full-scale tests frequently end before the
occurrence of failure. This implies that during the fatigue test, several parameters must be monitored to find at least indications about the damage
process. For instance, a stress redistribution in the surroundings of a localized
damage may be detected with a strain gauge. A general damage, widespread in
the volume of the structure, might be recognized by a change of stiffness or a
modification in the hysteresis loop. This information is necessary in order to
decide whether a structure will fail in a safe manner (Chapter 15).
10.2 Experimental details
A total offivefull-scale blade roots were tested. Prior to the fatigue test, static
tests were invariably performed under modest loads to record strain gauge and

Response of Blade Roots to High Bending Moments


Table 10.1

Test programme

Specimen
reference

Flange
type

Rotor
diameter
(m)

Maximum
bending moment
(kNm)

VL5
VLPOl
VLAE1
VLAE2
IKAE

Pin-hole
Pin-hole
Trumpet
Trumpet
'T' bolt

26
26
25
25
20

462
258
300
250
108

151

displacement readings. These readings form a sort of basic reference for the
initial status of the specimen. The fatigue tests were performed under constantamplitude loading; variable amplitude loading is described in Chapter 12. At
regular intervals, measurements were made to monitor eventual changes of
displacements, strains, and temperatures. The changes might be indicators of
fatigue damage.
Two servo-controlled fatigue testing rigs were available for tests on full-scale
blade roots. The capacity of these facilities are expressed in kNm, that is the
highest admitted bending moment. One rig had a capacity of 250 kNm and the
other could be loaded up to 500 kNm. The blade roots were installed like
cantilever beams and the load was applied transversely to the beam at the free
end. Any section was subjected to the same shear, which was many times
higher than in actual service.
A summary of the test programme is presented in Table 10.1 for three types
of flanges.
10.3 Pin-hole flange
Pin-holeflangeVL5
The specimen consisted of the root part of a central 'D' spar and the complete
flange, as shown in Figs 10.1 and 10.2. Thisflangewas of the pin-hole type. The
total test length was about 2.5 m. The connection to the hub was realized by
means of twenty-four bolts of 24 mm diameter. The pin-hole connection was
achieved with eighteen bolts of 30 mm diameter. The test spar was the central
part, without shells, of a blade used in a 26 m diameter rotor.
The material was a glass-fibre-reinforced-polyester with a specific mass of
1630 kg/m3, an 'E' modulus of 22.3 GPa, and a weight fraction W( = 50 percent.
The 'D' spar was wound by a tape winding process (Chapter 9), resulting in 80
percent UD material and 20 percent 45 degree material.
The specimen was instrumented with fifteen strain gauges in critical points
near the pin-hole area. Two displacements transducers were installed: one to
measure the overall displacement of the loading point, and the other to record
the displacement between point PS (Fig. 10.2) on the steel flange, and point PO

152

Design of Composite Structures Against Fatigue

Flange bolts M24

35mm

Fig. 10.1

Flange connection of VL5 rotor blade, pin-hole construction; location of strain gauges
R1-R8, K5-K10; dimensions in mm

^ GRP spar

Outer steel flange

Fig. 10.2

Sectional view of root flange VL5

Response of Blade Roots to High Bending Moments

153

on the composite. It was very likely that the latter would sense the fitness of the
pinhole structure.
The static test was performed under a maximum bending moment of 231
kNm. The maximum positive strain in the theoretically most stressed fibre was
577/strain; the highest negative strain in the opposite point was 633/strain.
These strains were measured in the : = 285 mm plane (Fig. 10.2), some 15 mm
above the upper edge of the flange collarpart.
The fatigue test was carried out under constantamplitude loading varying
from 23 kNm to a maximum of 231 kNm. This test was stopped after five million
cycles. The effect of these five million cycles, with regard to fatigue damage,
was apparently insignificant. The observations revealed: no visual damage, no
change in stiffness (load divided by displacement), no change of the surface of
the hysteresis loop in a loaddisplacement diagram, and no substantial changes
in the strain distribution as monitored by fifteen strain gauges. Though a
thermographic picture revealed a zone of about 150 150 mm with higher
surface temperatures, this zone was not very close to the pinhole area.
To gain more information about the fatigue strength, it was decided to
extend the test under a sensible higher loading. The specimen was moved to the
500 kNm rig, shown in Fig. 10.3. A static test showed that a continuation of the
fatigue test could be achieved under a maximum moment of 462 kNm. This
moment gave rise to the following: an overall deflection at the load point of
20.2 mm; a maximum positive strain of 1112//strain, and a maximum negative
strain of 1213 //strain (measured in the = 285 mm plane).
The second step of the fatigue test was performed under constantamplitude
loading, varying from 46 kNm to 462 kNm. After about 100000 cycles, a small
damaged area appeared in the GRP material. It was located on the tension side
in the loadingplane at about 620 mm from the flange. This local damage did
not influence the monitored parameters. Indeed, the strain gauges were not
situated in its vicinity, they were much closer to the flange. The damage was
apparently too low to influence the overall deflection. The surface of the
damaged area steadily increased. After 803000 cycles, the fatigue test was
temporarily stopped to perform a second static test at high bending moment. In
tension, 1116//strain, and in compression, 1182//strain, were found: practically
the same values as in the first test. The strain gauge readings from other
locations demonstrated the same behaviour. The overall deflection, however,
was slightly higher i.e. 20.5 mm compared to 20.2 mm in the first test. This
difference cannot be considered as significant for damage; rather, it might be
due to scatter.
Finally, at 1 727 589 cycles, the testing machine was automatically shut down,
because the maximum allowed displacement was exceeded. The final monitor
ing was performed at 1 720000 cycles. The measured variables did not show
significant variations.
A third static test could not be properly finished because a fracture propaga
tion, originating from the damaged area, was produced before reaching the

154

Design of Composite Structures Against Fatigue

Beam

a
-I-Beam
Load point

2
Fig. 10.3

Test rig for fatigue testing of blade roots up to 500 kNm

bending moment of 462 kNm. However, from the record at lower levels, the
actual stiffness could be obtained. This revealed a decrease of stiffness from
11.4 kN/mm (first static test) over 11.1 kN/mm (second test) to 10.3 kN/mm
(third test after the shut down). This means that a few cycles before final
rupture, the stiffness showed a reduction of 10 percent, whilst a few thousand
cycles before final rupture, the loss of stiffness was not significant at all.
The fracture is shown in Fig. 10.4. Surprisingly, the failure did not occur in
the pin-hole connection, but in the plain GRP material. The damage was not
properly monitored because the fracture area was anticipated to lie much
closer to the flange.
Pin-hole flange VLPO
The specimen consisted of a 4 m long root section, cut from a blade which was
fabricated according to the spar-shell type. The spar was practically identical to
the previous spar VL5. The fatigue test was aimed primarily to study the
bonding between the spar and shell.
The structure was instrumented with twenty-four strain gauges, as shown in

Response of Blade Roots to High Bending Moments

155

Fig. 10.4 Fracture in VL5 component

Fig. 10.5. At the load point, situated at r = 3540 mm, a displacement


transducer was installed to measure the overall deflection.
A general view of the specimen mounted in the 250 kNm test rig is shown in
Fig. 10.6.
The following results were obtained in a static test loaded to a bending
moment of 248 kNm. At point 2 (Fig. 10.5) a negative strain of 1772//strain was
recorded; at the opposite point 15, only 1108//strain was found. These results
demonstrate that the actual spar is sensibly less stiff than the spar VL5.
Relatively important strains, in the range of 1000 to 1300 //strain, were
measured in the points 5, 8, 9,17, and 18.
The fatigue test was carried out under constant-amplitude loading varying
from 13 kNm to 258 kNm. The ranges of the strains at some interesting points
are listed in Table 10.2.
The results listed in Table 10.2 demonstrate that strain redistributions took
place from the very beginning of the test. Between 40000 and 50000 cycles,
very important changes were recorded in point 22. After 60000 cycles, the

Design of Composite Structures Against Fatigue

156

3540
3400

2600

Concave side

900

Convex side

Detail A

Fig. 10.5

Location of strain gauges on VLPO component; r = 0 mm is axis of rotor

strain gauge amplifier of point 22 was saturated. The test was stopped for a
thorough visual inspection. A longitudinal crack in the leading edge at r = 3540
mm was observed. That crack was clearly due to a badly-designed load
transferring interface. The applied load acted almost as a point load; the spar
inside the blade was not able to keep the section undeformed. Unwanted local
deformations took place in the sections close to the loading point, which led to
the formation of the observed crack.
The repair and stiffening of the damaged sections were not complicated, but
it was decided to terminate the test.
10.4 Trumpet flange
Trumpet flange VLAE1
The specimen consists of a 2.5 m long blade root. A sectional view of the flange
is shown in Fig. 10.7. The flange structure was of the trumpet type, aluminium
being selected for the metallic part. The connection to the hub was performed
by means of twenty-four bolts of 24 mm diameter. The material around the bolt
holes was supported by means of 'plastic-collapse-tubes' of steel. The blade
root was manufactured by contact moulding in the same way as an entire blade,
though special arrangements were made for the load introduction. The blade

157

Response of Blade Roots to High Bending Moments

Fig. 10.6

General view of the VLPO component in test rig

Table 10.2 Strain ranges during fatigue of VLPOl


Number of
cycles

Point 15

0
10000
20000
30000
40000
50000
60000

1108
919
903
906
906
909
907

Point 16

Strains
( strain)
Point 18

Point 19

Point 22

731
793
783
782
784
792
794

1337
1209
1200
1206
1218
1221
1225

768
714
696
702
706
700
662

299
250
382
309
146
1020
1980

158

Design of Composite Structures Against Fatigue

550
Sections
AA, [BB]

Outer flange

Plastic collapse tubes

Fig. 10.7

Sectional view of the flange of component VLAE1 showing location of strain gauges

was of a stiffened shell type (Fig. 7.2) and normally serves in rotors of 25 m
diameter.
The material was a glass-fibre-reinforced-polyester consisting of 75 percent
UD fabric (56 percent fibre fraction by weight and 600 g/m2; E = 30 GPa, and
specific mass = 1.72 kg/km3), and 25 percent randomly oriented CSM (32
percent fibre fraction by weight and 300 g/m2; E = 7.7 GPa and specific mass =
1.42 kg/dm3).
The component was instrumented with seven strain gauges and two displace
ment transducers (Fig. 10.7). The location of the strain gauges is described in
Table 10.3. One displacement transducer measured the overall deflection and
the other recorded the displacements between the aluminium flange and the
adjacent GRP material at the 0 degree position. A general view of the specimen
mounted vertically in the 500 kNm rig is shown in Fig. 10.8. It is worth noting
here that the plane of loading coincides with the plane containing the joints
between the shells.
The static test was performed under a maximum bending moment of 300
kNm. The measured strains are not very consistent (Table 10.3); this might be
due to the fact that several strain gauges are actually located on the longitudinal
bonds.
The fatigue test was carried out under constant-amplitude loading ranging

Response of Blade Roots to High Bending Moments

159

Table 10.3 Static strains in VLAE1 under a bending moment


of 300 kNm and after repair at 250 kNm
Strain
gauge

Orientation
(degrees)

Strain (300)
(M)

Strain (250)
()

1
2
3
4
5
6

0,AA
180.AA
330.AA
30,AA
0,BB
180.BB

1880
-3310
2662
2039
1993
-2787

2099
-2808
1489
1471
968
-1258

Note: Gauges located circumferentially within sections AA


and BB: a few mm and 150 mm above steel flange (Fig.
10.7). Gauge 7 located on a rib of the aluminium flange
adjacent to bolt 20.

from 100 kNm to 300 kNm. In Table 10.4, the strains recorded during fatigue
are listed; the readings are in terms of load ranges.
The results shown in Table 10.4, reveal the occurrence of a serious strain
redistribution which does not affect the overall deflection. In the region of
compressive stress, the strains remain virtually constant (strains 2 and 6).
Between 5000 and 20000 cycles, an important drop is observed in the points 1
and 5. This drop is accompanied by an increase in the nearby points 3 and 4. At
the same time, the relative displacement decreases in the boundary between
the metal flange and the GRP material. A visual inspection revealed a
longitudinal crack along the bond at the side subjected to tensile stresses.
The flange region is not really axi-symmetric as it contains two joints. In real
service conditions, the most stressed fibres are not along the joints. Hence, it
was rather unjustified to test the root section under a bending moment which
induced the highest strains in these joints.
After repair and re-arrangement of the test rig, the specimen was re-installed
in a flat position (Fig. 10.9). The specimen was re-instrumented with strain
gauges placed in the same configuration as previously.
The static test was carried out under a maximum bending moment of 250
kNm, this being a reduction of 50 kNm compared to the previous situation.
These strains (Table 10.3) are more consistent than earlier, though the
negative strains 2 and 6 are sensibly higher than the positive strains in the
opposite points.
The fatigue test was performed under constant-amplitude loading varying
from 50 to 250 kNm. The strain gauges showed unexpected but analogous
responses as during the previous test. Between 5000 and 10000 cycles, very
serious changes occurred in strain gauges nos 1 and 3 (Fig. 10.10). The range of
no 1 dropped from 1540 to 673//strain, and that of no 3 from 1082 to 94//strain.
These drops were not balanced by increases in the other strain gauges,
although the number of strain gauges were too limited to find evidence of such
compensations.

160

Design of Composite Structures Against Fatigue

Fig. 10.8

Table 10.4

Number of cycles
Moment (kNm)
Displacement (mm)
Relative displacement (mm)
strain in no. 1
/( strain in no. 2
// strain in no. 3
strain in no. 4
strain in no. 5
strain in no. 6

Component VLAE1 mounted vertically in rig

Strain range during the fatigue of VLAE1

1000

5000

20000

30000

40000

45000

194.5
18.2
0.62
1108
1820
1569
1261
1211
1789

195.0
18.1
0.61
1132
1809
1579
1270
1226
1786

192.5
18.2
0.56
634
1787
1787
1424
848
1783

192.5
18.2
0.54
603
1787
1800
1452
811
1786

194.0
18.1
0.54
444
1799
1835
1475
805
1794

192.5
18.2
0.46
319
1789
2067
1682
598
1791

Response of Blade Roots to High Bending Moments

Fig. 10.9

161

Component VLAE1 mounted horizontally in rig

From 10 000 to 200 000 cycles, the strain range in no. 1 increased to a level of
1120 //strain. In this period, the strain range in no. 3 remained practically
constant at about 100//strain. From 200000 cycles, changes took place in nos 1
and 4, whilst the overall deflection did not show significant variation. After
230000 cycles, the test was stopped for a thorough visual inspection. Four
unexpected cracks were found in the upper part of the flange. Two crack paths
were apparently located in radial planes. Each plane ran straight through a bolt
hole. The sketch shown in Fig. 10.11 indicates the positions of these cracks (the
centre lines of the bolt holes lie in the plane of the paper, and each radical plane
is depicted perpendicular to that plane). The third and fourth crack were
observed in point (see Fig. 10.7) close to the ribs of bolt 20 and bolt 22,
respectively. The two latter cracks may be the extensions of the two main
cracks present in Fig. 10.11.
A change of colour was observed in the GRP material confined to the area

162

Design of Composite Structures Against Fatigue

IO4

IO5
Cycles
2.0

Gauge 4
1.5, F

" " ^ M

TS 1.0

3.0.5
IO5

Vf
Cycles

Q
Q I

'iff

o~oKX)oqiFn>cio
I
I _
Vf
Vf
Cycles

10'

Fig. 10.10 Maximum and minimum strains at gauges 1, 2, 3 and 4 of component VL AE1 being
cycled from 50 to 250 kNm

Gap (outer flange consists


f of two halves)
Inner flange
GRP
23
22
21
20
Outer flange
(Bolt number)
Fig. 10.11 L ocation of cracks in the flange of VLAE1 component, and location of gauges 1 , 3 , 4 , 5

Response of Blade Roots to High Bending Moments

"

IO4

Cycles

Vf

Vf

IO3

10'
Vf
- Cycles

163

10'

104

10'
Vf
Vf
w
Cycles
Cycles
Fig. 10.12 Maximum and minimum strains of gauges 1, 2, 3 and 4 of component VLAE2 being
cycled from 50 to 250 kNm
10'

adjacent to strain gauge no. 1. This might be an indication of a local delamina


tion. In such areas, the strains measured outside may be very different from
those inside. This may be the reason why a measured decrease of strain was not
compensated by an increase at other points.
Trumpet flange VLAE2
The component VLAE2 was identical to component VLAE1, but the 'plasticcollapse-tubes' were excluded (Fig. 10.7).
The specimen was mounted in the flat position and instrumented with seven
strain gauges and two displacement transducers in the same way as VLAE1.
The strain gauge readings in the static test were almost identical to those of
VLAE1 with the exception of strain gauge no 2: in the former test, 2808//strain
(compression) was found, and now 1854//strain was prsent, which is closer to
the 1971 //strain at the opposite point.
The fatigue test was performed under constant-amplitude loading, ranging
from 50 to 250 kNm. Again, as observed in previous tests, very similar strain
redistributions occurred between 10000 and 20000 cycles, as shown in Fig.
10.12. The range of gauge 1 dropped from 1470 to 786//strain, and that of gauge
3 from 990 to 186//strain. Further, the overall deflection remained constant. At
this stage, no cracks could be observed in the component, and the test was not

164

Design of Composite Structures Against Fatigue

Fig. 10.13

Location of cracks in the flange of VLAE2 component

interrupted. From 20 000 cycles, several changes of strains took place, which is
an indication of the evolution of a certain process. After 255 000 cycles, a small
crack was observed in P, very close to the rib of bolt 20 (Figs 10.7 and 10.11). As
shown in Fig. 10.13, this rib was instrumented with strain gauge 7 which,
indeed, showed clear evidence of neighbouring changes in the material.
It was believed that the detrimental effect of cracks in the upper part of the
flange was negligible. To test this hypothesis, it was decided to continue the test
at a reduced load varying from 40 to 200 kNm. This test was stopped after
5 000 000 cycles, and some results are presented in Fig. 10.14. The ranges of the
overall deflection (18.22 mm) and the local displacement (0.58 mm) remained
essentially constant. The same can be said about the strains in gauge 2 (979
//strain) and in gauge 6 (745 //strain), which were located at the compression
side. In gauge 1 and gauge 5, the strains, still in terms of ranges, showed final
reductions of about 6 to 8 percent as compared to their initial values. The strain
in gauge 3 stayed remarkably low (about 73 //strain). In the rib (gauge 7), a
rather rapid decrease of strain started at about 2 500 000 cycles from 372//strain
to 287 //strain. At the same time, a crack appeared in the upper part of the
flange between bolt nos 21 and 22. This crack slowly progressed during the
remainder of the test. It traversed an auxiliary hole. Apparently this crack did
not affect the behaviour of the root part.

165

Response of Blade Roots to High Bending Moments

12

*co-o-oooooo-

0i l

Fig. 10.14

10'

I I I I I ll

10'

Vf
Cycles

10'

10'

Strain ranges in VLAE2 for gauges 1-7; second load increment after 5,000,000 cycles

10.5 T-bolt flange


T-boltflangeIKEA1
The test specimen was the root part of a prestressed T-bolt type blade (IKEA).
The length was about 2.5 m and the sectional dimensions, locations of bolts and
strain gauges are shown in Fig. 10.15. The connection to the hub is done with
twenty-four stud bolts of 12 mm. Therefore, twenty-four special cylindrical
nuts (see Fig. 10.15) were passed through holes, drilled radially in the wall.
The material was a glass-fibre-reinforced-polyester consisting of 50 percent
of a balanced 0/90 degree fabric (E = 18 GPa) and 50 percent of 45 degree
fabric (E = 9.3 GPa).
The specimen was instrumented with six strain gauges and one displacement
transducer to measure the overall displacement.
When the bolts were fastened (torque = 60 Nm) to the desired prestress, the
balance between inside- and outside-strain was very good (Table 10.5).
A static test revealed, at a bending moment of 90 kNm, the large difference
between 1 and 5, which is surprising, and the spread of 2 and 6 was acceptable.
According to simple beam theory, the measured strains are on the low side,
which is typical for prestressed joints.
The fatigue test was carried out under constant amplitude loading, varying
from 30 to 90 kNm. To control the prestress, which is of utmost importance for
the joint, the test was halted after 103000 cycles. No loss of prestress was

Design of Composite Structures Against Fatigue

166

I Compression side
JA

Cylindrical nut

M holes

Tension side

View

-/Cylindrical nut

GRP blade

SH7
V//////ernten?

f^^pyzzp^za
^
\
\
\
M
GRP
XW>\N
b)ade

2XZ^

Prestressed steel bolt

Flange
(c)

Section AA
Fig. 10.15 Flange of IKAE1 component showing location of six strain gauges

observed; the changes were within 3 percent. The same controls, yielding the
same result, were done after 820 000 cycles and after 1 325 000 cycles. Finally,
after 3110 000 cycles, the test was stopped. Apparently, no damage was caused
during more than three million cycles. The load was increased by 20 percent
from 90 to 108 kNm and the minimum from 30 to 36 kNm. After 315 000
additional cycles, one of the most stressed bolts, located between strain gauges

Response of Blade Roots to High Bending Moments

167

Table 10.5 Strains in T-bolt flange after tightening


bolts and under a bending moment of 90 kNm
Strain gauge

Pre-stress
()

Storie testing
()

1
2
3
4
5
6

-1990
-1797
nm
nm
-2058
-1780

-1080
488
-1206
1126
-200
378

2 and 6, was broken. The fracture occurred in the thread at the side of the
internal cylindrical nut. This crack was clearly announced by the strain gauges 2
and 6. As the bolt lasted nearly 3.1 million cycles in the first part of the test, it
cannot be excluded that a part of the damage was induced during that first part.
New bolts were installed and the test was restarted under the higher loading.
An automatic shut-down occurred after 380 000 added cycles. Three bolts were
broken; one at the same position as in the preceding case, the two others on
either side of that bolt. These fractures were clearly sensed by strain gauges 2,6
and 4. Two fractures were located at the outside nuts, one at the internal
cylindrical nut.
After repair, the test was restarted under the same high loading. After
360000 added cycles, a fracture occurred again in the bolt between gauges 2
and 6. The fracture was sensed as in the former cases. The fracture was at the
inside nut.
10.6 Discussion
Five full-scale components were tested in two steps. Why this second step is
worth the added effort is explained further by means of two examples.
A rotor blade of good design must be able to withstand a certain static load as
well as a certain fatigue load. Various types of load-time histories may be
chosen for fatigue testing. The simplest loading is the classic constantamplitude loading, the selection of a proper magnitude of the loading being a
difficult step. Normally, it turns out that the time within which the test must be
finalized is a stringent factor. This means that the number of cycles N, after
which the test must be halted, is predetermined. must be high enough to
permit a fair chance of locating the fatigue critical points, as in realistic service
conditions. However, even very high numbers of cycles may fail to locate these
critical points. Indeed, this is the weak point of constant-amplitude testing,
where detrimental and beneficial effects resulting from the load-time history,
are ignored.
In the field of rotor blades, there are indications of a trend to keep
in the order of 1 to 10 million cycles. In this work, the suggestions given in

168

Design of Composite Structures Against Fatigue

reference (1) were partially adopted for the test on the VL 5 component. The
load is based on a thrust load of 300 P times the swept area of the turbine.
Here, a load somewhat higher was chosen and the pre-determined number of
cycles was set to five million.
The test on VL 5 progressed up to the predetermined five million cycles and
apparently no damage was induced. So it is clear that the blade fulfils the design
options. However, nothing is known about the remanent fatigue life nor the
location of the fatigue critical zone. More information may be gained by a
consecutive second step under a higher load, until failure occurs.
This supplementary information will not be precise, because many sources of
inaccuracy have to be introduced. Firstly, the shape and position of the S-N
curve, utilized as design curve, is required. A convenient analytical definition
of the S-N curve is:
o = S0(l-

Dlo
g N)

which represents a linear law in an 5-log/Vdiagram. So is a constant depending


on the static strength. D, the so-called degradation factor, is mainly in the
range of 0.08 to 0.12. A typical value for D is 0.1, as used by Mandell (2). When
results of tests are available, better estimates of D could be used when dealing
with components of a similar type.
Secondly, a fatigue damage accumulation law is required. Ignoring sequence
effects, it is generally recognized that Miner's rule is a convenient tool to yield
results which may serve as valuable estimates (Chapter 4).
When, for instance, the specimen fails in the second step after n2 cycles, an
S-N line may be defined which represents the failure S-N line of the specimen.
This line, compared to the design S-N line, defines the remanent fatigue life
after fulfilling the design options.
The failure S-N curve is defined by
rti//V] + n2/N2 = Miner's sum
tr, = o u (1 - D log Ni)

o2 =

ou(l-D\o
g N2)

where: n\ is the number of cycles in the first step under a stress 0\',
n2 is the number of cycles to failure in the second step under a stress o2;
N\ and N2 are the numbers of cycles to failure in a one step test under a load of,
respectively, and a2;
ou is the ultimate failure stress.
An appropriate estimate of o2 may be determined by a good estimate of the
ratio n\IN\ and a minimum requirement of the numbers of cycles in the second
step. It seems reasonable to ensure that one million cycles are attained in the
second step.

Response of Blade Roots to High Bending Moments

169

Table 10.6 Estimates of damage in component VL5

Miner's
sum

Damage
first step
(%)

Damage
second step
(%)

0.1
1
10

11
3.7
1.2

89
96.3
98.8

The results for VL5 (first step five million at 231 kNm, second step 1.728
million at 462 kNm) are listed in Table 10.6. The analysis, although approximate, provides useful information about damage.
The second example concerns the specimen IKAE1. This specimen lasted
3.11 million cycles under a constant-amplitude loading ranging from 30 to 90
kNm, without any apparent damage. A second step was envisaged under a
higher load to obtain a failure. This step was performed under a load ranging
from 36 to 108 kNm. After 315000 additional cycles, failure occurred in a
prestressed bolt. The situation here differs from that in the previous specimen,
because the failure is not in the GRP material. The damage calculations may
not be performed as the loading in the bolts is not known with sufficient
accuracy. Indeed, besides axial loading, the bolts may experience bending
moments. Nevertheless, the second step gives additional information about the
location of fatigue critical points.
Various techniques exist to evaluate damage accumulation with respect to
mechanical performance. When a damage process which is widespread
throughout the component is expected, overall measurements such as overall
deflection (stiffness) and the area of the hysteresis loop (damping), might be
suitable. The stiffness of all components and the area of hysteresis loop of some
of them were fully monitored. However, no definite damage could be observed, not even in the case of a measurement taken a few thousand cycles
before final failure.
Thermographic methods have been used in an explorative way. Heat
emission from damaged areas can easily be monitored. At a frequency of about
2 Hz, temperature differences of nearly 4C were observed. An alternative is to
install thermocouples. This was not successful, because the location was poorly
anticipated.
Strain gauges are very sensitive but the interpretation in terms of damage is
exacting. Redistributions of strains, for instance, are not always indicators of
fatigue damage.
Fatigue testing does not always lead to failure of the entire component. Some
load bearing parts can show fatigue cracks apparently without affecting the
performance of the component. This was demonstrated in the test on VLAE2.
The choice of a suitable failure criterion can therefore be difficult.

170

Design of Composite Structures Against Fatigue

10.7 Conclusions
Experience gained from performing full-scale tests on blade roots has shown
that damage monitoring is very exacting. There are various methods and each
will provide some information about damage accumulation. Testing in two
steps may perhaps provide the best information about damage build-up when a
failure is induced during the second step. Then, a fair chance exists to locate the
fatigue critical area and to obtain estimates about damage endured in each
step.
It is difficult to define the fatigue life of a full-scale component, as cracks, for
example, may appear in some parts of the flanges without affecting the load
bearing capacity in fatigue. Indeed, after the presence of such cracks, one
fatigue test was not stopped and after more than four million additional load
cycles, the component still performed in a satisfactory way.
References
(1) JENSEN, P. H., KROGSGAARD, J., LUNDSAGER, F., and RASMUSSEN, F., 1986,
Fatigue testing of wind turbine blades, European Wind Energy Association Conference and
Exhibition, Rome.
(2) MANDELL, J. F., 1975, Fatigue crack propagation rates in woven and nonwovenfiberglass
laminates, (ASTM, Philadelphia, USA).

CHAPTER 11

Influence of Moisture on GFRP Bolted


Joints
P. W. Bach*
In order to investigate the influence of moisture, fatigue tests on preconditioned GFRP bolted
joints are performed at 100 percent RH and compared to dry GFRP bolted joints.
At the highest stress level a reduction in lifetime of about one order of magnitude is observed
for the humid specimens. The reduction in fatigue strength by the moisture decreases with stress
level. The results of the variable amplitude WISPER tests are in accordance with the diminishing
effect of moisture in the low stress range, as found in a literature survey and observed in constant
amplitude tests.

11.1 Introduction
The effect of environment is important in many applications particularly if the
component is load bearing. As observed in Chapter 3, stress can lead to the
formation of cracks in GFRP, which in turn could allow the ingress of water.
Other environmental effects such as impact and abrasion due to hailstones are
discussed in Chapter 5.
Probably the most fatigue-critical part of a wind turbine is the rotor blade
connection to the hub, for which a bolted joint is often used. Although GFRP
blades are usually well protected against weathering by a gel-coat, moisture
may still find its way to the joint. It is, therefore, important to investigate the
sensitivity of bolted joints to moisture.
The design strength and stiffness of rotor blades must be maintained during
the lifetime of a wind turbine. Therefore, special attention was also paid to
monitoring the stiffness degradation of the specimens.
11.2 Effect of moisture
Although GFRP is generally found to have good weathering properties, the
effects of moisture on the fatigue properties are not negligible (1).
The water absorption process can be described as follows:
(a) diffusion of water molecules through the matrix to the glass/resin interface ;
*ECN, The Netherlands.

172
Table 11.1

Design of Composite Structures Against Fatigue


Indication of the effect of moisture on the mechanical properties of FRP. All data
in % (1)

Material
Glass/polyester
Glass/epoxy
Carbon/polyester
Carbon/epoxy
Glass/carbon epoxy

Maximum
moisture
absorption
(% H-r)

Stiffness
change

4
2

-10
-10

1.5
<2

(%)

UTS change

(%)

-15
-10
-5
-2
=3

Fatigue
strength
reduction
(103 cycles)

Fatigue
strength
reduction
(107 cycles)

-35
-20

-10
-5

(b) weakening of the interfacial fibre matrix bond through chemical and
physical attack on the coupling agent;
(c) exposure of the glass fibre surface following partial dissolution of the
coupling agent;
(d) chemical attack on the glass surfaces from which alkali silicates and silica
molecules are leached out.
The major effects of absorbed moisture on the laminates are:
(a) lower glass transition temperature of the matrix resin;
(b) damage of the interface between the fibres and the resin;
(c) reduction of the cure induced residual stresses through swelling which may
retard failure.
Hence the effects of moisture on the properties are not all negative.
Especially when loaded in a humid environment glass fibres are susceptible
to subcriticai crack growth, a phenomenon which is often referred to as static
fatigue. This will influence the time-under-load dependent properties such as
creep and cyclic fatigue.
Also some synergy with moisture and fatigue can be expected. A composite
subjected to water and fatigue loads will show increased matrix cracking which
will accelerate the uptake of water.
From a literature survey it was found that the effects of moisture absorbtion
on the mechanical properties of GFRPs are not known accurately (1). This is
mainly due to the large scatter in results which may sometimes be inconsistent.
The effect of moisture on the fatigue properties for both glass/polyester and
glass/epoxy laminates diminishes in the very high cycle and low stress range.
For wind turbine applications this is a positive observation.
Table 11.1 lists the effects of moisture on the mechanical properties of fibre
reinforced plastics, as estimated from the survey. From the viewpoint of known
moisture effects the combination of glass/epoxy or carbon/epoxy gives the best
fatigue properties.

Influence of Moisture on GFRP Bolted Joints


Table 11.2

173

Laminate lay-up and material

Lay-up

+ -45/3X0/+-45/1.5X0/S

Combination-fabric
+ or - 45

700 gm2 UD + 100 gm2 CSM


70 gm2

Ahlstrom 700/100
Ahlstrom R24-700-45

Resin

Iso-polyester

Synolite 593-E-l (DSM)

11.3 Materials and methods


Glass-polyester bolted joints
Glass polyester laminates 15 mm thick were fabricated and subsequently
bolted together. The materials were identical to those in Chapter 6 - the lay-up
is listed in Table 1-1.2. The laminate contained nine layers of combination fabric
and eight layers of 45 degree glass fibres in a symmetrical lay-up. The plate
was produced under industrial conditions by hand lay-up with a glass/polyester
ratio of 53% wt (33 v/o) (also see Tables 6.1 and 6.2).
In the specimens four holes were drilled with a diameter of 16 mm and with
an H8 fit. Figure 11.1 shows the details of the bolted specimens. It is well known
that full bearing strength in bolted composites is only attained when the width/
diameter and the end/diameter ratios are above minimum value (3). The
ratios used (w/d=3.5 and e/d=4) are above this minimum value.
The performance of bolted joints can be improved when lateral constraint is
applied, as is usual in the connection to the hub. Therefore, in order to fit
precisely in the steel flange, the specimens were thickened around the holes
from 15 mm to a total thickness of 20 mm with Synolite TP 347 VG epoxy(UTS
= 7 MPa). The ultimate tensile strength of dry specimens was measured with a
loading rate of about 10 kN/sec. The UTS was 118 kN per bolt hole, which
means a bearing stress of 490 MPa (3).
Constant amplitude tests
The constant amplitude tests were performed on load controlled servohydraulic testing machines (Instron 1343) with tension-tension or tensioncompression loading (stress ratio R = omin/omax = 0.1, 1, or - 0 . 5 ) . During
fatigue tests the fibre reinforced plastic specimens experience hysterectic
heating, the extent of which depends on the stress amplitude and the frequency. Due to the expected high cycle fatigue life, the tests of the coupon
specimens were performed with a frequency up to 5 Hz. The temperature of the
specimen was kept low with a cooled air flow of 15-18C.
After measurement of the initial stiffness, the stiffness degradation was
determined on-line during the test at regular intervals (100-10 000 cycles). The
stiffness was calculated as the difference in load divided by the difference in

,15,

Specimen

^.
"-*

2
2S

l l
filili s

y '/

\
\
\
\

/
/
/
/

a
S

ro

S2
to'
3
O

Steel plate
56.0

Steel plate

S!

>

-j-e

170

10

10

(a)

(b)

40
*

Fig. 11.1

c?
3

ro

2
o
c

17.

Width

',
/ HIHI
-c

136

Geometry of the GFRP bolted joint specimen

Influence of Moisture on GFRP Bolted Joints

175

displacement of the specimen as measured by the displacement transducer of


the actuator. The stiffness degradation of the specimens influences the load
performance in the closed-loop servo-hydraulic system. In order to have a
constant load until failure, the load was corrected on-line during the fatigue life
of a specimen.
Variable amplitude tests
Current design procedures for wind turbines are primarily based oil constant
amplitude data. However, this simple kind of loading is usually insufficient to
represent the interaction effects of high and low cycles that occur in a realistic
type of loading.
Especially in the aerospace field, variable amplitude fatigue testing is widely
used and standard loading sequences have been developed in order to predict
the material fatigue behaviour under realistic conditions. The WISPER spectrum represents a mean two-month period of flapwise loading of an imaginary
wind turbine blade (Chapter 4). As a result, WISPER does not simulate typical
loading in a blade of a particular turbine in a particular wind regime to the
required standard. This means that it can never be used instead of a design
spectrum.
The WISPER load sequence is intended for comparative purposes only, i.e.,
to evaluate materials, structural details, dimensions, design alternatives, and
lifetime prediction methods.
Testing
The variable amplitude tests were performed on load-controlled servohydraulic testing machines (Instron 1342 and Instron 1343) equipped with the
Harrier computer interface and a Hewlett-Packard 9816 micro-computer. In
order to control the fatigue machines in the WISPER sequence tests a special
software programme was developed. After measurement of the initial stiffness, the stiffness degradation was determined on-line at regular intervals.
It is known that for glass fibre-reinforced plastics the loading rate has an
effect on the (fatigue) strength due to the visco-elastic behaviour and subcritical crack growth. Further, the frequency and the load amplitude will affect the
heating due to hysteresis of the specimens. So there is some uncertainty in
performing WISPER tests.
In a previous investigation the influence of the loading parameters on the
WISPER fatigue lifetime was studied (4). It was observed that there was no
difference in fatigue life within the tested conditions (constant loading rate
versus constant frequency and a sine or a saw tooth as wave shape).
The WISPER fatigue tests were, therefore, performed with a constant
loading rate and a sinusoidal wave shape, because these are more realistic for
the stress cycles in a turbine blade.

Design of Composite Structures Against Fatigue

176

11.4 Results and discussion


Moisture ingress
The sensitivity of GFRP bolted joints to moisture has been investigated by
preconditioning the laminate in water at a temperature of 50C for fourteen
days and then testing under 100 percent RH. The initial conditioning of the
specimens was for several months at a temperature of about 22C and an RH of
40 percent.
The uptake of water by the preconditioning was determined by weighing the
specimens from the glass/polyester laminate. In Fig. 11.2 the curves of the
change in moisture content are shown. Curves were obtained for absorption in
water at 25C and 50C, and for desorption in an exsiccator at 25C. During the
preconditioning the moisture content in the laminate increased with 0.13% wt.
Even after 100 days in water at 50C, the ingress of moisture still continues,
though at a decreasing rate.
The desorption of the laminate shows the same kinetics as the absorption.
The bolted joint is especially sensitive to the moisture content near the hole
where an increased diffusion can take place due to edge effects and the damage
caused by the drilling of the holes (capillary action), particularly if the edges are

Time (days)"
Fig. 11.2

Absorption and desorption of water in the GFRP bolted joint laminate

10

Influence of Moisture on GFRP Bolted Joints

177

Fig. 11.3 Photograph ofa failed bolted joint. Note the elongation of the bolt holes in the direction of
the applied load

not subsequently sealed with resin. To simulate this phenomenon a 10 mm wide


strip was ground and preconditioned. The rate of moisture uptake was almost
doubled (Fig. 11.2). The change in moisture content near the holes in the
laminate was, therefore, about 0.26 percent.
The diffusion of moisture into composite materials can usually be described
by Fick's law. The curves are plotted with the square root of time on the
abscissa. For Fickian absorption a straight line can be expected when the
moisture content is far from saturation. The slope of the curves decreases
somewhat, so there is probably some deviation of the Fick's law due to edge
effects or the moisture content approaching the saturation level.
Fatigue
In order to investigate the influence of moisture, constant amplitude fatigue
tests on preconditioned GFRP bolted joints were performed at 100 percent RH
and compared to the results of dry GFRP bolted joints (Chapter 6).

Design of Composite Structures Against Fatigue

178
400

O o.i

300

0.5
0.1 drj'
+ 0.5 humid

200

E
s
E
M

10

10'

iv

IO'

10'
Cycles

to'

+ SI

10*

10'

</

io

(a)
400

O o.i
300

0.5
x 0.1 dry
+ 0.5 humid

S 200
E
3

oo

S 100

itf

i<P

IO*

vT

"ur"

10'

10"

Cycles

(b)
Fig. 11.4 Comparison of stress against cycles for dry and humid bolted joints: (a) fracture; (b) 10
percent reduction in stiffness. Runout points indicated by an arrow

Figure 11.3 shows a failed specimen. The damaged zone near the hole in the
diagonal direction is about 2 x the diameter after failure and is in accordance
with the stress distribution as predicted with a finite element method calcu
lation (5). The joint eventually failed by the degradation of the bearing
capability of the bolt hole. At half the lifetime of the specimen some damage
started to develop in the 45 degree layers of the laminate in the middle part of
the specimen.
There is some difference between the dry and preconditioned specimens in
the magnitude of the damaged zone. This zone is about 20 percent smaller for
the humid specimens.

Influence of Moisture on GFRP Bolted Joints

179

500

I 400

O Dry
+ Humid

< O O O

soo
S

O+CH- O

200

10'

\V

10'

(a)

10*
Cycles

10"

10*
Cycles

10*

10*

10'

10*

10'

10"

500
O Dry
+ Humid

400

I
J 300
3

200

(b)

10'

10"

10"

Fig. 11.5 Comparison of stress against cycles for dry and humid bolted joints using the WISPER
spectrum, (a) Fracture; (b) 10 percent reduction in stiffness

Initially, in the R = 0.1 humid tests at the lower stress level the bolts failed
due to corrosion-assisted fatigue. In a later stage zinc coated bolts were
applied. In Fig. 11.4 the number of cycles until fracture of the dry and humid
specimens versus the maximum initial bearing stress level is shown for R = 0.1
and R = -0.5. At the highest stress level a reduction in lifetime of about one
order of magnitude is observed for the humid specimens. At the lower stress
level there is no indication of the effect of moisture due to run-out of the
specimens. However, an indication of the effect can be found by the stiffness
degradation of the specimens as shown in Fig. 11.4(b). The reduction in fatigue
strength by the moisture decreases with stress level. This effect is more
pronounced for pure tensile than reversed loading.
The results of the variable amplitude WISPER tests are illustrated in Fig.
11.5 for the failure criteria fracture and 10 percent stiffness reduction. The tests
with the moisture condition were run at a rather low stress level, which has a
better predicting value for the stress levels used in practice. Within the limited
test results there is no difference in the behaviour of the dry and humid
specimens.

180

Design of Composite Structures Against Fatigue

The WISPER spectrum contains only a small minority of through zero cycles
and the fatigue behaviour will be largely governed by stress ratios R > 0
(Chapter 4). The results are, therefore, in accordance with the diminishing
effect of moisture in the low stress range, as found in the literature survey and
observed in the R = 0.1 constant amplitude tests.
11.5 Conclusions
Although glass fibre plastics are generally found to have good weathering
properties, the effects of moisture on the fatigue properties may not be
negligible.
Fatigue tests on wet preconditioned GFRP bolted joints performed at 100
percent RH at the highest stress level gave a reduction in lifetime of about one
order of magnitude. The reduction in fatigue strength caused by the moisture
ingress decreases with the stress level. This effect is more pronounced for pure
tensile loading. The results of the WISPER tests are in accordance with the
diminishing effect of moisture in the low stress range, as found in the literature
survey and observed in the R = 0.1 constant amplitude tests.
References
(1) BULDER, . . and BACH, P. W., 1991, A literature survey on the effects of moisture on the
mechanical properties of glass and carbon fibre reinforced laminates, ECN-C-91-033.
(2) GODWIN, E. W. and MATTHEWS, F. L., 1980, A review of the strength of joints.
Composites, 11, 155.
(3) BACH, P. W. and SCHAAP, . A. J., 1990, Vcrmociingsonderzock aan windturbincmaterialcn en bladvcrbindingcn, ECN-C-90-012.
(4) BACH, P. W, 1991, High cycle fatigue testing of glass fibre reinforced polyester and welded
structural details, ECN-C91-010.
(5) ROORDA, B., 1988, Analyse pen/gat verbindingen fase 2 vcrmociingsonderzock. Poly
marin, rapport R&D-880622.

CHAPTER 12

Influence of Complex Loading on


Blade-Root Joints
J. Lamris*
Although most of the material properties of glassfibre-reinforcedpolyester may be determined
from small coupons and/or joints, the effects of complex loading and the correctness of design
concepts may only be determined from large-scale fatigue tests on three-dimensional components. In this chapter fatigue tests on two types of blade-root joints arc described.

12.1 Introduction
A considerable part of the literature on fatigue tests is related to tests on
relatively small and simple coupon specimens. In such tests the material itself
rather than its shape is being investigated to determine the inherent fatigue
strength. Structural details like (prestressed) pin joints may be studied in uniaxially loaded plate specimens, but the conversion of the obtained fatigue
relationships or S-N curves is cumbersome and has a fairly low level of
reliability. In order to gain insight into the effects of complex loading, and to
make an assessment of the conversion of newly-developed fatigue relations
obtained in two-dimensional specimens, fatigue tests on large-scale components are vital. In the present study large-scale fatigue tests were conducted
on two types of blade-root concepts, each with their own objectives.
The first blade-root connection concept was a circular, prestressed, pin-hole
blade-root connection of an existing 300 kW wind turbine blade, in which the
glass fibre/polyester laminate of the 'D' spar was located between two steel
flanges (Fig. 12.1).
The main objective was to compare the effect of a variable amplitude
spectrum loading on the fatigue life with a similar blade-root connection
fatigue loaded under constant amplitude, (Chapter 10).
The rectangular blade-root connection was designed as proof of concept for
a 500 KW-type wind turbine based on the latest design ideas and certification
requirements. Primary design objectives were to obtain a relatively inexpensive blade-root connection with good maintenance and manufacturing properties. Secondary objectives were to study possible three-dimensional effects in a
bolt connection. Reference (1) describes in more detail the design assumptions
'NLR, The Netherlands.

182

Design of Composite Structures Against Fatigue

Pin and hole


bolts, 30

fl*

Load application
Fig. 12.1

Dimensions of the pin-hole b lade-root

and objectives for this rectangular blade-root design. It consists of two metal
couple plates on either side of a three meter long rectangular glassfibre/
polyester tube (Fig. 12.2) joined together by a single row of shear-loaded bolts.
12.2 Description of tests
Pin-hole flang e
Specimen shape and test configuration
This specimen consisted of the first 2.3 meter section of the 'D'-shaped spar of
an existing POLYMARIN 13 m rotor blade for a 300 kW wind turbine (Fig.
12.1). A transition from a circular to a 'D'-shaped section takes place between
270 and 1050 mm. Wall thickness varies between 40 mm at the root to 25 mm in
the 'D'-shaped section. The material for the filament wound spar was ' glass
fabric and polyester resin (Table 12.1).
The circular root section was connected to the hub by prestressed bolts or, in
the test facility, to the support rig (Fig. 12.3).
Load was applied by means of a 100 mm diameter pin through the locally
reinforced spar.
Loads

The specimen was loaded in one direction with a variable amplitude spectrum
load named WISPERX. The WISPERX spectrum is a derived version of the
WISPER spectrum, with almost identical loading properties but with shorter
testing times (Chapter 4, (2)). The maximum applied bending moment in the
WISPERX spectrum was chosen with the following considerations:

183

Influence of Complex Loading on Blade-Root Joints

3250
Left

ii

Load
application

-o

if

Plan

-I

Right

1*

A
Steel plate

GRP tub

? ui.....'..*..-.
s a * ^ * i JS3
mn
Side

-+X

Bottom Inimgaggaaa
VI,,.,
' --
' M l I
3150

(a)

19S
!
30_
to rig

24 to GRP

M u bolts

r E-SJJry^ GRP ,ube


i> I II I|
Ti-H-I i ! l-l27.5
IV
I I I I I I I I I I
-3

340
Steel couple plate

.J

i 3 3 3 3

-3

905
Section AA

(b)
Fig. 12.2

Dimensions of the rectangular bladeroot component B l

(a) identical damage development as constant amplitude fatigue tests (Chap


ter 10) and design load as calculated by the manufacturer (3).
(b) a realistic testing period.

184

Design of Composite Structures Against Fatigue


Table 12.1

Material properties

Test component A
Fibre type
Resin type
Laminate built-up
Thickness at joint
Production method

'E' glass NW 5/500


Synolite 593-E-l
(polyester)
80 percent 0 degrees,
20 percent 45 degrees
40 mm
Filament winding

Test component BMB2


'E' glass
BASF A24
(polyester)
([0 (600 g/m UD)] 2 ,[45 (400 g/m SB)3]5)S
27 mm
Hand lay-up, two halves, male mould

Fig. 12.3 Test configuration of the pin-hole blade-root

The design fatigue load was based on the current certification criterion
developed by RIS National Laboratory for fatigue loading ((4), (5)) in which
the blade should endure 107 load cycles fluctuating between zero and a fatigue
proof load without any apparent damage.
Similar to the approach in Chapter 10, this proof load was translated into a
bending moment in the flange section. Subsequently combinations of WISPERX loading sequences with maximum applied bending moment and number
of passes were calculated, giving identical damage as the RIS fatigue load,
based on the Palmgren-Miner rule and a Mandell/Appel-Olthoff S-N relation
(Chapter 4). Based on these results, a maximum bending moment of 350 kNm
and 300 WISPERX load sequences were chosen. The test was conducted in a
bi-axial vertical fatigue testing machine with one 200 kN actuator in the vertical
plane (Fig. 12.3). The test frequency was 2 Hz.

Influence of Complex Loading on Blade-Root Joints

185

Test parameters
The specimen was instrumented with nine strain gauges on the outside surface
of the spar, 220 mm above the metal flange and with six strain gauges on the
outer metal flange (Fig. 12.4). Most of the strain gauges were located in the
section which is mostly loaded under tension. In addition two bolts were
instrumented with four strain gauges to measure the stress in the prestressed
pins in the horizontal flanges.
Before and after the fatigue test, the specimen was subjected to a static test to
about 60 percent of the maximum range size occurring in the WISPERX load
sequence (117 kN and 60 kN) in order to measure the following:
- strains at selected places in the GRP laminate and in the steel parts;
- the stiffness of the component;
- the stress variation, due to external loading, in the preloaded bolts of the
flange connection.
In the course of the fatigue test periodical measurements were performed at
three load levels: +100 kN, 50 kN, (the extremes of the most often occurring
WISPERX cycles), and at zero load. Measurements were also made for the
vertical deflection of the spar at the point of load application.
Rectangular blade-root
Shape
Since the component was designed as a study for an improved blade-root
connection, the structural design was kept simple. It consisted of a rectangular
cross-section with internal dimensions 800 x 400 mm, and a total length of
3.06 m, representing a 80 percent scaled model of a 500 kW type wind turbine
(Fig. 12.2). An overall laminate thickness of 27 mm was kept through the entire
length for reasons of simplicity. Manufacture of GRP component was carried
out using conventional hand-layup techniques in combination with a mould.
The materials are listed in Table 12.1.
The blade-root connection consists of a double lap-joint with two 15 mm
steel plates with a single row of shear-loaded bolts at the top and bottom side of
the rectangular spar (Fig. 12.2). Design data for the bolted joints were derived
from coupon test results (Chapter 6). The support rig was adapted with a
special open-ended box bolted onto the rig (Fig. 12.5). Loads in two perpendicular directions were applied by two hydraulic actuators and introduced into
the specimen via an internal rib at a 3.0 m distance from the bolt-line of the
blade-root connection.
Loads
The design was loaded in two perpendicular directions (flap and lead-lag), at
constant amplitude at different levels in two different tests, each on a separate
specimen (Bl and B2). The biaxial loads were applied with the loads in phase.

186

Design of Composite Structures Against Fatigue

Bolt IO

Outer flange

Flange bolts

Pin and hole bolts

(a)

GRP spar

? Inner flange

/ 1 B3.B4
B1.B2.K8

i S Prestressed bolt
^ <*30

753
840

(b)
Fig. 12.4

Strain gauge instrumentation on pin-hole blade-root

187

Influence of Complex Loading on Blade-Root Joints

Fig. 12.5

Test configuration of the rectangular bladeroot component B l

Table 12.2

Specimen
Dimensions
Mfli,p. kNm
\-., kNm
Duration
(planned)
Duration
(actual)

Components and loading properties

Test A

Test Bl

Test B2 mended

Test Bl conducted

Bl

B2

0 84() mm (at flange) X 27(K)


mm
350
250 WISPERX sequences

800 4(X) 3060


mm
151 195
60 118
5.10 s cycles

8(X) X 4(X) X 3060


mm
109 141
4484
5.10* cycles

B2
(modified)
800 X 4<X) X 3060
mm. modified
104 135
42 8 2
1.I05 cycles

300 WISPERX sequences

100000 cycles

3(XK) cycles

Loads were derived from an actual design for a 500 kW wind turbine, scaled
down to the 80 percent scale model and in such a way that the lifetime could be
calculated for 500000 and 5 000000 cycles, respectively, in order to obtain a
reasonable testing period. For this purpose, the R ratio for the tests were
chosen to be those responsible for the largest amount of damage. Table 12.2
summarizes the applied test loads and duration.
Test parameters
The component was extensively instrumented with twenty strain gauges on the
couple plates (Fig. 12.2) in order to determine the load transfer in the plates.

188

Design of Composite Structures Against Fatigue


106

50

100

150
200
WISPERX sequences

250

300

350

Fig. 12.6 Stiffness variation of pin-hole blade-root during WISPERX fatigue loading

Ten rosette strain gauges were placed on the outside surface of the GRP
laminate to determine the strain distribution in the plates both behind the bolts
and in the laminate.
Before and after the fatigue test, both specimens were subjected to a static
test equal to the extreme values of the load cycle in horizontal (lead-lag),
vertical (flap), and combined loading mode.
In the course of the fatigue test (with a maximum test frequency of 0.6 Hz)
periodical measurements were performed at certain load levels and in certain
loading modes (flap, lead-lag, and combined loading), and the vertical deflection of the component at the point of load application was also measured.
12.3 Pin-hole flange
The effect of the fatigue test of 300 WISPERX load sequences with a maximum
bending moment of 350 kNm was totally insignificant for the induced damage:
no visual damage, no change in stiffness, no change in the hysteresis loop or
damping (as measured before and after the fatigue test), and no significant
change in the strain distribution as measured with the strain gauges. Figure 12.6
shows the relative stiffness of the component in positive (upward) and in
negative (downward) bending mode as a function of the number of applied
WISPERX sequences.
During the test, the component made loud noises and some sliding movement was observed between the steel flange and the GRP laminate in the
highest tensile loaded area. The pre-load in the instrumented bolt in this area
decreased by about 45 percent during the test. Post test inspection showed that
the laminate exhibited some wear at the bolt locations between the laminate
and the steel flange. Figure 12.7, taken from inside bolt 10, shows an eroded
outer layer zone with thickness of about 0.2-0.5 mm.

Influence of Complex Loading on Blade-Root Joints

Fig. 12.7

189

Eroded layers in laminate, pin-hole blade-root at bolt 10 location (interface of steel


flange and laminate)

12.4 Rectangular blade-root


Test component Bl
During the fatigue test, the strain distributions in the couple plates, as well as
on the laminate surfaces at several locations, were monitored together with the
applied load and the corresponding deflection in both directions.
Figure 12.8 presents the overall stiffness degradation of the component
(including the blade-root connection) as a function of the applied cycles.
Since the overall stiffness degraded by more than 20 percent, the fatigue test
was stopped after 100000 cycles, or 20 percent of the expected lifetime.
Although stiffness measurements outside the blade-root connection started
only after 15000 cycles, the stiffness of the blade-spar itself only degraded
towards the end of the test in contrast to the overall stiffness reduction. Besides
this stiffness degradation, some strain redistribution could be observed in the
lower couple plate IV (Fig. 12.2) and in the upper surface of the laminate close
to the blade-root, (Fig. 12.9).
Some damage could be observed at the location of the most highly loaded
bolts between the coupling plates and the laminate, where erosion of the glassfibre lateral constraint/fill plates occurred. Post test evaluation of the specimen
indicated large ovalization of the holes in the laminate along the resulting load
axis, (Fig. 12.10), indicating bearing stress failure.
The probable cause of the failure is that the outer bolts were more severely
loaded than was assumed in the design calculations, which were based only on a
consideration of the combined load distribution in phase and on a relatively

190

Design of Composite Structures Against Fatigue

6
Cycles ( 10s)

Vertical load

O Vertical load, blade root


-- Horizontal load

Vertic
a l loa d, spar box
O
Combined loads
Fig. 12.8 Stiffness va ria tion of blade-root component Bl during fatigue loading

L5

L6

Strain gauges

1^50

1.0
H80

m 15.5
g 90

43

Cycles (xlO)

0100

Fig. 12.9 Stra in redistribution in lower couple pla te IV during fa tigue loa ding of bla de-root
component I a t ma ximum combined loa ding; for loca tion of stra in ga uges see
Fig. 12.2(b)

Influence of Complex Loading on Blade-Root Joints

Fig. 12.10

191

Lower surface bladeroot component Bl after fatigue test

Steel plate
Left
!.1!.:I:W

^rzrtr.

!p

y#::!>:'y

Right
GRP tube
Top
+X

Bottom
3100
Fig. 12.11

Dimensions and shape of modified bladeroot component B2

simple engineering beam approach. The effect of the adjacent sidewalls on the
load transfer of the bolts was underestimated. B ased.on these test results, the
design team decided to modify the second specimen B2.
Test component B2
Before the modification, the original and possible alternatives for the modified
structure were investigated by finite element analysis. As a result, the com
ponent was changed according to Fig. 12.11.
Parts of the sidewalls at the bladeroot connection were removed to alleviate

192

Fig. 12.12

Design of Composite Structures Against Fatigue

Edge cracks/delaminations in left upper corner of cut-out component B2 after initial


static test at maximum vertical load

the loading of the most heavily loaded bolts by 40 percent. The loads were
calculated in such a way that failure at the most heavily loaded bolts could be
expected at 105 cycles.
After conducting the first static test to the extreme values of the constant
amplitude fatigue test (see Table 12.2), it was discovered that the edge of the
laminate at the left cut-out area (Fig. 12.12) exhibited several edge cracks
between 8-10 cm width above each other and with minimal depth.
Other edge cracks were also detected before the static tests at identical
locations of the cut-out on the right side of the component, but were smaller in
number and in size. This damage was most likely caused in the modification
process by removing material in highly curved areas. Subsequent fatigue
testing caused damage immediately in the left upper area of the cut-out and a
large delamination of about 17 x 17 cm became clearly visible with normal
background lightning. A temporarily fix with clamps was not successful in
preventing further delamination extension. The fatigue test was eventually
stopped after 850 cycles.
Although the objective of this fatigue test was not met and a possible repair
with bolts would interfere with the normal load distribution, it was decided to
continue the fatigue test without repair and to study the development of the
damage. In subsequent testing additional multiple delaminations started to
grow from the other corners, leading to ultimate failure at 3000 cycles when the
upper surface completely delaminated close to the couple plates, Fig. 12.13.
Besides the visual signs of delaminations, damage development was also
noticeable via a stiffness degradation in the vertical plane, measured with
respect to the hub as well as to the rectangular box only, see Fig. 12.14.

Influence of Complex Loading on Blade-Root Joints


.

. . : . ' - . '

-/ : --' .,' ; : :.' i

193

,%,

J K ~v*issta * Sf v * * i . . w v v ' , . ^ . . . , ^

Fig. 12.13 Delaminations in upper surface of bladeroot component B2 after 3 000 cycles

12.5 Conclusions
The fatigue testing on the Polymarin bladeroot concept of a circular cross
section with prestressed pinhole joints demonstrated an excellent fatigue
behaviour in the tested conditions. Based on this test and on the tests on similar
components by the University of Gent (Chapter 10), the circular prestressed
pinhole joint seems to be overdesigned, at least in reference to the applied
RIS certification norm. Unfortunately it was not possible to extend the
fatigue test to ultimate failure due to time constraints and to test facility
limitations. A direct comparison with the extended fatigue test conducted later
on a similar test component by the University of Gent (Chapter 10) is,
therefore, not possible.
The fatigue tests on the rectangular bladeroot concept clearly indicated that
a direct extrapolation of twodimensional results is not possible and that this
particular design for a more inexpensive blade root concept was ill fated due to
lack of sufficient insight into the actual load distribution in the blade root joint.
Secondly, the failure in the second rectangular specimen was caused by damage
originated during the removal of material in the sidewalls at highly curved
areas. This emphasizes the need for the designer to carefully evaluate possible
effects of cutouts, corners, and other sources of stressconcentrations on the
strength of components.
Besides visual observations, damage developments in the two rectangular
bladeroot components could also be observed by stiffness measurements of
the specimen with respect to the rig as well by stiffness measurements of the

194

Design of Composite Structures Against Fatigue

1.05

0.95
-.

^ *

I 0.90

' .

"

0.85

. 0 ^

...

0.80
\
\

0.75
0.70

0.5

O
&
e_
O

1
1.5
Cycles (3)

S
2.5

Vertical load
Vertical load, blade root
Horizontal load
Vertical load, spar box
Combined loads

Fig. 12.14 Stiffness variation during fatigue loading of bladeroot component B2

component outside the bladeroot connection. In this way, both types of


stiffness measurements could be used to identify the two distinct damage
locations (bearing stress failure in bolt holes versus delamination failure
outside joint).
References
(1) JOOSSE P. A. and TIKKEMEIJER R., 1995, Evaluation of fatigue test rectangular blade
root, STORK Product Engineering BV report SPE95029.
(2) TEN HAVE . ., 1992, WISPER and WISPERXfinaldefinition of two standardised fatigue
loading sequences for wind turbine blades, NLR report TP 91476 L.
(3) ROORDA, B., Polymarin, private communication.
(4) JENSEN P. H., KROGSGAARD P., LUNDSAGER P. and RASMUSSEN F., 1986.
Fatigue testing of wind turbine blades, European Wind Energy Association Conference and
Exhibition.
(5) JENSEN P. H., JENSEN M. W. and MADSEN P. H., 1988, Fatigue aspects of the Danish
licensing of wind turbines, RIS National Laboratory report.

CHAPTER 13

Evaluation of T-Bolt Root Attachment


C. W. Kensche* and K. Schultest
Rotor blades of a 12 m wind turbine are taken out of operation and investigated statically and in
fatigue under laboratory conditions. One of the blades is predamaged and repaired before the
test. The blade root design is of the T-bolt type which connects the GRP-hub with the root by
means of prestressed studs screwed into steel cross nuts in the GRP. The behaviour of the
repaired structure and of the system GRP-prestressed stud under extreme loads up to 2 x 106
cycles is described. No fatigue of the glass fibre composite structures is observed. In the residual
strength tests the blades fail due to buckling.

13.1 Introduction
When obtaining fundamental information for the design of rotor blades and
light weight, as well as cost-effective, blade root connections, in addition to
tests on small-scale specimens, basic research is also necessary on whole rotor
blades. Additionally, it is interesting to observe the behaviour of a structure
which has been in operation and has been damaged and repaired again. Two
rotor blades which met all of these requirements were made available by
ENERCON for testing.
The tests were not only of interest for the blade structure itself but also for
the design of the load introduction of the blade root into the hub. The IKEA(or T-bolt) principle was applied as a light weight replacement for conventional
heavy steel flanges. This is also an important factor for economic and safety
aspects.
The aims of the experiments were to investigate:
(a) the T-bolt connection up to higher load cycle numbers than had previously
been approved;
(b) the design rigging diagram;
(c) the fatigue and failure behaviour of the composite structure.
13.2 Load attachment principle. FEA-simuIation
In the area of light weight design and safety aspects, the T-bolt (or IKEA)
connection of the blade root to the hub of a WEC has been applied for more
than ten years on various types of composite rotor blades (Chapter 10). For
*DLR Stuttgart, Germany.
tAero-Dynamik-Consult, Germany.

196

Design of Composite Structures Against Fatigue

Rotorblade

^yyyyyvTFCX
GRP
Cross nut (steel)

f>?yy|yyyy
Steel stud
Metallic flange

Fig. 13.1

Principle of the T-bolt load attachment of GRP blade to a metallic flange

GRP, the principle was used for the first time on the.DLR-windturbine
DEBRA 25 which has been in operation since 1984 without any problems of
the load attachment (1). However, there are still doubts about the proof and
durability of the design.
Figure 13.1 demonstrates the principle of the T-bolt load introduction
system: prestressed high strength studs orientated parallel to the blade axis are
screwed into radially arranged cylindrical nuts which fit into bores within the
thickened laminate. The system, which is also used in the IKEA furniture
industry, looks like a T-bolt. To reduce the load amplitudes in the bolts, they
have to be built-in under an initial stress which gives additional safety against
failure by dynamic loads.
To find out more about the stresses in the thick laminate, a theoretical
analysis was carried out at the beginning of the investigation by means of a
finite element method(FEA)simulation (2). The model used for the composite
lay-up was the DEBRA 25 laminate design which is nearly identical to the
material evaluated in Chapter 5. It comprises a mixture of UD and 45 degree
laminates. The analytical results could be compared with those obtained from
tests on a two-dimensional component (3). The program system PERMAS was
used together with the pre- and post processing system PATRAN. Three types
of elements were applied: volume, shell, and bar elements. Thus, threedimensional stress effects induced by the thick laminate could be investigated.
By means of contact analysis the nonlinear structural behaviour which occurs
when the GRP loses contact with the steel flange was also modelled. As an
example of the results, Fig. 13.2 shows the longitudinal stresses in the UD
layers due to the external design load and the prestress of the bolt. Some
important conclusions for the 0 degree-layers are listed in the following
summary.
(1) The maximum compressive stress is about 70 percent above the maximum
tensile stress.
(2) The out-of-plane shear stresses 4 and 5 are very low compared to the inplane shear stress 6.

Evaluation of T-Bolt Root Attachment

Fig. 13.2

197

Stress in the UD layers in loading direction under the external design load in the vicinity
of the T-bolt joint

(3) The normal stress 3 which is perpendicular to the laminate plane, is very
low apart from a local stress concentration in the contact area.
(4) The stress character between the cross nut and theflangeis independent of
the external loading and differs only by the applied stress level.
(5) The maximum stress in the contact zone to the cross nut is proportional to
the forces in the prestressed screws.
For the 45 degree-layers it can be concluded that:
(a) the stress concentrations also occur in the 45 degree-layers;
(b) both maximum compressive stress in thefibredirection and the shear stress
are higher than those for the 0 degree-layers.
In the comparison between the analysis and the tests, described in (3), earlier
failure could have been expected in the experiments. Because of the chosen
failure hypotheses in the analysis (ply damage stress), the calculated high stress
concentrations within the contact area to the cross nuts predict individual ply
failures of about one third of the applied maximum test load. The reasons for
the non-critical global failure behaviour may be due to creep effects in the
polymeric composite. This was measured to be between 3 to 4 percent of the
original prestressed state (3). The other layers of the laminate can then reduce
the high stresses of the single layers.

198

Design of Composite Structures Against Fatigue


Upper spar cap

GRP skin

Trailing edge
Foam

\
Foam
Lower spar cap
Fig. 13.3 Principle of ENERCON rotor blade structure showing location of spar caps and shear
webs

13.3 Tests on ENERCON rotor blades


Two prototype ENERCON 12 metre rotor blades were provided by Aero
Dynamik Consult for fatigue testing the blades and their T-bolt load attachments. As the blades had already been in operation and one of them was
severely damaged, it had to be repaired before testing. Since the blades were to
be loaded by a single load at the 2.5 m position they were cut down to a length of
about 3 m.
The aims of the tests were:
(a) proof that the T-bolt load attachment would exceed 105 load cycles (3);
(b) investigation of the rigging diagram for the system GRP-steel bolts by
testing;
(c) fatigue and failure mechanism of the blade structure.
Description of the rotor blades
The blade structure consisted of a shell of GRP-foam sandwich. The spar caps
were made of pure GRP rovings, the shear web of PVC-foam (Fig. 13.3). The
matrix system was a hand lay-up epoxy resin, whilst the load introduction was
performed by the T-bolt connection.
There is little known about the preloading of the rotor blades. However, it is
supposed that there were frequent emergency stops during the testing phase of
the prototype.
One of the blades (T2) showed severe damage (Fig. 13.4) as follows:
(a) upper spar cap 50-60 percent broken at a position of 0.48 to 0.55 m from
the root;
(b) leading edge delaminated between 0.3 and 3 m from the root;
(c) shear web compressed beneath the position of spar cap failure;
(d) shear web broken from beginning of the web up to 2.5 m from the root.
The rotor blade was repaired in such a way that the structure fulfilled all
requirements on the strength of the blade. No attention was paid to the surface
or aerodynamics.

Evaluation of T-Bolt Root Attachment


Replacement outer sandwich shell

199

Repair after beginning of test

Repair cap

Renewal outer laminate


Fig. 13.4

Fig. 13.5

Predamage on test blade T2

Rotor blade on test bed, loaded by a single hydraulic actuator

(a) the broken shear web was rebonded again by resin injection or, where
damaged too severely, replaced (at the beginning of the web at about 0.6 m
from the root);
(b) the damage in the upper spar cap was simply repaired by overlaminating
sufficient roving material from the outside;
(c) the damaged zone in the leading edge was cut out and replaced by a
laminate of the same lay-up.
General testing procedure
The rotor blade was mounted in a cantilever manner on a steel plate on the wall
which simulated the hub connection. The test or design load was applied as a
single load by a servohydraulic cylinder at a distance of 2.5 m from the flange
(Fig. 13.5). The relation of this transverse load to the bending moments at the

200

Design of Composite Structures Against Fatigue


1

Detail W

-Load introduction
Fig. 13.6

Positions of strain gauges at various levels along the blade

position of the root corresponds to the load of an extreme wind gust that occurs
once every one hundred years. The maximum measured strains in the GRP
caused by this load are shown in Fig. 13.11.
In afirsttesting series the design load was applied as a sine wave in tensioncompression (R = 1) on blade Tl. In a second series blade T2 was fatigue
loaded only in the tensile mode, the maximum load exceeding the alternating
load of Tl by 36 percent.
By the nature of the single load principle, the transverse load and bending
moment distribution near the load introduction point differ from the realistic
loading during operation of the blade on a WEC, i.e., the bending moments are
relatively small and the transverse loads are too high. This causes relatively
high shear loading in the web compared to the actual loading.
The strain measurements were carried out by means of strain gauges (Fig.
13.6). The following signals were measured:
(a) longitudinal strains in the cross section level 1,2,3 and 4 (4 only for Tl ) at
those locations where maximum strains were expected from theory;
(b) the shear strains near the leading and trailing edge at levels 4 and 5 (Tl
only).
Blade Tl
The first test series was carried out on blade Tl which, from the outside, was
undamaged. It was planned to fatigue-load it in tension-compression with the
design load up to 106 cycles, i.e., R = 1 with a residual static test to conclude
this investigation. Several static tests were performed during fatigue loading to
monitor eventual changes in the stiffness behaviour.

Evaluation of T-Bolt Root Attachment

201

During the fatigue testing the following problems occurred and were solved:
(a) Due to the high load the single-load clamping device deformed and
delaminated partly from the blade shell although it was glued on it. It was
strengthened and freshly glued on.
(b) After about 16 000 load cycles the inner part of the web delaminated from
the shell. About 250 mm of the web were cut out and replaced by a new
piece. Beyond that the end of the web had a smooth transition into the
shell.
(c) By determining the load amplitude, the fibre structure of the blade was
monitored. However, the prestressed screws themselves exceeded their
fatigue life limit after about 200000 load cycles.
Between 200 000 and 250 000 load cycles a raising of the front plane of the
GRP blade root from the steelflangewas observed, but it was not recognized as
failure of the screws. The fatigue test was continued up to 4.4 x 105 load cycles
without inspection of the steel bolts. Inspection at that stage showed that three
screws were broken at the upper and three at the lower side of the root; these
were fixed by sixteen screws on the flange. After renewing all screws, further
testing showed that some of them had to be replaced again after a further
200 000 to 250 000 load cycles.
Blade T2
For the second test series, the repaired blade (T2) was fatigue loaded in
tension-tension. The maximum load corresponded to 136 percent, the minimum load to 23 percent of the testing load of the first blade Tl. Thus, the
loading ratio was R = 0.167 for the tensile loaded lower spar cap and R = 6 for
the compression loaded upper spar cap. It was intended to apply 2 x 106 load
cycles being interrupted by static inspection tests. During the whole test no
screw had to be replaced; however at the end, the mostly unloaded screw (at
the upper side of the blade root), had failed.
During the fatigue testing the following problems occurred and were solved:
(a) The blade root separated from the steel flange device during the first static
test because the front side of the blade foot was not absolutely plane. It was
equalized by cotton filled resin.
(b) During the first load cycle.s it was shown that:
- the sandwich in the leading edge area was partly delaminated (audible
internal buckling);
- the web was newly broken (area of 1.06-2.5 m from the root).
The delaminated sandwich part in the leading edge was cut out and
replaced by the corresponding part of blade Tl. The web was repaired by
means of a ply-wood web. The resulting increase in blade stiffness had no
effects on the blade deformations within the 1 m inner blade part which

202

Fig. 13.7

Design of Composite Structures Against Fatigue

Failure of rotor blade Tl during a residual strength test after fatigue testing to 2 X 106
cycles

remained in the original stage between the repair of the spar cap and the
strengthened part of the web.
(c) At the resumption of the tests, the blade was, by a fault in the hydraulic
system, overloaded by 2.4 times the design load. However, since no
damages could be observed, the test was continued without any further
problems.
Residual strength tests-blade Tl
After the fatigue loading of the Tl blade up to 106 load cycles, a static residual
strength test was carried out. At a load of 2.33 times the design load sudden
instability failure occurred at 0.4-0.5 m from the root (refer Figs 13.7 and 13.8).
During failure the upper spar cap buckled into the interior of the blade and
destroyed the web underneath. Additionally it had partly broken through.
After failure the blade could still sustain 31 percent of the design load. Beside
the local damage the web was broken over the whole area of the transverse
loading. No damage was observed at the load attachment area.
Residual strength tests-blade T2
The residual strength of blade T2 was tested after 2 x 106 load cycles. The
failure was similar to that of Tl. It occurred at 2.5 times the design load near the
repair location, at a distance of 0.7-0.95 m from the root (see Figs 13.9 and
13.10).
The compressive spar cap failed in a manner similar to that of Tl, the web

Evaluation of T-Bolt Root Attachment

Fig. 13.8

203

Failure pattern of blade TI showing location of cracks

Fig. 13.9 Failure of rotor blade T2 during a residual strength test after fatigue testing to IO6 cycles

showing severe local damage. After failure, 44 percent of the design load could
still be applied. No damage was observed at the load attachment.
13.4 Discussion
Fatigue tests
In neither of the rotor blades was damage observed in the GRP laminates
during the fatigue investigation. This corresponded well to expectation, since
the measured maximum strains were very low compared to fatigue data
measured in (4) on a similar material, see Fig. 13.11.

Design of Composite Structures Against Fatigue

204

Cracks

Areas of delamination

Fig. 13.10

Failure pattern of blade T2

*=1

O=0.1

UD (a)

D=1

X=0.1

UD/i450(b)

t_J_Jf____ zzrS'
=&
t

10
Fig. 13.11

10'

10*

IO4
Cycles

Vf

10'

10'

10'

Measured strains in rotor blades Tl and T2 at various levels and e- diagram for GlEp
for both tensile and reverse loading (4); materials: (a) UD |5p|; (b) (45 degree/
[2p]UD)s

Evaluation of T-Bolt Root Attachment

205

"^ FZ3^"Failure

J;fcfl

100
ce
Cu

Norm

Tension side
Compression
side

OJ)

io

10'
Fig. 13.12

(S

10'
Cycles

107

Alternating stress versus cycles for steel (7)

In the load introduction area, at the location of maximum loading, cracks


were observed in the gel coat and the filler underneath. However, the cracks
did not continue into the laminate.
The fatigue limit of the screws was in good agreement with the corresponding
Whler curves of the steel (Fig. 13.12).
After the fatigue failure of one or more screws, the remaining (and now
higher loaded) screws did not suddenly break during the continued fatigue
loading. Thus, the load attachment had sufficient tolerance. The failure of the
screws was always in the first part of the screwedin thread.
Also after a severe damage, a repair carried out in an appropriate manner
will lead to a structure with full structural integrity.
Rigging diagram
Even in the design phase of a Tbolt connection, the establishment of a rigging
diagram is necessary to obtain a good balance between the strains in the GRP
and in the steel bolt. During the blade tests it was possible to verify the design
by measuring the strains of the prestressed GRP in the area between the cross
nuts and the flange. A comparison of the measurements with an FEA
simulation was possible (5). The strain measurements were carried out at the

206

Design of Composite Structures Against Fatigue


Compression

Tension

- Calculated

steel

5S

'

Elongation GRP

Fig. 13.13 Rigging diagram for T-bolt prestressed to 30 kN under design load; fi, is elongation in
steel stud and is the elongation within prestressed area of GRP (Fig. 13.1)

inner and outer surface of the shell, assuming that the medium strain would
correspond to the results of the calculation. Asymmetries could lead to wrong
interpretations when, during the prestressing procedure, the strain is only
considered on one side. Beyond that, an increased strain will be measured in
the outer phase of the shell due to the bending of the loaded structure. In the
rigging diagram the prestress load was corrected corresponding to the
measured load for lifting the blade root from the steel flange (Fig. 13.13).
The mean strains measured in the stage of prestress and external load
correspond well to the calculated mean strains (Fig. 13.14). The steep gradient
between inside and outside cannot only be caused by pure structural bending
due to the classical beam theory. It appears plausible that additional local
deformations were induced due to shear effects of the relatively thick laminate
and due to the special conditions of the load introduction.
Residual strength
As expected, in the residual strength tests stability failure occurred. The results
for the two blades are close together, the scatter being only about 10 percent.
The difference may be influenced by the fact that the failure location of blade

Evaluation of T-Bolt Root Attachment

1.5

207

Prestress + .
load 30kN

1.0

External
load
29.2kN +'

ce
c

0.5
Calculated
+ Measured
Inside

Outside

Location

Fig. 13.14 Strains on GRP in load introduction area of studs at prestress and maximum external
load
T l is in the area where the cap of blade T2 was strengthened by the repair (Figs
13.7-13.10). The buckling of the spar caps appeared at about 85 percent of the
theoretical buckling load of a shell strip curved correspondingly to the local
form of the spar cap (6).
13.5 Conclusions
Rotor blades and T-bolt connections were investigated theoretically and by
experiment. In an FEA analysis it was shown that:
(a) the stress concentrations in the contact area between the GRP and the
cross nuts are proportional to the forces in the screw threads;
(b) the out-of-plane stresses induced by the thickness of the laminate are very
low compared to the in-plane stresses;
(c) the compressive and shear stress maxima in the 45 degree-layers are
higher than those in the 0 degree-layers;
(d) in comparison to experimental results values achieved by the simulated
stress concentrations were too high, since the smoothing creep effects of
the polymer compound were not considered.
Two prototype rotor blades from the ENERCON E12 turbine, one of them

208

Design of Composite Structures Against Fatigue

predamaged and repaired, were investigated statically and in fatigue. They


were designed with a Tbolt connection system. One of them was loaded up to
106 load cycles with the design load at R = 1 (tensioncompression), the other
to 2 x 106 load cycles at a 36 percent higher load at R = 0.167 (tension
tension). It was shown that:
(a) there was no fatigue damage in the GRP part of the load introduction
system, except for cracks in the surface protection near the cross nuts;
(b) the prestressed steel studs failed beyond the Whler curve given in (7)
(EC3 norm) for the stressconcentration index referring to the quality of
the screws;
(c) after failure of one or more studs, there was still a large enough safety
margin to carry the loads via the other screws for a considerably longer
time period;
(d) the blade foot must be exactly planar so that no lifting of the GRP from the
steel flange occurs;
(e) the rigging diagram could be verified by means of strain gauge measure
ments and under loading. However, the strains inside and outside the GRP
shell may differ from each other by bending due to the classical beam
theory and shear deformations;
(f) in the GRP part of the blades no fatigue damage could be observed;
delaminations in the sandwich, caused by the predamage, could be
repaired and did not occur again;
(g) the residual strength was not influenced by fatigue, but was determined by
the expected buckling failure; the safety factor of 2.3 and 2.5 for the two
blades related to the design load was sufficiently high to have an adequate
safety margin.
This type of load attachment design is attractive as it is of light mass, able to be
calculated and has builtin redundancy.
References
(1) HALD, H. and KENSCHE, Ch., 1985, Development and test of a light weight GRP rotor blade
for the DFVLR 100 kW WEC, Proceedings Wind Power'85, San Francisco, CA.
(2) WINKELMANN, P., 1992, Berechnung einer Krafteinleitung in ein dickes FVLaminat mit
Hilfe der Finiten Elemente, DLR, IB 43592/09.
(3) KENSCHE, Ch. W., 1991, Dehnbolzenverbindung in Faserverbundwerkstoffen, VDI Werk
stofftag Mnchen.
(4) KENSCHE, Ch. W., 1992, High cycle fatigue of glass fibre reinforced materials for wind
turbines, DLR Forschungsbericht, 92, 17.
(5) KLEINHANSL, S., 1992, FEBerechnung Rotorblattanschlu, EN 12, AeroDynamik
Consult, Stuttgart.
(6) WIEDEMANN, J., 1986, Leichtbau, Band /(Springer Verlag).
(7) Eurocode 3, Abschnitt 9: G renzzustand der Betriebsfestigkeit (Ermdung), Stahlbau
Verlagsgesellschaft mbH, Germany.

CHAPTER 14

Comparison of Fatigue Curves for Glass


Composite Laminates
A. T. Echtermeyer, E. Hayman, and K. O. Ronold*
The fatigue curves of the preceding chapters are collated and analysed by plotting the data in the
form of strain versus cycle curves. Standard curves can then be produced, which can be used as a
basis for design. Criteria for the application of these curves arc given.

14.1 Introduction
Knowledge of both materials properties and fatigue loading sequences is
needed to perform fatigue lifetime calculations. Lifetime calculations also
depend on the safety factors used in the analysis. It is important that these
general issues are clarified with customers, certification bodies, and authorities
before the design work is carried out. This chapter addresses the fatigue
resistance of fibre composites, which is an important part of structural design
calculations.
Characterizing fatigue of composites by a strain-life diagram was first
suggested by Boiler (1) and Talreja (2)(3). This method provides a conceptual
framework for the interpretation of fatigue results, based on matrix and fibre
properties. The method also allows for a direct comparison of fatigue data for
different laminates without the need to normalize data by the appropriate
value of the ultimate strength.
Fatigue data in the literature are, however, usually reported as S-N curves,
sometimes normalized by the static strength. Mandell et al. showed that a large
number of different glass fibre reinforced composites have the same fatigue
behaviour under tensile loading (R = 0.1) (4). The reduction of fatigue
strength is about 10 percent per decade of stress range relative to the actual
tensile strength.
The approach described here is similar, showing that many different glass
reinforced composites follow the same strain-life (-) curve. However, data
are not normalized by their static properties, tensile or compressive strength or
strain to failure. The fatigue curves obtained here are obtained from fatigue
data, without using values from static measurements. It was found that static
strains to failure can be quite different, but laminates still showed the same
"Det Norske Veritas Research AS, Norway.

210

Design of Composite Structures Against Fatigue

Data point
Residual

\-_A
^v
~ "^v

"c
2

S
95% tolerance '.
bound with 95% '.,
confidence

N^Mean
A

^ ^

Log (cycles)
Fig. 14.1

Schematic presentation of fitting fatigue data and obtaining the tolerance bound.
(Not to scale)

fatigue characteristics (Chapter 3). This means that static and fatigue perform
ance are not necessarily related to each other.
14.2 Linear regression analysis
Mean values
The curve expresses the relationship between strain amplitude, , and
number of cycles to failure, N. A model of the form
=-'"

(1)

is commonly used to represent this relationship, with being the dependent


variable and the independent, and and m are material coefficients. This
model implies the following relation
log,N = log,,, m log,,,

(2)

which gives a linear model in the data pairs (,A/) by


logioN = log,,, m log,,,, + e, i =1,2,...

(3)

where e is an error term (or residual) and is the total number of observations
of pairs (,).
Linear regression analysis of data pairs (,) by this model leads to least
squares estimates of the intercept log,,, and the slope m. An estimate s of the
standard deviation oof the residuals results as a byproduct of the regression. A
schematic of measurement data, linear regression line, and residuals is shown
in Fig. 14.1.

Comparison of Fatigue Curves for Glass Composite Laminates

211

Note that it is common to plot versus (or rather log,,, versus log,,, N)
and, accordingly, work with a relation
log, = logio k - a log, N

(4)

However, a regression according to this relation, with residuals in log I( ^ rather


than in log]0A7, leads to an estimation of coefficients \ogn)k and a that describe a
different line from the one described by the estimates of log,,, K and m obtained
by the linear regression, as described above. As long as N is the dependent
variable and is the independent variable, one must adhere to this estimation
of log,,, and m, and then calculate log,,, k and as follows
a = and
m

\ogwk

= log,/C

(5)

when their values are needed.


For regression, reference is made to standard text books on this subject, e.g.,
Neter et al. (5). The linear regression is also described in ASTM E739. A
method for obtaining an estimate 2 of the variance o2 of the normal distri
bution is also given. The estimate s of the standard deviation can be
calculated by

Tolerance bounds for random variables


The line described by equation (2) gives the best fit of the data and represents
the mean fatigue life. Assuming a normal distribution of the residuals of the
data with respect to the best-fit line, the significance of the mean is that there
will be 50 percent probability that the true fatigue life exceeds this mean and 50
percent probability that it falls short of this mean. Characteristic materials
data, only allowing for a low probability of failing short of the mean fatigue life,
have to be used for design. A common way of describing such fatigue data for
design is to determine a lower 'y-proportion' tolerance bound for the random
normal variable log,A/ from the - data with confidence 1 a as follows
(login) -a* = log,N - c,

,s

(6)

where log,,, is the estimated mean, expressed as a linear function of log,,^,


and s is the estimated standard deviation of log,,, based on observations of
log,,, N. The coefficient c, _ is solved from the probability equation (6)(7)
P{P(log, - c, _ a^

< log,,, A/) > ) = 1 -

(7)

The interpretation of the lower tolerance bound (log,,. A/), _,,, is that a
proportion of all future realizations of \ogn)N will exceed (log,,, ), _ ,, with
confidence 1 a. As log,,, and s are uncertain estimates, c, _ , , becomes the

212

Design of Composite Structures Against Fatigue

1 - a quantile in a non-central t distribution and depends on the number of


observations n of log,,, N.
In many cases the one-sided lower 95 percent tolerance bound with 95
percent confidence is calculated, i.e., = 95 percent and 1 - a = 95 percent.
The so-called 'A' basis and 'B' basis design allowables are also frequently used.
These have 99/95 percent and 90/95 percent tolerance bounds, respectively (8).
These can be calculated by the same method. (Note that the ASTM E739
standard (9) provides a method of calculating confidence bounds for the
median values of the fatigue curve. This is different from the tolerance bounds
presented here.)
The general equation (7) cannot be solved explicitly for the quantile c, _ a ,,.
It has been solved implicitly for the case in which the data are unconditioned
random Gaussian variables (10)-(12). This solution is the same as that
described in DIN 55303 (13) and it can, for example, be applied to analyse test
data of static tests. The method is also sometimes used in fatigue analysis.
However, strictly speaking, fatigue data are conditioned. Data are obtained in
a certain cycle range and depend on that range. The uncertainty in fatigue life
predictions increases for cycle ranges extrapolated beyond the ranges of the
measurements. For this case of conditioned fatigue data, no tabulated solution
exists. A solution method was developed by Ronold and Echtermeyer (14). If
the experimental data are distributed fairly uniformly over a log (strain) range
Lx, the c, _ a values can be found in Fig. 14.2. The figure can be used in the
following ways to find the 95 percent tolerance bound with 95 percent
confidence:
(a)
(b)
(c)
(d)

is the strain at which the tolerance bound is found;


is the mean of the logarithm of all strain measurements;
L x is the range of the logarithm of all strain measurements;
Ax is the distance from the mean measured strain to the strain level of
interest
Ax = | - log1 0 '

(e) c,. y can now be found for any AxlL% value: note that c,. a ,, depends on the
number of measurements;
(f) the number of cycles which, with 95 percent confidence, will be exceeded
by (at least) 95 percent of all fatigue life realizations at the strain can now
be found by using equation (6);
(g) a curve representing the 95 percent tolerance bound with 95 percent
confidence can be obtained by repeating steps 1-6 for many different strain
values .
The tolerance bound obtained by these steps is also shown in Fig. 14.1.
If tolerance bounds other than the 95/95 bound are to be found, equation (7)
has to be solved by computer simulation. Alternatively, a curve similar to Fig.
14.2 has to be developed. Both methods are described in reference (14).

Comparison of Fatigue Curves for Glass Composite Laminates


1

213

=10 . /
8
7

n=\SS/'

- 6

5
c
a

S^^n=VS
^ ^

/ \ s ^

^^^n=\W^~
O 4

3
2

Ar/,
Fig. 14.2

Quantile C\ _ y versus ratio /, for various numbers of observations n; 1 a = 0.95;


y = 0.95

14.3 Reverse loading


DA/VT? data (Chapter 3)
Ten different materials with some fibres in the load direction were tested
(Table 3.1). Note: only ten of eleven materials in Table 3.1 have fibres in the
load direction. The fatigue test results are presented in Fig. 14.3. Even though
there are differences between the materials (for more details see Chapter 3),
the similarity between the data points is striking. All data lay fairly close to
one curve. The fatigue performance of these laminates is fibre dominated. A
low percentage of fibres (of the order of 2 percent) in the load direction (as in
the MA90 laminate) is enough to ensure that fatal cracking of the matrix does
not occur.
Fibredominated fatigue is influenced by the quality of thefibresand the type
of weave or fabric. This may explain why the multiaxial fabrics perform slightly
better than the woven combilaminates, because fibres in the woven layups
experience more friction. However, these differences appear to be small,
especially at high cycle numbers (under small loads). It is, therefore, suggested
that all glass fibre reinforced laminates with some fibres in the load direction
are described by a single standard fatigue curve, see Echtermeyer (15).

214

Design of Composite Structures Against Fatigue

2.0

'S

\
\

1.6

95/95 limit

<^

2 12

Mean

\
\ 4

lA ""

.
00\

o
.o
\

1.0

1 0.8

S
0.6

0.4
0.2
0

/^\

Combimat
Multiaxial

o o

twK$?

^^

10

10"

10*

10'

10'

"

10'

10'

Cycles

Fig. 14.3 Strain' data and standard logarithmic fit curve for fatigue data at R = 1. The curve is
based on data measured at DNVR

The standard fatigue curve can be based on test data of the ten different
laminates with somefibresin the load direction with a total of 111 measurement
pairs (,). Under the assumption that the different laminates all adhere to the
same material model with the same coefficients log,0 and m, i.e., all 111 data
pairs belong to the same population, a regression can be carried out as
described above. Figure 14.4 shows the regression line which represents the
best fit to the data on a loglog scale. The regressed coefficients of the line
equation are given in Table 14.1. Figure 14.3 shows the corresponding curve in
an diagram with linearly scaled axes. The standard deviation OL0g(N) ^ t n e
residuals of log,0 is estimated by s = 0.438.
The bestfitsfor the data sets of the individual materials can be found in Table
3.6.
The onesided lower 95 percent tolerance bound with 95 percent confidence,
constructed as described above, is also shown in Figs 14.3 and 14.4. The
interpretation of this bound is that at least 95 percent of all future outcomes
of log,,, N, given , will fall to the right of this bound with 95 percent

Comparison of Fatigue Curves for Glass Composite Laminates

215

3.0
0 DNVR combi-mat
O DNVR multi-axial
A DLR dry
ECN

2.0

2 1.0

2 0.9
I 0.8

AAA

A A A

^0.7

| <

I 0.5
0.4
0.3

0.2

umi
10

L
10'

Vf

10'
Cycles

Vf

10'

IO7

10*

Fig. 14.4 Standard fatigue curve in comparison with data from different materials measured at
ECN and DLR
Table 14.1

Coefficients of - curve for reverse loading (R = 1)


Regressed coefficients*

is independent variable
is dependent variable
Source
Standard curve
DNVR

Converted regression
coefficients^

Chapter

Curve/limit

log,

logla k

Mean
95/95

3.552
2.681

7.838
7.838

0.453
0.342

0.128
0.128

Mean
95/95
Mean
Mean
95/95
Mean

3.534
2.653
3.775
3.808
3.163
4.583

8.039
8.039
10.204
9.615
9.615
8.333

0.440
0.330
0.370
0.396
0.329
0.550

0.124
0.124
0.098
0.104
0.104
0.120

Other regression results


DNVR + ECN
3+ 6
ECN
FACT

DLR

'Regression has been carried out with as independent and as dependent variable
tlogm k and are found by conversion (equation (5)).
95/95 tolerance board for 50 < < 10 6

216

Design of Composite Structures Against Fatigue

confidence. On a log-log scale, this 95/95 tolerance bound can be approximated by a straight line within the range of the measured data, 50 < N < IO6
(Table 14.1).
Note that only for the large number of specimens tested, the 95/95 tolerance
limit is nearly a straight line within the range of the measurements, and in the
present case it is about the same as the limit which is given by the mean line
shifted to the left by 2crLog (/v). For a smaller number of test specimens the shift
of the line would have to be larger than 2aLog(N) to maintain the same
tolerance, and the curved nature of the true 95/95 tolerance bound would be
more pronounced (14). For an infinite number of test specimens the 95/95
tolerance limit would coincide with a line obtained as the mean line shifted
1.64o"Log(N)t0 the left.
The deviation of the tolerance bound is based on an assumption of normality
of the residuals in logI0 N. The validity of this assumption should be checked. A
rough and simple test for an infinite number of test samples is to check whether
68.27 percent of the data points lie within the interval of plus/minus one
standard deviation of the mean, and 95 percent of the points lie within the
interval of plus/minus two standard deviations of the mean. The actual
percentages of points below each line is about as expected. From this it seems
reasonable to judge that the fatigue data follow a normal distribution. A more
accurate way to check whether data are normally distributed is to prepare a
normal probability plot (14) and check for linearity. This plot confirms that the
data seem to be normally distributed.
Comparison with other data
A glass 0/45 polyester matrix lay-up was tested at ECN at R = 1 (Chapter
6). The material yielded similar results to those described by the standard
fatigue curve obtained from DNVR data, though the slope was slightly less
(Table 14.1). The data are shown in Fig. 14.4.
The ECN data may be judged to belong to the same population as the DNVR
data. A regression of the 111 data DNVR data pairs pooled together with the
nineteen available ECN data pairs gives a regression line fairly close to the
standard curve which is based on the DNVR data alone (Table 14.1). The
standard deviation of the residuals in log, N by this regression is only slightly
reduced to 5 = 0.422.
A larger number of fatigue measurements on 0/45 glass reinforced laminates is stored in ECN's data base FAtigue of Composites for wind Turbines
(FACT) (16) (Table 14.1).
A lay-up with 0/45 glass epoxy was subjected to fatigue testing at DLR
(Chapter 5). The slope and intercept of the regression line are also given in
Table 14.1. In comparison with the specimens tested by DNVR the material
tested at DLR is superior. Whether the good properties were caused by the
type of fibres or the use of epoxy resin could not be determined. The DLR

Comparison of Fatigue Curves for Glass Composite Laminates

217

material was also tested under wet conditions; it then gave similar results to the
dry DNVR materials. Pre-damaged specimens were also tested at DLR
(Chapter 5).
Mandell et al. tested various triaxial and unidirectional laminates at R = 0.1.
Data also followed a general strain-life fatigue curve similar to the one shown
here (17). The unidirectional laminate tested had good properties, comparable
to the material measured at DLR.
The fatigue results of the other studies give slightly different fatigue curves
than the DNVR standard curve. This is to be expected, since there are
differences in the materials, which were also observed in the tests performed at
DNVR. A standard fatigue curve will, therefore, always depend on the
particular materials that the average was taken from. If test results from ECN
and DLR were incorporated into the DNVR standard curve, the curve would
change slightly. The presented standard curve is not an absolute physical law.
The advantage of a standard curve is that it is based on many measurements.
The amount of uncertainty in the data is less than for a few measurements taken
on one specific material. The standard curve is very useful for design studies
when exact materials data are not available. It is also useful as a standard curve
against which new materials can be compared. From a safety point of view it is
important to have a standard curve that is not only based on high-performance
materials. On the other hand, if high-performance materials are used in a
structure, credit should be given for their good properties, if sufficient data can
be provided to demonstrate that the material is better.
The conclusion from this is that the single - curve gives a fairly good
approximation of the individual - curves (reversed loading) with some fibres
in the load direction. This fit equation can be used as a basis for initial design.
More advanced materials, such as DLRs should be given more credit in design,
but good documentation is necessary to ensure that their properties really are
better. It is also necessary to check whether a new material is at least as good as
predicted by the standard curve. Only a few tests are necessary to establish
whether the minimum performance is met. These tests are discussed below.
No fibres in the load direction
Fatigue lifetime curves of test coupons with no fibres in the load direction are
much lower than the standard curve. In Fig. 14.5 the results from reverse load
tests on FFA 30 (Chapter 4) and DNVR MA-45 (Chapter 3) are compared
with the standard logarithmic fit curve of all the other DNVR specimens.
The short lifetimes of these materials compared to the standard curve may be
due to the testing method. The fibres run from edge to edge in the small test
coupons; they do not connect the load introduction points (grips). Therefore,
the loads have to be transferred by the much weaker polymeric matrix. The
fibres can be continuous in a structure or component, e.g., a filament wound
pipe with fibres in the 55 direction. In such a case fatigue lives may be longer

218

Design of Composite Structures Against Fatigue


2.0

1.8
FFA30 degrees
o DNVR145 degrees

1.6
1.41

| UK
2 1.0

I 0.8
"

m
S

0.6

0.4
0.2

Fig. 14.5

io

io1

IO'

10*
Vf
Cycles

10'

10'

10'

0
Comparison of Strain/V curves resulting from specimens with no fibres in the load
direction, with the common DNVR logarithmic fit curve

and can possibly be calculated by applying the standard fatigue curve to the
strains in the direction of the fibres. However, no experimental proof of this
hypothesis is known to the authors.
14.4 Tensile loading
The experimental results described above were all obtained at an R ratio of 1 ,
i.e., the maximum tensile and compressive stresses were the same and the
mean stress was zero. Fatigue lifetimes at other ratios will be different due to
the presence of nonzero mean stresses. This is important when a structure is
exposed to variable loading.
Typically, experimental data are obtained for R ratios of 1 , 0.1 (tension),
and 10 (compression). From such data a constant lifetime diagram can be
constructed from which fatigue lifetimes at other amplitudes and means can be
extrapolated, as discussed in Chapters 4 and 7.
Just as in the case of reverse loading, fibres in the load bearing direction
dominate the fatigue properties with tensile loading. The Ris data (Fig.
2.4(d)) clearly show how the zero degreefibresdominate the loading even with
some fibres oriented at 45 degrees. This is confirmed by data from ECN (Fig.

Comparison of Fatigue Curves for Glass Composite Laminates


1

219

2.0
* DLR
ECN
RIS
B enchmark

.*
s*
S
S

CS
c

>.

\ .

1.0 \

^ >*
0.8

^\.

57

^ ^v P
^

> V

VM

0.6 _

r/3

or>o A 0 < * ^

0.4

0.2

1
1

1
lir

1
lir

1
10"

S.

1
10s

>v
" ^

X^.^

1
10'

IV
IO7 IO8

Cycles
Fig. 14.6 Strain' data and standard logarithmic fit curve for fatigue data at R = 0.1. The curve is
based on data measured at Ris and ECN and DLR

6.3) on a 0/45 degree plied laminate. Fatigue dominated by the fibre


properties was also observed by Mandell et al. for many different glass fibre
reinforced laminates (4). They found a standard SN curve by normalizing
fatigue stresses to the static strength.
One can, therefore, derive a standard curve for this loading, and this is
shown for R = 0.1 in Fig. 14.6 based on data from tests by ECN, DLR, and
Ris. The data consist of ninetynine pairs (,). The coefficients of the
regressed line, which gives the bestfitto the data on a loglog scale, are given in
Table 14.2, and the standard deviation o,ogA/ of the residuals of log,,, by the
regression is estimated by s = 0.716. Since the number of materials and data is
relatively low, a standard curve based on more measurements may have
slightly different coefficients and a lower standard deviation.
The onesided lower 95 percent tolerance bound with 95 percent confidence,
constructed as described above, is also shown in Fig. 14.6. The interpretation
of this bound is that 95 percent of all future outcomes of log,,, N, given , will
fall to the right of this bound with 95 percent confidence. On a loglog scale,
this 95/95 tolerance bound can be approximated by a straight line within the
range of the measured data, 10 < < 107 (Table 14.2).

220

Design of Composite Structures Against Fatigue


Table 14.2

Coefficients of - curve for tensile loading (R = 0.1)

f is independent variable
N is dependent variable
Source
Standard curve
ECN. DLR
Ris
Benchmark

Regressed coefficients*

Converted regression
coefficients f

Chapter

Curve/limit

logn, K

log,t, k

6
2
8

Mean
95/95

2.105
0.678

10.194
10.194

0.206
0.066

0.098
0.098

Note: Ninetynine data pairs used in analysis comprising data from Ris 0, 0710, 0730, 0745;
ECN 0745; benchmark, combination fabric.
95/95 tolerance board for 10 < < IO7

14.5 Compressive loading


The data from Ris (Fig 2.4(e)) show that for this type of loading, having some
fibres in the zero direction does not seem to dominate the fatigue properties.
There is also a clear angle effect. Furthermore, the fibre composites are less
resistant to fatigue loading in compression than in tension (Fig. 2.4(b)).
Therefore, a standard curve is not possible
14.6 Flexural loading
Flexural testing by UWE (Chapter 9) uses coupons with a transverse filament
tape as the reinforcement. This puts almost all thefibresin the load direction (0
degrees). Testing was done in flexure with an R ratio of 0.25. In Fig. 14.7 the
data are plotted for coupons cut from flat plates. The data all lie higher than the
standard fatigue curves in Figs 14.4 and 14.6, which were obtained by applying
reverse and tensile loading (and other materials). This is analogous to the
benchmark tests (Chapter 8) where the flexural fatigue data lie above the data
for tensile and reverse loading. A combined fatigue curve can be based on the
data from all the tests by UWE. The data consist of twentyseven pairs (,).
The coefficients of the regressed line, which represents the best fit to the data
on loglog scale, are given in Table 14.3, and the standard deviation o,ogN of
the residuals of log,A/ is estimated by s = 0.322. Note that too few materials
were tested in a small range of cycle numbers to consider the combined fatigue
curve a standard fatigue curve.
The onesided lower 95 percent tolerance bound with 95 percent confidence,
constructed as described above, is also shown in Fig. 14.7. The interpretation
of this bound is that 95 percent of all future outcomes of log,,, N, given , will
fall to the right of this bound with 95 percent confidence. On a loglog scale,
this 95/95 tolerance bound can be approximated by a straight line within the
range of the measured data, 104 < < 106 (Table 14.3).
Higher strengths or longer lifetimes obtained by flexure tests are generally
attributed to the smaller volumes tested. The probability of finding a defect in a
small volume is low, which leads to higher strengths (18). Also, the minimum of

Comparison of Fatigue Curves for Glass Composite Laminates

221

3.0
Mean
2.0

95/95
limit

c
2

1.0

I 0.8
0.6

O.4I

10'

Fig. 14.7

I I I I I II

10'
Cycles

10*

10'

10'

Slrain-.V data for flexural testing in three-point bend. Data are fitted by one curve

Table 14.3

Coefficients of e-N curve for flexural loading (R = 0.25)

is independent variable
is dependent variable
Source
Standard curve
UWE
Benchmark

Chapter

Regressed coefficients"
Curve/limit

log,

Mean
95/95

6.539
5.654

Converted regression
coefficientst
log, k

7.400
7.400

0.884
0.764

0.143
0.143

Note: Twenty-seven data pairs used in analysis comprising data from UWE transverse filament
tape (plates 1 and 2); benchmark, combination fabric
95/95 tolerance board for 10 < N < 10"

a strength field over a small volume is higher than over a large volume, and this
minimum is the governing strength. Therefore, flexure data may overestimate
the strength of larger components if subject to tensile or compressive loading.
The results must be used with caution for fatigue design, especially if inplane
properties are needed.

222

Design of Composite Structures Against Fatigue

14.7 Other analysis methods

There is no agreement in the literature on how fatigue lifetime data should be


described. The linear model described above for the standard fatigue curve
based on equation (4) is commonly used
log,,, = log,,, - ctlog,,,

(4)

However, many other methods can be used to describe the fatigue data. Some
other methods to find means and confidence levels are discussed below.
log -log relations versus -log relations
In some cases data are fitted by a semi-logarithmic equation
= k - m*\ogl0(N)

(8)

The equation is often expressed as a stress-lifetime (S-N) curve (19)-(21) and


normalized to the tensile strength. Usually there is not enough data to decide
which way of fitting them is best. However, the data described here seem to
follow the log-log format of the standard curve (equations (2) and (9)) very
well (equation (2) is derived from Paris fatigue crack growth law). The high
cycle fatigue data presented here seem to indicate that the log-log equation (2)
can be used. Van Delft et al. (22) tested specimens for more than 108 cycles and
found that the log-log presentation of fatigue data was the best approach for
their GRP material. It should be noted, that it is important to describe the
fatigue properties of a material as well as possible in the strain range in which
the load sequences cause most damage.
Weibull statistics
Other methods try tofitthe fatigue data using Weibull statistics (23)-(26). This
method takes the residual strength of the materials into account, assuming that
the strength data follow a Weibull distribution. A good description of the
method is given by Sendeckyj (23). The method is applied to the complete set
of DNVR data (Chapter 3) used to obtain the standard fatigue curve (27). Data
and the Weibullfitare shown in Fig. 14.8. It can be seen that the Weibullfitand
the log-log regression fit yield very similar results, except for low cycle
numbers (N < 100). The 95/95 tolerance bound of the log-logfitis compared to
the 5 percent hazard curve for the Weibullfit.The hazard curve (28) is based on
a different theory from the tolerance bound.
14.8 Comparing new materials against the standard fatigue curve
The standard - curves for R = -1 and R = 0.1 can be used as a basis for
initial design, Figs 14.4 and 14.6. For application of the equations of these
standard curves, the material in question must satisfy the criteria outlined
below.

Comparison of Fatigue Curves for Glass Composite Laminates


3.0

~i

223

Weibull

0.2
Fig. 14.8

10

10'

10"

Cycles

10*

10'

10'

10'

Comparison of the standard fatigue curve with a Weibull fit based on the same data set.
Note that the tolerance bound for the log-log fit and the confidence limit for the Weibull
fit have a different statistical meaning

(1) Some continuous fibres must lie in the load direction. Two percent of the
fibres may be adequate, as this was the amount in the DNVR MA-90
material. This may prove sufficient in preventing fatal cracking of the
matrix; hence this material behaved in the same fashion as the other
materials with more fibres in the load direction.
(2) The static strain to failure should be measured based on a standard test
method, like ISO or ASTM. The standard curve is only valid where the
fatigue strain is lower than the static strain to failure.
(3) The fatigue test methods must yield valid results. This means that failure in
the grip area must be prevented, as grip failure may cause too short
lifetimes to be measured. Fatigue testing in flexure yields longer lives to
failure than other methods of testing and gives non-conservative values for
the fatigue lives under other types of loading than flexure.
(4) The material should be tested at three strain levels to ensure that the
resulting strain-A/ curve has the same character as the standard curve.
These strain levels should be chosen such that failure occurs after about
103,104, and 106 cycles, respectively. With reference to the standard curve
in Fig. 14.4, this would correspond to strain levels of 0.5, 0.9 and 1.2

224

Design of Composite Structures Against Fatigue

percent, respectively, for R = 1. Slightly different strain levels would be


needed for R = 0.1 (Fig. 14.6). At least three data points (,) should be
measured at each strain level to be able to apply the standard curve in initial
design.
Ideally, under an assumption of normality, no more than 2.5 percent of the
measured pointsshould lie below - 2, no more than 16 percent below - ,
and no more than 50 percent below the mean/*. If the data points for a material
do not fulfil these requirements, an individual fatigue curve has to be estab
lished. The equations for the standard fatigue curve (see above) and its
associated crand/t - 2 curves are:
for/? = - 1
log(emCan) = 0.453 - 0.1276*log(A/)
log(/i_f7) = 0.397-0.1276*log(A/)
log(e-2a) = 0.342 - 0.1276*log(A7)
for = 0.1:
logman) = 0.207-0.101* log(N)
log(v) = 0.135- 1.101* log(N)
log(/(.2a) = 0.063 - 1.101* log(A)
where is the strain in percent, and the indices are self explanatory.
Note that the , a, and 2 curves are used here to check whether a
new material can be described by the standard fatigue curve. The 95 percent
tolerance bound with 95 percent confidence is different; it should be obtained
as described above.
A material may have a high proportion of points above the mean curve (as is
the case with the DLR dry material. Fig. 14.4). The standard curve fit is then
conservative. Such advanced materials should be given more credit in design,
but good documentation is required to ensure that they are better. In order to
obtain an individual fatigue curve, at least five tests per strain level should be
measured, following common practice of fatigue testing. If the 95/95 tolerance
bound should be established as described above.
The effect of R ratio for a given type of loading has only been considered for
wood composites (Chapter 7). From the definition of the R ratio (Fig. 0.2),
increasing its value increases the minimum stress and, therefore, the mean
stress and decreases the strain amplitude. There is no reason to suppose that
other standard curves cannot be developed for other R ratios if there are some
fibres in the load bearing direction.
14.9 Conclusions
(a) Fatigue data of glass fibre reinforced composites can be described well by
common strain-life fatigue standard curves, as long as some fibres in the

Comparison of Fatigue Curves for Glass Composite Laminates

(b)
(c)
(d)
(e)

225

material run in the load direction for reverse loading,flexuralloading, and


tensile loading with similar R ratios.
These standard curves can be used as the basis for initial design studies.
A few materials show better fatigue characteristics. Credit should be given
to them if they are used in structures.
Fatigue properties of an unknown composite should be checked against
the standard curves for at least three strain levels. A detailed procedure is
described.
A method is shown to obtain the 95 percent tolerance bound with 95
percent confidence for fatigue data.

References
(1) B OLLER, K. H., 1969, Fatigue fundamentals for composite materials, ASTM STP 460,
(ASTM, Philadelphia), pp. 217235.
(2) TALREJA, R., 1981, Fatigue of composite materials and fatiguelife diagrams. Proc Royal
Soc Lond Ser. A 378 pp. 461^175.
(3) TALREJA, R., 1982, Damage models for fatigue of composite materials, (edited by H.
Lilholt and R. Talreja) Fatigue and creep of composite materials (Ris National Laboratory,
Denmark) pp. 137153.
(4) MANDELL, J. F., HUANG, D. D., and McGARRY, F. J., 1981, Proc 36th R P/C Inst,
Society of the Plastics Industry, Paper 10A.
(5) NETER, J., WASSERMAN, W., and KUTNER, M. H., 1990, Applied linear statistical
models, third edition, (R. D. Irwin Inc., Homewood, Illinois).
(6) MADSEN, H. O., KRENK, S., and LIND, N. C , 1986, Methods of structural safety,
(PrenticeHall, NJ), p. 39.
(7) LITTLE, R. E., 1981, Review of statistical fatigue life data using onesided lower statistical
tolerance limits. Statistical analysis offatigue data, ASTM STP 744, (edited by R. E. Little and
J. C. Ekvall), (ASTM, Philadelphia, USA), pp. 323.
(8) Military Standardization Handbook, Metallic materials and elements for aerospace vehicle
structures, MILHDB K5B . 1971. Department of Defense.
(9) ASTM, 1991, Standard practice for statistical analysis of linear or linearized stresslife ()
and strainlife () fatigue data, (ASTM, Philadelphia, USA).
(10) RESNIKOFF, G. J. and LIEB ERMANN, G. J., 1957, Tables of the noncentral t distri
bution, (Stanford University Press, California).
(11) OWEN, D. B., 1958, Tables offactors for onesided tolerance limits fora normal distribution,
(Sandia Corporation Monograph SCR13, Washington, DC).
(12) PEARSON, E. S., and HARTLEY, H. O., 1976, Biometrica tables for statisticians. Third
edition, (B iometrika Trust, London).
(13) DIN, Statistische auswertung von Daten, Bestimmung eines statistischen Anteilsbereichs
(Statistical interpretation of data; determination of a statistical tolerance interval), 1987,
DIN55303, Part 5, (in German).
(14) RONOLD, K. O. and ECHTERMEYER, A. T., 1995, Estimation of fatigue curves for
design of composite laminates. Composites.
(15) ECHTERMEYER, A. T. 1994, Fatigue of glass reinforced composites described by one
standard fatigue lifetime curve. Proceedings of the 5th European Wind Energy Conference,
Vol.1, pp. 391396.
(16) SMER, B . J. and DE, B ACH, P. W., 1994, Data base FACT; FAtiguc of Composites for
wind Turbines, ECN Report ECNC94045.
(17) MANDELL, J. F., REED, R. M.. SAMB ORSKY. D. D. and PAN, Q. R., 1993, Fatigue
performance of wind turbine blade composite materials, SEDVol. 14, Wind Energy
(ASME, USA).
(18) PIPES, R. B ., 1990, Delaware composites design encyclopaedia. Vol. 6, Test Methods,
(edited by L. A. Carlsson and J. W. Gillespie), (Tcchnomic Publishing Corporation, B asel),
pp. 354.

226

Design of Composite Structures Against Fatigue

(19) MANDELL, J. F., 1982. Fatigue behaviour of fibreresin composites, Developments in


reinforced plastics2, (edited by G. Pritchard) Chapter 4, (Applied Science, New York), pp.
67107.
(20) KIM, R. Y., 1988, Fatigue Behavior Composite design. Fourth edition, (edited by S. Tsai),
(Think Composites, Dayton OH).
(21) STINCHCOMB , W. W. and BAKIS, C. E., 1991, Fatigue behavior of composite laminates,
Fatigue of Composite Materials, (edited by K. L. Reifsnider), Composite Materials Series,
Vol. 4, (Elsevier, New York).
(22) van DELFT, D. R., RINK, H. D. and JOOSSE, P. ., 1994, Fatigue behaviour of fibreglass
wind turbine blade material in the very high cycle range. Proceedings of the 5th European
Wind Energy Conference, Vol. I, pp. 379384.
(23) SENDECKYJ, G. P., 1981, Test methods and design allowables for fibrous composites,
ASTM STP 734, (edited by C. C. Chamis), (ASTM, Philadelphia) p. 245.
(24) THOMAS, D. R. and BAIN, L. J., 1969, Technometrics, 11, 805.
(25) WHITNEY, J. M., 1981, Fatigue offibrouscomposite materials, ASTM, p. 133.
(26) HAHN, H. T. and KIM, R. Y., 1976, Fatigue behaviour of composite laminates, J.
Composite Mater., 10, 156.
(27) TURCIC, B., 1995, personal communication.
(28) MANN, N. R., SCHFER, R. E. and SINGPURWALLA, N. D., 1974, Methods for
statistical analysis or reliability and life data, (John Wiley, New York).

CHAPTER 15

Recommendations for Good Working


Practices, Norms, and Standards
R. M. Mayer*
A set of recommendations is outlined for design against fatigue. This can form the basis of a
check list for good working practices and possibly part of a structural design code.

15.1 Introduction
Existing rules and regulations relate primarily to the use of reinforced plastics
in boats (1)(2), aircraft (3), and wind turbine blades (4). None deal with the
requirements of fatigue in any depth, but they do form the basis on which one
can develop a set of recommendations for design against fatigue.
The recommendations cover material selection and characterization, design,
manufacture, structural testing of coupons and components, and type approval
(Table 15.1). Where possible, reference is made to existing guidance.
15.2 Material selection and characterization
rl) Identify fully the reinforcement
Fibres are characterized by their type, composition, size, finish, and tex; fabrics
are characterized by their construction, types of fibres, and their orientation
(5). It is important to identify the source of bothfibresand fabrics and to ensure
that the source and type are not changed without notification. A variety of
fibres and fabrics have been used in the present study (Table 1.3).
A variety of wood veneer species have been screened (Chapter 7); these
should be characterized by their ply thickness and the type/number of knots. In
production, plies should be visually graded and any sections having a high
density of knots should be rejected.
r2) Specify reinforcement
Numerous types of reinforcement fabrics are now available. In addition to the
traditional woven fabrics, the study has evaluated stitch bonded and inlaid
Sciotech, UK.

228

Design of Composite Structures Against Fatigue


Table 15.1

Recommendations for design against fatigue

Malerials' selection and characterization


(1)
(2)
(3)
(4)
(5)
(6)

Identify fully the reinforcement


Specify reinforcement
Demonstrate compatibility of resin/hardener system
Select additives and curing agents
Use limit state analysis to determine the safe working loads
Orientate and select reinforcement to withstand principal loadings

Manufacture
(7) Develop a manufacturing procedure to produce good quality components in a reliable
manner
(8) Check degree of cure of GRP components by Barcol hardness
(9) Check compatibility of the reinforcement for manufacture and component geometry
(10) Verify design requirements
(11)
(12)
(13)
(14)
(15)

Coupon testing
Validate test methods
Characterize relevant material properties
Evaluate damage accumulation and failure criterion
Measure response to fatigue loading
Determine design allowables

Structural testing
(16) Evaluate fatigue response of sections, components or sub-assemblies
(17) Verify structural integrity by fatigue testing
(18) Evaluate the accumulation of damage during testing and mode of failure
Type approval
(19) Develop type approval procedure for volume production

fabrics with the possible addition of a layer of random mat (combi-mat). The
reinforcement and lay-up sequence must be specified unambiguously, using a
systematic classification, such as used in this book, in order to characterize the
composite fully.
With wood, the veneers are normally laid up along the grain, but if any are
cut on the bias, this also needs to be specified.
r3) Demonstrate compatibility of resin/hardener system
The interface between the reinforcement (fibre or veneer) and resin is loaded
quite severely under fatigue.
It is, therefore, recommended that properties be measured which characterize (by stressing) the interface. This includes measuring interlaminar shear and
testing in flexure and compression (5). The property levels should be compared
with other similar material combinations (6).
A variety of resins can be used with the same fabric reinforcement, but this
may depend upon the fibre size and finish (Chapter 3).
As wood species differ in their ability to take up resin, the compatibility of

Recommendations for Good Working Practices, Norms, and Standards

229

the resin with the veneer should be checked, as well as the optimum ratio of
resin to wood.
r4) Select additives and curing agents
For epoxy resins, the hardener system and cure cycle is specified at the same
time as the base resin by the manufacturer. For polyester resins, a wide variety
of curing agents and additives (fillers, retardants, pigments) can be used. It is
essential that the selected combination can be fully cured to ensure a good level
of properties. The manufacturer's advice should be sought where necessary.
A variety of additives and curing agents have been used and there is no
indication of any systematic influence on properties.
15.3 Design
r5) Use limit state analysis to determine the safe working loads
The general design principles set out in ISO 2394 (7) have been adapted for use
with the structural Eurocodes and wind turbines. Though fatigue forms only
one part of the analysis, it could well form the critical aspect of the design if the
structure moves or rotates.
Where the partial safety factors are not specified in the codes or national
application documents, those design aspects that impinge on safety need to be
assessed (Chapter 1).
The use of finite element analysis is recommended to determine the stress
and strain fields caused by complex loading (Chapter 13).
r6) Orientate and select reinforcement to withstand principal loadings
Fatigue in fibre reinforced plastics tends to be dominated by the fibres, rather
than the matrix (Chapter 3). Under these circumstances, emplacing some
fibres in the principal load directions is very beneficial in suppressing matrix
cracking, thereby imparting good fatigue resistance (Chapters 2 and 14).
A good level of static strength properties generally leads to a good level of
fatigue properties. This allows various material combinations to be screened
rapidly. The combination giving the best level of static properties should then
be checked for its fatigue resistance.
15.4 Manufacture
r7) Develop a manufacturing procedure to produce good quality components
in a reliable manner
For a demanding duty cycle like fatigue loading, a suitable process has to be
evolved, which will produce a component with a consistent quality. Various
manufacturing processes (Table 15.2) have been evaluated and characterized

230

Design of Composite Structures Against Fatigue


Table 15.2

Manufacturing process evaluated

Process

Components

Contact moulding

Test plates and components


wooden blades shear webs
Test plates and
blade root
Test plates
Spars, blade root
GRP blades

Prcprcg moulding
Filament winding
Tape winding
Resin transfer moulding

Location
(chapter)
3,6,9,10,12
7
4
13
2
9
9

through testing coupons and components. Data have been derived on the basis
of fibre alignment and volume fraction and it can be concluded that all
processes are capable of producing components of a consistent quality for
fatigue loading.
For design verification, both non-destructive and destructive techniques may
be required. A good level of quality control is essential.
r8) Check degree of cure of GRP components by Barcol hardness
The component should be fully cured so that the resin is stable over long
periods of time and against thermal cycling. Barcol hardness, being a handheld micro-hardness indicator (8), enables the degree of cure to be monitored
over the component: its value can then be checked with the manufacturer's
recommendation. The indentor can also be used to check whether the component is ready to be released from the mould (Fig. 15.1).
The versatility of this technique was demonstrated in the case of spars
manufactured with transverse filament tape (Chapter 9) as the components
were postcured until the Barcol hardness reached the recommended level.
r9) Check compatibility of the reinforcement for manufacture and component
geometry
Fabrics should be selected for their ability to drape, handle, and pack as well as
for their fibre orientation. This can be ascertained by laying up the fabric dry.
After manufacture, sections should be examined to check that the reinforcement is located correctly with neither resin-rich nor fabric-rich areas. If
necessary, the process should be adjusted, the fabric changed, or the geometry
modified.
For example, the development of an inlaid fabric for the tape winding
process (Chapter 9) began with the design of the tape to suit the mandrel
geometry and ended with the design of a tape tensioning device to ensure that a
constant load was applied to the tape during the winding process.

Recommendations for Good Working Practices, Norms, and Standards

231

Fully cured

Fig. 15.1 Barcol hardness as a function of time

rlO) Verify design requirements


To verify the design requirements, checks should include:
(a) shape and dimensional accuracy;
(b) fibre distribution and volume fraction particularly in areas of high loading
or rapidly changing curvature;
(c) physical properties like mass and natural frequency;
(d) a strain survey by loading the component statically (Chapters 9 and 13).
This allows a direct verification with the design predictions; any significant
deviation should be checked.
15.5 Coupon testing
rll) Validate test methods
As there are few standard test methods for fatigue testing of coupons and none
for components, the requirement for validation is self-evident.
The benchmark method for coupons (Chapter 8) permits correlation of
methods for tensile loading. Similar tests have yet to be undertaken for
compressive testing and reverse loading.
Environmental testing is time consuming to undertake and the results on
moisture absorption, property retention, and impact set some limits on
property changes (Chapters 5 and 11). The translation of such results to

232

Design of Composite Structures Against Fatigue

components is more problematic as the free surface to volume ratio and


thickness is very different. For safety critical components in aircraft, tests have
had to be conducted on large panels as well as coupons (3).
rl2) Characterize relevant material properties
The relevant properties like tension, flexure, compression, and shear should be
measured along the principal fibre directions and in off-axis directions where
loading is appreciable. Results should always be compared with theoretical
values, those from the literature (6), or standard fatigue curves (Chapter 14).
Sets of data have been derived to characterize a range of composite materials
used in structural design. It is noted that the ratio of the tensile, flexural and
compressive properties is not always predictable. For wood, the compressive
strength is always lower than the tensile strength and both strengths are
dependent on the density of knots (Chapter 7).
rl3) Evaluate damage accumulation and failure criterion
Of the various techniques to monitor damage accumulation, visual inspection,
and reduction in stiffness are the most suitable and are therefore recommended.
The most significant finding is that the stiffness reduces slowly until close to
the point at which failure is initiated (Figs 2.5a, 3.6, 6.2, 9.7). These observations appear to be independent to a large extent of the number of cycles (Fig.
15.2) or of loading (Fig. 15.3). This is in accordance with the three stages of
damage proposed by Reifsnider (9) (Fig. 1.2).
This is a major difference, for example, with metals where the stiffness
remains constant until failure is imminent. Moreover, the modulus of glass
reinforced plastic materials is much lower than of metals. For GRP materials it
therefore follows that components are generally stiffness rather than strength
limited.
It may, therefore, be appropriate for a specified loss of stiffness to be a design
criterion that must be fulfilled. For example, this criterion has been used in
Chapter 9 and is shown in Figs 6.5 and 6.7.
The change in hysteresis damping on loading and unloading and the heating
due to enhanced internal friction are other possible methods. These are more
difficult to correlate with damage accumulation; internal heating has been
successfully used in early detection of possible damage areas (Chapters 7 and
10).
rl4) Measure response to fatigue loading
The measurement of a complete data set is time consuming because of the
many parameters.
It is recommended that the minimum requirements be the ultimate static

Recommendations for Good Working Practices, Norms, and Standards

233

0.95

0.90

S 0.85
at
0.80

0.75

Fig. 15.2

WIN
Reduction in stiffness as a function of fatigue life. Stiffness normalized to its initial
value and number of cycles to that of failure. A omox = 250 Mpa, cycles to failure 4,806,
o^na = 75 MPa, cycles to failure 23 106. Benchmark coupons tested under reverse
loading at ECN (Chapter 8)

properties in tension and compression and the fatigue properties of the most
appropriate loading regime. For rotor blades, for example, reverse loading
with equal and opposite magnitudes in tension and compression would be
appropriate.
The rationale for this approach is discussed in Chapter 7 (wood) and Chapter
4 for GRP. This limited data set may not always give conservative values, and
so additional data might be needed, depending upon the severity of the
loading. A method to evaluate, from a small set of data, whether standard
fatigue curves can be used is given in Chapter 14.
rl5) Determine design allowables
The detailed analysis in Chapter 14 indicates how data sets may be used to
establish values of the design allowables from coupons. The use of standard
fatigue curves to initiate a design is also discussed.
It is recommended that the values of the design allowables and partial
coefficients be determined according to the size and scatter of the data set in
accordance with standard statistical analysis. This process is described in
Chapter 5 where a design limit of 0.6 percent strain has been derived for a glass/
epoxy material.
The effect of spectral loading and its relation to constant amplitude loading
using Miner's rule has been considered in Chapter 4. Under some conditions,

234

Design of Composite Structures Against Fatigue

0.95

0.90

Life WIN,)

(a)

V OT

'

0.4

0.6

Life WIN,)

(b)
Fig. 15.3 Reduction in stiffness as a function of fatigue life for two different types of loading:
(a) flexure (UWE); (b) tension (ECN). Benchmark coupons (Chapter 8)

Recommendations for Good Working Practices, Norms, and Standards

235

the rule will lead to non-conservative values of the fatigue life. Thus, at present
such summations should be treated with caution, and verified by testing.
Other effects that should be considered are:
(a) multi-axial loading;
(b) scaling up from thin coupons to structures with thick sections;
(c) environment.
15.6 Structural testing
Testing of components and structures can be a lengthy process. In some
applications it might be essential, for example GRP leaf springs where four axis
rig tests with the appropriate spectral loading is deemed an essential requirement before vehicles can be track or road tested.
The pyramid of substantiation (Fig. 1.4) requires the largest number of tests
to be carried out on coupons, a smaller number on sections or sub-assemblies,
and ultimately a test of the entire structure. It is recommended that this also be
followed for other types of composite structures.
rl6) Evaluate fatigue response of sections, components, or sub-assemblies
It is prudent to evaluate the fatigue response of critical sections of a structure.
Such sections might comprise regions that
(a) are highly stressed as determined from the design or by a strain survey
when the structure is loaded;
(b) require a change in material type, such as a composite to a metal;
(c) possess sharp curvature which may cause difficulty with the drape of a
fabric or veneer;
(d) contain a joint.
Wood veneers have to be jointed and the type of joints have been evaluated
at coupon level by Ansell (Chapter 7) and on full size blades by Hancock (12).
It is for these reasons that a wide range of blade roots have been examined
(Table 1.2). All require a change in material type from composite to steel or
aluminium and load transfer across materials with different physical and
mechanical characteristics. Both commercially available and prototype designs
have been successfully evaluated.
Accelerated loading can be used on components (Chapters 9-13) to validate
the design or to determine the weakest link in new design concepts provided
that the higher imposed loads do not invoke different failure mechanisms.
rl7) Verify structural integrity by fatigue testing
For load-bearing structures, it is necessary to test a structure if its failure could
endanger life or result in damage to the environment. Composite structures can
be no exception to this rule. Evidence shows that structural testing is meaningful because it:

236

(a)
(b)
(c)
(d)

Design of Composite Structures Against Fatigue

creates confidence in the concept and the design;


verifies the manufacturing procedure;
indicates how damage accumulation can be detected;
establishes the failure mode(s)

Such testing can, therefore, be used as a means of demonstrating the safety


of the design and mitigating any public liability claims that might subsequently
arise if a mishap did occur. The codes allow a lower value of the partial safety
coefficient if the structure can be shown to fail in a safe manner (4). As
composites are often stiffness-limited (discussion in rl3), it should be possible
to design structures in these materials which will fail safe.
The magnitude of the fatigue stress and the number of load cycles depend
upon the nature of the loading. For structures such as rotor blades, a fatigue
test of at least one million cycles is recommended. If testing is carried out for a
shorter period at higher loads, failure could be initiated by a mechanism not
necessarily characteristic of failure at a lower stress (Chapter 9).
If the structure is still in a sound condition after the initial test, it is
recommended that testing should continue at a higher stress level (Chapters 9
and 10). This would have the added advantage that it would be possible to
check the prediction of Miner's rule of summing fatigue amplitudes (Chapter
10).
rl8) Evaluate the accumulation of damage during testing and mode offailure
It is recommended that the accumulation of both local and global damage
should be monitored during testing of a structure to determine whether failure
occurs in a fail safe manner.
Techniques which are useful in monitoring structural damage include:
(a) visual inspection to monitor damage nucleation and growth;
(b) strain gauging in areas of high strain;
(c) thermocouples - adjacent to strain gauges recording high strains (Chapter
9);
(d) overall stiffness;
(e) natural frequency and internal damping;
(f) infra-red emissions to identify damage areas (Chapter 10).
Using one or more of these techniques, it should be possible to monitor the
various damage stages and thereby determine when the structure has failed.
15.7 Type approval
rl9) Develop type approval procedure for volume production
In practice it has been found that volume produced composite structures have
not always performed as well as prototypes. As with boats (1), it is desirable to

Recommendations for Good Working Practices, Norms, and Standards

237

adopt a type approval scheme to ensure that good quality components can be
produced in volume.
With the advent of the single European market, such a scheme would need to
be European in origin in order to comply inter alia with the relevant requirements of the various EU directives and rules (5).
The techniques and data in this book could be used for such a scheme by
providing information to the designer, manufacturer, surveyor, and test
laboratory. The recommendations set out in this chapter would form only part
of such a procedure and would need to be adapted for a particular type of
structure and application.
It is likely that a European type approval scheme for rotor blades will evolve
in the next few years. This will form part of the plan to develop renewable
energies in order to reduce fossil fuel emissions and meet an increasing demand
for electricity with minimal effect on the environment.
15.8 Conclusions
A set of recommendations is proposed, covering the design of composite
structures against fatigue. They can be used to supplement methods already in
use for designing structures to withstand static loading. The relevance of such
recommendations will depend upon the application and the degree of structural integrity required.
References
(1) Rules for the classification and certification of vessels of less than 15 metres in length, 1983,
1990, (Det Norske Veritas, Oslo).
(2) Rules and regulations for the classification of yachts and small craft, 1980, (Lloyds Register,
London).
(3) Composite aircraft structures (section 2), Acceptable means of compliance of composites
made from reinforced plastics, 1986, Civil Aviation Authority, Gatwick.
(4) Wind turbine generator systems. Part 1 - safety requirements, 1994, IEC 1400-1, (IEC,
Geneva).
(5) MAYER, R. M., 1993, Design with reinforced plastics, (Chapman & Hall, London) pp. 102107.
(6) HANCOX, N. L. and MAYER, R. M., 1993, Design data for reinforced plastics, (Chapman
& Hall, London).
(7) General principles on reliability for structures, ISO 2394, 1986 (ISO, Geneva).
(8) Measurement of hardness by means of a Barcol impressor EN 59,1989, (BSI, Milton Keynes).
(9) REIFSNIDER, K. F. (editor), 1991, Fatigue of composite materials, (Elsevier, Amsterdam).
(10) Leighton Harris, 1994, Private communication.
(11) SCHNEIDER, K. and LANG, R. W., 1990, Secondary source qualification of carbon fibre
prepregs for primary and secondary Airbus structures, 1990, Eleventh International SAMPE
Conference.
(12) HANCOCK, M. et al., 1993, Fatigue testing of 12m wind turbine blades, Proceedings of the
Fourteenth BWEA Wind Energy Conference, (edited by B. R. Clayton) (MEP, London).
(13) Bart Roorda, 1994, Private communication.

Glossary
Accelerator
Additive

Barcol hardness
BS
Catalyst
CEN
Combination
fabric
Composite
Consolidation
Contact
CRP, CFRP
CSM
Delamination
DIN
Drape
DSC

A material that, when mixed with a catalyst or resin, speeds


up the curing process.
A substance added to the resin to either polymerize it, such
as accelerator, initiator, or catalyst; or to improve resin
properties, such asfillersor flame retardants.
Micro-hardness indentation technique.
A British standard.
A material that, when added in small quantities, increases
the rate of cure of a resin.
Comit Europen de Normalisation
Usually a layer of chopped fibre added to an underlying
fabric by powder bonding, stitching, or needling.
A generic term to describe the mixture of a fibrous reinforcement or wood veneer and resin matrix.
A process by which the composite is compressed, such as a
roller, vacuum or pressure to eliminate air bubbles and
achieve desired properties.
A process in which the composite is laid up in a mould.
Carbon fibre reinforced plastic.
Chopped strand mat, a form of fabric with randomly
orientated short length fibres.
Separation between two or more layers of a composite due
either to incorrect processing or subsequent degradation
during use.
Deutsches Institut fr Normung; a German standard.
The ability of a fabric to conform to a shape or surface.
Differential scanning calorimetry.

EN
Epoxy resin

European standard.
A thermosetting resin widely used with high-performance
fibres.

FEA
Fibre

Finite element analysis.


A material in filamentary form which has a small diameter
compared with its length.

240

Design of Composite Structures Against Fatigue

Filament
Filler

See Fibre.
An inert material added to a resin to alter its properties or
to lower cost or density; generally in the form of a fine
powder.
Material used to coat filament bundles; usually contains a
coupling agent to improve the fibre to resin bond, a lubricant to prevent abrasion, and a binder to preserve integrity
of a filament bundle.
Filament winding - a fabrication method involving winding
fibres or tapes onto a mandrel.

Finish

FW
Gel-coat
Gl-Ep
Gl-Up
GCRP
GRP, GFRP

Resin-rich layer on the surface of a GRP component.


Glass/epoxy composite.
Glass/polyester composite.
Glass/carbon hybrid fibre reinforced plastic.
Glass reinforced plastic.

Hardener
HDT

A substance added to a resin to promote curing.


Heat distortion temperature (note: varies according to load
applied during the test).
A composite containing two (or more) types of reinforcement.

Hybrid
I
IEC
ILSS
Interface
ISO
Iso-polyester
Iso-NPG
polyester
Mandrel
Mat
Matrix
Mould
NDE, NDT

Inlaid, a form of fabric manufacture in which the rovings


are not kinked or crimped.
International Electro-technical Commission.
Interlaminar shear stress.
The area between the fibre and the matrix.
International Standards Organization; an international
standard.
Isophthallic type of polyester resin.
Isophthallic neo pentyl glycol-type of polyester resin,
A form of mould (usually associated with a component
having cylindrical symmetry).
A material consisting of randomly orientated fibre bundles
which may be chopped or continuous, loosely held together
with a powder binder, by needling or by stitch bonding.
The resin in which the fibrous reinforcement is embedded.
The tooling in which the composite is placed to give the
correct shape to the article while the resin cures.
Non-destructive evaluation, non-destructive testing.

Glossary

241

NPG-polyester

A type of polyester resin formed using neopentyl glycol.

Ortho-polyester

Orthophthallic type of polyester resin.

Palmgren-Miner
Ply
Polyester
Prepreg

A rule for summing the damage for a loading spectrum.


A single layer in a laminate.
A thermosetting resin widely used with glass fibre.
An intermediate fibre reinforced plastic product (lightly
impregnated with resin) which is ready for manufacture
into a component, either in the form of sheet or tape.
Prepreg moulding A process by which prepreg material is moulded either by
autoclave or vacuum bag.

R
Reinforcement
Resin
RH
Roving
SB
Size

Ratio of minimum to maximum stress.


Fibres or wood veneers with desirable properties which,
when mixed together, increase the properties of the host
matrix.
The polymeric material in which the reinforcement is embedded.
Relative humidity.
A number of yarns, strands, or tows which are collected
into a parallel bundle with little or no twist.
Stitch bonding, a method of fabric manufacture in which
the rovings are held together by fine yarns.
A treatment applied to yarns or rovings to protect their
surface and facilitate handling.

TFT

A unit of linear density equal to the mass (in grams) of 1000


metres of filament or yarn.
Transverse filament tape.

UCS
UD
UFS
UTS

Ultimate compressive stress.


Unidirectionally aligned fibres or fabric.
Ultimate flexural stress.
Ultimate tensile stress.

Vinyl ester

A type of thermosetting resin with good mechanical


properties.
Fraction of fibres per unit volume of composite.

Tex

Volume fraction
W

Weaving, a method of fabric manufacture in which glass


rovings are interlocked.

242

Design of Composite Structures Against Fatigue

Warp
Weft
WISPER
WISPERX

Yarns running lengthwise in a fabric.


Yarns running transversely in a fabric.
A load spectrum characteristic of rotor blades.
A truncated version of WISPER in which the low load
cycles are omitted.
A plot of maximum stress versus the number of cycles in a
fatigue test.
Woven roving: a coarse woven fabric usually made from
glass fibre tows.

Whler plot
WR
Yarn

A bundle of filaments that have been twisted: generally


used for processing into fabrics.

Index
Angle-ply laminates 23-24
Anti-buckling devices 18, 20, 35, 54, 68,
74,77
Appel-Olthoff relationship 59, 61, 95
Baltic pine see Pine
Barcol hardness 230
Beech/birch in turbine blades 116-118
Benchmark tests 18,124-133
Blade root:
complex loading on 181-194
loads 185-187
test parameters 187-188
design 9-10
rectangular 185-192
testing 149-170
Miner's rule 168
Bolted joints 4, 11, 93,173-174
British Standard see Standards
Carbon fibres in GFRP 89-90,101
Chopped strand mats 15, 34
Components:
curing 230
evaluating fatigue response 235
Composites:
defined 1
properties 10-11
with long fibres 15-31
materials 16-18
Compression tests 79-111
Compressive loading, glass fibre
composites 220
Constant:
amplitude fatigue 55-58
glass-polyester joints 173-175
wood composites 110-115
life diagram 55-57,112
glass fibre laminates 95
Contact moulding 2
Coupon:
shape 17, 67, 91,109
testing 123, 137-138
tests -v- spar tests 133-147

Cracks 12
Critical fatigue modulus, FRP 41-42
Curing of components 230
Damage:
accumulation 4, 138-141, 143-146,
167-169, 232, 236
development 40
monitoring 138-141
Data analysis 11-12
Design:
against fatigue 6, 228
allowables, recommendations 85-86,
232-235
blade root 9-10,170,193
international code 8
of materials 10-11
recommendations 229-239
rotor blade 8-9
rules 2, 30-31, 47-49, 63, 83-86,112,
170
structural 6-10
ENERCON rotor blades, tests on 198203
Environment, effects of 65-89
Epoxy resin 10, 52, 65,172
Eurocodes 8
Fabric:
construction 10
used for benchmark tests 125-126
standards 230
Fabrics, effect on fatigue properties 4446
FACT 216
Failure 42, 97
Fatigue:
behaviour, GFRP 97-100
bolted joints 104,178-180
considerations 4-6
coupon -v- spar tests 133-147
curves, glass composites 209-225
data, recording 222
of glass fibre laminates 39-43, 94-100

244

Design of Composite Structures Against Fatigue

Fatigue cont.
humidity 79
hybrid fibre laminates 100-103
impact 80
life impact 44, 98
loading 95-97, 129-130, 232-233
properties of 38-40,137
standard curves 220-224
tests:
GFRP bolted joints 177-180
in harsh conditions 81-82
rotor blade roots 195-205
wood composites 110-120, 129-133
Fibre reinforced plastics see FRP
Fibres:
alignment 10, 15, 23-27
identifying 227
orientation 30-31
Fick's law 71
Flanges, blade root testing 151-167
Flexural:
loading, glass composites 220-222
tests 136-137
for benchmarking 127-128
FRP 33-49
cycling above/below linear limit 40-41
damage development 40
failure 47-49
fatigue properties 38-40
ply properties 37-38
stress-strain curves 35
testing 33-36
GFRP see Glass fibre composites
Glass:
/carbon polyester:
bolted joints 93
coupons 92-93
composite laminates, fatigue curves
209-225
/polyester coupons 100-192
fatigue 94-97
Glass fibre composites 15-31
carbon fibres in 89-90
compressive loading 220
constant life diagram 95
fabrication 17
fatigue of 94-100
GFRP 89
effect of moisture on joints 171-180
materials in 16-17
reinforced polyester see GFRP
test coupons 17
testing 52-54, 59-61

Glossary 239-242
Goodman diagram 95-96,112
Hailstones, effects of 65, 70-72
Hybrid:
fabrics 11
fibre laminates, fatigue 100-103
Humidity, effects of 10-11, 65, 68-70,
172
Hysteresis:
damping 12
loops, testing 28
IEC 1400-1 see Standards
ILSS 76-77
Impact damage 71-73
Infra-red:
emissions 12
monitoring of joints in wood 119-120
Inlaid fabrics 34, 135
International design code see Standards
Jointing 4
Joints, weakness 149
bolted joints 11
glass-polyester bolted joints 173-175
in wood composite blades 118-121
Khaya see Wood composites
Laminates, testing 33-36
failure 42-43
Linear:
fit curves, FRP 48
regression analysis 210-212
Load deflection curve 12
Loading 31
Local strains 12
Logarithmic fit curves, FRP 48
Mandell relationship 59, 95
Manufacturing standards 229-231
Material properties, recommendations
232
Materials 10-11
for long fibre composites 16-17
Matrix:
construction 10
cracking, FRP 41-42, 46
effect on fatigue properties 43-45, 48
Mechanical testing 18-30
Miner's rule 51, 55-56, 61-62
blade root tests 168
wood composite blades 115
Modulus, linear limit 36, 40

Index

Moisture:
content see Humidity, effects of
effect on joints 171-180
result of tests 176-180
Monotonie see Static
Non-destructive evaluation (NDE) 12-13
Palmgren-Miner's rule see Miner's rule
Partial safety coefficients 8
Pin-hole flanges, blade root 9, 151-156,
182,188
Pine in turbine blades 116-118
Polyester resin:
iso 10, 16, 34, 90,134
NPG-iso 34,124
ortho 10, 34
Ply properties 37-38
Prepreg moulding 2
Properties:
composites 1
with long fibres 15-31
ply, FRP 37-38
R ratio, effect of 57,114
Rainflow count, wood composites 115
Recommendations see Standards
Reinforcements 36
Residual strength tests 202-203, 206-207
Resin:
transfer moulding 4
Resins:
comparison of 36
compatability of 228-229
Resins, static properties of 36
Resonant frequency 12
Reverse loading 213-218
Rotor blades:
design 8-9, 89
ENERCON 198-199
see also Blade root
Sendeckyj method 82
SN curves:
glass composite laminates 209
wood composites 114-115
Spar tests -v- coupon tests 133-147
spar:
manufacture 134-136
testing 141-147
Spectral loading 11, 51-63
Standards 227-237
BS 7000 2

245

IEC 1400-1 8, 11
ISO 2394 8, 229
coupon testing 231-235
damage accumulation 232, 236
in design 229
design allowables 233
fatigue loading 232-233
in manufacture 229-231
structural testing 235-236
type approval 236
Static:
properties 21, 36, 74-78, 92-93, 109,
137
testing:
flat plates 137
glass fibre composites 21
Statistical analysis 80-83, 209-214
Stiffness reduction 12, 27-28, 33, 42, 45,
94-95
Stitch bonded fabrics 34
Strain:
gauges 141,152,156,166, 186, 200
life 46, 76
Stress 46
in harsh conditions 76
Stress-strain curves 22, 35, 38, 41
Structural:
design 6-10
materials selection 8
testing 235-236
Tape winding, spar manufacture 134
T-bolt:
flange, blade root tests 165-167
root attachment 9, 195-208
Tensile:
loading, glass composite laminates
218-219
testing, wood composites 126-127
Testing:
blade roots 149-170
flanges 151-167
procedure 199-200
D spars 141-145
ENERCON rotor blades 198-203
hysteresis loops 28
Testing of composites:
compression 73, 79-81
fatigue t. on wood composites 129-133
flexural testing 125-126
FRP 33-36
glass fibre 18-30
cyclic testing 21-27
design implications 30-31

246

Design of Composite Structures Against Fatigue

Testing of composites: glassfibrecont.


hysteresis loops, testing 28
monotonie testing 21
load conditions 51-63
tensile 93,126-129
using WISPER/WISPERX 59-61
Tropical conditions, simulating 71-72
Trumpetflanges9,156-165
Turbine blades see Rotor blades
Type approval 236-237
Units xiv
Variable amplitude tests, joints 175
Vinyl ester resin 10, 34, 56
WEG turbine blade 107-109

Weibull statistics 222


Wind Energy Group see WEG turbine
blade
Winding technique 2-4
glassfibrecomposites 17
WISPER/WISPERX 51-63,115,175
hybrid fibre laminates, tests 102
pin-hole flange tests 188
tests with 59-61
wood composites 115
Whler curves 56
Wood composites 107-121
constant amplitude fatigue 110-115
joints, behaviour of 118-120
size effect 113-114
species other than Khaya 116-118

Principal authors and addresses


S. I. Anderson
M. Ansell
P. Bach
A. Echtermeyer
C. Kensche
J. Lameris
H. Lilholt
G. Lowe
R. Mayer
M. Poppen
W. Sys

Materials Department, Ris National Laboratory, PO Box


49, DK-4000 Roskilde, Denmark.
Materials Science Department, University of Bath,
Ciaverton Down, Bath, BA7 7AY, United Kingdom.
Energie Centrum Nederland, Postbus 1, NL-1755 ZG
Petten, Netherlands.
Det Norsk Veritas Research, PO Box 300, N-1322 Hvik,
Norway.
Deutsche Forschungsanstalt fr Luft und Raumfahrt,
Pfaffenwaldring 38-40, D-7000 Stuttgart 80, Germany.
Nationaal Lucht-en Ruimtevaartlaboratorium, PO Box
153, NL-8300 AD Emmeloord, Netherlands.
Materials Department, Ris National Laboratory, PO Box
49, DK-4000 Roskilde, Denmark.
Engineering Department, University of the West of
England, Coldharbour Lane, Frenchay, Bristol, BS16
1QY, United Kingdom.
Sciotech, 9, Heathwood Close, Yateley, Camberley,
Surrey, GU17 7TP, United Kingdom.
Flygtekniska Forsokanstalten, PO Box 11021, S161 11
Bromma, Sweden.
University of Gent, Sint-Pieternieustraat 41, B-9000 Gent,
Belgium.

The use of composite materials for load bearing


structures is gaining wider acceptance as more
knowledge is gained about their performance
and durability. The idea is intrinsically simple to utilize two or more constituents, which
together have more attractive properties than
they possess individually.
Design of Composite Structures Against Fatigue
examines the use of two types of composite
materials - fibre reinforced plastics and wood
composites - and extends their range of
applications to structures which experience
both static and dynamic loading. The study is
applied to the inclusion of such materials into
wind turbine blades.
This volume has been prepared with authors
from seven European countries and includes:
Materials selection
Design verification
Component testing
These are covered systematically
principal material combinations.

for

the

With its low environmental impact, the


generation of electricity from renewable energy
sources is increasingly providing the needs for
the next millennium. The information contained
in Design of Composite Structures Against
Fatigue will allow the next generation of wind
turbines to be developed based on new designs
of rotor blades.
This
volume
can
wholeheartedly
be
recommended to engineers, designers, and
materials specialists, especially those involved in
the wind power field.
Dr Rayner Mayer is the leading consultant
with Sciotech, which is involved in product
innovation and development. He is a leading
researcher in the field of composite materials.
Widely published, he also collaborates across 12
European countries.
ISBN 0 85298 957 1

You might also like