You are on page 1of 86

DOCTORA L T H E S I S

Improving detection and diagnosis


of bearing failures in wind
turbine drivetrains

Daniel Strömbergsson

Machine Elements
Improving detection and diagnosis of bearing
failures in wind turbine drivetrains

Daniel Strömbergsson

Luleå University of Technology


Department of Engineering Science and Mathematics,
Division of Machine Elements
Cover figure: Illustration of the frequency spectrum of a vibration measure-
ment from a wind turbine with a failing gearbox bearing (Turbine image cour-
tesy of Nordex. Bearing damage photo courtesy of Henrik Renberg, Skellefteå
Kraft)

Improving detection and diagnosis of bearing failures in wind turbine


drivetrains

Copyright © Daniel Strömbergsson (2020). This document is freely available


at:

www.ltu.se
or by contacting Daniel Strömbergsson,

daniel.strombergsson@ltu.se
This document may be freely distributed in its original form including the cur-
rent author’s name. None of the content may be changed or excluded without
permissions from the author.

This document was typeset in LATEX 2ε

Printed by Lenanders Grafiska AB, Kalmar 2020

ISSN: 1402-1544
ISBN: 978-91-7790-621-6 (print)
ISBN: 978-91-7790-622-3 (pdf)

Luleå 2020

www.ltu.se
Abstract

Wind power has in the last 20 years grown into one of the main sources of
renewable energy in the world, with both the amount and size of the turbines
increasing substantially. One of the major challenges for the wind power in-
dustry is the premature failures of especially the drivetrain components. These
failures cause a lot of turbine downtime, which increases the operation and
maintenance costs of the turbines. Failures in the gearbox have been shown to
lead to the highest downtime and the multitude of bearings within that sub-
system is overrepresented in the total amount of component failures. Vibration-
based condition monitoring is considered the best method to find these types
of defects early and avoid prolonged turbine downtime. Previous research has
therefore been focused on the different aspects of condition monitoring; i.e.
measurement technologies, signal analysis of vibration measurements to im-
prove detection and diagnosis as well as the implementation of machine learn-
ing solutions. However, the majority of research work has yet to evaluate the
performance of new developments using wind turbine field data, and still no
fundamentally new developments have seen a large-scale implementation in the
industry. Further, it is known that the positioning of the accelerometer, used
to measure the vibrations, affects the ability to detect and diagnose defects.
However, it is not known how to optimally position the accelerometers to mon-
itor the individual drivetrain sub-systems. Also, previous research does not
show how the influence of the measurement properties of the field data affect
the ability to detect and diagnose component failures.
Therefore, this thesis provides a comprehensive evaluation of how to im-
prove the detection and diagnosis of bearing failures specifically in wind turbine
drivetrains. In this thesis, a simulation model was developed to study how the
accelerometer positioning affects the measurement quality. Vibration simula-
tions of a similar sized bearing to ones found in the wind turbine drivetrain
show an optimal accelerometer position as close to the primary loaded zone
of the bearings as possible. The current placement of the accelerometers in
the wind turbine drivetrain are often diametrically opposed to the loaded zone,
and the performance of the vibration monitoring with respect to the commonly
used signal analysis tools could thereby be increased. Further, wavelet-based
signal analysis has been evaluated using historical wind turbine drivetrain field
data. A new implementation of the wavelet packet transform to analyse en-

iii
veloped vibration measurements in the frequency domain was developed, where
the measurements were decomposed into packets matching the frequency reso-
lution of the fast Fourier transform, and analysing the packet energy spectra.
Finally, an anomaly detection solution utilizing an artificial neural network has
been implemented to separate the condition indicator values, used for detec-
tion and diagnosis, from their inherent variance due to the dynamic turbine
operation seen in the drivetrain rotational speed.
The results in this thesis show the inadequacy of the commonly stored vi-
bration measurements to the condition monitoring databases when used in post
failure investigations and application of research developments on available field
data. Measurements both taken over a long period of time and covering wide
frequency range should be stored, instead of the either/or of today. Otherwise,
the real-time monitoring of wind turbine drivetrain bearing failures cannot be
replicated and monitoring improvements not fully evaluated. By implement-
ing the wavelet packet transform, the detection and diagnosis performance was
increased 250% compared to the fast Fourier transform, in an example of gear-
box output shaft bearing failure. By implementing the anomaly detection by
the artificial neural network, the performance increased further and was able
to provide indications in a planet bearing failure case, which was not possible
before. For turbine owners, these results provide both practical actions to take
and provide an example of an easily implementable signal analysis tool to im-
prove the detection and diagnosis of drivetrain bearing failures. The anomaly
detection, which utilizes available historic data from healthy turbines and does
not require any amount of labelled data for all considered types of bearing fail-
ures, also shows promise to detect failures in the drivetrain components which
has been historically problematic. For the research community, the results
mainly provides guidance into using historic field data when evaluating new
developments. Also, they highlight potential pitfalls one can face using field
data and what data properties to look for to successfully show the potential of
your work.
Acknowledgements

The work within this doctoral thesis was carried out at the Division of Machine
Elements at Luleå University of Technology. I would first like to thank past
and present colleagues at the division for everyday creating an enjoyable and
fulfilling working environment. Going to work at the university, discussing
tribology, machine elements and whatever else has over five years newer been
even slightly difficult due to the people there. To Sergio, Fredrik, Juhamatti,
Stephan and the others at the SKF - University Technology Center, I need to
say thank you for your support and collaboration throughout the first years of
my Ph. D. studies.
From Skellefteå Kraft, Henrik Renberg deserves my gratitude for his support
throughout this whole project. Without his involvement I would not been able
to form the understanding of the wind turbine needed to perform the research
and arrive at the conclusions I did. At SKF, I’d like to thank Fabrice Drommi
and Lars-Erik Stacke for their support, Allan Thomson for our discussions and
collaboration during my time at the SKF in Livingston and above all to Per-
Erik Larsson at SKF in Luleå for his unending help, discussion and support
during the whole project.
To my supervisors Pär Marklund and Kim Berglund for introducing me
to the project, supporting and challenging me in ways for me to personally
improve without ever creating an even remotely bad working environment. I
am truly grateful for the work they put in over the last 5+ years to prepare
me to conduct my Ph. D. studies and for what’s ahead.
The most special thanks to my family, especially nieces and nephew for
allowing me to not care to take myself all to seriously and to be able to stop
being a grown up for a while.

Daniel Strömbergsson
Luleå, September 2020

v
Contents

I Comprehensive Summary 1
1 Introduction 3
1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Objectives of this thesis . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Thesis outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2 Wind turbines 7
2.1 The wind turbine drivetrain . . . . . . . . . . . . . . . . . . . . 7
2.2 Condition monitoring systems . . . . . . . . . . . . . . . . . . . 9
2.3 Drivetrain vibration monitoring . . . . . . . . . . . . . . . . . . 9
2.3.1 Other monitoring systems & techniques . . . . . . . . . 12

3 Bearing failure & vibration monitoring 13


3.1 Bearing failures found in the wind turbine drivetrains . . . . . 15
3.2 Vibration monitoring of bearing health . . . . . . . . . . . . . . 17
3.2.1 The enveloping technique . . . . . . . . . . . . . . . . . 17
3.2.2 Condition indicators of bearing health . . . . . . . . . . 18
3.3 Dynamic simulation of bearing vibrations . . . . . . . . . . . . 23

4 Vibration monitoring in research 29


4.1 Wavelet-based methods . . . . . . . . . . . . . . . . . . . . . . 29
4.1.1 Continuous & discrete wavelet transform . . . . . . . . . 30
4.1.2 Wavelet packet transform . . . . . . . . . . . . . . . . . 34
4.2 Influence of measurement properties . . . . . . . . . . . . . . . 38
4.3 Implementing a machine learning solution . . . . . . . . . . . . 45

5 Contributions 53
5.1 Paper A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
5.2 Paper B . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
5.3 Paper C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
5.4 Paper D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.5 Paper E . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

vii
6 Conclusions & Future Work 57
6.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
6.2 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

References 63

II Appended Papers 75
A Multi-body simulation of fault vibrations 77

B Mother wavelet selection in the DWT 95

C Bearing monitoring in the wind turbine drivetrain 121

D Wind turbine drivetrain condition monitoring Part A 147

E Wind turbine drivetrain condition monitoring Part B 183


List of appended papers

This doctoral thesis is composed of the following appended publications:

A D. Strömbergsson, P. Marklund, K. Berglund.


"Multi-body simulation and validation of fault vibrations from
rolling element bearings."
Submitted to Proc. IMechE, Part J: Journal of Engineering Tribology.

B D. Strömbergsson, P. Marklund, K. Berglund, J. Saari, A. Thom-


son.
"Mother wavelet selection in the discrete wavelet transform for
condition monitoring of wind turbine drivetrain bearings."
Wind Energy, 2019; vol. 22, Issue 11, p. 1581-1592.

C D. Strömbergsson, P. Marklund, K. Berglund, P.-E. Larsson.


"Bearing monitoring in the wind turbine drivetrain - A com-
parative study of the FFT and wavelet transforms."
Wind Energy, 2020; vol. 23, Issue 6, p. 1381-1393.

D D. Strömbergsson, P. Marklund, K. Berglund, P.-E. Larsson.


"Wind turbine drivetrain condition monitoring by vibration
analysis Part A: Property requirements of vibration measure-
ments."
Submitted to Wind Energy.

E D. Strömbergsson, P. Marklund, K. Berglund.


"Wind turbine drivetrain condition monitoring by vibration
analysis Part B: Anomaly detection using artificial neural net-
works."
Submitted to Wind Energy.

ix
Other publications
– J. Saari, D. Strömbergsson, J. Lundberg, A. Thomson.
"Detection and identification of windmill bearing faults using a
one-class support vector machine (SVM)."
Measurement, 2019; Vol. 137, p. 287-301.

– S. Martin del Campo Barraza, F. Sandin, D. Strömbergsson.


"A dictionary learning approach to monitoring of wind turbine
drivetrain bearings."
Submitted to International Journal of Computational Intelligence Sys-
tems.

– D. Strömbergsson, P. Marklund, E. Edin, F. Zeman.


"Acoustic emission monitoring of a mechanochemical surface
finishing process."
Tribology International, 2017 ; Vol. 112, p. 129-136.

– M. J. Gómez, P. Marklund, D. Strömbergsson, C. Castejón, J.


C. García-Prada.
"Analysis of vibration signals of drivetrain failures in wind tur-
bines for Condition Monitoring"
Accepted for publication in Experimental Techniques.

Conference contributions
– D. Strömbergsson, P. Marklund, K. Berglund.
"Bearing faults in the wind turbine drivetrain: Comparative
study of monitoring with FFT and the Discrete Wavelet Trans-
form."
World Tribology Congress, Beijing, China, September 17-22, 2017.

– D. Strömbergsson, P. Marklund, K. Berglund.


"Dynamic simulation of rolling element bearing vibrations us-
ing SKF BEAST and test rig validation."
22:nd International colloquium Tribology, Stuttgart, Germany, January
28-30, 2020.
Part I

Comprehensive Summary

1
Chapter 1

Introduction

1.1 Background
As an answer to the global demand for renewable energy sources, the wind
power sector has grown substantially in the last 20 years. At the turn of the
millennium, only 17.4 gigawatt (GW) of installed capacity existed which has
grown to a predicted 727.1GW in 2020 [1,2]. The cumulative installed capacity
of the time period 2000-2019 together with the predicted capacity in 2020 is
illustrated in Figure 1.1. In the EU, 15% of energy consumption was accounted
for by the wind power capacity [3]. An increase has also been seen in the size
of the wind turbines themselves and thereby the asset value of each producing
turbine.

GROWTH OF INSTALLED WIND POWER CAPACITY 2000 -2020

727,1

651
INSTALLED CAPACITY, [GW]

591
TOTAL CUMULATIVE

540
488
433
370
319
283
238
198
159
121
94
59 74
31 39 48
17,4 24

20 0 0 200 1 20 0 2 20 0 3 2 00 4 2005 20 06 20 07 200 8 200 9 20 10 2011 2 0 1 2 20 13 2 014 2 015 2016 2 0 17 2018 2019 2020

Figure 1.1: Growth of the total wind power capacity in the world between 2000-
2020, reproduced from [1, 2].

3
4 Chapter 1. Introduction

Premature failures of wind turbine drivetrains continue to plague the wind


power industry, where the operation and maintenance requirements has histori-
cally been reported to cover up to 20% of total wind energy generation costs [4].
More specifically, these costs are primarily found in the lost production over
the downtime of the turbine. Also, the drivetrain sub-system where failures
is associated with the longest downtime is the gearbox and the main critical
component leading to the highest amount of gearbox failures is rolling element
bearings [5]. Thereby, efforts in increasing the efficiency in the maintenance of
these sub-system and component failures is of high importance.
Condition-based maintenance has emerged as the main maintenance tool to
reduce the downtime of the turbine caused by a drivetrain failure. Here, the
health of the machine elements is continuously monitored, and an impending
failure can be detected at an early stage before the defect can cause extensive
damages to surrounding components [6]. Also, maintenance planning can start
at an early stage while the full lifetime of the component is utilized, compared
to predetermined maintenance intervals which can cause healthy components
to be discarded or corrective maintenance where the planning starts after the
complete failure of the component and allowing for the machine to stand still
for a long period of time. This has led to research into condition monitoring
emerging as a field unto itself, developing new methods and systems. Condition
monitoring research concerns several research fields such as sensor technology,
signal analysis, material science, tribology etc. and also connects these to create
novelty.
For rolling element bearings, vibration-based condition monitoring emerged
as the primary tool for continuous on-line monitoring [7]. This due to its ver-
satility to extract and separate information from multiple rotating component
sources such as bearings, gears, shafts and the structural sources. From a re-
search perspective, an abundance of tools for signal analysis of the vibration
measurements, such as the wavelet-based transforms [8–21], spectral kurto-
sis [22–27] etc., and automatic failure diagnosis from e.g. machine learning so-
lutions [28–34] have been created. However, cohesion of this development has
been lacking, and moving into field data studies has often been overlooked [35].
Thereby, issues surrounding non-stationarity of the operation, high amount of
structural noise, disturbance effects and non-ideal accelerometer placements
has not been taken into account when reporting the increase in failure detec-
tion and diagnosis improvements. The lack of detailed field data comparable
to the immensely popular Case Western Reserve University bearing vibration
data repository and benchmark study by Smith and Randall [36] used to rank
the developments made, is also a detriment for more field data studies be-
ing conducted. As an illustrative example, the review paper in wind turbine
bearing condition monitoring by de Azevedo in 2016 reported only 3 out of
38 relevant studies using field data instead of experimental data from scaled
down test rigs or idealized simulated measurement [37]. A more recent example
shows improvement with 19 out of 48 studies using field data by the review
by Liu in 2020 covering the same topic [38]. However, this development has
1.2. Objectives of this thesis 5

not been enough for a full-scale implementation of a newly developed signal


analysis, automatic detection and diagnosis solution to be implemented in the
wind turbine condition monitoring systems [39].

1.2 Objectives of this thesis


The goal of the research work summarized in this thesis is to comprehensively
improve detection and diagnosis of wind turbine drivetrain bearing failures.
Specifically to show the potential in research developed methods of signal analy-
sis and automatic detection when moving from a laboratory or purely simulated
environment into using wind turbine drivetrain field data. Moving beyond the
historical field data that is available, another objective is to comprehensively
identify how to ensure the vibration measurement in the turbine condition
monitoring systems contains an as high amount of relevant information for the
detection and diagnosis of bearing failures. From this, a number of research
questions related to vibration-based monitoring of bearing failures in the wind
turbine drivetrain, have been formulated:
Q1 Which improvements can be made to the current condition monitoring
methodology with respect to e.g. accelerometer position, vibration mea-
surement properties etc.?
Q2 Can signal analysis alternatives developed in the research field improve
the detection of historical wind turbine drivetrain bearing failures?
Q3 Can the sensitivity in detection and diagnosis be increased utilizing a
form of machine learning implementation?

1.3 Thesis outline


Part I of this thesis contains the compilation summary of the thesis work.
Here, Chapter 2 contains an introduction to the wind turbine drivetrain and
the condition monitoring systems used in the thesis work. Chapter 3 presents
the rolling element bearing, the main methodology in vibration condition mon-
itoring of the bearings as well as related dynamic simulation work. Chapter
4 presents the family of wavelet transforms which has been utilized in vibra-
tion condition monitoring research as signal analysis tools to improve upon the
FFT, as well as the artificial neural network machine learning algorithm and
its potential use to increase the sensitivity of the detections and diagnosis. The
results from the five papers this thesis is based on is summarized in Chapter 5.
Lastly, reflections and conclusions regarding the research questions above are
presented together with possible avenues of future work in Chapter 6. Part II
contains the five papers this thesis is based on in the order they are used in
Part I.
Chapter 2

Wind turbines

The first use of a wind turbine to generate electricity occurred in late 19th
century Scotland when Prof James Blyth built his vertical axis turbine, with
four canvas sails providing the rotation [40]. The generated power was stored
in accumulators and later used to power the lights in his cottage, making it
the first wind powered house. The first horizontal axis wind turbine, with
three lightweight blades placed upwind, with a capacity of multiple megawatts
(MW) was developed in Denmark in the late 1970s. This then became the
standardized model partly as a consequence of the simultaneous political drive
away from nuclear power in the country. Modern onshore wind turbines are
not uncommon to have a rotor diameter over 100 meters, and be rated at a 2.5-
3MW capacity. Also, the requirement for even wind flows has driven the tower
height, i.e. how high the main hub called nacelle is situated, up to 100-130
meters.

2.1 The wind turbine drivetrain


As the wind drives the blades, the rotor will start to slowly rotate. For the
turbine to generate electricity, this rotation will have to reach between 10 and
15.5rpm. Up until the recent developments in slow rotating, permanent magnet
generators and the direct drive wind turbine, this rotation will have to be
increased substantially to accommodate for the optimal rotational speed range
of the commonly found asynchronous generator to produce electricity. Thereby,
a gearbox is needed in between the main shaft from the rotor to the generator.
The three-stage gearbox consists of one or two planetary gears followed by two
or one helical gear stages respectively, yielding a total gear ratio of typically
77.4 or 120 to one in the turbines used in this work. Thereby, the maximum
rotational speed of the generator shaft ends up between 1200 and 1850rpm. A
schematic of a wind turbine drivetrain with two planetary gear stages and one
helical gear stage can be seen in Figure 2.1 with the rotor, main gear compon-

7
8 Chapter 2. Wind turbines

: Bearings
: Accelerometer Ring

Pinion
Output shaft
Generator

Planets
Ring

Planets
Rotor

Gear
Sun
Sun

Main shaft

Carrier

Helical gear
1st planetary stage
stage

2nd planetary
stage

Figure 2.1: A configuration schematic of a turbine with two planetary gear stages
followed by a helical gear stage, the triangle symbols denotes the bearing positions
and the positions of the accelerometers for the vibration measurement system by
arrows.

ents and generator shown. All rotating components throughout the drive-
train is supported by a large number of bearings, illustrated as triangles in
the schematic. Most are cylindrical roller bearings to support the radial loads
while axial loads are supported mainly through tapered or thrust roller bear-
ings, or in some cases ball bearings. The bore diameter of these bearings varies
from around 200 to 800mm, generally being reduced in size as the rotational
speed increases. The largest bearing supporting the main shaft is a double row
spherical roller bearing with a bore diameter of at least 850mm.
Not immediately apparent in the schematic is the multiple parallel bearings
supporting the planets in the planetary gear stages. In the first stage, each
of the four planets are used as the outer raceway of four identical bearings
mounted on the planet carrier shaft to carry the heavy radial load in the sub-
system. Now, an even distribution of the radial load assumes the alignment
of the whole planetary gear to be perfect. A misalignment of e.g. one of the
planets will detrimentally influence this balance and the majority, if not the
whole, radial load will have to be carried by one of the end bearings. Thereby,
the assumed load case when making the bearings selection in the designing of
the gearbox is no longer valid and the overloading will cause extensive failures
2.2. Condition monitoring systems 9

in the end bearings much faster than the assumed lifetime. This also goes be-
yond the derived issues with bearing failures throughout the whole drivetrain
caused by the dynamic operation. In each acceleration and de-acceleration
of the rotational speed, the rolling elements can start to slightly slip on the
raceways, increasing what in tribological terms is called the slide-to-roll ratio.
This introduced slip can cause comparatively high stresses on the surfaces and
sub-surfaces of the component, compared to the advantageous pure rolling ac-
tion when the rotational speed is constant. Thereby, the inevitable failure of
the rolling element bearings can be accelerated compared to the pure rolling
operation most likely assumed in the design phase of the gearbox with respect
to the bearing lifetime.
Specific for the bearings in the generator, is the eventuality for the electrical
insulation of the bearings to fail and allowing high currents to pass through
the rolling element/raceway contact. With the high amount of energy trying
to pass the small contact area, weld arcs can form and material transfer from
one component to the other creating defects on the surfaces.

2.2 Condition monitoring systems


2.3 Drivetrain vibration monitoring
The condition monitoring systems of the drivetrains consists of a vibration
measurement system. In the turbine drivetrains used this work, the SKF IMx-
W Multilog on-line system was used. The system is connected to the internet
and remotely stores measurements to a database, allowing for a real time on-
line monitoring beyond the location of the wind farm. Accelerometers are most
commonly used to take vibration measurements at a regular interval and are
placed throughout the drivetrain [41], with their approximate measurement
position indicated as arrows in the Figure 2.1 schematic. The accelerometer
monitoring the first planetary gear stage is mounted directly on the ring gear,
while all other accelerometers are mounted on easily accessible positions on
the gearbox housing close to the bearings. As the accelerometers are moni-
toring a large number of components each, the vibration measurements will
contain information on the condition of several components which need to be
separated and continuously monitored for signs of defects, more on this in
Chapter 3.2. Also associated measurements on the rotational speed and the
momentary produced power by generator are made for each vibration measure-
ment. The rotational speed is continuously measured with a tachometer and
the maximum, minimum and mean rotational speed during the measurement
time is recorded and stored as meta information. Through this, conditions on
the rotational speed and variance can be set when a vibration measurement
should be stored to the database. Typically, the rotational is limited to be
between e.g. 700 and 1850rpm for the turbines used in this work, while the
variance can be no more than 30rpm to allow for a high degree of stationarity
10 Chapter 2. Wind turbines

of the vibration measurement.


The detection of defects in rolling element bearings and gears through vi-
bration analysis makes use of a repetition behaviour. E.g. the repeated impulse
vibration response of a defect on the bearing raceway being rolled over by the
rolling elements, or each time a damaged gear tooth mesh with the countering
gear teeth. These impulses with repeatedly be emitted with a certain fre-
quency decided by the dimensional parameters of the machine such as shaft
rotational speeds, number of gear teeth, bearing dimensions etc. Thereby, the
paramount phenomena which is analysed for is not the specific frequency of
the impulse itself, but how often they reappear. For the vibration analysis to
register the impulses as a repetitive behaviour, the measurement must contain
a large number of repetitions. Quantifying an exact number is difficult, but
as a rule of thumb at least 20-25 repetitions is needed and the measurement
time must therefore be set long enough with this in mind. The other main
measurement property to take into account is the measured frequency range.
This need to include the theoretical repetition frequencies indicating defects
in the monitored bearings and gears, included with a number of harmonics.
Since the repetition frequencies are dependent on the shaft rotational speed,
intuitively the measured frequency range could be set solely with respect to
the rotational speed of the monitored components. However, the first incipient
signs of a defect appearing in a bearing or gear exists as high frequency res-
onances of the machine structural components excited by the impacts. These
resonances are demodulated with the enveloping technique, explained in more
detail later in Chapter 3.2.1, to extract the repetition frequency of these events.
The specific frequency range where the resonance frequencies exist is decided
by the stiffness of the system and should therefore be the primary deciding
parameter for the measured frequency range, not the rotational speed which is
a common misconception.
The majority of the vibration measurements stored in the databases are
regular, time synchronous measurements over a fixed time and set sample rate.
Measurements are stored to the databases often once or twice per day together
with the frequency spectrum created by applying the FFT. Also, the number
of samples in each measurement is usually set to 16384 and the number of
lines in the FFT spectrum is normally set to 6400 lines. Thereby, an inherent
trade-off between the frequency range and measurement time is created by
the set number of samples and a given sampling rate yielding either a high
frequency range or a long measurement time. The relationship between these
two measurement properties is described with the following equations:

N = t ∗ fs ,
fs (2.1)
fmax =
2
where N is number of samples, t is measurement time, fs is sampling rate, fmax
is the maximum reliable frequency defined by the Nyquist theorem. The mea-
2.3. Drivetrain vibration monitoring 11

surement systems include a hardware anti-aliasing filter to remove frequencies


above the fmax . Due to the spectral leakage over such a filter the factor of 2 has
in practice been extended to 2.56. Thereby, the frequency range denomination
is changed into be fs /2.56. In the low rotational speed sub-systems of the drive-
train, i.e. the main bearing and planetary gears, the measurement time needs to
be comparatively long to cover enough rotations of the components. Therefore,
the regular time synchronous measurements used to monitor these sub-systems
are most often taken with a sampling rate of 2.56ksamples/s, where each mea-
surement will be 6.4s long and with a frequency range of 0-1kHz. Similarly,
when monitoring the generator and gearbox output shaft sub-systems with a
comparatively high rotational speed it is less crucial to measure over a longer
time. Therefore, the sampling rate of the measurement from the accelerometers
placed here is typically set to 12.8ksamples/s, yielding a measurement time of
1.28s and a frequency range between 0-5kHz. Lately, functionality to regularly
store raw measurements taken over a long period of time covering a wide fre-
quency range has been introduced in the vibration measurement systems used
in the turbines. However, this functionality became available to late for these
types of measurements to be used in this thesis work.
However, the raw vibrations measurements are most often not sensitive
enough to find early signs of failures in the drivetrain bearings, which as men-
tioned exist by the resonance of the structural components emitted from the
impacts in the high frequency range. Therefore, the enveloping technique is
implemented digitally in the measurement systems of the wind turbines to find
the incipient signs of a defect appearing in a bearing or gear. Important to
understand here is the inherent difference in the measurement properties be-
tween these enveloped measurements and the original measurements. When
the enveloping is implemented in the measurement systems the sampling rate
is set much higher, to facilitate a wide frequency range, and the measurements
are taken over a longer period of time. After the enveloping technique has been
performed the measurement is downsampled to the regular 16384 samples and
analysed in the time and frequency domains. Thereby, measurements with
the same size can be used but with both a wide frequency range and a long
measurement time.
As standard practice, the enveloped measurements are primarily used in
the frequency domain for the online, real-time monitoring and not stored to
the databases. As these measurements are more sensitive to the vibrations
introduced due to the defect being present, the detection and diagnosis is most
often done using the FFT spectra of these measurements. This can easily lead
to the turbine being stopped for maintenance and the damaged components
replaced before signs of the defects start to appear in the original measurements,
which is available for post-failure investigations. Also, the envelope spectrum
cannot be replicated using the stored original measurements due to the inherent
difference in the measurement properties. Thereby, no possible post analysis
can be done with historic data to e.g. investigate alternative signal analysis
methods beyond the FFT.
12 Chapter 2. Wind turbines

Another implemented post-processing technique, but generally not utilized


in the database measurement, is ordertracking. Here, the acceleration samples
are sampled with a fixed increment in shaft angular position, instead of a
fixed time step. This by resampling an over-sampling a measurement with
respect to the rotational speed. Thereby, the changes in rotational speed as
the measurement is taken does not introduce the non-stationarity which can
cause issues in the frequency analysis.
From a condition monitoring perspective, the most problematic bearing
failures occur in the planetary gears and to a lesser extent in the main bearing.
The relatively slow rotational speed of these sub-systems leads to the impacts
from the defects passing through the rolling element/raceway contact to be less
violent. This creates difficulties for the vibrations to propagate over the long
distance to where the accelerometers are placed, at a significantly high strength
to be reliably monitored. Also, the large amount of components in the planetary
gears which the vibrations have to propagate through potentially attenuates the
vibrations to an even larger extent. In the case of defects appearing on the static
outer raceway of the main bearing, the accelerometer position can in the worst
case not be close enough to detect the defect vibrations until a catastrophic
failure occurs. For the gearbox output shaft and generator bearings however,
the high rotational speed together with the accessibility for the accelerometers
to be placed close to the bearings makes failures in these bearing to be seen as
relatively easy to detect and diagnose reliably.

2.3.1 Other monitoring systems & techniques


Alternative types of measurements which is used for the condition monitoring
of the wind turbine drivetrains exist in the supervisory control and data acquisi-
tion (SCADA) systems. E.g., deviations of the actual and expected generated
power with respect to the current wind speed and pitch angle of the blades
etc. can be used to indicate issues with rotor and the drivetrain. Two often
used parameters for condition monitoring is temperature and debris measure-
ments which are performed throughout the lubrication circulation system in the
gearbox. While this can be useful, diagnosing the specific component in the
drivetrain which is experiencing the failure is not possible. Also, lubricant sam-
ples taken a couple of times per year from the oil sump in gearbox is analysed
to monitor the condition of the lubrication, if any contamination has entered
the closed circulation system, deviating physical and chemical properties etc.
Optimally, a framework of integration between these different condition moni-
toring systems is needed, which as of today is not widely focused on [39]. Such
a framework would allow the turbine owners or researches to be able to make
full use and correlate the information of all systems to better build a complete
understanding of what happened in the turbine which led to a failure. Today,
this requires an often insurmountable amount of manual work to synchronize
the information, e.g. which operating and weather conditions occurred during
and between the vibration measurements.
Chapter 3

Bearing failure & vibration


monitoring

The origin of the modern rolling element bearing can be found in drawings by
Leonardo da Vinci in the 15th century [42]. Earlier examples of two surfaces
being separated by rolling elements either consisted of iron strapped balls or
rollers allowed to be in contact with each other, with the adverse sliding friction
being a source of high losses. The rolling element separator in da Vinci’s
drawings, seen in Figure 3.1a, eventually led to the bearing cage technology.
The increased development in manufacturing of precision steel components
after the industrial revolution led to the rolling element bearing as it looks
today. The invention of the self-aligning bearing by Sven Wingquist, patent
drawings seen in Figure 3.1b, at the start of the 20th century to allow for
small misalignment movements in a machine without the supporting bearings
experiencing a high increase of friction and catching fire, led to the reliability
of industrial machines to drastically increase. Today, rolling element bearings
in one form or another is one of the more important machine elements in
rotating, industrial machinery. The rolling element bearing mainly consist of
four primary components:
• Rolling elements, converting sliding motion between two surfaces to rolling
action reducing the friction in the contact.
• Inner and outer rings with the raceways of the rolling elements.

• Cage, separating the rolling elements and lowering the internal bearing
friction by inhibiting them from being in contact with each other.

13
14 Chapter 3. Bearing failure & vibration monitoring

(a) Rolling element separator. (b) Self-aligning bearing.

Figure 3.1: Drawings of da Vinci’s rolling element separator in a) and Wingquist’s


patent of the self-aligning bearing in b) [43].

The rolling elements and the rings are typically manufactured from hard-
ened ANSI 51200 high chromium steel while the cage is typically brass, low
carbon steel or more recently polymers [44]. A lubricant is used to separate
the surfaces in the contact and thereby reduce the internal friction within the
bearing. This is either done by a grease, which can provide lubrication while
staying put in within the bearing, or an oil circulating system. Rolling element
bearing designs exist in various different types dependent mainly on if the load
case being purely radial, purely axial or a combination of both. Examples in-
clude ball bearings, cylindrical, needle and spherical roller bearings etc. as well
as tapered and thrust bearing configurations. Examples of a ball and roller
bearings with the different components highlighted can be seen in Figure 3.2.
Also, the requirements necessary for a rolling element bearing to operate to its
fullest lifetime is [45]:
1. Be of correct design and size for the predetermined operation load cases.
2. Be properly mounted on the shaft and seat.

3. Include some form of lubrication.


4. Operated at described speeds allowing for the lubrication to work prop-
erly.
5. Protected from contaminants entering the rolling element/raceway con-
tact.
3.1. Bearing failures found in the wind turbine drivetrains 15

Outer ring

Rolling
elements

Inner ring

Raceway

Cage

Figure 3.2: Illustration of ball and cylindrical roller bearing components (Courtesy
of SKF).

3.1 Bearing failures found in the wind turbine


drivetrains
Even though the above listed requirements are ensured, eventually the lifetime
of a rolling element bearing will come to an end. Thereby, prolonged operation
will always lead to degradation and fatigue of the bearing components. Esti-
mations can be made with the bearing life models originating in the Lundgren
and Palmgren life theory model, which gives a statistical measurement of the
lifetime at the predetermined operation parameters. However, this is not able
to give an accurate time-wise prediction of the defect initiation in a singular
bearing, only provide guidance to roughly what lifetime to expect [46].
The phenomena attributed to the limit of the service life of a bearing, is
the rolling contact fatigue. Here, the repeated loading and unloading of the
raceway and rolling element surfaces will lead to the maximum shear stresses
occurring at a certain depth in the subsurface. Thereby, fatigue cracks are
initiated which then propagate upwards to the surface. The depth at which
the fatigue cracks are initiated is determined by the Hertzian contact theory at
0.49 times the contact length [44]. When the subsurface-initiated cracks have
propagated to the surface, a piece of material is broken of creating a spall. Also,
surface imperfections such as dents in the raceway can locally increase the stress
at the surface and initiate cracks there. These cracks can then propagate down
to the region of maximum shear stress before propagating upwards again and
forming a spall. Similarly, the spall itself can act as an origin of increased
surface stresses and consecutive spalls forming. As a result of this, large parts
of the surface can be removed in a ripple pattern along the rolling direction.
16 Chapter 3. Bearing failure & vibration monitoring

Such an extensive failure is called flaking. Examples of an indentation in the


raceway causing a spalling failure and an extensively progressed flaking failure
can be seen in Figure 3.3. Both of these failure types are commonly found
in examples of wind turbine drivetrain bearing failures, with especially flaking
seen in planetary gear bearings. This due to the difficulty of detecting defects
early at these components in the gearbox as discussed earlier, giving the defect
time to grow extensively.

(a) Surface initiated spalling. (b) Extensive flaking damage.

Figure 3.3: Two types of bearing failures with a spall having been initiated at an
indentation on the surface in a) [47] and extensive flaking damage in a planet
bearings from a wind turbine drivetrain in b).

Micro-pitting is a failure mechanism where high pressures at boundary or


mixed lubrication in a rolling-sliding contact causes surface-initiated cracks
at the surface roughness asperity level and leading to micro-spalls covering
extended patches of e.g. the bearing raceways. This will appear on the rolling
surfaces as matt grey areas. While this process can be halted and amended
by changing the tribological system, e.g. changing the lubricant to more fully
separate the surfaces, it can also act as a source of increased surface stresses and
to an accelerated rolling contact fatigue failure [48]. Large patches of micro-
pitting is not uncommon to find within the wind turbine drivetrain, with an
example in the failure cases used in this thesis work appearing on the rolling
surfaces throughout the first planetary gear stage.
Another type of wind turbine drivetrain bearing failure worth of note typ-
ically appears in the generator ball bearings. Here if the electrical insulation
of a bearings fails, the electrical currents in the generator can pass through
the bearing. The high amount of electrical energy which passes through the
small point contact between the balls and raceways will create weld arcs and
material transfer between the components. The defects created will then lead
to increased stresses and further surface-initiated cracks can form leading to
3.2. Vibration monitoring of bearing health 17

spalling. This constitutes the three of the possible failure modes which has been
found in the wind turbine drivetrain bearing failure cases used in the work for
this thesis. However, all possible failure modes in the ISO/DIS 15243 standard
must be considered in root cause investigations of similar failures [49].

3.2 Vibration monitoring of bearing health


As previously mentioned in Chapter 1.1, the primary method of component
condition monitoring in wind turbines is vibration monitoring due to the ver-
satility to extract and separate information from different sources. The mea-
sured vibration time waveform is used to calculate statistical parameters to try
and describe the measurement, and by tracking changes to these parameters
an identification of progressing defects in the machine can be made [50]. These
statistical parameters include peak-to-peak, root-mean-squared (RMS), crest
factor, skewness and kurtosis. Mainly the first two has been used extensively in
wind turbine condition monitoring as trendable values to get an understanding
on the overall health of the machine or sub-system. As defects start to influence
the vibration measurements the RMS, seen as a general measure of the overall
vibration strength, and the peak-to-peak values, i.e. the maximum and mini-
mum peak amplitudes, start to increase from a healthy baseline and indicate
possible issues. Skewness and kurtosis has mainly been reported favourably to
be used in condition monitoring of rolling element bearings in research [51–53],
but is not widely used in the condition monitoring systems of the wind turbine
drivetrains. However, as each measurement position is used to monitor a mul-
titude of component and no separation of the information is made, an increase
of either parameter can only detect an overall issue with the machine and not
a component diagnosis.
To separate the vibration information from the different components, the
measurements is converted into the frequency domain. To do this, the FFT has
been implemented as a standard in wind turbine drivetrain condition monitor-
ing. The FFT is implemented on a time waveform x(t) to frequency amplitude
values Y (f ) over a set number of lines f , by:
Z∞
Y (f ) = x(t) (cos (2πf t) + i sin (2πf t)) dt. (3.1)
−∞

3.2.1 The enveloping technique


In 1984, McFadden & Smith summarily reviewed what was then called the
high-frequency resonance technique [54]. This technique intends to separate
vibrational information from bearings at the early stages of experiencing a de-
fect with the noise introduced in a vibration measurement from the surround-
ing structural components. Repeated impacts from the defect entering e.g. a
18 Chapter 3. Bearing failure & vibration monitoring

rolling element/raceway contact will excite resonances in the system compo-


nents which can be identified and separated from the structural noise. This
is done by band-pass filtering the measurement, keeping the high frequency
resonance component, then rectifying the measurement and applying a low-pas
filter. Thereby, the repetition frequency of the excited resonances is separated
and can be analysed without the structural noise interfering. Today, this tech-
nique is called enveloping and an illustrative example can be seen in Figure 3.4.
Here, four evenly spaced high frequency events exist on a 2Hz dominant sinu-

Figure 3.4: Step-by-step example of the enveloping technique demodulating a


repeated high frequency event from a dominant low frequency wave.

soidal. By the enveloping steps described these events are separated and their
repetition frequency becomes dominant as a result. The combination of apply-
ing the enveloping technique and the FFT constitutes the primary monitoring
methodology of wind turbine drivetrain bearings as it stands today.

3.2.2 Condition indicators of bearing health


Central in vibration condition monitoring of rotating equipment is defining
condition indicators for the specific components from e.g. dimensional data.
These condition indicators need to be measurable, giving us a condition indica-
tor or CI-value, and in the case of a defect will change from a baseline healthy
reference. The previously mentioned RMS, peak-to-peak and kurtosis values
can be such condition indicators on a machine or sub-system level.
The main indicator used for diagnosing the presence of a defect in a specific,
degrading rolling element bearing out of the multiple components in rotating
3.2. Vibration monitoring of bearing health 19

equipment machinery, is the theoretical bearing fault frequencies. Three of


these describe the repetition frequency of a defect, being present on either
the raceways or rolling elements, entering the rolling element/raceway contact
and emitting a singular vibration event. These events being similar to the
attenuated bursts seen the illustrative example of the enveloping technique in
Figure 3.4. Therefore, these are applicable e.g. when a spall from the RCF
has appeared and not when a distributed defect is larger than the spacing of
two rolling elements. In the cases of the defect being present on one of the
raceways, this repetition frequency is the same as frequency of the equally
spaced rolling elements passing over a specific point on the raceway. Thereby,
these are denominated as the ball pass frequency of the inner (BPFI) and outer
(BPFO) rings for an inner ring and outer ring raceway defect respectively. In
the case of a defect appearing on the rolling element, the defect will enter a
raceway contact twice per rotation of the rolling element. This bearing fault
frequency is therefore denominated as the ball spin frequency (BSF). When a
defect has appeared on the bearing cage however, the imbalance caused will
appear as an increased fundamental rotational frequency of the cage, and is
therefore denominated as the fundamental train frequency (FTF). These four
frequencies depend solely on the relative speed between the raceways in the
bearing and the bearing dimensions, namely the number ni and diameter D of
the rolling element, the pitch diameter dm where the rolling element’s centre
points are situated from the bearing centre, and the contact angle α between the
normal of the raceway and the normal of the axial centreline. An illustration of
these dimensions can be seen in Figure 3.5a and each bearing fault frequency
is calculated by [44]:
 
ωni D
BP F I = 1+ cos (α)
2 dm
 
ωni D
BP F O = 1− cos (α)
2 dm
 2 !. (3.2)
ωdm D 2
BSF = 1− cos (α)
2D dm
 
ω D
FTF = 1− cos (α)
2 dm

Also, illustrative FFT spectra of each bearing fault frequency as it would appear
in a bearing mounted on a rotating shaft and a stationary outer ring can be
seen in Figure 3.5b. Included in the illustrative spectra is gradually attenuated
harmonics of the fault frequencies as well as harmonics of sidebands around the
BPFI and BSF peaks. These sidebands appear due to the defect being present
on a rotating component and the bearing fault frequency peak being modulated
by the rotational frequency of the rotating component. This change in the case
of the inner raceway defect equalling the rotational speed of the shaft and the
spacing in the spectrum between the BPFI peak and sideband peaks will be
20 Chapter 3. Bearing failure & vibration monitoring

Ødm

ω α

Ødm

Ødm
ØD

ØD
(a) Dimensions for calculating the bearing fault frequencies, with a cylindrical roller bear-
ing on the two left views and a tapered roller bearing on the right view.

(b) Illustrative examples of the bearing fault frequencies in a FFT spectrum.

Figure 3.5: Illustrations of the dimensions for calculating the bearing fault frequen-
cies in a) and illustrative examples as they would look in a FFT spectrum when a
defect has appeared in a bearing with the inner ring rotating.

the same as the shaft rotational frequency. The distance between a rolling
element and the measurement position will change with the rotational speed
3.2. Vibration monitoring of bearing health 21

of the cage, and the spacing of the sideband peaks from the BSF peak will be
the cage rotational frequency, i.e. the FTF.
To create robust CI-values sensitive to increases of the bearing fault fre-
quencies while being insensitive to disturbances, a summation methodology
of several harmonics was used in all the work compiled in this thesis. When
a defect appears, the peak acceleration amplitude the frequency spectrum of
all these harmonics would increase, generating a certain increase in CI-value.
Meanwhile, interfering vibrational sources being excited, e.g. the rotational
frequency of a nearby geared up shaft, in the same frequency range of one
harmonic would not yield the same level of increase. The specific peak acceler-
ation amplitude level pi,j of k amounts of harmonics are summed up together
with l amounts of sideband harmonics if the situation would allow for these to
theoretically exist. Thereby,
v
u k ±l
uX X 2
CIn = t pi,j . (3.3)
i=1 j=1

Generally, three sets of harmonics of both the fault frequencies and sidebands
was chosen as further harmonics most often did not appear in preliminary
inspections of FFT spectra of enveloped measurements from a diagnosed defect
in a wind turbine drivetrain bearing.
For distributed defects, e.g. large patches of micro-pitting on rotating bear-
ing raceways, these can manifest in the frequency domain as changes to the
amplitude level of the fundamental shaft rotational frequency. As material
is worn away the design geometry is lost over a part of the raceway and the
cyclic imbalance manifests itself as an increase of the rotational frequency of
the raceway. If the increased surface stress at a spall is allowed to grow the
defect into extended flaking over large parts of the raceway longer than the
rolling element spacing, the amplitude strength of the bearing fault frequency
will decrease. Instead, the material loss will cause the same imbalance of the
rotating component, and can therefore be detected through the rotational fre-
quency. However, this will only be able to indicate existing issues at a system
level as multiple bearings are mounted on each shaft and further inspection
have to be conducted to diagnose the failing component.
Also, in cases where weak impact vibrations from a bearing defect is unable
to be distinguished at the theoretical fault frequencies, the fundamental vibra-
tions from nearby machine components can be modulated from the otherwise
healthy state of the machine. One such example could be bearing defects in the
planetary gear bearings. Here, the vibrations from the bearing can hypotheti-
cally be detected in changes to one of the planetary gears’ gear mesh frequency.
However, the two planetary gears in the turbine from which the schematic in
Figure 2.1 is based on include in total 26 bearings. Therefore, changes to a
gear mesh frequency over time cannot be used for a diagnosis of the failing
component, only that the gearbox as a unit is experiencing some kind of defect
22 Chapter 3. Bearing failure & vibration monitoring

and further investigations such as endoscopy inspection need to be conducted.


A illustrative example of the improvements found by employing the en-
veloping technique to separate the impulses emitted from a defect in a wind
turbine drivetrain bearing, and calculating the CI-values from the enveloped
measurement spectrum compared to the original measurement, can be seen in
Figure 3.6. In this example, the electrical insulation to one of the ball bearings

Figure 3.6: Normalized CI-values calculated from FFT spectra of original and
enveloped measurements leading up to a wind turbine generator bearing inner ring
failure.

in the generator failed and a defect appeared on the inner raceway. Vibra-
tion measurements were collected from before the incipient failure until the
bearing was replaced. The theoretical BPFI and shaft speed sidebands peak
amplitudes were identified, with respect to the rotational speed of the gener-
ator during the measurement being taken, and the CI-values calculated with
eq. (3.3) from both the spectra, normalized to make them directly comparable
and trended.
The level of separation achieved by enveloping the measurements from be-
fore the defect appeared and after is significantly higher, by a factor of at least
10. This clearly shows the increased detectability of the defect when the en-
veloping is employed. To reduce the risk of subsequent damage to surrounding
components, the power output of the turbine was severely limited after the
defect had been detected. Thereby, the overall strength of the spectra will be
lowered as well, which can be seen in the results without the enveloping being
used where the CI-value level at measurement 167 drops down to the level be-
fore the damage. Meanwhile, the detection level of the enveloped measurements
is still clearly above the level before the defect appeared.
3.3. Dynamic simulation of bearing vibrations 23

3.3 Dynamic simulation of bearing vibrations


To better understand the operation, tribological phenomena and internal dy-
namic movements of rolling element bearings, dynamic simulation models has
been developed and used for the last 70 years starting with limited models by
Palmgren in 1948 and Harris in 1966 [55, 56]. Thereafter, a complete dynamic
model was presented by Gupta and the non-linearity and time variance was
developed by Fukata et al. in the 70s and 80s [57, 58]. The simulation models
often seen in literature consists of a complete representation of a bearing inves-
tigating the influence of singular or distributed defects on dynamic movements,
emitted vibrations, wear, friction etc. Different simulation models focusing on
the vibrational response of a bearing can arbitrarily be divided into analyti-
cal models [59–62], dynamic models [63–70] and finite element models [71, 72].
Focus on e.g. modal shapes influence on spectrum absolute values [59], defect
positions [60], extended defects [66], stiffness variations [68], dent shoulders [72]
etc. has been prevalent. Lately, interest in the vibrational event emitted when a
rolling element enters and exits a defect has increased and complex vibrational
modelling focusing on these events has been prevalent [73].
The ADORE simulation model by Gupta is one of the more influential full
3D-representation of a generalized bearing, while focusing on computational
speed, and possibilities in vary operating condition variation, geometry imper-
fections, power loss, wear etc. has made it widely used [74]. The limited used of
generalized bearings has somewhat limited its potential as more complex bear-
ing and specifically cage designs has become of interest to simulate. The 3D
multi-body SKF BEAring Simulation Tool (BEAST) was initially developed as
a calculation model of new cage designs, and focuses in simultaneous detailed
contact calculations and simulating the dynamic behaviour [75]. Further, it
has been used for studying individual components’ dynamic movements, stress
analysis at surface and sub-surface levels, fatigue analysis, lubrication phe-
nomena in elastohydrodynamically lubricated contacts, thermal conductivity
etc. The contact calculations are made by the intersection of two surface el-
ements, defined from the bodies centre coordinate system in a specific time
step yielding the contact forces, deformations and body accelerations. More
coordinate systems can be defined connected to the bodies or independently,
and stiffness/damping matrices can be introduced in ties between two coordi-
nate systems. An integrated finite element solver is used to calculate the body
deformations, and the structural surface movements measured through defined
coordinate systems.
However, the optimization of the accelerometer position to ensure an as
high signal-to-noise ratio as possible has not seen much focus from bearing
simulation research as of yet. Tackling the accelerometer position optimization
problem could increase the early detectability of defects in critical sub-systems
and components and thereby reduce the associated downtime costs, making it
interesting from a condition monitoring perspective [76]. Seeing as full scale
filed testing in an actual wind turbine drivetrain would be a nearly overwhelm-
24 Chapter 3. Bearing failure & vibration monitoring

ing logistical and economical challenge, moving towards a simulating test bench
could be beneficial.
Before dynamic simulation work of the wind turbine drivetrain bearings
can start, another situation where the load case and operation can be fully
controlled is needed. One such example of a BEAST simulation model with
a medium sized, double row, SKF 22220E spherical roller bearing seated in a
flange housing is shown in Figure 3.7. The bearing model is imported from a
repository with the exact design geometry and the flange seating specifically
designed within the software. Instead of defining another body representing
the shaft and defining the contact to the inner ring, a stiffness parameter is set
to the inner ring bore surface. Thereby, the shaft is represented but without
the need for another set of detailed contact-force calculations which increase
the computational effort considerably. The inner ring is only allowed to

Applied axial force

Applied radial force


Depth

(a) Simulation model overview with illustration of the modelled defect at the bottom left.
3.3. Dynamic simulation of bearing vibrations 25

Rot.
dir.

y z

(b) Simulation model xy-view (backside).

Figure 3.7: BEAST simulation model overview with bearing components, flange
housing, load cases and surface plot of a spline curve representing the defect on the
inner raceway in a) and xy-view of the simulation model backside with 72 equally
distributed coordinate systems and rotational direction in b).

rotate around the centre-axis (z), and is fixed in all other degrees of freedom.
Meanwhile, the rotation of the flange is constricted around the centre-axis
to counter the inner rings rotation. The load case is created through a spring
preload in the axial direction, acting on the back row of rollers, and a radial load
downwards, creating a loaded zone at the top of the bearing. The influence of
this non-rotational symmetry of the load case on the emitted bearing vibrations
goes beyond the methodology to only ensure a high signal-to-noise ratio by
placing the accelerometer as closely as possible to the bearing as a whole [77].
Also, a 0.5mm deep indentation 1mm long and 10mm wide is modelled on
the back row inner raceway. Thereby, the defect size encompasses the Hertzian
contact and when the defect passes the roller/raceway contact the components
no longer touch. Thereby, the impact vibrations emitted will be significantly
higher, compared to only the changes load carrying area if the defect is smaller,
allowing for a clear identification using the bearing fault frequencies in eq. (3.2).
72 coordinate systems equally distributed around the circumference measuring
the surface velocity in the axial direction is placed on the flange, thereby al-
lowing for the accelerations over time at different positions to be calculated.
The FFT spectrum of the enveloped surface acceleration at the measurement
position directly above the bearing loaded zone can be seen in Figure 3.8a while
26 Chapter 3. Bearing failure & vibration monitoring

the calculated CI-values, using eq. (3.3) with three BPFI harmonics and three
shaft speed sidebands, can be seen in Figure 3.8b. The optimal choice of ac-

(a) FFT spectrum of the simulated vibrations at θ = 0 in b).

0
330 30
40
300 30 60
20
10
270 0 90

240 120

210 150
180
(b) CI-values calculated from the simulated vibration spectra for all 72 measurement
positions.

Figure 3.8: Simulated vibration results with FFT spectrum from the measurement
position above the primary bearing loaded zone in a) and CI-values calculated for
all 72 measurement positions in b).

celerometer position is here as close to the loaded zone as possible, with an 20%
increased CI-value compared to the minimum. However, in the wind turbine
drivetrain the accelerometers are placed at the top of the components directly
opposed to the bearing loaded zones. Thereby, for an as optimal vibration
measurement as possible this would require the accelerometers to be placed on
the bottom surface of the gearbox instead. This location is not easily accessible
3.3. Dynamic simulation of bearing vibrations 27

when working in the nacelle and accessibility has most likely had a substan-
tial influence in mounting the accelerometers when the vibration measurement
system was installed.
A comparable spectrum and circumferential CI-value results from valida-
tion experiments in a test rig with the same load case, rotational speed and
artificial defect 1mm long and 10mm wide across the bearing raceway experi-
encing the axial load, is seen in Figure 3.9. 24 accelerometer positions around
the circumference is able to validate the circumferential CI-value results, and
enveloped measurements collected with an SKF IMx-W vibration measurement
system and a magnet to mount the accelerometer to the flange surface. While
the increase in CI-value does not extend beyond the position closest to the
bearing loaded zone as was the case in the simulated results, the same appar-
ent increase is present directly above the primary loaded zone with a reasonable
qualitative agreement. The same conclusion can be said about the simulated
and experiment spectra. While, there are differences in the noise level etc. the
same vibrational components exist from the rotational frequency of the shaft
and the BPFI with shaft speed sidebands. Also, the peak amplitude levels
of the rotational frequency and first BPFI harmonic, are comparable to the
simulated spectrum.

(a) FFT spectrum of the experimentally measured vibrations at θ = 0 in b).


28 Chapter 3. Bearing failure & vibration monitoring

0
330 30
60
300 40 60

20
270 0 90

240 120

210 150
180
(b) CI-values calculated from the experimentally measured vibration spectra for all 72
measurement positions.

Figure 3.9: Measured vibration results from test rig experiments with FFT spectrum
from the measurement position above the primary bearing loaded zone in a) and
CI-values calculated for all 24 measurement positions in b).
Chapter 4

Vibration monitoring in
research

With the operation of a wind turbine being dynamic in nature, the stationarity
requirement of the vibration measurements when applying the FFT is difficult
to fulfil. While the measurement systems have limits in variances of generator
rotational speed allowed for a measurement to be saved, the stored measure-
ments cannot be described as stationary. Therefore, the experiences gained in
vibration condition monitoring from fully controlled laboratory environment
and other rotating machinery such as paper machines where operation is more
stable, is not directly transferable [7]. The trend in development of new signal
analysis methods have been towards introducing more complexity to find more
details in simulated or laboratory measurements. The solutions created in the
research into new methodologies for signal processing, analysis, detection and
diagnosis of failures needs to instead be evaluated and shown to be able to
perform in wind turbine drivetrain field data before the industrial actors can
trust the new methods enough for a widespread implementation.

4.1 Wavelet-based methods


One technique of signal processing and analysis which has been used in research
extensively but has not seen a wide implementation in the wind turbine drive-
train condition monitoring systems is the wavelet transform. In this technique,
the similarities between a predefined wave shape, i.e. the so called mother- or
base wavelet, and a measured time waveform is analysed. By, translating the
mother wavelet in the time domain of the measurement as well as dilating and
contracting the mother wavelet, a 2D map of the time-frequency domain can
be analysed. Thereby, certain events similar to the mother wavelet in the mea-
surement can be found in both the time and frequency domain simultaneously.
The origins of the wavelet technique can be found in the mathematics field in

29
30 Chapter 4. Vibration monitoring in research

the early 20th century. The work by Alfred Haar in his dissertation led to a set
of rectangular basis functions of a positive pulse followed by a negative pulse,
thereby defining the simplest mother wavelet which later was denominated
the Haar wavelet. No major advancement was made until Jean Morlet in
the mid-1970s, when analysing acoustic echoes, introduced the dilation and
contracting as well as the translation of the analysis window function and
called the resulting waveforms wavelets.

4.1.1 Continuous & discrete wavelet transform

The theoretical wavelet transform, where a signal can be transformed into small
wavelets and back without loss of information, was proposed by Morlet together
with Alex Grossmann in 1984 [78]. The scale function, i.e. dilating and con-
tracting a wavelet, was improved and a multi-resolution analysis using wavelets
was introduced, increasing the applicability substantially [79, 80]. Thereby, fil-
ter banks could be created of a broken-down time waveform by implementing
orthonormal wavelets in a convolution operation, resulting in what now called
the continuous wavelet transform (CWT). A full presentation of the mathemat-
ical representation of the wavelet transforms can be found in Gao and Yan’s
compiled book on the subject [81]. The continuous wavelet transform is written
as:
Z∞
1
wt(a, b) = √ x(t)ψ ∗ (a, b)dt, (4.1)
a
−∞

where ψ (a, b) is the complex conjugate of the scaled and shifted mother wavelet

ψ:  
1 t−b
ψ ∗ (a, b) = √ ψ , (4.2)
a a
where x(t) is the measured time waveform, while a is representing the dilation
and contraction scale parameter and b is the translation shift parameter, both
which are varied continuously over the whole time waveform. However, the
continuous form of the wavelet transform requires a relatively high computa-
tional effort. By a discretization of the scale parameter, this could be reduced
making it easier implemented on e.g. the large amount of vibration measure-
ments in the wind turbine drivetrains online condition monitoring system. The
discretization of the scale parameter is done, with a linear relationship to the
shift parameter, by: (
a = 2j
. (4.3)
b = k2j
4.1. Wavelet-based methods 31

Thereby, eq. (4.1) is rewritten as:


Z∞
1
wt(j, k) = √ x(t)ψ ∗ (j, k)dt, (4.4)
2j
−∞

and the discretized mother wavelet:


 
1 t − k2j
ψ (j, k) = √ ψ

. (4.5)
2j 2j

This new form is called the discrete wavelet transform (DWT). Further, Mallat
introduced the scaling function, φ, used with the wavelet function in the dual
scale equation:
√ X
φ(t) = 2 h(n)φ(2t − n)
(4.6)
n
√ X .
ψ(t) = 2 g(n)φ(2t − n)
n

Here, g(n) and h(n) are scale independent high- and low-pass filters created
from the mother wavelet. The wavelet transform is thereby implemented as a
high- and low-pass filtering operation with a chosen number of repetitions on
the low frequency component, denominated as levels of decomposition. Also,
the outputs from the filtering operations are denominated as detail and approx-
imation coefficients for the high- and low-pass filtering operations respectively.
In the practical implementation, the coefficients are downsampled by a factor of
two in each level of decomposition. This is done to ensure the amount of infor-
mation having to be handled not being twice as much in each step. Thereby,
there is an inherent trade-off between the time and frequency resolutions to
consider when deciding on signal decomposition level. The DWT implemen-
tation procedure is illustrated in Figure 4.1. One of the first examples of the

Level 4
љ
Level 3 L cସ,஺
Ϯ
Level 2 љ
L cଷ,஺
Ϯ
љ љ
Level 1 L cଶ,୅ H cସ,஽
Ϯ Ϯ
Approximation љ
љ H
L coefficients cଷ,஽
Ϯ Ϯ
cଵ,୅ љ
Signal H cଶ,஽
Ϯ
љ Detail
H coefficients
Ϯ
cଵ,ୈ

Figure 4.1: Schematic of the DWT where L and H denotes the low- and high-pass
filters followed by a downsampling by half into wavelet coefficients cj,n .
32 Chapter 4. Vibration monitoring in research

DWT being used in bearing vibration research was performed by Mori et al.
in 1996 [82]. Here, the wavelet coefficients were shown to increase in strength
leading up to a spalling event in an artificiality dented thrust ball bearing
mounted in a test rig. Further studies have shown its ability to e.g. estimate
the defect size [83], and as a pre-processor for a machine learning algorithm
e.g. for fault classification [84].
The mother wavelet used can, compared to the base sinusoidal in the FFT,
be selected to better resemble the vibration response of an impact as a bearing
defect enters the rolling element/raceway contact. The mother wavelets are
divided into different types, or families, and is compiled in Table 4.1. While
this list constitutes all mother wavelets available in the Matlab repository to
be used in the DWT implementation, other families out off the sevel listed
exist but cannot be used to create the high- and low.pass filters needed. Also,
examples of the Haar, Daubechies of order 10 and Coiflet wavelet of order 3
can be seen compared to a sinusoidal in Figure 4.2.

Table 4.1: Full lists of mother wavelets available in the Matlab repository.

Wavelet family Number Denomination


Daubechies 1-45 db1 (i.e. haar), db2,..., db45
Symlets 46-89 sym2, sym3,..., sym45
Coiflets 90-94 coif1, coif2,..., coif5
bior-1.1, -1.3, -1.5, -2.2, -2.4, -2.6, -2.8,
Biorthogonal 95-109
-3.1, -3.3, -3.5, -3.7, -3.9, -4.4, -5.5, -6.8
rbio-1.1, -1.3, -1.5, -2.2, -2.4, -2.6, -2.8,
Reverse biorthogonal 110-124
-3.1, -3.3, -3.5, -3.7, -3.9, -4.4, -5.5, -6.8
"Discrete" Meyer 125 dmey
Fejer-Korovkin filters 126-130 fk4, fk6, fk8, fk14, fk22

Figure 4.2: FFT sinusoidal, haar, db10 and coif3 mother wavelet comparison.
4.1. Wavelet-based methods 33

The Daubechies wavelet family is most often used in bearing research, al-
though motivation for this choice is often non-existent or based on limitations
to the required computational effort [85]. Studies attempting to identify an
optimal choice of mother wavelet in the wavelet transforms when applied in
bearing condition monitoring has found no consensus [86–92]. Also, no such
study had been performed using field vibration data.
How the Haar wavelet perform in detecting a bearing failure is seen in Fig-
ure 4.3 together with comparable progression ratios of each mother wavelet in
Table 4.1. Here, a dataset with enveloped vibration measurements from a wind
turbine drivetrain condition monitoring system, leading up to the replacement
of a faulty output shaft bearing is used. The DWT is applied at three levels us-
ing all 130 mother wavelets individually. Thereafter, the FFT is applied to the
approximation coefficients containing all the theoretical bearing fault frequen-
cies indicating defects in the bearing. CI-values calculated from each spectrum
by eq. 3.3 using three harmonics of the BPFI and shaft speed sidebands, gives
comparable progression-to-failure trend curves for each mother wavelet. By
identifying a healthy period in the dataset, the first 120 measurements, and a
faulty period of the last 20 measurements a comparable value of the CI-values
progression ratio can be calculated. Thereby, the mother wavelets ability to
separate the bearing vibrations and identify the defect can be ranked. From

Figure 4.3: Progression-to-failure for a gearbox output shaft bearing failure and
progression ratio results for all mother wavelets, indicated as numbers following
the list in Table 4.1.

the progression ratio, all choices of mother wavelet will result in a clear identi-
fication of the bearing defect. However, a specific choice will have a negligible
influence as the progression ratio only varies 1.3%. The optimal choice in this
bearing failure case would be the Daubechies order 38 mother wavelet. Two
other bearing failure datasets, one gearbox output shaft and one planet bearing
failure, that was used for validation purposes showed similar tendencies, but
with the optimal choice of mother wavelet changing and thereby being case
34 Chapter 4. Vibration monitoring in research

specific.
Instead, the most important parameter to consider when implementing the
DWT turned is the computational effort needed. By measuring the time it
takes to perform the DWT operations with each mother wavelet, the com-
putational effort can be illustrated and is seen in Figure 4.4. The measured

Figure 4.4: Computational effort of the DWT dependant on mother wavelet used,
normalized to the Haar wavelets computational effort.

computational effort is normalized to the simplest Haar wavelet, and the scale
of the increase of all other mother wavelets is logarithmic. Taking the lack
of increase in progression ratio using the Symlet wavelets into account, the
use of high order Symlet wavelets can be excluded entirely as it requires an
unmanageable computational effort without any increased performance.

4.1.2 Wavelet packet transform


From the work by Mallat, the implementation of the DWT was later extended
to perform the signal decomposition over the whole frequency content in the
wavelet packet transform (WPT) [93]. Wickerhauser implemented the complete
signal decomposition from eq. 4.6 as:
X
dj+1,2n = h(m − 2k)dj,n
(4.7)
m
X
dj+1,2n+1 = g(m − 2k)dj,n
m
4.1. Wavelet-based methods 35

where j is level of decomposition. The resulting wavelet coefficients, with a


known frequency span, are stored in easily handled packets in a node tree, where
n is node tree position. The node tree is illustrated in Figure 4.5. The first use

Node, n
0,0
0 ! '()
()*

1,0 1,1
"#$% "#$%
0 ! !'()*
/ /

2,0 2,1 2,2 2,3


"#$% "#$% "#$% 4"#$% "#$% 4"#$%
0 ! ! !'()* !
3 3 / 3 / 3
Level, j

3,0
"#$%
%
0 !
& 3,4
3 4 3,6
3 6
3,2
4"#$% "#$% 2"#$% "#$% 1"#$% 3,7
! ! '()* ! 1"#$% 4"#$%
& / & / &
!
3,1
3 1 3,5
3 5 & 3
3,3
"#$% "#$% "#$% 4"#$% 4"#$% 2"#$%
! ! !
& 3 3 & 3 &
… … … … … … …
… … … … … …
… … … … …
j0
j,0
… … … …
0 ! '+,. j1
j,1 jj,n
'+,. ! 5'+,. '()* -'+,. ! '()*

Figure 4.5: Illustrated wavelet packet transform tree structure repeated j number
of levels and frequency range of specific nodes n.

of this method in bearing vibration condition monitoring research was reported


by Nikolaou & Antoniadis in 2002 [94]. Here, inner raceway defects were identi-
fied in simulated signals and in bearing test rig measurements. Further studies
have used the WPT for wind turbine generator bearing defect detection [95],
identification of the optimal decomposition for test rig bearing defects [96] as
well as a pre-processing tool for machine learning algorithms [97–101].
To find where the majority of information is situated in a decomposed signal,
the signal energy content of each packet is used. However, this has either just
been used as an identification tool on which wavelet coefficients to apply the
FFT for a detailed frequency analysis, or as a trended value to find changes
over time of a specific frequency span. At decomposition level j and tree node
n, the signal energy content of the wavelet packet coefficients dj,n is calculated
by:
N
X
Es (dj,n ) = |dj,n (i)|2 . (4.8)
i=1

With the theoretical fault frequencies of multiple components possibly experi-


encing a defect existing close together in a wind turbine drivetrain, no example
36 Chapter 4. Vibration monitoring in research

can be found where a clear diagnosis is made by only monitoring the wavelet
coefficient energy in literature. However, if the vibration measurements from
the wind turbine drivetrain condition monitoring systems is broken down to
such a level that each node’s frequency content matched the frequency content
per line when applying the FFT on the measurement, a comparable frequency
spectrum can be created. Also, by calculating the CI-values from the relevant
vibrational components and normalizing the progression to failure, the two
different methods to convert the enveloped measurement into the frequency
domain can be compared. In Figure 4.6 and 4.7 such comparisons are shown
in two wind turbine bearing failures, one gearbox output shaft bearing spalling
failure and one planet bearing flaking failure in the first planetary stage. Also,
the methodology previously described using the DWT, with the Daubechies
30 mother wavelet in three levels and applying the FFT on the low frequency
approximation coefficients, is included in the comparison. The results from
the gearbox output shaft failure dataset is seen in Figure 4.6, with 4.6b con-
taining the 45 measurements around the incipient CI-value increase. Here, the
CI-values for all measurements from the first 510 measurements in the dataset
follow each other and fall under normal operation, i.e. the defect has not
appeared yet. The variation in this time period broadly follows the associ-
ated generator shaft rotational speed measurement. From the detailed view
in Figure 4.6b, only the CI-values from the WPT spectra is able to increase
substantially over the previous variation. Actually, this first indication by the
WPT spectra reaches a level neither of the other two methods do before the
bearing was replaced. Thereby the WPT is able to indicate the defect a week

(a) CI-values for the full time-period.


4.1. Wavelet-based methods 37

(b) CI-values around the incipient increase.

Figure 4.6: Comparative results of normalized CI-values for a gearbox output shaft
bearing failure with the full time-period leading up to the bearing being replaced in
a) and the CI-values around the incipient increase in b). CI-values calculated from
the FFT spectra in blue, DWT followed by the FFT in red and the WPT spectra
in black.

earlier, allowing for the maintenance planning to start earlier and the downtime
reduced. After the incipient detection, the level of increase using the WPT is
roughly 2.5 times higher than the FFT. This added detail by using the WPT
can thereby increase the surety of an increase of CI-values happening due to a
defect and not from other sources disturbing the measurement. Also, the ad-
dition of the DWT has slightly lowered the level of increase compared to only
applying the FFT on the enveloped measurements. This happens due to the
frequency resolution being coarser as the wavelet coefficients are downsampled
and the BPFI peaks flattening. However, the implementation of the WPT can
work around this problem.
The trended CI-values for all three signal analysis methods when applied
on the enveloped measurement leading up to the planet bearing failure is seen
in Figure 4.7. Here, all three methods fail to identify the progressing bearing
failure at all. Either the defect vibrations are attenuated before reaching the
accelerometer, which might be alleviated if the accelerometer is moved closer
to the loaded zone in accordance with the results in Chapter 3.3. Alternatively,
the high frequency resonances of the impact vibrations exist higher than the
1kHz upper limit on the frequency range of the measurements available in the
condition monitoring database for this specific case.
38 Chapter 4. Vibration monitoring in research

Figure 4.7: Comparative results of normalized CI-values for a 1st stage planetary
gear planet bearing failure with CI-values calculated from the FFT spectra in blue,
DWT followed by the FFT in red and the WPT spectra in black.

4.2 Influence of measurement properties


The known situation with most of the online vibration monitoring being done
through post processed data from non-stored measurements, has a detrimen-
tal influence on the historical vibration data available for post-failure vibra-
tion condition monitoring research, as seen in the Figure 4.6 and 4.7. Also,
researchers looking for historical field data need to be made aware of these
constrictions in the existing data. Therefore, new recommendations on the
properties of the stored measurements, i.e. frequency range and measurement
time, has to be made. The aim of these recommendations is to optimize the
storage of vibration measurements to the databases for later use, with respect
to their measurement time and frequency range properties. I.e. investigate
if the properties of the currently stored measurements are good enough to be
able to perform post-failure investigations or if the storage of more detailed
measurements need to start. To help to categorize different measurements
with respect to the relationship between those properties, explained through
eq. (2.1), a division into four quadrants is used, illustrated in Figure 4.8.
The optimal choice is of course 1st quadrant measurements, measured over
a long period of time, to cover the rule-of-thumb 20-25 repetitions of e.g. the
defect passing through a rolling element/raceway contact, and covering a wide
frequency range to include the repetitive behaviour of the high frequency res-
onances emitted from the defect. Such measurements are often found in labo-
ratory environments where a smaller amount of long measurements taken with
4.2. Influence of measurement properties 39

II
(JUDZPHDVXUHPHQW
I
(JSUHHQYHORSHGPHDVXUHPHQW
VORQJ aVORQJ
&RYHULQJN+] &RYHULQJN+]

Measurement time
Frequency range

III IV
(JUDZPHDVXUHPHQW
VORQJ
&RYHULQJN+]

Figure 4.8: Relationship between the measurement properties divided into quad-
rants with increasing frequency range moving upwards and increasing necessary
measurement time moving to the left.

a high sampling rate, allowing for the wide the frequency range, can be han-
dled. However, limitations into the amount of data generated which needs to
be transferred from e.g. a complete wind farm makes the unfeasible in un-
treated raw form as field data is concerned. The measurements which has been
enveloped within the vibration measurement systems and thereafter downsam-
pled, fulfils the requirements of a 1st quadrant measurements while not being
any larger than the currently stored measurements, but are not generally stored
to the databases.
Instead, the decision into measurement length and covered frequency range
of the vibration measurements stored to the databases, is dependant on mis-
conceptions of the rotational speed being the only deciding factor, as previously
discussed in Chapter 2.3. It is assumed that raw measurements only covering
enough rotations of the component will yield early signs of a failing component
and can be stored as historical data to be used later in eventual post-failure
investigations. This has in turn led to only 2nd or 4th quadrant measure-
ments being stored to the databases. These measurements are, however, not
40 Chapter 4. Vibration monitoring in research

optimized for finding the repetitive behaviour of the high frequency resonances
emitted from a defect bearing. Thereby, the frequency range has been set
too low to include the high frequency resonances to be able to replicate the
real-time monitoring.
For the monitored drivetrain components normally operating at a low ro-
tational speed, i.e. the main bearing and planetary gears, this has led to the
measurement time being set relatively long and consequently the frequency
range to low, as we saw in figure 4.7. These values are typically 6.4s and 0-
1kHz, which in the arbitrary division in Figure 4.8 has been designated as 4th
quadrant measurements. For the high-speed side with the gearbox output shaft
and generator bearings, the higher rotational speed has allowed the measure-
ment time to be set lower and consequently the covered frequency range will
be higher. Here, these values are typically 1.28s and 0-5kHz, designated as 2nd
quadrant measurements in Figure 4.8.
The recommendations made in this work for condition monitoring of bear-
ings in wind turbine applications are based on a larger number of bearing failure
case datasets throughout the drivetrain. Included in this is bearing failure cases
from another turbine drivetrain set-up with a single planetary gear followed by
two helical gear stages, and an intermediate shaft in between. Thereby, rec-
ommendations on measurements for monitoring failures in the intermediate
shaft bearings can also be made. Again. the vibration measurements were
enveloped before the FFT and WPT was applied individually to create spec-
tra from which CI-values could be calculated. Also, datasets of 1st quadrant
measurements were available in some select cases and is analysed to investi-
gate if the either/or relationship between the measurement time and frequency
range of the regularly stored measurements causes failed monitoring of bear-
ing defects. Figures 4.9-4.11 shows the trended CI-values in selected dataset
throughout the wind turbine drivetrain, with the results from the remaining
bearing failure cases found in Paper D.
These bearing failures include a generator bearing failure in Figure 4.9a, a
gearbox output shaft failure in 4.9b, planet bearing failures from the 2nd plan-
etary stage in 4.10a and 4.10b, and a main bearing failure in 4.11a and 4.11b.
The two high speed side failure datasets consists of 2nd quadrant measurements
with a frequency range of 0-5kHz and 1.28s measurement time, the planet bear-
ing dataset 4th quadrant (0-1kHz frequency range and 6.4s measurement time)
and longer 4*th quadrant ordertracked measurements, and the two dataset
from the same main bearing failure consists of 2nd quadrant measurements
(0-2kHz frequency range and 3.2s measurement time) and 1st quadrant mea-
surements having been enveloped in the vibration measurement system before
being stored to the database encompassing a 0-10kHz frequency range.
For the two high speed side failure cases, Figure 4.9 show the CI-values
from the FFT and WPT spectra on the 2nd quadrant measurements. For both
cases, a substantial increase in CI-values appears detecting and diagnosing the
defects. This increase appear at the same time for the CI-values from the
FFT and WPT spectra, but noting the y-axis scale shows the level of increase
4.2. Influence of measurement properties 41

using the WPT to be at an substantially higher level. Thereby, the increased


detection and diagnosis performance shown in Figure 4.6 is shown again.

(a) Normalized CI-values for a generator bearing failure.

(b) Normalized CI-values for a gearbox output shaft bearing failure.

Figure 4.9: Normalized CI-values, from both both FFT and WPT spectra of 2nd
quadrant measurements, for a generator bearing failure in a) and a gearbox output
shaft bearing failure in b).

From this, the recommendations on stored measurements’ properties are:


42 Chapter 4. Vibration monitoring in research

• at least 2nd quadrant measurements for generator bearing failures. This


as the typically available 2nd quadrant measurements with a 0-5kHz fre-
quency range and 1.28s measurement time are enough for these types of
investigations.
• at least 2nd quadrant measurements for gearbox output shaft bearing
failures. This as the typically available 2nd quadrant measurements are
enough for these types of investigations. However, 1st quadrant measure-
ments not typically available can be needed for increased sensitivity.
• 1st quadrant measurements for intermediate shaft bearing failures. This
as typically available 2nd quadrant measurements can struggle to detect
the defects at all as seen in the results in Paper D.
In Figure 4.10a, the CI-values from the 4th quadrant measurements are
unable to yield a substantial increase at all and no apparent detection of the
bearing defect can be claimed, as was the similar case in Figure 4.7. For the
other planet bearing failure case, storage of the 4*th quadrant ordertracked
measurements, in Figure 4.10b, started as a reaction to the second planet
bearing defect being identified. Thereby, the defect is present for the first
1050 measurements and the last 240 once the gearbox had been replaced can
be used as healthy reference. The CI-values when the defect is present in the
4*th quadrant measurements are substantially higher than the healthy refer-
ence. Also when noting the y-axis scale in Figure 4.10b using the 4*th quadrant
measurements, the level of increase using the WPT spectra CI-values is at an
substantially higher level compared to the FFT spectra CI-values. Thereby, the
detection and diagnosis performance is increased by implementing the WPT to
convert the enveloped measurements into the frequency domain.

(a) Normalized CI-values for a planet bearing failure using 4th quadrant measurements.
4.2. Influence of measurement properties 43

(b) Normalized CI-values for a planet bearing failure using 4*th quadrant ordertracked
measurements.

Figure 4.10: Normalized CI-values, from both both FFT and WPT spectra, for two
2nd planetary stage planet bearing failures with 4th quadrant measurements in a)
and 4*th quadrant ordertracked measurements in b).

Here, the recommendations on stored measurements’ properties are:


• 1st quadrant measurements for planetary gear bearing failures. This as
the typically stored 0-1kHz frequency range and 6.4s long, 4th quad-
rant measurements are not enough for a successful detection. While,
4*th quadrant ordertracked measurements could provide indications of
the defect in Figure 4.10b, the used healthy reference on the replace-
ment gearbox introduce to much uncertainties to fully claim a successful
detection.
For the main bearing failure in Figure 4.11, both the 2nd and 1st quadrant
measurements are able to identify the defect, but the 1st quadrant measure-
ments at a substantially higher level. For the 2nd quadrant measurements in
Figure 4.11a, the CI-values from both the FFT and WPT spectra are able to
indicate the defect to the same level. When noting the y-axis scale in Fig-
ure 4.11b for the 1st quadrant measurements, however, the CI-values from the
WPT spectra show a significantly higher level of increase compared to the FFT.
From this, the recommendations on stored measurements’ properties are:
• 1st quadrant measurements for main bearing failures. While 0-2kHz fre-
quency range and 3.2s long, 2nd quadrant measurements were shown to
be useful, 1st quadrant measurements having been enveloped before the
measurement was stored to the database, was able to yield a much higher
detectability of the defect.
44 Chapter 4. Vibration monitoring in research

(a) Normalized CI-values for a main bearing failure using 2nd quadrant measurements.

(b) Normalized CI-values for a main bearing failure using 1st quadrant measurements.

Figure 4.11: Normalized CI-values, from both both FFT and WPT spectra, for
a main bearing failure with 2nd quadrant measurements in a) and 1st quadrant
enveloped measurements in b).
4.3. Implementing a machine learning solution 45

4.3 Implementing a machine learning solution


To aid with the detection, diagnosis and prediction of bearing failures, the im-
plementation of machine learning solutions has been a recently popular area of
research. Often, this is motivated to replace the vibration expert making main-
tenance decisions from small ocular differences in the vibration measurements
or CI-value trends. However, implementations of machine learning solutions in
the condition monitoring designed to increase the sensitivity of the detection
and diagnosis to help the decision making is highly advantageous. The two main
techniques which can be found in literature is implementations with artificial
neural networks (ANN) [97, 102–105] and support vector machines [106–108].
Similar to the development in new signal analysis methods by the research
community, a lot of work with these techniques introduces more complexion
instead of moving towards ensuring a successful field implementation. These
techniques have primarily been used as classifiers to find deviations in gen-
eralized descriptive values of vibration measurements from healthy data and
correlate these to which component that is experiencing the defects. However,
this requires labelled data for all of types considered failures which cannot be
argued, which is very difficult to obtain in available wind turbine drivetrain
field data. Also, studies using SCADA data such as wind speed, blade pitch
angle, drivetrain component temperatures, generated power etc. to detect de-
fects have been performed [109–111]. However, taking advantage of the benefits
in vibration monitoring compared to lubrication and temperature monitoring
for early detection is thereby lost due to the lack of integration between the
SCADA and condition monitoring systems [39].
From the previously shown CI-value results in Figure 4.6a, the inherent
variation of these values during normal operation when essentially describing
the noise level of the spectra, yielded difficulties in the incipient detection of
the defect. This variation is eerily similar to the associated rotational speed
measurements, from which there exist an inherent relationship to the CI-values.
This is further shown in the planet bearing CI-value results in Figure 4.10a,
where a slight change in behaviour is seen in the last 200 measurements. With
the historical knowledge that a defect is present, this can be perceived as a
detection of the defect. However, taking the rotational speed measurements
into consideration, these were consistently towards the highest limit of the
turbine and would inherently yield higher CI-values at that period. Thereby,
no detection of the defect could longer be confidently said to have happened.
One of the stated advantages of an ANN is its ability to design a descrip-
tive function from a complex relationship between sets of observations [98].
Thereby, an ANN implementation could be able to separate the perceived
variance in CI-values from the incipient signs of the defect and increase the
sensitivity of the condition monitoring solution. A major advantage of this is
the utilization of measured operational data directly from the vibration mea-
surement system already associated with the vibration measurements to find
behaviour abnormal to operation of the wind turbine when healthy. Also, this
46 Chapter 4. Vibration monitoring in research

implementation would only need labelled healthy data for the learning period
of the neural network.
The structure of an ANN is generally said to consist of three layer types,
the first layer with neurons corresponding to the input size to the network, a
chosen number of hidden layers consisting of a chosen number of neurons each
and finally the output layer. Also, each neuron in the layers are connected
to all neurons in the following layer, yielding a structure seen in Figure 4.12.
As the input values is passed through the network, these are multiplied with

Hidden Hidden
layer 1 Hidden layer N
Weights, layer 2 Weights,
" #
Input Weights,
Output
Weights, 1 1
layer layer
! $
1

2 2
1 1
2

… … …
… … …

2 2
… …


m n

y
x z

Figure 4.12: Schematic of an artificial neural network with N hidden layers.

a set of weights between each neuron. The weights are trained from initial
random values in the training period to as closely as possible match with the
known output target value. The training is performed by iteratively minimizing
the error between the output and target values through the backpropagation
algorithm [112].
The influence of implementing an ANN solution, to separate the perceived
variance in CI-values from the incipient signs of bearing defects, in the detec-
tion and diagnosis sensitivity can be shown to in the following examples from
both the gearbox output shaft bearing failure and 2nd stage planet bearing
failure datasets in Figure 4.6a and 4.10a respectively. In training the ANNs,
the first 300 CI-values are used as targets in training individual ANNs while
the associated rotational speed measurements are used as inputs. As this is
performed for the CI-values from both the FFT and WPT spectra, two ANNs
is created for each failure case as well. An ANN structure of three hidden lay-
ers with 64, 8 and 64 neurons are consistently used. This structure compared
similarly to single hidden layer structures and less neurons when evaluating the
training performance, without reducing the performance due to overfitting and
was able to yield higher repeatability. After training the ANNs, the rotational
4.3. Implementing a machine learning solution 47

speed values for the whole datasets is to predict the CI-values. Thereby, an
anomaly detector is created for when the true CI-values start to differentiate
from what the prediction indicates they would had been without the defect
being present.
From the gearbox output shaft bearing failure case, the CI-values’ rela-
tionship to the rotational speed and variance in rotational speed is seen in
Figure 4.13a and 4.13b respectively. Included in the datasets are stored vi-
bration measurements taken during non-normal operation. When the turbine
has been at a standstill, the generator needs to reach 880rpm in order to start
generating power again and the rotational speed can thereafter fall to 750rpm
before it stops. However, the lower limit on the rotational speed set in the con-
dition monitoring system for a vibration measurement to be stored is 700rpm.
Thereby, vibration measurements exist where the resistance in the drivetrain
is much lower, both as the rotational speed is below 750rpm and as the turbine
is accelerating towards 880rpm, and as a consequence the spectrum noise level
will be reduced. Also, if a measurement is taken during the turbine acceleration
or de-accelerating while the generator is online, the noise level of the spectrum
will be lower as well. These data-points can have a detrimental effect on the
training and performance of the ANN and need to be filtered out. As already
mentioned in Chapter 2.2, a limit on the variance in rotational speed during a
measurement is set in the condition monitoring system at 30rpm. However, the
CI-values from the three measurements taken at the highest rotational speed
deviated from the expected values as they had also been taken with a relatively
high variance in rotational speed, and might also need to be removed.

(a) CI-values dependent on rotational speed


48 Chapter 4. Vibration monitoring in research

(b) CI-values dependent on variance of rotational speed

Figure 4.13: Absolute CI-values dependent on rotational speed in a) and variance


in b). CI-values filtered out with respect to the generator not producing in red and
to high variance in black

The influence of filtering out measurements taken with a variance in rota-


tional speed above 8rpm is seen when comparing Figure 4.14a to Figure 4.14b.

<------------------------------------ Training data ------------------------------------>

(a) Predicted CI-values by the ANN compared to the true values at the top graph and
normalized difference at the bottom graph, before rotational speed variance data filtration.
4.3. Implementing a machine learning solution 49

(b) Normalized difference between ANN’s predicted and true CI-values with rotational
speed variance data filtration.

Figure 4.14: Comparison of predicted CI-values by the ANN in red and true gearbox
output shaft bearing failure CI-values from FFT spectra in blue as well as the
normalized difference between predicted and true values before filtering with respect
to rotational speed variance in a) and normalized difference between predicted and
true values after filtering with respect to rotational speed variance b).

Here, the ANN anomaly detection solution is implemented before and after
filtrating with respect to the variance in rotational speed on the gearbox output
shaft bearing failure CI-values from the FFT spectra in Figure 4.6a. In the top
graph Figure 4.14a, both the true CI-values in blue and the ANNs prediction
in red is shown, the bottom graph then contains the absolute value of the
difference between the true and predicted values. The three distinct peaks
around measurement number 270-370 correspond to the three measurements
taken during the highest rotational speed and with a high variance. The ANN
thereby predicts much higher values due to the rotational speed being high, and
when calculating the difference these peaks register as anomalies at the same
level of increase as when the defect has appeared at the end of the dataset.
However, by the additional data filtration before re-training the ANN these
peaks are eliminated and the level of increase when the defect has appeared is
higher. Also, the ANN solution implemented on the CI-values from the FFT
spectra is able to detect the defect earlier compared to the results in Figure 4.6.
Now, the detection can be made simultaneously as the CI-values from the WPT
spectra and the comparable sensitivity is increased from 4 in Figure 4.6a to 30
with the ANN solution implemented.

The implementation of the ANN anomaly detection on the CI-values from


the WPT spectra does not display the same detrimental influence by the vari-
ance in rotational speed, as seen in Figure 4.15a. By performing the filtration
however, the level of increase in the difference between the predicted and true
CI-values when the defect appears is increased further to 50 in Figure 4.15b.
Compared to the relative level of increase of the CI-values by themselves before
implementing the ANN, this has increased by a factor of five.
50 Chapter 4. Vibration monitoring in research

<------------------------------------ Training data ------------------------------------>

(a) Predicted CI-values by the ANN compared to the true values at the top graph and
normalized difference at the bottom graph, before rotational speed variance data filtration.

(b) Normalized difference between ANN’s predicted and true CI-values with rotational
speed variance data filtration.

Figure 4.15: Comparison of the predicted CI-values by the ANN in red and the true
CI-values, calculated from WPT spectra from the gearbox output shaft bearing fail-
ure results in Figure 4.6a, in black as well as the normalized difference between the
predicted and true values before filtering with respect to rotational speed variance
in a) and the normalized difference between the predicted and true values after
filtering with respect to rotational speed variance b).

When implemented on the 2nd stage planet bearing failure results in Fig-
ure 4.10a, the difference between the predicted and true CI-values from the
FFT spectra is negligible for the first 1024 measurements, seen in Figure 4.16a.
Thereafter follows a short period of increased difference before steadily increas-
ing for the final 200 measurements. This later steady increase is also seen in
the difference between the predicted and true CI-values from the WPT spectra.
Thereby, the implementation of the ANN solution as an anomaly detector is
able to detect a type of bearing failure in the wind turbine drivetrain which has
4.3. Implementing a machine learning solution 51

previously been impossible using the readily available vibration measurements.

<--------- Training data --------->

(a) ANN’s predicted compared to the true CI-values at the top graph and normalized
difference at the bottom graph, using CI-values from FFT spectra.

<--------- Training data --------->

(b) ANN’s predicted compared to the true CI-values at the top graph and normalized
difference at the bottom graph, using CI-values from WPT spectra.

Figure 4.16: Predicted CI-values by the ANN red and true CI-values, from a 2nd
planetary stage bearing failure, as well as normalized difference between the pre-
dicted and true CI-values from a) FFT spectra in blue and b) WPT spectra in
black.
52 Chapter 4. Vibration monitoring in research
Chapter 5

Contributions

This thesis is the summation of five research papers, in which more details on
the topics in the comprehensive summary as well as further findings can be
found. A short summary as well as my contribution to each appended paper
are as follows:

5.1 Paper A
Title: Multi-body simulation and validation of fault vibrations from rolling
element bearings.
Authors: D. Strömbergsson, P. Marklund, K. Berglund.
Published in: Submitted to Proc. IMechE, Part J: Journal of Engineering
Tribology.
Summary: In this paper, a dynamic simulation model of a radially loaded,
complete spherical roller bearing mounted in a seat housing is set up and used
to investigate the circumferential symmetry of the vibration response. Also,
an inner raceway defect is modelled and the axial direction surface velocities
on the housing component is measured at 72 circumferentially equal spaced ac-
celerometer positions. The FFT spectra of the enveloped measurements showed
the expected BPFI and sidebands clearly detection the defect. The CI-values
calculated from the bearing fault frequencies showed an optimal accelerometer
position as close to the primary loaded position as possible. In the wind turbine
drivetrain the accelerometers are placed directly opposed the loaded zone and
the simulation results shows a possibly for the monitoring performance to be
increased by changing the accelerometer position.
Contribution: I participated in formulating the idea, conducted the simu-
lation and experimental work, analysed the data and wrote the draft of the
manuscript.
Relevance: This paper partially addresses research question 1 and provides a
practical action to improve the condition monitoring of wind turbines.

53
54 Chapter 5. Contributions

5.2 Paper B
Title: Mother wavelet selection in the discrete wavelet transform for condition
monitoring of wind turbine drivetrain bearings.
Authors: D. Strömbergsson, P. Marklund, K. Berglund, J. Saari, A. Thom-
son.
Published in: Wind Energy, 2019; vol. 22, Issue 11, p. 1581-1592.
Summary: In this paper, a comprehensive analysis of the mother wavelet in-
fluence on the condition monitoring performance, using enveloped wind turbine
gearbox vibration measurements, is performed. Three level DWT operations
was applied on the measurements with each available mother wavelet individ-
ually before the FFT was applied and the relevant bearing fault frequencies
summarized into CI-values. While the results show the potential of the DWT
to be used for detection bearing defects, the choice of mother wavelet did not
have a significant influence on the increase of the CI-values. Also, the optimal
mother wavelet choice was shown to be case specific. Instead, the deciding
parameter was the computational effort to perform the DWT, with specifically
the Symlet family mother wavelets being particularly computationally heavy
whole not increasing performance.
Contribution: I participated in formulating the idea, collaborated with Dr.
Juhamatti Saari to develop the first part of the signal analysis on the vibration
measurements provided by Allan Thomson, conducted the wavelet analysis and
wrote the draft of the manuscript.
Relevance: This paper partially addresses research question 2 and provides
a guide to implement the DWT on wind turbine drivetrain vibration measure-
ments.

5.3 Paper C
Title: Bearing monitoring in the wind turbine drivetrain - A comparative
study of the FFT and wavelet transforms.
Authors: D. Strömbergsson, P. Marklund, K. Berglund, P.-E. Larsson.
Published in: Wind Energy, 2020; vol. 23, Issue 6, p. 1381-1393.
Summary: In this paper, the DWT implementation from Paper B and a novel
implementation of the WPT directly on enveloped vibration measurements is
evaluated against applying the FFT using two wind turbine gearbox bearing
failure cases. In the a gearbox output shaft bearing failure case, the previous
implementation with the DWT applied before the FFT on the enveloped vibra-
tion measurements were able to yield a lower level of increase of the CI-values
compared to applying FFT on the measurements without the DWT. This due
to the downsampling being performed in each step of the DWT coarsening the
5.4. Paper D 55

frequency resolutions of the spectra and flattening the bearing fault frequency
peaks. The WPT implementation was able to increase the detectability of the
failure roughly 2.5 times. Also, the incipient signs of the defect only rise above
the inherent variance of the CI-values leading up to the failure when the WPT
is applied. Thereby, the defect can be detected a week earlier and subsequently
the maintenance planning starting earlier to reduce the turbine downtime and
associated costs. In a planet bearing failure case however, all methods of signal
analysis is unable to detect and diagnose the defect, either due to the high res-
onance signs of the defect being higher than the measurement frequency range
or not reaching the accelerometer at all.
Contribution: I participated in formulating the idea, conducted the signal
analysis of the measurements and wrote the draft of the manuscript.
Relevance: This paper builds on Paper B, partially addresses research ques-
tion 2 and shows the potential of the WPT on wind turbine drivetrain vibration
measurements.

5.4 Paper D
Title: Wind turbine drivetrain condition monitoring by vibration analysis Part
A: Property requirements of vibration measurements.
Authors: D. Strömbergsson, P. Marklund, K. Berglund, P.-E. Larsson.
Published in: Submitted to Wind Energy.
Summary: In this paper, new recommendations on the properties of the vi-
brations measurements being stored to the wind turbine drivetrain condition
monitoring databases are made. 15 cases of bearing failures throughout the
whole drivetrain are used as basis with datasets covering different measure-
ment lengths and frequency ranges. Normally, measurements taken either over
a long period of time or with a wide frequency range is stored to the databases.
CI-values from FFT and WPT spectra on enveloped measurements are used
to evaluate the detection of the bearing defects. The resulting recommenda-
tion identified measurements taken both over a long period of time and with a
wide frequency range being preferred. In a generator bearing failure, the nor-
mally stored short measurements with a wide frequency range was seen to be
enough to replicate the monitoring when the failure was detected. In all other
sub-systems, measurements in accordance with the new recommendations are
needed for turbine owners to best utilize their systems for post-failure analysis
and for researchers to evaluate new developments of signal analysis tools.
Contribution: I participated in formulating the idea, conducted the signal
analysis of the measurements and wrote the draft of the manuscript.
Relevance: This paper builds on Paper C, partially addresses research ques-
tion 1 and 2. For turbine owners, the paper provides practical action to improve
the condition monitoring of wind turbines. For researchers, it provides guid-
ance into what type of field data measurements to look for.
56 Chapter 5. Contributions

5.5 Paper E
Title: Wind turbine drivetrain condition monitoring by vibration analysis Part
B: Anomaly detection using artificial neural networks.
Authors: D. Strömbergsson, P. Marklund, K. Berglund.
Published in: Submitted to Wind Energy.
Summary: In this paper, the CI-values from three of the bearing failure cases
are used to evaluate the potential of an ANN anomaly detection solution. The
networks was trained to generate functions describing the inherent relationship
between the generator rotational speed measurements, taken simultaneously
as the vibration measurements, and the CI-values before the defect appears.
Thereby, CI-values could be predicted as if the defect had not appeared, and
compared to the true CI-value. The difference between the predicted and true
CI-values showed a further increased detection sensitivity in the gearbox output
shaft bearing failure from Paper C and a similar failure from Paper D. Also,
in a planet bearing failure case from Paper D the anomaly detection is able to
indicate the defect appearing, which was not possible by purely evaluating the
CI-values by themselves.
Contribution: I formulated the idea, conducted the signal analysis of the
data and wrote the draft of the manuscript.
Relevance: This paper is directly connected and builds on Paper D, addresses
research question 3 and provides a guide to implement an ANN anomaly detec-
tion scheme to improve the sensitivity of the condition monitoring to a point
that one planet bearing failure could be detected and diagnosed.
Chapter 6

Conclusions & Future Work

This chapter contain the conclusions of the compilation thesis work, with re-
flections on research questions from Chapter 1.2 and also the extended results
of the appended papers. Further, possible avenues of future research and de-
velopment work is discussed
The goal of the research work summarized in the thesis was to compre-
hensively improve detection and diagnosis of wind turbine drivetrain bearing
failures. This as the wind turbine drivetrain still experience early failures lead-
ing to a long downtime of the turbine, which in extension lead to high amounts
of costs in lost production and maintenance. Thereby, an early detection is
required to reduce the downtime, facilitate taking action such as limiting the
turbine output to produce less but over a longer period of time, facilitate earlier
maintenance planning to help reduce costs etc. More in depth, the thesis work
consists of investigating how to find high signal-to-noise ratio measurements,
evaluating the potential and difficulties of research developed signal analysis
methods when applied on field data and finally utilizing a machine learning
solution to increase sensitivity of the monitoring to specific failures.

6.1 Conclusions
This thesis investigated the optimal accelerometer position, signal analysis
methods from research and machine learning anomaly detection to improve
detection and diagnosis of wind turbine drivetrain bearing failures. The results
provide comprehensive overview on practical actions to facilitate as clear vi-
bration measurements as possible to monitor the condition of the wind turbine
drivetrain bearings. Stated answers to the research questions in Chapter 1.2
are as follows:

Q1 Which improvements can be made to the current condition mon-


itoring methodology with respect to e.g. accelerometer position,
vibration measurement properties etc.?

57
58 Chapter 6. Conclusions & Future Work

Concerning the accelerometer position, vibration results from a dynamic


simulation model of a similar sized bearing to the smaller ones found in
the wind turbine drivetrain, would indicate that the accelerometers are
not optimally positioned. In the simulated case, the rotational frequency
of the shaft and bearing fault frequencies exist in the FFT spectra of
vibration measurements extracted from accelerometer positions on the
bearing housing. However, the relative strength of the condition indi-
cator value constructed from the bearing fault frequency harmonics is
increased closer to the primary loaded zone of the bearing. This primary
loaded zone was situated at the top of the simulated bearing, contrary to
most of the accelerometer positions in the wind turbine drivetrain where
they are placed in the circumferential opposite to the loaded zone on the
bottom side of the sub-systems. Thereby, the difficulties found primarily
in monitoring the main bearing and the bearings within the planetary
sets prone to experience extensive failures can in part be reduced by sim-
ply changing the position of the accelerometers to surfaces underneath
the bearings in the gearbox where to loaded zones are situated. Vali-
dation experiments in a bearing test rig showed reasonable qualitative
agreement to the simulated results with a high coherence of the shaft
rotational frequency and the bearing fault frequencies’ peak amplitude
values between the simulated and test rig spectra.

Concerning the vibration measurement properties, at the high rotational


speed side of the drivetrain the vibration measurements regularly stored
to the database are enough to be able to recreate the detectable signs
leading to a diagnosis of failures. This includes vibration measurement
used to monitor the generator and gearbox output shaft bearings. Repli-
cating the monitoring is not possible as there exists an inherent difference
in the measurements stored to the database and the measurements pri-
marily used for the online real-time monitoring. However, the frequency
range is wide enough in the stored measurements, while still being long
enough to not be detrimental, for a detection to be possible and the per-
formance of new solutions in signal analysis etc. can be evaluated against
the currently used methods. On the low rotational speed side, however,
the regularly stored measurements are not able to recreate the detectable
signs leading to a diagnosis of failures. In the monitoring of the bearings
within the planetary gears and the main bearing, the regularly stored
measurements taken over a relatively long period of time does not con-
tain nearly the frequency range needed to replicate the detection and
diagnosis from the real-time monitoring. Instead, measurements both
over a long period of time and taken with a high sampling frequency
to contain a wide frequency range has to be stored to the databases for
a successful monitoring both to be made in real-time and replicated in
post-failure analysis. This information is of high importance to both the
turbine owners, to make full use of their condition monitoring systems and
6.1. Conclusions 59

solutions, as well as researchers looking to source field data to investigate


the performance of newly developed signal and data analysis solutions.
If this is not taken into consideration, progress to address the difficulties
in monitoring problematic bearing failures is severely limited and on the
other does not give the optimal foundation to show the applicability of
new solutions to work towards a widespread industrial implementation.
Q2 Can signal analysis alternatives developed in the research field
improve the detection of historical wind turbine drivetrain bear-
ing failures?
While the DWT can be used with the implementation in this work to
detect and diagnose wind turbine bearing failures, with the enveloped
measurements being filtered down to its narrowest band possible and still
contain a number of bearing fault frequency harmonics, the specific choice
out of 130 tested mother wavelets will not have a significant influence.
The total variance in the level of the CI-value increase from measure-
ments without a defect present to measurement after it has appeared in
two gearbox output shaft bearing failures was no more than 1.3%. Also,
the optimal choice was shown to be case specific. Instead, computational
effort to perform the DWT with the selected mother wavelet becomes the
deciding parameter, e.g. with increasing detail of Symlet wavelets leading
to a logarithmic increase in computational effort.
Implementing the DWT contains an inherent disadvantage when com-
pared to applying the FFT directly on the enveloped measurements, how-
ever. The wavelet coefficient outputs are downsampled to half the size in
each step of signal decomposition to counter the amount of information
in the output otherwise being twice as much as the input. Thereby, the
frequency resolutions when applying the FFT on the wavelet coefficients
for a frequency domain analysis will be coarser and detail is lost, lead-
ing to the frequency peaks indicating the defect being present flattening
and being comparative lower than the reference without the DWT im-
plemented. This then shows as a slightly lower level of increase of the
CI-values calculated from the bearing fault frequencies.
By implementing the WPT on the enveloped measurements in such a way
that the inherent disadvantage in the downsampling within the DWT, the
level of increase when a gearbox output shaft bearing experiencing a de-
fect is more than double the case with FFT applied on the enveloped
measurements were able to produce. Thereby, the monitoring on this
type of bearings is less sensitive to false alarms occurring due to the in-
herent variance in the CI-values before the defect appears, which can
lead to unnecessary costs in investigations and maintenance activities.
The main advantage gained is the WPTs ability to differentiate the in-
cipient increase in CI-values from the inherent variance, which the FFT
is unable to produce. Thereby, the defect can be detected and diagnosed
four days in advance and action taken to reduce the turbine downtime
60 Chapter 6. Conclusions & Future Work

during the maintenance event as well as limiting the generator output if


the failure starts to increase the risk for consequent damages in surround-
ing components.
These results ware only valid for failures happening in the high-speed side
of the drivetrain, i.e. the gearbox output shaft and generator bearings.
As discussed in the Q1 reflections, the typically stored vibration mea-
surement from the planetary gear and main bearing accelerometers were
did not contain a wide enough frequency content to a substantial degree
include the repeated high frequency resonance from the defect passing
through the rolling element/raceway contact. Thereby, the increased sen-
sitivity of the WPT solution is void in an example case of planet bearing
failure as the signs of the defect does not exist in a substantial amount
to yield a definitive increase of the CI-values.

Q3 Can the sensitivity in detection and diagnosis be increased uti-


lizing a form of machine learning implementation?
Yes, by implementing an artificial neural network an anomaly detector
could be created increasing the sensitivity of the vibration monitoring
substantially. The relationship between the CI-values before the defect
appears and the associated rotational speed measurements, being a con-
tributor to the variance seen in the CI-value behaviour, was learned by
the ANN and subsequent predictions on new CI-values with respect to the
rotational speed could be made. When the incipient damage of the defect
appears and starts to influence the CI-values, the difference between the
predicted and true values will give an early detection and diagnosis. This
increased sensitivity adds on the previously discussed increase using the
WPT to convert the enveloped measurement into the frequency domain
and calculating the CI-values in the gearbox output shaft bearing failures.
In a second stage planet bearing failure where small differences in CI-value
behaviour, coinciding with a consistently high rotational speed leading up
to the maintenance event, was able to be amplified to a substantial degree
leading to a detection to be possible. Thereby, a condition monitoring
solution which can detect the problematic bearing failures in the plane-
tary gears and main bearing without a large number of historical cases
needed and can utilize the readily available field data has been shown.

6.2 Future work


During the conducted thesis work summarized in this thesis and from the con-
clusions drawn, new avenues of possible research activities has been identified.
These are presented in this section in three paragraphs regarding the simula-
tion work, the implementation of the wavelet packet transform to convert the
measurements into the frequency domain and finally the implementation of the
6.2. Future work 61

artificial neural network to increase the sensitivity of the failure detection and
diagnosis.
The bearing simulation model in the dynamic simulation work can be used
for a substantial amount of future research activities. However, the main focus
should first be to increase the agreement between the model itself and the
validation experiments from the test rig. The confidence of the minute details
in the results be significant in the actual application can thereby be higher and
their implications more certain. An example for the use of the simulation model
can then be to design a methodology to identify the optimal frequency range
where the high frequency resonance from the defect impacts exist. Thereby, the
use of the enveloping technique in the vibration condition monitoring of rolling
element bearings can be further refined at an early stage. After these steps,
the modelling work can move to a bearing from the wind turbine drivetrain.
One of the primary drivers of the lack of knowledge regarding the root cause
of several types of bearing failures in the drivetrain is in the unknown dynamic
load cases of the bearing, where a better understanding would lead to a more
optimized choice of bearing type and size in the design phase [113]. Thereby,
an optimal goal of the simulation model would be to investigate if the load
cases can be approximated with the vibrational behaviour found when running
the simulation model.
While ready for full scale implementation in a monitoring software, the
possible advantages by the wavelet packet transform to convert the measure-
ment into the frequency domain can be identified. The present implementation
was designed to separate the waveform behaviour of an attenuated impact,
as it would theoretically be seen when a defect passes through the rolling
element/raceway contact, and make full use of the theoretical bearing fault
frequencies as condition indicators. However, its potential to be used with al-
ternative condition indications which is not limited to the frequency domain
can be evaluated.
Regarding the implementation of an artificial neural network in the con-
dition monitoring solution, a lot of optimization research work can still be
performed. Firstly, the presented implementation made use of available func-
tions in the MathWorks Matlab software which limited the number of input
parameters which in turn could be correlated to a single output parameter. A
neural network Toolbox in the software could be used instead to facilitate this,
with the limitation to only a single hidden layer being available and the opti-
mization of the number of neurons being a lot more time consuming. However,
moving away from the available functions and toolboxes, or in fact into another
software, to write the calculation code from the ground up could facilitate more
input parameters and the use of more than one hidden layer. Thereby, the pre-
sented implementation could move away from the filtration of the available
measurements with respect to the variance in rotational speed, and instead
include this value as an input parameter. Also, an input value describing the
separation between the peaks constituting the CI-value and an overall increase
of the spectrum noise level beyond the rotational speed value could advanta-
62 Chapter 6. Conclusions & Future Work

geously be included. Thereby, the diagnosing ability would increase as possible


surrounding issues in the drivetrain which can influence the overall spectrum
noise level from the expected value will not appear as an abnormal behaviour
of the designed condition indicator for a specific type of bearing failure. A first
step could be to include the measurements RMS-value as an indicator to its
overall strength, but a more refined parameter could also be defined.
References

[1] GWEC, “Global Wind Report 2015,” tech. rep., Global Wind Energy
Council, 2016.

[2] GWEC, “Global Wind Report 2019,” tech. rep., Global Wind Energy
Council, 2020.

[3] EWEA, “Wind energy in Europe in 2019 - Trends and statistics,” tech.
rep., WindEurope, 2020.

[4] C. Walford, “Wind turbine reliability: understanding and minimizing


wind turbine operation and maintenance costs,” Tech. Rep. SAND2006-
1100, Sandia National Laboratories, 2006.

[5] S. Sheng, “Wind turbine gearbox reliability database, condition mon-


itoring, and operation and maintenance research update,” Tech. Rep.
No. NREL/PR-5000-68347, National Renewable Energy Lab (NREL).,
Golden, CO, 2017.

[6] M. C. A. Olde Keizer, S. D. P. Flapper, and R. H. Teunter, “Condition-


based maintenance policies for systems with multiple dependent compo-
nents: A review,” European Journal of Operational Research, vol. 261,
no. 2, pp. 405–420, 2017.

[7] Z. Hameed, Y. S. Hong, Y. M. Cho, S. H. Ahn, and C. K. Song, “Con-


dition monitoring and fault detection of wind turbines and related algo-
rithms: A review,” Renewable and Sustainable Energy Reviews, vol. 13,
pp. 1–39, 2009.

[8] C. J. Li and J. Ma, “Wavelet decomposition of vibrations for detection


of bearing-localized defects,” NDT & E International, vol. 30, no. 3,
pp. 143–149, 1997.

[9] R. Rubini and U. Meneghetti, “Application of the Envelope and Wavelet


Transform Analyses for the Diagnosis of Incipient Faults in Ball Bear-
ings,” Mechanical Systems and Signal Processing, vol. 15, no. 2, pp. 287–
302, 2001.

63
64 References

[10] P. W. Tse, Y. H. Peng, and R. Yam, “Wavelet Analysis and Envelope


Detection For Rolling Element Bearing Fault Diagnosis - Their Effec-
tiveness and Flexibilities,” Journal of Vibration and Acoustics, vol. 123,
no. 3, p. 303, 2001.

[11] J. Lin, M. J. Zuo, and K. R. Fyfe, “Mechanical Fault Detection Based on


the Wavelet De-Noising Technique,” Journal of Vibration and Acoustics,
vol. 126, no. 1, 2004.

[12] D. F. Shi, W. J. Wang, and L. S. Qu, “Defect Detection for Bearings


Using Envelope Spectra of Wavelet Transform,” Journal of Vibration and
Acoustics, vol. 126, no. 4, pp. 567–573, 2004.

[13] P. W. Tse, W. X. Yang, and H. Y. Tam, “Machine fault diagnosis through


an effective exact wavelet analysis,” Journal of Sound and Vibration,
vol. 277, no. 4-5, pp. 1005–1024, 2004.

[14] W. X. Yang and X. M. Ren, “Detecting impulses in mechanical signals by


wavelets,” Journal on Applied Signal Processing, vol. 8, pp. 1156–1162,
2004.

[15] W. Sun, G. A. Yang, Q. Chen, A. Palazoglu, and K. Feng, “Fault diag-


nosis of rolling bearing based on wavelet package transform and envelope
spectrum correlation,” Journal of Vibration and Control, vol. 19, no. 6,
pp. 924–941, 2012.

[16] C. K. E. Nizwan, S. A. Ong, M. F. M. Yusof, and M. Z. Baharom,


“A wavelet decomposition analysis of vibration signal for bearing fault
detection,” in 2nd International Conference on Mechanical Engineering
Research (ICMER 2013), vol. 50, 2013.

[17] Q. Yang and D. An, “EMD and Wavelet Transform Based Fault Diagnosis
for Wind Turbine Gear Box,” Advances in Mechanical Engineering, vol. 5,
2013.

[18] D. P. Jena and S. N. Panigrahi, “Precise measurement of defect width


in tapered roller bearing using vibration signal,” Measurement, vol. 55,
pp. 39–50, 2014.

[19] H. Bendjama, D. Idiou, K. Gherfi, and Y. Laib, “Selection of Wavelet


Decomposition Levels for Vibration Monitoring of Rotating Machinery,”
in ADVCOMP 2015: The Ninth International Conference on Advanced
Engineering Computing and Applications in Sciences:, (Nice, France),
pp. 96–100, 2015.

[20] J. Chen, J. Pan, Z. Li, Y. Zi, and X. Chen, “Generator bearing fault di-
agnosis for wind turbine via empirical wavelet transform using measured
vibration signals,” Renewable Energy, vol. 89, pp. 80–92, 2016.
References 65

[21] Z. Liu, L. Zhang, and J. Carrasco, “Vibration analysis for large-scale


wind turbine blade bearing fault detection with an empirical wavelet
thresholding method,” Renewable Energy, vol. 146, pp. 99–110, 2020.
[22] J. Antoni and R. B. Randall, “The spectral kurtosis: Application to the
vibratory surveillance and diagnostics of rotating machines,” Mechanical
Systems and Signal Processing, vol. 20, no. 2, pp. 308–331, 2006.
[23] J. Antoni, “Fast computation of the kurtogram for the detection of tran-
sient faults,” Mechanical Systems and Signal Processing, vol. 21, no. 1,
pp. 108–124, 2007.
[24] N. Sawalhi, R. B. Randall, and H. Endo, “The enhancement of fault de-
tection and diagnosis in rolling element bearings using minimum entropy
deconvolution combined with spectral kurtosis,” Mechanical Systems and
Signal Processing, vol. 21, no. 6, pp. 2616–2633, 2007.
[25] Y. Lei, J. Lin, Z. He, and Y. Zi, “Application of an improved kurtogram
method for fault diagnosis of rolling element bearings,” Mechanical Sys-
tems and Signal Processing, vol. 25, no. 5, pp. 1738–1749, 2011.
[26] L. Gelman, T. H. Patel, G. Persin, B. Murray, and A. Thomson, “Novel
Technology Based on the Spectral Kurtosis and Wavelet Transform for
Rolling Bearing Diagnosis,” International Journal of Prognostics and
Health Management, vol. 4, no. 2, pp. 1–7, 2013.
[27] Y. Kong, T. Wang, Z. Li, and F. L. Chu, “Fault feature extraction of
planet gear in wind turbine gearbox based on spectral kurtosis and time
wavelet energy spectrum,” Frontiers of Mechanical Engineering, vol. 12,
no. 3, pp. 406–419, 2017.
[28] J. Rafiee, F. Arvani, A. Harifi, and M. H. Sadeghi, “Intelligent condi-
tion monitoring of a gearbox using artificial neural network,” Mechanical
Systems and Signal Processing, vol. 21, no. 4, pp. 1746–1754, 2007.
[29] Q. Huang, D. Jiang, L. Hong, and Y. Ding, “Application of Wavelet Neu-
ral Networks on Vibration Fault Diagnosis for Wind Turbine Gearbox,”
in Advances in Neural Networks (F. Sun, J. Zhang, Y. Tan, J. Cao, and
W. Yu, eds.), pp. 313–320, Berlin, Germany: Springer Berlin Heidelberg,
2008.
[30] C. Murray, M. Asher, N. Lieven, M. Mulroy, C. Ng, and P. Morrish,
“Wind turbine drivetrain health assessment using discrete wavelet trans-
forms and an artificial neural network,” in 3rd Renewable Power Gener-
ation Conference, RPG 2014, (Napels, Italy), pp. 1–5, 2014.
[31] H. Heidari Bafroui and A. Ohadi, “Application of wavelet energy and
Shannon entropy for feature extraction in gearbox fault detection under
varying speed conditions,” Neurocomputing, vol. 133, pp. 437–445, 2014.
66 References

[32] C. Wang, M. Gan, and C. Zhu, “Intelligent fault diagnosis of rolling ele-
ment bearings using sparse wavelet energy based on overcomplete DWT
and basis pursuit,” Journal of Intelligent Manufacturing, vol. 28, no. 6,
pp. 1377–1391, 2017.

[33] C. Wang, M. Gan, and C. Zhu, “Fault feature extraction of rolling element
bearings based on wavelet packet transform and sparse representation
theory,” Journal of Intelligent Manufacturing, vol. 29, no. 4, pp. 937–
951, 2018.

[34] M. Gan, C. Wang, and C. Zhu, “Construction of hierarchical diagnosis


network based on deep learning and its application in the fault pattern
recognition of rolling element bearings,” Mechanical Systems and Signal
Processing, vol. 72-73, pp. 92–104, 2016.

[35] I. El-Thalji and E. Jantunen, “A summary of fault modelling and predic-


tive health monitoring of rolling element bearings,” Mechanical Systems
and Signal Processing, vol. 60-61, pp. 252–272, 2015.

[36] W. A. Smith and R. B. Randall, “Rolling element bearing diagnostics


using the case western reserve university data: A benchmark study,”
Mechanical Systems and Signal Processing, vol. 64, pp. 100–131, 2015.

[37] H. D. M. de Azevedo, A. M. Araújo, and N. Bouchonneau, “A review


of wind turbine bearing condition monitoring: State of the art and chal-
lenges,” Renewable and Sustainable Energy Reviews, vol. 56, pp. 368–379,
2016.

[38] Z. Liu and L. Zhang, “A review of failure modes, condition monitor-


ing and fault diagnosis methods for large-scale wind turbine bearings,”
Measurement: Journal of the International Measurement Confederation,
vol. 149, pp. 1–22, 2020.

[39] J. Singh, M. Azamfar, F. Li, and J. Lee, “A systematic review of machine


learning algorithms for PHM of rolling element bearings: fundamentals,
concepts, and applications,” Measurement Science and Technology, 2020.

[40] T. J. Price, “James Blyth - Britain’s first modern wind power pioneer,”
Wind Engineering, vol. 29, no. 3, pp. 191–200, 2005.

[41] W. Qiao and D. Lu, “A survey on wind turbine condition monitoring and
fault diagnosis - Part II: signals and signal processing methods,” IEEE
Transactions on Industrial Electronics, vol. 62, no. 10, pp. 6546–6557,
2015.

[42] D. Dowson, History of Tribology. London, UK: Professional Engineering


Publishing, 2nd ed., 1998.
References 67

[43] X. Ai, “Rolling Element Bearings, History,” in Encyclopedia of Tribol-


ogy (Q. J. Wang and Y.-W. Chung, eds.), pp. 2932–2937, Boston, MA:
Springer US, 2013.

[44] T. A. Harris and M. N. Kotzalas, Essential Concepts in Bearing Tech-


nology. Boca Raton, FL: CRC Press, 5th ed., 2006.

[45] T. A. Harris and M. N. Kotzalas, Advanced Concepts in Bearing Tech-


nology. Boca Raton, FL: CRC Press, 5th ed., 2006.

[46] J. Halme and P. Andersson, “Rolling contact fatigue and wear fundamen-
tals for rolling bearing diagnostics - state of the art,” Proc. IMechE, Part
J: Journal of Engineering Tribology, vol. 224, no. 4, pp. 377–393, 2010.

[47] X. Ai, “Failure Mechanisms of Rolling Element Bearings,” in Encyclope-


dia of Tribology (Q. J. Wang and Y.-W. Chung, eds.), pp. 1006—-1014,
Boston, MA: Springer US, 2013.

[48] A. Greco, S. Sheng, J. Keller, and A. Erdemir, “Material wear and fatigue
in wind turbine Systems,” Wear, vol. 302, no. 1-2, pp. 1583–1591, 2013.

[49] ISO Central Secretary, “Rolling bearings - damage and failures - terms,
characteristics and causes,” Standard ISO 15243:2017, International Or-
ganization for Standardization, Geneva, CH, 2017.

[50] W. H. Hannon, “Rolling Bearing Condition Monitoring,” in Encyclope-


dia of Tribology. (Q. J. Wang and Y.-W. Chung, eds.), pp. 2812–2820,
Boston, MA: Springer US., 2013.

[51] D. Dyer and R. M. Stewart, “Detection of Rolling Element Bearing Dam-


age by Statistical Vibration Analysis,” Journal of Mechanical Design,
vol. 100, pp. 229–235, 1978.

[52] H. R. Martin and F. Honarvar, “Application of statistical moments to


bearing failure detection,” Applied Acoustics, vol. 44, pp. 67–77, 1995.

[53] F. Honarvar and H. R. Martin, “New statistical moments for diagnos-


tics of rolling element bearings,” Journal of Manufacturing Science and
Engineering, vol. 119, no. 3, pp. 425–432, 1997.

[54] P. D. McFadden and J. D. Smith, “Vibration monitoring of rolling el-


ement bearings by the high-frequency resonance technique - a review,”
Tribology International, vol. 17, no. 1, pp. 3–10, 1984.

[55] A. Palmgren, Ball and Roller Bearing Engineering. SH Burbank & Co.
Philadelphia, PA: SH Burbank & Co. Inc., 1947.

[56] T. A. Harris, Rolling bearing analysis. New York: John Wiley & Sons
Inc., 1966.
68 References

[57] P. K. Gupta, “Transient Ball Motion and Skid in Ball Bearings,” Journal
of Lubrication Technology, vol. 97, no. 2, pp. 261–269, 1975.
[58] S. Fukata, E. H. Gad, T. Kondou, T. Ayabe, and H. Tamura, “On the
radial vibration of ball bearings,” JSME, vol. 28, no. 239, pp. 899–904,
1985.
[59] L. D. Meyer, F. F. Ahlgren, and B. Weichbrodt, “An analytic model
for ball bearing vibrations to predict vibration response to distributed
defects,” Journal of Mechanical Design, vol. 102, no. April, pp. 205–210,
1980.
[60] N. Tandon and A. Choudhury, “an Analytical Model for the Prediction
of the Vibration Response of Rolling Element Bearings Due To Localized
Defect,” Journal of Sound and Vibration, vol. 205, no. 3, pp. 275–292,
1997.
[61] A. Choudhury and N. Tandon, “A Theoretical Model to Predict Vibration
Response of Rolling Bearings to Distributed Defects Under Radial Load,”
Journal of Tribology, vol. 120, pp. 214–220, 1998.
[62] J. Liu, H. Wu, and Y. Shao, “A theoretical study on vibrations of a ball
bearing caused by a dent on the races,” Engineering Failure Analysis,
vol. 83, no. October 2017, pp. 220–229, 2018.
[63] J. Sopanen and A. Mikkola, “Dynamic model of a deep-groove ball bear-
ing including localized and distributed defects. Part 1: theory,” Proc.
IMechE, Part K: Journal of Multi-body Dynamics, vol. 217, pp. 201–211,
2003.
[64] J. Sopanen and A. Mikkola, “Dynamic model of a deep-groove ball bear-
ing including localized and distributed defects. Part 2 : implementation
and results,” Proc. IMechE, Part K: Journal of Multi-body Dynamics,
vol. 217, pp. 213–223, 2003.
[65] N. Sawalhi and R. B. Randall, “Simulating gear and bearing interactions
in the presence of faults Part I: The combined gear bearing dynamic
model and the simulation of localised bearing faults,” Mechanical Systems
and Signal Processing, vol. 22, pp. 1924–1951, 2008.
[66] N. Sawalhi and R. B. Randall, “Simulating gear and bearing interactions
in the presence of faults Part II: Simulation of the vibrations produced
by extended bearing faults,” Mechanical Systems and Signal Processing,
vol. 22, pp. 1952–1966, 2008.
[67] F. T. Wang, M. Q. Jing, J. Yi, G. H. Dong, H. Liu, and B. W. Ji,
“Dynamic modelling for vibration analysis of a cylindrical roller bearing
due to localized defects on raceways,” Proc. IMechE, Part K: Journal of
Multi-body Dynamics, vol. 229, no. 1, pp. 39–64, 2015.
References 69

[68] D. Petersen, C. Q. Howard, N. Sawalhi, A. M. Ahmadi, and S. Singh,


“Analysis of bearing stiffness variations, contact forces and vibrations in
radially loaded double row rolling element bearings with raceway defects,”
Mechanical Systems and Signal Processing, vol. 50-51, pp. 139–160, 2015.
[69] H. Liu, L. Wang, and Y. Li, “Simulation of rail wheel axle bearing vibra-
tion due to local damages on outer races,” Proc IMechE, Part K: Journal
of Multi-body Dynamics, vol. 233, no. 2, pp. 429–440, 2019.
[70] F. Kong, W. Huang, Y. Jiang, and W. Wang, “Research on effect of
damping variation on vibration response of defective bearings,” Advances
in Mechanical Engineering, vol. 11, no. 3, pp. 1–12, 2019.
[71] J. Liu, Y. Shao, and T. C. Lim, “Impulse vibration transmissibility char-
acteristics in the presence of localized surface defects in deep groove ball
bearing systems,” Proc. IMechE, Part K: Journal of Multi-body Dynam-
ics, vol. 228, no. 1, pp. 62–81, 2014.
[72] S. Singh, C. Q. Howard, C. H. Hansen, and U. G. Köpke, “Analytical
validation of an explicit finite element model of a rolling element bearing
with a localised line spall,” Journal of Sound and Vibration, vol. 416,
pp. 94–110, 2018.
[73] S. Singh, C. Q. Howard, and C. H. Hansen, “An extensive review of vi-
bration modelling of rolling element bearings with localised and extended
defects,” Journal of Sound and Vibration, vol. 357, pp. 300–330, 2015.
[74] P. K. Gupta, Advanced Dynamics of Rolling Elements. Springer-Verlag
New York, 1984.
[75] L.-E. Stacke, D. Fritzson, and P. Nordling, “BEAST - a rolling bearing
simulation tool,” Proc. IMechE, Part K: Journal of Multi-body Dynamics,
vol. 213, pp. 63–71, 1999.
[76] U. Kumar, A. Ahmadi, A. K. Verma, and P. Varde, Current Trends in
Reliability, Availability, Maintainability and Safety An Industry Perspec-
tive. Springer International Publishing, 2015.
[77] C. Wang and R. X. Gao, “Sensor placement strategy for in-situ bear-
ing defect detection,” Conference Record - IEEE Instrumentation and
Measurement Technology Conference, vol. 3, pp. 1463–1467, 2000.
[78] A. Grossmann and J. Morlet, “Decomposition of Hardy functions into
square integrable wavelets of constant shape,” SIAM Journal on Mathe-
matical Analysis, vol. 15, no. 4, pp. 723–736, 1984.
[79] S. G. Mallat, “A theory for multiresolution signal decomposition: the
wavelet representation,” IEEE transactions on pattern analysis and ma-
chine intelligence, vol. 11, no. 7, pp. 674–693, 1989.
70 References

[80] Y. Meyer, “Wavelets, Time-Frequency Methods and Phase Space,”


in Orthonormal wavelets. (J.-M. Combes, A. Grossmann, and
P. Tchamitchian, eds.), pp. 21–37, Springer-Verlag., 1989.

[81] R. X. Gao and R. Yan, Wavelets, Theory and Applications for Manufac-
turing. Boston, MA: Springer US, 2011.

[82] K. Mori, N. Kasashima, T. Yoshioka, and Y. Ueno, “Prediction of spalling


on a ball bearing by applying the discrete wavelet transform to vibration
signals,” Wear, vol. 195, pp. 162–168, 1996.

[83] S. Khanam, N. Tandon, and J. K. Dutt, “Fault Size Estimation in the


Outer Race of Ball Bearing Using Discrete Wavelet Transform of the
Vibration Signal,” Procedia Technology, vol. 14, pp. 12–19, 2014.

[84] S. Sheng, “Prognostics and Health Management of Wind Turbines -


Current Status and Future Opportunities,” in Probabilistic Prognostics
and Health Management of Energy Systems. (S. Ekwaro-Osire, A. C.
Gonçalves, and F. M. Alemayehu, eds.), pp. 33–47, Cham, Switzerland:
Springer International Publishing., 2017.

[85] J. Rafiee, M. A. Rafiee, and P. W. Tse, “Application of mother wavelet


functions for automatic gear and bearing fault diagnosis,” Expert Systems
with Applications, vol. 37, no. 6, pp. 4568–4579, 2010.

[86] R. Yan and R. X. Gao, “Base Wavelet Selection for Bearing Vibration
Signal Analysis,” International Journal of Wavelets, Multiresolution and
Information Processing, vol. 7, no. 4, pp. 411–426, 2009.

[87] P. K. Kankar, S. C. Sharma, and S. P. Harsha, “Rolling element bear-


ing fault diagnosis using wavelet transform,” Neurocomputing, vol. 74,
pp. 1638–1645, 2011.

[88] P. K. Kankar, S. C. Sharma, and S. P. Harsha, “Fault diagnosis of ball


bearings using continuous wavelet transform,” Applied Soft Computing,
vol. 11, pp. 2300–2312, 2011.

[89] F. Hemmati, W. Orfali, and M. S. Gadala, “Rolling element bearing con-


dition monitoring using acoustic emission technique,” in International
Conference on Noise and Vibration Engineering 2012, ISMA 2012, (Leu-
ven, Belgium), pp. 699–713, 2012.

[90] K. Deák, T. Mankovits, and I. Kocsis, “Optimal wavelet selection for


manufacturing defect size estimation of tapered roller bearings with vi-
bration measurement using Shannon Entropy Criteria,” Journal of Me-
chanical Engineering, vol. 63, no. 1, pp. 3–14, 2016.
References 71

[91] C. Wu, T. Chen, R. Jiang, L. Ning, and Z. Jiang, “A novel approach


to wavelet selection and tree kernel construction for diagnosis of rolling
element bearing fault,” Journal of Intelligent Manufacturing, vol. 28,
no. 8, pp. 1847–1858, 2017.

[92] P. G. Kulkarni and A. D. Sahasrabudhe, “Investigations on mother


wavelet selection for health assessment of lathe bearings,” Interna-
tional Journal of Advanced Manufacturing Technology, vol. 90, no. 9-12,
pp. 3317–3331, 2017.

[93] M. V. Wickerhauser, “Lectures on wavelet packet algorithms,” in Lecture


notes, INRIA, 1991.

[94] N. G. Nikolaou and I. A. Antoniadis, “Rolling element bearing fault di-


agnosis using wavelet packets,” NDT & E International, vol. 35, no. 3,
pp. 197–205, 2002.

[95] E. Mollasalehi, Q. Sun, and D. Wood, “Wind turbine generator bearing


fault diagnosis using amplitude and phase demodulation techniques for
small speed variations,” in Advances in Condition Monitoring of Machin-
ery in Non-Stationary Operations. (F. Chaari, R. Zimroz, W. Bartelmus,
and M. Haddar, eds.), pp. 385–397, Cham, Switzerland: Springer Inter-
national Publishing., 2016.

[96] H. Li, X. Lian, C. Guo, and P. Zhao, “Investigation on early fault clas-
sification for rolling element bearing based on the optimal frequency
band determination,” Journal of Intelligent Manufacturing, vol. 26, no. 1,
pp. 189–198, 2015.

[97] C. Castejón, O. Lara, and J. C. García-Prada, “Automated diagnosis of


rolling bearings using MRA and neural networks,” Mechanical Systems
and Signal Processing, vol. 24, pp. 289–299, 2010.

[98] Z. Zhang, Y. Wang, and K. Wang, “Fault diagnosis and prognosis us-
ing wavelet packet decomposition, Fourier transform and artificial neural
network,” Journal of Intelligent Manufacturing, vol. 24, no. 6, pp. 1213–
1227, 2013.

[99] C. Castejón, J. C. García-Prada, M. J. Gomez, and J. Meneses, “Auto-


matic detection of cracked rotors combining multiresolution analysis and
artificial neural networks,” Journal of Vibration and Control, vol. 21,
no. 15, pp. 3047–3060, 2015.

[100] D. Cabrera, F. Sancho, R.-V. Sánchez, G. Zurita, M. Cerrada, C. Li, and


R. E. Vásquez, “Fault diagnosis of spur gearbox based on random forest
and wavelet packet decomposition,” Frontiers of Mechanical Engineering,
vol. 10, no. 3, pp. 277–286, 2015.
72 References

[101] Z. Wang, Q. Zhang, J. Xiong, M. Xiao, G. Sun, and J. He, “Fault Diag-
nosis of a Rolling Bearing Using Wavelet Packet Denoising and Random
Forests,” IEEE Sensors Journal, vol. 17, no. 17, pp. 5581–5588, 2017.

[102] B. Samanta and K. R. Al-Balushi, “Artificial neural network based fault


diagnostics of rolling element bearings using time-domain features,” Me-
chanical Systems and Signal Processing, vol. 17, no. 2, pp. 317–328, 2003.

[103] M. Unal, M. Onat, M. Demetgul, and H. Kucuk, “Fault diagnosis of


rolling bearings using a genetic algorithm optimized neural network,”
Measurement: Journal of the International Measurement Confederation,
vol. 58, pp. 187–196, 2014.

[104] L. F. de Almeida, J. W. P. Bizarria, F. C. P. Bizarria, and M. H. Mathias,


“Condition-based monitoring system for rolling element bearing using a
generic multi-layer perceptron,” Journal of Vibration and Control, vol. 21,
no. 16, pp. 3456–3464, 2015.

[105] G. Li, C. Deng, J. Wu, Z. Chen, and X. Xu, “Rolling bearing fault
diagnosis based on wavelet packet transform and convolutional neural
network,” Applied Sciences, vol. 10, no. 3, 2020.

[106] A. Widodo and B. S. Yang, “Support vector machine in machine con-


dition monitoring and fault diagnosis,” Mechanical Systems and Signal
Processing, vol. 21, no. 6, pp. 2560–2574, 2007.

[107] A. Soualhi, K. Medjaher, and N. Zerhouni, “Bearing health monitoring


based on hilbert-huang transform, support vector machine, and regres-
sion,” IEEE Transactions on Instrumentation and Measurement, vol. 64,
no. 1, pp. 52–62, 2015.

[108] J. Saari, D. Strömbergsson, J. Lundberg, and A. Thomson, “Detection


and identification of windmill bearing faults using a one-class support
vector machine (SVM),” Measurement: Journal of the International Mea-
surement Confederation, vol. 137, pp. 287–301, 2019.

[109] J. Tautz-Weinert and S. J. Watson, “Using SCADA data for wind tur-
bine condition monitoring - a review,” IET Renewable Power Generation,
vol. 11, no. 4, pp. 382–394, 2017.

[110] W. Yang, R. Court, and J. Jiang, “Wind turbine condition monitoring


by the approach of SCADA data analysis,” Renewable Energy, vol. 53,
pp. 365–376, 2013.

[111] P. Bangalore, S. Letzgus, D. Karlsson, and M. Patriksson, “An artificial


neural network-based condition monitoring method for wind turbines,
with application to the monitoring of the gearbox,” Wind Energy, vol. 20,
no. 8, pp. 1421–1438, 2017.
References 73

[112] D. E. Rumelhart, G. E. Hinton, and R. J. Williams, “Learning internal


representations by error propagation,” tech. rep., California Univ. San
Diego La Jolla Inst. for Cognitive Science, 1985.
[113] S. Struggl, V. Berbyuk, and H. Johansson, “Review on wind turbines with
focus on drive train system dynamics,” Wind Energy, vol. 18, pp. 567–
590, 2015.
Department of Engineering Sciences and Mathematics
Division of Machine Elements

ISSN 1402-1544
ISBN 978-91-7790-621-6 (print)
ISBN 978-91-7790-622-3 (pdf)

Luleå University of Technology 2020

You might also like