You are on page 1of 29

This article was downloaded by: [Dr. L. F. M.

da Silva]
On: 08 September 2014, At: 13:16
Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

The Journal of Adhesion


Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/gadh20

Composite Repair in Wind Turbine


Blades: An Overview
a b b c
K. B. Katnam , A. J. Comer , D. Roy , L. F. M. da Silva & T. M.
b
Young
a
School of Mechanical, Aerospace and Civil Engineering (MACE),
University of Manchester , Manchester , UK
b
Irish Centre for Composites Research (IComp), Materials and
Surface Science Institute (MSSI), University of Limerick , Limerick
City , Ireland
c
Department of Mechanical Engineering , Faculty of Engineering,
University of Porto , Porto , Portugal
Accepted author version posted online: 10 Apr 2014.Published
online: 05 Sep 2014.

To cite this article: K. B. Katnam , A. J. Comer , D. Roy , L. F. M. da Silva & T. M. Young (2015)
Composite Repair in Wind Turbine Blades: An Overview, The Journal of Adhesion, 91:1-2, 113-139,
DOI: 10.1080/00218464.2014.900449

To link to this article: http://dx.doi.org/10.1080/00218464.2014.900449

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the
“Content”) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or
howsoever caused arising directly or indirectly in connection with, in relation to or arising
out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &
Conditions of access and use can be found at http://www.tandfonline.com/page/terms-
and-conditions
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014
The Journal of Adhesion, 91:113–139, 2015
Copyright # Taylor & Francis Group, LLC
ISSN: 0021-8464 print=1545-5823 online
DOI: 10.1080/00218464.2014.900449

Composite Repair in Wind Turbine


Blades: An Overview

K. B. KATNAM1, A. J. COMER2, D. ROY2, L. F. M. DA SILVA3,


and T. M. YOUNG2
1
School of Mechanical, Aerospace and Civil Engineering (MACE),
University of Manchester, Manchester, UK
2
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

Irish Centre for Composites Research (IComp), Materials and Surface Science
Institute (MSSI), University of Limerick, Limerick City, Ireland
3
Department of Mechanical Engineering, Faculty of Engineering,
University of Porto, Porto, Portugal

Renewable energy sources such as wind energy—together with


energy-efficient technologies—are essential to meet global energy
demands and address climate change. Fiber-reinforced polymer
composites, with their superior structural properties (e.g., high
stiffness-to-weight) that allow lightweight and robust designs, play
a significant part in the design and manufacture of modern wind
turbines, especially turbine blades, for demanding service
conditions. However, with the current global growth in onshore=
offshore wind farm installations (with total global capacity of
282 GW by the end of 2012) and trend in wind turbine design
(7–8 MW turbine capacity with 70–80 m blade length for
offshore installations), one of the challenges that the wind energy
industry faces with composite turbine blades is the aspect of struc-
tural maintenance and repair. Although wind turbines are
typically designed for a service life of about 20 years, robust struc-
tural maintenance and repair procedures are essential to ensure
the structural integrity of wind turbines and prevent catastrophic
failures. Wind blades are damaged due to demanding mechanical
loads (e.g., static and fatigue), environmental conditions (e.g.,
temperature and humidity) and also manufacturing defects. If
material damage is not extensive, structural repair is the only
viable option to restore strength since replacing the entire blade

Received 8 December 2013; in final form 27 February 2014.


Address correspondence to K. B. Katnam, School of Mechanical, Aerospace and Civil
Engineering (MACE), University of Manchester, G12, Pariser Building, PO BOX 88, Manchester
M60 1QD, UK. E-mail: Kali-Babu.Katnam@manchester.ac.uk.

113
114 K. B. Katnam et al.

is not cost-effective, especially for larger blades. Composite repairs


(e.g., external and scarf patches) can be used to restore damaged
laminate=sandwich regions in wind blades. With composite mate-
rials in the spar (30–80 mm thick glass=carbon fiber laminates)
and aerodynamic shells (sandwich sections with thin glass fiber
skins and thick foam=wood as core), it is important to have reliable
and cost-effective structural repair procedures to restore damaged
wind blades. However, compared to aerospace bonded repairs,
structural repair procedures in wind blades are not as well
developed and thus face several challenges. In this regard, the area
of composite repair in wind blades is broadly reviewed to provide
an overview as well as identify associated challenges.
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

KEYWORDS Composite materials; Non-destructive testing;


Structural repairs; UV curing; Wind turbine blades

1. WIND ENERGY INDUSTRY: COMPOSITE MATERIALS

Reliable energy supply is critical to global socio-economic development


and sustainable economic growth. Fossil fuels (e.g., coal, oil, gas) currently
supply most of the global energy needs [1,2]. However, to combat the
global issues of climate change, greenhouse gases and energy insecurity,
renewable energy sources (e.g., wind, solar, bio, hydro, and marine ener-
gies) and energy-efficient technologies are essential to meet the increasing
global energy demands [3,4]. In recent decades, wind energy has been
recognized as one of the key renewable energy sources. Wind energy
can play an important role in reducing the dependency on fossil fuels
and provide opportunities for green electricity generation [5–9]. The wind
energy industry has in recent years seen a significant growth through
onshore and offshore wind farms—with individual turbine capacity ranging
from 1 to 7 MW (e.g., Vestas V-90 3.0 MW, Gamesa G-128 4.5 MW, Siemens
SWT6-120 6 MW, Samsung S7-171 7.0 MW and Enercon E-126 7.5 MW) and
the development of larger turbines with 8–10 MW for offshore installation
(e.g., Vestas V-164 8 MW). The 2012 global wind power market grew by
more than 10% (compared to 2011); nearly 45 GW of new installed wind
power represents investments of about 456 billion, according to the global
wind energy council (GWEC) [10]. The total global installed wind power
capacity at the end of 2012 was 282.5 GW (109.6 GW in Europe,
97.57 GW in Asia, 67.57 GW in North America, and 7.76 GW in the rest of
the world) [10].
Wind flows from the regions of high pressure to low pressure (due to
such factors as uneven solar heating, the Coriolis effect, and local geographical
Composite Repair in Wind Turbine Blades 115

conditions). The kinetic energy in air is converted to electrical energy


(P ¼ 0.5aqAv3, where a is an aerodynamic efficiency constant (which is
0.593, the Betz limit), q is the density of air, A is the swept area of blades,
and v is the wind velocity) by using wind turbines [11]. The three-blade hori-
zontal-axis wind turbine (HAWT), see Fig. 1(a), is the most commonly used
design for power generation. The major components of a HAWT turbine are
a tower, rotor blades, and nacelle (houses components such as the gear box
and generator). As the capacity of the wind turbine depends on the swept area
of blades, manufacturing of lighter, larger, and reliable turbine blades is critical
to increase the efficiency and capacity of modern wind turbines. For this
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

FIGURE 1 Modern composite wind turbine blades: (a) a HAWT with three composite blades
and (b) a typical cross section of an adhesively bonded composite blade.
116 K. B. Katnam et al.

reason, wind turbine blades grew in size significantly in recent decades [12,13].
Currently the largest wind turbine blade is 75 m long with a swept area of
18,600 m2 (Siemens SWT-6-154 6 MW), while even bigger blades for offshore
turbines are under development (e.g., Vestas V-164 with 80 m blades and
21,100 m2 of swept area).
Polymer composite materials, which can be used to manufacture
lighter and larger turbine blades, offer several opportunities in the wind
energy industry. From a structural design viewpoint, composite materials
have many advantages such as high strength-to-weight ratio, high
stiffness-to-weight ratio, fatigue tolerance, corrosion resistance, formability
(i.e., easily mouldable to complex shapes), tailored mechanical properties,
and low thermal expansion. Wind blades are typically manufactured using
E-glass fibers (with 10–15 mm diameter, 2.5–2.6 g=cm3 density, 70–
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

75 GPa Young’s modulus and 3500 MPa tensile strength) or combined


E-glass and carbon fibers (with 6–8 mm diameter, 1.7–1.8 g=cm3 density
and 220–240 GPa Young’s modulus and 4000 MPa tensile strength) with
epoxy or polyester resins (with 1.1–1.3 g=cm3 density, 2–4 GPa Young’s
modulus and 50–100 MPa tensile strength) [14–16]. Core materials (with
0.05–0.25 g=cm3 density) such as balsa, polyvinyl chloride foam, styrene
acrylonitrile foam, or polyethylene terephthalate foam are employed in
sandwich panels [17]. Resin infusion and pre-preg processes are predomi-
nantly used to manufacture composite wind blades [14–19]. Thermoset
resins are the most widely used matrix materials (together with glass
fibers) in the manufacture of wind blades. For larger blades, carbon fibers
are often used in the regions of high load levels (e.g., spar caps) to
exploit their superior mechanical properties (i.e., strength-to-weight and
stiffness-to-weight factors) [20]. However, for structural affordability and
safety, some of the major challenges for large composite wind blades
are reducing material and manufacturing costs, ensuring manufacturing
quality (i.e., repeatable and defect-free processes), developing efficient
joining technologies, developing reliable design rules, preventing and
detecting in-service damage, and improving structural maintenance and
repair technologies [14–19]. Globally, research and development activities,
with core objectives such as cost-effectiveness, reliability, and automation,
are currently underway to improve the existing manufacturing approaches
with novel processes and material technologies for producing larger wind
turbines.
The objective of this paper is to provide an overview of structural
bonded repair requirements and challenges in composite wind blades. An
overview of wind blade design and integrity is presented in Section 2 (includ-
ing blade design and service loads, and common damage modes); the chal-
lenges and opportunities associated with the aspects of damage detection
and bonded repair are outlined in Section 3 (including damage assessment
and repair materials and fabrication).
Composite Repair in Wind Turbine Blades 117

2. COMPOSITE WIND BLADES: STRUCTURAL INTEGRITY

2.1. Blade Design and Service Loads


Modern composite wind blades are typically designed by combining aerody-
namic shells with a load-carrying spar; see Fig. 1(b) for a cross section of a
typical wind blade with adhesively bonded top and bottom shells (note: inte-
grated design without adhesive bondlines is also in use) [14–16]. Structural
adhesives (e.g., epoxies or polyurethanes) are commonly used for bonding
the shells at the leading and trailing edges, and the spar caps to the shells
(see Fig. 2(b)) [21]. Wind blades are often subject to complex loads (e.g.,
static, fatigue, transport=installation loads) under demanding service con-
ditions (e.g., temperature, moisture, erosion, lightning strike, bird strike).
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

As the turbines get bigger, the magnitude of these service loads also increases
significantly (e.g., the weight of the blade significantly increases with increas-
ing length, thus increasing gravity and inertial loads). Moreover, for offshore
installations, service conditions will be considerably different to that of
onshore installations. A wind blade is typically subject to flap-wise and
edge-wise bending, gravitational loads (mainly generate edge-wise bending),
torsional loads (because the shear resultants do not go through the shear
center of the blade section), axial loads due to the rotation of the blade
(inertia forces), and loads due to pitch decelerations and accelerations
[14–16,18,22].

FIGURE 2 Some common damage modes in composite wind blades.


118 K. B. Katnam et al.

The lift force generated by the aerodynamic profile of the blade causes
flap-wise bending (i.e., either static or dynamic). The natural variations in
wind spend cause variations in flap-wise bending moments and thus cause
blade fatigue. Normally, the maximum wind speed for operation of the wind
turbine is about 25 m=s (i.e., storm conditions) [14–16]. Above this speed
limit, the rotor is brought to a standstill by turning the rotor blades out of
the wind by rotation about the longitudinal axis of the blade. This position
exposes the rotor blades to the wind, coming in a right angle to blade planar
surface, and causes flap-wise bending [18]. The blades are also subjected to
gravitational forces, which are most pronounced when the blades are in a
horizontal position, resulting in edge-wise bending moments, which vary
with blade rotation and cause fatigue. From a design viewpoint, the global
stiffness of the blade must be sufficiently high in order to avoid collision with
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

the tower during service [16]. In addition, the local stiffness should ensure
that the aerodynamic profile of the blade is maintained for aerodynamic
efficiency. Flap-wise and edge-wise loads drive the structural design and
the blade cross sections. The spar carries most of the flap-wise bending;
the edge-wise bending is primarily carried by the leading and trailing edges
of the aerodynamic profile. For the design of modern wind turbines, the
International Electrotechnical Commission (IEC) established the IEC 61400
series of standards to consolidate different local schemes (e.g., developed
in Denmark, Germany, and The Netherlands) [16].

2.2. Composite Failure Behavior


Structural design rules for advanced composites, in comparison to metallic
materials, are not yet fully mature [23]. As a fiber-reinforced composite
material is a microstructure in itself, with several glass or carbon fibers held
together by a polymer material, the mechanical properties of the fiber,
matrix, and fiber matrix interface mainly contribute to the composite proper-
ties and failure mechanisms [24]. A weak fiber-matrix interface can lead to a
low stiffness and strength but high resistance to fracture, whereas a strong
interface produces high stiffness and strength but often a low resistance to
fracture [25]. The failure behavior depends not only on inherent heterogen-
eity and anisotropy but also on possible failure modes and their interactions
[26,27]. This complex failure behavior is a major issue associated with the
development of a robust failure criterion that incorporates all possible failure
mechanisms with accuracy [28]. It is also worth noting that the failure beha-
vior of composite materials strongly depends on strain rates as well as
environmental conditions [29,30]. Damage in composite laminates can be at
a lamina scale (i.e., matrix-cracking, fiber-matrix debonding, fiber-breakage),
laminate scale (i.e., delamination), or structural scale (i.e., extensive compo-
nent damage), which can occur due to mechanical and environmental con-
ditions during service. Importantly, impact loads often induce considerable
Composite Repair in Wind Turbine Blades 119

subsurface damage (i.e., at intra-lamina and inter-lamina levels) in composite


structures, with very limited visible surface damage. It is thus important to
consider different length scales (i.e., intra-lamina, inter-lamina, laminate,
and component levels) in the analysis and design of composite structures
[31,32]. Understanding the failure mechanisms and their interactions at each
length scale is critical for the development of robust design rules [33]. More
importantly, cost-effective manufacturing processes often introduce defects
(e.g., wrinkles, fiber misalignment, dry spots, disbonds, voids [16]), which
are common in composite wind blades, and thus can lead to complex failure
behavior. For repeatability and reliability, robust design rules and controlled
manufacturing processes are vital for composite wind blades—considering
the growth rate of wind industry as well as the blade size.
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

2.3. Blade Reliability and Maintenance


Wind blades are manufactured at production sites, and subsequently trans-
ported and then installed at wind farms. In addition to manufacturing defects,
wind blades can accumulate material damage during transportation and
installation. Non-destructive inspection of wind blades, at the production site
as well as at the tower before and after installation, help improve blade
reliability in service. During service, structural damage can initiate from
manufacturing defects, occur due to mechanical loads (e.g., extreme wind,
fatigue, impact), and=or environmental loads (e.g., moisture, temperature,
lightning strike) [15,16]. Some of the common damage modes in wind blades
are shown in Fig. 2. Erosion of the leading edge is a common issue. Although
gelcoat erosion is often considered as non-structural damage, it will not only
reduce the aerodynamic efficiency of the blades considerably, but can also
lead to moisture ingression in structural elements and consequently degrade
the material properties. Lightning strikes cause serious damage to composite
materials such as surface damage, delamination, and heat damage [34–36].
Moreover, manufacturing defects such as wrinkles, misalignment, and
porosities can lead to failure mechanisms [37] such as fiber fracture, interface
cracking, and matrix cracking, which are sources for micro-buckling,
transverse cracking, and delamination [16,38].
Wind blades can be exposed to impact loads during transportation,
installation, and in service. In general, composite structures are inherently
brittle (the fibers are brittle, and so is the matrix when compared with ductile
metals) and can only absorb energy in elastic deformation and through
damage mechanisms—making them sensitive to impact damage [39,40]. A
considerable reduction in compressive, tensile, and shear strength is often
caused by impact damage, depending on the impact energy and impactor
diameter [41,42]. Blunt impacts can induce subsurface damage without
visible surface damage [43]; it is thus difficult to identify such damage during
visual inspections [44]. Impact loads can cause delamination, skin-core
120 K. B. Katnam et al.

debonding, core failure as well as micro-damage such as fiber-breakage and


matrix-cracking [18]. Impact damage can cause significant reduction in local
stiffness and strength. Fiber-breakage, unlike matrix-cracking, can be critical
as mechanical properties are dominated by fiber reinforcement. Structural
damage to composite wind blades needs to be assessed by using suitable
non-destructive techniques to determine the extent and location of the dam-
age [45]. If damage is not widespread and extensive, structural repair is the
only feasible solution as replacing a damaged blade is not cost-effective
[46,47]. Non-destructive inspection methods used in a manufacturing
environment can be more stringent than the methods that can be practically
suitable for in situ inspection [48]. Non-contact non-destructive techniques
that can scan large areas with high speed and detect subsurface damage
are ideal for wind blade inspection [45]. Moreover, as the reduction in
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

material strength depends on the type and size of damage, accurate damage
detection and quantification are essential for structural repair. In addition,
continuous structural health monitoring (SHM) and regular maintenance
checks, together with timely structural repairs, are therefore critical to ensure
the integrity of composite wind blades and turbine as a whole.

3. STRUCTURAL REPAIR OF COMPOSITE WIND BLADES

Wind turbines are expected to have a service life of about 20 years [15,16].
But structural maintenance is critical to extend the life of ageing=damaged
wind turbines and also to ensure that they last the expected service life
[49]. The composite blades, being the largest dynamic structural components
in a wind turbine, are especially prone to structural damage during service.
Minor damage such as surface erosion can considerably reduce aerodynamic
efficiency and thus power generation; but subsurface structural damage
could lead to an unpredictable blade failure during service. From an econ-
omic viewpoint, regular maintenance and repair of ageing wind blades are
essential to prevent serious blade failures as well as lower the cost of wind
energy in order compete with the cost of energy generated from fossil fuels
[50]. However, structural maintenance and repair of large wind blades
poses several challenges: accessibility for inspection and repair (e.g., in situ
inspection), damage identification and assessment (e.g., non-destructive
testing), repair conditions (e.g., weather and wind speeds), repair materials
and procedures (e.g., patch fabrication and curing), and prior knowledge
of the blade’s structural and material details (i.e., the OEM’s design and
material data). Moreover, in comparison to onshore wind turbines, offshore
installations present significant logistical challenges for maintenance and
repair [51].
In general, when a defect or damage is found either at production site or
prior to installation or in service, an appropriate structural repair should be
Composite Repair in Wind Turbine Blades 121

performed to improve the reliability of wind blades. The primary objective


of any structural composite repair is to restore the strength and stiffness of
the defect or damaged region, and bring back its original service condition
(i.e., structural and operational efficiency) [47,48]. Depending on the type
and location of the defect or damage as well as repair site (e.g., up tower
or on ground), structural composite repair can be performed using bonded
doublers or scarf patches. Bonded doubler repairs can restore structural
strength, but they cannot offer an aerodynamically smooth surface. How-
ever, scarf repair can offer structural strength as well as a flush surface,
and thus have greater potential for composite repair. Manufacturing defects
such as excessive fiber waviness and dry spots (i.e., resin-free fiber
regions), and in-service damage such as delamination, can be repaired
using scarf patches. Structural repair of damaged sandwich shells often
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

require replacing the damaged core, and bonding a replacement core


and skin sheet. Delamination in spar caps can structurally be critical and
thus require scarf repairs. External doublers and single-side scarf patches,
as shown in Fig. 3, are particularly suitable for wind blade repairs as access
to damage regions is limited to only one side.

FIGURE 3 External doubler and single-side scarf repairs in composite laminates (a, b) and
sandwich panels (c, d).
122 K. B. Katnam et al.

External doubler repairs are relatively easy to perform as no material is


removed from the parent structure, but they are more suitable for thin cross
sections and do not provide flush surfaces. Scarf patches are, in general, suit-
able for permanent repair of thick cross sections; however, the repair needs
accurate processing techniques and trained technicians to precisely
implement repair procedures [48]. Single-side scarf repairs can be conducted
with access to one side of the damaged region, and they can also provide
effective stress transfer and a good aerodynamic surface finish. However,
structural repair in composite wind blades—because of their size, location,
damage modes, and cross-sectional dimensions (e.g., the thickness of shells
and spar caps)—has several challenges [49] such as damage assessment,
material removal (e.g., scarfing the laminates=cutting the core), surface
preparation (i.e., clean and active surface for bonding), repair materials
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

(e.g., easy to store and process), and equipment to achieve the required cure
temperature and pressure.

3.1. Damage Assessment


3.1.1. VISUAL INSPECTION
Composite wind blades need to be assessed for structural damage to perform
an adequate bonded repair. Visual inspection is the most obvious approach
to assess surface damage (e.g., visible cracks, dents, and punctures). But vis-
ual inspection of in-service blades (with tower heights typically 2–3 times
of the blade length, i.e., 50–150 m) is not straightforward and is often car-
ried out from ground level to identify visible damage (e.g., leading edge ero-
sion, damage caused by lightning strike, excessive trailing edge debonding)
using binoculars or digital cameras with long-range lenses [52]. Visual inspec-
tions can also be performed at close range using either ropes suspended
from the top of the nacelle or automated tower climbing machines [53,54].
However, damage assessment using external visual inspection techniques
cannot identify the internal damage in the blades and thus require internal
inspection techniques, which are complicated to perform on in-service
blades.

3.1.2. ULTRASONIC TECHNIQUES


When ultrasonic waves travel through composite materials, the wave propa-
gation is influenced by internal damage (e.g., delamination or disbonding),
which acts as discontinuities and introduces a local change in acoustic impe-
dance. Ultrasonic testing is now an established non-destructive technique for
detecting subsurface damage in composite materials [55,56]. In direct ultra-
sonic techniques, high-frequency waves are generated by a transducer and
transmitted through a test component; then the reflected or transmitted
Composite Repair in Wind Turbine Blades 123

waves are received by the same or a second transducer [57]. High-frequency


waves are more sensitive to defects, whereas, low-frequency waves can pen-
etrate to greater depths [58]. With contact transducers, a thin layer of couplant
is used between the test surface and the transducer to transmit waves without
large attenuation. From a composite repair viewpoint, conventional ultra-
sonic techniques offer several advantages. Ultrasonic testing can detect dif-
ferent defects (e.g., delamination, voids, or disbonding) and indicate the
depth of a defect in a laminate. Scanning systems can be portable to inspect
large areas. Scans with pulse-echo mode can be used with access to only one
side of the test surface. However, one major disadvantage of direct ultrasonic
techniques is the need for a liquid couplant to overcome the acoustic impe-
dance mismatch between air and composite materials [59]. Without a cou-
plant, the majority of the sound energy is lost, and very little is transmitted
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

into the test component. Scanning with a couplant can be complex and often
not practical for in situ inspection of large structures such as wind blades [60].
In contrast to direct techniques, indirect ultrasonic techniques such as
laser-based ultrasound do not require conventional transducers to generate
ultrasonic waves at the test surface for transmission and conversely on recep-
tion [61,62]. A short pulse laser is used to generate waves; a long pulse or a
continuous wave laser is used as an indirect approach to detect the waves.
When a laser pulse strikes a composite material, it rapidly expands the
material locally and creates a thermo-elastic ultrasonic pulse [63]. Interfer-
ometry (e.g., Fabry–Perot) can be used to detect ultrasound waves [64].
The main advantages of laser ultrasound technique are non-contact inspec-
tion, couplant-free scanning, inspection with access to only one side of the
test component, and inspection of complex shapes from greater distances.
On the other hand, a laser-based ultrasonic inspection system, which
requires two laser systems and interferometer, is expensive in comparison
to the conventional transducer-based systems [61]. Moreover, as material-
induced wave phenomena are used to optically detect damage, sensitivity
of laser ultrasound techniques could be an issue for wind blade inspection.

3.1.3. DIGITAL SHEAROGRAPHY


Optical non-destructive testing techniques, when compared to conventional
ultrasonic techniques, allow large surfaces to be inspected relatively quickly,
and are thus appropriate for composite wind blades. Shearography, which is
an interferometric technique, uses coherent laser illumination for surface
deformation measurements (i.e., displacements and displacement deriva-
tives) non-destructively [65,66]. Compared to holography, which measures
surface displacements, shearography measures derivatives of surface
displacements and thus provides surface strains [67]. The technique has
significantly been improved in recent years with advancements in charge-
coupled device (CCD) cameras, lasers, and computing hardware. Digital
124 K. B. Katnam et al.

shearography is a non-destructive testing technique suitable for composites


due to its ability to provide non-contact, full-field measurements. The
technique uses the coherent, monochromatic properties of laser light to-
generate speckle patterns on the test surface. The speckle patterns on the test
surface are recorded—one image when the specimen is unstressed and one
with an applied stress. The technique requires an image shearing device
(e.g., Michelson interferometer) in front of the CCD camera [68]. When the
test component is subject to an applied load, subsurface damage will exhibit
strain anomalies compared with the regions that are free from damage. Digi-
tal shearography can be used to detect the damage-induced surface strain
anomalies. Common loading techniques include thermal loading (e.g., high
power flash lamps), pressure loading (e.g., vacuum), or vibrational exci-
tation. The inspection results are shown in real time to the user, and systems
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

can be configured to automatically detect=measure defects in composite


wind blades—including delamination, disbonding, impact damage, wrinkles,
and dry spots [65]. Digital shearography is particularly effective in revealing
damage such as delamination and skin-core debonding in the shells of com-
posite wind blades. Digital shearography increases inspection speed of large
structures such as wind blades [45,60] and also enables non-destructive
testing of bonded repair patches [69]. However, as shearography directly
measures surface strain anomalies in the component, the success in detecting
damage=defects depends on their size and location (large defects in thin
sections are sensitive to surface loading) [65–70]. Subsurface damage in thick
laminate regions (e.g., spar caps can be 30–80 mm thick, depending on the
size of the blade and the location of the cross section) can be insensitive to
surface loading and thus cannot be detected using shearography.

3.1.4. PASSIVE AND ACTIVE THERMOGRAPHY


Similar to digital shearography, thermography allows large surface scans in
relatively quickly and is thus well suited for the inspection of composite wind
blades. Thermography is a non-contact technique based on infrared radiation
for detecting material damage or defects [71]. In thermography, which can be
either passive or active, temperature gradients are measured to detect
material damage non-destructively [67]. While passive thermography is used
for composite components that are at a different temperature than its
surroundings, active thermography involves heating the component surface
rapidly by using an external heat source and observing how the temperature
decays with time [72,73]. As composite materials possess a relatively low ther-
mal conductivity, thermographic techniques are well suited for the inspection
of composite structures such as wind blades [56,74]. Infrared cameras are
used to capture thermal images; advanced software is then used to process
these images for detecting subsurface defects. In relation to wind blade
repair, subsurface damage in composites (i.e., delamination, debonding,
Composite Repair in Wind Turbine Blades 125

cracks, or moisture) affects local thermal conductivity and manifests in local


temperature gradients within the damaged region [71]. The detected subsur-
face damage can be quantified in terms of depth and size. The changes in
temperature during day and night could potentially enable wind blade
inspection using passive thermography for detecting damage [75,76]. The
major advantages of thermography are non-contact non-destructive inspec-
tion with access to only one side, inspection of large and complex surfaces
in quick time, and data processing in pictorial format for rapid decisions.
However, active thermography is relatively expensive with current tech-
nology as highly sensitive thermal cameras and external heat sources are
required. In addition, the technique could be less sensitive to subsurface
defects in thicker laminates common in composite wind blades.
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

3.1.5. SHM SYSTEMS


Unlike non-destructive testing (often scheduled inspection), SHM systems for
wind blades can provide continuous real-time data directly to a central
control=monitoring room often located at a significant distance from the
wind farm [34,35], and thus help identify damaged blades and plan
condition-based structural repairs [77]. Blade SHM systems typically feature
contact sensors (as opposed to non-contact sensors) attached directly to
either the blades or the rotor shaft. Data acquisition hardware is located
within the turbine nacelle or tower or rotor hub, and the signals from the
blade sensors can be monitored remotely over the turbine communication
network. For example, strain sensors attached directly to the blade are sensi-
tive to local changes in strain. As structural blade damage affects the bending
moments induced through changes in the stiffness of the blades, any
phenomena affecting both the flap-wise and edge-wise bending moments
close to the root of the blades can be detected if the sensors are strategically
positioned [78]. However, with regard to realizing a fully integrated real-time
SHM system for blades, there are some major technical challenges
[15,16,17,34,35,77,78]: installation of sensors (during manufacture or retrofit-
ting of sensors); reliability of sensors and associated hardware for a lifetime
of about 20 years; and robust systems for data communication, storage, pro-
cessing (local=remote), and analysis. The integrated nature of blade SHM sys-
tems means that they need to be installed during manufacture of the turbine.
Retrofitting of sensors in the blades is not desirable because of the difficulty
involved.
Modern blade SHM systems employ sensors to measure bending and
torsional loads in the main shaft which can be indicative of issues with the
rotor blades; or measure strains at specific locations in the blades; or monitor
vibration=noise characteristics of the blades under operational loads. Typical
options include fiber optics, strain gauges, acoustic emission sensors, and
vibration sensors. Fiber optics have been chosen over traditional strain
126 K. B. Katnam et al.

gauging technology mainly because they are immune from high cycle fati-
gue, lightening-induced damage and electromagnetic interference [79]. But
drawbacks include interference with the manufacturing process (typically
infusion) as the fiber-optic sensors are typically embedded within the
laminates [80,81]. Moreover, measurement is local, and the sensors may
not be sensitive to damage that is remote from the sensor location. Similar
to fiber optics, strain gauges are typically employed to measure local surface
strains at the root of the blade [82], but they suffer environmental and fatigue
damage leading to disbond. Acoustic emission sensors can be strategically
employed to locate and monitor damage initiation and propagation
[83–85], but significant user experience is required to analyze and interpret
data [86]. Moreover, vibration sensors can be used to detect damage by
comparing the pre-damage signature measurements to the post-damage
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

measurements [34,35,87], but data processing and interpretation is not


straightforward. In this regard, non-contact approaches for monitoring wind
blades are under development. For example, photogrammetry can be used
to determine the displacement coordinates of a wind blade under normal
operating conditions by capturing images continuously from different loca-
tions and processing simultaneously [88]. Retroreflective markers are used
at various locations on the surface of the blade and are used as displacement
sensors. It is however important to note that SHM techniques often help
identify the structural condition of large wind blades, but may not be
able to accurately locate the damage regions and material failure modes.
Condition-based non-destructive inspection is often required to assess
damage and perform structural repairs.

3.2. Repair Preparation


The damaged material needs to be removed from the regions to be repaired.
Depending on the type of damage modes identified, this may include remov-
ing surface layers (e.g., gelcoat), damaged laminates (skin) and damaged
core (in sandwich regions) [53,54]. While external doubler repairs can be per-
formed without extensive material removal, scarf repairs require machining
of the damaged laminates to achieve tapered sections for patch fabrication.
Manual tools such as sanders and routers are typically used to remove
damaged materials. But machining of composite laminates=sandwich skins
can introduce material damage (e.g., delamination near machined edges)
because of their inherent heterogeneity and anisotropy [89–91]. Moreover,
manual machining of curved surfaces is complex to remove damage and
achieve desirable scarf angles. On the other hand, repair processes (e.g., dust
and debris from machining) and pre-repair conditions (e.g., moisture present
in damaged skins and core) often contaminate composite surfaces prior
to bonding. As surface properties (i.e., physical and chemical surface
conditions) are important for promoting strong interface bonds, surface
Composite Repair in Wind Turbine Blades 127

preparation is essential in order to have a surface that is conducive for


bonded repair.
The drying, cleaning, and preparation of composite surfaces prior to
bonding ensure strong interface adhesion. In general, factors such as low
surface energy, chemical inertness, surface contamination, and a weak
boundary layer often contribute to poor wettability and weak interface
adhesion in polymer composites [92]. A misconception about surface prep-
aration prior to bonding is that a clean surface (e.g., free from dust, lubri-
cants, or other surface contaminants) is the only requirement for better
adhesion. A clean surface is only a necessary condition for adhesion, but
not sufficient to achieve strong interface bonds. Good adhesion requires
very close contact; adhesive needs to flow and wet the surfaces to be
bonded [93]. To ensure good surface wetting, the surface energy of the
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

adherend must be higher than that of the adhesive used for bonding. Sur-
face preparation is thus essential to modify the physical and chemical
properties of the surfaces, increase surface energy, and remove contami-
nants and weak boundary layers. Appropriate surface preparation will pro-
mote adherend–adhesive interface bond strength and improve structural
performance (i.e., strength and durability) of adhesively bonded joints. Sur-
faces that are either untreated or insufficiently prepared prior to bonding
could lead to adherend–adhesive interface fracture reducing the joint
efficiency.
With regard to composite repair, the machining process used will
remove material damage and create new surfaces (e.g., tapered scarf
surfaces). The surfaces often require mechanical abrasion to achieve
surface uniformity and solvent cleaning to remove debris, dust, or any sur-
face contaminants. But the type of surface preparation can depend on the
machining process used in creating the surfaces. In general, a variety of
surface preparation (e.g., abrasion=solvent cleaning, grit blasting,
low-pressure plasma treatment) can be used to enhance surface energy,
increase surface roughness, activate surface chemistry, and thereby
increase bond strength and durability of adhesively bonded composite
joints [93,94]. Abrasion=solvent cleaning is the most commonly used tech-
nique for composite adherends to increase the mechanical interlocking of
the adhesive into the adherend by removing contaminants and improving
surface roughness [95,96]. Abrasive processes do not guarantee high sur-
face wettability and surface energy that are required to provide intimate
contact between the adherend and adhesive. Solvent cleaning can remove
dust and debris, but solvent residue could adversely affect the adherend–
adhesive interface properties. Moreover, manual surface preparation can
lead to non-uniform and inconsistent surface properties, and may also
contaminate the surface. In this context, advanced surface treatments
(e.g., atmospheric plasma and laser ablation) may offer opportunities for
wind blade repairs.
128 K. B. Katnam et al.

3.3. Repair Strength and Durability


Bonded composite repairs in wind blades should restore the structural
response (e.g., strength and stiffness) of the damaged region as close to
the original properties as possible. But the structural response of bonded
repairs heavily relies on processing factors (i.e., surface properties, adhesion,
and curing conditions) rather than mechanics (i.e., stress and strain) alone
[48,97]. In general, the stiffness and strength of a bonded repair patch depend
on the type of raw materials, geometrical parameters, and process para-
meters that are used in the repair. The design of bonded repairs requires a
comprehensive analysis of the stress transfer phenomena at the interface as
well as in the two adherends under service loads [98]. The design must also
ensure that the bond between the two adherends transmits the service loads
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

(e.g., tension, compression, fatigue) under environmental (i.e., humidity and


temperature) service conditions [99,100]. Structural repair with bonded
patches should ideally maximize the repair efficiency and minimize the risk
of structural failure by effectively transferring different service loads (e.g.,
static and fatigue) through it. As the structural performance of bonded com-
posite patches depends on material, geometrical and processing parameters,
predicting the behavior of bonded patches is thus very complicated. In
comparison to the response of an external bonded patch, the response of
the structural response of a bonded scarf patch is relatively more complex
because of the stiffness variation along the parent–patch interface [101].
The stress distributions (i.e., peel and shear stresses) vary significantly along
the parent–patch interface, depending on laminate thickness, material
properties, and the stacking sequence [102,103]. In general, the composite
laminate strength in the overlap section of the scarfed lap joint is usually
reduced when compared with the parent laminate strength—due to discon-
tinuous fibers over the overlap length as well as stress concentrations in the
bondline and laminates. In addition, the bond strength could vary, which is a
characteristic feature of adhesively bonded joints (especially with brittle
interfaces), due to variations of the associated material, geometrical, and
processing properties. The repair processes could also introduce different
types of defects (i.e., bondline porosity, laminate porosity, kissing bonds)
[104–109]. It is thus difficult to fabricate bonded patches that can withstand
demanding service conditions without robust repair materials and
procedures.
In general, bonded composite joints may in theory be designed such
that the interface can sustain loads greater than the strength of the adherends,
ensuring that it will be able to sustain the service loads of the original struc-
ture. However, many bonded joints fail in service because of inconsistent
processing methods, while other factors such as deficient design or poor
materials selection could also play a role [99]. Variability in process con-
ditions, which is highly relevant to bonded repairs (especially for in situ
Composite Repair in Wind Turbine Blades 129

patch fabrication in wind blades), can significantly contribute to the poor


performance of bonded patches. Process-induced defects will degrade the
mechanical properties of the patch laminate and bondline, and also generate
stress concentrations in the patch when subject to external loads. As failure
often initiates at stress concentrations, it is critical to use appropriate process
conditions (i.e., uniform surface treatment, uniform cure temperature, and
consolidation) to reduce process-induced defects. As the mechanical proper-
ties of brittle resins=adhesives are often sensitive to porosity and stress con-
centrations, adhesives with low modulus and high ductility could minimize
bondline stress concentrations and thus improve the joint strength and
reduce its variation [110–113]. Toughened structural adhesives can offer
opportunities to enhance damage tolerance of bonded repairs [114,115].
The failure behavior of bonded joints with composite adherends largely
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

depends on the transverse tensile strength (i.e., through the thickness of the
laminate) of the two adherends [96]. As the transverse mechanical properties
of composite laminates (with no through-the-thickness reinforcing fibers) are
relatively lower in comparison to the in-plane mechanical properties of the
laminate, the low transverse tensile strength, which is of the same order or
even lower than that of the matrix, could lead to the adherend failure in
transverse tension before the failure of the interface because of induced peel
stresses in addition to shear stresses along the interface [112]. The peel stres-
ses could damage (e.g., by delamination) the parent or patch laminates and
thereby adversely affect the stress transfer capacity of the repair patch. It is
thus essential to keep the peel stresses below the transverse tensile strength
of the two adherends by using appropriate geometrical parameters and
adhesive types, which offer an enhanced stress distribution in the patch
region.
Bonded repairs are exposed to harsh environmental conditions during
their service life. The long-term performance of composite bonded patches
depends on the structural response of the interface and adherends to fatigue
and environmental conditions. When subject to cyclic loading conditions,
adhesives and resins can accumulate damage (e.g., crazing, shear yielding,
micro-cracking) near stress concentrations and initiate fatigue cracks [116–
118]. In addition, the mechanical properties (e.g., elastic modulus, tensile
strength) of thermosetting adhesives and resins can, in general, be consider-
ably deteriorated when exposed to harsh environments (e.g., humidity, tem-
perature), which will affect the durability of bonded repairs [119–121]. The
absorbed moisture can lead to both reversible and irreversible effects (e.g.,
plasticization, swelling, and degradation) [116]. Adhesives=resins that are less
sensitive to environmental service conditions could improve the long-term
performance of bonded repairs. Furthermore, manual composite repair pro-
cedures are prone to human error, depending on the skills and knowledge of
repair technicians. Human errors and inconsistencies in repair processes can
significantly influence the structural strength and durability of bonded
130 K. B. Katnam et al.

composite repairs. Post-repair non-destructive evaluation of bonded patches


can detect physical disbonds and voids, but cannot identify weak interface
bonds (i.e., kissing bonds). Under-cured adhesive=resin regions are also
difficult to identify. Weak interface bonds could initiate considerable
disbonding of repair patches over time and influence the patch integrity
during service.

3.4. Repair Materials and Processing


Suitable repair materials (i.e., pre-preg, fabrics, replacement core, resins, and
adhesives) and controlled curing conditions (i.e., temperature and pressure)
are essential to fabricate composite repairs in wind blades. The time
required, which is directly related to the turbine downtime, to fabricate
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

and consolidate repairs will have an impact on the associated economical


and operational aspects [52]. Material systems that can be stored at ambient
temperature, cured at low temperature and with short cure cycle time are
therefore ideal for repair applications [53,54], especially for in situ fabrication.
In addition, suitable fabrication techniques are required for adequate proces-
sing, curing, and consolidation of patches without internal defects [122–125].

3.4.1. REPAIR MATERIALS


The material selection process for repairs will however depend on the
material systems used in the manufacturing of wind blades to achieve a fully
compatible patch. The repair materials for wind blades include resins (e.g.,
epoxies and polyester), fiber fabrics=pre-preg (e.g., glass and carbon), and
adhesives (e.g., epoxies and polyurethane). The material requirement also
varies, depending on the extent of damage or the kind of repair. The most
important requirements for blade repair materials are controllable or adjust-
able pot life, low viscosity, low exotherm, low temperature cure and short
cure time [49]. To suit the adversities of on-site repair (i.e., weather con-
ditions, temperature, and location of the damaged blade on high-mounted
tower), suitable material processing conditions are required along with
sufficient strength and adhesion of the cured material. Thermoset resins such
as epoxy, vinylester, unsaturated polyester, and polyurethanes can be used
as repair materials [47]. The possibility of low temperature cure and low vis-
cosity make them suitable candidate for wind blade repairs. Epoxy systems
are stronger with high tensile and flexural strength, while unsaturated
polyester resins have easy processability. Thermoset resins typically do not
reach their ultimate mechanical properties unless processed with appropriate
cure conditions (e.g., low temperature cure) [126]. For example, some grades
of epoxy resins, when cured at low ambient temperatures and high ambient
humidity, can lead to permanent under-cure of the resin and poor
inter-laminar adhesion [127,128].
Composite Repair in Wind Turbine Blades 131

3.4.2. PATCH FABRICATION


Damaged regions can be repaired with a composite patch (e.g., external
doubler or scarf patch, see Fig. 3) fabricated in situ with pre-preg or wet
lay-up (i.e., soft patch approach). In a soft pre-preg patch approach, pre-preg
laminae are cut to required shapes (e.g., to match the cavity for a scarf repair)
and used in conjunction with an adhesive to fabricate a patch [122]. As the
fiber orientation and lay-up influence the mechanical properties of the patch,
each lamina needs to be accurately cut and located while fabricating the
patch. But pre-preg materials and film adhesives usually need to be stored
at very low temperatures to prevent undesirable curing at ambient con-
ditions. Dry fiber fabric can also be used to fabricate a soft patch with wet
lay-up or in situ resin infusion. The resin will also act as an adhesive, which
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

may however lack required interface properties such as ductility and tough-
ness, to bond the patch. The soft patch approaches require in situ curing of
the patch and often need elevated temperatures and pressure to achieve
desired patch consolidation [128]. But, as the curing process would often
have to use a vacuum bag together with a heat blanket, which can result
in low fiber-volume fraction and porosity in the patch and also voids in
the bondline, it can be difficult to achieve patch properties that match those
of the parent component [123–125]. Based on the damage conditions, the
blades can either be repaired up-tower or on ground. For up-tower repairs,
the location of damage=repair and weather conditions could affect the man-
oeuvrability required to fabricate a soft patch. Wrinkling of the co-cured
patch can also be an issue with a soft patch approach. Moisture absorbed
by either the parent materials or repair materials prior to curing can also
introduce considerable defects in the patch and bondlines [129].

3.4.3. CURE TEMPERATURE AND PRESSURE


Cure temperature is a key parameter that can have a direct influence on the
quality of a bonded patch. Controlled heating is required to cure adhesives
and co-cure composite patches. In case of humid conditions, it is necessary
to dry the surfaces and remove the moisture in the sandwich core prior to
repair at elevated temperatures without over heating the damaged region.
Composite bonded repairs, especially when performed in situ, can be diffi-
cult to process as elevated temperatures are often required for curing the
structural adhesives and resins used in the repair patch. For example, a
single-sided heat source is often used to transfer heat through the full thick-
ness of the repair patch to achieve uniform cure of the adhesive and co-cured
patch. But composite laminates typically exhibit poor thermal conductivity in
the through thickness direction, which may lead to a thermal gradient when a
single-side heat source is used during repair [130]. The issue of thermal gradi-
ent can be significant if a substructure beneath the repair patch acts as a heat
132 K. B. Katnam et al.

sink. A thermal gradient could lead to non-uniform cure of the adhesive and
co-cured patch, and consequently introduce non-uniform stress transfer
making the bonded repair ineffective. Complex cure temperature gradients
may also increase the potential for process-induced warpage, residual stres-
ses, and matrix micro-cracking, and micro-delamination of the repair patch
[131–133]; additionally, overheating may locally degrade the parent compo-
nent (e.g., skin-core debonding in sandwich structures). Heat blankets,
which are commonly used for in situ bonded repairs in aerospace industry,
can be difficult for use in wind blade repairs and thus non-conventional
methods such as UV curing can offer new opportunities for wind blade
repairs.
Cure pressure, which is also an important parameter for bonded repairs,
must be adequate to ensure proper bondline thickness, minimize bondline
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

porosity, and cause the resin=adhesive to flow and properly wet the surfaces
[134]. For co-cured composite patches, pressure is required to consolidate the
composite in order to obtain the desired mechanical properties. Pressure can
be applied by using a vacuum bag or mechanical approach. Vacuum
bagging, which is common in aerospace bonded repairs because of its
convenience, can conform to any surface, apply uniform pressure, remove
volatiles, and hold a heat blanket in place. Vacuum pressure will help
remove air entrapped during fabrication and volatile gases during cure, thus
reducing porosity [135]. However, it is not possible to achieve high pressure
by using vacuum bagging alone. This could lead to inadequate consolidation
of co-cured patches and consequently affect the fiber-volume fraction
and mechanical properties. On the other hand, approaches to mechanically
apply pressure need complicated fixtures and must be held against the
structure by a support mechanism [136,137]. Moreover, the curvature of the
repair region could adversely affect the pressure distribution and may not
be uniform.

3.4.4. UV-CURABLE RESINS


UV-curable resins and pre-preg can offer opportunities for shorter cure times
(typically in minutes rather than hours) in bonded repairs and thus reduce
the downtime of the damaged wind turbines [49,54]. Radiation-induced poly-
merization can be used for fast generation of highly cross-linked polymer
networks by adding a photoinitiator that generates reactive species when
exposed to radiation such as UV light (100–400 nm wavelength)
[138,139]. Irradiation can produce free radicals, for radical polymerization
of monomers such as unsaturated polyester, or cations, for cationic polymer-
ization of multifunctional epoxies [138]. As there is no gel time for UV-curable
resins, the processing time is flexible and reduces the waste of raw materials.
UV curing can also offer an improved working environment due to lower
styrene emissions [140,141]. During UV curing, the reactive species are
Composite Repair in Wind Turbine Blades 133

mainly generated at the top layer and initiate resin cure at the top surface,
which can prevent styrene emission. As the photoinitiator gets destroyed
upon UV exposure in the top layer, the incident radiation penetrates deeper
into the laminate and allow the polymerization front to move progressively
towards the inner layers [138]. For non-structural repairs such as leading-edge
erosion, surface defects, or small holes, resin pastes can be applied and cured
within a few minutes with a UV source (e.g., mercury arc lamp) [49]. For
structural repairs, UV-curable pre-preg can be cut to the required size and
used in multiple layers and cured together. In addition, UV-curable resins
can also be used in a wet lay-up process to fabricate a repair patch.
Because UV light can only penetrate optically transparent materials such
as glass fibers and limited depths, UV-curing resins are suitable for wind
blade repairs where glass fiber laminates are used (e.g., sandwich skins).
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

Opaque material regions, such as spar caps (with carbon fibers) and sand-
wich cores, cannot directly be repaired using UV-curable resins [142]. The
resins can be fully cured in the presence of glass fibers because UV light
can penetrate more in glass fiber laminates due to light scattering and wave
guiding effects from the fibers [142–144]. However, the concentration of
photoinitiator used in the resin formulation, the wavelength of the UV light
used, and the irradiation time have a pronounced effect on the rate of poly-
merization, extent of penetration, and cross-link density in the cured resin
[138,139]. While shorter wavelengths (100–300 nm, i.e., UV-C) provide
higher energy, longer wavelengths (300–400 nm, i.e., UV-A) give better
penetration into the laminate [138]. UV-curable resins need to be shielded
from sunlight or other sources of actinic radiation to avoid reactions that
can increase viscosity and eventually lead to polymerization. UV light must
also reach the entire repair region to cure effectively and uniformly [145].
Variables such as lamp wavelength, choice of resin, amount of photoinitiator,
exposure time, and distance from lamp to laminate surface have to be
carefully considered when using UV-curable resins.

4. CONCLUSIONS

With the current growth in global wind farm installations and increasing
wind turbine sizes, structural maintenance and repair is critical for ensuring
structural safety and longevity of wind turbines and also in reducing the cost
of wind energy generation. The safety and efficiency of composite wind
blades will largely depend on SHM, non-destructive testing, and repair tech-
niques, because of the current not-fully-matured composite design rules,
manufacturing processes and joining technologies used in the wind industry
as well as the demanding service conditions that the blades are often
exposed to. But there is a strong need for improving the current composite
wind blade repair technologies in some key areas such as non-destructive
134 K. B. Katnam et al.

testing for damage assessment, surface treatments for interface bonding,


repair materials for patch fabrication, and repair design for patch strength
and durability. Non-destructive inspection is a challenge for accurate damage
assessment because of the blade size (i.e., large surface area and thick cross
sections), turbine location (i.e., onshore=offshore weather conditions) and
different damage modes (e.g., skin delamination, skin-core debonding, core
failure, leading=trailing edge debonding). Contact=non-contact ultrasonics,
passive=active thermography and digital shearography can however offer
opportunities for damage assessment in wind blades. SHM techniques
(e.g., strain sensors, acoustic emission, and photogrammetry) can help per-
form condition-based (rather than scheduled) non-destructive inspections.
In bonded repairs, surface preparation is essential to modify surface proper-
ties, increase surface energy, and remove contaminants. Surfaces that are
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

either untreated or insufficiently prepared prior to bonding could lead to


interface failures and thus reduce the repair efficiency. But wind blade repair
conditions can be a challenge for achieving conducive surface properties
prior to bonding. In addition, suitable materials and cure conditions are also
essential to fabricate reliable repairs. The time required to fabricate and con-
solidate bonded repairs can significantly influence the downtime of damaged
wind turbines. The material systems that can be stored at ambient tempera-
ture and cured at low temperature, with short cycle time, are ideal for wind
blade repairs, especially for in situ fabrication. Soft patch approaches (e.g.,
pre-preg, wet lay-up, and resin infusion processes) and non-conventional
curing techniques (e.g., UV-curable resins) can help address some of the
challenges related to patch fabrication. Moreover, structural adhesives and
resins that can improve interface ductility and toughness can enhance
strength and long-term durability of bonded repairs. But it is worth noting
that the performance of bonded patches depends on processing, material,
and geometrical parameters. Human errors and inconsistencies in repair pro-
cesses can also significantly influence the strength and durability of com-
posite repairs. Post-repair non-destructive evaluation of bonded patches
can detect physical disbonds and voids, but it cannot identify weak interface
bonds and under-cured regions.

ACKNOWLEDGEMENTS

The authors would like to thank Dr. S. Grant, Dr. P. Wales, Dr. T. Murmu,
Dr. A. Gandhi, and Dr. D. Modi for useful discussions.

FUNDING

The authors (KBK, AJC, DR, and TMY) acknowledge Enterprise Ireland for
research funding (CC-2009-0001).
Composite Repair in Wind Turbine Blades 135

REFERENCES

[1] Dorian, J. P., Franssen, H. T., and Simbeck, D. R., Energy Policy 34, 1984–1991
(2006).
[2] Dresselhaus, M. S. and Thomas, I. L., Nature 414, 332–337 (2001).
[3] Asif, M. and Muneer, T., Renewable Sustainable Energy Rev. 11, 1388–1413
(2007).
[4] Armaroli, N. and Balzani, V., Angew. Chem. Int. Ed. 46, 52–66 (2007).
[5] Islam, M. R., Mekhilef, S., and Saidur, R., Renewable Sustainable Energy Rev.
21, 456–468 (2013).
[6] Herbert, G. M., Iniyan, S., Sreevalsan, E., and Rajapandian, S., Renewable
Sustainable Energy Rev. 11, 1117–1145 (2007).
[7] Quarton, D. C., Wind Energy 1, 5–24 (1998).
[8] Esteban, M. D., Diez, J. J., López, J. S., and Negro, V., Renewable Energy 36,
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

444–450 (2011).
[9] Kaldellis, J. K. and Zafirakis, D., Renewable Energy 36, 1887–1901 (2011).
[10] Fried, L., Sawyer, S., Shukla, S., and Qiao, L., Annual Market Update 2012,
Global Wind Report, (Global Wind Energy Council, Brussels, 2013).
[11] Hansen, M. O. L., Aerodynamics of Wind Turbines, (Earthscan, London, 2008).
[12] Marsh, G., Refocus 6, 22–28 (2005).
[13] Red, C., Composite Technology, (Composites World, Web Publisher, June,
2008).
[14] Veers, P. S., Ashwill, T. D., Sutherland, H. J., Laird, D. L., Lobitz, D. W., Griffin,
D. A., Mandell, J. F., Musial, W. D., Jackson, K., Zuteck, M., Miravete, A., Tsai,
S. W., and Richmond, J. L., Wind Energy 6, 245–259 (2003).
[15] Brøndsted, P., Lilholt, H., and Lystrup, A., Annu. Rev. Mater. Res. 35, 505–538
(2005).
[16] Haymen, B., Wedel-Heinen, J., and Brøndsted, P., MRS Bull. 33, 343–353
(2008).
[17] Sloan, J., Composites Technology, (Composites World, Web Publisher, June,
2010).
[18] Thomsen, O. T., J. Sandwich Struct. Mater. 11, 7–26 (2009).
[19] Campbell, F. C., Manufacturing Processes for Advanced Composites, (Elsevier,
Oxford, 2004).
[20] Wood, K., Composites Technology, (Composites World, Web Publisher, June,
2012).
[21] Subrahmanian, K. P. and Dubouloz, F., Reinf. Plast. 53, 26–29 (2009).
[22] Risø=DNV., Guidelines for Design of Wind Turbines, (Jysk Centraltrykkeri,
Denmark, 2002).
[23] Soden, P. D., Kaddour, A. S., and Hinton, M. J., Compos. Sci. Technol. 64,
589–604 (2004).
[24] Barbero, E. J., Introduction to Composite Materials Design, (CRC Press, Boca
Raton, FL, 2011).
[25] Soutis, C., Mater. Sci. Eng. A 412, 171–176 (2005).
[26] Echaabi, J., Trochu, F., and Gauvin, R., Polym. Compos. 17, 786–798 (1996).
[27] Orifici, A. C., Herszberg, I., and Thomson, R. S., Compos. Struct. 86, 194–210
(2008).
136 K. B. Katnam et al.

[28] Catalanotti, G., Camanho, P. P., and Marques, A. T., Compos. Struct. 95, 63–79
(2013).
[29] Daniel, I. M., Werner, B. T., and Fenner, J. S., Compos. Sci. Technol. 71,
357–364 (2011).
[30] Selzer, R. and Friedrich, K., Composites Part A 28, 595–604 (1997).
[31] Llorca, J., González, C., Molina-Aldareguı́a, J. M., Segurado, J., Seltzer, R., Sket,
F., Rodrı́guez, M., Sádaba, S., Mu~ noz, R., and Canal, L. P., Adv. Mater. 23,
5130–5147 (2011).
[32] Mishnaevsky, L., Comput. Mech. 50, 195–207 (2012).
[33] Spearing, S. M., Lagace, P. A., and McManus, H. L. N., Appl. Compos. Mater. 5,
139–149 (1998).
[34] Ciang, C. C., Lee, J. R., and Bang, H. J., Meas. Sci. Technol. 19, 1–20 (2008).
[35] Ghoshal, A., Sundaresan, M. J., Schulz, M. J., and Pai, P. F., J. Wind Eng. Ind.
Aerodyn. 85, 309–324 (2000).
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

[36] Amirat, Y., Benbouzid, M. E. H., Al-Ahmar, E., Bensaker, B., and Turri, S.,
Renewable Sustainable Energy Rev. 13, 2629–2636 (2009).
[37] Toft, H. S., Branner, K., Berring, P., and Sorensen, J. D., Eng. Struct. 33,
171–180 (2011).
[38] Cantwell, W. J. and Morton, J., J. Strain Anal. Eng. Des. 27, 29–42 (1992).
[39] Davies, G. A. O. and Zhang, X., Int. J. Impact Eng. 16, 149–170 (1995).
[40] Richardson, M. O. W. and Wisheart, M. J., Composites Part A 27, 1123–1131
(1996).
[41] Abrate, S., Appl. Mech. Rev. 44, 155–190 (1991).
[42] Prichard, J. C. and Hogg, P. J., Composites 21, 503–511 (1990).
[43] Kumar, P. and Rai, B., Compos. Struct. 23, 313–318 (1993).
[44] Polimeno, U. and Meo, M., Compos. Struct. 91, 398–402 (2009).
[45] Mason, K., High Performance Composites, (Composites World, Web Publisher,
May, 2006).
[46] Wood, K., High Performance Composites, (Composites World, Web Publisher,
November, 2008).
[47] Armstrong, K. B., Cole, W., and Bevan, G., Care and Repair of Advanced
Composites, (SAE International, London, 2005).
[48] Katnam, K. B., Da Silva, L. F. M., and Young, T. M., Prog. Aerosp. Sci. 61, 26–42
(2013).
[49] Wood, K., Composites Technology, (Composites World, Web Publisher, April,
2011).
[50] Blanco, M. I., Renewable Sustainable Energy Rev. 13, 1372–1382 (2009).
[51] Henderson, A. R., Morgan, C., Smith, B., Sorensen, H. C., Barthelmie, R. J., and
Boesmans, B., Wind Energy 6, 35–52 (2003).
[52] Wallace, J. and Dawson, M., Renewable Energy Focus 10, 36–41 (2009).
[53] Marsh, G., Reinf. Plast. 55, 32–36 (2011).
[54] Cripps, D., Reinf. Plast. 55, 28–32 (2011).
[55] Roach, D., High Performance Composites, (Composites World, Web Publisher,
July, 2008).
[56] Scott, I. G. and Scala, C. M., NDT Int. 15, 75–86 (1982).
[57] Blitz, J. and Simpson, G., Ultrasonic Methods of Non-destructive Testing,
(Chapman and Hall, London, 1996).
Composite Repair in Wind Turbine Blades 137

[58] Garnier, C., Pastor, M. L., Eyma, F., and Lorrain, B., Compos. Struct. 93, 1328–
1336 (2011).
[59] Castaings, M., Cawley, P., Farlow, R., and Hayward, G., J. Nondestr. Eval. 17,
37–45 (1998).
[60] Cawley, P., Composites 25, 351–357 (1994).
[61] Dubois, M. and Drake, T. E., Nondestr. Test. Eval. 26, 213–228 (2011).
[62] Blouin, A., Neron, C., Campagne, B., and Monchalin, J. P., Insight 52, 130–133
(2010).
[63] Davies, S. J., Edwards, C., Taylor, G. S., and Palmer, S. B., J. Phys. D Appl. Phys.
26, 329–348 (1993).
[64] Monchalin, J. P., Appl. Phys. Lett. 47, 14–16 (1985).
[65] Hung, Y. Y., Composites Part B 30, 765–773 (1999).
[66] Steinchen, W., Yang, L., Kupfer, G., and Mackel, P., Proc. Inst. Mech. Eng. Part
G J. Aerosp. Eng. 212, 21–30 (1998).
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

[67] Hung, Y. Y., Chen, Y. S., Ng, S. P., Liu, L., Huang, Y. H., Luk, B. L., Ip, R. W. L.,
Wu, C. M. L., and Chung, P. S., Mater. Sci. Eng. 64, 73–112 (2009).
[68] Gryzagoridis, J. and Findeis, D., Insight 52, 248–251 (2010).
[69] Hung, Y. Y., Luo, W. D., Lin, L., and Shang, H. M., Compos. Struct. 50, 353–362
(2000).
[70] Gryzagoridis, J. and Findeis, D., Insight 50, 249–252 (2008).
[71] Avdelidis, N. P., Hawtin, B. C., and Almond, D. P., NDT & E Int. 36, 433–439
(2003).
[72] Almond, D. P. and Peng, W., J. Microsc. 201, 163–170 (2001).
[73] Avdelidis, N. P., Almond, D. P., Dobbinson, A., Hawtin, B. C., Castanedo, C. I.,
and Maldague, X., Prog. Aerosp. Sci. 40, 143–162 (2004).
[74] Dattoma, V., Marcuccio, R., Pappalettere, C., and Smith, G. M., NDT & E Int.
34, 515–520 (2001).
[75] Yang, B. and Sun, D., Renewable Sustainable Energy Rev. 22, 515–526 (2013).
[76] Meinlschmidt, P. and Aderhold, J., Thermographic Inspection of Rotor Blades,
in 9th European Conference on NDT, Berlin, Germany, 2006.
[77] Hameed, Z., Hong, Y. S., Cho, Y. M., Ahn, S. H., and Song, C. K., Renewable
Sustainable Energy Rev. 13, 1–39 (2009).
[78] Schroeder, K., Ecke, W., Apitz, J., Lembke, E., and Lenschow, G., Meas. Sci.
Technol. 17, 1167–1172 (2006).
[79] Majumder, M., Gangopadhyay, T. K., Chakraborty, A. K., Dasgupta, K., and
Bhattacharya, D. K., Sens. Actuators A 147, 150–164 (2008).
[80] Zhou, G. and Sim, L. M., Smart Mater. Struct. 11, 925–939 (2002).
[81] Guemes, J. A. and Menendez, J. M., Compos. Sci. Technol. 62, 959–966 (2002).
[82] Overgaard, L. C. T., Lund, E., and Thomsen, O. T., Composites Part A 41,
257–270 (2010).
[83] Giordano, M., Calabro, A., Esposito, C., D’Amore, A., and Nicolas, L., Compos.
Sci. Technol. 58, 1923–1928 (1998).
[84] Joosse, P., Blanch, M., Dutton, A., Kouroussis, D., Philippidis, T., and Vionis,
P., J. Sol. Energy Eng. Trans. ASME 124, 446–454 (2002).
[85] Schubel, P. J., Crossley, R. J., Boateng, E. K. G., and Hutchinson, J. R.,
Renewable Energy 51, 113–123 (2013).
[86] Blanc, M. J. and Dutton, A. G., Key Eng. Mater. 245, 475–482 (2003).
138 K. B. Katnam et al.

[87] Zou, Y., Tong, L., and Steven, G. P., J. Sound Vib. 230, 357–378 (2000).
[88] Ozbek, M., Rixen, D. J., Erne, O., and Sanow, G., Energy 35, 4802–4811 (2010).
[89] König, W., Wulf, C., Gra, P., and Willerscheid, H., CIRP Ann. Manufact.
Technol. 34, 537–548 (1985).
[90] Sheikh-Ahmad, J. Y., Machining of Polymer Composites, (Springer, New York,
2009).
[91] Sloan, J., High Performance Composites, (Composites World, Web Publisher,
May, 2010).
[92] Kinloch, A. J., Adhesion and Adhesives: Science and Technology, (Chapman
and Hall, London, 1987).
[93] Da Silva, L. F. M., Öchsner, A., and Adams, R. D., Handbook of Adhesion Tech-
nology, (Springer, Berlin Heidelberg, 2011).
[94] Wingfield, J. R. J., Int. J. Adhes. Adhes. 13, 15–156 (1993).
[95] Baldan, A., J. Mater. Sci. 39, 1–49 (2004).
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

[96] Banea, M. D. and Da Silva, L. F. M., Proc. Inst. Mech. Eng. Part L J. Mater. Des.
Appl. 223, 1–18 (2009).
[97] Kinloch, A. J., Proc. Inst. Mech. Eng. Part G J. Aerosp. Eng. 211, 307–335
(1997).
[98] Van Straalen, I. J. J., Wardenier, J., Vogelesang, L. B., and Soetens, F., Int.
J. Adhes. Adhes. 18, 41–49 (1998).
[99] Davis, M. and Bond, D., Int. J. Adhes. Adhes. 19, 91–105 (1999).
[100] Harman, A. B. and Wang, C. H., Compos. Struct. 75, 132–144 (2006).
[101] Wang, C. H. and Gunnion, A. J., Compos. Sci. Technol. 68, 35–46 (2008).
[102] Baker, A. A., Chester, R. J., Hugo, G. R., and Radtke, T. C., Int. J. Adhes. Adhes.
19, 161–171 (1999).
[103] Chalkley, P. and Baker, A. A., Int. J. Adhes. Adhes. 19, 121–132 (1999).
[104] Kimiaeifar, A., Toft, H., Lund, E., Thomsen, O. T., and Sørensen, J. D., Eng.
Struct. 35, 281–287 (2012).
[105] Dıaz, J., Pereira, F., Romera, L., and Hernandez, S., Int. J. Adhes. Adhes. 37,
70–78 (2012).
[106] Sutherland, L. S. and Soares, C. G., Reliability Engineering and System Safety
56, 183–196 (1997).
[107] Adams, R. D. and Drinkwater, B. W., NDT & E Int. 30, 93–98 (1997).
[108] Katnam, K. B., Comer, A. J., Stanley, W. F., Buggy, M., Ellingboe, A. R., and
Young, T. M., Int. J. Adhes. Adhes. 31, 679–686 (2011).
[109] Katnam, K. B., Stevenson, J. P. J., Stanley, W. F., Buggy, M., and Young, T. M.,
Int. J. Adhes. Adhes. 31, 666–673 (2011).
[110] Marques, E. A. S. and Da Silva, L. F. M., J. Adhes. 84, 915–934 (2008).
[111] Da Silva, L. F. M. and Lopes, M. J. C. Q., Int. J. Adhes. Adhes. 29, 509–514
(2009).
[112] Da Silva, L. F. M. and Adams, R. D., Int. J. Adhes. Adhes. 27, 227–235 (2007).
[113] Katnam, K. B., Dhote, J. X., and Young, T. M., J. Adhes. 89, 486–506 (2013).
[114] Kinloch, A. J. and Shaw, S. J., J. Adhes. 12, 59–77 (1981).
[115] Katnam, K. B., Comer, A. J., Stanley, W. F., Buggy, M., and Young, T. M., Int.
J. Adhes. Adhes. 37, 3–10 (2012).
[116] Kinloch, A. J. and Young, R. J., Fracture Behaviour of Polymers, (Applied
Science, London and New York, 1983).
Composite Repair in Wind Turbine Blades 139

[117] Dessureault, M. and Spelt, J. K., Int. J. Adhes. Adhes. 17, 183–195 (1997).
[118] Quaresimin, M. and Ricotta, M., Compos. Sci. Technol. 66, 176–187 (2006).
[119] Crocombe, A. D., Int. J. Adhes. Adhes. 17, 229–238 (1997).
[120] Wahab, M. M. A., Ashcroft, I. A., Crocombe, A. D., and Shaw, S. J., J. Adhes. 77,
43–80 (2001).
[121] Katnam, K. B., Crocombe, A. D., Sugiman, S., and Khoramishad, H., Int.
J. Damage Mech. 20, 1217–1242 (2011).
[122] Whittingham, B., Baker, A. A., Harman, A., and Bitton, D., Composites Part A
40, 1419–1432 (2009).
[123] Sherwin, G. R., Int. J. Adhes. Adhes. 19, 155–159 (1999).
[124] Grunenfelder, L. K. and Nutt, S. R., Compos. Sci. Technol. 70, 2304–2309
(2010).
[125] Pearce, P. J., Arnott, D. R., Camilleri, A., Kindermann, M. R., Mathys, G. I., and
Wilson, A. R., J. Adhes. Sci. Technol. 12, 567–584 (1998).
Downloaded by [Dr. L. F. M. da Silva] at 13:16 08 September 2014

[126] Loos, A. C. and Springer, G. S., J. Compos. Mater. 17, 135–169 (1983).
[127] Brostow, W. and Glass, N. M., Mater. Res. Innovations 7, 125–132 (2003).
[128] Rai, N. and Pitchumani, R., Polym. Compos. 18, 566–581 (1997).
[129] Elaldi, F. and Elaldi, P., J. Reinf. Plast. Compos. 30, 749–755 (2011).
[130] Yang, Z. L. and Lee, S., Mater. Manuf. Processes 16, 541–560 (2001).
[131] Davies, L. W., Day, R. J., Bond, D., Nesbitt, A., Ellis, J., and Gardon, E., Compos.
Sci. Technol. 67, 1892–1899 (2007).
[132] Guo, Z. S., Du, S., and Zhang, B., Compos. Sci. Technol. 65, 517–523 (2005).
[133] Djokic, D., Johnston, A., Rogers, A., Lee-Sullivan, P., and Mrad, N., Composites
Part A 33, 277–288 (2002).
[134] Chester, R. J. and Roberts, J. D., Int. J. Adhes. Adhes. 9, 129–138 (1989).
[135] Tzetzis, D. and Hogg, P. J., Polym. Compos. 29, 92–108 (2008).
[136] Olson, S. Wind Blade Spar Cap Laminate Repair, (US Patent, US 7927077 B2,
2011).
[137] Anasis, G., Bogaert, J., Arellano, R., and Flobeck, G., Method of Repairing a
Wind Turbine Blade, (US Patent, US 0167633 A1, 2011).
[138] Endruweit, A., Johnson, M. S., and Long, A. C., Polym. Compos. 27, 119–128
(2006).
[139] Decker, C., Polym. Int. 45, 133–141 (1998).
[140] Li, G., Pourmohamadiam, N., Cygan, A., Peck, J., Helms, J., and Pang, S.,
Compos. Struct. 60, 73–81 (2003).
[141] Compston, P., Schiemer, J., and Cvetanovska, A., Compos. Struct. 86, 22–26
(2008).
[142] Gupta, A. and Ogale, A. A., Polym. Compos. 23, 1162–1170 (2002).
[143] Shi, W. and Ranby, B., J. Appl. Polym. Sci. 51, 1129–1139 (1994).
[144] Pang, S., Li, G., Jerro, D., and Peck, J. A., Polym. Compos. 25, 298–306 (2004).
[145] Black, S., Composite Technology, (Composites World, Web Publisher, April,
2004).

You might also like