Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Modeling, Identification and Control, Vol. 43, No. 1, 2022, pp.

21–38, ISSN 1890–1328

RMS Based Health Indicators for Remaining


Useful Lifetime Estimation of Bearings
A. Klausen 1 H.V. Khang 1 K.G. Robbersmyr 1

1
Department of Engineering Sciences, Mechatronics Group, University of Agder, N-4879 Grimstad, Norway. E-
mail: kjell.g.robbersmyr@uia.no

Abstract

Estimating the remaining useful life (RUL) of bearings from healthy to faulty is important for predictive
maintenance. The bearing fault severity can be estimated based on the energy or root mean square
(RMS) of vibration signals, and a stopping criterion can be set based on a threshold given by an ISO
standard. However, the vibration RMS is often not monotonically increasing with damage, which renders
a challenge for predicting the RUL. This study proposes a novel method for splitting the vibration signal
into multiple frequency bands before RMS calculations to generate multiple health indicators. Monotonic
health indicators are identified using the Spearman coefficient, and the RUL is afterward estimated for each
indicator using a suitable model and parameter update scheme. Historical failure data is not required to
set any parameters. The proposed method is tested with the Paris’ law, where parameters are updated by
particle filters. Experimental results from two test rigs validate the performance of the proposed method.

Keywords: Ball bearings, Remaining useful life, Particle filter, Paris’ law, Vibration measurement

1 Introduction (2016b), whitening methods Peeters et al. (2017), and


minimum entropy deconvolution Abboud et al. (2019).
Bearings are widely used in rotating machines, and Continued machine operation after detecting the initial
bearing defects result in increased vibration, tempera- bearing fault is beneficial for planning maintenance or
ture, and friction. Up to 44% of failures in the most even necessary if the machine should halt in a con-
common motors, namely induction motors, are due trolled manner. As such, an estimation of the bearing
to bearing faults Zhang et al. (2011). The vibration remaining useful life (RUL) is crucial to select between
from severe bearing faults may damage other machine regulated or emergency stop. Here follows a general
components, such as gears, stators, and pump seals. methodology for estimating the RUL Lei et al. (2018):
Unscheduled stops can cause extended downtimes and
1. Produce a health indicator (HI) that correlates to
huge expenses due to maintenance and productivity
the component wear
losses. Monitoring the bearing health condition is
essential to avoid emergency shutdowns and to plan 2. Divide the HI into multiple health stages (HSs)
maintenance.
3. Identify the transition from the first to the second
A condition monitoring (CM) system can detect
HS
faults and enhance machine reliability. In rotating ma-
chines, such systems can detect faults using vibration 4. Predict future HI trend using a suitable model
signals, typically by using the envelope spectrum Mc-
Fadden and Smith (1984). More advanced methods 5. Determine the remaining time left, i.e. RUL, until
for fault diagnosis apply machine learning Lu et al. the prediction reaches a failure threshold (FT)

doi:10.4173/mic.2022.1.3 c 2022 Norwegian Society of Automatic Control


Modeling, Identification and Control

The bearing HI can be assessed by examining the require training data from previous failures, which may
level of wear on the rollers and raceways. However, not always be available.
this wear is impractical to determine, as an offline in- The estimated RUL is the predicted time until the
spection is required after dissembling the bearing Lei model reaches an FT, hence determining this thresh-
et al. (2018), resulting in productivity loss. Instead, old accurately is essential for the estimation accuracy.
the HI can be estimated using less invasive methods. For VHIs, historical failure data from a similar setup
The vibration signal acquired using an accelerometer is often necessary to create a suitable FT Khan et al.
is a good substitute, as wear on the rollers or raceways (2018); Zhu et al. (2019); Lei et al. (2016a). However,
increases the vibration energy Singleton et al. (2017). for machines where no historical failure data is avail-
HIs can be categorized into physical HIs (PHIs) and able, many of the methods reported in the literature
virtual HIs (VHIs) Hu et al. (2012). PHIs are gener- may not be applicable. In Klausen et al. (2018), the
ated from primarily physical signals and are directly FT is set based on the vibration RMS as guided by the
related to the physics of failure. Examples are the vi- ISO standard 10816-3 ISO (1998), without involving
bration root mean square (RMS) Li et al. (2015), kur- historical failure data.
tosis Lei et al. (2016b), and characteristic bearing fault Estimating the RUL using vibration RMS is difficult
frequency amplitudes Gebraeel et al. (2004). On the as this HI does not always increase monotonically with
other hand, VHIs do not correlate directly with the damage. There are methods for extracting the hidden
physics of failure Hu et al. (2012) and can be calcu- degradation state of a system using phase-space warp-
lated by processing multiple PHIs. The Mahalanobis ing Qian et al. (2017). However, the proposed RUL
distance was applied in Wang et al. (2016) to fuse 14 estimation method in Qian et al. (2017) requires set-
PHIs into a single VHI, and principal component anal- ting parameters using historical failure data. Another
ysis (PCA) was used in Lu et al. (2016a) to estimate method Qian and Yan (2015) uses enhanced particle
the principal component of multiple PHIs. filter to estimate the RUL, but the failure threshold
The bearing degradation can be divided into two or for the RP entropy is set without justification. Arti-
multiple HSs Lei et al. (2018). During the first HS, ficial intelligence methods, and most particularly deep
there is no apparent degradation increase in the vi- learning methods, have also been used to estimate the
bration signal, while the signal may increase rapidly RUL of bearings Akkad and He (2019); Ma and Mao
during proceeding ones Wang (2002). This transition (2019); Wang et al. (2019); Cheng et al. (2020); Pan
can be detected using baseline measurements of the vi- et al. (2020). However, these methods require super-
bration RMS signal Wang (2002) and kurtosis value Li vised learning techniques using historical failure data
et al. (2015). For example, the HI mean and standard to train the model. Consequently, there is a challenge
deviation can be determined during baseline measure- in estimating the RUL on systems where historical fail-
ments, and a substantial deviation suggests the transi- ure data is not available.
tion from the first to the second HS.
This paper proposes a method to address the chal-
After transitioning to the second HS, the HI can be lenges above for bearing RUL estimation. The phys-
predicted with a mathematical model that resembles ical meaning of the RUL estimation is preserved by
the failure physics. Analytic models such as the expo- basing the HIs on RMS, and the FTs are based on
nential model Li et al. (2015); Wang and Tsui (2017), ISO 10816-3 ISO (1998). Historical failure data is not
Brownian motion Wang et al. (2016), and Paris’ law Lei needed to determine either monotonic VHIs or FTs in
et al. (2016a) have been used to model the HI. The the proposed method. The vibration signal is first di-
Kalman filter (KF) Singleton et al. (2015) and parti- vided into multiple frequency bands to construct mul-
cle filter (PF) Qian and Yan (2015) can be applied to tiple RMS-based HIs. Instead of using a digital filter
update parameters for the models. Also, numerical op- bank, the proposed approach utilizes only the result-
timization can be used to minimize the error between ing bins of a discrete Fourier transform (DFT) of the
measurements and a model Ahmad et al. (2018). vibration signal. This approach yields no energy loss
As an alternative to analytic models, data-driven and is less computationally expensive than digital finite
methods can perform predictions without knowing the impulse response (FIR) filters. The Spearman coeffi-
physics of failure, such as the Gaussian process model cient is used to determine which of the HIs are suitable
(GPM) Aye and Heyns (2017), least squares support for RUL estimation. FTs for the generated HIs are
vector machine (LSSVM) Lu et al. (2016a); Manjurul computed by extending the method in Klausen et al.
Islam et al. (2021) and transfer learning method ap- (2018). Statistically significant changes in each HI are
plied to a multiple layer perceptron network Zhu et al. compared to baseline measurements to determine the
(2020). These models are then used to predict the fu- transition from the first to second HS. For each mono-
ture trend of the HI. However, most data-driven models tonic HI, the RUL is estimated by projecting the future

22
Klausen et.al., “RMS Based Health Indicators for RUL Estimation of Bearings”

trend using a suitable model. The RUL estimation for lower than 2 Hz. Afterwards, the acceleration signal
each trend is then finally combined to a single weighted is integrated digitally using the trapezoidal rule:
RUL estimation, that is more accurate than either es- (
timation alone. It should be noted that the proposed 0 i=0
xv,i = xi−1 +xi (3)
method correlates energy increase/decrease to varia- xv,i−1 + dt else
2
tions in component health, and does not change in
operating conditions. Hence, for variable speed/load where xv,i is the velocity signal at a discrete time step
operations, other changes also need to be considered. i. The mean value of xv is finally removed after in-
Vibration signals from two bearing degradation tests tegration. For brevity, the following notation is used:
are used to validate the proposed method. PFs are Ra = RMS(x) and Rv = RMS(xv ).
used to update model parameters of the Paris’ law to In many practical cases, when monitoring bearing
predict the future trend of HIs. faults, Rv is not suitable for RUL estimation. Ini-
The rest of the paper is organized as follows. Practi- tial bearing defects generate high-frequency resonance
cal differences between velocity- and acceleration-based vibration, and integration reduces the effect of high-
RMS are discussed in Section 2. Next, the proposed frequency components Klausen et al. (2018). For that
approach for subdividing the vibration signal into mul- reason, the acceleration-based RMS value is better
tiple RMS HIs is detailed in Section 3. Afterwards, the suited as an HI. However, the ISO standard ISO (1998)
algorithm used for RUL estimation is presented in Sec- is only defined for a velocity-based RMS signal. To al-
tion 4, and the experimental results from two test rigs leviate this limitation, the FT is experimentally trans-
are given in Section 5. Finally, conclusions are drawn formed into the acceleration domain before estimating
in Section 6. bearing RUL. This procedure was proposed in Klausen
et al. (2018) and is briefly described in the following
paragraph.
2 Velocity versus acceleration RMS The mean velocity and acceleration RMS values are
first calculated at a known steady state (baseline) and
ISO Standard 10816-3 ISO (1998) defines levels of
are denoted as R̄v and R̄a , respectively. Afterwards,
velocity-based vibration RMS values for rotating ma-
the ratio between the velocity-based FT and the base-
chines with the rating over 15 kW. Four levels are
line is used to scale the acceleration-based baseline into
given, ranging between A - D: A =“standard accep-
a suitable threshold R̂a using
tance for new machines”; B =“unlimited operation is
possible”; C =“short term operation allowed”; D =“vi- R̂v
bration causes damage”. A machine should stop before R̂a = R̄a (4)
R̄v
reaching level D. Hence, the boundary between levels
C and D can be used as a FT. where the value of R̂v is taken from the ISO standard.
The RMS of a sampled vibration signal x can be Vibration data collected by the NSF I/UCR Center
calculated with for Intelligent Maintenance Systems (IMS) Qiu et al.
v
u n
u1 X (2006) is used as a numerical example of this calcu-
RMS(x) = t x2 (1) lation. Four Rexnord ZA-2115 double row bearings
n i=1 i are run to failure by applying a constant load of 26.6
kN at a constant speed of 2000 revolutions per minute
where n is the vibration signal length, and xi is the i’th
(rpm). The test is stopped when the circulating oil is
vibration sample. The vibration signal in this study is
thoroughly contaminated with metal debris. Every 10
measured with an accelerometer, which gives the vi-
minutes, 1 second of vibration data was measured on
bration in acceleration units (m/s2 ). Through inte-
each bearing housing with a sample rate of 20480 Hz.
gration, the velocity signal xv (t) is acquired, which is
An index k is defined to keep track of which 1-second
used for checking vibration severity according to ISO
vibration signal is in use. Accelerometer data from the
10816-3 ISO (1998). Before integrating, the vibration
second run-to-failure test, consisting of 982 vibration
signal is filtered with a first-order infinite impulse re-
measurements, are used to calculate the acceleration-
sponse (IIR) high-pass filter. The cut-off frequency is
based failure threshold. An outer race fault was ob-
set based on shaft speed fs ISO (1998):
( served in bearing number 1 after this test.
10 Hz if fs ≥ 10 Hz The first 30 hours of vibration data (k ∈ [1, 180] vi-
Cut-off frequency = (2) bration signals) are used as a baseline, and the mean
2 Hz if 10 > fs ≥ 2 Hz
RMS values are R̄a = 0.757 m/s2 and R̄v = 0.755
It should be noted that the standard ISO (1998) does mm/s. The motor size is assumed smaller than 15 kW,
not define a cut-off frequency for shaft frequencies and therefore the FT is set slightly lower than level D

23
Modeling, Identification and Control

5 3 Proposed RMS health indicator


4
3.1 Filter bank RMS
RMS [mm/s]

3 Rv
R̂v
A digital filter bank can be used to subdivide a sig-
2 nal into multiple components, where each component
contains a frequency sub-band of the signal. The filter
1
bank can be created by iteratively passing the signal
0 through a low- and high-pass FIR filter and decimating
0 20 40 60 80 100 120 140 160 each output signal to half the signal frequency. Such
Time [hours]
(a) a procedure has some drawbacks: Digital FIR filters
8 use kernel convolution to filter the signal, and only the
Ra
R̂a
overlapping part between the filter kernel and the sig-
6 nal should be preserved to avoid adding signal artifacts.
RMS [m/s2 ]

A high order FIR filter kernel would then remove much


4 of the signal energy, and the loss is exponential for each
level of filtering. However, a low order FIR filter ker-
2 nel results in a considerable frequency overlap between
the signal components. Computing a digital filter bank
0 is also computationally taxing if the input signal is
0 20 40 60 80 100 120 140 160
long, and multiple frequency levels are required. Such
Time [hours]
(b) a filter-bank is, however, not required to calculate the
RMS in different frequency regions. The following ex-
Figure 1: Comparison between velocity- and plains how the vibration RMS in a particular frequency
acceleration-based RMS for the IMS band can be determined with the DFT spectrum bins.
dataset. (a) velocity-based RMS and its FT; An alternative representation of RMS is given by the
(b) acceleration-based RMS and its FT. energy E of the signal, such as
r
E(x)
RMS(x) = (5)
T
where T is the vibration signal length in seconds. The
signal energy can be calculated in both time and fre-
in the ISO standard ISO (1998), at R̂v = 4 mm/s. Us- quency domain with
ing (4) results in R̂a = 4.01 m/s2 . Fig. 1 (a) shows Rv n n
and R̂v , while Fig. 1 (b) shows Ra and the resulting E(x) =
X
|xi |2 ∆t, E(X) =
X
|Xi |2 ∆f (6)
FT R̂a . As observed in Fig. 1 (a), Rv starts increas- i=1 i=1
ing from the mean value near the end of useful life (t ≈
160 hours), giving a short time to stop the machine and where ∆t is the vibration sampling period, X =
plan maintenance. However, Ra starts increasing much (X1 , . . . , Xn ) is the DFT, and ∆f is the frequency step
sooner (t ≈ 90 hours), which better allows for predict- between each DFT bin.
ing RUL and plan maintenance. The two trends reach To explain the procedure, the DFT is assumed or-
their respective FTs near the end of life, which shows dered from 0 Hz to Nyquist frequency, and from nega-
that the transformation can be used for this purpose. tive Nyquist up to 0 Hz. Let XL = (XLP , Zn/2 , XLN ),
where XLP contains the DFT values of the lower pos-
itive frequencies (i.e. 0 Hz to half the Nyquist fre-
Ra in Fig. 1 (b) is, however, not suitable for RUL quency), Zn/2 contains n/2 zeros, and XLN contains
estimation due to non-monotonic oscillations. For ex- the DFT values of the lower negative frequencies. Sim-
ample, the sudden increase of Ra at t ≈ 117 hours, and ilarly, let XH = (Zn/4 , XHP , XHN , Zn/4 ), where XHP
the subsequent decrease, render a challenge for deter- and XHN are the DFT values of higher positive and
mining the actual degradation trend. A more stable negative frequencies, respectively. This arrangement
HI should be extracted from the vibration signal to represents a single level filter bank that splits X at half
achieve better RUL estimation. In the next section, Nyquist frequency completely. Because of the added
a new approach for extracting more useful RMS data zeros in the array, the two components XH and XL
from the vibration signal is proposed. can be directly summed element-wise to re-create X.

24
Klausen et.al., “RMS Based Health Indicators for RUL Estimation of Bearings”

The total energy can afterwards be calculated as the energy of bins belonging to frequency band i is used
n to calculate Ri,k using (11). The next step is to obtain
FTs for all Ri .
X
E(X) = E(XL + XH ) = |XL,i + XH,i |2 ∆f (7)
i=1

where XL,i and XH,i are the i’th bin of XL and XH ,


3.2 Filter bank RMS failure thresholds
respectively. Let the spectrum bins be given by their To get an FT for each Ri , (12) is reconsidered such
complex values, i.e. XL,i = aL,i + jbL,i , XH,i = aH,i + that v
jbH,i , then (7) becomes u Nb
uX
R̂a = t R̂i2 (13)
n i=1
X
E(X) = |aL,i + aH,i + j(bL,i + bH,i )|2 ∆f (8)
where R̂i is the FT defined for Ri . To solve (13) for
i=1
n
any number of frequency bands, it is necessary to make
some assumptions about the vibration signal. It is as-
X
2 2
= aL,i + 2aL,i aH,i + aH,i
i=1
sumed that the RMS values can be modeled as nor-
mally distributed noise with a mean value µi and stan-
+ b2L,i + 2bL,i bH,i + b2H,i ∆f

(9)
dard deviation σi during baseline measurements, such
Given that XL and XH represent a complete sig- that
nal separation at half Nyquist frequency, the overlap Ribaseline ∼ N (µi , σi2 ) (14)
between XL and XH is zero. Hence, 2aL,i aH,i = Bearing wear increases vibration energy at the reso-
2bL,i bH,i = 0, and (9) is reduced to nance frequencies of the bearing and the machine it-
n
X self. There are possibly many resonance regions on
E(X) = (a2L,i + a2H,i + b2L,i + b2H,i )∆f a rotating machine Klausen et al. (2017b), hence the
i=1 increased energy spreads over multiple Ri ’s. Some fre-
n
X quency bands may have relatively high energy before
= (|aL,i + jbL,i |2 )∆f bearing wear, while others have low ones. Therefore,
i=1 the FTs are based on scaling from the initial baseline
n
X value. It is also assumed that this energy increase from
+ (|aH,i + jbH,i |2 )∆f bearing wear is proportional to the baseline value of Ri .
i=1
With these assumptions, the FTs are defined as
= E(XL ) + E(XH ) (10)
R̂i = mµi (15)
Eq. (10) shows that the energy of the entire signal can
be calculated by the energy of separate DFT bins. In where m is the constant scaling factor. To determine
effect, the RMS value within a particular frequency m, the conservation of energy is considered, such that
band can be calculated directly from the DFT, without R̂i in (15) is substituted into (13) resulting in
any loss of energy. This information is used to split the
v
vibration signal into equally-sized frequency bands and u Nb
uX
calculate the RMS value for each of them using R̂a = t (mµi )2 (16)
r i=1
E(Xb,i ) Nb
Ri = RMS(Xb,i ) = (11)
T
X
R̂a2 = m2 µ2i (17)
where Xb,i contains the DFT bins within frequency i=1
band i. Since the energy of each frequency band can be
Eq. (17) is finally solved for m as
summed to the total vibration energy, the same applies
for RMS such that v ,
u
u XNb
µ2i
v tR̂2
u Nb
uX m = a (18)
RMS(X) = t R 2 (12) i=1
i
i=1
With the acquired m, the resulting FTs for each RMS
where Nb is the number of equally-sized frequency band are determined with (15).
bands. The rest of this section contains an example of the
In summary, the DFT X is first obtained for the proposed RMS filter bank. The IMS dataset intro-
entire vibration signal x at time index k, and afterward, duced in Section 2 is subdivided into equally sized

25
Modeling, Identification and Control

R12 R9 1.25 R5 R2
2.0 1.5 1.0
R̂12 R̂9 1.00 R̂5 R̂2
0.8
RMS [m/s2 ]

RMS [m/s2 ]

RMS [m/s2 ]

RMS [m/s2 ]
1.5
1.0 0.75
0.6
1.0
0.50 0.4
µ12 + 5σ12 µ9 + 5σ9 µ5 + 5σ5 µ2 + 5σ2
0.5
0.5 0.25 0.2

0.0 0.0 0.00 0.0


0 40 80 120 160 0 40 80 120 160 0 40 80 120 160 0 40 80 120 160
Time [hours] Time [hours] Time [hours] Time [hours]
(a) (b) (c) (d)

Figure 2: Collection of RMS trends with FTs for the IMS dataset. (a) R12 [4400, 4800] Hz; (b) R9 [3200, 3600]
Hz; (c) R5 [1600, 2000] Hz; (d) R2 [400, 800] Hz.

bands with 400 Hz bandwidth (Nb = 25), and Ri is cal- where cov(·, ·) is the covariance of two trends, rank(·)
culated for each band using (11). The mean value and is the rank of a signal, and STD(·) is the standard
standard deviation of each Ri are calculated using the deviation. The running Spearman coefficient for Ri at
baseline vibration values within the first 30 hours (k ∈ time index k is defined as
[1, 180]). With these values, the FTs for each Ri are (
calculated using (15) after determining m with (18). Spearman(Ri,kd,i :k , t) if k > kd,i
ρi,k = (20)
Additionally, the initial degradation alarm is triggered 0 else
at time index kd,i when Ri,kd,i > µi + 5σi Qian et al.
(2014). where Ri,kd,i :k = (Ri,kd,i , . . . , Ri,k ). Eq. (20) is used to
Fig. 2 shows four RMS trends acquired from differ- continually check whether Ri increases monotonously
ent frequency ranges. Fig. 2 (a) shows R12 , contain- over time. The next section describes how to identify
ing RMS from frequency band [4400, 4800] Hz, and monotonic RMS trends.
the oscillations are similar to Ra shown in Fig. 1 (b).
On the other hand, R2 ([400, 800] Hz) in Fig. 2 (d) 3.4 RMS trend selection
is monotonically increasing after the initial alarm is
triggered, thus suitable for RUL estimation. The two The Spearman coefficient is used to describe the mono-
other RMS trends in Figs. 2 (b) and (c) share sim- tonicity of RMS trends. These values are used to de-
ilarities with the two other trends. R12 triggers the termine which of the RMS trends are useful for RUL
initial degradation alarm first, i.e., the transition from estimation. Three criteria must be met by a certain Ri
the first to second HS. This observation supports the at time index k to be used for RUL estimation:
model that initial bearing damage is the most promi- 1. k ≥ kd,i + ns , where ns is the minimum number
nent in the high-frequency region. Further, the FTs, of samples used to calculate the Spearman coeffi-
shown as blue lines, are reached near the end of useful cient.
life, and are therefore considered appropriate for RUL
estimation. However, it is not simple to know which 2. ρi,k ≥ ρ̂, where ρ̂ is the Spearman threshold.
frequency band is best suited for RUL estimation be-
3. i < min(Ip ), where Ip is a set of indices i belonging
forehand. The next section details how suitable RMS
to Ri trends previously passing these criteria. Ip
trends can be selected online and used to estimate the
initially starts as an empty list.
RUL.
The reasons for these criteria are as follows. 1) Several
samples are necessary to calculate a stable Spearman
3.3 Spearman coefficient coefficient, and ns is set as the minimum number of
The Spearman correlation coefficient Spearman (1904) samples. In practice, ns should be based on the num-
is a rank-based correlation between two signals, being ber of samples acquired within a time frame signifi-
used to describe how monotonic a signal is over time. cantly shorter than the assumed time from initial fault
The Spearman coefficient between a signal x and time until breakdown. This time should be estimated based
t is calculated as on the bearing load and shaft speed. 2) The running
Spearman coefficient must be higher than the threshold
cov(rank(x), rank(t)) ρ̂ to accept monotonic trends. The threshold should be
Spearman(x, t) = (19) at least 0.5 to capture most of the monotonic trends,
STD(rank(x))STD(rank(t))

26
Klausen et.al., “RMS Based Health Indicators for RUL Estimation of Bearings”

Figure 3: Flowchart of the proposed method. The stapled line indicates the proposed method for identifying
monotonic RMS trends and calculating the weighted RUL.

even if the signal-to-noise ratio is low. This value can them. There is an opportunity to combine multiple es-
be set higher to track fewer, more influential trends. 3) timations into a single, more accurate estimation. The
Bearing energy increases more in low-frequency bands next subsection shows how these RUL estimations can
during advanced fault stage, as can be evidenced since be combined by utilizing the Spearman coefficient as a
Rv in Fig. 1 (a) increases rapidly near the end of useful weight.
life. In effect, it is assumed that RMS trends of high-
frequency bands trigger the initial degradation alarm 3.6 Weighted remaining useful lifetime
first, and the energy increases at low-frequency bands
later. The index i of new trends must be smaller than The vibration signal is split into Nb frequency bands,
all the other trends previously passing these three cri- where each band is used to calculate Ri . Theoretically,
teria since they add more useful information during ad- up to Nb trends can be monitored, and all RUL esti-
vanced bearing fault stage. The next section describes mations should be used in a weighted decision. Since
how the RUL estimation from multiple RMS trends monotonic trends are better suited for RUL estimation,
can be combined. these get a higher weight. Then, let the weight of each
Ri be
Wi,k = ρ3i,k − ρ3L 1 − ρ3L
 
(21)
3.5 Remaining useful lifetime estimation
where the Spearman coefficient is cubed to prioritize
The RUL is estimated after capturing each vibration monotonic trends, and ρL is a limiting value.
signal. There are many reported methods in the liter- Afterwards, the weighted RUL ¯lk PDF at time index
ature for tracking a monotonic trend and determin- k is defined as
ing the time until failure by forward propagating a
Nb
degradation model, such as the Paris’ crack propaga- X
p(¯lk | R1:Nb ) = Wi,k p(li,k | Ri,kd,i :k ) (22)
tion model. When a new RMS trend passes the cri-
i=1
teria to be tracked, the RUL is estimated using the
designated model for this trend. The estimated RUL where R1:Nb = (R1 , . . . , RNb ). Operators can moni-
is here assumed probabilistic to comply with PFs and tor the median and confidence intervals (CIs) of p(¯lk |
KFs. Let p(li,k | Ri,kd,i :k ) denote the probability den- R1:Nb ) over time.
sity function (PDF) of RUL estimation li,k for RMS
trend i at time index k. Most of the methods reported
3.7 Overview
in literature only track a single trend over time, and
only estimate the RUL once per new vibration signal. A flowchart of the proposed method is shown in Fig. 3.
In contrast, the proposed method allows for tracking First, a vibration sensor is installed on the rotating ma-
up to Nb trends and estimating the RUL for each of chine, and baseline measurements are made. Based on

27
Modeling, Identification and Control

ISO 10816-3 ISO (1998), the velocity threshold is set. 4 Application with prediction
Using baseline data, the vibration failure threshold R̂a
and thresholds for all RMS bands R̂i are determined.
model
Afterwards, new measurements are split into mul- This section describes the implementation of a PF used
tiple RMS-based HI trends Ri ∀i ∈ [1, Nb ]. For each to update parameters of a Paris’ crack propagation
trend, the running Spearman coefficient ρi,k is calcu- model Lei et al. (2016a). Combined with the proposed
lated and checked for monotonicity. If the trend passes method, this implementation is applied to test the RUL
the criteria for being monotonic, a new model is initial- estimation capabilities on vibration data from two ex-
ized. If it is already initialized, the model parameters perimental tests.
are updated, and the RUL is estimated. Once all trends
are checked, the weighted RUL estimation p(¯lk | R1:Nb )
4.1 Degradation model
is determined, and related statistics can be monitored
over time. The bearing degradation level is assumed
monotonously increasing and never self-healing.
Also, as an increasing number of defects develop in the
3.8 Number of frequency bands bearing, the vibration level increases, resulting in an
exponential degradation rate. A study Rycerz et al.
Setting Nb requires some considerations. A high num- (2017) shows that bearing crack propagation may
ber of bands results in a threshold R̂a divided on a be modeled with the Paris’ law Paris and Erdogan
large number of RMS trends R̂i . Consequently, there (1963). This model describes the crack propagation
is a higher chance that a few of these trends increase rate in materials under cyclic load, and is given by
by more than others. Thus, setting a high number of
da √
Nb may result in a more conservative RUL estimation = c(∆k)m , ∆k = γ∆σ πa (23)
as some Ri may pass the threshold faster. Besides, dnc
a larger number of monotonic trends may need to be where a is the crack size, nc is the cycle number, c,
tracked using parameter update schemes. Thus, there m and γ are the material constants, and ∆σ is the
is a potential increased computational load by increas- cyclic load amplitude. Henceforth, a will be referred
ing the number of bands. On the other hand, setting to as the health√state of the bearing. For convenience,
a too low number for Nb may limit the ability to iden- let α = c(γ∆σ π)m and β = m/2, then the modified
tify frequency bands with monotonic energy increase, Paris’ law is Lei et al. (2016a)
restricting the ability to estimate the RUL. Nb should
da
be in the range [20, 40] to strike a balance between the = αaβ (24)
RUL estimation ability and computational burden. dnc
Eq. (24) is afterwards re-written in the form of a state-
space model as Lei et al. (2016a)
3.9 Robustness against noise
= ak−1 + αk−1 aβk−1 ∆nc

ak

In reality, measurements are always contaminated by αk = αk−1 (25)
noise, and increased noise will increase the signal en- 
Ri,k = ak + νh

ergy. With increased signal energy during baseline
measurements, the gap between the baseline mean and where ak is the health state at time index k, αk−1 ∼
ISO threshold is reduced. Increased noise in the time N (µα , σα2 ) is a normally distributed random variable,
domain may make the RUL estimation more conser- β is a constant, ∆nc is the number of cycles since
vative by decreasing the needed energy to classify the last update and νh ∼ N (0, σh2 ) is the measurement
system as damaged. noise. The health state ak is hard to determine in
If the noise is manifested in the energy domain, i.e., practice and is generally represented with an HI An
the energy fluctuates between measurements such that et al. (2013). Each RMS trend serves as the health
σi increases, the initial degradation alarm is activated state in this paper, which is contaminated by measure-
at a later stage. Such a change would not affect the fail- ment noise. Since the measurement noise is generally
ure threshold since R̂i is only based on the mean value small An et al. (2013), ak can be approximated by Ri,k .
of Ri during baseline measurements, which would not A PF is applied to update the model parameters
be affected by uniform noise. Besides, the experimental based on new samples, but the initial values are set
results in this paper are not from simulated data, but first. The modified Paris’ law in (25) has five unknown
rather from real sensor data of two test rigs, including parameters (a1 , µα , β, σα2 , σh2 ), which must be identi-
both mechanical and measurement noises. fied.

28
Klausen et.al., “RMS Based Health Indicators for RUL Estimation of Bearings”

4.2 Initial parameter setting update the parameters. General PF theory is given
in Lei et al. (2016a); An et al. (2013); Arulampalam
Three model parameters Θ1 = (a1 , µα , β) are identified
et al. (2002). A set of initial particles zkj ∀ j ∈ [1, Np ]
by minimizing an objective function using non-linear
are initialized at time index k with
least squares (NLS). The objective function is the mean    2 
square error between measurements and model simula- j ainit σh 0
zk ∼ N , (32)
tion. Let the measurement available for optimization µα 0 σα2
be H1:no = (Ri,k−kd,i , . . . , Ri,k ). To obtain the objec-
tive function, the ordinary differential equation in (24) where Np is the number of particles and ainit =
is first solved for a as P (no , Θ1 ). The particle weights are initialized as wkj =
1/Np .
a = (C1 (1 − β) + αnc (1 − β))1/(1−β) , β 6= 1 (26) The next model state is predicted using the particles
where C1 is based on the initial conditions, being de- with
termined by solving (26) with initial values a = a1 and  j  j
+ (µjα + να,k )(ajk−1 )β ∆T

a a
nc = 0. After the substitution of C1 , the equation zkj = jk = k−1 (33)
µα µjα
becomes
1/(1−β) j
where να,k ∼ N (0, σα2 ) is the process noise added due

P (nc , Θ1 ) = a = αnc (1 − β) + a1−β 1 (27)
to the uncertainty of bearing load.
Then, the goal is to minimize the error between the The health state ak is approximated with the RMS
measurements and the model, which is given by trend which is the observed variable. When a new mea-
surement Ri,k is available, the particle weights are up-
f (Hj , nc , Θ1 ) = Hj − P (nc , Θ1 ) (28) dated and normalized with
The unknown parameters Θ1 are finally identified using
w̃kj = wk−1
j
p(Ri,k | zkj ) (34)
the NLS minimization routine as described below , Np
wkj = w̃kj
X j
arg min F (H, Θ1 ) (29) w̃k (35)
Θ1
 j=1
0 ≤ a1 ≤ max(H1:no )

j
subject to 0 ≤ µα ≤ ∞ (30) where w̃k is the un-normalized particle weight and

0.5 ≤ β ≤ 1.25
 " #
j 1 (Ri,k − ajk )2
1X
no p(Ri,k | zk ) = p exp − (36)
where F (H, Θ1 ) = f (Hj , (j − 1)∆T, Θ1 )2 (31) 2πσh2 2σh2
2 j=1
A re-sampling step is performed if necessary to deal
Note that the constraints for β are set to avoid an with particle degeneracy. The effective number of par-
unstable exponential rate, and the number of cycles ticles is determined with
is replaced by time with (j − 1)∆T where ∆T is the , Np
time period between vibration measurements (i.e. be- X j
tween each k index). The trust region reflective algo- Neff = 1.0 wk (37)
i=1
rithm Branch et al. (1999) is used in this research to
minimize (29). When Neff < Np /2, particles are re-sampled accord-
The variances Θ2 = (σα2 , σh2 ) are set based on base- ing to a systematic re-sampling approach, and parti-
line data and initial values in Θ1 . The measurement cle weights are re-initialized as wkj = 1/Np Hol et al.
noise variance σh2 is simply identified as the variance (2006).
of Ri during the baseline measurements, i.e. σh2 = σi2 . A Gaussian importance re-sampling step is added
The process noise heavily influences a PF, and a greater to lessen particle impoverishment An et al. (2019).
variance results in a wider search space of the PF. The If there are less than N/3 unique particles after the
mean value µα is here used as the variance, such that systematic re-sampling step, all particles are re-drawn
σα2 = µ2α . The next section describes the PF used for from a continuous Gaussian distribution such that
updating the Paris’ law parameters online.
zkj ∼ N (µp , pΣp )) (38)
4.3 Particle filter
where µp and Σp are the mean and covariance matrix
After initializing the model parameters, a sequential of the particle states, respectively, and p is a scalar
importance sampling (SIS) PF is applied to further greater than 1.0.

29
Modeling, Identification and Control

The RUL estimated by particle j at time index k is


given by Table 1: Initial PF parameter values for test rig 1 ex-
periment.
lkj = inf{lkj : aj (lk + tk ) ≥ R̂i | aj1:k } (39) i µα σα2 σh2 β
j −6 −12 −4
where a (lk +tk ) is the predicted state value for particle 12 4.63 · 10 2.39 · 10 1.02 · 10 0.5
j at time tk + lk and aj1:k = (aj1 , . . . , ajk ). To solve (39), 9 3.17 · 10−6 1.11 · 10−12 5.57 · 10−5 0.96
the state of each particle j is simulated using the state 5 1.08 · 10−5 1.30 · 10−11 2.47 · 10−5 1.25
transition function given by (33) up to the time tk when 2 8.72 · 10−6 8.44 · 10−12 3.29 · 10−5 1.25
aj (lk + tk ) ≥ R̂i . With the estimated RUL and weight
for each particle, the RUL PDF is approximated by
Np are practically equal, the converged PF output is esti-
wkj δ(lk lkj ) mated and shown in column 3. Initial PF parameters
X
p(lk | Ri,1:k ) = − (40)
j=1
are given in Table 1.
At t ≈ 97 hours (k = 585), a new PF is initial-
where δ() is the delta-dirac function. The next sec- ized for R12 , indicating a change in bearing health.
tion includes experimental results using the proposed This RMS trend, shown in Fig. 4 (a), is from a high-
method combined with the PF for estimating the RUL. frequency band [4400, 4800] Hz. Initially, R12 increases
monotonously, but at t ≈ 117 hours, the trend in-
creases and decreases cyclically. This causes the run-
5 Experimental results ning Spearman calculation to decrease in value, and the
trend loses its significance in RUL estimation. µα has
5.1 Test rig 1 converged at t ≈ 110 hours as seen in Fig. 4 (b), and
To validate the proposed method, the IMS bearing the estimated trend will continue until the end of RUL.
dataset Qiu et al. (2006) is utilized. Details of the Fig. 4 (c) shows the median and 95% CI output of P F12
datasets are given in Section 2. The vibration spec- when µα converges. The future trend of the PF sug-
trum of each file is split into evenly sized bands of 400 gests an earlier failure than reality. However, the RUL
Hz, giving a total of Nb = 25 RMS trends. The baseline estimated by this trend is not the only contribution
measurements are calculated using data from the first to the weighted RUL in the proposed method. Other
30 hours. It is assumed that the bearing can last for monotonic RMS trends also contribute to the weighted
at least 5 hours after the initial fault. Hence the min- RUL decision. A vertical black line in Fig. 4 (a) shows
imum number of samples for Spearman calculation is when the running Spearman value is less than the re-
determined based on this. Given there are 6 files mea- quirement of 0.7. At this time, the trend is determined
sured per hour, ns = 30. The minimum value of Spear- to be unsuitable for RUL estimation after all, and its
man is set as ρ̂ = 0.9 to monitor mostly monotonic weight W12 is 0 as defined by (21).
trends. Non-monotonic trends should not contribute R9 passes the criteria for RUL estimation at t ≈ 110
significantly to the weighted RUL. Thus ρL = 0.7 is hours, and the initial output of the PF is shown in
the lower Spearman threshold. Np = 1000 particles Fig. 4 (d). The parameter µα in P F9 also converges
are used in each PF. To make sure the initial parame- as indicated in Fig. 4 (e). After this time, the con-
ters of the PFs are determined based on monotonically verged PF output trend continues until the Spearman
increasing values, the number of values for optimiza- coefficient value goes below 0.7 at t ≈ 135 hours. The
tion is set to no = ns . The value for p in the Gaussian converged output of P F9 is shown in Fig. 4 (f). The
re-sampling step is set to p = 2 in an attempt to de- converged PF gives a non-conservative RUL estima-
crease particle degeneracy. tion, as the value undershoots the FT.
Fig. 4 shows the results of using the proposed At t = 115 hours, P F5 is initialized, and the initial
method on this dataset. Subplots in the first column output is shown in Fig. 4 (g). R5 has fewer oscilla-
show the identified monotonic RMS trends and the out- tions than the two previously identified RMS trends,
put of the initialized PF. The median and 95% CI of the and therefore the Spearman coefficient never gets be-
initialized PF output are shown as red and red-stapled low 0.7. The parameter µα converges over 10 hours
lines, respectively, and black lines show Ri . The blue as shown in Fig. 4 (h), and results in a lower value
lines indicate the FT R̂i . Each row in Fig. 4 corre- of µα , which gives a less conservative RUL estimation
sponds to a single index of i, which is shown in the compared to the initial estimation. The predicted out-
upper left corner of the leftmost subplot column. The put after convergence shown in Fig. 4 (i) will reach the
second column shows the median and 95% CI of the threshold a bit early, but is relatively accurate.
µα parameter. When the median and 95% CI of µα When t ≈ 145 hours, P F2 is initialized, and R2 to-

30
Klausen et.al., “RMS Based Health Indicators for RUL Estimation of Bearings”

R̂i Ri Median 95% CI

2.0 i = 12 ρ12 ≤ 0.7 7e-06 2.0


RMS [m/s2 ]

RMS [m/s2 ]
5e-06

µα
1.0 1.0
2e-06
0.0 0.0
100 120 140 160 100 120 140 160 100 120 140 160
Time [hours] Time [hours] Time [hours]
(a) (b) (c)
i=9 ρ9 ≤ 0.7
RMS [m/s2 ]

RMS [m/s2 ]
5e-06
1.0 1.0
µα

2e-06

0.0 0.0
100 110 120 130 140 150 160 120 130 140 150 160 100 110 120 130 140 150 160
Time [hours] Time [hours] Time [hours]
(d) (e) (f )
i=5 2e-05
RMS [m/s2 ]

RMS [m/s2 ]
1.0 1.0
µα

0.5 1e-05 0.5

0.0 0.0
110 120 130 140 150 160 120 130 140 150 160 110 120 130 140 150 160
Time [hours] Time [hours] Time [hours]
(g) (h) (i)
1.0 i=2 1.0
RMS [m/s2 ]

RMS [m/s2 ]
2e-05
µα

0.5 0.5
1e-05

0.0 0.0
140 150 160 150 160 140 150 160
Time [hours] Time [hours] Time [hours]
(j) (k) (l)

Figure 4: Identified RMS trends, and output of the corresponding PFs. Rows 1-4 indicate i = [12, 9, 5, 2].
(column 1) Ri , FT R̂i , and median and 95% CI of initial PF output; (column 2) µα over time for the
initiated PF; (column 3) predicted PF trend when µα converges.

gether with the PF output are shown in Fig. 4 (j). R2 timate is defined as
is the most monotonic trend thus far, but the initial PNb PNp j j
PF is undershooting the FT and initially gives a poor i=1 Wi,k j=1 wi,k li,k
RUL estimation. However, the median and 95% CI of Weighted mean(k) = PN PNp j (41)
b
i=1 Wi,k j=1 wi,k
µα is continually increasing as seen in Fig. 4 (k) be-
cause of the Gaussian re-sampling step. Once the PF
which can be a number in between discrete values of
has converged at about t ≈ 151 hours, the PF out-
the weighted RUL PDF.
put in Fig. 4 (l) is following the future trend well, and
The weighed RUL mean, 95% CI and 30% error bars
reaches the FT at almost the correct time.
for this dataset are shown in Fig. 5 (a). Here, the true
RUL is shown as a solid black line, while the weighted
mean and 95% CIs are the red and red-stapled lines,
Combining the RUL estimations of each PF will give respectively. The 30% error is assumed as an accept-
a better estimate of the actual RUL. The median of this able limit for checking the performance of the RUL
combined discrete RUL PDF, however, can only take estimation. Ideally, the estimated RUL should always
discrete values that are present in the PDF. Conse- be within these error limits. The weighted mean oscil-
quentially, the median cannot take a value in between lates around the true RUL until the end of life, and is
two large discrete PDFs. To alleviate this constraint, only outside of the 30% error in a brief time between
and to reduce instability, the weighted mean RUL es- t ≈ [142, 152] hours.

31
Modeling, Identification and Control

100
95% CI
90
Weighted
80 mean
Remaining useful life [hours]

True RUL
70 30% error
60

50

40

30

20

10 Figure 6: Test rig 2 Klausen et al. (2017a).


0

95 100 105 110 115 120 125 130 135 140 145 150 155 160 165 170
Time [hours] 4
(a)
1.0

RMS [mm/s]
3
Rv
0.8
R̂v
Trend weights

2
0.6
W12 1
0.4
W9
W5 0
0.2
W2 0 20 40 60 80 100 120 140 160
0.0 Time [hours]
(a)
95 100 105 110 115 120 125 130 135 140 145 150 155 160 165 170 2.5
Time [hours]
(b)
2.0
RMS [m/s2 ]

Figure 5: Estimated RUL of the IMS dataset using the 1.5 Ra


proposed method. (a) weighted mean, 95% R̂a
1.0
CI, true RUL and 30% error margins; (b)
weights for each PF prediction. 0.5

0.0
0 20 40 60 80 100 120 140 160
When t = 125 hours, three PFs have a similar weight
Time [hours]
as shown in Fig. 5 (b), because of their high Spearman (b)
coefficient. The weighted mean calculated using (41)
is somewhere between these three RUL estimates, re- Figure 7: Comparison of velocity- and acceleration-
sulting in an estimate close to the true RUL. The lower based RMS of the in-house test rig dataset.
95% CI shows the lowest estimation of the trends and (a) velocity-based RMS; (b) acceleration-
is considered a conservative estimate of the RUL. Af- based RMS.
terward, the weights W12 and W9 decrease towards 0
due to a low Spearman coefficient, and model predic-
tions of R5 and R2 provide a reasonable estimate of race fault. The motor was operated at 700 rpm during
the RUL. measurements, and a planetary gearbox with a ratio
of 1:7 caused the test bearing on the output shaft to
rotate at 100 rpm. Vibration data was sampled ev-
5.2 Test rig 2 — in-house setup
ery 12 minutes (5 times per hour) at 51200 Hz for 6
Data from a second test rig is used to validate the seconds. The 154 last hours of operation are used to
proposed method further. A 6008-type roller element verify the performance of the proposed method. The
bearing is worn naturally in an accelerated life test by test rig is shown in Figure 6, and more details are given
applying radial and axial loads. The dynamic capacity in Klausen et al. (2017a).
of the bearing is 17.8 kN, and the static capacity is 11 The velocity- and acceleration-based RMS are shown
kN. With constant radial and axial loads of 9 kN and in Figs. 7 (a) and (b), respectively. The velocity-based
5 kN, respectively, the bearing lasted approximately FT is set to R̂v = 4.0 mm/s because a 1.1 kW motor is
34.6 million revolutions before failing due to an outer used, and the vibration signal was high-pass filtered at

32
Klausen et.al., “RMS Based Health Indicators for RUL Estimation of Bearings”

0.03 0.08
R23 R15 R11 R1
R̂23 R̂15 0.03 R̂11 0.3 R̂1
0.06
RMS [m/s2 ]

RMS [m/s2 ]

RMS [m/s2 ]

RMS [m/s2 ]
0.02
0.02 0.2
0.04 µ15 + 5σ15
µ23 + 5σ23 µ11 + 5σ11 µ1 + 5σ1
0.01
0.01 0.1
0.02

0.00 0.00 0.00 0.0


0 40 80 120 160 0 40 80 120 160 0 40 80 120 160 0 40 80 120 160
Time [hours] Time [hours] Time [hours] Time [hours]
(a) (b) (c) (d)

Figure 8: Collection of RMS trends with FTs for the in-house test rig dataset. (a) R23 [8800, 9200] Hz; (b) R15
[5600, 6000] Hz; (c) R11 [4000, 4400] Hz; (d) R1 [0, 400] Hz.

R̂i Ri Median 95% CI

i=1
0.3 0.3
RMS [m/s2 ]

RMS [m/s2 ]
4e-06
0.2 0.2
µα

2e-06
0.1 0.1

0.0 0.0
120 130 140 150 130 140 150 120 130 140 150
Time [hours] Time [hours] Time [hours]
(a) (b) (c)

Figure 9: R1 and initial median and 95% CI output of the corresponding PF. (a) R1 with marked part as
optimization input, blue line for FT, and red lines for median and 95% CI of PF; (b) α over time for
the initiated PF; (c) predicted PF trend when µα converges.

70
Table 2: Initial PF parameter values for test rig 2 ex- 95% CI
60 Weighted
periment. mean
Remaining useful life [hours]

True RUL
µα σα2 σh2 β 50 30% error
−6 −12 −5
P F1 3.21 · 10 1.14 · 10 2.25 · 10 0.5
40

30
10 Hz because of the motor speed of 700 rpm. Other
hyper-parameters (such as ρ̂, Np , ρL , etc.) are the 20

same as in experimental test 1, except where noted in


10
the following.
The velocity-FT R̂v is almost reached at the end 0
of useful life, as shown in Fig. 7 (a). On the other
hand, the transformed acceleration-based FT R̂a is far 110 115 120 125 130 135 140 145 150 155 160
Time [hours]
from reached in Fig. 7 (b). The test was stopped due
to a high torque requirement, indicating the bearing
was severely damaged. If the test was run for a few Figure 10: Estimated RUL of the in-house test rig
more measurement cycles, R̂a might have been reached. dataset using the proposed method. The
Nevertheless, it is assumed that the end of life is the weighted mean, 95% CI, true RUL and 30%
time of final measurement in this case. error margins are shown.
The proposed method is used to split the vibration
signal into Nb = 25 frequency bands, resulting in a fre-
quency bandwidth of 400 Hz. Some of the RMS trends and hence not suitable for RUL estimation. Addition-
are shown in Fig. 8. R23 in Fig. 8 (a) is very noisy, ally, R15 and R11 in Figs. 8 (b) and (c) do not increase

33
Modeling, Identification and Control

R̂i Ri Median 95% CI

6.0 4e-06 6.0


RMS [m/s2 ]

RMS [m/s2 ]
4.0 3e-06 4.0

µα
2e-06
2.0 2.0
1e-06
0.0 0.0
90 100 110 120 130 140 150 160 110 120 130 140 150 160 90 100 110 120 130 140 150 160
Time [hours] Time [hours] Time [hours]
(a) (b) (c)

Figure 11: R1 and initial median and 95% CI output of the corresponding PF. (a) R1 with marked part as
optimization input, blue line for FT, and red lines for median and 95% CI of PF; (b) α over time for
the initiated PF; (c) predicted PF trend when µα converges.

340
95% CI
Table 3: Comparison of error. The best values are 320
Weighted
300
shown in bold. mean
280
Remaining useful life [hours]

True RUL
e1 e2 260
30% error
240
Proposed method mean 7.88 10.33 220
200
Proposed method best 3.81 6.25 180
PF on full RMS 142.96 144.78 160
PF on MD-based HI 140
22.64 25.14 120
- Klausen et al. (2018) 100
Multi-scale approach 80
4.61 5.15
- Qian et al. (2017) 60
40
20
0
steadily. R1 in Fig. 8 (d), on the other hand, increases 95 100 105 110 115 120 125 130 135 140 145 150 155 160 165 170
almost linearly after the degradation alarm is triggered, Time [hours]
which makes the trend suitable for RUL estimation.
The first 30 hours in the dataset are used to cal- Figure 12: Estimated RUL of the IMS dataset using
culate the baseline values. 5 hours data is utilized to PF on RMS. The weighted mean, 95% CI,
determine the Spearman coefficient and optimization true RUL and 30% error margins are shown.
of initial values, e.g. ns = no = 25. The first and only
identified RMS trend is R1 at t = 116.5 hours using the
proposed RUL estimation algorithm. The initial PF the average error and root mean square error using the
parameters are given in Table 2. The initial PF out- following equations Qian et al. (2017):
put and R1 are shown in Fig. 9 (a), and µα is shown M
in Fig. 9 (b). The trend is re-drawn in Fig. 9 (c) to 1 X ˆ ˜
e1 = |li − li | (42)
show the PF prediction when µα has converged. Since M i=1
the predicted trend is similar to new samples, the esti- v
u
u 1 XM
mated RUL is accurate.
e2 = t (ˆli − ˜li )2 (43)
The weighted RUL in Fig. 10 is close to the actual M i=1
RUL during the rest of the lifetime. The main reason
is that R1 increases almost linearly with time, and the
where ˆli is the estimated RUL at index i (weighted
final measurements are very close to the FT. mean of p(¯li |R1:Nb ) in case of PF) and ˜li is the true
RUL at index i. The range i ∈ [1, M ] corresponds
5.3 Comparisons to all the values for which the method estimates the
RUL. The results from combining the proposed method
The proposed method’s performance is compared to with a Gaussian importance re-sampling PF are non-
other methods that use RMS and do not rely on his- deterministic, meaning that runs with different random
torical failure data. The performance is evaluated by seeds produce slightly different results of e1 and e2 .

34
Klausen et.al., “RMS Based Health Indicators for RUL Estimation of Bearings”

R̂i Ri Median 95% CI

3e-05
MD-based HI [-]

MD-based HI [-]
10.0 10.0

2e-05

µα
5.0 5.0
1e-05

0.0 0.0
90 100 110 120 130 140 150 160 110 120 130 140 150 160 90 100 110 120 130 140 150 160
Time [hours] Time [hours] Time [hours]
(a) (b) (c)

Figure 13: R1 and initial median and 95% CI output of the corresponding PF. (a) R1 with marked part as
optimization input, blue line for FT, and red lines for median and 95% CI of PF; (b) α over time for
the initiated PF; (c) predicted PF trend when µα converges.

Hence, the same RUL estimation as given in Section 5.1 100


95% CI
is carried out 1000 times with different random seeds to 90
Weighted
acquire a mean value of error. The mean and minimum 80
mean
Remaining useful life [hours]

True RUL
errors of the proposed method and comparisons are
70 30% error
available in Table 3.
60
The first comparison utilizes the PF as described in
Section 4.3 directly on the full RMS dataset. The ini- 50

tial degradation alarm is still determined based on the 40


mean and STD of RMS value during the first 180 mea- 30
surements. The results are shown in Fig. 11. The PF 20
is initialized at t ≈ 100 hours as seen in Fig. 11 (a),
10
and the PF follows initial trend of the signal. However,
once the PF converges at t ≈ 110 hours, the PF out- 0
put is not directed at the failure threshold, but rather 95 100 105 110 115 120 125 130 135 140 145 150 155 160 165 170
below it, as shown in Fig. 11 (c). The result is an Time [hours]
overestimated RUL, as is shown in Fig. 12. In this fig-
ure, the weighted mean of the PF is much greater than Figure 14: Estimated RUL of the IMS dataset using
the true RUL. In effect, the error for this estimation is PF on MD-based HI. The weighted mean,
considerable, as pointed out in Table 3. 95% CI, true RUL and 30% error margins
The second comparison utilizes the Mahalanobis- are shown.
distance-based HI (MD-based HI) to create a more lin-
ear HI compared to standard RMS. The method of gen-
erating the HI is well described in Wang et al. (2016), failure data to work. This historical data is used to
and the FT for this HI is calculated using the method determine two parameters, namely D1 and D2 in Qian
in Klausen et al. (2018). The PF with Paris’ law is uti- et al. (2017). Therefore, while the error can be com-
lized to predict the RUL using this MD-based HI. The pared between the proposed method and Qian et al.
results are shown in Fig. 13. At roughly 100 hours, (2017), it should be noted that the proposed method
the PF is initialized as indicated in Fig. 13 (a). Once does not require historical failure data.
converged, the PF output is directed straight at the
FT, as shown in Fig. 13 (c). This convergence results
in a very conservative RUL estimation, as seen in Fig. 6 Conclusions
14. The error is lower than standard RMS, as shown
in Table 3, but not as low as the proposed method. This paper proposes a novel method for subdividing the
The final comparison is the resulting RUL reported vibration signal into multiple frequency bands for root
in Qian et al. (2017) for the IMS dataset. The error mean square (RMS) calculations. The method utilizes
for this one is also given in Table 3. While e2 is lower a single discrete Fourier transform (DFT) per signal,
for the method given in Qian et al. (2017), it must and individual bins are used to acquire the signal en-
be noted that the method described requires historical ergy within a frequency band. The Spearman coeffi-

35
Modeling, Identification and Control

cient identifies monotonic RMS trends suitable for the life of slow speed bearings based on acoustic emis-
remaining useful life (RUL) estimation. It is observed sion. Mechanical Systems and Signal Processing,
that low-frequency RMS bands are most monotonic, 2017. 84:485–498. doi:10.1016/j.ymssp.2016.07.039.
while higher frequency RMS bands show earlier signs
of degradation. The failure threshold (FT) for vibra- Branch, M. A., Coleman, T. F., and Li, Y. A Sub-
tion RMS, developed in earlier research, has been ex- space, Interior, and Conjugate Gradient Method for
tended for multiple RMS trends. A framework for com- Large-Scale Bound-Constrained Minimization Prob-
bining the RUL estimation for each monotonic RMS lems. SIAM Journal on Scientific Computing, 1999.
trend has been presented. The Gaussian importance 21(1):1–23. doi:10.1137/s1064827595289108. Pub-
re-sampling particle filter (PF) is applied to determine lisher: Society for Industrial & Applied Mathematics
the Paris’ law parameters for estimating the RUL on (SIAM).
two experimental tests. The results show that the pro- Cheng, C., Ma, G., Zhang, Y., Sun, M., Teng,
posed method produces reasonable RUL estimations F., Ding, H., and Yuan, Y. A Deep Learning-
without the use of historical failure data. The perfor- Based Remaining Useful Life Prediction Ap-
mance is comparable to another method reported in proach for Bearings. IEEE/ASME Transac-
the literature that requires historical failure data. tions on Mechatronics, 2020. 25(3):1243–1254.
doi:10.1109/TMECH.2020.2971503. Conference
Name: IEEE/ASME Transactions on Mechatronics.
References
Gebraeel, N., Lawley, M., Liu, R., and Parmeshwaran,
Abboud, D., Elbadaoui, M., Smith, W., and Ran- V. Residual Life Predictions From Vibration-Based
dall, R. Advanced bearing diagnostics: A compara- Degradation Signals: A Neural Network Approach.
tive study of two powerful approaches. Mechanical IEEE Transactions on Industrial Electronics, 2004.
Systems and Signal Processing, 2019. 114:604–627. 51(3):694–700. doi:10.1109/TIE.2004.824875.
doi:10.1016/j.ymssp.2018.05.011.
Hol, J. D., Schon, T. B., and Gustafsson, F. On Re-
Ahmad, W., Khan, S. A., and Kim, J.-M. A Hybrid sampling Algorithms for Particle Filters. In 2006
Prognostics Technique for Rolling Element Bearings IEEE Nonlinear Statistical Signal Processing Work-
Using Adaptive Predictive Models. IEEE Transac- shop. IEEE, Cambridge, UK, pages 79–82, 2006.
tions on Industrial Electronics, 2018. 65(2):1577– doi:10.1109/NSSPW.2006.4378824.
1584. doi:10.1109/TIE.2017.2733487.
Hu, C., Youn, B. D., Wang, P., and Taek Yoon, J.
Akkad, K. and He, D. A Hybrid Deep Learning Based Ensemble of data-driven prognostic algorithms for
Approach for Remaining Useful Life Estimation. robust prediction of remaining useful life. Reliability
2019. pages 1–6. doi:10.1109/ICPHM.2019.8819435. Engineering & System Safety, 2012. 103:120–135.
doi:10.1016/j.ress.2012.03.008.
An, D., Choi, J.-H., and Kim, N. H. Prognos-
ISO. Mechanical vibration Evaluation of machine vi-
tics 101: A tutorial for particle filter-based prog-
bration by measurements on non-rotating parts Part
nostics algorithm using Matlab. Reliability En-
3 (Standard no. 10816-3). 1998.
gineering & System Safety, 2013. 115:161–169.
doi:10.1016/j.ress.2013.02.019. Khan, S. A., Prosvirin, A. E., and Kim, J.-M. Towards
bearing health prognosis using generative adversar-
An, H., Wang, G., Dong, Y., Yang, K., and Sang, L. ial networks: Modeling bearing degradation. In 2018
Tool life prediction based on Gauss importance re- International Conference on Advancements in Com-
sampling particle filter. The International Journal of putational Sciences (ICACS). IEEE, Lahore, pages
Advanced Manufacturing Technology, 2019. 103(9- 1–6, 2018. doi:10.1109/ICACS.2018.8333495.
12):4627–4634. doi:10.1007/s00170-019-03934-5.
Klausen, A., Folger, R. W., Robbersmyr, K. G.,
Arulampalam, M., Maskell, S., Gordon, N., and and Karimi, H. R. Accelerated Bearing Life-time
Clapp, T. A tutorial on particle filters for online Test Rig Development for Low Speed Data Acquisi-
nonlinear/non-Gaussian Bayesian tracking. IEEE tion. Modeling, Identification and Control: A Nor-
Transactions on Signal Processing, 2002. 50(2):174– wegian Research Bulletin, 2017a. 38(3):143–156.
188. doi:10.1109/78.978374. doi:10.4173/mic.2017.3.4.
Aye, S. and Heyns, P. An integrated Gaussian pro- Klausen, A., Robbersmyr, K. G., and Karimi,
cess regression for prediction of remaining useful H. R. Autonomous Bearing Fault Diagnosis

36
Klausen et.al., “RMS Based Health Indicators for RUL Estimation of Bearings”

Method based on Envelope Spectrum. IFAC- McFadden, P. D. and Smith, J. D. Model for the vi-
PapersOnLine, 2017b. 50(1):13378–13383. bration produced by a single point defect in a rolling
doi:10.1016/j.ifacol.2017.08.2262. element bearing. Journal of Sound and Vibration,
1984. 96(1):69–82. doi:10.1016/0022-460x(84)90595-
Klausen, A., Van Khang, H., and Robbersmyr,
9. Publisher: Elsevier BV.
K. G. Novel Threshold Calculations for Re-
maining Useful Lifetime Estimation of Rolling Pan, Z., Meng, Z., Chen, Z., Gao, W., and Shi,
Element Bearings. In 2018 XIII Interna- Y. A two-stage method based on extreme learn-
tional Conference on Electrical Machines (ICEM). ing machine for predicting the remaining useful
IEEE, Alexandroupoli, pages 1912–1918, 2018. life of rolling-element bearings. Mechanical Sys-
doi:10.1109/ICELMACH.2018.8507056. tems and Signal Processing, 2020. 144:106899.
doi:10.1016/j.ymssp.2020.106899.
Lei, Y., Li, N., Gontarz, S., Lin, J., Radkowski, S.,
and Dybala, J. A Model-Based Method for Re- Paris, P. and Erdogan, F. A Critical Analysis of Crack
maining Useful Life Prediction of Machinery. IEEE Propagation Laws. Journal of Basic Engineering,
Transactions on Reliability, 2016a. 65(3):1314–1326. 1963. 85(4):528–533. doi:10.1115/1.3656900. Pub-
doi:10.1109/TR.2016.2570568. lisher: ASME International.
Lei, Y., Li, N., Guo, L., Li, N., Yan, T., and Lin, J. Peeters, C., Guillaume, P., and Helsen, J. A
Machinery health prognostics: A systematic review comparison of cepstral editing methods as
from data acquisition to RUL prediction. Mechanical signal pre-processing techniques for vibration-
Systems and Signal Processing, 2018. 104:799–834. based bearing fault detection. Mechanical Sys-
doi:10.1016/j.ymssp.2017.11.016. tems and Signal Processing, 2017. 91:354–381.
Lei, Y., Li, N., and Lin, J. A New Method doi:10.1016/j.ymssp.2016.12.036.
Based on Stochastic Process Models for Machine
Qian, Y. and Yan, R. Remaining Useful Life Predic-
Remaining Useful Life Prediction. IEEE Transac-
tion of Rolling Bearings Using an Enhanced Par-
tions on Instrumentation and Measurement, 2016b.
ticle Filter. IEEE Transactions on Instrumenta-
65(12):2671–2684. doi:10.1109/TIM.2016.2601004.
tion and Measurement, 2015. 64(10):2696–2707.
Li, N., Lei, Y., Lin, J., and Ding, S. X. An Improved doi:10.1109/TIM.2015.2427891.
Exponential Model for Predicting Remaining Useful
Qian, Y., Yan, R., and Gao, R. X. A multi-time scale
Life of Rolling Element Bearings. IEEE Transactions
approach to remaining useful life prediction in rolling
on Industrial Electronics, 2015. 62(12):7762–7773.
bearing. Mechanical Systems and Signal Processing,
doi:10.1109/TIE.2015.2455055.
2017. 83:549–567. doi:10.1016/j.ymssp.2016.06.031.
Lu, C., Chen, J., Hong, R., Feng, Y., and Li,
Y. Degradation trend estimation of slewing bear- Qian, Y., Yan, R., and Hu, S. Bearing Degradation
ing based on LSSVM model. Mechanical Sys- Evaluation Using Recurrence Quantification Anal-
tems and Signal Processing, 2016a. 76-77:353–366. ysis and Kalman Filter. IEEE Transactions on In-
doi:10.1016/j.ymssp.2016.02.031. strumentation and Measurement, 2014. 63(11):2599–
2610. doi:10.1109/TIM.2014.2313034.
Lu, C., Wang, Y., Ragulskis, M., and Cheng, Y.
Fault Diagnosis for Rotating Machinery: A Method Qiu, H., Lee, J., Lin, J., and Yu, G. Wavelet filter-
based on Image Processing. PLOS ONE, 2016b. based weak signature detection method and its appli-
11(10):e0164111. doi:10.1371/journal.pone.0164111. cation on rolling element bearing prognostics. Jour-
nal of Sound and Vibration, 2006. 289(4-5):1066–
Ma, M. and Mao, Z. Deep Recurrent Convolutional 1090. doi:10.1016/j.jsv.2005.03.007.
Neural Network for Remaining Useful Life Pre-
diction. In 2019 IEEE International Conference Rycerz, P., Olver, A., and Kadiric, A. Propagation
on Prognostics and Health Management (ICPHM). of surface initiated rolling contact fatigue cracks in
pages 1–4, 2019. doi:10.1109/ICPHM.2019.8819440. bearing steel. International Journal of Fatigue, 2017.
97:29–38. doi:10.1016/j.ijfatigue.2016.12.004.
Manjurul Islam, M., Prosvirin, A. E., and Kim, J.-M.
Data-driven prognostic scheme for rolling-element Singleton, R. K., Strangas, E. G., and Aviyente, S.
bearings using a new health index and variants of Extended Kalman Filtering for Remaining-Useful-
least-square support vector machines. Mechanical Life Estimation of Bearings. IEEE Transactions
Systems and Signal Processing, 2021. 160:107853. on Industrial Electronics, 2015. 62(3):1781–1790.
doi:10.1016/j.ymssp.2021.107853. doi:10.1109/TIE.2014.2336616.

37
Modeling, Identification and Control

Singleton, R. K., Strangas, E. G., and Aviyente, S. Wang, Y., Peng, Y., Zi, Y., Jin, X., and Tsui, K.-L. A
The Use of Bearing Currents and Vibrations in Life- Two-Stage Data-Driven-Based Prognostic Approach
time Estimation of Bearings. IEEE Transactions for Bearing Degradation Problem. IEEE Transac-
on Industrial Informatics, 2017. 13(3):1301–1309. tions on Industrial Informatics, 2016. 12(3):924–932.
doi:10.1109/TII.2016.2643693. doi:10.1109/TII.2016.2535368.
Spearman, C. The Proof and Measurement of Associ-
ation between Two Things. The American Journal Zhang, P., Du, Y., Habetler, T. G., and Lu, B.
of Psychology, 1904. 15(1):72. doi:10.2307/1412159. A Survey of Condition Monitoring and Protec-
Publisher: JSTOR. tion Methods for Medium-Voltage Induction Motors.
IEEE Transactions on Industry Applications, 2011.
Wang, B., Lei, Y., Li, N., and Yan, T. Deep
47(1):34–46. doi:10.1109/TIA.2010.2090839.
separable convolutional network for remaining use-
ful life prediction of machinery. Mechanical Sys-
tems and Signal Processing, 2019. 134:106330.
Zhu, J., Chen, N., and Peng, W. Estimation of
doi:10.1016/j.ymssp.2019.106330.
Bearing Remaining Useful Life Based on Multiscale
Wang, D. and Tsui, K.-L. Statistical Modeling Convolutional Neural Network. IEEE Transactions
of Bearing Degradation Signals. IEEE Trans- on Industrial Electronics, 2019. 66(4):3208–3216.
actions on Reliability, 2017. 66(4):1331–1344. doi:10.1109/TIE.2018.2844856.
doi:10.1109/TR.2017.2739126.
Wang, W. A model to predict the residual Zhu, J., Chen, N., and Shen, C. A new data-
life of rolling element bearings given monitored driven transferable remaining useful life prediction
condition information to date. IMA Journal approach for bearing under different working con-
of Management Mathematics, 2002. 13(1):3–16. ditions. Mechanical Systems and Signal Processing,
doi:10.1093/imaman/13.1.3. 2020. 139:106602. doi:10.1016/j.ymssp.2019.106602.

38

You might also like