ArticlePDF AvailableLiterature Review

Non-Destructive Techniques for the Condition and Structural Health Monitoring of Wind Turbines: A Literature Review of the Last 20 Years

MDPI
Sensors
Authors:

Abstract and Figures

A complete surveillance strategy for wind turbines requires both the condition monitoring (CM) of their mechanical components and the structural health monitoring (SHM) of their load-bearing structural elements (foundations, tower, and blades). Therefore, it spans both the civil and mechanical engineering fields. Several traditional and advanced non-destructive techniques (NDTs) have been proposed for both areas of application throughout the last years. These include visual inspection (VI), acoustic emissions (AEs), ultrasonic testing (UT), infrared thermography (IRT), radiographic testing (RT), electromagnetic testing (ET), oil monitoring, and many other methods. These NDTs can be performed by human personnel, robots, or unmanned aerial vehicles (UAVs); they can also be applied both for isolated wind turbines or systematically for whole onshore or offshore wind farms. These non-destructive approaches have been extensively reviewed here; more than 300 scientific articles, technical reports, and other documents are included in this review, encompassing all the main aspects of these survey strategies. Particular attention was dedicated to the latest developments in the last two decades (2000–2021). Highly influential research works, which received major attention from the scientific community, are highlighted and commented upon. Furthermore, for each strategy, a selection of relevant applications is reported by way of example, including newer and less developed strategies as well.
Content may be subject to copyright.


Citation: Civera, M.; Surace, C.
Non-Destructive Techniques for the
Condition and Structural Health
Monitoring of Wind Turbines: A
Literature Review of the Last 20
Years. Sensors 2022,22, 1627.
https://doi.org/10.3390/s22041627
Academic Editors: Adam Glowacz,
Jose A Antonino-Daviu and
Wahyu Caesarendra
Received: 29 December 2021
Accepted: 15 February 2022
Published: 18 February 2022
Publisher’s Note: MDPI stays neutral
with regard to jurisdictional claims in
published maps and institutional affil-
iations.
Copyright: © 2022 by the authors.
Licensee MDPI, Basel, Switzerland.
This article is an open access article
distributed under the terms and
conditions of the Creative Commons
Attribution (CC BY) license (https://
creativecommons.org/licenses/by/
4.0/).
sensors
Review
Non-Destructive Techniques for the Condition and Structural
Health Monitoring of Wind Turbines: A Literature Review of
the Last 20 Years
Marco Civera * and Cecilia Surace
Department of Structural, Geotechnical and Building Engineering (DISEG), Politecnico di Torino, Corso Duca
degli Abruzzi 24, 10129 Turin, Italy; cecilia.surace@polito.it
*Correspondence: marco.civera@polito.it
Abstract:
A complete surveillance strategy for wind turbines requires both the condition monitoring
(CM) of their mechanical components and the structural health monitoring (SHM) of their load-
bearing structural elements (foundations, tower, and blades). Therefore, it spans both the civil
and mechanical engineering fields. Several traditional and advanced non-destructive techniques
(NDTs) have been proposed for both areas of application throughout the last years. These include
visual inspection (VI), acoustic emissions (AEs), ultrasonic testing (UT), infrared thermography (IRT),
radiographic testing (RT), electromagnetic testing (ET), oil monitoring, and many other methods.
These NDTs can be performed by human personnel, robots, or unmanned aerial vehicles (UAVs);
they can also be applied both for isolated wind turbines or systematically for whole onshore or
offshore wind farms. These non-destructive approaches have been extensively reviewed here; more
than 300 scientific articles, technical reports, and other documents are included in this review,
encompassing all the main aspects of these survey strategies. Particular attention was dedicated
to the latest developments in the last two decades (2000–2021). Highly influential research works,
which received major attention from the scientific community, are highlighted and commented upon.
Furthermore, for each strategy, a selection of relevant applications is reported by way of example,
including newer and less developed strategies as well.
Keywords:
structural health monitoring; condition monitoring; damage detection; fault diagnostics;
non-destructive testing; artificial intelligence; wind turbine; wind farm; blade monitoring
1. Introduction
There is a general consensus from technicians, political leaders, and public opinion
alike that the worldwide energy sector should shift to more sustainable sources.
Wind power is widely considered one of the best options in this sense. As for any
energy source, it has its own advantages and limitations; for instance, it is an intermittent
source, not dispatchable on demand but rather subject to the fluctuating nature of meteoro-
logical conditions. Nevertheless, it is fully renewable and highly sustainable, with minimal
environmental impact when compared to traditional fuel power. However, wind turbines
(WTs) come with both worker and public safety concerns.
In case of mechanical faults, turbine nacelle fires may erupt. Due to their height,
these can be dangerous to extinguish, while releasing toxic flumes and potentially causing
secondary fires in their immediate surroundings.
The risks are even more evident for structural collapses. These can be due to global or
local failure mechanisms. The first case can be caused by a failure at any point along the
tower height or its complete toppling due to foundation issues. In this instance, not only
the wind turbine but also nearby structures can be damaged in the collision.
Even in the case of local failures, the consequences may be particularly severe, espe-
cially for turbines located near highly-populated areas. These failures include the detach-
Sensors 2022,22, 1627. https://doi.org/10.3390/s22041627 https://www.mdpi.com/journal/sensors
Sensors 2022,22, 1627 2 of 52
ment of the rotor from the nacelle or of the whole rotor–nacelle ensemble from the tower.
However, the most well-known structural risk concerns blade brakes.
Indeed, according to Ref. [
1
], blade failure is the most common risk, accounting for
65% of all incidents (when the detachment of both the full blade or part thereof are
accounted for together). Even the loss of some smaller blade components may endanger
people’s safety, due to their potential high impact velocity and the long distances they
can cover when carried by strong winds. For this reason, WT blade monitoring will be
particularly detailed in this discussion.
Thus, wind turbines require constant integrity and safety monitoring. This can be
achieved with automated approaches, based on Artificial Intelligence (AI). However, these
AI-based diagnostics strategies require damage-sensitive features for data-driven anomaly
detection. These should ideally be retrieved from the operating WT in a non-destructive,
non-invasive fashion.
To this aim, this paper reports a broad overview of non-destructive evaluation (NDE)
approaches and the respective techniques (NDTs) for structural health monitoring (SHM)
and condition monitoring (CM). Due to the very large quantity of published documents on
this subject, vibration- and SCADA-based approaches will not be included here; instead,
they will be covered in a dedicated review in the near future.
The remainder of this paper is structured as follows. Section 2describes the historical
and current context of the WT and wind industry. Section 3recalls the main components
of any generic wind turbine and discusses their implicit risks and the main causes of
damage/failure. Section 4briefly describes the main sub-fields of applications for both
CM and SHM in WTs. Section 5lists the main NDE strategies for the structural and
mechanical components. Section 6reports some final discussions and suggestions. Finally,
the conclusions end this paper.
2. Context: The Worldwide Politics and Economics of Wind Turbines
2.1. Climate Change and the Political Stance on Sustainable Energy Sources
The need for sustainable and renewable energy sources (also known as green energy)
originated mainly in the last decades. This was due to the raising concerns about the mid-
and long-term effects of human activities on the environment, on a global scale.
For this reason, in 1980 the World Meteorological Organization (WMO) organized the
first world conference on climate “to prevent potential man-made changes in climate that
might be adverse to the well-being of humanity” [2].
In the following years and up to the present day, numerous UN climate conferences
and meetings have been held. Among these, the 1997 UNFCCC Climate Change Confer-
ences resulted in the famous Kyoto Protocol, where binding milestones were set for the
reduction in harmful emissions by industrialized countries. In 2015, these meetings led
to the Paris Agreement, where 196 countries agreed to the goal of limiting the increase in
the global temperature to less than two Celsius degrees above pre-industrial levels. The
most recent UN conference on climate, named COP26, was ongoing at the time of writing.
Therefore, there is pressure from both public opinion and decision-makers to subsidize
green energy producers to extend their market share.
In this regard, the International Renewable Energy Agency (IRENA) recently published
in its latest annual report [
3
] the statistics on the global renewable energy generation capac-
ity. According to the data collected, the global renewable generation capacity amounted
to 2799 GW at the end of 2020. In the same year, wind energy took second place with a
capacity of 733 GW (26%), preceded only by hydropower sources (1332 GW or 48%) and
ahead of solar energy (714 GW or 25%, including both photovoltaic and concentrated solar
power). In more detail, onshore WTs and wind farms produced 699 GW, almost on par with
the total solar energy production by itself, while offshore sites added a further 34 GW.
Furthermore, solar and wind energy continued to dominate the expansion of renew-
able capacity. Out of a total increase in the total renewable generation capacity by 261 GW
(+10.3%) in 2020, solar energy continued to drive capacity expansion, with an increase of
Sensors 2022,22, 1627 3 of 52
127 GW (+22%), yet very closely followed by wind energy with 111 GW (+18%). Indeed,
in most industrialized countries, due to the limited possibility of new large hydroelectric
installations, the increased demand for renewable energy is largely covered by these two
sources, with wind energy production exceeding its solar counterpart in many European
countries, e.g., Italy, according to the most recent data [
4
]. From a historical perspective
in the European Union, the wind power generation capacity has been well above solar
photovoltaic capacity since the early 2000s, overtaking the fuel oil capacity in 2007, nuclear
energy in 2013, hydroelectric in 2015, and coal in 2016, remaining only behind natural gas
as of 2017 [5].
2.2. The Current and Near-Future Economic Impact of Wind Power
At the world scale, according to the latest data from BloombergNEF, wind power
developers around the world commissioned a record 96.7 GW of installations in 2020, up
59% from 60.7 GW installed in 2019. This increase in capacity was mainly due to the surge
in installations in China and the United States [
6
]. Specifically, 2020 set a record year of
wind growth for China, with a 36% year-on-year increase in WT installations [
7
]. In the
same year, the USA installed a record 14.2 GW, more than in any other year so far [
8
].
Other non-European high-GDP countries strongly committed to the transition towards
wind energy include Japan (where the Japan Wind Power Association declared 2020 as the
best year for capacity addition in the country’s history [9]), Australia (where wind energy
supplied 35.9% of the clean energy in 2020, remaining the leading renewable source and
setting a record-breaking year [
10
]), India (with a year-on-year increase of +5.9%, reaching
a total of 37.7 GW installed nationwide [
11
]), Brazil (where the installed capacity increased
from 1 to 18 GW from 2010 to 2020 [
12
]), and South Korea (where the government Green
New Deal, announced in July 2020, set a goal of 12 GW of wind capacity by 2030, also
announcing the largest offshore wind farm in the world, to be built in the South Jeolla
province [13]).
At the European scale, the 5-year outlook reported in Ref. [
14
] considered a growth
of 15 GW p.a. as a “realistic expectation”. It should be stated, furthermore, that new
installations (as of 2020) of offshore WTs strongly exceeded onshore ones in many North
Sea countries (1493 MW vs. 486 MW in the Netherlands, 706 MW vs. 152 MW in Belgium,
and 483 MW vs. 115 MW in the UK [
14
]); the aggregate onshore and offshore installations
are portrayed in Figure 1. This makes economic sense due to the more stable and steady
wind flows in the open sea and the lower acoustic and visual impact in comparison to their
onshore counterparts. However, with the operation and maintenance (O&M) costs much
larger for offshore production facilities, structural health and condition monitoring will be
even more important.
Figure 1.
Percentage of new WT installations in 2020 (both on- and offshore), in terms of produced
MW capacity. Based on data retrieved from [14].
Sensors 2022,22, 1627 4 of 52
In Mediterranean and southern European countries, this can act as a driving economic
force for less-developed areas. For instance, Italy was the fifth country in Europe in terms
of installed wind capacity, with more than 10 thousand MW of plants installed as of 2019
(mostly onshore), for a total of about 670 WTs [
15
]. Most of the wind farms (over 90%)
are concentrated in the South and the islands of Sicily and Sardinia, due to the greater
availability in these regions of adequately windy sites. At the moment, not even 1 MW
of offshore installations are operative. However, the installation of the first offshore wind
farm in Italy (in the Sicilian channel) is expected in the coming years.
The prospects in the medium to long term are very positive, especially due to the
unique opportunity given by the Recovery Plan for Europe. According to the National
Integrated Plan for Energy and Climate (Piano Nazionale Integrato per l’Energia e il Clima,
PNIEC [
16
])], the installed wind energy capacity in Italy should reach approximately
19,300 MW by 2030, of which approximately 900 MW will come from offshore wind. This
would guarantee an annual production of electricity equal to 40 TWh, which is 10% of the
national gross electricity consumption.
2.3. Expected Returns and Benefits from WT Monitoring
For all the reasons discussed above, the impact of CM and SHM cannot be neglected
from the perspective of more reliable, more cost-effective wind power production in the
coming decades. From a risk management point of view, the NDE strategies and their
related NDTs are required to reduce the number of both minor incidents and (fatal or non-
fatal) severe accidents. From the economic perspective, the same approaches are intended to
improve, rationalize, and automate as much as possible the maintenance routine. In general
terms, the end goal is a paradigm shift from time-scheduled maintenance to condition-based
maintenance, which is more flexible and avoids unnecessary deployments on-site. This
can save man-hours and transportation costs. Fewer human operators on-site would also
mean less exposure to dangerous conditions (especially for offshore wind farms). All these
beneficial consequences would lower the O&M costs and therefore the total energy price,
thus making the renewable energy from wind farms more cost-competitive in comparison
to other traditional alternatives, such as fossil fuels. In this sense, the cost-efficiency of WTs
is commonly evaluated in terms of (yearly or life-long) levelized cost of energy (LCOE),
i.e., the total cost (per year or over lifetime) divided by the (annual or lifetime) energy
production [
17
,
18
]. Decreasing O&M costs automatically increases the LCOE of new and
already-existing installations [19].
In particular, for mechanical faults, the mean downtime can vary in a range from
6 to 15 days onshore. The cost of a gearbox bearing exchange ranges from EUR 15000
for a simple up-tower replacement to more than EUR 1 million for the substitution of
larger (5 MW) gearboxes [
20
]. These figures are indicative of onshore installations; as
mentioned earlier, failures in offshore WTs would result in even longer interruptions and
costlier reparations [
21
]. McMillian & Ault [
22
] provided a detailed quantitative analysis of
the impact of condition monitoring on O&M costs. In this regard, Figure 2reports some
indicative estimates.
Regarding (partial or total) structural collapse, while likely much rarer, this occurrence
would result in the complete loss of the whole asset, plus the collateral damage; therefore,
an SHM apparatus should be always considered jointly to the more common CM systems.
For all these applications, the expected gains of implementing an SHM/CM strategy
can be evaluated in terms of its value of information (VoI [
23
]). Some examples can be found
in Ref. [
24
] for the SHM of WT blades and in Refs. [
25
,
26
] for the CM of WT gearboxes and
generators (in the same order).
Sensors 2022,22, 1627 5 of 52
Figure 2.
Bar chart of the estimated investment (
left
) and O&M (
right
) costs per MW of an onshore
HAWT, according to the class of mean electric power produced. Based on data retrieved from Ref. [
27
].
3. Wind Turbines: Structural and Mechanical Components
There are several different typologies of “wind turbines”, the most common type
being the so-called Horizontal Axis Wind Turbine (HAWT) systems. In this configuration,
the rotation axis of the rotor is parallel to the ground. Specific attention must be paid to
the orientation with respect to the wind direction, which is different from other types of
wind turbines such as those with a vertical axis, whose orientation is independent of the
prevailing wind direction. The upwind configuration is the most common one, with the
rotor facing the incoming wind.
In this review, only HAWT will be considered; the term “wind turbine” will therefore
be used exclusively for this specific device hereinafter. In more detail, the only structural
configuration of interest will be the classic towered HAWT, with one or more blades (usually
three). Both on- and offshore turbines are included. Except for small wind turbines for
private on- or off-grid energy production, any tower size and blade length are considered.
From an engineering perspective, any wind turbine is made up of:
(i)
static, load-bearing components;
(ii)
moving/rotating parts, needed to harness the wind’s kinetic energy and turn it
into electricity.
The elements in (i) are generally referred to as the support structure. The components
of (ii) can be further divided between slowly rotating elements (blades) and high-speed
rotating mechanisms. These latter ones are all included in the rotor–nacelle assembly on
top of this support structure.
These distinctions are essential since the blades and the support structure are fields
of application for structural health monitoring, while condition monitoring, according to
the common definition of the term, deals with machinery and rapidly moving components
such as gears and bearings in the gearbox and generator.
Indeed, elements in both (i) and (ii) are subject to naturally occurring use and con-
sumption and, therefore, can develop structural damage in the long run. Damage in the
external structure will cause (partial and localized or global) collapse, while damage in the
internal mechanisms will cause faults, disruption in the energy production, and potentially
fire, explosions, oil leakage, or other events.
Hereinafter, the main structural components and pieces of machinery of a towered,
multi-bladed HAWT are reported and briefly recalled. A more detailed description can be
found in the textbook of Hau & Von Renouard [28].
Sensors 2022,22, 1627 6 of 52
3.1. Components under Structural Health Monitoring
3.1.1. The Tower
The tower is the main component of the support structure. Apart from foundation
costs (which, as it will be discussed later, vary noticeably for on- and offshore structures),
its cost can be up to one-fourth of the total (Figure 3).
The main parameter of the tower is its height. This is typically about 1.5 times the
rotor diameter; generally, it is never lower than 20 m and can reach up to 150 m or more (for
10–12 MW outputs). In absolute terms, the higher the tower, the better the wind conditions
in terms of intensity and constancy. The tower can be lattice or tubular. This second design
choice has been more common since the mid-1980s. In this case, the tower is made of
thin-walled steel conical parts of varying diameters and diameter-to-wall thickness ratios.
These offer a practical and safer way for the survey teams to access the nacelle. Moreover,
in comparison to lattice structures, there are fewer bolted joints to inspect and maintain.
The tower diameter (maximum at its basis and minimum at its top) increases with the
tower height; e.g., a typical 50 m-tall HAWT will have a diameter ranging from 3.5 m to
0.4 m [29].
From a vibrational and SHM perspective, the stiffness of the tower is the main param-
eter to be taken into consideration in evaluating the global dynamics of WTs due to the
possibility of coupled vibrations between the tower and the rotor.
Figure 3.
Estimated costs of a HAWT, as a percentage of the total and excluding foundations. Based
on data retrieved from Ref. [30].
3.1.2. The Substructure
It must be mentioned that offshore HAWTs, differently from their onshore counterparts,
include a further group of structural elements, which are included below the platform
and above the sea floor (Figure 4). These components are particularly at risk due to their
location under water or—even worse—in the splash zone, immediately above/below the
mean water level and highly subject to corrosion. Furthermore, being submerged, they
cannot easily be visually inspected if not using divers or manned/unmanned underwater
inspection robots. They are also subject to marine growth and other potentially damaging
environmental conditions such as wave, tidal, and current forces.
3.1.3. The Foundations
On- and offshore foundations differ sensibly. However, in both cases, the choice of the
specific structural design depends on the location and site conditions. For example, the
quality and strength of the soil are the main determinants affecting the size and shape of
onshore foundations, while the depth of the water and the distance from the coast are the
key factors for offshore turbines.
Sensors 2022,22, 1627 7 of 52
Figure 4. Structural components of a fixed offshore HAWT according to IEC 61400-3-1.
For onshore installations, both surface (shallow) and deep foundations are frequently
used. In the wind farms dating back to the 1990s, square foundations with a constant
thickness were commonly utilized. This solution, however, can lead to the formation of
localized damage. Hexagonal and octagonal shapes subsequently became more common,
even with variable thicknesses. The most modern designs use circular shapes, which allow
the reinforcing bars to be positioned more homogeneously. This ensures a better dispersion
of the forces induced by the soil–structure (or soil–pile–structure) interactions.
The foundation design, construction, and monitoring in offshore wind farms are more
complex and challenging. The costs are much higher as well, absorbing a large percentage
of the total expense of an offshore wind turbine. Specifically, Ref. [
31
] mentioned the
offshore foundation costs to be 35% of the total project expenses. Refs. [
31
,
32
] estimated a
cost from 352 EUR/kW (
19.6% of the total construction cost) for a water depth between 10
and 20 m, up to 900 EUR/kW (
35.8%) for transitional waters (40 to 50 m deep). For smaller
HAWT closer to the coast (1–2 MW, with A water depth <30 m), monopile foundations are
often encountered. Jacket/tripod substructures are more frequent between 25 and 50 m
and for larger structures (2–5 MW) [
33
]. For deep waters (50–120 m and beyond), floating
structures, anchored to the sea floor, are preferable. Conversely, gravity-based foundations
(to the right side in Figure 4) can be found for offshore installations in very shallow waters
but they are not widely used [31].
3.1.4. The Rotor
The rotor is a crucial component of the wind turbine. Indeed, it is the most expensive
mechanical component, up to circa one-fifth of the total cost [
34
]. Its design is one of the
most critical and delicate phases, especially in terms of expected performances. Therefore,
the economic feasibility of the whole system depends on these aspects.
The rotor consists of the blades and the hub from which they branch off. Generally,
the rotor can be single-, double-, or three-bladed. The most common turbines include three
blades arranged at 120
from each other; this is conventionally considered the optimal design.
The length of the blades determines the capability to convert high wind speed to low
rotational speed and finally to electrical energy. As mentioned before, the diameter of
the rotor governs the tower height and thus the overall size of the structure, including
the foundations. Indeed, a larger rotor will produce a consequent increase in the energy
Sensors 2022,22, 1627 8 of 52
produced but, on the other hand, will require a wider tower cross-section and more massive
and/or deep foundations, thus increasing the construction costs.
For monitoring purposes, a “smart rotor” system should include several sensors
(accelerometers, strain gauges, pitot tubes, pressure tabs,etc.), embedded and distributed
along the blades [
35
]. Indeed, due to the prominence of the rotor blades, these components
deserve to be discussed on their own.
3.1.5. The Blades
WT blades are generally made of glass or carbon fibre reinforced polymeric (GFRP
and CFRP) materials. E-glass fibres are particularly used as the main reinforcement in the
composite material.
The blade design is actually quite complex, with several different components and
materials as pictorially described in Figure 5. This complexity makes them particularly
susceptible to manufacturing defects, which indeed were estimated in Ref. [
36
] to account
for
51% of all blade damages (with debonding and voids in skin core being the most
common defects at 20% and 18%, respectively).
Figure 5.
Key components of a typical wind turbine blade. The upper and lower surfaces are also
known as the suction (or windward) and pressure (or lee) sides, respectively. The blade root bolt
connection shown here is a classic T-bolt type.
From a geometric perspective, the cross-section profile varies from root to tip, often
also rotating around its main axis; all these aerodynamical aspects of their airfoil design
characterize the blade lift-to-drag ratio and, by consequence, the wind-to-rotor efficiency.
The internal reinforcements may include different sorts of load-carrying structural elements,
such as shear webs, closed shells, box spars, or other geometries. This design choice affects
the space available for the maintenance workers to operate.
As briefly mentioned earlier, the structural integrity of WT blades is of the foremost
importance, due to the potential impacts of fully or partially detached blades with neigh-
bouring structures.
3.2. Components under Condition Monitoring
3.2.1. Drive Train (and Other Components Inside the Nacelle)
Positioned on the top of the tower, the nacelle is a cover housing, intended to shelter all
the mechanical and electrical components installed inside from the external environment.
These mechanisms include (for a conventional HAWT) the gearbox, rotor shaft, brake,
and generator, all assembled together (Figure 6, adapted from Ref. [
37
]). These pieces
Sensors 2022,22, 1627 9 of 52
are necessary for energy conversion and represent roughly between one-third and half of
the HAWT total cost on their own [
34
]. For these reasons, they will be discussed in more
detail separately.
Figure 6. Mechanical and electrical components inside the nacelle of a conventional HAWT.
Other mechanisms included in the tower and the nacelle are, for instance, the rotor
yaw system and the control and power electronics systems.
The yaw system is responsible for the orientation of the nacelle–rotor assembly towards
the wind, rotating 360
around the vertical axis. The control and power electronics systems
are tools of fundamental importance to control the operation of the machine, manage the
supply of electricity, and stop the system beyond certain wind speeds for safety reasons due
to the excessive heat (generated by the friction of the rotor on the axis) and/or mechanical
stresses. Therefore, they maximize the life of the system by ensuring the limitation of
the fatigue of all components (fatigue that can result due to changes in wind speed and
direction, the presence of turbulence, shutdown, and start-up cycles of the turbine, etc.).
These systems include speed, position, temperature, and voltage sensors; mechanical
or electrical controllers; actuators; plus valves; switches; microprocessors; and many
other components.
3.2.2. The Gearbox
The main rotating machinery in a typical HAWT is the gearbox, which is also the
element most prone to mechanical faults. Therefore, it is the component of major interest
for Condition Monitoring.
Gearboxes in HAWTs are generally multi-stage, with one or more sequential planetary
stages, followed by one or more parallel stages (i.e., helical gears). The rationale is that
the speed of the rotor axis (in the order of magnitude of tens of revolutions per minute,
rpm, depending on the wind) is not sufficient for the generator to produce electricity
cost-efficiently.
In the typical drive train configuration, the gearbox acts as a rotation speed multiplier,
connecting a low-rpm, high-torque shaft on the rotor side (known as the input, slow, or
main shaft) to a high-speed (output) shaft on the generator side [
28
]. The support bearings,
mechanical brake, and rotating parts of the generator make up the rest of this common
configuration. This can increase the revolutions per minute from 12–30 up to 1200–1800.
The efficiency of the gearbox is also linked to the lubricant conditions, as will be discussed in
the following sections. These refer to both the oil cleanliness in terms of particle content and
the oil viscosity (which influences the thickness of the oil film in the gears and bearings).
Sensors 2022,22, 1627 10 of 52
3.2.3. The Generator
Located behind the gearbox and driven by the high-speed shaft, the electric generator
accommodates the mechanical and electrical components needed to convert the incom-
ing rotation into electricity. In comparison to other generators, the ones in use for wind
power need to adapt to a fluctuating mechanical power (torque) source. Furthermore, for
wind turbines directly or indirectly connected to local or national grids, synchronous or
asynchronous alternators are required to produce electrical energy at the grid frequency
(generally 50 Hz or 60 Hz, depending on the national standards). Traditionally, squirrel-
cage induction machines and synchronous machines have been used for small scale WTs;
doubly-fed induction generators are considered the dominant technology for larger mod-
els [
38
]. Other technologies include the permanent-magnet, switched reluctance, and
high-temperature superconducting generators. A comparative analysis of the benefits and
limitations of some options can be found in Ref. [39].
3.3. Incidence and Main Causes of Structural Collapse
In general terms, structural components are much less likely statistically to be subject
to damage and failure than mechanical and electrical components. This is shown in detail
in the statistics in Figure 7. For the tower, typical failure mechanisms are caused by bolt
loosening induced by dynamic loads [
40
]. However, as can be seen from Figure 7b, WT
blades are the structural elements most affected by damage. This is due to their higher
complexity, particular materials, and exposition to strong dynamic loads. The static load-
carrying elements (tower structure and substructure, plus foundations) are considered to
be less damage-prone. Among the many potential causes of their catastrophic collapse, the
structure of the metallic tower is subject to all the classic risks of thin-walled shells, e.g.,
buckling. Even concrete-made towers may suffer total collapse due to damage at the base
(at least one case was reported in Germany in 2000 [41]).
Figure 7.
Damage and failure statistics. (
a
) Percentage of unforeseen malfunctions as recorded in
Germany for 1500 wind turbines. Based on data retrieved from Ref. [
42
], collected over 15 years
(34,582 events). (
b
,
c
) percentage distribution of the total number of failures and downtime for WTs.
Based on data retrieved from Ref. [
43
], collected from several sources in Sweden, totalling about
600 WTs from 2000 to 2004 (1202 events, 156,202 h). The nomenclature used in the original sources is
reproduced for all charts.
Sensors 2022,22, 1627 11 of 52
During their operational lifespan (commonly 20–25 years), the blades are subject
to rapidly changing dynamic loads, which can cause fatigue damage [
44
,
45
], plus other
environmental conditions such as rain, humidity, and abrasive wind-carried dust or other
kinds of particulate matter. Even the simple uninterrupted exposition to sun-radiated
ultraviolet rays causes durability issues in the long term. All these eventualities might
cause surface damage and/or corrosion.
Moreover, the blades can be hit by larger foreign objects, such as hailstone and bird
strikes, and (commonly) by lightning. These two latter factors account (in the same order)
for 16% and 20% of the total damage causes according to Ref. [
36
], making them the most
common causes after manufacturing defects. It is estimated that one-third of all damages
on blades happens during storm events [46].
The blade tip and trailing edge (refer to Figure 5) are generally designed and shaped
to minimize aerodynamic noise, which makes them thin and particularly prone to the
occurrence of mechanical damage. A full-scale static test showed that a 40 m-long E-
glass/epoxy composite WT blade can bear tip deflections up to 11 m under flap-wise
loading before structural collapse [
47
]. This large flexibility, however, may cause the blade
tip to hit the HAWT tower under exceptional wind conditions. Due to its cross-section, the
blade is much stiffer edge-wise. In conclusion, seven categories of damage are considered
as the most common ones in composite blades; these are reported in Table 1.
Table 1. Typical damage typologies in WT blades, according to Refs. [48,49].
Damage Type Description
#1 Damage formation and growth in the adhesive layer joining the skin
and main spar flanges (skin/adhesive debonding and/or the main
spar/adhesive layer debonding).
#2 Damage formation and growth in the adhesive layer joining the up-
and downwind skins along leading and/or trailing edges (adhesive
joint failure between skins).
#3 Damage formation and growth at the interface between the face and
core in the sandwich panels in skins and the main spar web (sandwich
panel face/core debonding).
#4
Internal damage formation and growth in laminates in the skin and/or
main spar flanges, under a tensile or compression load (delamination
driven by a tensional or a buckling load).
#5
Splitting and fracture of separate fibres in the laminates of the skin and
main spar (fibre failure in tension; laminate failure in compression).
#6 Buckling of the skin due to damage formation and growth in the bond
between the skin and main spar under a compressive load. *
#7 Formation and growth of cracks in the gel coat; debonding of the
gel-coat from the skin (gelcoat cracking and gel-coat/skin debonding).
* Type #6 can be seen as a particular case of Type #1 damage [50].
3.4. The Incidence and Main Causes of Mechanical Failure
Recalling Figure 7b,c, one can see that the mechanical failures (including brakes, gears,
drive train, and generator) are preponderant in terms of both total failures and downtime.
For this reason, the majority of scientific research focuses on the CM of these components
housed inside the nacelle.
In the context of mechanical components, the concept of the “failure rate” is generally
applied. The standard time-scheduled maintenance, therefore, should follow the classic
bathtub curve, focusing on early-stage monitoring (to assess “infant mortality”, due to
defective components) and in the long run, where wear out failures start to become predom-
inant. During its life cycle, any wind turbine is however subject to the random occurrence
of unexpected faults; this represents a statistical constant along time. In general, gearbox
Sensors 2022,22, 1627 12 of 52
failures happen slightly more frequently due to late wear out failures compared to early
infant mortality [
21
]. A study over a 13-year timeframe showed a failure rate from 0.10 to
0.15 mechanical failure/turbine/year for land-based European WTs [51].
Bearings are, arguably, the most critical component of the gearbox. A 2013 report
from the U.S. National Renewable Energy Laboratory (NREL) [
52
] estimated that about
70% of all gearbox failures are due to bearing failures. More specifically, the bearing life of
the parallel stages is generally considered to be very high, in many cases over 1 million
hours [
20
], even if bearing failures at this stage are nevertheless not uncommon (see, e.g.,
Ref. [
53
]). On the other hand, bearings in the planetary stages often do not fulfil their
design lifespan [
20
]. This has also been assessed in many experimental campaigns [
53
] and
it is quite understandable due to the large torque applied at the planetary gear stages. The
reliability and durability of these bearings have always been considered one of the major
issues for WTs [
54
]. Many failures are observed in the bearing inner rings of the planetary
stages since these are exposed to 5 times more high load cycles than the outer rings [20].
Table 2reports a brief description of the most common bearing failure typologies, according
to the nomenclature reported in Ref. [55].
Table 2. Typical damage typologies in bearings, according to Ref. [55].
Damage Type Description Possible causes
Flaking Creation of regions with a rough and
coarse texture due to the splitting off of
small pieces from the raceway surface.
Rolling fatigue, caused in turn by excessive load,
misalignment, poor lubrification, water or debris
inclusions, unsuitable bearing clearance, unevenness in
housing rigidity, rust, corrosion pits, dents.
Peeling Light wear and dull spots on the surface,
with micrometric cracks and
minor flaking.
Poor or unsuitable lubricant, debris intrusion in
the lubricant.
Scoring Straight lines on the surface,
circumferentially on the raceway surface.
Generated by accumulated small seizures, caused in
turn by sliding under improper lubrication or
excessive/improper loads and conditions (shaft
bending, the inclination of inner and outer rings, etc).
Smearing Surface damage, with the formation of
rough and partially melted material.
Generated by accumulated small seizures between
bearing components, caused in turn by oil film rupture
(because of poor/improper lubrication or high speeds
with very light loads).
Fracture
Small pieces broke off due to shock loads
or stress accumulation. Impacts during mounting/dismounting, excessive
loads, progression of surface cracks.
Cracks Formation of surface cracks on the
raceway rings and/or rolling elements.
Excessive loads, progression of flaking damage,
creep-induced heating, inappropriate shaft (e.g., poor
taper angle).
Cage damage Cage deformation, fracture, and/or wear
(considering the cage guide surface,
pocket surface, and cage pillars).
Excessive speed, sudden acceleration/deceleration, high
temperature, poor lubrication, excessive vibrations,
bearing misalignment.
Denting Small dents on the surface of raceway
rings or rolling elements. Caused by metallic particles or other very small debris
caught in the surface during rolling.
Pitting Pitted surface on the raceway rings or
rolling elements. Poor lubricant, debris in the lubricant, or exposure
to moisture.
Wear Surface deterioration on the raceway
rings, rolling elements, cage pockets,
and/or roller end faces.
Sliding friction between two surfaces, caused in turn by
an irregular motion of the rolling elements, poor
lubrication, debris intrusions in the lubricant, or as a
progression from chemical or electrical corrosion.
Sensors 2022,22, 1627 13 of 52
Table 2. Cont.
Damage Type Description Possible causes
Fretting
Corrosion happening at the contact area
between the raceway ring and the rolling
elements. It may happen at regular roller
pitch intervals.
Repeated sliding on the fitting surface.
False brinelling
Hollow spots that resemble Brinell dents.
Caused by wear, induced in turn by vibration and
swaying at the contact points between the raceway and
the rolling elements, especially with poor lubrication.
Creep Shiny appearance on the fitting surface,
potentially coupled with scoring
and wear.
Relative slipping at the fitting surfaces, due to a loose fit
or insufficient sleeve tightening.
Seizure Softened, deformed, and/or melt
material in the raceway rings, rolling
elements, or cage.
Excessive load, speed, shaft bending, poor housing or
lubrication, small internal clearance.
Electrical corrosion Corrugations resulting from locally
melted material.
Melting by arcing, induced by the passage of electric
currents. In turn, these are induced by the electrical
potential between the inner and the outer rings.
Pit corrosion Pits on the surface of raceway rings or
rolling elements due to
chemical corrosion.
Entry of corrosive gas or liquid, improper lubricant,
moisture, high humidity, improper handling and
storage conditions.
Mounting flaws
Scratches on the surface of raceway rings
or rolling elements caused by
mounting/dismounting.
Incorrect mounting/dismounting (impulse loads, the
inclination of inner or outer rings, etc).
Discolouration Discolouration of the cage, rolling
elements, or raceway rings. Poor lubrication and/or high temperature.
On the other hand, the gearbox gears can be generally assumed to fulfil their design
life as well, even if their reliability is often lower than expected and depends on the
manufacturer’s design, tooth profile, and material quality [
20
]. Surface fatigue cracks and
tooth bending fatigue are among the most common causes [20].
Outside of the gearbox, the main shaft bearings are considered to be much less critical
and are generally capable of fulfilling their design life [
56
]. The generator presents some
rotating components that require CM as well. This can be performed with vibration-
based inspection (VBI) and signal processing approaches (two examples are reported in
Refs. [57,58]) or using alternative methods such as temperature trend analysis [59].
In conclusion, Table 3reports the main causes of mechanical failure according to the
scientific literature. However, one should remember that, apart from purely mechanical
failures, generators are (obviously) at risk of the failure of electrical materials [60].
Table 3. Typical failure modes in wind turbine mechanical components, according to Ref. [61].
Mechanical Component Common Failure Modes
Gearbox and drive train
Gear tooth damages, high- or low-speed shafts faults, gearbox bearing failures.
Generator Generator stator failure, generator rotor failure, generator bearing failure.
Main bearing Bearing failure, bearing rubs, bearing looseness
Pitch gears Pitch Gear tooth damages.
Yaw gears Yaw Gear tooth damages.
3.5. Survey and Maintenance Policy for Offshore Wind Farms
Due to their greater logistic complexity and higher O&M costs, a brief discussion
about the maintenance strategies for on- and offshore wind farms is needed.
Sensors 2022,22, 1627 14 of 52
“Corrective” or “reactive” maintenance is, by definition, performed after failure detec-
tion, aiming at the restoration of an asset to a condition in which it can perform its intended
function. It is both unscheduled and unplanned. Thus, corrective maintenance is often
unavoidable, with the maintenance teams having to respond to equipment breakdown
or failure immediately with little to no pre-alarm and often without a totally clear under-
standing of the exact cause of damage/failure. “Preventive” maintenance, on the contrary,
aims to carry out an overhaul, replacement, or repair before the component fails, in a
pre-planned, predictable fashion. This can help to reduce downtimes while also improving
efficient resource planning.
Preventive maintenance can be further divided into pre-determined (time-scheduled)
and predictive (condition-based) maintenance. This scheme is the one currently set by the
current European standard, as dictated by the norm EN 13306:2017 (depicted in Figure 8).
Figure 8. Maintenance strategies according to EN 13306:2017.
The current trend in the wind industry is to transition from time-scheduled to condition-
based monitoring, using embedded sensors and continuous, permanent monitoring sys-
tems to detect early signs of anomalies at a global level. This is mostly achieved through
vibration-based signal processing, SCADA data analysis, and statistical pattern recognition,
applying artificial intelligence strategies (mainly machine learning and artificial neural
networks). As mentioned, these aspects will be all addressed in future, dedicated works.
However, once an anomaly is detected (or even localized) at a global level, on-site main-
tenance is still required for in-depth, localized analysis. Therefore, local NDE and global
vibrational approaches are not mutually exclusive. On the contrary, they both allow fast
and cost-efficient maintenance planning. This concept is a key component of the “Intelligent
Maintenance” framework, based on a globally optimum trade-off between prevention and
repair costs [37].
As already mentioned before, the issue of proper maintenance planning is even more
prominent for offshore installations. Table 4reports the main factors that affect the selection
of the maintenance and survey strategy for offshore wind farms, according to selected
authors from the recent scientific literature.
Table 4. Main factors influencing the choice of the maintenance strategy for offshore wind farms.
Study Year Mentioned Factors
Henderson et al. [62] 2003 Accessibility of the offshore platform and reliability of the
monitoring strategy.
Nielsen et al. [63] 2011 Weather conditions, total power generation, repair strategies,
transportation strategies.
Dinwoodie et al. [64] 2012 Repair time, wave height, wind speed, number of wind turbines in the
wind farm, ship availability, availability of spare parts stocks.
Scheu et al. [65] 2012 Expected typologies of component failures, ship fleet size, ship type,
travel time, number of maintenance workers on staff.
Sensors 2022,22, 1627 15 of 52
Table 4. Cont.
Study Year Mentioned Factors
Besnard et al. [66] 2013
Location of accommodation facilities for maintenance staff, vessels for
the transfer of crew (type and number), availability of helicopters,
organization of work shifts, management of spare parts stocks, technical
support, availability of cranes (purchase or contract), environmental
conditions (depending on weather and season), economic parameters
(electricity prices, ship rental costs).
Halvorsen-Weare et al. [67] 2013
Investment costs, ship costs (fixed and variable costs), failure probability,
downtime costs, meteorological data.
Hofmann & Sperstad [68] 2013 Weather conditions (including uncertainty), breakdown rates, electricity
price, ship price (costs, fleet composition, type, quantity), workers (shift
length, quantity), location of the maintenance base of operations.
Perveen et al. [69] 2014 Protection methodologies, occurrence of cable and component failures,
repair strategy, wind speed predictions, and condition
monitoring systems.
Endrerud et al. [70] 2015 Weather conditions, ships (availability, operating limits, costs),
availability of maintenance technicians, repair times, wind farm layout,
cost of spare parts, logistics (warehousing and other costs).
Nguyen & Chou [71] 2018 Duration of maintenance (downtime), expected loss of production
during maintenance time, the market price of electricity, location of the
wind farm.
4. Main Applications for NDE Techniques in Wind Turbines
Non-destructive evaluation strategies and techniques can be applied for SHM, CM, or
both. They are used to evaluate the (global or local) structural integrity of the load-bearing
and mechanical components, as well as to detect growing cracks, manufacturing defects,
or the deleterious consequences of external actions or prolonged operating conditions.
However, some specificities depend on the particular system under investigation. For
instance, the SHM of foundations (especially deep ones) has its own specific needs and
more compelling limitations. WT blades, for economic reasons, are generally made of
fibre-reinforced plastic. However, these composite materials are quite prone to develop
manufacturing defects and are more difficult to assess in detail than other more conven-
tional building materials (e.g., structural steel). Finally, the drive train and other mechanical
components inside the nacelle are subject to a high torque and/or speed rotation and/or
temperatures, differently from the wind turbine external structure.
For these reasons, it is possible to define some well-defined fields of application,
described hereinafter.
4.1. Condition Monitoring of the Mechanical Components
Arguably, this is the most relevant aspect due to the statistical prominence of mechani-
cal faults over structural collapses in WTs, as discussed before. Indeed, it can be considered
as the main focus for the whole wind industry [
72
]. Vibration-based approaches are one
of the main strategies in this ambit, even codified by specific requirements (see e.g., ISO
13373-1:2002 and ISO 61400-25-6:2016). SCADA data analysis is also highly estimated for
this aim. However, as mentioned earlier, due to the vastness of this specific argument, these
approaches are not included here and will be deferred to a dedicated future work. Oil
monitoring is another very common approach, while other NDTs, often performed locally,
include acoustic emissions, infrared thermography, and electromagnetic testing. These will
be all discussed in detail in the next sections.
Sensors 2022,22, 1627 16 of 52
4.2. SHM of the Wind Turbine Blades (Blade Monitoring)
Probably the second most important aspect, again due to the relative frequency of
(partial or total) blade failures. To account for all the potential sources of damage enlisted
in the previous section, acoustic emissions and strain measurements are commonly applied,
both on- or off-site, at different life stages of the blade (quality checks after manufacturing,
routine maintenance, extraordinary repairs, and forensic studies on collapsed specimens).
Many studies involve the dynamic and/or static characterisation of the WT blades, off-
site or in situ. Of course, the first case generally produces more reliable results, yet it requires
detaching the blades from the rotor and carrying them in a laboratory, with all the related
costs and delays. On-site inspection, on the other hand, can be performed (depending on
the technology) on stationary blades, temporarily halting the energy production, or during
operations, without interruptions.
The National Wind Technology Center (NWTC) of the NREL performed a remarkable
full-scale fatigue testing of two 9 m-long CX-100 WT blades (one pristine benchmark and
one with purposely inserted defects) [
73
,
74
]. Strain gauges, piezoelectric transducers, and
accelerometers were used to measure the vibrational response of the two systems. These
were analyzed to identify the onset of fatigue injury and to study its progression with a
variety of different signal processing techniques.
Generally, many different sensing techniques are jointly applied for a complete as-
sessment; e.g., the European project ReNEWiT developed a lightweight, automated multi-
sensor gantry system for the monitoring of stationary WTs (up to 40 m-long) GFRP blades.
It included an X-ray Compton backscattering system, dual laser shearography, pulsed
thermography, and phased array ultrasonic testing. These (and other) approaches will be
presented separately in the rest of this discussion.
Regarding the sensor placement, they can be located close to the blade “hot spots”,
i.e., the locations where damage is most likely to occur. These were enlisted in Ref. [
50
] as:
1. at the blade root (where the mechanical stress is maximized);
2. between 30% and 35% of the chord length;
3. at 70% of the same;
4. at the maximum chord section (subject to potential buckling);
5.
on the upper flange of the spar, at different chord lengths depending on the pitch
angle (thus on the current wind speed).
4.3. SHM of the Supporting Structure and Substructure
While overall a rare occurrence, the structural failure of the tower itself has a high level
of risk due to the severity of the potential consequences (total asset loss). In this sense, both
global and local monitoring are required. The first case deals with the overall structural
integrity and stability; this is performed, generally, with vibration or strain measurements,
even if Global Position Systems (GPSs) sensors and inclinometers are commonly used as
well to monitor potential risks of subsidence or capsizing, especially offshore. Periodic
controls to avoid tower misalignment are required by many regulators as well.
Local monitoring focuses on specific high-risk components. These damage-prone hot
spots include:
1.
welted, grouted, and bolted joints, due to their relative fragility, in particular to
fatigue damage; e.g., on tripod offshore structures, the upper central joint is the most
critical location;
2.
location exposed to an aggressive environment (e.g., underwater or, even worse,
in the splash zone). Corrosion monitoring is particularly requested in these most
endangered locations.
Thus, while much less developed than drive train CM, wind turbine SHM is gradually
gaining industrial applications as well. For instance, the German BSH standards (in force
since 2007 [
75
]) mandate that at least 1 out of every 10 offshore WTs is equipped with an
SHM apparatus dedicated to the support structure.
Sensors 2022,22, 1627 17 of 52
4.4. SHM of the Foundations
The concepts expressed for the load-bearing structure can be extended to the founda-
tions, with the additional risk induced by a higher degree of uncertainty. This derives from
both the complexity of soil–structure interactions (especially underwater on the sea bed)
and the difficulty to access for periodic checks. Indeed, deep foundations are often needed
even onshore, yet they are implicitly inaccessible. They can be assessed via above-ground
Remote Sensing (RS) techniques, at the cost of lower reliability, or through lengthy exca-
vation campaigns (when possible), at the cost of suspending wind power generation [
76
].
Both choices are sub-optimal from a cost-benefit point of view.
For offshore installations, in addition to this problem, scour monitoring is strictly
required. In particular, for monopile HAWTs, a scour-induced reduction in the foundation
integrity over time may lower the first natural frequency of the support structure, making
it dangerously closer to the frequencies where most of the broadband wave and gust
energy is contained [
77
]. Thus, more wave energy may become resonant with the structure,
increasing material fatigue [
78
]. This is even more dangerous in seismic areas due to the
potential joint action of seismic and wave loads on monopile structures [79].
Inclinometers, strain gauges, and optical fibre sensors are all commonly used for
foundation monitoring. To protect them from the harsh external environment, they can
be embedded in the concrete pour (as tested e.g., in Ref. [
80
] with fibre Bragg gratings).
The same sensors can be used to detect surface cracks in shallow foundations as well [
76
].
Displacement sensors, such as laser or infrared telemeters and tachymeters [
81
], can be
located in the foundation to monitor vertical movements considered as precursors of certain
failure mechanisms [82].
5. Non-Destructive Techniques (NDTs)
In this section, the main NDTs used for HAWT integrity monitoring are reviewed
and discussed. Please consider that, while all techniques are reported here separately for
better comprehension, they can (and should) be applied synergistically to compensate one
another for their limitations. As an example, He et al. [
83
] proposed combined ultrasound
and vibrothermography testing for the damage detection of very small surface damages in
composite panels.
For all the tables reported in this and the following section, the following inclusion
criteria have been applied. Due to the large range of different aspects, as well as the
vastness of the scientific literature on each subtopic, only a selection of relevant, recent, and
well-recognized scientific articles is included.
In more detail, all articles (except wherever specified differently) have been published
in the 2000–2021 period and had at least
30 mentions as of 1 December 2021. Please
consider that the review should not be considered exhaustive, and it is therefore only
indicative of the most prominent and recent advancements in each field of research.
For each reported study, its main declared field of application will be indicated as SHM
(external structure, blades, foundations) or CM (any high-speed rotating
machinery component).
5.1. Traditional, Enhanced, and Automatic Visual Inspection (VI)
Simple Visual Inspection is still regarded as the most common form of maintenance
survey for both the tower, the exterior of the nacelle, and (especially) the WT blades. This is
generally performed from the tower bottom through binoculars or, during blade cleaning or
inspection routines, by roped maintenance staff and/or from a gantry or lift system. These
latter solutions expose the personnel to a certain level of danger and are not feasible during
bad weather conditions. Due to the limited space, the VI inspection of the mechanical
components inside the nacelle is more inconvenient, yet still feasible (visual walkarounds).
More detailed surveys can be performed thanks to borescope inspection or through the
gearboxes’ inspection pots.
Sensors 2022,22, 1627 18 of 52
The main advantage of VI is that, as a non-contact technique, it is implicitly non-
invasive, and thus does not alter the structural conditions of the monitored system (even
if a closely-related technique, dye penetrant inspection, applies some external liquid to
enhance the visual contrast of surface cracks). On the other hand, it is obviously limited to
surface damages and defects alone. Due to the scale of HAWTs, it is rather time-consuming.
Even more importantly, it is a qualitative approach; the accuracy of the assessment results
varies highly depending on the inspectors’ skill and is hampered by human errors. These
assessments are neither easily comparable between different maintenance teams.
It is possible to classify the VI strategies according to two main factors:
I if they require human personnel on-site or not;
II if they add any kind of support to the human eyesight.
The classic visual inspection is generally performed periodically, on-site, by one or
more technicians without any form of AI support or advanced device.
In case I, it is also possible to have maintenance staff members using manned platforms
(e.g., small submarines, helicopters, etc.). Otherwise, it is possible to replace the human
personnel with (autonomous or not) robot platforms that can fly, swim, crawl, hike, etc.,
to the needed location. These are essential e.g., in narrow tubes and ducts, or very cost-
efficient (when adjusted for risk) for the inspection of components located underwater or
at dangerous heights. Several examples of potential designs for a climbing robot can be
found in Ref. [
84
]. Finally, it is possible to deploy a permanent system of closed-circuit
cameras for a constant VI (as proposed e.g., with a computer-controlled pan/tilt zoom
camera system in Ref. [85]).
Case II is linked to the use or not of computer vision approaches. Generally, survey
teams are commonly equipped with hand-held devices (e.g., digital cameras). However,
these are often intended for visual documentation rather than quantitative data analysis.
These and other optical instruments can both enhance human eyesight or replace it com-
pletely. Two examples of the first case are the use of the line or edge detection methods
such as the ones applied for WT blade surface crack detection in, respectively, Ref. [
86
] and
Ref. [
87
]. In this sense, the Canny algorithm was found to be the most efficient choice in
many similar studies [
88
,
89
]. Similar techniques were utilized to estimate the extension of
the surface area damaged by machining processes (e.g., drilling) in composite laminates for
WT blades [90].
In this second case, the term automated visual inspection is used. Automatic computer
vision approaches can be image- or video-based. These allow for the inspection of large
surfaces rapidly and reliably, even for the smallest surface defects, including e.g., barely
visible impact damages (BVIDs) due to bird strikes on composite materials.
To summarize, non-conventional VI strategies can be classified (as in Table 5) de-
pending on the (manned or unmanned) platform and the potential use of any Computer
Vision techniques.
Table 5.
Some notable and recent examples of advanced and automated VI strategies applied for the
NDE of wind turbines.
Study Year Platform
Computer Vision/Video or
Image Processing1
Technique
Application
Stokkeland et al. [91] 2015 Digital camera-equipped
multi-copter UAV.
Computer Vision was also utilized for
autonomous navigation (moving along the
blades to acquire pictures). SHM
Park et al. [92] 2015 Fixed Digital camera
(laboratory test only)
Image segmentation, canny edge detection,
and Hough Transform are applied to
evaluate the angle changes in the nuts. The
method is proposed for bolt loosening
monitoring in the ring flange joints in
WT towers.
SHM
Sensors 2022,22, 1627 19 of 52
Table 5. Cont.
Study Year Platform
Computer Vision/Video or
Image Processing1
Technique
Application
Wang et al. [93] 2017 Remotely-controlled,
digital
camera-equipped UAV.
Cascading classifiers (several variants)
were applied to detect and locate pixel
regions containing cracks in the images. SHM
Reddy et al. [94] 2019 Digital camera-equipped
multi-copter UAV. A convolutional neural network
(CNN) architecture. SHM
Shihavuddin et al. 1[95]. 2019 Digital camera-equipped
multi-copter UAV.
Deep learning-based damage detection and
classification. Specifically, the authors used
the well-established faster region-based
CNN (R-CNN) algorithm [96] and
compared it to other similar architectures
(R-CNN, Fast R-CNN, SSD, and R-FCN).
SHM
Yang et al. [97] 2021 Digital
camera-equipped UAV.
The authors used the pre-trained CNNs
described in Ref. [98] after integrating
them with their image dataset via
Transfer Learning 2.
SHM
1
A dataset of several hundred UAV-taken pictures of a single wind turbine has been released linked to this
study [
99
].
2
Few pictures of wind turbine blades are available easily within common annotated image datasets
such as ImageNet and AlexNet.
Regarding the use of robotized inspection, unmanned aerial vehicles (UAVs) are
the predominant choice. Other non-flying alternatives are more uncommon and have
received less attention from practitioners and researchers alike. For instance, Lim et al. [
100
]
proposed the concept of an inchworm-like robot, with telescopic motion, for blade VI. The
concept is intended to carry several sensors, including a camera. Damage detection and
localisation are achieved for surface cracks using computer vision and blob labelling.
Regarding computer vision and automatic VI, the most common approach is to use
deep neural networks (DNNs), especially some sort of region-based convolutional neural
networks such as the Fast R-CNN [
101
], faster R-CNN [
96
], or similar variants, to au-
tomatically detect, localize, and estimate the severity of surface cracks. Often, reliable
R-CNNs architectures, pre-trained over some generic image dataset, are employed after
transfer learning, fine-tuning, and/or retraining with specific datasets, many of which are
already available online. Depending on the typologies of damage included in the training
dataset, AI can assess different damages such as cracks, breakages, oil stains, etc. [
102
] Some
recently-developed and fast-growing pre-trained models include VGG-16 [
103
], ResNet-
v2 [
104
], Inceptionv4 [
105
], YOLOv5 [
106
], and EfficientNet [
107
]. Several variants of these
models are available, including hybrid architectures such as e.g., Inception-ResNet-v2 [
105
].
One recent example can be found in Ref. [108].
However, VI is not the only non-contact approach for NDT. Other RS techniques
include, for instance, optical measurement techniques, laser-based approaches, and video
spectroscopy.
5.2. Optical Methods
Image and video processing are wide research fields. Some optical measurement
technologies are well-established and extensively used; the classic example is digital image
correlation (DIC), which is well-established for the experimental testing of materials in
general [
109
] and nowadays is quite common for the full-field strain inspection of WT
blades in particular (see e.g., Refs. [
110
,
111
]; other highly cited works are reported in
Table 6). If multiple cameras are available, stereophotogrammetry can be used for strain
measurements as well [
112
]. The method has been tested as well for drone-borne images
taken by a multi-copter UAV [113].
Apart from blade monitoring, DIC measurements have been (less commonly) applied
to the tower structure as well (e.g., in Ref. [
114
]). Instead, no relevant applications for CM
were found during this literature review.
Sensors 2022,22, 1627 20 of 52
A more recent video processing technique, the phase-based motion magnification
(PBMM [
115
]), is also gaining researchers’ attention as a means to extract modal parameters
from imperceptible vibrations [
116
,
117
]. This technique has been proven to be feasible
for SHM purposes [
88
,
89
] and applied in combination with stereophotogrammetry to WT
blades [
118
]. All these techniques resort, in some ways, to the pixels’ brightness (amplitude)
and/or phase and are therefore generally limited to the visible band of the electromagnetic
spectrum. They can be preferably used off-site under controlled laboratory conditions
(illumination etc.) but have been validated for outdoor investigation on-site as well, with
natural illumination.
Table 6.
Some notable and recent examples of DIC and other optical techniques applied for the NDE
of wind turbines.
Study Year Technique Notes Application
Baqersad et al. [119] 2012 3D DIC
The authors used two stereoscopic
high-speed cameras to record the
vibrations of a WT blade with
optical targets attached to its
surface (excited with
hammer hits).
SHM
LeBlanc et al. [120] 2013 3D DIC
The full-field displacement and
strain fields of one CX-100
9 m-long WT blade were
estimated. The damaged areas
were located from discontinuities
in the curvature shapes.
SHM
Winstroth et al. [121] 2014 3D DIC and
point tracking
A random black-and-white dot
pattern was applied at four
different radial positions on one
blade of a three-bladed rotor. The
tests were performed in situ on the
operating HAWT.
SHM
Carr et al. [122] 2016 DIC and 3D Dynamic
Point Tracking (3DPT)
The authors compared the
dynamic stress and strain fields
obtained with their
video-extracted measurements
with the readings from attached
strain gauges.
SHM
Another optical method is optical coherence tomography (OCT), firstly proposed for
the noninvasive cross-sectional imaging of biological tissues [
123
] and then further refined
for high depth resolution in Ref. [
124
]. The method is based on an external broad bandwidth
light source, a charge-coupled device camera, and a series of other components (the setup
change accordingly to the specific implementation, several variants exist). The technique
has been proven to be suitable for the ultrahigh-resolution imaging of internal defects in
fibre-reinforced polymers such as the ones currently used for WT blade manufacturing
(CFRP and GFRP) [
125
]. For these materials, the method was compared to X-ray scanning
in [
126
]. Very recently, Ref. [
127
] used near- and mid-infrared ultrahigh-resolution OCT
for subsurface defects on metal samples covered with marine coatings. This application is
well-suited for the undersea parts of the substructure of an offshore HAWT.
One highly cited work (>30 citations as of 1 December 2021) was found in the scientific
literature for OCT; Liu et al. [
128
] studied the delamination growth in a GFRP WT blade. A
tridimensional geometric model of the crack surfaces was reconstructed.
Sensors 2022,22, 1627 21 of 52
5.3. Laser-Based Measurement Techniques (LDV, LiDAR, and Shearography)
Some applications of laser technology in HAWT monitoring include laser vibrometry
and laser scanning [
129
]. However, apart from the specific acquisition techniques, 1D,
2D, or 3D laser Doppler velocimeters (LDVs) will return velocity time series that can be
used for vibration-based SHM, which will be discussed elsewhere. Another example of
laser-based approaches is the light detection and ranging (LiDAR) technology, which was
proposed e.g., in Schäfer et al. [
130
] to be mounted on a multicopter UAV prototype for the
3D mapping of HAWTs in situ.
Finally, shearography testing (ST), is another optical technique that applies coherent
laser illumination for surface deformation measurements. It can be considered as a sort
of speckle interferometry with laser point patterns; thus, it is a short-range laser-based
NDT. ST offers full-field and fast (even real-time) non-contact imaging; it can then be
used for defect and damage detection by searching for deformation anomalies [
131
]. The
technique has been investigated since the early 1990s [
132
], yet is used mainly for indoor
industrial applications or under controlled laboratory conditions only (e.g., Ref. [
133
]), even
if portable devices for in situ testing are nowadays available. However, this equipment is
generally more expensive and complex than the ones for other NDTs [133].
Shearography has been applied to detect delamination, debonding, impact damage,
wrinkles, and dry spots in composite WT blades [
131
]; to monitor their response under
quasi-static failure tests [
132
]; to evaluate the soundness of bonding in laminated compos-
ites [
134
]; and to assess bird strike BVID in carbon/glass fibre reinforced plastic sandwich
panels [
135
]. Its results have been compared to the ones from active infrared thermography
(in Refs. [
136
,
137
]) and the ones from image correlation, acoustic emission, fibre-optic
strain sensing, and piezoelectric sensing (in Ref. [
138
]). From these and other studies, it
has emerged that the ST technology may still require future research [
139
]. Nevertheless,
some related studies have been published very recently (e.g, Ref. [
140
]), even including
remotely-controlled robotic platforms [141].
5.4. Video Spectroscopy
Even not considering laser-based approaches, optical methods are not strictly limited
to the visible portion of the electromagnetic spectrum (i.e., wavelengths of 380–700 nm, see
Figure 9). The Near-Infrared (NIR, ~700–~900 nm), as well as the Short-, Medium-, and
Long-Wavelength infrared ranges (SWIR, ~900–~1700 nm, MWIR, ~3000–~5000 nm, and
LWIR, ~8000–~14000 nm, respectively) have all been successfully applied for RS in several
engineering applications. Furthermore, the combined use of visible and NIR-SWIR-MWIR
can be easily achieved by combining standard cameras with multi- or hyper-spectral sensors
and thermographic cameras. For instance, the sensor payload proposed in Ref. [
142
] for a
low altitude land survey was intended to allow for synchronous acquisitions over a band
of more than 10
4
nm, with overlaps between the different sensors. This presents several
practical advantages since visible, NIR, and SWIR electromagnetic emissions are mainly
reflected radiations, while MWIR and LWIR are mostly emitted from the object itself. The
former group can provide information regardingthe chemical composition of the irradiated
object, according to its absorbed bands, while the latter depend instead on its thermal state
and energy content. For these reasons, thermographic cameras will be discussed on their
own in a dedicated subsection. Short and very short wavelengths (X-rays and gamma rays)
will be treated separately as well.
Regarding multi- and hyper-spectral imaging, the difference between the two terms
lies solely in the resolution in terms of wavelength bands. Multispectral imaging considers
a small number (typically 3 to 15) of spectral bands, while commercially available hyper-
spectral sensors can discern 244 bands in the VNIR range (sampling 2.3 nm per band) and
254 in the SWIR one (5.8 nm per band) [
143
]. In both cases, the spectral signature of each
material can be used for supervised or unsupervised classification; many algorithms have
been developed for this aim.
Sensors 2022,22, 1627 22 of 52
Figure 9.
The range of the electromagnetic spectrum that can be covered by common optical tech-
niques (digital, multi/hyperspectral, and thermographic cameras), as well as Gamma-ray, X-ray,
microwave, and terahertz testing technologies.
This research field is, however, still underdeveloped for WT applications.
Rizk et al. [144]
discussed the potential offered by a hyperspectral imaging system in detecting damage
to WT blades and icing events. Several damage typologies were considered; the results
demonstrated that hyperspectral imaging could detect surface and subsurface defects, as
well as icing events in their early onset stages. A similar approach was then further tested
for blade defect detection [145].
Finally, considering even longer wavelengths (>1 mm), microwave sensors have been
validated for remote sensing testing, as will be discussed later in a dedicated section.
5.5. Infrared Thermography (IRT) and Other Temperature Measurements
Infrared thermography, also known as thermal imaging, is one of the most common
optical techniques. Indeed, it is applied extensively both for SHM purposes to the external
structure and blades and for the CM of the rotating machinery inside the nacelle.
The IR sensors can be single-point transmitters or cameras with different levels of
spatial definition (generally in the order of the hundreds or thousands of pixels). Many
commercial cameras offer a dual view (visible and IR) simultaneous recording, both with
a pan/tilt control or fixed angle of view. They can also be deployed in parallel and
directly connected to a SCADA system. They do not need to be in direct contact with the
target; this is especially useful for hot machines. However, normally, the sensor must be
mounted to a close distance from the object to obtain reliable readings. Because of this
requirement, it is very rare to temperature monitor more than one target with a single IR
transmitter/camera. Due to the implicit line-of-sight limitations, IRT is generally limited
to a single surface; however, 3D scanning and computer-aided model reconstruction also
allow for 3D thermography over all the external surfaces of a single object. This was proven
to be feasible for WT blades in Ref. [146].
5.5.1. Passive IRT
Inside the nacelle (indoor), temperature readings are particularly useful for the CM of
the gearbox, mechanical brake, generator, main shaft bearing, the yaw and pitch systems,
and the pump motor of the hydraulic system. In Ref. [
147
], it was suggested that any
temperature rise between +1 and +10
C should be considered as a potential index of
minor damage. The same guidelines suggest planning for a repair in 2 to 4 weeks for an
increase between +10 and +35
C, in 1–2 days for +35–+75
C, and immediately if higher
than +75
C (all these indications must be corrected according to the environmental and
Sensors 2022,22, 1627 23 of 52
operating conditions). Thermal images can be used for supervised learning, training a
classifier with readings from the healthy conditions as completed e.g., for brushless DC
motors in Ref. [
148
]. Furthermore, these can be used for the detection of non-mechanical
failures such as fire detection and the monitoring of the high voltage transformer and other
electrical systems (power electronics, control system, etc.). All these uses are well-described
in Refs. [149,150].
Apart from damage detection, IRT is especially well-established outdoor for ice de-
tection on WT blades (see e.g., Ref. [151]). This is relevant from an SHM perspective since
freezing conditions are known to increase the stiffness of the structure [152] and therefore
can change the vibrational response of the same. In turn, this might cause false alarms for
vibration-based anomaly detection, if these confounding influences (i.e., damage-unrelated
environmental effects) are not actively depurated from the recorded output. For temper-
ature monitoring under operating conditions, Ref. [
153
] recently proposed a line laser
thermography approach, reading a fixed point along the rotating blade. IRT can even
be used to monitor the blade de-icing systems (generally an embedded heating wire) for
electrical breakdowns [154].
5.5.2. Active IRT
Indoor and outdoor IRT are not limited to passive thermography. Active heating and
cooling thermography techniques are available for different purposes; a large set of options
were investigated on WT blade samples in Ref. [
155
]. However, these approaches need
some sort of controllable thermal excitation. This can be induced through, e.g., thermal
emitters, ultrasounds, microwaves, eddy currents, flash lamps, etc.
In the case of surface heating (flash lamps [
156
], lasers [
157
], etc.), the properties of
surface defects (such as their depth [
158
]) can be estimated based on the heat conduction
from the surface inward. This approach is known as surface heating thermography (SHT).
It has been also referred to as optical thermography [
159
] and it is (generally) performed
in reflection mode, i.e., with the IR camera and the source heat on the same side of the
target surface. Pulsed, pulsed phase, stepped, modulated, lock-in, and line scanning
thermography are all feasible within SHT.
On the other hand, microwaves, ultrasounds, and high-frequency induction currents
can be used for volume heating thermography (VHT), i.e., heating inside out. These
applications are also known as non-optical thermography. A more detailed discussion
about optical and non-optical excitation sources can be found in Ref. [160].
According to the energy source, it is possible to further classify the VHT techniques as:
1.
eddy current (EC) thermography (or inductive thermography), based on the heath re-
leased by resistive losses to the eddy currents induced by electromagnetic pulses [
161
];
2.
microwave thermography, based on the well-known principles of microwave heating.
The electromagnetic energy is absorbed volumetrically by the target object, favouring
uniform and rapid self-heating;
3. vibrothermography (or thermo-sonic testing), with mechanical waves.
Note that, due to their low electrical conductivity and magnetic permeability, fibre-
reinforced polymers have a high penetration depth (about 50 mm for CFRP under 100 kHz
excitation) and thus can be volumetrically heated with ECs [162].
Some VHT approaches include EC pulsed [
163
], pulsed phase [
162
,
164
], stepped [
165
],
and lock-in [
166
] thermography, as well as microwave pulsed and lock-in thermogra-
phy [
167
]. Most of these strategies can be performed with ultrasonic waves too; another less
widespread alternative is the ultrasound-burst-phase thermography [
168
]. The links be-
tween these VHT techniques and their SHT counterparts, mentioned before, are graphically
displayed in Figure 10.
Sensors 2022,22, 1627 24 of 52
Figure 10. The main IRT techniques available as of 2021.
For completeness’ sake, the main aspects of the most prominent active IRT techniques
will be recalled in the next paragraph, considering both surface and volume heat sources.
As the name suggests, pulsed thermography is based on short-duration energy pulses,
while the pulsed phase is based on a phase analysis in the frequency domain and thermal
wave conduction [
169
]. These two techniques are, arguably, the most common alterna-
tives. The former can be used to detect several damage typologies, e.g., air bubbles [
170
],
inclusions of foreign matter [
171
], deficiency in adhesive bonding [
170
], and other glue
faults [
172
] in GFRP WT blades. The latter was tested and validated, again on GFRP WT
blades, with flash lamps for deep defects and delaminations [173].
Regarding non-optical pulsed approaches, EC-based pulsed thermography has shown
several successful applications in the last decades, especially for CFRP materials, e.g., for
delamination depth evaluation [
174
], impact damage [
175
,
176
], damage classification [
177
],
etc. However, it remains limited by technical issues such as the lift-off effects [
178
] and
the difficulty to discern subsurface defects in the inner layers from debonding at outer lay-
ers [
179
]. EC pulsed phase thermography was tested on CFRP laminates in Refs. [
180
,
181
].
Microwave pulsed thermography was tested on both CFRP specimens (with artificially-
added damages in Ref. [
182
], for delamination detection in Ref. [
183
], and for comparison
to laser-based pulsed thermography in [184]) and GFRP [185].
For what concerns the other (non-pulsed) active IRT techniques, stepped thermog-
raphy uses discrete increases in the temperature, while modulated thermography uses
frequency-modulated thermal waves. Lock-in thermographic analysis uses periodic input
energy waves to detect and localize internal inhomogeneities. This is achieved thanks to the
interferences between incoming and reflected heat waves; thus, the strategy is similar to ul-
trasonic testing, as will be discussed later. The technique was successfully tested with both
optical and non-optical heat sources. In Ref. [
186
], 1400 W halogen lamps were utilized for
the detection of skin–skin and skin–core delamination in a CX-100 WT blade. A commercial
microwave over was used in Ref. [
187
] with a CFRP specimen. Lock-in vibrothermography
was tested for delamination detection of CFRP specimens in Ref. [
188
]. Vibrothermography
can be also performed with high-energy bursts in the 10–50 kHz range [189].
Finally, line scanning thermography is a NASA patented technique that uses an IR
sensor moving in synchronous with the heat source to dynamically investigate metallic or
composite surfaces. It has been tested for impact damages on CFRP panels [
190
] and to
assess manufacturing defects in GFRP blades [191].
Some relevant studies for both active and passive IRT are reported in Table 7. Most
of the scientific articles encountered in this review discussed SHM applications for WT
Sensors 2022,22, 1627 25 of 52
blades; only a few works discussed applications for rotating machinery components such
as, e.g., fatigue-damaged gears [
192
]. Several of these techniques can be applied to metal-
lic/ferromagnetic components as well, for the structural monitoring of HAWT towers (e.g.,
using pulsed EC to detect corrosion [192]).
Table 7.
Some notable and recent examples of IRT techniques applied for the NDE of wind turbines.
Study Year Technique Notes Application
Rumsey & Musial [193] 2001 Passive IRT
Infrared thermography was applied by
the National Wind Technology Center at
the National Renewable Energy
Laboratory for the testing of full-size WT
blades. One of the tests performed was a
fatigue test in which a cyclic load was
applied to the WT blade until failure.
SHM
Dattoma et al. [194] 2001 Active IRT (external
heating and readings
during the cooling phase)
The IRT procedure was experimentally
tested on a WT blade sandwich panel,
taken from the box spar. Glue infiltration,
water ingress, and skin–core debonding
were tested.
SHM
Hahn et al. [195] 2002 Thermoelastic stress
analysis
Used to monitor the stress distribution
on a GFRP blade during static and
fatigue tests. Strain gauges were applied
as well to assess the integrity of the
root section.
SHM
Cheng & Tian [196] 2011 Inductive IRT (pulsed
eddy current
thermography)
The proposed method is based on
inductive thermography for the
inspection and assessment of
CFRP components.
SHM
Pan et al. [197] 2012 Pulsed eddy current
The inductor and the IR camera were
placed on opposite sides to detect
damage in the heat transmission mode
on CFRP specimens intended for
WT blades.
SHM
Cheng & Tian [198] 2013 Pulsed eddy current
Detected surface cracks, impact cracks,
defects, and delaminations from
transient thermal images or videos on
CFRP specimens.
SHM
Dattoma & Giancane [
199
]
2013 Passive IRT during
fatigue tests
Compared DIC and IRT results on a
GFRP specimen employed for
WT blades. SHM
Galleguillos et al. [200] 2015 Passive IRT from a
UAV platform
Performed in situ surveys on rotating
WT blades (in-service) with passive IRT
from an unmanned rotorcraft. SHM
Gao et al. [201] 2016 Pulsed eddy current
Developed a multidimensional tensor
model based not only on the analysis of a
single physical field such as heat
conduction (conventional approach) but
also on the inclusion of other properties
such as electrical conductivity and
magnetic permeability as well.
SHM
Paulmbo et al. [202] 2016 Lock-in IRT analysis (heat
source: halogen lamps)
The technique was tested for the
debonding of GFRP joints and compared
to ultrasonic testing. SHM
Yang et al. [203] 2016 Pulsed eddy current
Combined eddy current pulsed
thermography and thermal-wave-radar
analysis for the assessment of
delamination on CFRP blades.
SHM
Palumbo et al. [204] 2017 Thermoelastic phase
analysis
The study focused on the fatigue damage
analysis on GFRP specimens, analysing
the thermal signal in the
frequency domain.
SHM
Sensors 2022,22, 1627 26 of 52
5.5.3. Physically-Attached Temperature Sensors
While conventional or advanced IRT are more prominent for blade inspection, tem-
perature changes are often utilized for monitoring rotating machinery components as well.
However, this is generally achieved not by RS but rather by employing physicallyattached
sensors, such as thermocouples or similar electrical devices. Usually, at least three tempera-
ture sensors are installed at the main shaft support bearing and the high-speed shaft bearing,
and for the lubricant oil [
205
]. The operating temperatures of the generator, converter,
and transformer are often monitored as well. These readings are generally included in the
SCADA dataset along with ambient temperature [
206
]; therefore, they are generally pro-
cessed along with several other thermophysical measurements to detect anomalies. Some
noteworthy examples for CM focused solely and/or prominently on temperature readings
can be found in Guo et al. [
59
], Guo & Bai [
207
], Cambron et al. [
208
], and
Astolfi et al. [209]
.
Even more specifically, approaches based on oil temperature measurements were recently
reviewed by Touret et al. [210].
5.6. Radiographic Testing (RT)
Radiography is a conventional approach for the internal inspection of structures and
mechanical systems. In this regard, the most common radioactive sources available for
industrial applications are X- and Gamma-rays. However, the distinction between X-rays
and Gamma rays is not always very clearly defined. Depending on the conventions,
wavelengths between
10
10
and
10
12
m are generally considered X-rays. Shorter
wavelengths are classified as Gamma rays. Neutron radiography is also a viable application,
even if there is a very limited number of available examples encountered in the scientific
literature (e.g., in Ref. [211]).
Computed Tomography (CT), specifically, saw a relevant increase in popularity in the
last decade, mainly due to the increased availability of X-rays, improvements in spatial
resolution, and the reduced acquisition time [
212
]. These CT scanners can be 2D or 3D and
portable (hand-held or via manned or unmanned platforms such as in Ref. [
213
]), even if
the largest ones are fixed and available at specific laboratory testing locations. These allow
sections of WT blades up to 4.5 m-long to be scanned in a single take [214].
The main concept is to transmit ionising radiation throughout a dense material, mea-
suring its attenuation along the photon path. For a homogeneous material, the amount
of total attenuation would be a spatial constant, only depending on the thickness and
density of the material. This is visually presented as pixel readouts, generally in a greyscale.
Flaws and density inhomogeneities can be detected and located as anomalies in the image
outputs, with extremely good spatial resolution thanks to the very short wavelengths
utilized. This technology is particularly efficient for voids and other discontinuities that lay
parallel to the ray beam. On the other hand, the energy of photons is proportional to their
frequency (thus, inversely proportional to wavelength). This makes RT one of the most
energy-demanding NDTs. Therefore, it is a relatively expensive technology, and harmful
for biological tissue (since the rays can interfere and damage them at the molecular scale).
Thus, the operators must be adequately protected and the procedure is inherently more
hazardous than other options.
For this and other practical reasons, RT is more common for laboratory experiments
rather than in situ applications. Thus, for wind turbine applications, RT can be better used
for material testing, research on fracture mechanics, and forensic analysis after structural
failure. For instance, Mishnaevsky Jr. et al. [
215
] used an X-ray CT to investigate the
erosion mechanisms on the leading edge of WT blades at a microscopic scale. Jespersen
and Mikkelsen [
216
] and Baran et al. [
217
] used the same technique to investigate, respec-
tively, the evolution of fatigue damage and manufacturing defects (fibre misalignment
and porosities) in GFRP specimens intended for blade manufacturing. For the gearbox
and other mechanical components, an X-ray CT was used on laboratory experiments by
Gould et al. [218]
to map the distribution of White Etching Cracks (WECs) networks within
failed bearings, to assess the effects of subsurface steel inclusions as initiation sites. Gegner
Sensors 2022,22, 1627 27 of 52
& Nierlich [
219
] used an X-ray diffraction-based residual stress analysis on operating WT
gearbox bearings to investigate the effects of vibration loading on WECs.
5.7. Microwave and Terahertz Testing
Microwave and terahertz (THz) testing are based on electromagnetic radiation with
long and very long wavelengths, ranging from 1 mm up to 1 m and beyond. These corre-
spond to frequencies between 100 MHz and100 GHz, for standard microwaves, or 100 GHz–
10,000 GHz (0.1–10 THz) for THz approaches. Differently from the short-wavelength
wave-based approaches seen before (and from ultrasonic tests, which will be discussed
later), microwaves can penetrate dielectric materials and thus interact with their inner
structure with limited signal attenuation [
159
]. THz waves can penetrate even thicker
layers of low- and very-low-conductive polymers such as GFRP and other fibre-reinforced
polymers used for the manufacturing of WT blades (except for CFRP, which is instead too
conductive to be inspectable with THz testing [220]).
Indeed, both microwaves and THz techniques are particularly applied for blade
monitoring. Indeed, the only scientific article of this group that met the selection criteria
(>30 citations as of 1 December 2021) to be considered highly influential was the work
of Kuei Hsu et al. [
221
]. They used time-domain spectroscopy and terahertz radiation
to detect damages inserted by sawing small cuts into GFRP laminates (intended for the
manufacturing of WT blades). Martin et al. [
222
] compared a terahertz inverse synthetic
aperture radar system with X-ray and IRT imaging for the inspection of GFRP WT blades’
spar caps.
Some other recent applications to GFRP WT blades—which also include THz-inducted
active thermography—are reported in Refs. [
223
,
224
]. Applications for defect detection on
both CFRP and GFRP composites are discussed in Refs. [
225
,
226
]. Im et al. [
227
] reported
an application of THz testing to their trailing edge. On the other hand, there does not yet
seem to be any relevant study on the microwave or THz inspection technologies for the
condition monitoring of WT rotating machinery elements.
For HAWT towers, microwave technologies can be used to inspect the metallic surfaces
when covered with epoxy or other insulating paints, as it is often applied for long-term
steel protection in severely corrosive atmospheric conditions. This has been proven to be
feasible in Ref. [
228
] for fire protect-coated steel panels, using frequencies between 8 and
12 GHz. THz waves were also proved to be able to detect corroded metals under thick
insulating layers, plus water intrusion in sandwich panels [
229
]. THs imaging was proven
to provide a higher resolution than, e.g., ultrasound testing; on the other hand, it is limited
by a lower penetration capability [230].
Considering a different approach, Pieraccini et al. [
231
] used a portable, high-speed
continuous-wave step-frequency interferometric radar, which transmits continuous mi-
crowaves (central frequency: 16.75 GHz) at discrete frequency values, to remotely record
the dynamic behaviour of an onshore WT.
5.8. Electromagnetic Testing (ET)
Electromagnetic Testing, especially using the already cited ECs, relies on the use of
changes in the electric conductivity to detect and localize damages in metallic and non-
metallic components. A complete review, accounting for several fields of application, can
be found in Ref. [
232
]. This strategy is well-known and widespread in manufacturing
industries. Thus, these techniques are viable for the tower structure and substructure. For
instance, they are widely used for weld inspection [
233
]; hence, they can be applied for the
inspection of the circumferentially and longitudinally welded connections in the tubular
steel components.
EC testing can be used for conductive composites as well, e.g., for CFRP WT blades,
enabling the detection of both surface and subsurface damages and defects. However, while
the concept has proven to be feasible (see, for instance, Ref. [
234
]), the relatively low conduc-
Sensors 2022,22, 1627 28 of 52
tivity of these composites might compromise the detection accuracy for certain typologies
of damages such as delamination [234], making EC thermography a preferable option.
Pulsed ECs [
235
] have also been used to detect steel corrosion [
236
] and low-energy
impacts in CFRP composites [
237
]. The first application is apt for the monitoring of the
tower structure and substructure, especially in critical areas such as the splash zone. The
second technique can be advantageous for blade inspection e.g., after a bird strike.
Finally, radio frequency EC testing has been suggested for less conductive materials
such as CFRP [
237
], being applied for the detection of fibre misalignment, gaps, and local
polymer degradation [238,239].
However, despite the several potential applications, no SHM approach based solely
on EC seems to have obtained noteworthy attention from the scientific community. Hence,
the preferred use of ECs remains for volume heating and IRT.
Electromagnetism-based strategies are less commonly found for CM, except for some
applications of electrostatic monitoring of WT gearboxes [
240
], often resorting to a sin-
gle [
240
,
241
] or multiple [
242
] oil-line electrostatic sensors. The concept has been tested for
WECs in WT gearboxes as well [243].
5.9. Acoustic Emissions (AEs)
The key concept of AE is that, when an internal crack propagates, it releases energy in
form of acoustic waves. Debonding, delamination, crushing, and other kinds of damage
produce localized, transient changes in the stored elastic energy as well. These elastic
waves are therefore also known as stress release waves [
50
] and travel inside the material.
AE testing is based on their detection. This basic procedure is sketched in Figure 11.
Figure 11. The basic concept of AE event detection.
These acoustic waves are generally too weak to be heard by a human bystander, yet
they can be easily detected from (one or more) sensing devices attached to the surface of
the inspected element. There are, however, two main issues:
(1)
not all the typologies of damage emit strong AE;
(2)
even more importantly, many damage-unrelated phenomena emit AEs.
Therefore, especially for in situ testing, these confounding influences may exceed
crack-related emissions.
Assuming their correlation with damage, once detected, these non-audible emissions
can be then also used to estimate and (with multiple sensors) locate the origin of the damage.
This last task is generally performed via time-of-flight triangulation, even if alternatives
with less than three sensors have been proposed, e.g., in Ref. [244].
In SHM, AE event detection is particularly common for WT blade laboratory testing
since AEs occurring during loading conditions are very likely indicative of the presence
of crack propagation. In a load-hold test [
245
], the WT blade is loaded slightly above the
highest service load and then held in position for around 10 min. For an undamaged fibre
composite structure, AEs will occur only during the first loading and not be re-emitted
significantly on subsequent reloading to the same level. Therefore, sustained emission
Sensors 2022,22, 1627 29 of 52
during a load-hold is considered indicative of damage [
245
]. This was documented as early
as in the 1990s during the loadings of blade fatigue tests [246,247].
The most relevant examples of As applications for WT blade monitoring are included
in the second part of Table 8; some other less cited, but still noteworthy studies include the
research completed by the Centre for Renewable Energy Sources (CRES) on an NM48/750
NEG-MICON WT blade, monitored in-service [248,249].
For CM, the basic concept is that faulty mechanisms (e.g., bearing defects) disrupt the
AE waveform, causing a detectable divergence from the readings under normal operating
conditions. Historically, these changes in the AE signatures have been considered to be
observable earlier than significant alterations in the vibrational signatures of the same
pieces of rotating machinery, thus allowing for early damage detection and prognosis [
250
].
AE techniques have been proved feasible for the monitoring of ball bearings, standard
roller bearings [
251
], and tapered roller bearings [
252
]. Soua et al. [
253
] evaluated their
feasibility for gearboxes and generator shafts. Purarjomandlangrudi & Nourbakhsh [
254
]
tested AEs to detect a fault in the outer race bearing in a low-speed shaft rig test, simulating
the internal components of a WT drive train.
The first part of Table 8reports several relevant studies of the last 20 years for CM.
A comprehensive review of previous works about the AE-based structural diagnosis of
bearing defects, gearbox faults, and pumps can be found in Mba & Rao [255].
Table 8.
Some notable and recent examples of AE techniques applied for the NDE of wind turbines.
Study Year Technique Notes Application
Eftekharnejad & Mba [256] 2009 AE waveforms. Applied for the detection of seeded
tooth root cracks in one helical gear
of the wind turbine gearbox. CM
Elforjani & Mba [257] 2010 Continuous AE energy
monitoring.
The authors applied AEs for the CM
of low-speed shafts and bearings
(separately) also considering
different conditions such as lubricant
starvation. The bearing test
demonstrated the AE’s efficiency in
detecting crack initiation
and propagation.
CM
Eftekharnejad et al. [258] 2011 Kurtogram (spectral
kurtosis).
Compared the effectiveness of
applying the kurtogram to AEs and
for a roller bearing on a laboratory
test bench.
CM
Qu et al. [259] 2012 Time synchronous
averaging (TSA)
and kurtosis.
The heterodyne technique used in
telecommunication was used to
pre-process AE signals, reducing the
sampling frequency from MHz
to kHz.
CM
Niknam et al. [260] 2013 PAC-energy (Physical
Acoustic Corporation
PCI-2 AE system).
This study focused on wind turbine
drive trains subject to rotor
unbalances. These unbalances may
be caused by manufacturing defects
or non-uniform accumulation of ice,
dust, moisture, or even damage on
rotor blades.
CM
Ferrando Chacon et al. [261] 2016
Root Mean Square Error,
Peak Value, Crest Factor,
and Information Entropy
of AE waveforms.
The confounding influences induced
by different operating conditions
(load and torque) on the AE
signature of a wind turbine gearbox
were investigated.
CM
Sensors 2022,22, 1627 30 of 52
Table 8. Cont.
Study Year Technique Notes Application
Zhang et al. [262] 2017
Damage localisation was
performed via
triangulation (delays in
the time of arrival).
The first attempt of mechanical fault
localisation for CM inside a wind
turbine gearbox. CM
Joosse et al. [245] 2002 Load-hold test. An early application of AEs off-site
on a detached WT blade. SHM
Anastassopoulos et al. [263] 2002 Load-hold test.
Machine Learning (specifically,
Unsupervised Pattern Recognition)
was applied to AE data from ten
WT blades.
SHM
Blanch & Dutton [264] 2003
Load-hold, stationary, and
operating tests.
AEs applied on-site to attached
blades (both stationary and rotating
during normal operating conditions).
SHM
Paquette et al. [265] 2007 Three-point bending test.
The article documented a 5-year long
project performed at Sandia National
Laboratories (USA) to characterize
WT blades made of carbon fibres.
SHM
Zarouchas & Van
Hemelrijck [266]2011
Peak frequency analysis of
AEs and Digital Image
Correlation.
AEs were used to characterize the
crack growth at different scales in
laboratory specimens, treated with
an adhesive used for WT blades
composites. Tensile and compression
tests were executed. DIC was used to
compare the strain measurements
with the recorded acoustic activity.
SHM
Han et al. [267] 2013 Static loading test.
AEs and strain measurements of a
WT blade inner shear web were
compared, to correlate acoustic
emissions and stress conditions.
SHM
Bouzid et al. [268] 2014 Ambient
excitation(naturally
occurring AEs).
Proposed a Wireless Sensor Network
(WSN) architecture for damage
localisation in the blades of operating
wind turbines (via triangulation).
SHM
Tang et al. [269] 2016 Pencil lead break test.
The acoustic emissions were
generated by breaking a pencil lead
in the blade surface. Proved the
feasibility of damage severity
assessment and growth tracking.
SHM
Gómez Muñoz & García
Márquez [270]2016
Pencil lead break test.
Damage localisation was
performed via
triangulation (delays in
the time of arrival).
Three macro-fibre composite
transducers were applied on the
surface of a WT blade. SHM
Tang et al. [271] 2017 21-day long fatigue test. Unsupervised Pattern Recognition
was applied to a very large dataset of
recorded AEs. SHM
5.10. Ultrasonic Testing (UT)
Differently from AE testing, where the acoustic waves naturally originate at the dam-
age location and propagate freely throughout the whole structure, in UT high-frequency
waves are generated at a damage-unrelated, user-defined origin. The centre frequency is
generally included between 0.1 and 15 MHz but can reach up to 50 MHz; yet, a variety
of frequencies can be used, allowing optimization for resolution and/or penetration. As
for any portable device, the UT can be performed by manned or unmanned platforms,
such as climbing robots or others. Skaga [
272
] investigated the feasibility of UAV-carried
UT sensors, testing it for WT blade monitoring and comparing the results with hand-held
instrumentation.
Sensors 2022,22, 1627 31 of 52
The source of the ultrasonic waves can be either in contact or not with the target
surface. For contact techniques, most of the conventional transducers are piezoelectric;
thus, they require a wet film (gel, oil, or water) as a couplant between them and the test
object. Water immersion is used for this aim in laboratory testing when the test materials
allow it. Nevertheless, some technologies, such as electromagnetic acoustic transducers,
do not require the use of a couplant. Silicon membranes or other solid couplants can be
used for dry coupling as well, even if they are often limited in the range of frequencies
that can transmit. Air-coupled and laser-borne ultrasonic tests are other viable options
for water-incompatible materials. However, the air has a very low acoustic impedance, so
only a very limited amount of acoustic energy is transmitted. The use of high power pulse
lasers (such as Q-switched Nd: YAG and CO
2
lasers [
159
]) bypasses this issue while also
providing long-range capabilities.
Once emitted, the ultrasonic waves then travel along a well-defined direction through
the material thickness or along its surface (guided wave ultrasonic testing), even over long
distances. In this latter case, generally, slightly lower frequencies are applied
(10 kHz–1 MHz)
.
This allows a much larger inspected area to be covered, also known as the insonified
portion of the tested system. Simple transducers (inclined through a plexiglass wedge)
or ring transducers can be used. Nevertheless, guided waves are slightly less common
than other techniques for wind turbine applications; some applications for NDT can be
found in Refs. [
273
,
274
]. They are often applied for in-service inspection as they can cover
a larger area at once than other UT methods. They have also been recently proposed for ice
monitoring on WT blades [275].
The working principles of conventional and guided wave UT technologies are repre-
sented in Figure 12. Both can be used in reflection mode (with one transducer, analysing the
ultrasound echoes) or transmission mode (with one transmitter and one or more receivers).
The key concept is that when the (volume or surface) wave interacts with an inhomogeneity,
it is partially reflected backwards and partially transmitted forward, with the transmitted
wave being attenuated in its amplitude and delayed in its phase (due to the detour around
the inhomogeneous area or the different propagation velocity inside it). These differences
in amplitude and phase are also quantitatively related to the damage extension [276].
Figure 12.
The basic concept of UT. (
a
) conventional, i.e., through the thickness (reflection mode),
(b) guided waves. (through transmission mode).
Conventional UT can be applied to a wide range of materials. For the tower structure
(and substructure, for offshore installations), it can be used for thickness measurement,
monitoring the risk of material loss due to corrosion. These techniques are even more
important for WT blades due to their composite materials. Indeed, ultrasound waves
are particularly well-suited for fibre-reinforced panels, as the random distribution of the
components in the matrix requires high spatial resolution. However, the presence of such
fibres leads to sound scattering and directional (anisotropic) damping in GFRP, CFRP, and
similar materials [
277
]. The ultrasonic must travel through several centimetres of these
fibre-reinforced polymers; thus, a high voltage ultrasonic pulse is needed to send enough
Sensors 2022,22, 1627 32 of 52
energy into the material. For cross-ply CFRP plates, the effects of lamination and anisotropy
must be considered for damage localisation [278].
Apart from the already-mentioned air-coupled and laser ultrasonics, conventional UT
can be further classified into many categories. These include pulse-echo UT, phased/linear
array UT, local resonance spectroscopy, and others.
In a pulse–echo test, a short-duration ultrasonic pulse is sent into the test specimen
using an ultrasonic transducer. The waves then travel through the specimen and reflect at
the opposite end of the material, in the absence of inhomogeneities, or at cavities (flaws)
and/or discontinuities (delaminations) within the material, if present. The reflected waves
are recorded using the same transducer (in sensor mode). The difference in the travelled
distances results in different times of arrivals (also known as time-of-flight), allowing the de-
fect or damage to be located through the thickness of the inspected structure. The amplitude
and other waveform characteristics can be used as well for further characterisations.
Phased array ultrasonic testing (PAUT) is a growing and very promising technology.
By using many small ultrasonic transducers, each one pulsed independently, it can produce
a quasi-flat ultrasonic beam. This can be steered electronically by changing the time delay
between the transducers, thus allowing different angles to be inspected without physically
moving or turning the portable device. A comparison of advantages compared to standard
UT can be found in Ref. [
279
]. For wind turbine applications, it has been applied for blade
monitoring by the researchers of Sandia National Laboratories [
280
], Lamarre [
281
], and
Zhang et al. [
282
]. However, PAUT is still under development, and several improvements
have been achieved recently (see, for instance, Ref. [283]).
For local ultrasonic resonance spectroscopy, a portion of a large component (e.g., a WT
blade) is excited with ultrasounds in a broadband frequency range. The vibrational response
is recorded with a nearby sensor and is used to obtain the local material, geometrical, and
mechanical properties, for each of the scanning-grid locations [
284
]. This can be seen as the
ultrasound equivalent of the classic hammer and tap tests used for standard local resonance
spectroscopy. This NDT is also viable for nonlinear UT e.g., in presence of breathing
cracks [285].
Other options can be found in the scientific literature. Ultrasonic pulse velocity was
proposed in Ref. [
286
], specifically intended for monopile HAWT foundations monitoring
in combination with other NDE techniques. Single-sided inspection via air-coupled UT
guided Rayleigh waves was proposed in Ref. [
287
] to detect waviness in WT blades. Table 9
reports the main applications found for the timeframe of interest (2000–2021). These are
mostly limited to blade inspection, off- or on-site. UT seems to be rarely considered for
CM; one of the very few mentions can be found in a very recent PhD thesis [
288
] for WT
gearbox bearings.
Table 9.
Some notable and recent examples of IRT techniques applied for the NDE of wind turbines.
Study Year Technique Notes Application
Jørgensen et al. [289] 2004 Ultrasonic immersion test
An early example of UT for the
detection of damages and
manufacturing defects. The skin,
glue, laminate, and sandwich layers
were all clearly visible from
the scans.
SHM
Jasiünienéet al. [290] 2008 Ultrasonic immersion test
with moving
water container
A particular type of ultrasonic
immersion test (contact pulse–echo
immersion testing) was used to
assess internal defects in a WT blade.
The geometry of the defects was
recognized from the ultrasound
images obtained.
SHM
Sensors 2022,22, 1627 33 of 52
Table 9. Cont.
Study Year Technique Notes Application
Raišutis et al. [291] 2008 Air-coupled guided wave
ultrasonic test
The authors used an ultrasonic
air-coupled technique to transmit
guided waves, locating internal
defects in a WT blade.
SHM
Jüngert [292] 2008 Guided wave
ultrasonic test
It compared acoustic waves (from
hammer tests, using local resonance
spectroscopy) with ultrasonic guided
waves. Acoustic waves were found
to be less subject to scattering and
damping while travelling through
the fibre-reinforced material but less
sensitive to small damages (due to
their larger wavelength).
SHM
Jüngert & Grosse [293] 2009 Contact pulse-echo tests
Compared local resonance
spectroscopy (from hammer tests)
with contact pulse–echo UT on
sandwich composites and pristine
and delaminated GFRP. Ultrasonic
waves correctly detected debonding
at adhesive areas.
SHM
Jasiünienéet al. [294] 2009
Air-coupled ultrasonic
tests, ultrasonic
immersion tests with
moving water container,
and contact
pulse–echo tests
UT and radiographic techniques
were compared on WT blade
specimens. The ultrasonic
techniques proved to be more
efficient in terms of implementation
as they only require access from one
side. The best imaging results,
however, were obtained by
combining RT and UT techniques.
SHM
Lee et al. [295] 2011 Long distance laser
ultrasonic test
To overcome the attenuation due to
air travelling, a portable laser-based
device was proposed for
long-distance UT, up to 40 m (indoor
laboratory conditions).
SHM
Park et al. [296] 2013 Long distance laser
ultrasonic test
It proposed a new laser ultrasonic
imaging technique, specifically
intended for rotating blades SHM
Ye et al. [297] 2014 Pulse-echo test
A portable device for 2D (surface)
and 3D (volume) UT scanning was
proposed and tested on GFRP WT
blade specimens.
SHM
Park et al. [298] 2014 Long-distance laser
ultrasonic test
Delamination and debonding were
successfully visualized in a GFRP
composite wind blade structure. SHM
Park et al. [299] 2015 Laser ultrasonic
propagation
imaging system
A two-step UT imaging strategy was
proposed, with an initial coarse
scanning followed by a second
refined one limited to the areas
deemed of major interest after the
first step. Tested on a 10 kW GFRP
WT blade.
SHM
García Marquez & Gómez
Muñoz [300]2020 Macro fibre composite
transducers and
sinusoidal shaped signals
Cross-correlation and wavelet
analysis were applied to detect,
assess, and localize delaminations in
WT blades.
SHM
5.11. Oil Monitoring
Since the early 2000s, oil monitoring and lubricant contamination analyses became
very common techniques for machine condition monitoring, also for WTs. Indeed, their
usefulness is twofold. On the one hand, it is important to assess the quality of the oil,
to prevent mechanical failure due to e.g., lubricant scarcity or solid intrusions (mainly
Sensors 2022,22, 1627 34 of 52
iron or soot particle contamination). On the other hand, several lubricant parameters
are considered indicative of the potential occurrence of mechanical faults (see e.g., the
previous Table 2). These parameters include the oil viscosity (at +40
C and +100
C),
potential water content, wear particles in parts per million (or mg/L), the presence of
dissolved solvents or gases in the lubricant, and the oil acidity/alkalinity (for potential
oxidation) [
301
]. They can be used for the quantitative assessment of the health conditions
of the WT gearbox, hydraulic system, compressor, etc. [
301
]. Specifically, the presence of
wear debris formed in rubbing helps to detect and estimate the severity of wear mechanisms.
Total oil contamination can give general information on oil lubricity [
302
]. Surface fatigue
damage of bearing and gear rolling elements, bearing spalling, and gear teeth pitting are
typical mechanical failures of WT gearboxes that result in the release of metallic debris
particles in the oil lubrication system [
303
]. In this regard, oil monitoring is quite well
developed and codified; e.g., ISO 4406 provides proper thresholds for solid particle content
according to their size distribution ((
4
µ
m/
6
µ
m/
14
µ
m). Table 10 reports the most
influential publications in the field of WT oil monitoring from the last 20 years.
The scientific literature also reports several examples of sensors developed for oil
monitoring. These cannot be reported here in full due to space concerns; one can refer
to Hamilton & Quail [
304
] for a dedicated overview. Only as an example, a couple of
noteworthy publications may be recalled. A microacoustic sensor was proposed in [
305
] for
oil viscosity monitoring; Mignani et al. [
306
] applied wide-range absorption spectroscopy,
fluorescence spectroscopy, and scattering measurements to estimate the oil acidity, presence
of water infiltrations, and phosphorus content. This latter parameter is considered a proxy
of wear since it is generally utilized in anti-wear additives but then absorbed by metallic
surfaces over time.
Table 10.
Some common typologies of oil monitoring strategies encountered in the scientific literature.
Study Year Technique Notes Application
Myshkin et al. [302] 2003 Optical ferroanalyzer
The document presented the
operating principle of the optical
ferroanalyzer, a sensing device for
the estimation of total lubricant oil
contamination, for
condition monitoring.
CM
Dupuis [303] 2010 Oil debris monitoring
The technique is based on counting
debris particles and measuring their
size to assess the severity of the
gearbox failure.
CM
Zhu et al. [301] 2013 Several sensing devices
A total of 10 sensors and
6 performance parameters related to
oil oxidation, water contamination,
and particle contamination
were discussed.
CM
Coronado &
Kupferschmidt [307]2014
Water content, particle
concentration, particle
count, dielectric constant,
viscosity, oil colour, and
oil density sensors
The paper mainly described a highly
accelerated stress screening test
chamber to assess the performance
of oil properties sensors under
extreme ambient temperature and
vibration levels. The oil parameters
are intended as considered as proxies
of wind turbine gearbox conditions.
CM
Zhu et al. [308] 2015 Particle filtering, plus
viscosity and dielectric
constant sensors
Related to the previous paper by the
same authors [301], it applied online
oil monitoring for fault detection and
remaining useful life prediction.
CM
Sheng [309] 2016
2.5-MW dynamometer test
facility at U.S. National
Renewable Energy
Laboratory (fully
described in Ref. [310])
The laboratory tests were performed
on full-scale wind turbine gearboxes
in three configurations: run-in,
healthy, and damaged conditions.
CM
Sensors 2022,22, 1627 35 of 52
5.12. Static Strain Measurements
Static and dynamic strain measurements are both widely utilized for SHM purposes,
especially for monitoring the blade static and dynamic response. The specific uses for
vibration recordings will not be discussed here; the use of displacement time histories
is not different from the one of velocity or acceleration time series. Static strains, on the
other hand, can provide insight about the deflection of the WT blades. During laboratory
testing, this is useful to characterize the operational and failure behaviour of the specimens.
On-site, this allows to both monitor their shape (to avoid collisions with the tower) and to
estimate their stress field.
The sensors are generally deployed at the locations of maximum strain, i.e., close to
the clamped cross-sections, that is to say, at the blade roots (potentially on both surfaces
and directed in both the flap- and the edge-wise directions, totalling four sensing devices)
and at the bottom of HAWT towers.
Optical fibres have been extensively investigated and applied for WT strain monitoring.
This derives from their several advantages, such as their immunity to electromagnetic
interference and their good accuracy.
In particular, fibre Bragg grating (FBG) sensors are commonly used for strain mea-
surement, especially for WT blades. These can be interrogated with different types of
optoelectronic instrumentation. They are advantageous since have the same size and me-
chanical properties as the original fibre. They can be placed in series, performing many
measurements along a single fibre (multiplexing). Deploying many FBG sensors in parallel
and perpendicular lines allow, e.g., to detect and track crack growth or to locate an impact.
Other similar devices include microbend fibres, which were proposed for crack de-
tection in adhesive joints [
48
,
311
], and transverse optical fuses, proposed in Ref. [
312
] for
low-energy impact damage detection in laminated panels. The downside is that fibre-optic
methods are still an expensive technology and are difficult to implement en masse.
Strain memory alloys have been proposed for strain measurement and shape sensing as
well. Verijenko & Verijenko [
313
] suggested their use for SHM, also for WT blades. For this
application, they can be conveniently embedded into the laminate during manufacturing.
On the one hand, the scanning for magnetic susceptibility, needed to enable this technique,
is labour intensive. On the other hand, these systems can be deployed as actuators as well
for aerodynamic load control [314], thus utilized for both tasks at different moments.
To conclude, some high-impact examples of applications based on strain measure-
ments are reported in Table 11.
Table 11.
Some notable and recent examples of strain measurements applications for wind turbines.
Study Year Notes Application
Papadopoulos et al. [315] 2000
An early study on the feasibility of static strain
measurements for WT blades. The main
potential causes of error were discussed and
their impact was experimentally estimated.
SHM
Kim et al. [316] 2011
FBG sensors were embedded into a 1/23 scale of
the 750 kW composite blade to evaluate
its deflection. SHM
Dimopoulos et al. [317] 2012
The authors used strain measurements from
strain gauges to experimentally investigate the
buckling behaviour of the thin steel cylindrical
shells which make up the HAWT tower.
SHM
Choi et al. [318] 2012
FBG sensors were applied to estimate the static
tip deflection of a 100 kW GFRP blade. This
shape sensing is intended to avoid potential
collisions with the nearby tower.
SHM
Sensors 2022,22, 1627 36 of 52
Table 11. Cont.
Study Year Notes Application
Kim et al. [319] 2013
Similar to Choi et al. [
318
], the authors suggested
installing FBG sensors at the bonding line
between the shear web and spar cap SHM
Sierra-Pérez et al. [320] 2016
Compared strain measurements taken from
strain gauges, FBG sensors, and Optical
Backscatter Reflectometer (OBR) sensors on a
prototype GFRP WT blade.
SHM
5.13. Other NDE Approaches
Other less-common NDE strategies and/or applications include:
I.
Dynamometer testing, performed off-site on the whole drive train system, to
assess for potential slipping behaviour in the high-speed shaft tapered roller
bearings [321];
II.
Sound-based monitoring, using audio speakers to ensonify the internal cavities
of WT blades and arrays of external microphones to detect pattern changes in the
airborne sound radiation [322,323];
III.
Short-Range Doppler Radar, very recently tested for the son-site SHM of WT
blades [324];
IV.
Multi-sensor apparatuses, such as e.g., the one proposed in Ref. [
325
] (with optical,
acoustical, and vibrational sensing devices) to detect bird and bad strikes.
plus several other techniques that, however, are limited to very few or even only one
peer-reviewed scientific articles. These NDE methods are less established than their more
classic counterparts reviewed previously. In some cases, this is due to their (relatively)
novelty. All these techniques present several interesting qualities but are hampered by
specific limitations as well.
6. Discussion
Tables 12 and 13 summarize the main points of this review, reporting the most common
application for each NDT (indicated by
X
in Table 12) and their main advantages and
limitations (Table 13). This discussion is only limited to the NDT reviewed here; for many
applications, please remember that in many cases, vibrational and SCADA data analyses are
considered convenient and effective alternatives. These are not included in this discussion
and are postponed to a more detailed future work.
Table 12.
Most common monitoring strategies for the different load-bearing and rotating components
of a wind turbine, according to the literature review (in particular [
61
,
326
,
327
]) and considering both
on- and off-site (laboratory) inspection.
SHM CM
Tower Foundations Blades Bearings Shaft Generator Gearbox
VI X X X (limited visibility) X(limited visibility) X(limited visibility)
Optical
measurements X X
Shearography X
IRT X X X X X X
Temperature,
non IRT X X X X
X-ray CT X X X X
ET X X (CFRP only) X
AEs X X X X X
UT X X X
Oil Monitoring X X X
Static strain X X X
Sensors 2022,22, 1627 37 of 52
Table 13. Advantages and disadvantages of each NDT strategy.
Method Advantages Disadvantages
VI
Non-contact
Very simple
Low cost
Does not require extensive training or
specific instruments (man-made VI)
Can be automated (Computer Vision and
autonomous unmanned platforms)
Limited to surface damages and defects.
Safety hazard for the personnel
(if man-made).
Low accuracy and highly subjective
(if man-made).
Optical Measurements and Shearography
Non-contact
Full-field
Relatively fast to perform
High sensitivity to damage
Shearography requires a specific (and
expensive) setup.
Difficult to quantify the extension
of damage.
Some techniques (e.g., DIC) require
surface treatment.
IRT
Non-contact (except vibrothermography)
Full-field
Relatively fast to perform (except lock-in
thermography; depends on the thickness
of the material for pulsed and EC pulsed
thermography)
High sensitivity to damage
Many options (surface and volumetric
heating, different inputs, etc.)
Relatively simple setup (except
microwave thermography)
Highly standardized (e.g., ISO
10880:2017)
Good spatial resolution (depends on the
specific option)
Active IRT requires an active source
of heating.
Only microwaves ensure uniform
volumetric heating.
Only microwaves and
vibrothermography allow
selective heating.
Only lock-in and pulsed phase
thermography are
emissivity independent.
Surface heating thermography is limited
to the outermost layers of the material.
Eddy currents cannot be applied to all
materials (depending on
their conductivity).
Pulsed phase thermography requires
extensive signal processing to analyze
the results.
Damage-unrelated factors may cause a
rise in temperature.
Cannot provide a very accurate
damage diagnosis.
Temperature, non IRT Highly standardized (e.g., ISO
15312:2018).
Requires an embedded sensor (subject to
sensor faults).
Damage-unrelated factors may cause
temperature rise.
X-ray CT Non-contact
Very high spatial resolution Radiation hazard
Complex (and expensive) setup
ET Non-contact Relatively low-cost. Sensitive to lift-off
Limited by the material conductivity
Requires specific instruments.
AEs
Passive (no input required)
Able to detect early-stage cracks and
small defects.
Can be applied on-site and in-service
Can be applied also to low-speed
rotating machinery.
Can cover relatively large areas/volume.
High signal-to-noise ratio.
Frequency range far from
load perturbation.
Relatively expensive.
Requires a very high sampling rate.
Acoustic wave attenuation in the material.
Only detect damages at their inception or
during their growth.
Difficult to quantify the extension
of damage
In general, very noisy and difficult
to interpret.
Sensors 2022,22, 1627 38 of 52
Table 13. Cont.
Method Advantages Disadvantages
UT
Can be applied on-site and in-service
Many options
Can cover relatively large areas/volume,
also with complex geometries
Require an active source of ultrasounds
Coupling issues (especially for
water-incompatible materials)
Ultrasound attenuation in the material
The analysis of the results requires an
expert user
Oil
Monitoring
Easy to install.
Enables the direct characterisation of
several oil parameters.
The results are easy to interpret
Only viable for mechanical systems with
a closed-loop oil supply system.
Expensive for continuous
online monitoring.
Static strain
Can provide both damage detection and
shape sensing capabilities.
Conventional strain gauges are easy
to install.
Can be used to monitor dynamic strain as
well (for vibration-based inspection).
Fibre optics are still expensive and
difficult to install.
Regarding the prevalence of these techniques. as reported in Figure 13 (adapted from
Ref. [
328
], based in turn on data from several sources [
329
331
]), visual inspection remains
the most widespread approach. Temperature analysis is also very common, most probably
due to the very low cost of physical temperature sensors. UT, AE, and IRT are all (in
general) more expensive and thus less frequently deployed. Therefore, it is understandable
that the trend is quite clearly decreasing with increasing operational costs. Permanent
monitoring apparatuses, on the other hand, represent an outlier relative to this general
trend. The reason behind this larger-than-expected deployment despite the higher costs
derives from the perceived usefulness of embedded systems. Moreover, the total costs
account for both the installation—which is relatively expensive for a global system—and
the operating expenses, which instead are relatively convenient. Thus, a vibrational-based,
ML-based SHM apparatus becomes more and more convenient throughout the years, when
the structure/system as well becomes more prone to damage due to ageing and normal use
and consumption. However, as mentioned in Section 3.5, global and continuous monitoring
is intended to enable condition-based maintenance. Therefore, it is not an alternative but
rather a complement to the local damage assessment capabilities of the NDE approaches
reviewed here.
Figure 13.
Qualitative distribution of costs and deployment levels of different NDTs. Based on data
from Refs. [329331].
Sensors 2022,22, 1627 39 of 52
7. Conclusions
Despite the great achievements and growth of the wind industry in the last decades,
the reliability of wind turbines is challenged by premature mechanical faults, blade failures,
and even structural collapses. These issues involve both on- and offshore installations,
isolated or grouped in large and dense wind farms. Their consequences are, however, even
greater for offshore wind farms, due to the logistic of maintenance in the open sea.
Structural Health Monitoring (SHM) is concerned with the overall load-bearing struc-
ture of an asset, while Condition Monitoring (CM) is focused on the fault detection within
the subsystems and components of rotating machinery. Therefore, both are essential for the
correct functioning of wind turbines.
Here, all the main Non-Destructive Techniques (NDTs) and Evaluation (NDE) strate-
gies for wind turbines SHM and CM have been reviewed and thoroughly discussed. This
included the findings of more than 300 documents published in the last 20 years. These
covered all the related aspects, for the convenience of both academic researchers and
industry practitioners.
Overall, it is evident that no single option is superior to the others. The synergies of
contact and non-contact measurement techniques, especially when applied to the estimation
of different physical quantities, should be preferred. Therefore, the main conclusion is that
a large set of different sensing techniques can balance out the limitations and drawbacks
of each single NDT. This allows detecting damage/fault-related anomalies, as well as
discerning them from unrelated operating and environmental variations. This holistic
approach can provide a significant economic and safety benefit to the wind industry.
The NDE approaches reviewed here must be seen in the broader context of Intelligent
Maintenance. Thus, their use should be integrated with permanent monitoring apparatuses,
embedded in the structure and mechanical components and eventually integrated within a
Supervisory Control And Data Acquisition (SCADA) system. These apparatuses, which
are becoming more and more widespread and standard in the industry, will be reviewed in
future works.
Author Contributions:
Conceptualization, M.C. and C.S.; methodology, M.C. and C.S.; writing—
original draft preparation, M.C.; writing—review and editing, C.S.; visualization, M.C.; supervision,
C.S.; project administration, C.S. All authors have read and agreed to the published version of
the manuscript.
Funding: This research received no external funding.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement:
All data reported in this review are available at the referred original sources.
Acknowledgments:
The authors would like to thank Calogero Bonetti for his precious help and support.
Conflicts of Interest: The authors declare no conflict of interest.
Abbreviations
The following abbreviations are used in this manuscript:
AE Acoustic Emission.
AI Artificial Intelligence
BVID Barely Visible Impact Damage
CFRP Carbon Fibre Reinforced Polymer
CM Condition Monitoring
CT Coherence Tomography
DIC Digital Image Correlation
EC Eddy Current
Sensors 2022,22, 1627 40 of 52
ET Electromagnetic testing
FBG Fibre Bragg Grating
GFRP Glass Fiber Reinforced Polymer
HAWT Horizontal Axis Wind Turbine
IRT Infrared Thermography
LCOE Levelized Cost of Energy
LDV Laser Doppler Velocimeter
NDE Non-Destructive Evaluation
NDT Non-Destructive Technique
OCT Optical Coherence Tomography
O&M Operation and Maintenance (cost)
RS Remote Sensing
RT Radiographic Testing
SCADA Supervisory Control And Data Acquisition (system)
SHM Structural Health Monitoring
SHT Surface Heating Thermography
UAV Unmanned Aerial Vehicle
UT Ultrasonic Testing
VBI Vibration-Based Inspection
VHT Volume Heating Thermography
VI Visual Inspection
WEC White Etching Cracks
WT Wind Turbine
References
1.
Brouwer, S.R.; Al-Jibouri, S.H.; Cardenas, I.C.; Halman, J.I. Towards analysing risks to public safety from wind turbines. Reliab.
Eng. Syst. Saf. 2018,180, 77–87. [CrossRef]
2.
World Meteorological Organization. Proceedings of the World Climate Conference. A Conference of Experts on Climate and Mankind;
Secretariat of the World Meteorological Organization: Geneva, Switzerland, 1979.
3.
International Renewable Energy Agency (IRENA). Renewable Capacity Statistics; International Renewable Energy Agency (IRENA):
Abu Dhabi, United Arab Emirates, 2021.
4.
Gestore Servizi Energetici (GSE). Sviluppo e Diffusione Delle Fonti Rinnovabili di Energia in Italia e in Europa—Anno 2019; Gestore
Servizi Energetici (GSE): Rome, Italy, 2021.
5. Wind Europe. Wind in Power—2016 European Statistics; Wind Europe: Brussels, Belgium, 2017.
6.
BloombergNEF. Global Wind Industry Had a Record, Near 100GW, Year as GE, Goldwind Took Lead from Vestas. Available
online: https://about.bnef.com/blog/global-wind-industry-had-a-record-near-100gw-year-as-ge-goldwind-took-lead-from-
vestas/ (accessed on 22 November 2021).
7.
Pek, A. A Gust of Growth in China Makes 2020 a Record Year for Wind Energy. Available online: https://gwec.net/a-gust-of-
growth-in-china-makes-2020-a-record-year-for-wind-energy/ (accessed on 10 February 2022).
8.
Bowers, R.; Comstock, O. The United States Installed More Wind Turbine Capacity in 2020 Than in Any Other Year. Available
online: https://www.eia.gov/todayinenergy/detail.php?id=46976 (accessed on 10 February 2022).
9.
RenewablesNow. Japan Ends 2020 with 4.37 GW of Wind Power. Available online: https://renewablesnow.com/news/japan-
ends-2020-with-437-gw-of-wind-power-731205/ (accessed on 10 February 2022).
10.
Clean Energy Council Australia. Wind Generation. Available online: https://www.cleanenergycouncil.org.au/resources/
technologies/wind#:~{}:text=Wind%20was%20again%20Australia\T1\textquoterights%20leading,breaking%20year%20for%
20the%20sector (accessed on 10 February 2022).
11.
Ministry of New & Renewable Energy—Government of India. Current Status of the Wind Industry. Available online: https:
//mnre.gov.in/wind/current-status/ (accessed on 10 February 2022).
12.
Agência ABEEólica. Dia Mundial do Vento: Brasil Completa 19 GW de Capacidade Instalada de Energia Eólica. Available
online: http://abeeolica.org.br/noticias/dia-mundial-do-vento-brasil-completa-19-gw-de-capacidade-instalada-de-energia-
eolica/ (accessed on 10 February 2022).
13.
Adler, K.; Soo Cheng, P.L. South Korea Proposes World’s Largest Offshore Wind Farm. Available online: https://cleanenergynews.
ihsmarkit.com/research-analysis/south-korea-proposes-worlds-largest-offshore-wind-farm.html (accessed on 10 February 2022).
14. Wind Europe. Wind Energy in Europe—2020 Statistics and the Outlook for 2021–2025; Wind Europe: Brussels, Belgium, 2020.
15. Italy—National Reports. 2021. Available online: thewindpower.net (accessed on 10 February 2022).
16.
Ministero dello Sviluppo Economico. Piano Nazionale Integrato per l’Energia e il Clima; Ministero dello Sviluppo Economico: Rome,
Italy, 2019.
Sensors 2022,22, 1627 41 of 52
17.
Ueckerdt, F.; Hirth, L.; Luderer, G.; Edenhofer, O. System LCOE: What are the costs of variable renewables? Energy
2013
,63,
61–75. [CrossRef]
18.
Myhr, A.; Bjerkseter, C.; Ågotnes, A.; Nygaard, T.A. Levelised cost of energy for offshore floating wind turbines in a life cycle
perspective. Renew. Energy 2014,66, 714–728. [CrossRef]
19.
Costa, Á.; Orosa, J.; Vergara, D.; Fernández-Arias, P. New tendencies in wind energy operation and maintenance. Appl. Sci.
2021
,
11, 1386. [CrossRef]
20.
Ukonsaari, J.; Bennstedt, N. Wind Turbine Gearboxes Maintenance Effect on Present and Future Gearboxes for Wind Turbines; Energiforsk:
Stockholm, Sweden, 2016.
21.
Faulstich, S.; Hahn, B.; Tavner, P.J. Wind turbine downtime and its importance for offshore deployment. Wind Energy
2011
,14,
327–337. [CrossRef]
22.
McMillan, D.; Ault, G.W. Quantification of condition monitoring benefit for offshore wind turbines. Wind Eng.
2007
,31, 267–285.
[CrossRef]
23.
Zonta, D.; Glisic, B.; Adriaenssens, S. Value of information: Impact of monitoring on decision-making. Struct. Control Health
Monit. 2014,21, 1043–1056. [CrossRef]
24.
Sønderkaer, J.; Dalgaard, M.; Nielsen, J.S.; Tcherniak, D.; Ulriksen, M.D. Quantifying the value of SHM for wind turbine blades.
In Proceedings of the 9th European Workshop on Structural Health Monitoring (EWSHM 2018), Manchester, UK, 10–13 July 2018.
25.
Van Horenbeek, A.; van Ostaeyen, J.; Duflou, J.R.; Pintelon, L. Quantifying the added value of an imperfectly performing
condition monitoring system—Application to a wind turbine gearbox. Reliab. Eng. Syst. Saf. 2013,111, 45–57. [CrossRef]
26.
Mcmillan, D.; Ault, G. Towards quantification of condition monitoring benefit for wind turbine generators. In Proceedings of the
European Wind Energy Conference & Exhibition, Milan, Italy, 7–10 May 2007.
27. Gestore Servizi Energetici (GSE). Il Punto Sull’eolico; Gestore Servizi Energetici (GSE): Roma, Italy, 2017.
28. Hau, E.; von Renouard, H. Wind Turbines; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2006.
29.
Guo, L.; Uang, C.-M.; Elgamal, A.; Prowell, I.; Zhang, S. Pushover analysis of a 53 m high wind turbine tower. Adv. Sci. Lett.
2011
,
4, 656–662. [CrossRef]
30. ABB SACE. Quaderni di Applicazione Tecnica N.13 Impianti Eolici; ABB SACE: Zurich, Switzerland, 2011.
31.
Esteban, M.D.; López-Gutiérrez, J.-S.; Negro, V. Gravity-based foundations in the offshore wind sector. J. Mar. Sci. Eng.
2019
,
7, 64. [CrossRef]
32.
Sun, X.; Huang, D.; Wu, G. The current state of offshore wind energy technology development. Energy
2012
,41, 298–312.
[CrossRef]
33.
European Wind Energy Association. Deep Water: The Next Step for Offshore Wind Energy; European Wind Energy Association:
Brussels, Belgium, 2013.
34.
Fingersh, L.; Hand, M.; Laxson, A. Wind Turbine Design Cost and Scaling Model; National Renewable Energy Laboratory: Golden,
CO, USA, 2006.
35.
Barlas, T.; van Kuik, G. Review of state of the art in smart rotor control research for wind turbines. Prog. Aerosp. Sci.
2010
,46,
1–27. [CrossRef]
36.
NENUPHAR. NENUPHAR Report 2010. In Wind Farms Designed for the Offshore Environment; NENUPHAR SARL: Lille,
France, 2010.
37. Tchakoua, P.; Wamkeue, R.; Ouhrouche, M.; Slaoui-Hasnaoui, F.; Tameghe, T.A.; Ekemb, G. Wind turbine condition monitoring:
State-of-the-art review, new trends, and future challenges. Energies 2014,7, 2595–2630. [CrossRef]
38. Cao, W.; Xie, Y.; Tan, Z. Wind turbine generator technologies. Adv. Wind Power 2012,1, 177–204. [CrossRef]
39.
U.S. Department of Energy. Advanced Wind Turbines Drivetrain Concepts; U.S. Department of Energy: Washington, DC, USA, 2010.
40.
Nguyen, T.-C.; Huynh, T.-C.; Yi, J.-H.; Kim, J.-T. Hybrid bolt-loosening detection in wind turbine tower structures by vibration
and impedance responses. Wind Struct. Int. J. 2017,24, 385–403. [CrossRef]
41.
Ashley, F.; Cipriano, R.J.; Breckenridge, S.; Briggs, G.A.; Gross, L.E.; Hinkson, J.; Lewis, P. Report from the Bethany Wind Turbine
Study Committee. 2007. Available online: https://docs.wind-watch.org/bethany-windturbinestudycommittteereport.pdf
(accessed on 10 February 2022).
42.
Bethold, H.; Durstewitz, M.; Rohrig, K. Reliability of Wind Turbines: Experiences of 15 Years with 1500 WTs; Springer:
Berlin/Heidelberg, Germany, 2007.
43.
Ribrant, J.; Bertling, L.M. Survey of failures in wind power systems with focus on swedish wind power plants during 1997–2005.
IEEE Trans. Energy Convers. 2007,22, 167–173. [CrossRef]
44.
Shokrieh, M.M.; Rafiee, R. Simulation of fatigue failure in a full composite wind turbine blade. Compos. Struct.
2006
,74, 332–342.
[CrossRef]
45.
Marín, J.; Barroso, A.; París, F.; Cañas, J. Study of fatigue damage in wind turbine blades. Eng. Fail. Anal.
2009
,16, 656–668.
[CrossRef]
46.
Caithness Windfarm Information Forum. Wind Turbine Accident Data to December 31st 2005; 2005. Available online: http:
//www.caithnesswindfarms.co.uk/accidents.pdf (accessed on 10 February 2022).
47.
Yang, J.; Peng, C.; Xiao, J.; Zeng, J.; Xing, S.; Jin, J.; Deng, H. Structural investigation of composite wind turbine blade considering
structural collapse in full-scale static tests. Compos. Struct. 2013,97, 15–29. [CrossRef]
Sensors 2022,22, 1627 42 of 52
48.
Sorensen, B.F.; Lading, L.; Sendrup, P.; Mcgugan, M.; Debel, C.P.; Kristensen, O.J.D.; Larsen, G.; Hansen, A.M.;
Rheinlander, J.
;
Rusborg, J.; et al. Fundamentals for Remote Structural Health Monitoring of Wind Turbine Blades—A Preproject; Risø National
Laboratory: Roskilde, Denmark, 2002.
49.
Sørensen, B.F.; Jørgensen, E.; Debel, C.P.; Jensen, F.M.; Jensen, H.M.; Jacobsen, T.K.; Halling, K.M.; Jensen, H.M. Improved Design of
Large Wind Turbine Blade of Fibre Composites Based on Studies of Scale Effects (Phase 1)—Summary Report; Risø National Laboratory:
Roskilde, Denmark, 2004.
50.
Ciang, C.C.; Lee, J.-R.; Bang, H.-J. Structural health monitoring for a wind turbine system: A review of damage detection methods.
Meas. Sci. Technol. 2008,19, 122001. [CrossRef]
51.
Tavner, P. How are we going to make offshore wind farms more reliable? In Proceedings of the SUPERGEN Wind 2011 General
Assembly, Durham, UK, 20 March 2011.
52.
Sheng, S. Report on Wind Turbine Subsystem Reliability—A Survey of Various Databases (Presentation); National Renewable Energy
Laboratory (NREL): Golden, CO, USA, 2013.
53.
Martin-Del-Campo, S.; Sandin, F.; Strömbergsson, D. Dictionary learning approach to monitoring of wind turbine drivetrain
bearings. Int. J. Comput. Intell. Syst. 2020,14, 106. [CrossRef]
54.
Musial, W.; Butterfield, S.; McNiff, B. Improving wind turbine gearbox reliability. In Proceedings of the 2007 European Wind
Energy Conference & Exhibition, Milan, Italy, 7–10 May 2007.
55.
NSK Company. New Bearing Doctor: Maintenance of Bearings. 2009. Available online: https://www.nsk-literature.com/en/
new-bearing-doctor-maintenance/ (accessed on 10 February 2022).
56.
Liang, Y.; An, Z.; Liu, B. Fatigue life prediction for wind turbine main shaft bearings. In Proceedings of the 2013 International
Conference on Quality, Reliability, Risk, Maintenance, and Safety Engineering (QR2MSE), Chengdu, China, 15–18 July 2013;
pp. 888–893.
57.
Yang, W.; Tavner, P.; Wilkinson, M. Condition monitoring and fault diagnosis of a wind turbine with a synchronous generator
using wavelet transforms. In Proceedings of the 4th IET International Conference on Power Electronics, Machines and Drives
(PEMD 2008), York, UK, 2–4 April 2008.
58.
Yang, W.; Tavner, P.; Wilkinson, M. Condition monitoring and fault diagnosis of a wind turbine synchronous generator drive
train. IET Renew. Power Gener. 2009,3, 1–11. [CrossRef]
59.
Guo, P.; Infield, D.; Yang, X. Wind turbine generator condition-monitoring using temperature trend analysis. IEEE Trans. Sustain.
Energy 2011,3, 124–133. [CrossRef]
60.
Alewine, K.; Chen, W. A review of electrical winding failures in wind turbine generators. IEEE Electr. Insul. Mag.
2012
,28, 8–13.
[CrossRef]
61.
Gowda, M.M.; Mallikarjun, N.; Gowda, P.; Chandrashekhar, R. Improvement of the performance of wind turbine generator using
condition monitoring techniques. In Proceedings of the 2013 7th International Conference on Intelligent Systems and Control
(ISCO), Coimbatore, India, 4–5 January 2013; pp. 495–501.
62.
Henderson, A.R.; Morgan, C.; Smith, B.; Sørensen, H.C.; Barthelmie, R.J.; Boesmans, B. Offshore wind energy in Europe—A
review of the state-of-the-art. Wind Energy 2003,6, 35–52. [CrossRef]
63.
Nielsen, J.J.; Sørensen, J.D. On risk-based operation and maintenance of offshore wind turbine components. Reliab. Eng. Syst. Saf.
2011,96, 218–229. [CrossRef]
64.
Dinwoodie, I.; Quail, F.; McMillan, D. Analysis of offshore wind turbine operation & maintenance using a novel time domain
meteo-ocean modeling approach. In Proceedings of the ASME Turbo Expo 2012, Copenhagen, Denmark, 11–15 June 2012.
65.
Scheu, M.; Matha, D.; Hofmann, M.; Muskulus, M. Maintenance strategies for large offshore wind farms. Energy Procedia
2012
,24,
281–288. [CrossRef]
66.
Besnard, F.; Fischer, K.; Tjernberg, L.B. A model for the optimization of the maintenance support organization for offshore wind
farms. IEEE Trans. Sustain. Energy 2013,4, 443–450. [CrossRef]
67.
Halvorsen-Weare, E.E.; Gundegjerde, C.; Halvorsen, I.B.; Hvattum, L.M.; Nonås, L.M. Vessel fleet analysis for maintenance
operations at offshore wind farms. Energy Procedia 2013,35, 167–176. [CrossRef]
68.
Hofmann, M.; Sperstad, I.B. NOWIcob—A tool for reducing the maintenance costs of offshore wind farms. Energy Procedia
2013
,
35, 177–186. [CrossRef]
69.
Perveen, R.; Kishor, N.; Mohanty, S.R. Off-shore wind farm development: Present status and challenges. Renew. Sustain. Energy
Rev. 2014,29, 780–792. [CrossRef]
70.
Endrerud, O.-E.V.; Liyanage, J.P.; Keseric, N. Marine logistics decision support for operation and maintenance of offshore wind
parks with a multi method simulation model. In Proceedings of the 2014 Winter Simulation Conference, Savannah, GA, USA,
7–10 December 2014.
71.
Nguyen, T.A.T.; Chou, S.-Y. Maintenance strategy selection for improving cost-effectiveness of offshore wind systems. Energy
Convers. Manag. 2018,157, 86–95. [CrossRef]
72.
Fischer, K.; Coronado, D.A. Condition Monitoring of Wind Turbines: State of the Art, User Experience and Recommendations; Fraunhofer-
IWES Bremerhaven; VGB PowerTech Service GmbH: Essen, Germany, 2015; pp. 1–89.
73.
Farinholt, K.M.; Taylor, S.G.; Park, G.; Ammerman, C.M. Full-scale fatigue tests of CX-100 wind turbine blades. Part I: Testing. In
Industrial and Commercial Applications of Smart Structures Technologies; SPIE: Bellingham, WA, USA, 2012; Volume 8343, pp. 213–220.
[CrossRef]
Sensors 2022,22, 1627 43 of 52
74.
Taylor, S.G.; Jeong, H.; Jang, J.K.; Park, G.; Farinholt, K.M.; Todd, M.D.; Ammerman, C.M. Full-scale fatigue tests of CX-100 wind
turbine blades. Part II: Analysis. In Industrial and Commercial Applications of Smart Structures Technologies; SPIE: Bellingham, WA,
USA, 2012; Volume 8343, pp. 221–230.
75.
Bundesamt für Seeschifffahrt und Hydrographie. Konstruktive Ausführung von Offshore-Windenergieanlagen/Design of Offshore Wind
Turbines; Bundesamt für Seeschifffahrt und Hydrographie: Hamburg, Germany, 2007.
76.
Perry, M.; McAlorum, J.; Fusiek, G.; Niewczas, P.; McKeeman, I.J.; Rubert, T. Crack monitoring of operational wind turbine
foundations. Sensors 2017,17, 1925. [CrossRef] [PubMed]
77.
Weijtjens, W.; Verbelen, T.; de Sitter, G.; Devriendt, C. Foundation structural health monitoring of an offshore wind turbine—A
full-scale case study. Struct. Health Monit. 2016,15, 389–402. [CrossRef]
78.
Devriendt, C.; El-Kafafy, M.; de Sitter, G.; Jordaens, P.J.; Guillaume, P. Continuous dynamic monitoring of an offshore wind
turbine on a monopile foundation. In Proceedings of the ISMA Conference, Leuven, Belgium, 17–19 September 2012.
79.
Zheng, X.Y.; Li, H.; Rong, W.; Li, W. Joint earthquake and wave action on the monopile wind turbine foundation: An experimental
study. Mar. Struct. 2015,44, 125–141. [CrossRef]
80.
Rubert, T.; Perry, M.; Fusiek, G.; McAlorum, J.; Niewczas, P.; Brotherston, A.; McCallum, D. Field demonstration of real-time
wind turbine foundation strain monitoring. Sensors 2017,18, 97. [CrossRef]
81.
Resnik, B. Monitoring of concrete bases of wind turbines with modern inclinometers: Approaches and experience. In Proceedings
of the FIG Working Week 2017, Helsinki, Finland, 29 May–2 June 2017.
82.
Currie, M.; Saafi, M.; Tachtatzis, C.; Quail, F. Structural health monitoring for wind turbine foundations. Proc. Inst. Civ.
Eng.-Energy 2013,166, 162–169. [CrossRef]
83.
He, Y.; Chen, S.; Zhou, D.; Huang, S.; Wang, P. Shared excitation based nonlinear ultrasound and vibrothermography testing for
CFRP barely visible impact damage inspection. IEEE Trans. Ind. Inform. 2018,14, 5575–5584. [CrossRef]
84.
Franko, J.; Du, S.; Kallweit, S.; Duelberg, E.; Engemann, H. Design of a multi-robot system for wind turbine maintenance. Energies
2020,13, 2552. [CrossRef]
85.
Kim, D.Y.; Kim, H.-B.; Jung, W.S.; Lim, S.; Hwang, J.-H.; Park, C.-W. Visual testing system for the damaged area detection of wind
power plant blade. In Proceedings of the IEEE ISR 2013, Seoul, Korea, 24–26 October 2013; pp. 1–5.
86.
Zhang, H.; Jackman, J. A feasibility study of wind turbine blade surface crack detection using an optical inspection method. In
Proceedings of the 2013 International Conference on Renewable Energy Research and Applications (ICRERA), Madrid, Spain,
20–23 October 2013; pp. 847–852.
87.
Zhang, H.; Jackman, J. Feasibility of automatic detection of surface cracks in wind turbine blades. Wind Eng.
2014
,38, 575–586.
[CrossRef]
88.
Civera, M.; Zanotti Fragonara, L.; Surace, C. An experimental study of the feasibility of phase-based video magnification for
damage detection and localisation in operational deflection shapes. Strain 2020,56, e12336. [CrossRef]
89.
Civera, M.; Fragonara, L.Z.; Antonaci, P.; Anglani, G.; Surace, C. An experimental validation of phase-based motion magnification
for structures with developing cracks and time-varying configurations. Shock Vib. 2021,2021, 5518163. [CrossRef]
90.
Nagarajan, V.; Rajadurai, J.S.; Kumar, T.A. A digital image analysis to evaluate delamination factor for wind turbine composite
laminate blade. Compos. Part B Eng. 2012,43, 3153–3159. [CrossRef]
91.
Stokkeland, M.; Klausen, K.; Johansen, T.A. Autonomous visual navigation of unmanned aerial vehicle for wind turbine
inspection. In Proceedings of the 2015 International Conference on Unmanned Aircraft Systems (ICUAS), Denver, CO, USA, 9–12
June 2015; pp. 998–1007.
92.
Park, J.-H.; Huynh, T.-C.; Choi, S.-H.; Kim, J.-T. Vision-based technique for bolt-loosening detection in wind turbine tower. Wind
Struct. Int. J. 2015,21, 709–726. [CrossRef]
93.
Wang, L.; Zhang, Z. Automatic detection of wind turbine blade surface cracks based on UAV-taken images. IEEE Trans. Ind.
Electron. 2017,64, 7293–7303. [CrossRef]
94.
Reddy, A.; Indragandhi, V.; Ravi, L.; Subramaniyaswamy, V. Detection of cracks and damage in wind turbine blades using
artificial intelligence-based image analytics. Measurement 2019,147, 106823. [CrossRef]
95.
Shihavuddin, A.; Chen, X.; Fedorov, V.; Christensen, A.N.; Riis, N.A.B.; Branner, K.; Dahl, A.B.; Paulsen, R.R. Wind turbine
surface damage detection by deep learning aided drone inspection analysis. Energies 2019,12, 676. [CrossRef]
96.
Ren, S.; He, K.; Girshick, R.; Sun, J. Faster R-CNN: Towards real-time object detection with region proposal networks. IEEE Trans.
Pattern Anal. Mach. Intell. 2017,39, 1137–1149. [CrossRef]
97.
Yang, X.; Zhang, Y.; Lv, W.; Wang, D. Image recognition of wind turbine blade damage based on a deep learning model with
transfer learning and an ensemble learning classifier. Renew. Energy 2021,163, 386–397. [CrossRef]
98.
Krizhevsky, A.; Sutskever, I.; Hinton, G.E. Image, net classification with deep convolutional neural networks. In Advances in
Neural Information Processing Systems; Curran Associates, Inc.: Red Hook, NY, USA, 2012.
99.
Shihavuddin, A.S.M.; Chen, X. DTU—Drone Inspection Images of Wind Turbine. 2018. Available online: https://data.mendeley.
com/datasets/hd96prn3nc/2 (accessed on 10 February 2022).
100.
Lim, S.; Park, C.-W.; Hwang, J.-H.; Kim, D.-Y.; Kim, T.-K. The inchworm type blade inspection robot system. In Proceedings of the
2012 9th International Conference on Ubiquitous Robots and Ambient Intelligence (URAI), Daejeon, Korea, 26–28 November
2012; pp. 604–607.
Sensors 2022,22, 1627 44 of 52
101.
Girshick, R. Fast R-CNN. In Proceedings of the 2015 IEEE International Conference on Computer Vision (ICCV), Santiago, Chile,
7–13 December 2015; pp. 1440–1448. [CrossRef]
102.
Mao, Y.; Wang, S.; Yu, D.; Zhao, J. Automatic image detection of multi-type surface defects on wind turbine blades based on
cascade deep learning network. Intell. Data Anal. 2021,25, 463–482. [CrossRef]
103. Simonyan, K.; Zisserman, A. Very deep convolutional networks for large-scale image recognition. arXiv 2014, arXiv:1409.1556.
104.
He, K.; Zhang, X.; Ren, S.; Sun, J. Identity mappings in deep residual networks. In European Conference on Computer Vision;
Springer: Cham, Switzerland, 2016; pp. 630–645.
105.
Szegedy, C.; Ioffe, S.; Vanhoucke, V.; Alemi, A.A. Inception-v4, inception-res, net and the impact of residual connections on
learning. In Proceedings of the Thirty-First AAAI Conference on Artificial Intelligence (AAAI-17), San Francisco, CA, USA,
4–9 February 2017.
106.
Jocher, G.; Stoken, A.; Chaurasia, A.; Borovec, J.; Xie, T.; Kwon, Y.; Michael, K.; Changyu, L.; Fang, J.; Abrahim, V.; et al.
Ultralytics/yolov5: v6.0—YOLOv5n “Nano” Models, Roboflow Integration, Tensor, Flow Export, Open, CV DNN Support.
Available online: https://doi.org/10.5281/ZENODO.5563715 (accessed on 10 February 2022).
107.
Tan, M.; Le, Q.V. Efficient, net: Rethinking model scaling for convolutional neural networks. In Proceedings of the International
Conference on Machine Learning, Long Beach, CA, USA, 9–15 June 2019; pp. 6105–6114.
108.
Qiu, Z.; Wang, S.; Zeng, Z.; Yu, D. Automatic visual defects inspection of wind turbine blades via YOLO-based small object
detection approach. J. Electron. Imaging 2019,28, 043023. [CrossRef]
109.
Chu, T.C.; Ranson, W.F.; Sutton, M.A. Applications of digital-image-correlation techniques to experimental mechanics. Exp. Mech.
1985,25, 232–244. [CrossRef]
110.
Leblanc, B.; Niezrecki, C.; Avitabile, P.; Chen, J.; Sherwood, J.A.; Hughes, S.A. Full-field inspection of a wind turbine blade
using three-dimensional digital image correlation. In Industrial and Commercial Applications of Smart Structures Technologies; SPIE:
Bellingham, WA, USA, 2011; Volume 7979.
111.
Wu, R.; Zhang, D.; Yu, Q.; Jiang, Y.; Arola, D. Health monitoring of wind turbine blades in operation using three-dimensional
digital image correlation. Mech. Syst. Signal Process. 2019,130, 470–483. [CrossRef]
112.
Baqersad, J.; Niezrecki, C.; Avitabile, P. Full-field dynamic strain prediction on a wind turbine using displacements of optical
targets measured by stereophotogrammetry. Mech. Syst. Signal Process. 2015,62–63, 284–295. [CrossRef]
113.
Khadka, A.; Dong, Y.; Baqersad, J. Structural health monitoring of wind turbines using a digital image correlation system on a
UAV. In Dynamic Behavior of Materials; Springer Science and Business Media LLC: Berlin/Heidelberg, Germany, 2019; Volume 6,
pp. 85–91.
114.
Curt, J.; Capaldo, M.; Hild, F.; Roux, S. Modal analysis of a wind turbine tower by digital image correlation. J. Phys. Conf. Ser.
2020,1618, 022002. [CrossRef]
115.
Wadhwa, N.; Rubinstein, M.; Durand, F.; Freeman, W.T. Phase-based video motion processing. ACM Trans. Graph.
2013
,32, 1–10.
[CrossRef]
116.
Chen, J.G.; Wadhwa, N.; Cha, Y.-J.; Durand, F.; Freeman, W.T.; Buyukozturk, O. Modal identification of simple structures with
high-speed video using motion magnification. J. Sound Vib. 2015,345, 58–71. [CrossRef]
117.
Yang, Y.; Dorn, C.; Mancini, T.; Talken, Z.; Kenyon, G.; Farrar, C.; Mascareñas, D. Blind identification of full-field vibration modes
from video measurements with phase-based video motion magnification. Mech. Syst. Signal Process.
2017
,85, 567–590. [CrossRef]
118.
Poozesh, P.; Sarrafi, A.; Mao, Z.; Avitabile, P.; Niezrecki, C. Feasibility of extracting operating shapes using phase-based motion
magnification technique and stereo-photogrammetry. J. Sound Vib. 2017,407, 350–366. [CrossRef]
119.
Baqersad, J.; Carr, J.; Lundstrom, T.; Niezrecki, C.; Avitabile, P.; Slattery, M. Dynamic characteristics of a wind turbine blade using
3D digital image correlation. In Health Monitoring of Structural and Biological Systems 2012; SPIE: Bellingham, WA, USA, 2012;
Volume 8348, pp. 711–719.
120.
Le Blanc, B.; Niezrecki, C.; Avitabile, P.; Chen, J.; Sherwood, J. Damage detection and full surface characterization of a wind
turbine blade using three-dimensional digital image correlation. Struct. Health Monit. 2013,12, 430–439. [CrossRef]
121.
Winstroth, J.; Schoen, L.; Ernst, B.; Seume, J.R. Wind turbine rotor blade monitoring using digital image correlation: A comparison
to aeroelastic simulations of a multi-megawatt wind turbine. J. Phys. Conf. Ser. 2014,524, 012064. [CrossRef]
122.
Carr, J.; Baqersad, J.; Niezrecki, C.; Avitabile, P. Full-field dynamic strain on wind turbine blade using digital image correlation
techniques and limited sets of measured data from photogrammetric targets. Exp. Tech. 2016,40, 819–831. [CrossRef]
123.
Huang, D.; Swanson, E.A.; Lin, C.P.; Schuman, J.S.; Stinson, W.G.; Chang, W.; Hee, M.R.; Flotte, T.; Gregory, K.;
Puliafito, C.A.; et al.
Optical coherence tomography. Science 1991,254, 1178–1181. [CrossRef] [PubMed]
124.
Drexler, W.; Morgner, U.; Kärtner, F.X.; Pitris, C.; Boppart, S.; Li, X.D.; Ippen, E.P.; Fujimoto, J.G.
In vivo
ultrahigh-resolution
optical coherence tomography. Opt. Lett. 1999,24, 1221–1223. [CrossRef] [PubMed]
125.
Stifter, D.; Wiesauer, K.; Wurm, M.; Schlotthauer, E.; Kastner, J.; Pircher, M.; Götzinger, E.; Hitzenberger, C. Investigation of
polymer and polymer/fibre composite materials with optical coherence tomography. Meas. Sci. Technol.
2008
,19, 074011.
[CrossRef]
126.
Kastner, J.; Schlotthaue, E.; Burgholzer, P.; Stifter, D. Comparison of X-ray computed tomography and optical coherence
tomography for characterisation of glass-fibre polymer matrix composites. In Proceedings of the World Conference on Non
Destructive Testing, Montreal, QC, Canada, 30 August–3 September 2004; pp. 71–79.
Sensors 2022,22, 1627 45 of 52
127.
Petersen, C.; Rajagopalan, N.; Markos, C.; Israelsen, N.; Rodrigo, P.; Woyessa, G.; Tidemand-Lichtenberg, P.; Pedersen, C.;
Weinell, C.
; Kiil, S.; et al. Non-destructive subsurface inspection of marine and protective coatings using near- and mid-infrared
optical coherence tomography. Coatings 2021,11, 877. [CrossRef]
128.
Liu, P.; Groves, R.M.; Benedictus, R. 3D monitoring of delamination growth in a wind turbine blade composite using optical
coherence tomography. NDT E Int. 2014,64, 52–58. [CrossRef]
129.
Marks, R.; Gillam, C.; Clarke, A.; Pullin, R. Barely visible impact damage detection in a composite turbine blade using 3D
scanning laser vibrometery. In Proceedings of the British Society for Strain Measurement Conference 2015, Edinburgh, UK,
1–3 September 2015.
130.
Schafer, B.E.; Picchi, D.; Engelhardt, T.; Abel, D. Multicopter unmanned aerial vehicle for automated inspection of wind turbines.
In Proceedings of the 2016 24th Mediterranean Conference on Control and Automation (MED), Athens, Greece, 21–24 June 2016;
pp. 244–249.
131.
Hung, Y. Applications of digital shearography for testing of composite structures. Compos. Part B Eng.
1999
,30, 765–773.
[CrossRef]
132.
Sutherland, H.; Beattie, A.; Hansche, B.; Musial, W.; Allread, J.; Johnson, J.; Summers, M. The Application of Non-Destructive
Techniques to the Testing of a Wind Turbine Blade; Sandla National Laboratories: Albuquerque, NM, USA, 1994.
133.
Amenabar, I.; Mendikute, A.; Lopez-Arraiza, A.; Lizaranzu, M.; Aurrekoetxea, J. Comparison and analysis of non-destructive
testing techniques suitable for delamination inspection in wind turbine blades. Compos. Part B Eng.
2011
,42, 1298–1305. [CrossRef]
134.
Hung, Y.; Luo, W.; Lin, L.; Shang, H. Evaluating the soundness of bonding using shearography. Compos. Struct.
2000
,50, 353–362.
[CrossRef]
135.
Gryzagoridis, J.; Findeis, D. Impact damage detection on composites using optical NDT techniques. Insight-Non-Destr. Test. Cond.
Monit. 2010,52, 248–251. [CrossRef]
136.
Hung, M.Y.Y.; Chen, Y.S.; Ng, S.P.; Shepard, S.M.; Hou, Y.; Lhota, J.R. Review and comparison of shearography and pulsed
thermography for adhesive bond evaluation. Opt. Eng. 2007,46, 051007. [CrossRef]
137.
Hung, Y.; Chen, Y.; Ng, S.P.; Liu, L.; Huang, Y.; Luk, B.L.; Ip, R.; Wu, L.; Chung, P. Review and comparison of shearography and
active thermography for nondestructive evaluation. Mater. Sci. Eng. R Rep. 2009,64, 73–112. [CrossRef]
138.
Niezrecki, C.; Avitabile, P.; Chen, J.; Sherwood, J.A.; Lundstrom, T.; Leblanc, B.; Hughes, S.A.; Desmond, M.; Beattie, A.;
Rumsey, M.; et al.
Inspection and monitoring of wind turbine blade-embedded wave defects during fatigue testing. Struct. Health
Monit. 2014,13, 629–643. [CrossRef]
139.
Márquez, F.P.G.; Chacón, A.M.P. A review of non-destructive testing on wind turbines blades. Renew. Energy
2020
,161, 998–1010.
[CrossRef]
140.
Li, Z.; Tokhi, M.O.; Marks, R.; Zheng, H.; Zhao, Z. Dynamic wind turbine blade inspection using micro-polarisation spatial phase
shift digital shearography. Appl. Sci. 2021,11, 10700. [CrossRef]
141.
Li, Z.; Tokhi, M.O.; Zhao, Z.; Gao, J.; Zheng, H. A compact laser shearography system integrated with robotic climber for on-site
inspection of wind turbine blades. In Proceedings of the CLAWAR 2020—23rd International Conference on Climbing and Walking
Robots and the Support Technologies for Mobile Machines, Moscow, Russia, 24–26 August 2020.
142.
Gili, P.; Civera, M.; Roy, R.; Surace, C. An unmanned lighter-than-air platform for large scale land monitoring. Remote Sens.
2021
,
13, 2523. [CrossRef]
143.
AISA DUAL Hyperspectral Sensor Technical Data Sheet. Available online: http://www.adept.net.au/cameras/specim/systems/
pdf/Aisa_Dual.pdf (accessed on 8 March 2021).
144.
Rizk, P.; Al Saleh, N.; Younes, R.; Ilinca, A.; Khoder, J. Hyperspectral imaging applied for the detection of wind turbine blade
damage and icing. Remote Sens. Appl. Soc. Environ. 2020,18, 100291. [CrossRef]
145.
Rizk, P.; Younes, R.; Ilinca, A.; Khoder, J. Defect detection using hyperspectral imaging technology on wind turbine blade. Remote
Sens. Appl. Soc. Environ. 2021,22, 100522. [CrossRef]
146.
Hellstein, P.; Szwedo, M. 3D thermography in non-destructive testing of composite structures. Meas. Sci. Technol.
2016
,27, 124006.
[CrossRef]
147. Holst, G.C. Common Sense Approach to Thermal Imaging; SPIE Optical Engineering Press: Bellingham, WA, USA, 2000.
148. Glowacz, A. Thermographic fault diagnosis of ventilation in BLDC motors. Sensors 2021,21, 7245. [CrossRef] [PubMed]
149.
Anjar, B.; Dalberg, M.; Uppsäll, M. Feasibility Study of Thermal Condition Monitoring and Condition Based Maintenance in Wind
Turbines; Elforsk Rapport; Elforsk AB: Stockholm, Sweden, 2011; p. 11.
150.
Carlsson, F.; Eriksson, E.; Dahlberg, M. Damage Preventing Measures for Wind Turbines. Phase 1—Reliability Data; Elforsk: Stockholm,
Sweden, 2010.
151.
Muñoz, C.Q.G.; Márquez, F.P.G.; Tomás, J.M.S. Ice detection using thermal infrared radiometry on wind turbine blades. Measure-
ment 2016,93, 157–163. [CrossRef]
152.
Peeters, B.; de Roeck, G. One-year monitoring of the Z24-Bridge: Environmental effects versus damage events. Earthq. Eng. Struct.
Dyn. 2001,30, 149–171. [CrossRef]
153.
Hwang, S.; An, Y.-K.; Sohn, H. Continuous-wave line laser thermography for monitoring of rotating wind turbine blades. Struct.
Health Monit. 2018,18, 1010–1021. [CrossRef]
154.
Zeng, Z.; Tao, N.; Feng, L.; Li, Y.; Ma, Y.; Zhang, C. Breakpoint detection of heating wire in wind blade moulds using infrared
thermography. Infrared Phys. Technol. 2014,64, 73–78. [CrossRef]
Sensors 2022,22, 1627 46 of 52
155.
Sanati, H.; Wood, D.; Sun, Q. Condition monitoring of wind turbine blades using active and passive thermography. Appl. Sci.
2018,8, 2004. [CrossRef]
156.
Shepard, S.M. Flash thermography of aerospace composites. In Proceedings of the IV Conferencia Panamericana de END, Buenos
Aires, Argentina, 22–26 October 2007.
157.
Li, T.; Almond, D.P.; Rees, D.A.S. Crack imaging by scanning laser-line thermography and laser-spot thermography. Meas. Sci.
Technol. 2011,22, 035701. [CrossRef]
158. Sun, J.G. Analysis of pulsed thermography methods for defect depth prediction. J. Heat Transf. 2005,128, 329–338. [CrossRef]
159.
Yang, R.; He, Y.; Zhang, H. Progress and trends in nondestructive testing and evaluation for wind turbine composite blade. Renew.
Sustain. Energy Rev. 2016,60, 1225–1250. [CrossRef]
160.
Yang, R.; He, Y. Optically and non-optically excited thermography for composites: A review. Infrared Phys. Technol.
2016
,75,
26–50. [CrossRef]
161.
Netzelmann, U.; Walle, G.; Lugin, S.; Ehlen, A.; Bessert, S.; Valeske, B. Induction thermography: Principle, applications and first
steps towards standardization. Quant. InfraRed Thermogr. J. 2015,13, 170–181. [CrossRef]
162.
He, Y.; Tian, G.; Pan, M.; Chen, D. Eddy current pulsed phase thermography and feature extraction. Appl. Phys. Lett.
2013
,
103, 084104. [CrossRef]
163.
Gao, B.; Bai, L.; Woo, W.L.; Tian, G.Y.; Cheng, Y. Automatic defect identification of eddy current pulsed thermography using
single channel blind source separation. IEEE Trans. Instrum. Meas. 2013,63, 913–922. [CrossRef]
164.
Yang, R.; He, Y.; Gao, B.; Tian, G.Y. Inductive pulsed phase thermography for reducing or enlarging the effect of surface emissivity
variation. Appl. Phys. Lett. 2014,105, 184103. [CrossRef]
165.
He, Y.; Pan, M.; Chen, D.; Tian, G.; Zhang, H. Eddy current step heating thermography for quantitatively evaluation. Appl. Phys.
Lett. 2013,103, 194101. [CrossRef]
166.
Riegert, G.; Zweschper, T.; Busse, G. Lockin thermography with eddy current excitation. Quant. Infrared Thermogr. J.
2004
,1,
31–32. [CrossRef]
167.
Zhang, H.; Yang, R.; He, Y.; Foudazi, A.; Cheng, L.; Tian, G. A review of microwave thermography nondestructive testing and
evaluation. Sensors 2017,17, 1123. [CrossRef]
168. Gleiter, A.; Riegert, G.; Zweschper, T.; Busse, G. Ultrasound lock-in thermography for advanced depth resolved defect selective
imaging. Insight-Non-Destr. Test. Cond. Monit. 2007,49, 272–274. [CrossRef]
169. Maldague, X.; Marinetti, S. Pulse phase infrared thermography. J. Appl. Phys. 1996,79, 2694–2698. [CrossRef]
170.
Zeng, Z.; Tao, N.; Feng, L.C.; Li, Y.S.; Zhang, C.L. Relative thermal contrast analysis in the inspection of wind turbine blades
using pulsed thermography. Adv. Mater. Res. 2011,301, 591–596. [CrossRef]
171.
Bin Zhao, S.; Zhang, C.-L.; Wu, N.-M.; Duan, Y.-X.; Li, H. Infrared thermal wave nondestructive testing for rotor blades in wind
turbine generators non-destructive evaluation and damage monitoring. In International Symposium on Photoelectronic Detection and
Imaging 2009; SPIE: Bellingham, WA, USA, 2009; Volume 7383.
172.
Tao, N.; Zeng, Z.; Feng, L.; Li, X.; Li, Y.; Zhang, C. The application of pulsed thermography in the inspection of wind turbine
blades. In International Symposium on Photoelectronic Detection and Imaging 2011; SPIE: Bellingham, WA, USA, 2011; Volume 8193.
173.
Avdelidis, N.P.; Ibarra-Castanedo, C.; Maldague, X.P.V. Infrared thermography inspection of glass reinforced plastic (GRP) wind
turbine blades and the concept of an automated scanning device. In Proceedings of the SPIE Defense, Security, and Sensing,
Baltimore, MA, USA, 29 April–3 May 2013.
174.
Yi, Q.; Tian, G.; Malekmohammadi, H.; Zhu, J.; Laureti, S.; Ricci, M. New features for delamination depth evaluation in carbon
fiber reinforced plastic materials using eddy current pulse-compression thermography. NDT E Int.
2019
,102, 264–273. [CrossRef]
175.
Ren, W.; Liu, J.; Tian, G.Y.; Gao, B.; Cheng, L.; Yang, H. Quantitative non-destructive evaluation method for impact damage using
eddy current pulsed thermography. Compos. Part B Eng. 2013,54, 169–179. [CrossRef]
176.
He, Y.; Tian, G.; Pan, M.; Chen, D. Impact evaluation in carbon fiber reinforced plastic (CFRP) laminates using eddy current
pulsed thermography. Compos. Struct. 2014,109, 1–7. [CrossRef]
177.
Cheng, L.; Gao, B.; Tian, G.Y.; Woo, W.L.; Berthiau, G. Impact damage detection and identification using eddy current pulsed
thermography through integration of PCA and ICA. IEEE Sens. J. 2014,14, 1655–1663. [CrossRef]
178. Tian, G.Y.; Sophian, A. Reduction of lift-off effects for pulsed eddy current NDT. NDT E Int. 2005,38, 319–324. [CrossRef]
179.
Pan, M.; He, Y.; Tian, G.; Chen, D.; Luo, F. PEC frequency band selection for locating defects in two-layer aircraft structures with
air gap variations. IEEE Trans. Instrum. Meas. 2013,62, 2849–2856. [CrossRef]
180.
Ishikawa, M.; Shiiya, Y.; Kogo, Y.; Hatta, H.; Habuka, Y.; Fukui, R.; Utsunomiya, S. Non-destructive inspection of CFRP laminates
using phase-transformed induction heating thermography. Trans. JSME Jpn. 2015,81, 15–170. [CrossRef]
181.
Yang, R.; He, Y. Eddy current pulsed phase thermography considering volumetric induction heating for delamination evaluation
in carbon fiber reinforced polymers. Appl. Phys. Lett. 2015,106, 234103. [CrossRef]
182.
Palumbo, D.; Ancona, F.; Galietti, U. Quantitative damage evaluation of composite materials with microwave thermographic
technique: Feasibility and new data analysis. Meccanica 2014,50, 443–459. [CrossRef]
183.
Foudazi, A.; Donnell, K.M.; Ghasr, M.T. Application of active microwave thermography to delamination detection. In Proceedings
of the 2014 IEEE International Instrumentation and Measurement Technology Conference (I2MTC), Montevideo, Uruguay, 12–15
May 2014; pp. 1567–1571.
Sensors 2022,22, 1627 47 of 52
184.
Keo, S.-A.; Defer, D.; Breaban, F.; Brachelet, F. Comparison between microwave infrared thermography and CO
2
laser infrared
thermography in defect detection in applications with CFRP. Mater. Sci. Appl. 2013,4, 600–605. [CrossRef]
185.
Cheng, L.; Tian, G.Y.; Szymanik, B. Feasibility studies on microwave heating for nondestructive evaluation of glass fibre reinforced
plastic composites. In Proceedings of the 2011 IEEE International Instrumentation and Measurement Technology Conference,
Hangzhou, China, 10–12 May 2011; pp. 1–6.
186.
Manohar, A.; di Scalea, F.L. Detection of defects in wind turbine composite blades using statistically enhanced lock-in thermogra-
phy. Struct. Health Monit. 2013,12, 566–574. [CrossRef]
187.
Galietti, U.; di Bari, P.; Palumbo, D.; Galietti, U.; Palumbo, D.; Calia, G.; Pellegrini, M. Non destructive evaluation of composite
materials with new thermal methods damage assessment of metallic material. In Proceedings of the ECCM15—15th European
Conference on Composite Materials, Venice, Italy, 24–28 June 2012.
188.
Salazar, A.; Mendioroz, A.; Apiñaniz, E.; Oleaga, A.; Venegas, P.; Sáez-Ocáriz, I. Characterization of delaminations by lock-in
vibrothermography. J. Phys. Conf. Ser. 2010,214, 012079. [CrossRef]
189. Shepard, S.M.; Ahmed, T.; Lhota, J.R. Experimental considerations in vibrothermography. Def. Secur. 2004,5405, 332–335.
190.
Ley, O.; Butera, M.; Godinez, V.; Chung, S. Assessing impact damage on composites using line scanning thermography. Adv.
Mater. Process. 2012,170, 20–26.
191.
Ley, O.; Butera, M.; Godínez, V. Inspection of composite structures using line scanning thermography. SPIE Def. Secur. Sens.
2012
,
8354, 835406.
192.
Gao, B.; Yin, A.; Tian, G.; Woo, W. Thermography spatial-transient-stage mathematical tensor construction and material property
variation track. Int. J. Therm. Sci. 2014,85, 112–122. [CrossRef]
193.
Rumsey, M.A.; Musial, W. Application of infrared thermography nondestructive testing during wind turbine blade tests. J. Sol.
Energy Eng. 2001,123, 271. [CrossRef]
194.
Dattoma, V.; Marcuccio, R.; Pappalettere, C.; Smith, G. Thermographic investigation of sandwich structure made of composite
material. NDT E Int. 2001,34, 515–520. [CrossRef]
195.
Hahn, F.; Kensche, C.W.; Paynter, R.; Dutton, A.G.; Kildegaard, C.; Kosgaard, J. Design, fatigue test and NDE of a sectional wind
turbine rotor blade. J. Thermoplast. Compos. Mater. 2002,15, 267–277. [CrossRef]
196.
Cheng, L.; Tian, G.Y. Surface crack detection for carbon fiber reinforced plastic (CFRP) materials using pulsed eddy current
thermography. IEEE Sens. J. 2011,11, 3261–3268. [CrossRef]
197.
Pan, M.; He, Y.; Tian, G.; Chen, D.; Luo, F. Defect characterisation using pulsed eddy current thermography under transmission
mode and NDT applications. NDT E Int. 2012,52, 28–36. [CrossRef]
198.
Cheng, L.; Tian, G.Y. Transient thermal behavior of eddy-current pulsed thermography for nondestructive evaluation of
composites. IEEE Trans. Instrum. Meas. 2013,62, 1215–1222. [CrossRef]
199.
Dattoma, V.; Giancane, S. Evaluation of energy of fatigue damage into GFRC through digital image correlation and thermography.
Compos. Part B Eng. 2013,47, 283–289. [CrossRef]
200.
Galleguillos, C.; Zorrilla, A.; Jiménez, A.; Diaz, L.; Montiano, Á.L.; Barroso, M.; Viguria, A.; Lasagni, F. Thermographic
non-destructive inspection of wind turbine blades using unmanned aerial systems. Plast. Rubber Compos.
2015
,44, 98–103.
[CrossRef]
201.
Gao, B.; He, Y.; Woo, W.L.; Tian, G.Y.; Liu, J.; Hu, Y. Multidimensional tensor-based inductive thermography with multiple
physical fields for offshore wind turbine gear inspection. IEEE Trans. Ind. Electron. 2016,63, 6305–6315. [CrossRef]
202.
Palumbo, D.; Tamborrino, R.; Galietti, U.; Aversa, P.; Tatì, A.; Luprano, V.A.M. Ultrasonic analysis and lock-in thermography for
debonding evaluation of composite adhesive joints. NDT E Int. 2016,78, 1–9. [CrossRef]
203.
Yang, R.; He, Y.; Mandelis, A.; Wang, N.; Wu, X.; Huang, S. Induction infrared thermography and thermal-wave-radar analysis
for imaging inspection and diagnosis of blade composites. IEEE Trans. Ind. Inform. 2018,14, 5637–5647. [CrossRef]
204.
Palumbo, D.; de Finis, R.; Demelio, G.P.; Galietti, U. Study of damage evolution in composite materials based on the thermoelastic
phase analysis (TPA) method. Compos. Part B Eng. 2017,117, 49–60. [CrossRef]
205. Zhao, H.-S.; Zhang, X.-T. Early fault prediction of wind turbine gearbox based on temperature measurement. In Proceedings of
the 2012 IEEE International Conference on Power System Technology (POWERCON), Auckland, New Zealand, 30 October–2
November 2012; pp. 1–5.
206.
Nie, M.; Wang, L. Review of condition monitoring and fault diagnosis technologies for wind turbine gearbox. Procedia CIRP
2013
,
11, 287–290. [CrossRef]
207.
Guo, P.; Bai, N. Wind turbine gearbox condition monitoring with AAKR and moving window statistic methods. Energies
2011
,4,
2077–2093. [CrossRef]
208.
Cambron, P.; Tahan, A.; Masson, C.; Pelletier, F. Bearing temperature monitoring of a wind turbine using physics-based model. J.
Qual. Maint. Eng. 2017,23, 479–488. [CrossRef]
209.
Astolfi, D.; Scappaticci, L.; Terzi, L. Fault diagnosis of wind turbine gearboxes through temperature and vibration data. Int. J.
Renew. Energy Res. 2017,7, 965–976.
210.
Touret, T.; Changenet, C.; Ville, F.; Lalmi, M.; Becquerelle, S. On the use of temperature for online condition monitoring of geared
systems—A review. Mech. Syst. Signal Process. 2018,101, 197–210. [CrossRef]
211.
Fantidis, J.G.; Potolias, C.; Bandekas, D.V. Wind turbine blade nondestructive testing with a transportable radiography system.
Sci. Technol. Nucl. Install. 2011,2011, 347320. [CrossRef]
Sensors 2022,22, 1627 48 of 52
212.
Garcea, S.; Wang, Y.; Withers, P. X-ray computed tomography of polymer composites. Compos. Sci. Technol.
2018
,156, 305–319.
[CrossRef]
213.
Sattar, T.P.; Marques, V.; Anvo, R.N.; Garrido, G.G.; Kaur, A.P.; Routledge, P.; Markham, K. Climbing robot to perform radiography
of wind blades. In Climbing Robot to Perform Radiography of Wind Blades; Springer: Cham, Switzerland, 2021; pp. 165–176. [CrossRef]
214.
Holub, W.; Haßler, U. XXL X-ray computed tomography for wind turbines in the lab and on site. In Proceedings of the
International Workshop on Smart Material and Structures, Calgary, AB, Canada, 7–10 October 2013.
215.
Mishnaevsky, L.; Fæster, S.; Mikkelsen, L.P.; Kusano, Y.; Bech, J.I. Micromechanisms of leading edge erosion of wind turbine
blades: X-ray tomography analysis and computational studies. Wind Energy 2019,23, 547–562. [CrossRef]
216.
Jespersen, K.M.; Mikkelsen, L.P. Three dimensional fatigue damage evolution in non-crimp glass fibre fabric based composites
used for wind turbine blades. Compos. Sci. Technol. 2017,153, 261–272. [CrossRef]
217.
Baran, I.; Straumit, I.; Shishkina, O.; Lomov, S.V. X-ray computed tomography characterization of manufacturing induced defects
in a glass/polyester pultruded profile. Compos. Struct. 2018,195, 74–82. [CrossRef]
218.
Gould, B.; Greco, A.; Stadler, K.; Xiao, X. An analysis of premature cracking associated with microstructural alterations in an AISI
52100 failed wind turbine bearing using X-ray tomography. Mater. Des. 2017,117, 417–429. [CrossRef]
219.
Gegner, J.; Nierlich, W. Service loading analysis of wind turbine gearbox rolling bearings based on X-ray diffraction residual
stress measurements. Mater. Sci. Forum 2013,768, 723–732. [CrossRef]
220.
Hsu, D.K.; Im, K.-H.; Chiou, C.-P.; Barnard, D.J.; Thompson, D.O.; Chimenti, D.E. An exploration of the utilities of terahertz
waves for the NDE of composites. In AIP Conference Proceedings; American Institute of Physics: College Park, MA, USA, 2011;
Volume 1335, pp. 533–540.
221.
Hsu, D.K.; Lee, K.-S.; Park, J.-W.; Woo, Y.-D.; Im, K.-H. NDE inspection of terahertz waves in wind turbine composites. Int. J.
Precis. Eng. Manuf. 2012,13, 1183–1189. [CrossRef]
222.
Martin, R.W.; Sabato, A.; Schoenberg, A.; Giles, R.H.; Niezrecki, C. Comparison of nondestructive testing techniques for the
inspection of wind turbine blades’ spar caps. Wind Energy 2018,21, 980–996. [CrossRef]
223.
Li, Z.; Soutis, C.; Haigh, A.; Sloan, R.; Gibson, A.; Karimian, N. Microwave imaging for delamination detection in T-joints of wind
turbine composite blades. In Proceedings of the 2016 46th European Microwave Conference (EuMC), London, UK, 4–6 October
2016; pp. 1235–1238. [CrossRef]
224.
Li, Z.; Haigh, A.; Soutis, C.; Gibson, A.; Sloan, R. Microwaves sensor for wind turbine blade inspection. Appl. Compos. Mater.
2016,24, 495–512. [CrossRef]
225.
Im, K.-H.; Lee, K.-S.; Yang, I.-Y.; Yang, Y.-J.; Seo, Y.-H.; Hsu, D.K. Advanced T-ray nondestructive evaluation of defects in FRP
solid composites. Int. J. Precis. Eng. Manuf. 2013,14, 1093–1098. [CrossRef]
226.
Park, J.-W.; Im, K.-H.; Yang, I.-Y.; Kim, S.-K.; Kang, S.-J.; Cho, Y.-T.; Jung, J.-A.; Hsu, D.K. Terahertz radiation NDE of composite
materials for wind turbine applications. Int. J. Precis. Eng. Manuf. 2014,15, 1247–1254. [CrossRef]
227.
Im, K.-H.; Kim, S.-K.; Jung, J.-A.; Cho, Y.-T.; Wood, Y.-D.; Chiou, C.-P. NDE characterization and inspection techniques of trailing
edges in wind turbine blades using terahertz waves. J. Mech. Sci. Technol. 2019,33, 4745–4753. [CrossRef]
228.
Zhang, H.; Gao, B.; Tian, G.Y.; Woo, W.L.; Bai, L. Metal defects sizing and detection under thick coating using microwave NDT.
NDT E Int. 2013,60, 52–61. [CrossRef]
229.
White, J.; Zimdars, D. Time domain terahertz non destructive evaluation of water intrusion in composites and corrosion under
insulation. In Proceedings of the 2007 Quantum Electronics and Laser Science Conference, Baltimore, MD, USA, 6–11 May 2007;
pp. 1–2.
230. Zhong, S. Progress in terahertz nondestructive testing: A review. Front. Mech. Eng. 2019,14, 273–281. [CrossRef]
231.
Pieraccini, M.; Parrini, F.; Fratini, M.; Atzeni, C.; Spinelli, P. In-service testing of wind turbine towers using a microwave sensor.
Renew. Energy 2008,33, 13–21. [CrossRef]
232.
García-Martín, J.; Gómez-Gil, J.; Vázquez-Sánchez, E. Non-destructive techniques based on eddy current testing. Sensors
2011
,11,
2525–2565. [CrossRef] [PubMed]
233.
Gao, P.; Wang, C.; Li, Y.; Cong, Z. Electromagnetic and eddy current NDT in weld inspection: A review. Insight-Non-Destr. Test.
Cond. Monit. 2015,57, 337–345. [CrossRef]
234.
Mook, G.; Lange, R.; Koeser, O. Non-destructive characterisation of carbon-fibre-reinforced plastics by means of eddy-currents.
Compos. Sci. Technol. 2001,61, 865–873. [CrossRef]
235.
He, Y.; Pan, M.; Luo, F.; Tian, G. Pulsed eddy current imaging and frequency spectrum analysis for hidden defect nondestructive
testing and evaluation. NDT E Int. 2011,44, 344–352. [CrossRef]
236.
He, Y.; Tian, G.; Zhang, H.; Alamin, M.; Simm, A.; Jackson, P. Steel corrosion characterization using pulsed eddy current systems.
IEEE Sens. J. 2012,12, 2113–2120. [CrossRef]
237.
He, Y.; Tian, G.; Pan, M.; Chen, D. Non-destructive testing of low-energy impact in CFRP laminates and interior defects in
honeycomb sandwich using scanning pulsed eddy current. Compos. Part B Eng. 2014,59, 196–203. [CrossRef]
238.
Li, Z.; Meng, Z. A review of the radio frequency non-destructive testing for carbon-fibre composites. Meas. Sci. Rev.
2016
,16,
68–76. [CrossRef]
239.
Heuer, H.; Schulze, M.; Pooch, M.; Gäbler, S.; Nocke, A.; Bardl, G.; Cherif, C.; Klein, M.; Kupke, R.; Vetter, R.; et al. Review
on quality assurance along the CFRP value chain—Non-destructive testing of fabrics, preforms and CFRP by HF radio wave
techniques. Compos. Part B Eng. 2015,77, 494–501. [CrossRef]
Sensors 2022,22, 1627 49 of 52
240.
Liu, R.; Zuo, H.; Sun, J.; Wang, L. Electrostatic monitoring of wind turbine gearbox on oil-lubricated system. Proc. Inst. Mech. Eng.
Part C J. Mech. Eng. Sci. 2016,231, 3649–3664. [CrossRef]
241.
Mao, H.; Zuo, H.; Wang, H. Electrostatic sensor application for on-line monitoring of wind turbine gearboxes. Sensors
2018
,
18, 3574. [CrossRef]
242.
Liu, R.; Zhou, J.; Wang, W.; Yao, X. Electrostatic monitoring of wind turbine gearbox under variable operating conditions. In
Proceedings of the 2018 Prognostics and System Health Management Conference (PHM), Chongqing, China, 26–28 October 2018;
pp. 134–137.
243.
Esmaeili, K.; Zuercher, M.; Wang, L.; Harvey, T.; Holweger, W.; White, N.; Schlücker, E. A study of white etching crack bearing
failure detection using electrostatic sensing in wind turbine gearboxes. Int. J. Cond. Monit. 2018,8, 82–88. [CrossRef]
244.
Kijanka, P.; Manohar, A.; di Scalea, F.L.; Staszewski, W.J. Damage location by ultrasonic lamb waves and piezoelectric rosettes. J.
Intell. Mater. Syst. Struct. 2015,26, 1477–1490. [CrossRef]
245.
Joosse, P.A.; Blanch, M.J.; Dutton, A.G.; Kouroussis, D.A.; Philippidis, T.P.; Vionis, P.S. Acoustic emission monitoring of small
wind turbine blades. J. Sol. Energy Eng. 2002,124, 446–454. [CrossRef]
246.
Wei, J.; McCarty, J. Acoustic emission evaluation of composite wind turbine blades during fatigue testing. Wind Eng.
1993
,17,
266–274.
247.
Beattie, A. Acoustic emission monitoring of a wind turbine blade during a fatigue test. In Proceedings of the 35th Aerospace
Sciences Meeting and Exhibit, Reno, NV, USA, 6–9 January 1997.
248.
Tsopelas, N.; Kourousis, D.; Ladis, I.; Anastasopulos, A.; Lekou, D.; Mouzakis, F. Health monitoring of operating windt urbine
blades with acoustic emission. In Emerging Technologies in Non-Destructive Testing V; Taylor & Francis Group: London, UK, 2012;
pp. 347–352.
249.
Papasalouros, D.; Tsopelas, N.; Anastasopoulos, A.; Kourousis, D.; Lekou, D.J.; Mouzakis, F. Acoustic emission monitoring of
composite blade of NM48/750 NEG-MICON wind turbine. J. Acoust. Emiss. 2013,31, 36–49.
250. Yoshioka, T.; Fujiwara, T. A new acoustic emission source locating system for the study of rolling contact fatigue. Wear 1982,81,
183–186. [CrossRef]
251.
Nienhaus, K.; Boos, F.D.; Garate, K.; Baltes, R. Development of acoustic emission (AE) based defect parameters for slow rotating
roller bearings. J. Phys. Conf. Ser. 2012,364, 012034. [CrossRef]
252.
Usgame Sandoval, H.M.; Pedraza Ramirez, C.A.; Quiroga Mendez, J.E. Acoustic emission-based early fault detection in tapered
roller bearings. Ing. Investig. 2013,33, 5–10.
253.
Soua, S.; van Lieshout, P.; Perera, A.; Gan, T.-H.; Bridge, B. Determination of the combined vibrational and acoustic emission
signature of a wind turbine gearbox and generator shaft in service as a pre-requisite for effective condition monitoring. Renew.
Energy 2013,51, 175–181. [CrossRef]
254.
Purarjomandlangrudi, A.; Nourbakhsh, G. Acoustic emission condition monitoring: An application for wind turbine fault
detection. Int. J. Res. Eng. Technol. 2013,2, 907–918.
255.
Mba, D.; Rao, R.B.K.N. Development of acoustic emission technology for condition monitoring and diagnosis of rotating machines;
bearings, pumps, gearboxes, engines and rotating structures. Shock Vib. Dig. 2006,38, 3–16. [CrossRef]
256.
Eftekharnejad, B.; Mba, D. Seeded fault detection on helical gears with acoustic emission. Appl. Acoust.
2009
,70, 547–555.
[CrossRef]
257.
Elforjani, M.; Mba, D. Condition monitoring of slow-speed shafts and bearings with acoustic emission. Strain
2011
,47, 350–363.
[CrossRef]
258.
Eftekharnejad, B.; Carrasco, M.; Charnley, B.; Mba, D. The application of spectral kurtosis on acoustic emission and vibrations
from a defective bearing. Mech. Syst. Signal Process. 2011,25, 266–284. [CrossRef]
259.
Qu, Y.; Bechhoefer, E.; He, D.; Zhu, J. A new acoustic emission sensor based gear fault detection approach. Int. J. Progn. Health
Manag. 2013,4, 32–45. [CrossRef]
260.
Niknam, S.A.; Thomas, T.; Hines, J.W.; Sawhney, R. Analysis of acoustic emission data for bearings subject to unbalance. Int. J.
Progn. Health Manag. 2013,4, 80–89. [CrossRef]
261.
Ferrando Chacon, J.L.; Andicoberry, E.A.; Kappatos, V.; Papaelias, M.; Selcuk, C.; Gan, T.-H. An experimental study on the
applicability of acoustic emission for wind turbine gearbox health diagnosis. J. Low Freq. Noise Vib. Act. Control
2016
,35, 64–76.
[CrossRef]
262.
Zhang, Y.; Lu, W.; Chu, F. Planet gear fault localization for wind turbine gearbox using acoustic emission signals. Renew. Energy
2017,109, 449–460. [CrossRef]
263.
Anastassopoulos, A.; Kouroussis, D.A.; Nikolaidis, V.N.; Proust, A.; Dutton, A.G.; Blanch, M.J.; Jones, L.E.; Vionis, P.;
Lekou, D.J.
;
van Delft, D.R.V.; et al. Structural integrity evaluation of wind turbine blades using pattern recognition analysis on acoustic
emission data. In Proceedings of the 25th European Conference on Acoustic Emission Testing—EWGAE, Prague, Czech Republic,
11–13 September 2002; pp. 11–13.
264.
Blanch, M.; Dutton, A. Acoustic emission monitoring of field tests of an operating wind turbine. Key Eng. Mater.
2003
,245,
475–482. [CrossRef]
265.
Paquette, J.; van Dam, J.; Hughes, S. Structural testing of 9 m carbon fiber wind turbine research blades. In Proceedings of the
45th AIAA Aerospace Sciences Meeting and Exhibit, Reno, NV, USA, 8–11 January 2007; Volume 14, pp. 9922–9932.
Sensors 2022,22, 1627 50 of 52
266.
Zarouchas, D.; van Hemelrijck, D. Mechanical characterization and damage assessment of thick adhesives for wind turbine
blades using acoustic emission and digital image correlation techniques. J. Adhes. Sci. Technol. 2014,28, 1500–1516. [CrossRef]
267.
Han, B.-H.; Yoon, D.-J.; Huh, Y.-H.; Lee, Y.-S. Damage assessment of wind turbine blade under static loading test using acoustic
emission. J. Intell. Mater. Syst. Struct. 2013,25, 621–630. [CrossRef]
268.
Bouzid, O.M.; Tian, G.Y.; Cumanan, K.; Moore, D. Structural health monitoring of wind turbine blades: Acoustic source
localization using wireless sensor networks. J. Sens. 2015,2015, 139695. [CrossRef]
269.
Tang, J.; Soua, S.; Mares, C.; Gan, T.-H. An experimental study of acoustic emission methodology for in service condition
monitoring of wind turbine blades. Renew. Energy 2016,99, 170–179. [CrossRef]
270.
Gomez Munoz, C.Q.; Garcia Marquez, F.P. A new fault location approach for acoustic emission techniques in wind turbines.
Energies 2016,9, 40. [CrossRef]
271.
Tang, J.; Soua, S.; Mares, C.; Gan, T.-H. A pattern recognition approach to acoustic emission data originating from fatigue of wind
turbine blades. Sensors 2017,17, 2507. [CrossRef]
272.
Skaga, S.K. Feasibility study of Unmanned Aerial Vehicles (UAV) Application for Ultrasonic Non-Destructive Testing (NDT) of
Wind Turbine Rotor Blades. Preliminary Experiments of Handheld and UAV Utrasonic Testing on Glass Fibre Laminate. Master’s
Thesis, UiT The Arctic University of Norway, Tromsø, Norway, 2017.
273.
Makaya, K.; Burnham, K.; Tuncbilek, K. Structural assessment of turbine blades using guided waves. In Proceedings of the 10th
European Conference on Non-Destructive Testing, Moscow, Russia, 7–11 June 2010.
274.
Gan, T.-H.; Soua, S.; Dimlaye, V.; Burnham, K. Real-time monitoring system for defects detection in wind turbine structures
and rotating components. In Proceedings of the 18th World Conference on Nondestructive Testing, Durban, South Africa,
16–20 April 2012.
275.
Wang, P.; Zhou, W.; Bao, Y.; Li, H. Ice monitoring of a full-scale wind turbine blade using ultrasonic guided waves under varying
temperature conditions. Struct. Control Health Monit. 2018,25, e2138. [CrossRef]
276.
Yang, J.; He, J.; Guan, X.; Wang, D.; Chen, H.; Zhang, W.; Liu, Y. A probabilistic crack size quantification method using in-situ
lamb wave test and Bayesian updating. Mech. Syst. Signal Process. 2016,78, 118–133. [CrossRef]
277.
Shen, Y.; Cesnik, C.E.S. Hybrid local FEM/global LISA modeling of damped guided wave propagation in complex composite
structures. Smart Mater. Struct. 2016,25, 095021. [CrossRef]
278. Eremin, A.; Glushkov, E.; Glushkova, N.; Lammering, R. Guided wave time-reversal imaging of macroscopic localized inhomo-
geneities in anisotropic composites. Struct. Health Monit. 2019,18, 1803–1819. [CrossRef]
279.
Taheri, H.; Hassen, A.A. Nondestructive ultrasonic inspection of composite materials: A Comparative advantage of phased array
ultrasonic. Appl. Sci. 2019,9, 1628. [CrossRef]
280.
Roach, D.; Neidigk, S.; Rice, T.; Duvall, R.; Paquette, J.A. Development and assessment of advanced inspection methods for wind
turbine blades using a focused WINDIE experiment. In Proceedings of the 33rd Wind Energy Symposium 2015, Kissimmee, FL,
USA, 5–9 January 2015.
281.
Lamarre, A. Improved inspection of composite wind turbine blades with accessible advanced ultrasonic phased array technology.
In Proceedings of the 15th Asia Pacific Conference for Non-Destructive Testing (APCNDT 2017), Singapore,
13–17 November 2017
.
282.
Zhang, N.; Bian, W.; Pan, B. The application of ultrasonic phased array technology in wind turbine blade defect detection system.
IOP Conf. Ser. Earth Environ. Sci. 2021,675, 012077. [CrossRef]
283.
Mohammadkhani, R.; Fragonara, L.Z.; Padiyar, M.J.; Petrunin, I.; Raposo, J.; Tsourdos, A.; Gray, I. Improving depth resolution of
ultrasonic phased array imaging to inspect aerospace composite structures. Sensors 2020,20, 559. [CrossRef] [PubMed]
284.
Rus, J.; Grosse, C.U. Local ultrasonic resonance spectroscopy: A demonstration on plate inspection. J. Nondestruct. Eval.
2020
,39,
1–12. [CrossRef]
285.
Wang, J.; Shen, Y.; Rao, D.; Xu, W. An instantaneous-baseline multi-indicial nonlinear ultrasonic resonance spectral correlation
technique for fatigue crack detection and quantification. Nonlinear Dyn. 2021,103, 677–698. [CrossRef]
286.
Iliopoulos, A.; van Hemelrijck, D.; Vlassenbroeck, J.; Aggelis, D. Assessment of grouted samples from monopile wind turbine
foundations using combined non-destructive techniques. Constr. Build. Mater. 2016,122, 855–862. [CrossRef]
287.
Chakrapani, S.K.; Dayal, V.; Hsu, D.K.; Barnard, D.J.; Gross, A.; Thompson, D.O.; Chimenti, D.E. Characterization of waviness in
wind turbine blades using air coupled ultrasonics. AIP Conf. Proc. 2011,1335, 956–962. [CrossRef]
288. Nicholas, G. Development of Novel Ultrasonic Monitoring Techniques for Improving the Reliability of Wind Turbine Gearboxes; Elsevier:
Amsterdam, The Netherlands, 2021.
289.
Joergensen, E.R.; Borum, K.K.; McGugan, M.; Thomsen, C.L.; Jensen, F.M.; Debel, C.P.; Soerensen, B.F. Full Scale Testing of Wind
Turbine Blade to Failure—Flapwise Loading; Risø National Laboratory: Roskilde, Denmark, 2004.
290.
Jasi
¯
unien
˙
e, E.; Raišutis, R.; Šliteris, R.; Voleišis, A.; Jakas, M. Ultrasonic NDT of wind turbine blades using contact pulse-echo
immersion testing with moving water container. Ultragarsas 2008,63, 28–32.
291.
Raišutis, R.; Jasi
¯
unien
˙
e, E.; Žukauskas, E. Ultrasonic NDT of wind turbine blades using guided waves. Ultragarsas/Ultrasound
2008,63, 7–11.
292.
Jüngert, A.; Damage Detection in Wind Turbine Blades Using Two Different Acoustic Techniques. NDT Database J. NDT
2008. Available online: https://www.researchgate.net/profile/Anne-Juengert/publication/228576960_Damage_Detection_
in_wind_turbine_blades_using_two_different_acoustic_techniques/links/02e7e538da98b7fae9000000/Damage-Detection-in-
wind-turbine-blades-using-two-different-acoustic-techniques.pdf (accessed on 10 February 2022).
Sensors 2022,22, 1627 51 of 52
293.
Jüngert, A.; Grosse, C.U. Inspection techniques for wind turbine blades using ultrasound and sound waves. In Proceedings of the
Non-Destructive Testing in Civil Engineering, Nantes, France, 30 June–3 July 2009.
294.
Jasinien, E.; Raiutis, R.; Literis, R.; Voleiis, A.; Vladiauskas, A.; Mitchard, D.; Amos, M. NDT of wind turbine blades using adapted
ultrasonic and radiographic techniques. Insight-Non-Destr. Test. Cond. Monit. 2009,51, 477–483. [CrossRef]
295.
Lee, J.-R.; Shin, H.-J.; Chia, C.C.; Dhital, D.; Yoon, D.-J.; Huh, Y.-H. Long distance laser ultrasonic propagation imaging system for
damage visualization. Opt. Lasers Eng. 2011,49, 1361–1371. [CrossRef]
296.
Park, B.; Sohn, H.; Yeum, C.-M.; Truong, T.C. Laser ultrasonic imaging and damage detection for a rotating structure. Struct.
Health Monit. 2013,12, 494–506. [CrossRef]
297.
Ye, G.; Neal, B.; Boot, A.; Kappatos, V.; Selcuk, C.; Gan, T.-H. Development of an ultrasonic NDT system for automated in-situ
inspection of wind turbine blades. In Proceeding of the EWSHM—7th European Workshop on Structural Health Monitoring,
Nantes, France, 8–11 July 2014.
298.
Park, B.; An, Y.-K.; Sohn, H. Visualization of hidden delamination and debonding in composites through noncontact laser
ultrasonic scanning. Compos. Sci. Technol. 2014,100, 10–18. [CrossRef]
299.
Park, B.; Sohn, H.; Malinowski, P.; Ostachowicz, W. Delamination localization in wind turbine blades based on adaptive
time-of-flight analysis of noncontact laser ultrasonic signals. Nondestruct. Test. Eval. 2017,32, 1–20. [CrossRef]
300.
Marquez, F.P.G.; Muñoz, C.Q.G. A new approach for fault detection, location and diagnosis by ultrasonic testing. Energies
2020
,
13, 1192. [CrossRef]
301.
Zhu, J.; Yoon, J.M.; He, D.; Qu, Y.; Bechhoefer, E. Lubrication oil condition monitoring and remaining useful life prediction with
particle filtering. Int. J. Progn. Health Manag. 2013,4, 124–138. [CrossRef]
302.
Myshkin, N.; Markova, L.; Semenyuk, M.; Kong, H.; Han, H.-G.; Yoon, E.-S. Wear monitoring based on the analysis of lubricant
contamination by optical ferroanalyzer. Wear 2003,255, 1270–1275. [CrossRef]
303.
Dupuis, R. Application of oil debris monitoring for wind turbine gearbox prognostics and health management. In Proceedings of
the Annual Conference of the PHM Society, Portland, OR, USA, 10–16 October 2010. [CrossRef]
304.
Hamilton, A.; Quail, F.J. Detailed state of the art review for the different online/inline oil analysis techniques in context of wind
turbine gearboxes. J. Tribol. 2011,133, 044001. [CrossRef]
305.
Agoston, A.; Ötsch, C.; Jakoby, B. Viscosity sensors for engine oil condition monitoring—Application and interpretation of results.
Sens. Actuators A Phys. 2005,121, 327–332. [CrossRef]
306.
Mignani, A.G.; Ciaccheri, L.; Díaz-Herrera, N.; Mencaglia, A.A.; Ottevaere, H.; Thienpont, H.; Francalanci, S.; Paccagnini, A.;
Pavone, F.S. Optical fiber spectroscopy for measuring quality indicators of lubricant oils. Meas. Sci. Technol.
2009
,20, 034011.
[CrossRef]
307.
Coronado, D.; Kupferschmidt, C. Assessment and validation of oil sensor systems for on-line oil condition monitoring of wind
turbine gearboxes. Procedia Technol. 2014,15, 747–754. [CrossRef]
308.
Zhu, J.; Yoon, J.M.; He, D.; Bechhoefer, E. Online particle-contaminated lubrication oil condition monitoring and remaining useful
life prediction for wind turbines. Wind Energy 2015,18, 1131–1149. [CrossRef]
309.
Sheng, S. Monitoring of wind turbine gearbox condition through oil and wear debris analysis: A full-scale testing perspective.
Tribol. Trans. 2016,59, 149–162. [CrossRef]
310.
Musial, W.; McNiff, B. Wind turbine testing in the NREL dynamometer test bed. In Proceedings of the AWEA’s Wind, Power 2000
Conference, Palm Springs, CA, USA, 2 May 2000.
311.
Lading, L.; Mcgugan, M.; Sendrup, P.; Rheinlander, J.; Rusborg, J. Fundamentals for Remote Structural Health Monitoring of Wind
Turbine Blades-a Pre-Project. Annex B-Sensors and Non-Destructive Testing Methods for Damage Detection in Wind Turbine Blades; Risø
National Laboratory: Roskilde, Denmark, 2002.
312. Kolle, J.J.; Kolle, J.J. Transverse optical fuse for composite damage detection. SAMPE Q. 1993,24, 35–40.
313.
Verijenko, B.; Verijenko, V. A new structural health monitoring system for composite laminates. Compos. Struct.
2005
,71, 315–319.
[CrossRef]
314.
Lara-Quintanilla, A.; Hulskamp, A.W.; Bersee, H.E. A high-rate shape memory alloy actuator for aerodynamic load control on
wind turbines. J. Intell. Mater. Syst. Struct. 2013,24, 1834–1845. [CrossRef]
315.
Papadopoulos, K.; Morfiadakis, E.; Philippidis, T.P.; Lekou, D.J. Assessment of the Strain Gauge Technique for Measurement of
Wind Turbine Blade Loads. Wind Energy Int. J. Prog. Appl. Wind Power Convers. Technol. 2000,3, 35–65. [CrossRef]
316.
Kim, S.-W.; Kim, E.-H.; Rim, M.-S.; Shrestha, P.; Lee, I.; Kwon, I.-B. Structural performance tests of down scaled composite wind
turbine blade using embedded fiber bragg grating sensors. Int. J. Aeronaut. Space Sci. 2011,12, 346–353. [CrossRef]
317.
Dimopoulos, C.; Gantes, C. Experimental investigation of buckling of wind turbine tower cylindrical shells with opening and
stiffening under bending. Thin-Walled Struct. 2012,54, 140–155. [CrossRef]
318.
Choi, K.-S.; Huh, Y.-H.; Kwon, I.-B.; Yoon, D.-J. A tip deflection calculation method for a wind turbine blade using temperature
compensated FBG sensors. Smart Mater. Struct. 2012,21, 25008. [CrossRef]
319.
Kim, S.-W.; Kang, W.-R.; Jeong, M.-S.; Lee, I.; Kwon, I.-B. Deflection estimation of a wind turbine blade using FBG sensors
embedded in the blade bonding line. Smart Mater. Struct. 2013,22, 125004. [CrossRef]
320.
Sierra-Pérez, J.; Torres-Arredondo, M.A.; Güemes, A. Damage and nonlinearities detection in wind turbine blades based on strain
field pattern recognition. FBGs, OBR and strain gauges comparison. Compos. Struct. 2016,135, 156–166. [CrossRef]
Sensors 2022,22, 1627 52 of 52
321.
Helsen, J.; Guo, Y.; Keller, J.; Guillaume, P. Experimental investigation of bearing slip in a wind turbine gearbox during a transient
grid loss event. Wind Energy 2016,19, 2255–2269. [CrossRef]
322.
Aizawa, K.; Poozesh, P.; Niezrecki, C.; Baqersad, J.; Inalpolat, M.; Heilmann, G. An acoustic-array based structural health
monitoring technique for wind turbine blades. In Structural Health Monitoring and Inspection of Advanced Materials, Aerospace, and
Civil Infrastructure 2015; SPIE: Bellingham, WA, USA, 2015; pp. 450–466.
323.
Poozesh, P.; Aizawa, K.; Niezrecki, C.; Baqersad, J.; Inalpolat, M.; Heilmann, G. Structural health monitoring of wind turbine
blades using acoustic microphone array. Struct. Health Monit. 2017,16, 471–485. [CrossRef]
324.
Zhao, H.; Chen, G.; Hong, H.; Zhu, X. Remote structural health monitoring for industrial wind turbines using short-range doppler
radar. IEEE Trans. Instrum. Meas. 2021,70, 1–9. [CrossRef]
325.
Hu, C.; Albertani, R.; Suryan, R.M. Wind turbine sensor array for monitoring avian and bat collisions. Wind Energy
2018
,21,
255–263. [CrossRef]
326.
Lu, B.; Li, Y.; Wu, X.; Yang, Z. A review of recent advances in wind turbine condition monitoring and fault diagnosis. In
Proceedings of the 2009 IEEE Power Electronics and Machines in Wind Applications, Lincoln, NE, USA, 24–26 June 2009; pp. 1–7.
327.
Qiao, W.; Lu, D. A survey on wind turbine condition monitoring and fault diagnosis—Part II: Signals and signal processing
methods. IEEE Trans. Ind. Electron. 2015,62, 6546–6557. [CrossRef]
328.
Ren, Z.; Verma, A.S.; Li, Y.; Teuwen, J.J.; Jiang, Z. Offshore wind turbine operations and maintenance: A state-of-the-art review.
Renew. Sustain. Energy Rev. 2021,144, 110886. [CrossRef]
329.
Yang, W.; Tavner, P.J.; Crabtree, C.; Wilkinson, M. Cost-effective condition monitoring for wind turbines. IEEE Trans. Ind. Electron.
2010,57, 263–271. [CrossRef]
330.
Yang, W.; Tavner, P.J.; Crabtree, C.; Feng, Y.; Qiu, Y. Wind turbine condition monitoring: Technical and commercial challenges.
Wind Energy 2012,17, 673–693. [CrossRef]
331.
Lian, J.; Cai, O.; Dong, X.; Jiang, Q.; Zhao, Y. Health monitoring and safety evaluation of the offshore wind turbine structure: A
Review and discussion of future development. Sustainability 2019,11, 494. [CrossRef]
... The displacement monitoring technology for monopiles can be divided into noncontact and contact monitoring [6,7]. Noncontact displacement sensors, such as laser telemeters and tachymeters [8], can achieve high-frequency and accurate displacement monitoring of cer-tain points in a monopile. However, these sensors are limited in their ability to capture displacement data of the submerged portion of a monopile. ...
... In addition, because the UWFBG is affected by both temperature and strain, the strains on the two opposite sides, H 1 and H 2 (Fig. 1), can be expressed as Eq. (8). H 1 is on the side opposite to the main wind direction, whereas H 2 is on the side of the main wind direction. ...
... Condition monitoring (CM) techniques, if present, have fewer and less sophisticated sensors for remote monitoring of performance and maintenance alerts. Generally, rooftop wind turbines are smaller, simpler, and designed for lower power output with a focus on integration with building structures and minimizing their vibrations [80][81][82]. Nevertheless, mechanical imbalances and component misalignment are the primary sources of vibration in smaller wind turbines. ...
... At the same time, seismometers measure ground vibrations to assess the impact on the building structure hosting the rooftop turbine. Civera and Surace [82] presented a literature review of the last 20 years of using nondestructive techniques for structural health monitoring of wind turbines. Non-destructive techniques have advanced significantly in condition and structural health monitoring of wind turbines over the last 20 years, offering practical solutions for both mechanical and civil engineering fields. ...
Article
The recent tendencies of drastic variations of environmental conditions, inundations, severe winds, and gusts besides heavy and unexpected long-duration rains showed the world an image of what could happen if emissions and global warming are not adequately controlled. Implementing new energy solutions as fast as possible is essential to cope with climate change, one of the biggest threats to our survival. These alarming signs intensified research and development efforts to replace fossil-based activities with renewable, eco-friendly energy resources to ensure reduced emissions and global warming effects. Wind energy stands out as one of the favorable renewable energy sources for decarbonization and energy transition of buildings and urban areas. The installation of small and medium-sized windmills in urban areas and on roofs of buildings attends the recommendations for sustainable energy transition, and the UN Sustainable Development Goals, SDG 7, which ensures equity and justice to affordable and reliable clean energy. Installation of windmills of small and medium capacities in urban areas and buildings roofs faces significant challenges, including noise, mechanical vibration, efficiency, and mechanical security necessary for healthy ambient and social acceptance by the population. This review provides a general evaluation of the design and performance of moderate capacity horizontal and vertical axis wind turbines, geometry optimization, current installations on rooftops of buildings and urban districts, studies on horizontal axis and vertical axis wind turbines for rooftops of buildings and metropolitan areas, vibration, noise, and aerodynamic induced mechanical forces, end of life of wind turbine and blades waste recycling and reusing methods. Finally, the review provides conclusions based on the findings from the review and future research and development prospects.
... Due to various adverse factors-such as exposure to harsh environmental conditions [12], aerodynamic and hydrodynamic forces, and prolonged use-these elements can suffer different types of damage that impact their structural integrity and performance. Therefore, non-destructive testing (NDT), including ultrasonic testing [13] and infrared thermography inspection [14], is proposed for defect identification. Additionally, the Laser-Induced Breakdown Spectroscopy (LIBS) technique is recommended for evaluating the EoL matrix compounds. ...
Article
Full-text available
Citation: Rodriguez-Ortiz, L.A.; Suárez, E.A.T.; Muíños-Landín, S.; Boga, D.C.; Araújo, A.; Carneiro, C.; Santos, R.M.; Johst, P.; Böhm, R. Abstract: Research on sustainable business models (BMs) rooted in the circular economy is expanding in industry and academia, encouraging companies to enhance their impact on profits, people, and the planet. However, developing sustainable BMs is complex due to the conflicting objectives of sustainability dimensions and competition with traditional models. This work addresses this challenge by proposing a Sustainable Business Guide (SBG) supported by artificial intelligence (AI) to assist in decision-making. The SBG supports the design and assessment of business models by integrating inspection; business opportunity exploration; technical, economic, and environmental analyses; an AI-based decision support system; and Porter's Five Forces, focusing on the aeronautical and wind energy sectors.
... This chapter has presented the analysis of data extracted from selected SHM studies (Sofi et al. 2022). Thematic analysis identified key themes and trends, while comparative analysis offered insights into the advantages and drawbacks of various methodologies (Civera and Surace, 2022). Gap analysis highlighted areas requiring further investigation, guiding future research and development efforts in SHM (Wu et al. 2003) This chapter discusses the findings from the data analysis, highlighting the implications of the results and their relevance to the field of SHM (Martinez et al. 2016). ...
Article
Full-text available
Structural Health Monitoring (SHM) protects and extends the structure's lifespan. This review consolidates recent developments in SHM technologies and methods, emphasizing key advancements while pinpointing ongoing research challenges. We examine more than 130 studies, highlighting advancements in sensor technologies, including vibration sensors, fiber optics, piezoelectric transducers, and wireless networks. These technologies have been effectively applied across aerospace, civil, and mechanical engineering domains. Despite significant progress, challenges persist in integrating machine learning (M.L) and deep learning (D.L) for improved Structural assessment, managing large datasets, validating sensor reliability, and mitigating the effects of environmental factors on sensor performance. Although advancements have been made, there remains a significant gap in research focusing on SHM for timber structures, heritage buildings, and agricultural infrastructure. Additionally, optimizing sensor placement and developing robust multi-sensor approaches are critical areas needing further exploration. Standardizing SHM protocols and incorporating emerging technologies into comprehensive frameworks are essential for advancing the field. This analysis offers a comprehensive overview of the current status of SHM research, pinpoints crucial shortcomings, and proposes avenues for further investigation. Filling these voids will enable the creation of more efficient and dependable Structural Health Monitoring (SHM) systems, thereby enhancing the safety and durability of infrastructure globally.
... The blade root load (BRL) reflects the operating state of the wind turbine and is very important to the safety of blades [5]. But there are many challenges in the monitoring process of BRL. ...
Article
Full-text available
The monitoring of wind turbine blade root load (BRL) has always been a valuable yet highly challenging problem. BRL monitoring is often measured directly using strain sensors, which are expensive to install and poses safety risks. Thus, it is necessary to explore an indirect measurement method that establishes a mapping relationship between BRL and easily measurable data. This paper presents a non‐intrusive BRL monitoring method based on the vibrational signals at bearings (including main bearing, gearbox bearing, and generator bearing) and a new proposed AE‐LSTM‐Attention neural network. The proposed neural network incorporates an autoencoder (AE) pre‐training process, a long short‐term memory (LSTM) neural network, and a channel attention mechanism to overcome difficulties in extracting and fusing long‐term, high‐frequency, and multichannel signal features. Additionally, considering the wind turbine transmission mechanism, an optimal LSTM step number selection criterion is proposed. The wind turbine's physical features are innovatively bridged with the neural network training parameters, e.g., setting the transmission ratio of the gearbox as the number of LSTM steps, by which the most cost‐effective prediction performance can be achieved. Experiments were conducted using data from actual operating wind turbines, and the results show that the proposed method can successfully achieve non‐intrusive monitoring of BRL characteristics and outperform traditional methods in terms of predictive performance. The mapping relationship between wind turbine vibration signals and BRL characteristics can be established efficiently and accurately.
... There are often two main categories around which structure inspections are carried out: (a) nondestructive testing (NDT), and (b) vibration-based structural health monitoring (SHM). NDT inspection methods, such as visual inspections, 6 optical inspection, 7-9 vibrothermography, 10 digital image correlation, 11 thermal inspections, ultrasonic inspections, [12][13][14] are widely employed but can be visually subjective, time-consuming, limited to specific areas, 15 and uncertainty in interpretation can also be high. Consequently, effective use of NDT can be quite challenging for wind turbine composite blades, where damages can occur widely, crack growth may be unpredictable, and the detection requirements are linked to their specific operational aspects. ...
Article
Full-text available
There exists several methods for feature extraction from monitored infrastructure data. However, for wind turbine monitoring, there exists a dearth in the analysis of full-scale tests for damage detection, guidance for feature extraction, and expected performance levels. This paper attempts to address this through a benchmark study of feature selection and reduction methods, and their performance levels using a full-scale experimental study. Optimum feature selection is achieved through vibration-based damage detection of wind turbine composite blades using acceleration measurements. Experimental measurements of accelerations at 11 different locations on an operating Vestas V27 wind turbine is used. Variable operating conditions for three different rotor speeds corresponding to one healthy, three damaged, and one repaired condition were analyzed. Feature selection and related benchmarking identified the minimal feature set with the best capability to distinguish between multiple damage states. Time, frequency, and time–frequency features were extracted from the accelerations for feature estimation. Feature selection performance was tested via seven machine learning approaches using impulse and ambient responses. Results demonstrate how high classification accuracies of up to 95% can be achieved with only top 10 time–frequency features selected from the analysis of postimpact impulse responses. However, a larger set of optimal features between 30 and 50 are required for effective damage detection with ambient responses. The work, for the first time, presents such feature selection and classification performance for real-time wind turbine blade damage and as a function of optimum selected features. It also demonstrates how a reduced set of features is useful in distinguishing between damage states while a larger feature set may not be useful or even lead to reduced accuracy. The work provides a much needed performance benchmark for structural health monitoring practitioners and researchers for choice, comparison, validation, and use of features for wind turbines.
... Zhang et al. [22] found that the accuracy of VRSs in industrial environments exhibited an error margin of only ±0.3%. Civera and Surace [23] demonstrated that environmental factors, such as high temperatures and vibrations, impact sensor accuracy in wind turbines but can be mitigated through additional mechanical protection. While these findings indicate promising performance, further research is needed to address challenges such as measurement variability under more complex conditions. ...
Article
Full-text available
Accurate monitoring of turbine speed is essential for ensuring operational stability and efficiency inpower generation systems, particularly within the context of low-carbon and renewable energy integration. Thisstudy evaluates the performance of three Variable Reluctance Sensors (VRSs)—VRS1, VRS2, and VRS3—used forreal-time speed monitoring of the Steam Turbine Generator (STG) 1.0 at the Tambak Lorok Combined Cycle PowerPlant (CCPP). The evaluation was conducted using statistical methods, including Root Mean Square Error (RMSE),standard deviation, and two-factor Analysis of Variance (ANOVA) without replication, to assess the accuracy andconsistency of the sensors under varying operational conditions. The operational conditions were simulated througha motor controlled by a Variable Speed Drive (VSD), which allows for precise control over speed variations. Theresults indicate that the VRSs exhibit high accuracy and reliability, with RMSE values ranging from 0.08% to0.28%. Among the three sensors, VRS3 demonstrated the highest performance, achieving minimal variability, witha standard deviation of 0.000 at a frequency of 50.00 Hz. ANOVA revealed no significant differences in performancebetween the three sensors (P-value = 1.000), suggesting uniformity in their measurement capabilities. These findingssubstantiate the suitability of VRSs for turbine speed monitoring in power plants, ensuring operational stability andsupporting the integration of renewable energy technologies. The results reinforce the potential of VRSs as a reliabletool for improving the efficiency of sustainable energy systems Keywords: Variable Reluctance Sensor (VRS); Standard deviation; Root Mean Square Error (RMSE); TambakLorok Combined Cycle Power Plant (CCPP)
Article
Robots are used to conduct non-destructive defect detection on wind turbine blades (WTBs) and to monitor their integrity over time. However, current inspection robots are often bulky and heavy, and struggle to detect defects in the blade’s main beam, thus presenting difficulties in portability and effectiveness. To address these issues, we designed a wheel-wing composite robot equipped with a curved surface-adaptive phased array ultrasonic detection device for the detection of defects in the WTB’s main beam. We determined the pose equation under different section characteristics and identified the robot’s stable range of motion, thus developing a model of its kinematics. A detection device adapted for variable curvature surfaces was designed to ensure tight coupling between the robot’s probe and the blade. Additionally, element differential and least-square ellipse-fitting methods were employed to analyze blades with irregular sections. The simulation results demonstrated that the prototype can stably traverse an area with a vertical angle of ±14.06° at a speed of 0.25 m/s, fully covering the main beam area of the blade during walking operations. Moreover, the robot can scan the main beam area at a speed of 0.10 m/s, enabling the accurate detection of defects.
Article
Full-text available
Offshore wind turbines (OWTs) are subject to waves, currents, seabed shifts, and corrosion in harsh marine environment, posing significant challenges to structural integrity and durability. A prevalent issue of OWTs is the scour of foundations and cables, which can cause structural damage, interrupting the operation of OWTs and even leading to catastrophic failure. Timely detection and assessment of scour are crucial for maintaining the structural integrity and managing the operation of OWTs. This paper reviews the main methods utilized to assess scour for OWT. Three main types of scour assessment methods are reviewed, which are direct methods, vibration-based methods, and unmanned vehicle-based methods. For each type of methods, the reviewed contents mainly include the assessment principles, laboratory tests, and field applications, as well as data processing and interpretation. The limitations of existing methods and the new opportunities are discussed. This research promotes improvement of the monitoring and maintenance of OWTs.
Article
Full-text available
The maintenance of wind turbines is of growing importance considering the transition to renewable energy. This paper presents a multi-robot-approach for automated wind turbine maintenance including a novel climbing robot. Currently, wind turbine maintenance remains a manual task, which is monotonous, dangerous, and also physically demanding due to the large scale of wind turbines. Technical climbers are required to work at significant heights, even in bad weather conditions. Furthermore, a skilled labor force with sufficient knowledge in repairing fiber composite material is rare. Autonomous mobile systems enable the digitization of the maintenance process. They can be designed for weather-independent operations. This work contributes to the development and experimental validation of a maintenance system consisting of multiple robotic platforms for a variety of tasks, such as wind turbine tower and rotor blade service. In this work, multicopters with vision and LiDAR sensors for global inspection are used to guide slower climbing robots. Light-weight magnetic climbers with surface contact were used to analyze structure parts with non-destructive inspection methods and to locally repair smaller defects. Localization was enabled by adapting odometry for conical-shaped surfaces considering additional navigation sensors. Magnets were suitable for steel towers to clamp onto the surface. A friction-based climbing ring robot (SMART— Scanning, Monitoring, Analyzing, Repair and Transportation) completed the set-up for higher payload. The maintenance period could be extended by using weather-proofed maintenance robots. The multi-robot-system was running the Robot Operating System (ROS). Additionally, first steps towards machine learning would enable maintenance staff to use pattern classification for fault diagnosis in order to operate safely from the ground in the future.
Article
Full-text available
Shearography, as a novel non-destructive evaluation technique, has shown notable ability in the detection of composite materials. However, in current shearography practices, the phase shifting and loading methods applied are mainly static. For instance, vacuum hood or force loading facilities are often used in phase-shifting shearography, and these are hard to realise with robotic control, especially for on-board inspection. In this study, a dynamic process for detecting defects in the subsurface of a wind turbine blade (WTB) using spatial phase shift with dynamic thermal loading was proposed. The WTB sample underwent a dynamic thermal loading operation, and its status is captured by a Michelson interferometric-based spatial phase shift digital shearography system using a pixelated micro-polarisation array sensor. The captured images were analysed in a 2D frequency domain and low-pass filtered for phase map acquisition. The initial phase maps underwent a window Fourier filtering process and were integrated to produce a video sequence for realisation of visualising the first derivative of the displacement in the process of thermal loading. The approach was tested in experimental settings and the results obtained were presented and discussed. A comparative assessment of the approach with shearography fringe pattern analysis and temporal phase shift technique is also presented and discussed.
Article
Full-text available
Thermographic fault diagnosis of ventilation in BLDC (brushless DC) motors is described. The following states of BLDC motors were analyzed: a healthy BLDC motor running at 1450 rpm, a healthy BLDC motor at 2100 rpm, blocked ventilation of the BLDC motor at 1450 rpm, blocked ventilation of the BLDC motor at 2100 rpm, healthy clipper, and blocked ventilation of the clipper. A feature extraction method called the Common Part of Arithmetic Mean of Thermographic Images (CPoAMoTI) was proposed. Test thermal images were analyzed successfully. The developed method, CPoAMoTI is useful for industry and society. Electric cars, trains, fans, clippers, computers, cordless power tools can be diagnosed using the developed method.
Article
Full-text available
In this study, Computer Vision and Phase-Based Motion Magnification (PBMM) are validated for continuous Structural Health Monitoring (SHM) purposes. The aim is to identify the exact instant of occurrence for damage or abrupt structural changes from video-extracted, very low amplitude (barely visible) vibrations. The study presents three experimental datasets: a box beam with multiple saw cuts of different lengths and angles, a beam with a full rectangular cross section and a mass added at the tip, and the spar of a prototype High-Aspect-Ratio wing. Both mode-shape- and frequency-based approaches are considered, showing the potential to identify the severity and position of the damage as well A high-definition, high-speed camera and a low-cost commercial alternative have been successfully utilised for these video acquisitions. Finally, the technique is also preliminarily tested for outdoor applications with smartphone cameras.
Article
Full-text available
Near- and mid-infrared optical coherence tomography (OCT) is evaluated as a non-destructive and non-contact reflection imaging modality for inspection of industrial and marine coatings. Near-infrared OCT was used to obtain high-resolution images (~6/2 µm lateral/axial) of hidden subsurface cracks and defects in a resin base coating, which had been exposed to high pressure and high temperature to study coating degradation in hostile environments. Mid-infrared OCT was employed for high-resolution (~15/8.5 µm lateral/axial) subsurface inspection of highly scattering marine coatings, demonstrating monitoring of wet film thickness and particle dispersion during curing of a 210 µm layer of antifouling coating, and detection of substrate corrosion through 369 µm of high-gloss alkyd enamel. Combining high-resolution and fast, non-invasive scanning, OCT is therefore considered a promising tool for studying coating performance and for industrial inspection.
Article
Full-text available
The concept and preliminary design of an unmanned lighter-than-air (LTA) platform instrumented with different remote sensing technologies is presented. The aim is to assess the feasibility of using a remotely controlled airship for the land monitoring of medium sized (up to 107 m2) urban or rural areas at relatively low altitudes (below 1000 m) and its potential convenience with respect to other standard remote and in-situ sensing systems. The proposal includes equipment for high-definition visual, thermal, and hyperspectral imaging as well as LiDAR scanning. The data collected from these different sources can be then combined to obtain geo-referenced products such as land use land cover (LULC), soil water content (SWC), land surface temperature (LSC), and leaf area index (LAI) maps, among others. The potential uses for diffuse structural health monitoring over built-up areas are discussed as well. Several mission typologies are considered.
Chapter
The RADBLAD project aims to demonstrate a step change in the accurate and fast in-situ non-destructive testing of wind turbine blades located off-shore using remote robotic deployment of X-ray radiography at great heights to image the full thickness of a blade. This paper describes the development of the robot system that deploys an X-ray source and detector around a blade by climbing up a turbine tower. Robot design is concentrated on developing a lightweight modular robot system easily transported to test sites and assembled quickly but capable of carrying a large payload. Four tower climbing crawlers drive a ring structure (constructed around the tower) up a wind turbine tower to place an X-ray source and digital detector on either side of a blade with long arms constituted by a bridge structure. The system is actuated by fourteen electric motors, six pneumatic actuators on the ring structure and sixteen pneumatic actuators on the crawlers. The focus of the paper is on describing the remote control of these actuators via a CANopen bus and a DC power bus that aims to minimize the size of umbilical power and communication cables.
Article
Wind-generated electricity has increased continuously for the last decades and will continue to progress for the next years. Wind energy production is a mature technology with over 600GW installed capacity in the world. The operation of wind farms presents critical challenges to maintain the lowest possible cost of energy. It is essential to detect early or predict wind turbine breakdowns due to different factors such as material degradation, electrical or mechanical failure, faults or damage and others. Wind turbine blades constitute the costliest and most exposed parts and suffer from many shortcomings, mainly cracks and erosion, reducing their performance. Hence, there is an essential requirement for using non-destructive diagnostic of wind turbine blades. This paper reviews the current non-destructive techniques for wind turbine blades analysis, their applicability, advantages and drawbacks: Acoustic Emission Testing, Infrared, Fiber Grating, Ultrasonic, Guided Wave, Thermal Imaging, Terahertz Imaging, Visual Inspection, Tap Test, Electromagnetic Testing and Vibration Analysis.Nevertheless, these methods face drawbacks that can be overcome by remote sensing. Hyperspectral imaging is a spectral imaging technique that integrates imaging and spectroscopy, enables the analysis and identification of distinctive spectral signatures, and assigns them to the examined sample elements. Thus, this paper describes hyperspectral imaging's implementation in image acquisition, handling, and flaw recognition and detection of cracks and erosion. This technique's field output results show that blade defect detection's accuracy and precision are significantly enhanced.
Article
Operations and maintenance of offshore wind turbines (OWTs) play an important role in the development of offshore wind farms. Compared with operations, maintenance is a critical element in the levelized cost of energy, given the practical constraints imposed by offshore operations and the relatively high costs. The effects of maintenance on the life cycle of an offshore wind farm are highly complex and uncertain. The selection of maintenance strategies influences the overall efficiency, profit margin, safety, and sustainability of offshore wind farms. For an offshore wind project, after a maintenance strategy is selected, schedule planning will be considered, which is an optimization problem. Onsite maintenance will involve complex marine operations whose efficiency and safety depend on practical factors. Moreover, negative environmental impacts due to offshore maintenance deserve attention. To address these issues, this paper reviews the state-of-the-art research on OWT maintenance, covering strategy selection, schedule optimization, onsite operations, repair, assessment criteria, recycling, and environmental concerns. Many methods are summarized and compared. Limitations in the research and shortcomings in industrial development of OWT operations and maintenance are described. Finally, promising areas are identified with regard to future studies of maintenance strategies.