Anda di halaman 1dari 9

International Journal of Heat and Mass Transfer 68 (2014) 8593

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


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

Numerical investigation of the airoil two-phase ow inside an oil-jet


lubricated ball bearing
Jibin Hu, Wei Wu , Mingxing Wu, Shihua Yuan
National Key Laboratory of Vehicular Transmission, Beijing Institute of Technology, Beijing 100081, PR China

a r t i c l e

i n f o

Article history:
Received 3 June 2013
Received in revised form 7 September 2013
Accepted 8 September 2013
Available online 3 October 2013
Keywords:
Ball bearings
Oil-jet
Two-phase ow
Volume of uid
CFD
Precise lubrication

a b s t r a c t
The airoil two-phase ow inside an oil-jet lubricated ball bearing was investigated by the CFD method.
The VOF method for multiphase ow was used to track the airoil two-phase ow. The sliding mesh
plane was applied between the ow eld inside the bearing and the ow eld on both outsides of the
bearing. The results suggest that the airoil distribution inside the oil-jet lubricated ball bearing is not
uniform. The lowest oil volume fraction appears in the upstream side near the nozzle. The ow velocity
and the pressure of the airoil phase increase inside a faster bearing. The average oil volume fraction
becomes lower as the speed increases. With the increase of the oil ow rate, the oil volume fraction in
the bearing and the drag against the rotation of the rolling elements and cage increase. The nonuniformity of the airoil distribution is also enhanced. The nonuniform airoil distribution should be considered for the heat transfer analysis of the oil-jet lubricated ball bearing. Optimizing the oil-jet
lubrication according to the airoil distribution inside the oil-jet lubricated ball bearing is signicant
for the high-speed rotation to achieve lower drag loss and greater cooling effect.
2013 Elsevier Ltd. All rights reserved.

1. Introduction
The rolling bearing is commonly used as an important component in various types of machinery. For most low-speed machinery
applications, grease and simple splash lubrication are the main
lubrication means [1]. For high-speed systems, such as aircraft engines, turbomachinery, and drive trains, oil-jet lubrication method
is applied to the bearing lubrication [2]. The development of the
vehicular transmission is now focused on lighter and more compact designs. A higher rated speed is benecial to a greater power
density. To ensure the high-speed capability of main shaft bearings
of the vehicular transmission, an effective lubrication is necessary.
Further, coupled with the compact design requirement of the
vehicular transmission, there is no special bearing chamber commonly used in aircraft engines [3]. The operation speed of a vehicular transmission often circulates not only among the high speed
but also among the low speed, which is a challenge for improving
the lubrication efciency. To a large extent, the lubrication system
has been based on empirical correlations derived from research
and knowledge gathered through years of experience and from
time-consuming experiments. To improve the lubrication efciency and achieve precise lubrication, a further investigation into

Corresponding author. Address: Room 412, Building 9, School of Mechanical


Engineering, Beijing Institute of Technology, Beijing 100081, PR China. Tel.: +86 10
68914786; fax: +86 10 68944487.
E-mail address: wuweijing@bit.edu.cn (W. Wu).
0017-9310/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ijheatmasstransfer.2013.09.013

the oil-jet lubrication performance of the rolling bearing is


important.
The lubrication performance of the rolling bearing has been
investigated a lot by the single-phase method [46]. The non-Newtonian lubrication characteristics are considered [7]. The works
mainly focus on the characterization of point contact oil lm, such
as contact force [8], friction [9], and testing [10]. Regarding the oilow through a bearing compartment, only a relatively small part of
the oil is used to lubricate the bearings. Most of the oil is used to
provide sufcient cooling capability and control the bearing temperature by removing generated heat [4]. The evolution of temperature in the bearing is mainly estimated by the method presented
by Harris [11]. The convective heat transfer coefcients are given
considered the lubricant oil-ow and the air-ow, respectively
[12]. However, the ow eld inside the oil-jet lubrication rolling
bearing is complicated because there is a strong interaction
between the air-ow and the lubricant oil-ow which is greatly
affected by the bearing rotary speed. Further, a larger amount of
oil mass ow leads to a higher power loss since the drag against
the rotation of the rolling elements and cage is increased. Too small
oil mass ow may cause a lubrication failure. A sufcient
assessment of the oil-jet lubrication rolling bearing regarding the
air-ow and the lubricant oil-ow is of particular importance. This
paper investigates the process when one-phase (oil) ow turns to
the airoil two-phase ow inside an oil-jet lubricated ball bearing.
A CFD modeling method is presented for the airoil two-phase ow
inside the rotary ball bearing. It is aimed to increase the level of

86

J. Hu et al. / International Journal of Heat and Mass Transfer 68 (2014) 8593

Nomenclature
d
Db
F
g
M
n
na
nb
nw
p
P
Q
Saoil
t
tw
T

section circle diameter of bearing


ball diameter
external force
gravity acceleration
moment applied by the airoil two-phase ow
rotary speed of the inner ring
surface normal at the wall
rotary speed of the ball and cage
unit normal vector of the wall surface
pressure
power of the spindle motor
oil ow rate
source term
time
unit tangent vector of the wall surface
temperature

understanding of the ow eld inside the oil-jet lubricated ball


bearing. The results can be used for the precise oil-jet lubrication
design to control the oil volume fraction inside the ball bearing
at a later stage.
2. Oil-jet lubricated ball bearing
A schematic diagram of the oil-jet lubrication method for the
ball bearing is shown in Fig. 1. In the oil-jet lubrication, the lubricant oil is injected into the ball bearing through a small-diameter
nozzle. The nozzle is aimed directly at the inner raceway surface
and is located near the face of the ball bearing. Further, the
oil-jet lubrication method recirculates the oil used as the lubricant.
The location and number of nozzles, orice size, jet distance from
the bearing, and jet velocity affect the oil-jet lubrication
performance [13]. The removal of the lubricant oil from the bearing
is important for the bearing temperature and power loss [2]. The
experience with the oil-jet lubrication has also shown that the
measured bearing temperature after the bearing reaches thermal
equilibrium is a function of the bearing load, speed, oil-in temperature and oil volume fraction inside the bearing.
3. Mathematical modeling
3.1. Two-phase ow model
The analysis of the whole ow passage inside the oil-jet lubricated ball bearing involved the air and oil two-phase ow. To track

Greek symbols
a
bearing contact angle
hw
angle between the wall and the tangent of the interface
at the wall
l
dynamic viscosity
lair
air dynamic viscosity
loil
oil dynamic viscosity
q
mixture density
qair
air density
qoil
oil density
uair
air volume fraction
uoil
oil volume fraction
s
shear force applied by the airoil two-phase ow
t
velocity vector
x
circumferential azimuth

the airoil two-phase ow, the volume of uid (VOF) method for
multiphase ow is used. The VOF method was developed by Hirt
and Nichols [14] in 1981. It is a surface tracking technique and
the interface between two or more immiscible uids is of interest.
The summation of the volume fractions of all phases in each control volume is set to unity. The VOF method has reasonable accuracy, is less computationally intensive, is relatively simple to use,
and can solve highly complex ows. The VOF method is probably
one of the most convenient methods relative to other two-phase
ow modeling methods [15].
For the two-phase ow modeling inside the oil-jet lubricated
ball bearing, a volume fraction uoil is used to mark the volume fraction of the oil phase in the VOF method. Then, uoil = 0 represents a
cell that is empty of the oil and uoil = 1 represents a cell that is full
of the oil. If 0 < uoil < 1, it represents the interface between the oil
phase and the air phase. Thus, the uoil is dened by

8
uoil 1
in the oil
>
>
>
<u 0
in the air
oil
>
0
<
u
<
1
at the interface
>
oil
>
:

The subscript oil represents the oil phase properties. The interface between the oil and air phases is tracked by solving a continuity equation for the volume fraction of the oil phase. The continuity
equation for the volume fraction of the oil phase is given by:

@
u q r  uoil qoil~
t Saoil
@t oil oil

where qoil is the oil density, ~


t is the velocity vector. The mass source
term, Saoil , on the right-hand side is zero for the VOF method signifying that there is no mass transfer across the interface [16,17].
Because there are two phases, Eq. (2) for the volume fraction is
solved only for the secondary phase that is dened as the oil phase.
The volume fraction for the primary phase that is dened as the air
phase is obtained by the following constraint:

uoil uair 1

where uair is the air volume fraction. The subscript air represents
the air phase properties. In the VOF method, a single momentum
equation is solved. The elds for all variables and properties are
shared by both phases according to the following equation:

Fig. 1. Oil-jet lubrication method for the ball bearing.

@
F
q~
t r  q~
t~
t rp r  lr~
t r~
tT  q~
g ~
@t

J. Hu et al. / International Journal of Heat and Mass Transfer 68 (2014) 8593

where q is the mixture density, l is the dynamic viscosity, p is the


F is the external force due to
pressure, ~
g is the gravity acceleration, ~
the surface tension at the interface [18]. The properties that appear
in the momentum equation are volume-fraction-averaged properties. The density and the dynamic viscosity for the ail-oil two-phase
ow are given by

q uoil qoil uair qair

l uoil loil uair lair

where qair is the air density, lair is the air dynamic viscosity, loil is
the oil dynamic viscosity.
A wall adhesion angle in conjunction with the surface tension
model is considered in the simulation, since most lubricant oil
ows along the surfaces of the inner race, rolling elements and
cage of the high-speed ball bearing. The curvature of the surface
near the wall, where the airoil interface meets the solid surface,
is adjusted. The local curvature of this interface is determined by
the static contact angle, hw, which is the angle between the wall
and the tangent of the interface at the wall. The surface normal
at the wall is given by,

~
t w sin hw
na ~
nw cos hw ~

t w are the unit


where ~
na is the surface normal at the wall. ~
nw and ~
normal vector and tangent vector of the wall surface, respectively.
The oil-only Reynolds number is 2428 at an inner ring speed of
2000 r/min. Further, the value of the air-only Reynolds number
reaches 4154 at an inner ring speed of 2000 r/min. Thus, the air
oil two-phase ow eld cannot be treated as a laminar ow in
the high-speed rotation. Further, considering the inuence of the
high-speed rotation and the swirl ow, the ke renormalization
group (RNG) turbulence model is employed in this study. The
RNG ke model considers the inuence of high strain rate, large
curvature overowing and other factors, which can improve the
accuracy under rotational ow [19]. All of the above governing
equations were implemented in the ANSYS Fluent code.
3.2. Boundary conditions
The mesh images of the ow eld are shown in Fig. 2. Since the
lubricant oil ows out from both sides of the bearing, two pressure-outlets are set to fully observe the lubricant oil outlet situation. The mesh model includes inner ring, outer ring, rolling
elements, cage, volume of the pressure-outlets, and nozzles. The
ow eld inside the bearing was divided by a tetrahedral unstructured mesh. The nozzles and the ow area on both outsides of the
bearing were divided by a structured hexahedral mesh, as shown
in Fig. 2. The mesh density was designed following the research

87

purpose. A mesh scheme with 118,709 elements and 97,032 nodes


was nally adopted for the ow eld. Considering that the ball
bearing is a rotary component, static and dynamic interferences
are formed between the ow eld inside the bearing and the ow
area on both sides of the bearing. Thus, the sliding mesh plane was
used at the edge of the ow eld inside the bearing to complete the
data transfer in the entire computation domain.
The nozzle was set as the only mass ow inlet of the entire computation domain and the mass ow value is determined by different operations. The pressure of both pressure-outlets was a
standard atmospheric pressure. Due to collision of the lubricant
oil with the rolling elements and cage, some lubricant oil splashes
back. The oil spill outlet boundary condition was set as the right
side ow eld end face of the bearing. Different walls were also
added to the rotary boundary conditions. A no-slip boundary condition was adopted at the wall of low Reynolds number and the
standard wall function was adopted for near-wall treatment.
3.3. Numerical method
The rotation of the inner ring drives the rolling elements and
cage. After lubricant oil enters the bearing, it rotates around the inner ring due to the drive of the rolling elements and cage. It is necessary to calculate the speed of the rolling elements and cage. The
rotary speed of the rolling elements and cage [11] is given by

nb



1
Db cos a
n 1
2
d

where nb is the rotary speed of the rolling elements and cage, n is


the rotary speed of the inner ring, Db is the ball diameter, a is the
bearing contact angle, d is the pitch diameter of the bearing.
Many CFD solutions have been proposed to solve the complicated ow problems [20,21]. The airoil two-phase ow inside
the oil-jet lubricated ball bearing was investigated numerically
using the ANSYS Fluent code. The nite volume method was used
to discretely solve the governing equation. The central difference
scheme was adopted for the diffusion and pressure terms of the
momentum equation. The second-order upwind difference was
adopted for the convection terms. The pressure was adopted for
the PRESTO! (PREssure STaggering Option) format. The semi-implicit method for pressure-linked equations-consistent (SIMPLEC)
method was adopted for the coupling solution of pressure and
velocity. In the calculation, the air phase was set as the compressible main phase. The oil phase was set as the incompressible second phase. The volume fraction of the oil phase in the nozzle
inlet was set as 1.0. The volume fraction of the oil phase in the
other ow eld region was initialized to zero. The geometric reconstruction scheme from the work of Youngs was used to represent

Fig. 2. Calculating grid for the whole domain.

88

J. Hu et al. / International Journal of Heat and Mass Transfer 68 (2014) 8593

the interfaces between two uids, which is the most accurate one
for interface tracking [22]. The mass ow conservation between
the inlets and outlets was used as the calculation convergence
condition. The simulated results of different speeds, ow rates,
oil viscosities and nozzles will be presented in detail.
4. Experimental apparatus
To investigate the lubricant behavior of the ball bearing, an
experimental apparatus has been designed and constructed, as
shown in Fig. 3. The experimental apparatus is designed for installing rolling bearing to operate at variable speeds up to 15,000 r/min.
The speed of motor could be set independently by a variable frequency drive. The motor and the test bearing are connected by
support bearings and a shaft. Fig. 3(a) shows a schematic view of
the test rig. The oil-jet lubrication is used for the support bearings
in the experimental apparatus. The test bearing is lubricated by the
oil-jet lubrication on the unloaded side of the inner ring. A hydraulic system is separately designed to apply the external radial and
axial forces on the test bearing. As Fig. 3(b) shows, there are four
temperature sensors attached to the outer ring of the test bearing.
One more temperature sensor is mounted on the inner ring.
Another temperature sensor is placed near the oil-jet nozzle to
measure the oil temperature. A data acquisition system is also employed to collect and transfer the data to a computer for further
analysis. The technical data of the experimental apparatus is presented in Tables 1 and 2 lists the specications of the tested ball
bearing.
5. Results and discussion
5.1. Nonuniform airoil distribution and validation
Fig. 4 shows the simulated airoil distribution of different number of nozzles. The speed of the inner ring is 10,000 r/min and the
oil-ow rate is 3.0 L/min. The oil viscosity is 0.02 Pa s in the calculation. It is seen that the circumferential airoil distribution in the

Fig. 3. Experimental apparatus. (a) A schematic view of the test rig. (b) Sensors and
nozzles of the test rig.

Table 1
Technical data of the experimental apparatus.
Apparatus and sensor

Technical data

Motorized spindle
Temperature sensor
Oil ow transducer
Vibration transducer
External radial force
External axis force

015000 r/min
Pt1000, 70500 C
FT-110, 1.010 L/min
JHT-II-B, 15 g
Hydraulic loading, 030 kN
Hydraulic loading, 030 kN

Table 2
Specications of the test ball bearing.
Bearing type

SKF 7210 BECBY

Inner diameter (mm)


Outer diameter (mm)
Width (mm)
Ball diameter (mm)
Number of balls
Contact angle (deg.)

50
90
20
12.186
14
40

bearing is not uniform. The volume fraction of the oil phase quite
close to the nozzle is much more than other positions inside the
bearing. The oil volume fraction decreases along the rotation direction of the bearing. Further, the radial airoil two-phase ow also
distributed unevenly due to the centrifugal force, as shown in
Fig. 4(c). The airoil phase with larger density is thrown to the outer race of the bearing. The cooling effect of the oil-jet lubrication on
the inner ring becomes low, especial at higher speeds. The circumferential airoil distributions under different injection methods are
demonstrated in Fig. 4(d) and (e). The measured temperatures of
the outer ring are also proposed. Only the axial load was applied
during the test. Compared with the single-nozzle jet, the airoil
distribution becomes more uniform with the dual-nozzle jet at
the same oil-ow rate. The measured bearing temperature has followed the same trend. The measured temperature of the outer ring
becomes more homogeneous under the dual-nozzle jet. The lowest
volume fraction of the oil phase under both the single-nozzle jet
and the dual-nozzle jet appears in the upstream side near the nozzle. It is because that most of the lubricant oil ows along the circumferential direction due to the agitation of the rolling elements
and cage. The lubricant oil leaves the bearing from the end surface
of the outer ring at the same time. Thus, the oil volume fraction decreases continuously in the circumferential direction. However, the
oil volume fraction increases near the nozzle since some of the lubricant oil splashes out onto the upstream side near the nozzle.
The values of lubricant oil circumferential, radial and axial
velocities are all not zero after the lubricant oil is injected into
the bearing. The circumferential movement is derived from the
great agitation of the rolling elements and cage. The radial movement is mainly caused by the centrifugal force. The axial movement is mostly determined by the initial jet velocity of the
lubricant oil. The lubricant oil swirls around the inner ring and
ows to the outer ring by the centrifugal force. And then the lubricant oil ows out from the end face of the bearing. Due to the nonuniform airoil distribution, the convection heat transfer
coefcient inside the ball bearing changes with the circumferential
azimuth. The bearing temperature distribution is affected by the
nonuniform airoil distribution. The measured outer ring temperatures have represented the nonuniform characteristics, as shown
in Fig. 4(d) and (e). If even faster bearings are desired, the temperature difference of the outer ring should be considered. When the
bearing runs at higher speeds, the preload increases because of the
thermal expansion caused the higher temperature. Further, the
temperature difference makes the thermal expansion nonuniformity. Bearing life is an inverse function of the temperature

J. Hu et al. / International Journal of Heat and Mass Transfer 68 (2014) 8593

89

Fig. 4. Nonuniform airoil distribution inside the bearing. (a) Oil volume fraction distribution under the single-nozzle oil-jet lubrication. (b) Oil volume fraction distribution
under the dual-nozzle oil-jet lubrication. (c) Oil volume fraction distribution due to the centrifugal force. (d) Average oil volume fraction and measured outer ring
temperature under the single-nozzle oil-jet lubrication. (e) Average oil volume fraction and measured outer ring temperature under the dual-nozzle oil-jet lubrication.

difference [2]. To increase the calculation efciency of the proposed CFD model, the effect of the temperature difference on the
oil viscosity is neglected. The small temperature difference affects
the oil viscosity little at higher temperatures.
5.2. Effects of rotational speed
The variation of the airoil distribution with different speeds is
shown in Fig. 5. The oil-ow rate is 3.0 L/min and the oil viscosity is

0.02 Pa s in the calculation. It seems that the oil volume fraction


becomes lower when the speed increases. The airoil distribution
becomes more uniform in the dual-nozzle jet conguration at high
speeds. The airoil distribution inside the bearing tends to be uniform. However, the interaction between the airoil ow and the
bearing components becomes stronger with the higher inner ring
speed. It results in the increase of the pressure of the airoil phase
inside the bearing, as shown in Fig. 6(a). The ow velocity of the
airoil phase in the bearing also increases. The airoil phase leaves

90

J. Hu et al. / International Journal of Heat and Mass Transfer 68 (2014) 8593

the bearing faster. The oil volume fraction in the bearing remains at
a low level at a higher speed, which has an adverse effect on the
bearing heat diffusion. The bearing power loss increases with a
higher speed and more heat needs to be diffused in the bearing.
The measured power of the spindle motor at different speeds is
shown in Fig. 6(b). The oil-ow rate is 3.0 L/min and the oil-in temperature is 30 C.No external load is applied during the test. However, the average oil volume fraction inside the bearing declines.
This makes the cooling capability decrease. Reducing the speed effect on the ow velocity of the airoil phase is useful to improve
the oil volume fraction inside the bearing at a higher speed. It
can increase the dwell time of the airoil phase in the bearing
and ensure a sufcient heat exchange between the airoil phase
and the bearing components. For the oil-jet lubricated ball bearing,
a more efcient mechanism is required to make full use of the
lubricant oil.
5.3. Effects of lubricant oil

Fig. 5. Simulated airoil distribution inside the bearing at different speeds. (a)
Average oil volume fraction under the single-nozzle oil-jet lubrication. (b) Average
oil volume fraction under the dual-nozzle oil-jet lubrication.

Fig. 6. Bearing performances at different speeds. (a) Simulated airoil phase


pressure and velocity inside the bearing at different speeds. (b) Measured power of
the spindle motor at different speeds.

Fig. 7 shows the airoil distribution at different oil-ow rates.


The speed of the inner ring is 10,000 r/min and the oil viscosity
is 0.02 Pa s in the calculation. It seems that the oil volume fraction
in the bearing increases with the rise of the oil-ow rate. The nonuniform of the airoil distribution is also enhanced. The increase of
the oil-ow rate is conducive to the bearing heat diffusion. However, the churning loss also increases and the operation efciency
of the bearing reduces.
The effect of different oil-ow rates on the bearing performance
is shown in Fig. 8. It is seen that the shear force and the moment
applied by the airoil phase increase linearly with the oil-ow rate.
It results in the increase of the bearing churning loss the caused by
the interaction of the rolling components and the lubricant oil, as
shown in Fig. 8(b). The speed is 5000 r/min and the oil-in

Fig. 7. Simulated airoil distribution inside the bearing at different oil-ow rates.
(a) Average oil volume fraction under the single-nozzle oil-jet lubrication. (b)
Average oil volume fraction under the dual-nozzle oil-jet lubrication.

91

J. Hu et al. / International Journal of Heat and Mass Transfer 68 (2014) 8593

Fig. 8. Bearing performances at different oil-ow rates. (a) Simulated shear force
and moment applied on the bearing at different oil-ow rates. (b) Measured power
of the spindle motor at different oil-ow rates. (c) Measured equilibrium temperatures of the bearing at different oil-ow rates.

Fig. 9. Airoil distribution inside the bearing at different oil viscosities. (a) Average
oil volume fraction under the single-nozzle oil-jet lubrication. (b) Average oil
volume fraction under the dual-nozzle oil-jet lubrication. (c) Measured power of the
spindle motor at different oil-in temperatures.

Table 3
Measured temperatures of the bearing under an axial uniform load.

temperature is 50 C. No external load is applied during the test.


Fig. 8(c) shows the measured average outer ring temperature under different oil-ow rates after the bearing reaches thermal equilibrium. The test speed is 10,000 r/min. The axial load is 5.0 kN and
no radial load is applied. It seems that the bearing equilibrium
temperature decreases obviously with the increase of the oil-ow
rate at rst. Both the heat dissipation capacity and the power loss
increase with the rise of the oil volume fraction inside the bearing.
However, the former trends to be the main effect at this stage.
With the increase of the oil volume fraction, the temperature has
a slight increase then, as shown in Fig. 8(c). A glut of the lubricant
oil reduces the operation efciency of the bearing. However, it is
seen that the churning of the oil at higher ow rates does not account for the major power loss in the bearing at higher speeds in
the oil-jet lubrication. Both the oil churning and the oil cooling
are increased at higher ow rates. The numerical calculation of

Temperature sensor

No. 1

No. 2

No. 3

No. 4

Temperatures at 4000 r/min, C


Temperatures at 6000 r/min, C
Temperatures at 8000 r/min, C

63.7
69.3
76.2

63.4
69.0
75.9

63.9
70.2
77.5

64.3
70.7
78.0

the oil volume fraction inside the bearing proposes a method of


nding the equilibrium between the oil churning and the oil cooling in the oil-jet lubrication.
Fig. 9 shows the airoil distribution at different oil viscosities.
The oil-ow rate is 3.0 L/min with the single-nozzle method and
6.0 L/min with the dual-nozzle method. The speed of the inner ring
is 10,000 r/min in the calculation. In the test, the oil-ow rate is
3.0 L/min and no external load is applied. It seems that the oil volume fraction distributions at different oil viscosities have the same

92

J. Hu et al. / International Journal of Heat and Mass Transfer 68 (2014) 8593

Table 4
Measured temperatures of the bearing under different speeds.
4000 r/min
Temperature sensor
Temperatures of Single-nozzle (1#), C
Temperatures of Dual-nozzle (1#, 3#), C

No. 1
67.3
59.3

No. 2
68.6
58.3

6000 r/min
No. 4
70.7
59.8

No. 5
75.4
63.8

No. 1
71.0
69.0

8000 r/min

No. 2
71.4
68.3

No. 4
74.0
70.3

No. 5
82.5
78.7

No. 2
89.0
85.2

No. 4
93.7
88.0

No. 5
105.6
99.1

No. 1
78.5
68.2

No. 2
79.0
67.9

No. 4
82.8
69.0

No. 5
93.0
83.1

No. 2
91.2
87.3

No. 4
95.9
89.6

No. 5
106.8
101.7

Table 5
Measured temperatures of the bearing under different lubricant oil ow rates.
3.0 L/min
Temperature sensor
Temperatures of Single-nozzle (1#), C
Temperatures of Dual-nozzle (1#, 3#), C

No. 1
87.2
84.0

4.0 L/min
No. 2
88.1
83.7

No. 4
92.0
86.7

No. 5
105.4
100.6

trends along the circumferential direction. Because the owability


of the oil decreases, the oil volume fraction along the circumferential direction changes less with a larger viscosity. The nonuniformity of the airoil distribution is also weakened.
5.4. Nonuniform temperature distribution
For the oil-jet lubricated ball bearing, the airoil distribution affects the heat transfer ability of the bearing. The effects on the heat
transfer ability of the bearing are directly reected in the temperature distribution of the bearing. Since the measurement of the
airoil distribution inside the bearing is difcult, the bearing temperature distributions under different operation conditions are
measured by the experimental apparatus.
Tables 3 and 4 shows the temperature distribution under different speeds. Only the axial load was applied and the dual-nozzle jet
with the Nos. 1 and 3 nozzles were used during the test in Table 3.
The value of the axial load is 5.0 kN and the lubricant oil-ow rate
is 3.0 L/min. It seems that the nonuniform temperature distribution of the bearing appears under the axial uniform load. The lower
temperature of the outer ring is at the position close to the nozzle.
The higher temperature of the outer ring appears at the lowest oil
volume fraction position, which is the upstream side near the nozzle. Both the axial load of 5.0 kN and the radial load of 10.0 kN are
applied during the tests in Table 4. It is seen that the same temperature distribution of the bearing as the results in Table 3 is measured. The inner ring temperature is the highest due to the
centrifugal effect of the airoil phase. It can be readily deduced that
the nonuniform airoil distribution should be considered for the
precise lubrication analysis of the oil-jet lubricated ball bearing.
Table 5 shows the temperature distribution under different lubricant oil-ow rates. Both the axial load and the radial load are
applied during the tests. The value of the axial load is 5.0 kN and
the radial load is 10.0 kN. The tested speed is 10,000 r/min. It
seems that the nonuniform temperature distribution of the bearing
becomes more obvious. The highest temperature of the outer ring
is measured by the No. 4 temperature sensor. It is because the radial load is applied on the upstream side of the No.4 temperature
sensor. Further, the oil volume fraction in the position of the No.
4 temperature sensor is lower. It is important to optimize the nozzle position according to the load distribution and the airoil distribution for the oil-jet lubricated ball bearing. The nozzle should
be placed to guarantee the uniform injection of the lubricant oil
at higher speeds. The regulation of the lubricant oil through each
nozzle according to the loaded area is also important to increase
the high-speed capability of the rolling bearing. The higher loaded
area requires a little more lubricant oil to remove the generated
heat.

No. 1
88.7
85.9

5.0 L/min
No. 1
91.1
88.5

6. Conclusions and future work


The airoil two-phase ow inside the oil-jet lubricated ball
bearing was investigated by the CFD method. The VOF method
for multiphase ow was used to track the airoil two-phase ow
inside the oil-jet lubricated ball bearing. The sliding mesh plane
was used, considering that the ball bearing is a rotary component.
The results suggest the following.
(1) The airoil distribution inside the oil-jet lubricated ball
bearing is not uniform. The lowest oil volume fraction
appears in the upstream side near the nozzle. The oil volume
fraction close to the inner ring is extremely small.
(2) The ow velocity and the pressure of the airoil phase
increase when the inner ring speed increases. The average
oil volume fraction becomes lower inside the bearing.
(3) The average oil volume fraction in the bearing increases with
the higher oil ow rate. The shear force and the moment
applied by the airoil phase also increase linearly with the
oil ow rate. The nonuniformity of the airoil distribution
is also enhanced.
(4) The nonuniform airoil distribution should be considered for
the heat transfer analysis of the oil-jet lubricated ball bearing. It is important to optimize the nozzle position according
to the load and airoil distribution for the oil-jet lubricated
ball bearing.
The results can be used for the precise lubrication design for
main shaft ball bearings of the high-speed vehicular transmission.
The convection heat transfer characteristics of the airoil
two-phase ow needs more effort to study. Work on these topics
is currently underway in the National Key Laboratory of Vehicular
Transmission at the Beijing Institute of Technology.
Acknowledgments
This work is supported by the National Basic Research Program
of China (Grant no. 2011CB706600). The authors also wish to thank
referees for improving the quality of the paper.
References
[1] S. Jiang, H. Mao, Investigation of the high speed rolling bearing temperature
rise with oilair lubrication, J. Tribol. Trans. ASME 133 (2) (2011) 19.
[2] S.I. Pinel, H.R. Signer, E.V. Zaretsky, Comparison between oil-mist and oil-jet
lubrication of high-speed, small-bore, angular-contact ball bearings, Tribol.
Trans. 44 (3) (2001) 327338.
[3] J. Aidarinis, D. Missirlis, K. Yakinthos, A. Goulas, CFD modeling and LDA
measurements for the air-ow in an aero engine front bearing chamber, J. Eng.
Gas Turbines Power-Trans. ASME 133 (8) (2011) 18.

J. Hu et al. / International Journal of Heat and Mass Transfer 68 (2014) 8593


[4] T.A. Harris, R.M. Barnsby, M.N. Kotzalas, A method to calculate frictional effects
in oil-lubricated ball bearings, Tribol. Trans. 44 (4) (2001) 704708.
[5] H. Zhou, G. Luo, G. Chen, F. Wang, Analysis of the nonlinear dynamic response
of a rotor supported on ball bearings with oating-ring squeeze lm dampers,
Mech. Mach. Theory 59 (2013) 6577.
[6] P. Gloeckner, F. Ebert, Micro-sliding in high-speed aircraft engine ball bearings,
Tribol. Trans. 53 (3) (2010) 369375.
[7] D. Dong, T. Komoriya, T. Endo, Y. Kimura, Formation of EHL lm with
grease in ball bearings at low speeds, J. Jpn. Soc. Tribol. 57 (8) (2012)
568574.
[8] N. Biboulet, L. Houpert, Hydrodynamic force and moment in pure rolling
lubricated contacts. Part 2: Point contacts, Proc. Inst. Mech. Eng. Part J J. Eng.
Tribol. 224 (8) (2010) 777787.
[9] C. Brecher, S. Baeumler, J. Rossaint, Calculation of kinematics and friction of a
spindle bearing using a local EHL friction model, Tribol. Trans. 56 (2) (2013)
245254.
[10] M.K.W. Ibrahim, D. Gasni, R.S. Dwyer-Joyce, Proling a ball bearing oil lm
with ultrasonic reection, Tribol. Trans. 55 (4) (2012) 409421.
[11] T.A. Harris, M.N. Kotzalas, Rolling Bearing Analysis, fth ed., Taylor & Francis,
New York, 2006.
[12] J. Takabi, M.M. Khonsari, Experimental testing and thermal analysis of ball
bearings, Tribol. Int. 60 (2013) 93103.
[13] E.V. Zaretsky, H. Signer, E.N. Bamberger, Operating limitations of high-speed
jet-lubricated ball bearings, J. Lubr. Technol. Trans. ASME 98 (1) (1976) 32
39.

93

[14] C. Hirt, B. Nichols, Volume of uid (VOF) method for the dynamics of free
boundaries, J. Comput. Phys. 39 (1) (1981) 201225.
[15] A. Mehdizadeh, S.A. Sherif, W.E. Lear, Numerical simulation of thermouid
characteristics of two-phase slug ow in microchannels, Int. J. Heat Mass
Transfer 54 (1516) (2011) 34573465.
[16] M. Ramdin, R. Henkes, Computational uid dynamics modeling of Benjamin
and Taylor bubbles in two-phase ow in pipes, J. Fluids Eng. Trans. ASME 134
(4) (2012) 18.
[17] R. Banerjee, K.M. Isaac, L. Oliver, W. Breig, Features of automotive gas tank
ller pipe two-phase ow: experiments and computational uid dynamics
simulations, J. Eng. Gas Turbines Power-Trans. ASME 124 (2) (2002) 412420.
[18] E. Da Riva, D. Del Col, Numerical simulation of laminar liquid lm
condensation in a horizontal circular minichannel, J. Heat Transfer-Trans.
ASME 134 (5) (2012) 18.
[19] J.L. Xiao, E.Q. Zhu, G.D. Wang, Numerical simulation of emergency shutdown
process of ring gate in hydraulic turbine runaway, J. Fluids Eng. Trans. ASME
134 (12) (2012) 19.
[20] D.A. Bompos, P.G. Nikolakopoulos, CFD simulation of magnetorheological uid
journal bearings, Simul. Model. Pract. Theory 19 (4) (2011) 10351060.
[21] J. Gylys, L. Paukstaitis, R. Skvorcinskiene, Numerical investigation of the drag
force reduction induced by the two-phase ow generating on the solid body
surface, Int. J. Heat Mass Transfer 55 (2526) (2012) 76457650.
[22] M. Mohammadi, S. Shahhosseini, M. Bayat, Direct numerical simulation of
water droplet coalescence in the oil, Int. J. Heat Fluid Flow 36 (2012) 5871.

Anda mungkin juga menyukai