Sie sind auf Seite 1von 26

Journal of Sound and Vibration 357 (2015) 207–232

Contents lists available at ScienceDirect

Journal of Sound and Vibration


journal homepage: www.elsevier.com/locate/jsvi

A systematic study of ball passing frequencies based


on dynamic modeling of rolling ball bearings
with localized surface defects
Linkai Niu, Hongrui Cao n, Zhengjia He, Yamin Li
State Key Laboratory for Manufacturing Systems Engineering, Xi’an Jiaotong University, Xi’an 710049, PR China

a r t i c l e i n f o abstract

Article history: Ball passing frequencies (BPFs) are very important features for condition monitoring and
Received 4 March 2015 fault diagnosis of rolling ball bearings. The ball passing frequency on outer raceway (BPFO)
Accepted 6 August 2015 and the ball passing frequency on inner raceway (BPFI) are usually calculated by two well-
L.G. Tham
known kinematics equations. In this paper, a systematic study of BPFs of rolling ball
Available online 24 August 2015
bearings is carried out. A novel method for accurately calculating BPFs based on a
Keywords: complete dynamic model of rolling ball bearings with localized surface defects is
Ball passing frequency proposed. In the used dynamic model, three-dimensional motions, relative slippage, cage
Rolling ball bearing effects and localized surface defects are all considered. Moreover, localized surface defects
Fault diagnosis
are modeled accurately with consideration of the finite size of the ball, the additional
Dynamic model
clearance due to material absence, and changes of contact force directions. The reason-
ability of the proposed method for the prediction of dynamic behaviors of actual ball
bearings with localized surface defects and for the calculation of BPFs is discussed by
investigating the motion characteristics of a ball when it rolls through a defect. Parametric
investigation shows that the shaft speed, external loads, the friction coefficient, raceway
groove curvature factors, the initial contact angle, and defect sizes have great effects on
BPFs. For a loaded ball bearing, the combination of rolling and sliding in contact region
occurs, and the BPFs calculated by simple kinematical relationships are inaccurate,
especially for high speed, low external load, and large initial contact angle conditions
where severe skidding occurs. The hypothesis that the percentage variation of the spacing
between impulses in a defective ball bearing was about 1–2% reported in previous
investigations can be satisfied only for the conditions where the skidding effect in a
bearing is slight. Finally, the proposed method is verified with two experiments.
& 2015 Elsevier Ltd. All rights reserved.

1. Introduction

Rolling element bearings are widely used in aero-engines, high-speed spindles and other rotational machinery. Because
of material failure and adverse operating conditions, bearing faults often occur during operations, which may lead the whole
system to failure. Therefore, condition monitoring and fault diagnosis of rolling element bearings are crucial for the
prevention of system failure.

n
Corresponding author. Tel./fax: þ86 29 82663689.
E-mail address: chr@mail.xjtu.edu.cn (H. Cao).

http://dx.doi.org/10.1016/j.jsv.2015.08.002
0022-460X/& 2015 Elsevier Ltd. All rights reserved.
208 L. Niu et al. / Journal of Sound and Vibration 357 (2015) 207–232

In practice, vibration signals are widely used for the fault diagnosis of rolling element bearings. Fault features are
extracted from measured vibration signals with the aid of advanced signal processing techniques. In recent decades, many
advanced signal processing techniques have been proposed for extracting bearing fault features from measured vibration
signals, such as envelope analysis, wavelet [1,2], multiwavelet [3], dual-tree complex wavelet [4], empirical mode
decomposition [5], cyclostationarity [6], spectral kurtosis [7,8], ensemble empirical mode decomposition [9] and stochastic
resonance [10]. The most powerful bearing diagnostic techniques depend on detecting and enhancing the impulsiveness of
the signals [11]. When the bearing signals are enhanced and separated from other mechanical components and background
noise, ball passing frequencies (BPFs) are adopted as key indicators in determining whether a bearing has a fault or not, and
where the fault is. As is well known, for a ball bearing with stationary outer raceway, the ball passing frequency on outer
raceway (BPFO) and the ball passing frequency on inner raceway (BPFI) are expressed as Eqs. (1) and (2), respectively [11].
 
z fs D
BPFO ¼ 1 cos α (1)
2 dm
 
z fs D
BPFI ¼ 1þ cos α (2)
2 dm
where z is the number of rolling elements, f s is the shaft rotation frequency, D is the diameter of rolling element, dm is the
pitch diameter of the bearing, and α is the contact angle.
Moreover, the BPFs expressed by Eqs. (1) and (2) are widely used to model impulse trains induced by bearing defects. The
period between the impulses is the reciprocal of the BPF. The impulse trains are then used to mathematically model
vibration responses of defective bearings to predict the signals and to verify the diagnosis techniques. McFadden and Smith
[12,13] proposed a mathematical model of localized single and multiple point defects in a bearing under radial loads.
Tandon and Choudhury [14] proposed an analytical model for predicting the amplitudes of significant frequency
components due to localized defects for rolling element bearings. Choudhury and Tandon [15] modeled impact force trains
due to localized bearing defects based on BPFs. The impact force trains were used as excitations for a discrete spring-mass-
dashpot model of a rotor-bearing system to investigate the vibration responses. Stack et al. [16] developed a signal model for
inner raceway defects to design an inner raceway fault detection scheme. Cong et al. [17] proposed a signal model of rolling
element bearing in a rotor-bearing system for fault diagnosis. Cao et al. [18] simulated vibration responses of a machine tool
spindle system with localized bearing surface defects.
However, Eqs. (1) and (2) are obtained using kinematical relationships based on simple rolling motion and pure rolling
assumptions [19]. When a load occurs between a rolling ball and a raceway, a contact surface is formed. When the ball
rotates relative to the deformed surface, the simple rolling motion does not occur, rather, a combination of rolling and
sliding motions occurs [19]. As a result, the impulse trains are not strictly periodic, but stochastic due to the slip effect [11].
Ho and Randall [20] modeled a series of impulse responses for a bearing, and the model incorporated slight random
variations in the time between pulses. Antoni and Randall [21] proposed a bearing diagnosis technique in the presence of
strong interfering gear signals by a stochastic impulse train. Later, Antoni and Randall [22] modeled the impact process by
stochastic process to provide a better understanding of the recognized “envelope analysis” technique. Moreover, Sawalhi
and Randall [23,24] defined the slippage as a percentage variation of the mean frequency of impact (between 1% and 2%).
However, in these papers [20–24], the slippage effect was assumed as random numbers, and the root cause that introduces
the slippage were not investigated thoroughly.
In a word, the BPFs strongly depend on bearing parameters, operating conditions and lubricant characteristics which play
an important role in the contact condition and the relative slippage between a ball and a raceway. Brie [25] pointed out that
the assumption of the strict periodicity of the impulse train was questionable because of the complex dynamic phenomena.
However, these factors are not taken into account in Eqs. (1) and (2). Therefore, more complex models without kinematic
constraints but considering the relative slippage between bearing elements are necessary for calculating the BPFs exactly,
and enhancing the effectiveness of the diagnosis.
By now, many complex models of rolling ball bearings have been proposed, and they can be classified into quasi-static
model and dynamic model [26]. The quasi-static model was firstly proposed by Jones [27]. In this model, a “raceway control”
hypothesis (a type of kinematic constraints) which restricts the ball to spin either on the outer raceway or the inner raceway
was commonly used. Moreover, force and moment equilibrium equations were given for raceways and rolling elements.
These equations include centrifugal forces and gyroscopic moments. Based on the quasi-static model, the effects of
distributed faults of bearing raceways on system vibration responses were studied by Jang and Jeong [28] and Bai and Xu
[29]. The vibration responses of a spindle with defective ball bearings were studied by Cao et al. [18] based on Jones's model.
However, real behaviors of lubricant and relative slippage cannot be investigated by using this model due to its ‘raceway
control’ hypothesis.
In dynamic models, equilibrium equations used in quasi-static models are replaced by differential equations of motion
for each bearing component. All transient behaviors and lubrication effects can be simulated by dynamic models. Many
researchers [30–37] investigated the behaviors of rolling element bearings with localized surface defects using dynamic
models in recent years. These investigations were mainly focused on dynamic contact forces [30–33,36], varying stiffness of
a bearing system [31,32], time varying contact stiffness at contacting surfaces [37], and bearing accelerations [34] when a
roller rolls through a defect. Moreover, a dynamic modeling approach for wear evolution was proposed in Ref. [35].
L. Niu et al. / Journal of Sound and Vibration 357 (2015) 207–232 209

Fig. 1. Interactions between a ball and raceways. The symbols in this figure are for the interaction between a ball and the inner raceway. The interaction
between a ball and the outer raceway is similar to that between a ball and the inner raceway as shown in this figure. This figure is a simplified version of
the figure which was originally shown in Ref. [39] for investigating the interaction between a ball and a raceway.

However, in most of these investigations, the roller just has plane motions, and cage effects and relative slippage at
contacting surfaces, which largely influence the orbital speeds of rollers and BPFs, were not considered. Therefore, more
complex dynamic models are needed for predicting behaviors of defective bearings more reasonably and accurately.
The first complete three-dimensional dynamic model for rolling ball bearings called ADORE was proposed by Gupta
[26,38,39]. This model completely considered three-dimensional and time-varying transient motions of each bearing
component. The lubrication effect and relative slippage were also included. Based on Gupta's model, Wang et al. [40]
investigated vibration responses of a cylindrical roller bearing with localized surface defects, and Li et al. [41] proposed a
general dynamic modeling method for ball bearing-rotor systems. Niu et al. [42] recently developed a dynamic model for
rolling ball bearings with localized surface defects. In this model [42], the variations when a ball rolls through a defect, i.e.,
additional clearances due to material absence, changes of Hertzian contact coefficients and contact force directions were all
considered, and these characteristics were then integrated into Gupta's model to investigate the vibration responses of
rolling ball bearings with localized raceway defects. However, the finite size of a ball, which was recognized as an important
factor for the modeling of the interaction between a roller and a defect [33], was not considered in Refs. [40,42]. Moreover,
the cage effects were also not modeled in Refs. [40,42].
In this paper, based on the program ADORE [26], a novel dynamic model of defective ball bearings is proposed for
predicting the dynamic behaviors of defective ball bearings and calculating BPFs more reasonably and accurately. Compared
with available models for defective ball bearings [30–37,40,42], the effect of finite size of a ball, additional clearance at ball/
raceway contact, changes of contact force directions are all considered. By adopting Gupta's model, the relative slippage and
cage effects, which are important for the dynamics of a defective ball bearing but were not modeled in Refs. [30–37,40,42],
can also be considered in the proposed model. The proposed model is then used to investigate the dynamic behaviors of
defective ball bearings and BPFs systematically. The reasonability of the proposed model for the prediction of dynamic
behaviors of actual defective ball bearings is discussed by investigating the motion characteristics of a ball when it rolls
through a defect. Parametric studies are conducted to investigate the influences of operating conditions, lubrication
characteristics, bearing parameters, and defect sizes on BPFs. The simulation results are validated with two experiments.
The rest part of this paper is organized as follows. In Section 2, the modeling process of defective ball bearing is
introduced. In Section 3.1, simulations are carried out to discuss the reasonability of the proposed model for the prediction
of dynamic behaviors of actual defective ball bearings and for the calculation of BPFs. In Section 3.2, the influences of
operating conditions, bearing parameters and defect sizes on BPFs are investigated parametrically. These factors are shaft
speed, external loads, friction coefficient, raceway groove curvature factors, initial contact angle, and defect width.
Experiment verification is given in Section 4. Limitations of the proposed method are discussed in Section 5. Finally,
conclusions are presented in Section 6.

2. Dynamic modeling of rolling ball bearings with localized surface defects

2.1. Dynamic modeling of rolling ball bearings

In this section, the interactions between bearing components in a ball bearing are calculated based on Gupta's model
(ADORE [26]). Therefore, the basic concept used in ADORE is provided firstly in this section. In ADORE, models for simulating
the interactions between bearing components are developed in terms of the geometrical interactions, relative sliding
210 L. Niu et al. / Journal of Sound and Vibration 357 (2015) 207–232

Fig. 2. Traction model. This figure shows a simplified version of the typical traction curve of lubricant which was adopted by Gupta to investigate the
dynamic behaviors of rolling bearings in Ref. [26].

velocities, and the resulting normal and traction forces. Relative position vectors give the geometrical interactions between
two components. Then, the normal contact force can be obtained by using load-deflection relationship, such as Hertzian
contact theory. Similarly, together with the deformed geometry between the two components, the absolute velocities of
them will give their relative slip velocities at the contacting surface. By using the obtained relative slip velocity and contact
pressure, traction coefficient can be calculated based on the traction model of lubricant. Both the normal force and traction
force can now be added to determine the net force vector. The net moment vector is the cross product of the vector that
locates the point of interaction relative to the mass center and the net force vector. Finally, based on the obtained force and
moment vectors, the dynamic equations of these bearing components can be determined. Integrating the dynamic
equations numerically will give positions and velocities at the next time step.

2.1.1. Ball/raceway interactions


Take ball/inner raceway interactions for instance, as shown in Fig. 1. Similar method can be used to calculate ball/outer
raceway interactions. It should be noted that Fig. 1 is a simplified version of the figure which was originally shown in Ref.
[39] for investigating the interaction between a ball and a raceway. Three coordinate frames are established in Fig. 1.
Raceway fixed frame Or xr yr zr is established to describe the position of the raceway center in inertial frame Oi xi yi zi , and
azimuth frame Oa xa ya za is used to determine the radial and orbital positions of the ball in the inertial frame.
In order to calculate the interaction between a ball and a raceway, two position vectors, i.e., the position vector locates
the center of the raceway and the position vector locates the center of the ball are determined firstly. As shown in Fig. 1,
these two vectors are expressed as r r and r b , respectively.
Now, the geometrical interaction between a ball and a raceway can be determined by locating the ball center relative to
the raceway center. This vector is denoted by r b r in Fig. 1 and is given as

rb r ¼ rb  rr (3)

The relative position between the raceway groove curvature center and the ball center is given by

r bc ¼ rbr  r cr (4)

where vector r cr locates the raceway groove curvature center relative to the raceway center.
Contact angle is an important parameter for a ball bearing. Under unloaded conditions, the raceways have the same
contact angle. In this paper, the contact angle of the bearing under unloaded conditions is called ‘initial contact angle’ [43].
However, under dynamic conditions, contact angles of raceways are time-varying. Moreover, the contact angle largely
depends on operating conditions, such as external loads, shaft speeds, and lubrication characteristics. When the bearing is
operated at high speeds, the contact angle of inner raceway is larger than that of outer raceway due to large centrifugal force
[19,26]. The contact angle of a raceway under dynamic conditions can be determined based on the intersection angle
between vector r bc and plane ya za which is perpendicular to the bearing axis of rotation, as shown in Fig. 1.
 
r
α ¼ arctan bc1 (5)
r bc3

where subscripts 1 and 3 denote the first and the third components of vector r bc . In Eq. (5), vector r bc should be described in
azimuth frame Oa xa ya za , as shown in Fig. 1.
Next, the elastic deformation between the ball and the raceway can be obtained.

δ ¼ rbc3  ðf  0:5ÞD (6)

where f is the raceway groove curvature factor which is defined as the ratio of the radius of raceway groove to the ball
diameter. In Eq. (6), vector r bc should be described in contact frame Ok xk yk zk , as shown in Fig. 1. Contact frame can be
established based on azimuth frame and contact angle. More details can be found in Refs. [26,39].
L. Niu et al. / Journal of Sound and Vibration 357 (2015) 207–232 211

Fig. 3. Relative slip velocity urb at contact ellipse. (a) Two pure rolling points, and (b) one pure rolling point. In this figure, the relative slip velocity is shown
only along the major axis of the contact ellipse. Refs. [19,26] provide a more detailed description of the relative slip velocity which varies along both the
minor and the major axes of the contact ellipse.

Fig. 4. Ball/Cage pocket interaction. This figure is a simplified version of the figure which was originally provided in Ref. [39] for investigating the
interaction between a roller and a cage pocket.

The contact force between the ball and the raceway is given by Hertzian point contact theory [43]:
(
Kδ δ40
1:5
Q¼ (7)
0 δr0
where K is the Hertzian contact stiffness coefficient which largely depends on the curvature radii of two contacting bodies.
Moreover, it can be found from Eqs. (6) and (7) that raceway groove curvature factors have certain effect on the
contact force.
The friction forces between balls and raceways mainly rely on relative slip velocities. For any point in the contact ellipse
between a ball and a raceway, the relative slip velocity can be described as
urb ¼ upr  upb (8)

where upr and upb are velocities of the raceway and the ball in the contact ellipse, respectively.
Friction coefficient can be obtained by the lubricant model.
μ ¼ μðurb ; P r ; T Þ (9)

where P r is the contact pressure and T is the temperature. Fig. 2 shows a simplified curve of the typical traction curve of
lubricant which was adopted by Gupta to investigate the dynamic behaviors of rolling bearings in Ref. [26]. In this traction
model, the friction coefficient depends only on the relative slip velocity. This type of traction curve was adopted by various
researchers to study the dynamic performances of rolling bearings [40–42,44–48].
Generally, the contact ellipse is divided into several strips to determine the friction force. The obtained friction
coefficient, when multiplied by the incremental normal contact force dQ in a strip, gives the incremental friction force df T .
The incremental force vector dF can be given as
dF ¼ dQ þdf T (10)

Then, the incremental moments about mass centers for balls and raceways can be obtained.
When the incremental force dF and the incremental moment dM are integrated over the entire contact ellipse, the net
force F and the net moment M can be obtained. It can be found from Eqs. (8)–(10) that the friction force, the net force and
the net moment largely depend on relative slip velocities, friction coefficients and normal contact forces.
212 L. Niu et al. / Journal of Sound and Vibration 357 (2015) 207–232

Fig. 5. Model of bearing housing and the numbering of balls. Initially, the orbital position of the first ball (ball 1) is 0 degree with respect to axis zi .

Moreover, for rolling ball bearings, there may exist 0, 1, or 2 pure rolling points in the contact ellipse [19,26]. Fig. 3 shows
the relative slip velocity in the contact ellipse. This figure is a simplified version of figures in Refs. [19,26] in which the
relative slip velocity in the contact area was shown along both the major and the minor axes of the contact ellipse. As the
contact ellipse is always fairly narrow, the traction force is only determined along the major axis [39]. As shown in Fig. 3,
when the number of pure rolling points is 1 or 2, these pure rolling points will divide the contact ellipse into 2, or 3 sub-
regions, and the neighboring sub-regions have opposite relative slip velocities. As a result, the resulting friction forces in
these neighboring sub-regions are also opposite. According to the rolling direction of the ball, some friction forces will lead
the ball to accelerate (these forces can be recognized as traction forces), while other opposite friction forces will lead the ball
to decelerate.

2.1.2. Ball/cage interactions with raceway guidance


The cage investigated in this paper is assumed to be cylindrical. Similar to the approach used in ball/raceway interaction
calculations, the interaction between a cage pocket and a ball is determined by the relative position between the cage
pocket center and the ball center.
Fig. 4 is a simplified version of the figure which was originally provided in Ref. [39] for investigating the interaction
between a roller and a cage pocket. In Fig. 4, cage frame Oca xca yca zca is established to determine the azimuth and the radial
position of the cage in inertial frame Oi xi yi zi . As shown in Fig. 4, position vectors r b , r ca , and r cp are the position vectors
locating the center of the ball in inertial frame Oi xi yi zi , the center of the cage in inertial frame Oi xi yi zi , and the center of the
cage pocket in cage frame Oca xca yca zca , respectively. The relative position between the ball center and the cage pocket center
can be given as
r bcp ¼ r b  r ca  r cp (11)

It should be noted that the vectors on the right side of Eq. (11) should be transformed to proper frames before calculating.
More details can be found in Ref. [26].
The geometrical interaction between the ball and the cage pocket can be determined on the basis of vector r bcp and ball/
cage pocket clearance δbp . Next, the contact force between the ball and the cage pocket can be calculated using Hertzian
point contact theory. Moreover, the determinations of relative slip velocities, friction forces and moments between the ball
and the cage pocket are similar to those discussed in Section 2.1.1.
Additionally, the interaction between the cage and the guiding raceway is modeled as contacts between two cylinders.
More details can also be found in Ref. [26].

2.1.3. Dynamic equations


The translational motion of the mass center of any bearing component (balls, cage, and raceways) can be described by
Netwon's law of motion.
mr€ ¼ F (12)
where m is the mass of the component, and r€ is the acceleration vector.
L. Niu et al. / Journal of Sound and Vibration 357 (2015) 207–232 213

Fig. 6. Ball/defect interactions. (a) Interaction shown in plane xk zk , (b) interaction shown in plane yk zk , (c) contact force, and (d) defect size (wd and hd
correspond to the width and the depth, respectively). This schematic is shown for the interaction between a ball and a defect located in outer raceway. The
interaction between a ball and a defect located in inner raceway is similar to that shown in this figure.

The rotational motion of any bearing component (balls, cage, and raceways) can be described using Euler equations of motion:
8
< I 1 ω1  ðI 2 I 3 Þω2 ω3 ¼ M 1
> _
I2 ω
_ 2  ðI 3 I 1 Þω3 ω1 ¼ M 2 (13)
>
:I ω
3 _ 3  ðI 1 I 2 Þω1 ω2 ¼ M 3

where ðI 1 ; I 2 ; I 3 Þ are the principal moments of inertia, ðω1 ; ω2 ; ω3 Þ are the three components of angular velocity vector, and
ðM1 ; M2 ; M3 Þ are the three components of applied moment vector.
Moreover, another two translational dofs along yi and zi directions are added to model vibrations of the bearing housing,
as shown in Fig. 5.
(
mh y€ h þchy y_ h þ khy yh ¼ F hy
(14)
mh z€ h þ chz z_ h þ khz zh ¼ F hz

where mh is the mass of the housing, yh and zh are the translational displacements, khy and khz denote the stiffness, chy and
chz represent the damping coefficients, and F hy and F hz are the loads acting on the housing.

2.2. Localized surface defect model

When a ball rolls through a defect, two important variations arise, i.e.,

 Additional clearance between a ball and a raceway is introduced due to material absence in defect zone (Section 2.2.1).
 An additional tangential component of the contact force in defect zone is generated compared with original normal
conditions due to the change of the direction of contact force (Section 2.2.2).
214 L. Niu et al. / Journal of Sound and Vibration 357 (2015) 207–232

2.2.1. Additional clearance introduced by localized surface defect


As a ball rolls through a defect, the contact deformation between the ball and the raceway (δd ) is changed due to the
additional clearance introduced by the defect. The determination of δd is important for the calculation of the contact force
between the ball and the defect, and correspondingly the vibration responses of the bearing. In most of the available
literature [30–32,34,35,37,40,42], the ball is modeled as a point mass, and this results in overestimated contact force
between a roller and a defect [33]. A proper way to deal with this problem is to take account of the finite size of the ball as it
rolls through the defect [33]. In this paper, the model proposed in Ref. [42] is improved by considering the effect of the finite
size of a ball on the basis of Ref. [33].
As shown in Fig. 6, defect frame Od xd yd zd is established to determine the position of the defect center. The contacting
point between a ball and a raceway under normal conditions is Ok , and a frame called contact frame Ok xk yk zk is established
at this point. The zk axis is along the direction of vector r bc , and the yk axis is along the opposite direction of the orbital speed
of the ball. Moreover, as shown in Fig. 6(a), vector r bc intersects with the raceway axis of rotation at point m, and the
position vector locating the raceway groove center c relative to point m is r mc . Based on the geometry property of the
bearing, vector r mc can be described in contact frame Ok xk yk zk as
n oT
r
r mc ¼ 0 0 coscr3 α (15)

where r c r3 is the third component of vector r cr . In Eq. (15), vector r cr should be described in azimuth frame Oa xa ya za , as
shown in Fig. 6(a).
For any point (such as point p in Fig. 6(b)) on the surface of a ball, the position vector locating this point relative to ball
center b can be described in contact frame Ok xk yk zk as
n   d   oT
r p ¼ 0  2d sin φ 2 cos φ (16)

where φ is the angle between vector r p and axis zk .


Moreover, the position of point p relative to point m can be given as
r mp ¼ r mc 7 r bc þ r p (17)

where signs “ þ ” and “  ” refer to outer raceway and inner raceway respectively. In Fig. 6(b), vector r mp intersects with the
raceway at point q.
The geometrical interaction between the ball and the raceway at point p, δbd , can be determined based on the length of
vector r mp and the geometry property of the raceway:
   
δbd ¼ rmp   R θR (18)
   
where r mp  is the length of vector r mp , and R θR is the length of vector r mq as a function of the angle between vector r mp
and axis zd (i.e., θR , as shown in Fig. 6(b)).
Eq. (18) should be calculated for every point on the ball in plane yk zk , and the maximum positive δbd (δbd þ ) is adopted as
the geometrical interaction between the ball and the defect:
 
δd ¼ max δbd þ (19)

The contact force between the ball and the defect can be obtained by Hertzian contact theory:

Q bd ¼ K bd δd
1:5
(20)

where K bd is the Hertzian contact stiffness coefficient. Here, we assume that K bd is the same as that under normal
conditions. Indeed, non-Hertzian contact occurs between the ball and the defect due to the geometrical singularity of the
defect. The non-Hertzian contact characteristic will be investigated in the future.

2.2.2. Changes of directions of contact forces


When the geometrical interaction δd is determined using Eq. (19), the corresponding contact force between the ball and
the defect, Q bd , can be calculated based on Hertzian point contact theory. However, when the ball interacts with the defect,
the direction of contact force Q bd is not coincident with the line which connects the ball center b and the original raceway
groove curvature center c, and an additional force Q bd1 along axis xk can be decomposed from Q bd . In other words, contact
force Q bd can be decomposed into two components, i.e., tangential and normal components (Q bd1 and Q bd2 in Fig. 6(c)) with
respect to the original normal raceway. These effects should be considered in the dynamic model to accurately model the
behaviors of actual defective ball bearings. Therefore, the contact force is transformed to the original contact frame:
n     oT
Q kbd ¼ 0 Q bd sin φ Q bd cos φ (21)

where angle φ should satisfy Eq. (19).


It will be shown in Section 3 that the second component of Q kbd largely influences the orbital speed of a ball and the BPFs.
L. Niu et al. / Journal of Sound and Vibration 357 (2015) 207–232 215

Fig. 7. Orbital positions of a ball and a defect. In this figure, the origin of the inertial frame is assumed at the bearing center. The relationship between the
orbital positions can be clearly shown in plane yi zi .

Fig. 8. Flowchart of the numerical simulation.


216 L. Niu et al. / Journal of Sound and Vibration 357 (2015) 207–232

3. Simulations

The defect model discussed in Section 2.2 is integrated into the bearing model provided in Section 2.1 to simulate
dynamic performances of defective ball bearings. The general solutions of the dynamic equations (Eqs. (12)–(14)) are solved
numerically by using fourth-order Runge–Kutta–Fehlberg scheme with step-changing criterion. In program ADORE [26], all
of the variables were non-dimensionalized to enhance the numerical precision and the efficiency of the numerical
integration. In this paper, the variables are not non-dimensionalized. In the current investigation, the computer program has
been coded in FORTRAN 95 language. All of the variables in the program are double-precision floating-point. The authors
think that the numerical precision can be satisfied by the above programming scheme. Moreover, in the numerical
integration procedure, the relative orbital position between a ball and the defect is checked to determine whether the ball
rolls into the defect or not. As shown in Fig. 7, the orbital positions of the ball and the defect center are θb and θd ,
respectively. Moreover, in Fig. 7, θe is half the angle of the defect in the circumference of the raceway. It can be found that
when the difference between θb and θd (i.e., θbd in Fig. 7) is smaller than θe , the ball rolls into the defect. Detailed simulation
procedure is given in Fig. 8. Moreover, the descriptions of θbd and θe can be found in Ref. [42].
Then, envelope analysis is carried out on the simulated vibration responses to calculate BPFs. In this paper, the frequency
resolution is less than 0.1 Hz. As the time step size is step-changing in the numerical integration, the vibration response is
re-sampled before the envelope analysis. Moreover, the simulation results obtained by the dynamic model are compared
with those obtained by Eqs. (1) and (2) to show the reasonability of the proposed method. The initial contact angle is
adopted when Eqs. (1) and (2) are used to calculate BPFs as most usual applications. Here, Eqs. (1) and (2) are used based on
the assumption that the raceways have the same contact angle, i.e., the initial contact angle.
In Section 3.1, the reasonability of the dynamic model for the prediction of the behaviors of actual defective ball bearings
and for the calculation of BPFs is discussed. In Section 3.2, parametric studies are carried out to investigate the influences of
operating conditions, lubrication characteristics, bearing parameters, and defect sizes on BPFO and BPFI. These factors
include shaft speed Ω i , axial load F a , radial load F r , friction coefficient μ1 (refer to Fig. 2), ball/cage pocket clearance δbp ,
raceway groove curvature factors of inner raceway (f i ) and outer raceway (f o ), initial contact angle α0 , and defect width wd
(refer to Fig. 6(d)). The other geometrical parameters of the simulated bearing are listed in Table 1. In Table 1, the ball
number, and ball and pitch diameters are the same as those in Ref. [26] for investigating general dynamic motions of rolling
ball bearings.

3.1. Reasonability of the proposed model

As mentioned above, compared with available models for defective ball bearings, cage effects and relative slippage at
contacting surfaces are considered in the proposed model. Moreover, the effect of the defect on changes of directions of
contact force when a ball rolls through the defect is also taken into account. These considerations make the proposed model
more reasonable for the dynamic behavior predictions of defective ball bearings. In this section, the reasonability of the
proposed model for the prediction of dynamic behaviors of actual defective ball bearings and for the calculation of BPFs is
discussed.
In ball bearings, every ball exhibits similar motion characteristics when a pure axial load is applied. Therefore, in order to
investigate the general motion of defective ball bearings, a pure axial load of 2000 N is applied on the inner raceway.
Moreover, in order to demonstrate this issue better, a rectangular shaped defect is located in the outer raceway in this
simulation. The raceway groove curvature factors f i and f o are both set as 0.52. Moreover, the initial contact angle is 30
degree, the friction coefficient μ1 is 0.08, and the shaft speed is 15,000 r min  1. Besides the friction coefficient μ1 , the
lubrication parameter um (refer to Fig. 2) is set as 1 m s  1. The corresponding shaft rotation frequency f s is 250 Hz. The BPFO
calculated by Eq. (1) is 1475.04 Hz.
The dynamic behaviors of a ball when it rolls through the defect are shown in Fig. 9. The contact forces of the ball are
shown in Fig. 9(a). Fig. 9(b) is the enlarged view of Fig. 9(a). It can be seen that when the ball rolls into the defect
(corresponds to the ‘entry into’ point in Fig. 9), the contact forces decrease gradually due to material absence. However,

Table 1
Parameters of the simulated bearing system.

Number of balls z 14
Ball diameter D (m) 12.7  10  3
Pitch diameter dm (m) 70  10  3
Radial clearance (m) 0
Guiding raceway/cage clearance (m) 0.25  10  3
Young's modulus (Pa) 2.0  1011
Poisson ratio 0.25
Density (kg m  3) 7.75  103
Damping coefficient at ball/raceway (N s m  1) 20
Damping coefficients of bearing housing chy, and chz (N s m  1) 1800
Stiffness coefficients of bearing housing khy, and khz (N m  1) 15  106
L. Niu et al. / Journal of Sound and Vibration 357 (2015) 207–232 217

2 2
Defect zone inner raceway Defect zone
Contact force (kN)

Contact force (kN)


outer raceway
1.5 1.5 inner raceway
outer raceway
1 1
Entry into Exit
0.5 0.5
Impact

0 0
71.9 71.95 72 72.05 71.9 71.91 71.92 71.93 71.94
Number of shaft revolutions Number of shaft revolutions

50 50

Acceleration (m s )
Acceleration (m s )
-2

-2
0 0

-50 -50 Entry into Impact

-100 -100 Exit


Low frequency
-150 Defect zone High frequency -150 Defect zone

71.9 71.95 72 72.05 71.9 71.91 71.92 71.93 71.94


Number of shaft revolutions Number of shaft revolutions
Orbital speed of a ball (rad s )
Orbital speed of a ball (rad s )

-1
-1

690
Defect zone
Defect zone
688
688
687 Entry into Impact
Exit
686 686

685
684
71.9 71.95 72 72.05 71.9 71.91 71.92 71.93 71.94
Number of shaft revolutions Number of shaft revolutions

Fig. 9. Dynamic behaviors of a ball when it rolls through a defect located in outer raceway. (a) Contact force, (b) enlarged view of (a), (c) acceleration of the
bearing housing in zi direction, (d) enlarged view of (c), (e) orbital speed of a ball, (f) enlarged view of (e), and (g) schematic view of a ball when it rolls
through a defect located in outer raceway (δbp ¼ 0.1 mm, Ωi ¼ 15,000 r min  1, F a ¼ 2000 N, F r ¼ 0 N, α0 ¼ 30 degree, f i ¼ 0.52, f o ¼0.52, μ1 ¼0.08,
um ¼ 1 m s  1, wd ¼ 1.0 mm, hd ¼ 0.1 mm). In (g), the raceways are spread out for better showing.

although the ball center leaves the front side of the defect, the front side of the defect can also contact with the left side of
the ball, as shown in Fig. 9(g) (The raceways are spread out in Fig. 9(g) for better showing). As a result, the contact force at
ball/outer raceway cannot drop to zero suddenly. As the ball rolls into the defect, a low frequency can be found in
accelerations of the bearing housing (Fig. 9(c) and (d)). When the ball keeps on moving, the right side of the ball impacts the
back side of the defect (“impact” point in Fig. 9) although the center of the ball does not reach the back side of the defect, as
shown in Fig. 9(g). This can be attributed to the finite size of the ball [33]. It can be seen that a large contact force at outer
raceway is generated when the ball impacts the back side of the defect (Fig. 9(a) and (b)), and this force can be recognized as
the impact force between the ball and the defect. Moreover, after the ball impacts the back side of the defect, the contact
218 L. Niu et al. / Journal of Sound and Vibration 357 (2015) 207–232

800
800 Inner raceway
Inner raceway Defect zone
Defect zone Outer raceway

Contact force (N)


Outer raceway
Contact force (N)
600 600
Exit
Entry into
400 400

200 200
Impact
0 0
54.35 54.4 54.45 54.5 54.4 54.41 54.42 54.43 54.44
Number of shaft revolutions Number of shaft revolutions

40 40
Defect zone Defect zone
Acceleration (m ⋅s )

Acceleration (m ⋅s )
-2

-2
20 20
Entry into
0 0
Exit
-20 -20
Low frequency
High frequency Impact
-40 -40
54.35 54.4 54.45 54.5 54.4 54.41 54.42 54.43 54.44
Number of shaft revolutions Number of shaft revolutions
Orbital speed of a ball (rad⋅s )

Orbital speed of a ball (rad⋅s )


-1

-1

689.6 689.6
Defect zone
689.4 689.4
Exit
689.2 689.2
Impact
Entry into
689 689
Defect zone
688.8 688.8
54.4 54.45 54.5 54.55 54.4 54.41 54.42 54.43 54.44
Number of shaft revolutions Number of shaft revolutions

Fig. 10. Dynamic behaviors of a ball when it rolls through a defect located in the inner raceway. (a) Contact force, (b) enlarged view of (a), (c) acceleration of
the bearing housing in zi direction, (d) enlarged view of (c), (e) orbital speed of a ball, (f) enlarged view of (e), and (g) schematic view of a ball when it rolls
through a defect located in the inner raceway (δbp ¼ 0.1 mm, Ωi ¼15,000 r min  1, F a ¼ 2000 N, F r ¼0 N, α0 ¼ 30 degree, f i ¼ 0.52, f o ¼ 0.52, μ1 ¼ 0.08,
um ¼ 1 m s  1, wd ¼ 1.0 mm, hd ¼ 0.1 mm). In (g), the raceways are spread out for better showing.

forces at outer raceway and inner raceway increase and decrease successively. This shows that the ball jumps between the
two raceways after it impacts the defect. Additionally, when the ball jumps between the raceways, a high frequency motion
can be found in the acceleration of the bearing housing (Fig. 9(c)). These phenomena were also observed experimentally
[49,50].
An important dynamic performance which largely influences the BPFs is the orbital speed of a ball (θ_ b ). The orbital speed
of the ball when it rolls through the defect is shown in Fig. 9(e). Fig. 9(f) is the enlarged view of Fig. 9(e). It can be seen that
θ_ b increases slightly after the “entry into” point. This can be attributed to the tangential component of Q bd (Q bd1 as shown in
Fig. 9(g)) when the ball impacts the front side of the defect. The direction of this force is the same as the direction of θ_ b ,
which leads the ball to accelerate. However, because the force Q bd is relatively small at this stage, the acceleration is small
L. Niu et al. / Journal of Sound and Vibration 357 (2015) 207–232 219

100 100
With cage Without cage With cage Without cage
Acceleration (m⋅s )

Acceleration (m⋅s )
-2

-2
0 0

-100 -100

-200 -200
54 54.2 54.4 54.6 54.8 55 54.1 54.12 54.14 54.16 54.18 54.2
Number of shaft revolutions Number of shaft revolutions
Orbital speed of a ball (rad⋅s )
-1

692
With cage Without cage
690

688

686

684

682
50 60 70 80
Number of shaft revolutions

0.5 0.5
With cage 1533.9 Hz With cage
Without cage 1530.9 Hz Without cage
Amplitude (m⋅s )

0.4 0.4
Amplitude (m⋅s )

-2
-2

0.3 0.3

0.2 0.2

0.1 0.1

0 0
0 1000 2000 3000 4000 5000 1400 1500 1600 1700
Frequency (Hz) Frequency (Hz)
Fig. 11. Effects of the cage. (a) Acceleration of the bearing housing in zi direction, (b) enlarged view of (a), (c) orbital speed of a ball, (d) spacing between
two balls, (e) envelope spectrum, and (f) enlarged view of (e) (δbp ¼0.1 mm, Ωi ¼ 15,000 r min  1, F a ¼2000 N, F r ¼ 0 N, α0 ¼30 degree, f i ¼0.52, f o ¼0.52,
μ1 ¼ 0.08, um ¼1 m s  1, wd ¼1.0 mm, hd ¼0.1 mm).

correspondingly. When the ball keeps on moving, it contacts neither with inner raceway nor outer raceway, and the contact
forces of both raceways are all zero (‘hanging’ stage of the ball). As a result, θ_ b decreases gradually due to insufficient
traction force. When the ball impacts the back side of the defect, a large contact force is generated as discussed before. It can
be seen in Fig. 9(g) that the tangential component of the contact force (Q bd1 in Fig. 9(g)) at this stage is along the opposite
direction of θ_ b . This causes the ball to decelerate suddenly. It can be seen that the orbital speed of a ball when it rolls
through a defect is rather complex. However, in most of the available models [18,28–37], θ_ b is assumed as a constant. The
results show that the proposed model can reasonably predict the orbital speed of a ball when it rolls through the defect.
Fig. 10 shows the dynamic motions of a ball when it rolls through a defect in inner raceway. Similar analysis procedure
discussed above can be used to investigate the dynamic motions shown in Fig. 10. It should be noted that the contact force at
outer raceway cannot reduce to zero under this condition due to high centrifugal force (Fig. 10(a) and (b)). Moreover,
attentions should be paid to the direction of the contact force when the ball rolls through the inner raceway defect (Fig. 10
(g)) which is different from the condition of outer raceway defect.
Another important factor which was not considered in available models [18,28,29,31-33,35-37,40,42] is the cage.
However, the cage has a great influence on dynamic behaviors of a bearing [44]. In actual practice, the spacing between balls
is limited by the cage. The relative spacing between balls has some deviations due to the ball/cage pocket clearance, and this
is the reason for the random variation in BPFs which is experienced in practice. Similarly, take outer raceway defects for
instance. The radial accelerations of the bearing housing when some adjacent balls roll through the defect with and without
220 L. Niu et al. / Journal of Sound and Vibration 357 (2015) 207–232

750 10
Orbital speed (rad s )
-1
Cage Cage
Ball Ball

Skidding (%)
700 5

650 0
-1
Stable soulutions are around 687.09 rad s Stable soulutions are around 3.79%
600 -5
0 10 20 30 40 0 10 20 30 40
Number of shaft revolutions
Contact angle (degree) Number of shaft revolutions

40

30

20
Outer raceway,dynamic model
10 Inner raceway, dynamic model
Initial contact angle
0
0 10 20 30 40
Number of shaft revolutions
Fig. 12. Vibration responses of a bearing without raceway defects. (a) Ball and cage orbit speeds, (b) ball and cage skidding, and (c) contact angles
(δbp ¼ 0.1 mm, Ωi ¼15,000 r min  1, F a ¼ 2000 N, F r ¼0 N, α0 ¼ 30 degree, f i ¼ 0.52, f o ¼ 0.52, μ1 ¼ 0.08, um ¼1 m s  1).

cage are shown in Fig. 11. It can be seen that there are some deviations between these two conditions (Fig. 11(a) and (b)). As
shown in Fig. 11(c), when the cage is considered, θ_ b has very high frequency oscillations due to excessive impact between
the ball and the cage pocket, and the averaged amplitude of θ_ b is smaller than that when the cage is not considered.
Moreover, the spacing between two adjacent balls nearly remains constant when the cage is not considered (Fig. 11(d)).
However, in actual practice, the impacts between balls and cage pockets are excessive, and the spacing between balls
depends on the ball/cage pocket clearance as shown in Fig. 11(d). It can be seen that the maximum and the minimum of the
spacing between two balls correspond with two extreme relative positions between these two balls. The envelope spectrum
is shown in Fig. 11(e). Fig. 11 (f) is the enlarged view of Fig. 11(e). It can be found that the BPFO of the bearing without cage is
relatively larger than that when the cage is considered, and this deviation may be more severe under other complex
operating conditions. Moreover, when the cage is considered, some frequency smearing phenomena can be found in the
envelope spectrum (Fig. 11(e)) due to the impact between balls and cage pockets. These show the importance of the cage for
the prediction of dynamic behaviors of defective bearings.
It can be found that the BPFO calculated by Eq. (1) and that obtained by dynamic model are different from each other. The
reasons are discussed here. Eqs. (1) and (2) are constructed using kinematical relationships based on simple rolling motion
and pure rolling assumptions. However, when a ball rotates relative to the deformed surface, the simple kinematical
relationship used in Eqs. (1) and (2) does not occur, rather, a combination of rolling and sliding motions occurs [19]. In the
dynamic model, as mentioned in Section 1, no kinematic constraints and assumptions are used. The dynamic model
provides a real-time simulation of bearing performances. The vibration response of a healthy bearing is shown in Fig. 12 for
better understanding the skidding effect. As shown in Fig. 12, after several revolutions, traction forces (these forces are not
considered in Eqs. (1) and (2)) are adequate to produce a stable motion [26], and the ball orbit speed θ_ b and the cage orbit
speed θ_ c are stabilized to a certain value (some oscillations can be found in the stable solution of 687.09 rad s  1 due to the
impact between the ball and the cage pocket, as shown in Fig. 12(a)). Moreover, for a ball bearing, certain skidding exists
between a ball and a raceway due to relative slippage (this effect is not considered in Eqs. (1) and (2)). The skidding effect
can be evaluated by Eq. (22) [51] where θ_ p is the orbital speed calculated on pure rolling assumptions and θ_ dy is the orbital
speed calculated by the dynamic model considering relative slippage:

θ_ dy  θ_ p
Skidding ¼  100% (22)
θ_ p

The skidding effect calculated by the dynamic model is shown in Fig. 12(b). A positive percentage skidding can be found
under high speed conditions. This phenomenon was also reported in Ref. [52]. The relative slippage alters the overall
L. Niu et al. / Journal of Sound and Vibration 357 (2015) 207–232 221

Effects of shaft speed and axial load


-25
15 -1
BPFO 20 Fa=500 N, Ωi=15000 r min

Ball skidding (%)


BPFI
Relative error (%)

10 15
500 N 1000 N
5 10
2000 N Fa=2000 N, Ωi=3000 r min-1
5
0
2000 N 0
-5 1000 N
Fr=0 N 500 N -5
Fr=0 N
-10 -10
0 5000 10000 15000 40 45 50 55 60
-1
Shaft speed (r min ) Number of shaft revolutions
Fig. 13. Effects of shaft speed when the bearing is loaded by different axial loads. (a) Relative error, and (b) ball skidding (δbp ¼0.1 mm, F r ¼0 N, α0 ¼ 30
degree, f i ¼ 0.52, f o ¼ 0.52, μ1 ¼ 0.08, um ¼ 1 m s  1, wd ¼ 0.5 mm, hd ¼ 0.1 mm).

Effects of shaft speed and radial load


15 15
15000 r⋅min-1
10 -1
10000 r⋅min
Relative error (%)

-1 Ball skidding (%) 10 -1


5
6000 r⋅min -1 Fr=0 N, Ωi=15000 r⋅min
3000 r⋅min
5 -1
0 Fr=2000 N, Ωi=3000 r⋅min
-1
-5 3000 r⋅min
-1
6000 r⋅min 0
-10 -1
10000 r⋅min BPFO
15000 r⋅min
-1
BPFI Fa=1000 N
-15 -5
0 500 1000 1500 2000 40 45 50 55 60
Radial load (N) Number of shaft revolutions
Fig. 14. Effects of shaft speed when the bearing is loaded by different radial loads. (a) Relative error, and (b) ball skidding (δbp ¼ 0.1 mm, F a ¼ 1000 N, α0 ¼30
degree, f i ¼ 0.52, f o ¼ 0.52, μ1 ¼ 0.08, um ¼ 1 m s  1, wd ¼ 0.5 mm, hd ¼ 0.1 mm).

dynamic characteristics of the bearing, such as the contact angles shown in Fig. 12(c). It should be noted that different
external loads and lubricants will result in different BPFs. In a word, dynamic analysis can provide more reasonable BPFs.

3.2. Parametric studies

In this section, the influences of operating conditions, lubricant characteristics, bearing parameters and defect sizes on
BPFO and BPFI are discussed. The investigated factors include shaft speed Ω i , friction coefficient μ1 , axial load F a , radial
load F r , ball/cage pocket clearance δbp , raceway groove curvature factors f i and f o , initial contact angle α0 , and defect width
wd . In order to evaluate the influences of these factors, the relative error Δf (Eq. (23)) between
  the results calculated by Eqs.
(1) and (2) (f p ) and the results calculated by the proposed model (f d ) is provided. Large Δf  (j U j denotes the absolute value)
means that the influence of the parameter is significant:
f d f p
Δf ¼  100% (23)
fp

3.2.1. Effects of shaft speed, friction coefficient, and external load


BPFs are mainly dominated by orbital speeds of balls. Moreover, when the shaft speed is fixed, the orbital speeds are
largely determined by the traction forces at contacting surfaces. When traction forces are insufficient, the skidding effect
becomes severe, and Δf  becomes large. Additionally, the traction forces are dominated by friction coefficients, contact
forces and relative slip velocities at contacting surfaces. Therefore, shaft speeds, axial loads, radial loads, and friction
coefficients are combined together to investigate the effects of operating conditions on BPFs in this section.
 Fig.
 13 shows the relative errors under different shaft speeds and axial loads. It can be found that when the axial load is fixed,
Δf  increases with the increase of shaft speed. This phenomenon can be explained that when the axial load is fixed, the skidding
between a ball and a raceway
 becomes severe when the shaft speed increases [51,53]. Moreover, it can be seen that when the
shaft speed is fixed, Δf  increases with the decrease of axial load, which can be attributed to the severe skidding when the axial
load is low. To better demonstrate these effects, two extreme conditions in Fig. 13(a) (Ω i ¼ 15,000 r min  1 and Fa ¼500 N, and
222 L. Niu et al. / Journal of Sound and Vibration 357 (2015) 207–232

Effects of friction coefficient and shaft speed 20


20 -1
15000 r⋅min 15

Ball skidding (%)


-1 μ∞=0.005, Ω i=15000 r⋅min-1
Relative error (%)

10 3000 r⋅min 6000 r⋅min-1 -1 10


10000 r⋅min
μ∞=0.08, Ω i=3000 r⋅min-1
5
0
0
-10
-1 BPFO -5
10000 r⋅min 15000 r⋅min
-1
BPFI Fa=1000 N, Fr=0 N
-20 -10
0 0.02 0.04 0.06 0.08 40 45 50 55 60
Friction coefficient μ Number of shaft revolutions

Effects of friction coefficient and axial load 15


15
500 N
μ∞=0.005, Fa=500 N

Ball skidding (%)


10
Relative error (%)

10
1000 N
5
2000 N 5 μ∞=0.08, Fa=2000 N
0
2000 N
0
-5 1000 N
500 N
Ω i=10000 r⋅min-1, Fr=0 N
-10 -5
0 0.02 0.04 0.06 0.08 40 45 50 55 60
Friction coefficient μ∞ Number of shaft revolutions

Fig. 15. Effects of friction coefficient μ1. (a) Relative errors when the bearing is operated at different shaft speeds (F a ¼ 1000 N and F r ¼ 0 N), (b) ball
skidding of two conditions shown in (a), (c) relative errors when the bearing is loaded by different axial loads (Ωi ¼10,000 r min  1 and F r ¼0 N), and (d)
ball skidding of two conditions shown in (c) (δbp ¼ 0.1 mm, α0 ¼ 30 degree, f i ¼ 0.52, f o ¼ 0.52, um ¼1 m  s  1, wd ¼ 0.5 mm, hd ¼ 0.1 mm).

Ω i ¼3000 r min  1 and F a ¼2000 N) are adopted to investigate the skidding effect, as shown in Fig. 13(b). The ball skidding is
given for healthy bearings in the rest part of this paper for better understanding the skidding effect. It can be found that the ball
skidding is relatively slight when Ω i ¼3000 r min  1 and F a ¼2000 compared with Ω i ¼15,000 1
  r min and F a ¼ 500 N. Fig. 13
 
shows that low axial load and high speed conditions can result in severe skidding and large Δf . Additionally, it can be seen that
the relative error of BPFI is negative compared with BPFO at relative high speeds, which means that f p (Eq. (23)) is larger than f d
when calculating BPFI, while f p is smaller than f d when calculating BPFO. The main reason may be due to the contact angle. The
contact angle of inner raceway increases and the contact angle of outer raceway decreases due to high ratio of centrifugal force to
contact force under high speed conditions [19,53,54].
The influences of radial
  load at different shaft speeds can be found in Fig. 14. It can be found in Fig. 14(a) that when the
radial load decreases, Δf  increases correspondingly. This can also be attributed to severe skidding when the radial load is
small, as shown in Fig. 14(b). The relationship between ball skidding
  and radial load was also reported in [55]. Moreover, it
can also be found in Fig. 14 that when the radial load is fixed, Δf  increases with the increase of shaft speed due to severe
skidding under high speed conditions.
Another important factor which influences the traction force between a ball and a raceway is the friction coefficient. For
different types of lubricants, the friction coefficient varies in a wide range. For some solid lubricants, the friction coefficient
is in the range of about 0.02–0.6 [56]. For some oil lubricants, this value ranges approximately from 0.005 to 0.01 [47], and
from 0.05 to 0.1 for boundary lubrication conditions. It should be noted that the friction coefficient largely depends on
material properties and operating conditions, and detailed investigation on the lubricant behaviors is beyond the scope of
the current analysis. For comparison, four different friction coefficients (0.005, 0.01, 0.05, and 0.08) are chosen in this
simulation. The effects of friction coefficient μ1 can be found in Fig. 15. The absolute value of relative error becomes large
when the friction coefficient is small and the shaft speed is high (Fig. 15(a)). Small μ1 results in small friction force, and this
will lead to insufficient traction forces and severe skidding [19]. The ball skidding under two extreme conditions in Fig. 15(a)
is shown in Fig. 15(b). Severe skidding can be found when μ1 ¼0.005 and Ω i ¼15,000 r min  1. Although the relative error
becomes severe with the decrease of μ1 , the shaft speed also plays the dominant role in the relative error in Fig. 15(a) due to
high ratio of centrifugal force to contact force. Fig. 15(c) and (d) show that the relative error becomes severe when μ1 and
axial load are both small. This can also be attributed to insufficient traction forces and severe skidding under small μ1 and
axial load conditions, as shown in Fig. 15(d). However, the axial load has a much greater effect than μ1 on the skidding and
relative error.
L. Niu et al. / Journal of Sound and Vibration 357 (2015) 207–232 223

Effects of ball/cage pocket clearance and radial load


15
0N

Relative error (%)


10
BPFO
5 BPFI 1000 N 2000 N
0

-5
0N
-10
0.1 0.2 0.3 0.4 0.5
Ball/cage pocket clearance

27 26.5
Fr=1000 N Fr=2000 N δbp=0.1 mm
δbp=0.1 mm δbp=0.25 mm
26.5
θb2-θb1 (degree)

θb2-θb1 (degree)
26
26

25.5
25.5
25
δbp=0.25 mm δ =0.45 mm
δbp=0.45 mm bp
24.5 25
90 92 94 96 98 100 130 135 140 145 150
Number of shaft revolutions Number of shaft revolutions
Fig. 16. Effects of ball/cage pocket clearance when the bearing is loaded by different radial loads. (a) Relative error, (b) spacing between two adjacent balls
when F r ¼1000 N, (c) spacing between two adjacent balls when F r ¼2000 N (Ωi ¼ 10,000 r min  1, F a ¼ 500 N, α0 ¼ 30 degree, f i ¼0.52, f o ¼0.52, μ1 ¼ 0.08,
um ¼1 m s  1, wd ¼0.5 mm, hd ¼0.1 mm).

From the above investigations (Figs. 13–15), it can be found that the shaft speeds and the external loads are the dominant
operating parameters which largely influence BPFs. These parameters largely affect traction forces and skidding of a ball.
When the skidding effect of the investigated
  bearing is severe (shaft speed is larger than 10,000 r min  1, and axial and
radial loads are less than 500 N), Δf  are in the range of about 4.65–13.76% for BPFO and are in the range of about 2.24–
1
9.53% for BPFI, respectively. When the shaft speed is low (less than 3000r min  ), and the external load is high (larger than
2000 N), the skidding in the investigated bearing becomes slight, and Δ f  is about less than 1% for both BPFO and BPFI.

Additionally, Δf  of BPFO is larger than that of BPFI. It should be noted that different bearings will exhibit different skidding
characteristics, and the effects of these factors should be determined for different bearings specially.

3.2.2. Effects of ball/cage pocket clearance


The ball/cage pocket clearance δbp plays an important role in cage instability. When the ratio of δbp to cage/guiding
raceway clearance is larger than one, the cage may exhibit certain instability which leads the cage to failure [44]. In this
paper, the cage/guiding raceway clearance is set as 0.25 mm, and three different δbp (0.1 mm, 0.25 mm, and 0.45 mm) are
chosen to investigate the effects of δbp . Generally, the spacing between two adjacent balls are limited by the cage, and the
interaction at ball/cage pocket becomes more severe when a bearing is loaded by both axial and radial loads where orbital
speeds of balls are periodic with the rotation of the bearing. Therefore, in this section, the effects of δbp on BPFs are
investigated when the bearing is loaded by different radial loads.
The effect of δbp on BPFs is shown in Fig. 16. It can be seen that when the radial
 load is zero, the relative error is barely
affected by δbp . When the radial load increases, such as 1000 N in Fig. 16(a), Δf  increases slightly with the increase of δbp .
 
When δbp are 0.1 mm, 0.25 mm, and 0.45 mm, the corresponding Δf  are 0.63%, 1.26%, and 1.83% for BPFO, and are 0.48%,
0.92%, and 1.36% for BPFI. However, when the radial load increases up to 2000 N, the relative error does not affected by δbp
when δbp increases to a certain value (0.25 mm). When δbp are 0.1 mm, 0.25 mm, and 0.45 mm, the corresponding Δf  are
0.552%, 0.33%, and 0.33% for BPFO, and are 0.4%, 0.26%, and 0.26% for BPFI. This shows that when the traction forces are
sufficient to weaken the skidding effect, the influences of δbp on BPFs become insignificant.
An important issue which was investigated by previous researchers [11–13,20–25] is the impulse train of force generated
in defective bearings. Most of the researchers [11,20–25] realized that the spacing between the impulses in a defective
bearing is not strictly periodic because of the slippage at contacting surfaces. Sawalhi and Randall [23,24] reported that the
variation in the spacing is about 1–2%. In this section, variations in the spacing between impulses are studied by
investigating the spacing between two adjacent balls in a defective ball bearing. The spacing between two adjacent balls can
224 L. Niu et al. / Journal of Sound and Vibration 357 (2015) 207–232

Effects of raceway groove curvature factors


and axial load
15
BPFO
BPFI 500 N

Relative error (%)


10

5 1000 N

2000 N
0 2000 N

-5 1000 N

500 N
-10
0.5 0.52 0.54 0.56
Raceway groove curvature factors (f and f )
i o

Fig. 17. Effects of raceway groove curvature factors when the bearing is operated by different axial loads (Ωi ¼10,000 r min  1, F r ¼ 0 N, α0 ¼ 30 degree,
μ1 ¼ 0.08, um ¼ 1 m s  1, wd ¼ 0.5 mm, hd ¼ 0.1 mm, δbp ¼0.1 mm).

Effects of initial conact angle and axial load


15 15 -1
BPFO Fr=0 N, Ωi=10000 r⋅min
BPFI Ball skidding (%)
Relative error (%)

10 1000 N 500 N
2000 N 10
5 α0=30 degree, Fa=500 N
0
5
-5 2000 N α0=10 degree, Fa=2000 N
-1
Fr=0 N, Ωi=10000r⋅min 1000 N 500 N
-10 0
10 15 20 25 30 40 45 50 55 60
Initial contact angle (degree) Number of shaft revolutinns

Effects of initial contact angle and radial load 15


15
Ball skidding (%)

BPFO 10
Relative error (%)

10 0N
BPFI α0=30 degree, Fr=0 N
5
1000 N 2000 N 5
α0=10 degree, Fr=2000 N
0
0
-5 -1 -1
Fa=500 N, Ωi=10000 r⋅min 0N Fa=500 N, Ωi=10000 r⋅min
-10 -5
10 15 20 25 30 40 45 50 55 60
Initial contact angle (degree) Number of shaft revolutions
Fig. 18. Effects of initial contact angle. (a) Relative errors when the bearing is loaded by different axial loads (F r ¼ 0 N), (b) ball skidding of two conditions in
(a), (c) relative errors when the bearing is loaded by different radial loads (F a ¼ 500 N), and (d) ball skidding of two conditions in (c) (Ωi ¼ 10,000 r min  1,
f i ¼0.52, f o ¼ 0.52, μ1 ¼0.08, um ¼ 1 m s  1, wd ¼ 0.5 mm, hd ¼0.1 mm, δbp ¼ 0.1 mm).

be investigated by the difference between orbital positions of these two balls (θb2  θb1 in Fig. 16, where θb1 and θb2 are the
orbital positions of the first and the second balls, respectively. The numbering of balls in the bearing is shown in Fig. 5). As
shown in Fig. 16(b), when the radial load is 1000 N, the percentage deviations of θb2  θb1 from their mean values for the
three different δbp are about 71.4%, 72.5%, and 73.3%, respectively. However, when the radial load increased to 2000 N as
shown in Fig. 16(c), the percentage deviations of θb2  θb1 from their mean values for the three δbp are about 71.3%, 71.9%,
and 70.4%, respectively. This shows that the spacing between two adjacent balls largely depends on the operating
conditions and δbp . Additionally, from the discussions in Section 3.2.1, it can be found that the relative error of BPFs is
smaller than 2% only when the skidding effect in the bearing is slight. As a result, the hypothesis of 1–2% can be satisfied
only when the skidding effect is insignificant.
L. Niu et al. / Journal of Sound and Vibration 357 (2015) 207–232 225

1.5 15
BPFO BPFO
BPFI BPFI

Relative error (%)


Relative error (%)

1 -1 10
6000 r⋅min -1
15000 r⋅min
0.5 5
-1
3000 r⋅min
-1
0 0 10000 r⋅min
-1
3000 r⋅min
-0.5 -5 -1
-1 15000 r⋅min
6000 r⋅min
-1 -10
0 1 2 3 4 5 0 1 2 3 4 5
Defect width (mm) Defect width (mm)
Orbital speed of a ball (rad⋅s )
-1

480
Defect in outer raceway
0.5 mm 3 Ω =10000 r⋅min-1 3.5 mm

Contact force (kN)


i
460 5.0 mm
2 Impact force

440 1.5 mm 1.5 mm


1
Ω i=10000 r⋅min-1 3.5 mm 0.5 mm
Defect in outer raceway 5.0 mm
420 0
84 84.2 84.4 84.6 84.8 85 84.9 84.95 85 85.05 85.1
Number of shaft revolutions Number of shaft revolutions
Orbital speed of a ball (rad⋅s )
-1

470 2.5
Defect in inner raceway Defect in inner raceway
-1
Ω i=10000 r⋅min-1 Ω i=10000 r⋅min
Contact force (kN)

2 5.0 mm
465 0.5 mm
Impact force
1.5
460 3.5 mm
1.5 mm
1
0.5 mm 1.5 mm
455 5.0 mm 3.5 mm
0.5

450 0
92 92.2 92.4 92.6 92.8 93 92.15 92.2 92.25
Number of shaft revolutions Number of shaft revolutions
Fig. 19. Effects of defect width. (a) Relative errors when the shaft speeds are 3000 r min  1 and 6000 r min  1, (b) relative errors when the shaft speeds are
10,000 r min  1 and 15,000 r min  1, (c) ball orbit speeds when the ball rolls through a defect located in outer raceway, (d) contact forces at ball/outer
raceway when the ball rolls through a defect located in outer raceway, (e) ball orbital speed when the ball rolls through a defect located in inner raceway,
and (f) contact forces at ball/inner raceway when the ball rolls through a defect located in inner raceway (F a ¼ 1000 N, F r ¼ 0 N, α0 ¼ 30 degree, f i ¼0.52,
f o ¼0.52, μ1 ¼ 0.08, um ¼1 m s  1, hd ¼ 0.1 mm, δbp ¼0.1 mm). In (d) and (f), the maximum contact force is recognized as the impact force between the ball
and the defect.

3.2.3. Effects of raceway groove curvature factors


The raceway groove curvature factor is between about 0.515 and 0.53 for the most single row deep groove ball bearings
and is between about 0.52 and 0.53 for the most single row angular contact ball bearings [43]. However, this value will
between 0.54 and 0.58 in other special applications [57]. As a result, five different raceway groove curvature factors (0.51,
0.52, 0.53, 0.54, and 0.55) are chosen in this simulation. Moreover, raceway groove curvature factors mainly affect the
contact forces. Although the external load is fixed, the contact forces at ball/raceways vary with the change of raceway
groove curvature factors. As a result, the effect of raceway groove curvature factor is investigated when the bearing is loaded
by different axial loads.
The influences of raceway groove  curvature
 factors are shown in Fig. 17. It can be seen from Fig. 17 that when raceway
groove curvature factors increase, Δf  increases correspondingly. As mentioned in Section 2.1 that the raceway groove
curvature factor plays an important role in calculating the contact force (Eqs. (6) and (7)). The contact force influences the
friction forces in contact zones, and also the skidding effect. As a result, the calculated BPFs change according to different
raceway groove curvature factors.
226 L. Niu et al. / Journal of Sound and Vibration 357 (2015) 207–232

Table 2
Effects of defect width at different shaft speeds.

Ωi (r min  1) wd (mm) BPFO (Hz) BPFI (Hz)

Dynamic model Calculated by Eq. (1) Dynamic model Calculated by Eq. (2)

3,000 0.5 295.82 295.01 404.14 404.99


3,000 1.5 295.68 295.01 404.14 404.99
3,000 3.5 295.61 295.01 404.15 404.99
3,000 5.0 295.75 295.01 404.08 404.99
6,000 0.5 596.05 590.02 803.86 809.99
6,000 1.5 595.83 590.02 803.95 809.99
6,000 3.5 594.20 590.02 804.25 809.99
6,000 5.0 594.40 590.02 804.66 809.99
10,000 0.5 1029.05 983.36 1304.25 1349.98
10,000 1.5 1026.65 983.36 1305.66 1349.98
10,000 3.5 1004.90 983.36 1306.90 1349.98
10,000 5.0 1009.50 983.36 1311.80 1349.98
15,000 0.5 1642.90 1475.04 1856.65 2024.96
15,000 1.5 1641.85 1475.04 1856.44 2024.96
15,000 3.5 1552.05 1475.04 1854.85 2024.96
15,000 5.0 1586.50 1475.04 1889.08 2024.96

It can be found from Fig. 17 that, the axial load is more significant than raceway groove curvature factors for
determinations
  of the skidding and the relative error. Under low axial load conditions (such as 500 N) where severe skidding
occurs, Δf  changes from 8.59% to 12.16% for BPFO, and changes from 7.85% to 8.86% for BPFI when raceway groove
curvature
 factors change from 0.51 to 0.55. However, when the axial load is large (such as 2000 N) where the skidding is
slight, Δf  just changes from 1.07% to 1.24% for BPFO, and only changes from 0.78% to 0.86% for BPFI.

3.2.4. Effects of initial contact angle


The carry capacity of a ball bearing largely depends on the initial contact angle. Bearings having large initial contact
angles can support heavier thrust loads. Generally, the initial contact angle does not exceed 40 degree [43]. Therefore, three
different initial contact angles (10 degree, 20 degree, and 30 degree) are chosen in this section. Moreover, different initial
contact angles are designed to support different combinations of radial and axial loads. As a result, the effects of initial
contact angle are investigated for a bearing loaded by different radial and axial loads.  
The effects of initial contact angle are shown in Fig. 18. It can be found from Fig. 18(a) that Δf  increases when the initial
contact angle increases. The effects of axial load on BPFs become more significant when the initial contact angle increases.
The reason is discussed here. When a bearing is loaded by pure axial load, the contact force of a ball can be determined
statically as F a =ðz sin α0 Þ [43] where α0 is the loaded contact angle. Moreover, larger α0 leads to larger α0 . It can be seen that
when α0 increases, the contact force decreases correspondingly, which results in high ratio of centrifugal force to contact
force when the shaft speed is fixed. As a result,  the skidding effect becomes severe when α0 increases (Fig. 18(b)). Under low
axial load conditions (such as F a ¼500 N), Δf  varies from 0.84% to 11.38% for BPFO, and changes from 0.59% to 8.25% for
BPFI when α0 changesfrom  10 degree to 30 degree. However, under large axial load conditions (such as F a ¼2000 N) where
the skidding is slight, Δf  varies from 0.65% to 1.09% for BPFO, and changes from 0.45% to 0.78% for BPFI when α0 changes
from 10 degree to 30 degree.
Moreover, the effects of initial contact angle on BPFs under different radial loads (Fig. 18(c)) can also be attributed to the
skidding effect (Fig. 18(d)). Under low radial load conditions (such as F r ¼0 N), Δf  varies from 0.84% to 11.38% for BPFO, and
varies from 0.59% to 8.25% for BPFI when α0 changes from  10 degree to 30 degree. However, under large radial load
conditions (such as F r ¼2000 N) where the skidding is slight, Δf  varies from 0.077% to 0.56% for BPFO, and changes from
0.015% to 0.4% for BPFI when α0 changes from 10 degree to 30 degree.

3.2.5. Effects of defect sizes


As discussed in Section 3.1, when a ball rolls through a defect, the orbital speed of the ball varies due to the impacts with the
defect. Therefore, the defect size, which influences the orbital speed of a ball, plays an important role in the BPFs calculated by
using dynamic models. In this section, the effects of defect width wd (Fig. 6(d)) on BPFs are discussed. As the defect mainly
affects the orbital speed of a ball, the bearing investigated in this section is operated at different shaft speeds. Moreover, the
bearing is loaded by an axial load of 1000 N. The depth of the maximum orthogonal shear stress, which is the main cause of
subsurface fatigue failure [43], is about 0.07 mm when this bearing is loaded by an axial load of 1000 N. Therefore, the defect
depth is sect as 0.1 mm to take account of the damage growth after a spall is generated on the raceway [35].
The effect of defect width is shown in Fig. 19. It can be found that the effect of defect width on BPFO is more  significant

than that on BPFI. The variations of relative error (i.e., the difference between the maximum and the minimum Δf  when wd
varies from 0.5 mm to 5.0 mm) of BPFO are 0.071%, 0.31%, 2.45%, and 8.19% when the shaft speeds are 3000 r min  1,
6000 r min  1, 10,000 r min  1, and 15,000 r min  1, respectively. For BPFI, the variations of relative error are about 0.0012%,
L. Niu et al. / Journal of Sound and Vibration 357 (2015) 207–232 227

Fig. 20. The setup of experiment 1.

Table 3
Parameters of the tested bearing (experiment 1).

Number of balls z 8
Ball diameter D (m) 7.938  10  3
Pitch diameter dm (m) 33.477  10  3
Initial contact angle ɑ0 (degree) 9.08
Radial clearance (m) 0

0.099%, 0.56%, and 1.69% when the shaft speeds are 3000 r min  1, 6000 r min  1, 10,000 r min  1, and 15,000 r min  1,
respectively. Therefore, when the shaft speed is low (less than 6000 r min  1), the influence of wd on the relative error is
below 0.3% for BPFO, and is below 0.1% for BPFI. Under high speed conditions (larger than 10,000 r min  1) where the
skidding becomes severe, the influence of wd on the relative error is larger than 2% for BPFO, and is larger than 0.5% for BPFI.
It can be found from Fig. 19(a) and (b) that the relative error of BPFO decreases first, and then increases with the increase
of wd , which shows that BPFs vary with the change of wd . The BPFs calculated by both the dynamic model and Eqs. (1) and
(2) are listed in Table 2. From Table 2, it can be seen that BPFO decreases when wd increases from 0.5 mm to 3.5 mm, which
shows that the orbital speed of the ball decreases when wd increases from 0.5 mm to 3.5 mm. However, when wd increases
to 5.0 mm, the BPFO increases slightly, which means that the orbital speed of a ball increases slightly (Fig. 19(c)). The
decrease of ball orbital speed when wd varies from 0.5 mm to 3.5 mm can be attributed to the impact force between the ball
and the back side of the defect which reduces the ball orbital speed (Fig. 19(d)). The impact force increases with the increase
of wd when wd varies from 0.5 mm to 3.5 mm. When wd increases to 5.0 mm, the slight increase of BPFO is because that the
impact force decreases slightly when wd increases from 3.5 mm to 5.0 mm. Similar trend of impact force with the change of
defect width was also reported in Ref. [42]. Additionally, BPFI increases with the increase of wd . This may be due to the
impact force between the ball and the back side of the defect (Fig. 10(g)) which makes the ball to accelerate in the defect
zone, as shown in Fig. 19(e). This impact force increases with the increase of wd , as shown in Fig. 19(f). Moreover, as shown
in Fig. 19(e), the decrease of the orbital speed of a ball outside the defect zone with the increase of wd , which means that the
relative speed between the inner raceway and the ball increases with the increase of wd , may be another reason for the
relationship between BPFI and wd .
Moreover, it can be found in Fig. 19(a) and (b) that the most important factor that affects the BPFs is the shaft speed
which affects the skidding effect. When the shaft speed is fixed, the main reason for the variation of relative error is the
impact force when a ball rolls through the defect.

4. Experimental verification

In order to validate our proposed method, two experiments are presented in this section. The first experiment was
carried out on a Spectra Quest Inc.'s mechanical fault simulator and the second experiment was carried out on an aerospace
bearing test rig.

4.1. Experiment 1

The Spectra Quest Inc's mechanical fault simulator is shown in Fig. 20. The tested bearing was radial loaded by a heavy
mass (about 50 N). The data sampling rate was set as 6.4 kHz. The tested bearing is a 3/4 inch rotor bearing whose
parameters are listed in Table 3. Material parameters are assumed the same as those listed in Table 1. The tested bearing is
lubricated by grease.
228 L. Niu et al. / Journal of Sound and Vibration 357 (2015) 207–232

Table 4
Experiment and simulation results (BPFO, experiment 1).

fs (Hz) Ωi (r min  1) Experiment result (Hz) fd (Hz) Error of fd (%)

9.79 588 30.22 30 0.73


19.74 1185 60.69 59.90 1.30
29.72 1783.2 90.9 90.80 0.11
39.72 2383.2 121.7 121.76 0.049

Table 5
Experiment and simulation results (BPFI, experiment 1).

fs (Hz) Ωi (r  min  1) Experiment result (Hz) fd (Hz) Error of fd (%)

9.82 589.7 49.11 48.34 1.57


19.78 1187.8 98.47 97.68 0.80
29.76 1786.3 147.66 146.85 0.55
39.73 2384 196.83 196.01 0.42

Fig. 21. Setup up of the aerospace bearing test rig.

Table 6
Parameters of the tested bearing H7018C.

Number of balls z 27
Ball diameter D (m) 11.113  10  3
Pitch diameter dm (m) 115  10  3
Curvature factor of inner raceway fi 0.56
Curvature factor of outer raceway fo 0.54
Initial contact angle α0 (degree) 15
Radial clearance (m) 0

In the simulation procedure, the groove curvature factors of inner raceway and outer raceway were both set as 0.52
which lies within the range of the most angular contact ball bearings [43]. Moreover, the friction coefficient was set as 0.02.
This value is in the reasonable range of some grease lubrications under low external loads [58]. Besides the friction
coefficient, the other lubrication parameter um was assumed as 1 m s  1. The determination of the parameter um is beyond
the scope of the current analysis. The clearances at ball/cage pocket and cage/guiding raceway were set as 0.1 mm and
0.25 mm, respectively. The width and depth of the defect were set as 0.5 mm and 0.1 mm, respectively. From the above
investigations, it can be found that the ball/cage clearance and the defect width have relatively slight effects on BPFs at low
shaft speeds.
In order to verify the proposed method, the experiment was carried out under different shaft speeds. The maximum
shaft frequency of the experiment was about 40 Hz due to the limitation of the test rig. The experiment and simulation
results of BPFO and BPFI are listed in Tables 4 and 5. It can be seen in Tables 4 and 5 that the BPFs calculated by dynamic
model are all close to the experiment results. For BPFO, the errors of the proposed method are all less than 1.3 %. For BPFI,
the maximum error is 1.57 % and the minimum error is 0.42 %. The errors of the simulation results may be attributed to the
simple lubrication model used in this model which cannot simulate the behaviors of grease accurately. Moreover, the
instantaneous angular speed variations of the motor were not considered.
L. Niu et al. / Journal of Sound and Vibration 357 (2015) 207–232 229

Fig. 22. A spall on the inner raceway.

1.5

99.8Hz
Amplitude (m s-2)

1 1474.7 Hz
1574.5Hz

0.5 1374.9Hz

0
0 500 1000 1500 2000
Frequency (Hz)

Fig. 23. Experiment result (envelope spectrum, experiment 2).

1.5 100.6
Amplitude (m s )
-2

0.5 1474.1
1574.7
1373.4

0
0 500 1000 1500 2000
Fig. 24. Simulation result (envelope spectrum, corresponding to experiment 2).

4.2. Experiment 2

The second experiment was carried out on an aerospace test rig shown in Fig. 21. This experiment was a part of the
bearing life prediction program held in Xi’an Jiaotong University [59,60]. The tested bearing is a H7018C angular contact ball
bearing. Parameters of the tested bearing are listed in Table 6. Similarly, the material parameters were assumed the same as
those listed in Table 1, and the lubrication parameter um (refer to Fig. 2) was also assumed as 1 m s  1. The friction coefficient
was set as 0.03 in the simulation. The clearances at ball/cage pocket and cage/guiding raceway were both set as 0.1 mm. Two
accelerometers were mounted on the sleeve that was connected with the outer ring of the bearing. The sampling rate was
20 kHz. The shaft rotation speed was about 6000 r min  1.
A radial load of 11 kN and an axial load of 2 kN were applied to the test bearing by the hydraulic loading system. After
about 146 h, a spall was generated on the inner raceway as shown in Fig. 22. The width of the spall is about 1.5 mm. The
envelope spectrum of the experiment result is shown in Fig. 23. The shaft rotation frequency f s (99.82 Hz) is evident in this
spectrum. Furthermore, a component 1474.7 Hz which has sidebands spaced at f s exists. This frequency (1474.7 Hz) is
related to the BPFI of the current experiment.
The simulation result is 1474.1 Hz as shown in Fig. 24. The possible reason for the difference between the simulation
result (1474.1 Hz) and the experiment result (1474.7 Hz) is discussed here. In the experiment, the bearing is lubricated by
230 L. Niu et al. / Journal of Sound and Vibration 357 (2015) 207–232

aviation engine oil. The shear stress at contacting surface largely depends on external loads, shaft speeds and thermal
effects. In this situation, the friction forces between the ball and the raceway should be calculated by elastohydrodynamic
lubrication method. However, a simple lubrication model shown in Fig. 2 was adopted in the simulation. This simple
lubrication model cannot describe the behaviors of lubricating oil exactly. Moreover, the instantaneous angular speed
variations of the motor were also not considered, this may be the reason why the shaft frequency in the envelope spectrum
obtained by the simulation (100.6 Hz) is different from the experiment result (99.82 Hz). Other reasons for the difference
between the shaft frequency calculated by simulation and that obtained by experiment may be that the re-sampling
procedure introduces some errors into the envelope analysis, and may also be the positive percentage skidding in the
bearing under relative high speed conditions.

5. Discussions on limitations of the proposed method

The main limitation of the proposed method is that a simple lubrication model is adopted to calculate BPFs. As
mentioned in Ref. [26], this lubrication model may overly simplify the behavior of any lubricating oil, in which cases the
lubrication parameters will be a function of temperature, rolling speeds, Hertzian stresses, etc., and such variations may
often be quite significant. It is, therefore, necessary to adopt elastohydrodynamic lubrication model for simulating bearings’
dynamic characteristics. Moreover, the main focus of the proposed model was on the localized surface defects. Spalls and
dents are the two main kinds of the fault patterns modeled in this paper. However, some other more complex fault patterns
can be formed in the wear evolution of a bearing [35]. The effects of these fault patterns on vibrations and BPFs of a ball
bearing are not modeled. Additionally, the effects of non-Hertizian contact characteristic and transfer path are not
considered. These topics will be investigated in our future work.

6. Conclusions

 For a loaded ball bearing, the combination of rolling and sliding in contact region occurs, and the BPFs calculated by
simple kinematical relationships are inaccurate, especially for severe skidding conditions where the shaft speed Ωi is
high, the axial load F a and the radial load F r are low, the friction coefficient μ1 is small, and raceway groove curvature
factors (f i and f o ) and the initial contact angle α0 are large. The dominant parameters which largely influence BPFs are
Ωi , F a , F r and α0 . These parameters largely affect the skidding effect, and they are more significant than μ1 , f i and f o for
the calculation of BPFs. For conditions where Ωi is larger than 10,000 r min  1, F a and  F r are less than 500 N, and α0 is
larger than 20 degree, the skidding effect is severe for the investigated bearing, and Δf  is larger than 3% for both BPFO
and BPFI. However, for conditions where Ωi is less than 3000 r min  1, F a and F r are larger than 2000 N, and α0 is less
than 20 degree, the skidding in the investigated bearing becomes slight, and Δf  is about less than 1%.
 The spacing between two adjacent balls largely depends on the operating conditions and ball/cage pocket clearance δbp .
The hypothesis of 1–2% variation of the impulse spacing in a defective ball bearing reported by previous investigations
can be satisfied only for the conditions where the skidding effect is slight.
 When the shaft speed is low (about less than 6000 r min  1 for the investigated bearing), the skidding in the bearing is
slight, and the influence of wd on the relative error is below 0.3% for BPFO, and is below 0.1% for BPFI. For high speed
conditions (about larger than 6000 r min  1 for the investigated bearing) where the skidding becomes severe, the
influence of wd on the relative error is larger than 2% for BPFO, and is larger than 0.5% for BPFI. Moreover, the effect of wd
on BPFO is more significant than that on BPFI.

Acknowledgments

We would like to thank Prof. Xuefeng Chen, Dr. Zhongjie Shen and Dr. Chuang Sun for providing the aerospace bearing
test bench and the measured data. This work is supported by National Natural Science Foundation of China (no. 51421004),
National Basic Research Program of China (no. 2011CB706606), the National Science and Technology Major Project (no.
2014ZX04001-191-01), and the Fundamental Research Funds for the Central University.

References

[1] R. Yan, R.X. Gao, X. Chen, Wavelets for fault diagnosis of rotary machines: a review with applications, Signal Processing 96 (2014) 1–15.
[2] S. Wang, W. Huang, Z.K. Zhu, Transient modeling and parameter identification based on wavelet and correlation filtering for rotating machine fault
diagnosis, Mechanical Systems and Signal Processing 25 (4) (2011) 1299–1320.
[3] H. Sun, Z. He, Y. Zi, J. Yuan, X. Wang, J. Chen, S. He, Multiwavelet transform and its applications in mechanical fault diagnosis—a review, Mechanical
Systems and Signal Processing 43 (1-2) (2014) 1–24.
[4] C. Zhang, B. Li, B. Chen, H. Cao, Y. Zi, Z. He, Periodic impulsive fault feature extraction of rotating machinery using dual-tree rational dilation complex
wavelet transform, Journal of Manufacturing Science and Engineering 136 (5) (2014) 051011.
[5] R. Yan, R.X. Gao, Rotary machine health diagnosis based on empirical mode decomposition, Journal of Vibration and Acoustics 130 (2) (2008) 021007.
[6] J. Antoni, Cyclostationarity by examples, Mechanical Systems and Signal Processing 23 (4) (2009) 987–1036.
[7] J. Antoni, The spectral kurtosis: a useful tool for characterising non-stationary signals, Mechanical Systems and Signal Processing 20 (2) (2006) 282–307.
L. Niu et al. / Journal of Sound and Vibration 357 (2015) 207–232 231

[8] J. Antoni, R.B. Randall, The spectral kurtosis: application to the vibratory surveillance and diagnostics of rotating machines, Mechanical Systems and
Signal Processing 20 (2) (2006) 308–331.
[9] Z. Feng, M.J. Zuo, R. Hao, F. Chu, J. Lee, Ensemble empirical mode decomposition-based teager energy spectrum for bearing fault diagnosis, Journal of
Vibration and Acoustics 135 (3) (2013) 031013.
[10] J. Li, X. Chen, Z. Du, Z. Fang, Z. He, A new noise-controlled second-order enhanced stochastic resonance method with its application in wind turbine
drivetrain fault diagnosis, Renewable Energy 60 (2013) 7–19.
[11] R.B. Randall, J. Antoni, Rolling element bearing diagnostics—a tutorial, Mechanical Systems and Signal Processing 25 (2) (2011) 485–520.
[12] P.D. McFadden, J.D. Smith, Model for the vibration produced by a single point defect in a rolling element bearing, Journal of Sound and Vibration 96 (1)
(1984) 69–82.
[13] P.D. McFadden, J.D. Smith, The vibration produced by multiple point defects in a rolling element bearing, Journal of Sound and Vibration 98 (2) (1985)
263–273.
[14] N. Tandon, A. Choudhury, An analytical model for the prediction of the vibration response of rolling element bearings due to a localized defect, Journal
of Sound and Vibration 205 (3) (1997) 275–292.
[15] A. Choudhury, N. Tandon, Vibration response of rolling element bearings in a rotor bearing system to a local defect under radial load, Journal of
Tribology 128 (2) (2006) 252–261.
[16] J.R. Stack, T.G. Habetler, R.G. Harley, Fault-signature modeling and detection of inner-race bearing faults, IEEE Transactions on Industry Applications 42
(1) (2006) 61–68.
[17] F. Cong, J. Chen, G. Dong, Michael Pecht, Vibration model of rolling element bearings in a rotor-bearing system for fault diagnosis, Journal of Sound and
Vibration 332 (8) (2013) 2081–2097.
[18] H. Cao, L. Niu, Z. He, Method for vibration response simulation and sensor placement optimization of a machine tool spindle system with a bearing
defect, Sensors 12 (7) (2012) 8732–8754.
[19] T.A. Harris, M.N. Kotzalas, Rolling Bearing Analysis: Advanced Concepts of Bearing Technology, Taylor & Francis, Boca Raton, 2007.
[20] D. Ho, R.B. Randall, Optimisation of bearing diagnostic techniques using simulated and actual bearing fault signals, Mechanical Systems and Signal
Processing 14 (5) (2000) 763–788.
[21] J. Antoni, R.B. Randall, Differential diagnosis of gear and bearing faults, Journal of Vibration and Acoustics 124 (2) (2002) 165–171.
[22] J. Antoni, R.B. Randall, A stochastic model for simulation and diagnostics of rolling element bearings with localized faults, Journal of Vibration and
Acoustics 125 (3) (2003) 282–289.
[23] N. Sawalhi, 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 22 (8) (2008) 1924–1951.
[24] N. Sawalhi, 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 22 (8) (2008) 1952–1966.
[25] D. Brie, Modelling of the spalled rolling element bearing vibration signal: an overview and some new results, Mechanical Systems and Signal Processing
14 (3) (2000) 353–369.
[26] P.K. Gupta, Advanced Dynamic of Rolling Elements, Springer-Verlag, New York, NY, 1984.
[27] A.B. Jones, A general theory of elastically constrained ball and radial roller bearings under arbitrary load and speed conditions, Journal of Fluids
Engineering 82 (2) (1960) 309–320.
[28] G. Jang, S.W. Jeong, Vibration analysis of a rotating system due to the effect of ball bearing waviness, Journal of Sound and Vibration 269 (3-5) (2004)
709–726.
[29] C. Bai, Q. Xu, Dynamic model of ball bearings with internal clearance and waviness, Journal of Sound and Vibration 294 (1-2) (2006) 23–48.
[30] S. Singh, U. Köpke, C. Howard, D. Petersen, Analyses of contact forces and vibration response for a defective rolling element bearing using an explicit
dynamics finite element model, Journal of Sound and Vibration 333 (2014) 5356–5377.
[31] D. Petersen, C. Howard, Z. Prime, Varying stiffness and load distributions in defective ball bearings: analytical formulation and application to defect
size estimation, Journal of Sound and Vibration 337 (2015) 284–300.
[32] D. Petersen, C. Howard, N. Sawalhi, A. Moazen Ahmadi, 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 50-51 (2015) 139–160.
[33] A. Moazen Ahmadi, D. Petersen, C. Howard, A nonlinear dynamic vibration model of defective bearings—the importance of modelling the finite size of
rolling elements, Mechanical Systems and Signal Processing 52-53 (2015) 309–326.
[34] G. Kogan, R. Klein, A. Kushnirsky, J. Bortman, Toward a 3D dynamic model of a faulty duplex ball bearing, Mechanical Systems and Signal Processing 54-
55 (2015) 243–258.
[35] I. El-Thalji, E. Jantunen, Dynamic modelling of wear evolution in rolling bearings, Tribology International 84 (2015) 90–99.
[36] S. Khanam, J.K. Dutt, N. Tandon, Impact force based model for bearing local fault identification, Journal of Vibration and Acoustics 137 (2015) 051002.
[37] J. Liu, Y. Shao, W.D. Zhu, A new model for the relationship between vibration characteristics caused by the time-varying contact stiffness of a deep
groove ball bearing and defect sizes, Journal of Tribology 137 (2015) 031101.
[38] P.K. Gupta, Dynamics of rolling-element bearings—Part I: Cylindrical roller bearing analysis, Journal of Tribology 101 (3) (1979) 293–302.
[39] P.K. Gupta, Dynamics of rolling-element bearings—Part III: Ball bearing analysis, Journal of Tribology 101 (3) (1979) 312–318.
[40] F. Wang, M. Jing, J. Yi, G. Dong, H. Liu, B. Ji, Dynamic modelling for vibration analysis of a cylindrical roller bearing due to localizes defects on raceways,
Proceedings of the Institution of Mechanical Engineers, Part K: Journal of Multi-body Dynamics 229 (2015) 39–64.
[41] Y. Li, H. Cao, L. Niu, X. Jin, A general method for the dynamic modeling of ball bearing-rotor systems, Journal of Manufacturing Science and Engineering
137 (2) (2014) 021016.
[42] L. Niu, H. Cao, Z. He, Y. Li, Dynamic modeling and vibration response simulation for high speed rolling ball bearings with localized surface defects in
raceways, Journal of Manufacturing Science and Engineering 136 (4) (2014) 041015.
[43] T.A. Harris, M.N. Kotzalas, Rolling Bearing Analysis: Essential Concepts of Bearing Technology, Taylor&Francis, Boca Raton, 2007.
[44] N. Ghaisas, C.R. Wassgren, F. Sadeghi, Cage instabilities in cylindrical roller bearings, Journal of Tribology 126 (4) (2004) 681–689.
[45] A. Ashtekar, F. Sadeghi, A new approach for including cage flexibility in dynamic bearing models by using combined explicit finite and discrete
element methods, Journal of Tribology 134 (2012) 041502.
[46] N. Weinzapfel, F. Sadeghi, A discrete element approach for modeling cage flexibility in ball bearing dynamics simulations, Journal of Tribology 131
(2009) 021102.
[47] J. Takabi, M.M. Khonsari, On the influence of traction coefficient on the cage angular velocity in roller bearings, Tribology Transactions 57 (5) (2014)
793–805.
[48] M.D. Brouwer, F. Sadeghi, A. Ashtekar, J. Archer, C. Lancaster, Combined explicit finite and discrete element models for rotor bearing dynamic
modeling, Tribology Transactions 58 (2015) 300–315.
[49] I.K. Epps, An Investigation into Vibrations Excited by Discrete Faults in Rolling Element Bearings, School of Mechanical Engineering, The University of
Canterbury, Christchurch, New Zealand, 1991 Ph.D. Thesis.
[50] S. Singh, C. Howard, C. Hansen, An extensive review of vibration modelling of rolling element bearings with localised and extended defect, Journal of
Sound and Vibration (2015), http://dx.doi.org/10.1016/j.jsv.2015.04.037i.
[51] L. Cui, L.Q. Wang, D.Z. Zheng, L. Gu, Study on critical load of skidding in high speed ball bearing, Journal of Aerospace Power 22 (11) (2007) 1971–1976.
in Chinese.
[52] H. Prashad, The effect of cage and roller slip on the measured defect frequency response of rolling-element bearings, ASLE Transactions 30 (3) (1987)
360–367.
232 L. Niu et al. / Journal of Sound and Vibration 357 (2015) 207–232

[53] N.T. Liao, J.F. Lin, Ball bearing skidding under radial and axial loads, Mechanism and Machine Theory 37 (1) (2002) 91–113.
[54] H.R. Cao, Z.J. He, Y.Y. Zi, Modeling of a high-speed rolling bearing and its damage mechanism analysis, Journal of Vibration and Shock 31 (19) (2012)
134–140. in Chinese.
[55] X. Liu, Dynamics Analysis Model of High-speed Rolling Bearings and Dynamic Performance of Cages, Dalian University of Technology, Dalian, PR China,
2011 in Chinese.
[56] P.K. Gupta, Traction coefficients for some solid lubricants for rolling bearing dynamics modeling, Tribology Transactions 43 (4) (2000) 647–652.
[57] S.E. Deng, Q.Y. Jia, Y.S. Wang, Design Principles for Roller Bearings, Standards Press of China, Beijing, 2008 in Chinese.
[58] J. Chen, Research on Preparation of Titanium Complex Grease and its Tribological Properties and Biodegradability Ph.D. Thesis, Harbin Institute of
Technology, Harbin, P. R. China, 2010 in Chinese.
[59] C. Sun, Z. Zhang, Z. He, Z. Shen, B. Chen, Novel method for bearing performance degradation assessment—a kernel locality preserving projection-based
approach, Proceedings of the Institution of Mechanical Engineers, Part C: Journal of Mechanical Engineering Science 228 (3) (2014) 548–560.
[60] Z.J. Shen, The Research of Multivariable Support Vector Machine with Application in Life Prediction Ph.D. Thesis, Xi’an Jiaotong University, Xi’an, PR China,
2012 in Chinese.

Das könnte Ihnen auch gefallen