You are on page 1of 16

This article appeared in a journal published by Elsevier.

The attached
copy is furnished to the author for internal non-commercial research
and education use, including for instruction at the authors institution
and sharing with colleagues.
Other uses, including reproduction and distribution, or selling or
licensing copies, or posting to personal, institutional or third party
websites are prohibited.
In most cases authors are permitted to post their version of the
article (e.g. in Word or Tex form) to their personal website or
institutional repository. Authors requiring further information
regarding Elsevier’s archiving and manuscript policies are
encouraged to visit:
http://www.elsevier.com/copyright
Author's personal copy
ARTICLE IN PRESS

Tribology International 41 (2008) 1190– 1204

Contents lists available at ScienceDirect

Tribology International
journal homepage: www.elsevier.com/locate/triboint

CFD analysis of journal bearing hydrodynamic lubrication by


Bingham lubricant
K.P. Gertzos, P.G. Nikolakopoulos, C.A. Papadopoulos 
Department of Mechanical Engineering and Aeronautics, University of Patras, GR-26504 Patras, Greece

a r t i c l e in f o a b s t r a c t

Article history: Design of smart journal-bearing systems is an important issue that opens up the possibility for semi-
Received 6 June 2007 active dynamic control of bearing behavior. Recent studies show that there is an increasing interest in
Received in revised form designing hydrodynamically lubricated bearings using electro-rheological fluids (ERFs) or magneto-
27 January 2008 rheological fluids (MRFs). Both smart fluids behave like Bingham fluids, and thus the Bingham plastic
Accepted 10 March 2008
model is used to describe the grease and the electro-rheological (ER) and magneto-rheological (MR)
Available online 9 May 2008
fluids behavior of the non-Newtonian fluid flow. The performance characteristics of a hydrodynamic
Keywords: journal bearing lubricated with a Bingham fluid are derived by means of three-dimensional
Journal bearing
computational fluid dynamics (3-D CFD) analysis. The FLUENT software package is used to calculate
Bingham fluid
the hydrodynamic balance of the journal using the so-called ‘‘dynamic mesh’’ technique. The results
Yield stress
Computational fluid dynamics (CFD)
obtained from the developed 3-D CFD model are found to be in very good agreement with experimental
FLUENT and analytical data from previous investigations on Bingham fluids.
Journal-bearing performance characteristics, such as relative eccentricity, attitude angle, pressure
distribution, friction coefficient, lubricant flow rate, and the angle of maximum pressure, are derived
and presented for several length over diameter (L/D) bearing ratios and dimensionless shear numbers T0
of the Bingham fluid. The above diagrams are presented in the form of Raimondi and Boyd charts, and
can easily be used in the design and analysis of journal bearings lubricated with Bingham fluids. The
core profile formed in the bearing is also calculated and presented for various bearing eccentricities, L/D
ratios, and shear numbers T0, and found to be in very good agreement with previous experimental and
theoretical investigations. The analysis presented here leads to charts that could be used by the designer
engineer to design smart journal bearings.
& 2008 Elsevier Ltd. All rights reserved.

1. Introduction thickness. The extent of this core formation depends only on the
geometrical condition at the bearing and the dimensionless yield
The behavior of many grease lubricants, as well as electro- stress T0. Batra [4] studied only the case of attached cores in a
rheological (ER) and magneto-rheological (MR) fluids [1] pro- journal bearing, but showed that both floating and attached cores
posed as ‘‘smart’’ lubricants, is well described by the Bingham may occur. He found that the load capacity and the moment of
model of non-Newtonian fluid flow. A major difference from the friction of the bearing with a Bingham material are larger than
Newtonian fluid flow is that the Bingham model is characterized that with a Newtonian material.
by two parameters: (a) yield stress and (b) viscosity. When the Wada et al. [5] developed the general theory of a Bingham solid
stress on the lubricant is less than the yield stress, the material is in hydrodynamic lubrication, establishing the core formation and
rigid and a region called the ‘‘core’’ is formed; exceeding the yield also applied their theory to a step bearing [6]. Their theory finds
stress leads to a quasi-Newtonian flow. both a floating rigid core and cores that adhere to surfaces.
Studies of Bingham fluid lubrication date back to the experi- Hayashi and Wada [7] developed a more general modified
mental work of Cohen et al. [2], on greases which act like Bingham Reynolds equation that includes the effects of correlation of shear
materials. Milne [3] examined a journal-bearing model both stresses with the velocity gradients in the circumferential and
experimentally and theoretically and concluded that cores are axial directions. Finally, they applied their theory to a journal
formed near the bearing at the region of maximum film thickness bearing and obtained the theoretical core profile formed in the
and near the moving shaft at the region of the minimum film bearing as well as the bearing performance [8]. They also
observed the core formation on a bearing made of a transparent
material and measured both the pressure distribution and the
 Corresponding author. Tel.: +30 261 96 9426; fax: +30 261 99 6258. journal displacements. The experimental results of these works
E-mail address: chris.papadopoulos@upatras.gr (C.A. Papadopoulos). are used in this paper to validate the obtained numerical results

0301-679X/$ - see front matter & 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.triboint.2008.03.002
Author's personal copy
ARTICLE IN PRESS

K.P. Gertzos et al. / Tribology International 41 (2008) 1190–1204 1191

Nomenclature t* maximum core to film thickness ratio, t* ¼ max{t(y)/


h(y)}
a ratio of the bearing surface occupied by core, ~
u fluid velocity vector
a ¼ (Ab+Af)/(2pRbL) (Fig. 8b) ut tangential component of the lubricant velocity (m s1)
Ab area of the core adhered to bearing (Fig. 8b) (m2) W external force (N)
Af area of the floating core (Fig. 8b) (m2) w* dimensionless load carrying capacity w ¼
2 3
Aj area of the core adhered to journal (Fig. 8b) (m2) WC =ðm0 oRj LÞ
Ah cross sectional area of the lubricant film Ah ¼ Lh (m2) U journal velocity, parallel to the film
C clearance C ¼ Rb  Rj
D journal diameter (m) Subscripts
e eccentricity (m)
f friction coefficient 0 indicates the position of maximum film thickness at

f* normalized friction coefficient f ¼ f ðR=CÞ y¼0
Ffr frictional force (N) 1 indicates the position of minimum film thickness at
F fr normalized frictional force F fr ¼ F fr C=ðm0 oR2j LÞ y¼p
Fp lubricant pressure force (load carrying capacity) (N) b indicates the bearing
~
g acceleration of gravity vector j indicates the journal
h film thickness (m) f indicates floating core
Ī¯ unit tensor
k consistency factor (kg sn2 m1) Superscripts
L bearing length (m)
n dimensionless power-law index
* indicates dimensionless quantity
N rotational speed of journal (rps)
p pressure (Pa)
Greek symbols
pa ambient pressure (Pa)
p* dimensionless pressure p ¼ ðp  pa Þ  C 2 =ðm0  o  R2j Þ
pmax maximum pressure on the journal surface (Pa) g_ shear rate (s1)
pm mean pressure on the journal surface (Pa), pm ¼ e relative eccentricity  ¼ e=C
W=ðLDj Þ y bearing angle (1)
Q lubricant flow rate (m3 s1) ypmax angle of maximum pressure (1)
q* dimensionless lubricant flow rate, q ¼ Q =ðRj CNLÞ m Newtonian dynamic viscosity (Pa s)
Qs lubricant side leakage flow rate (m3 s1) ma non-Newtonian viscosity (local ‘‘apparent’’ viscosity)
qs side leakage to maximum lubricant flow rate ratio, m0 yielding viscosity (Pa s)
qs ¼ Q s =Q 0 r lubricant density (kg m3)
Rj, Rb journal and bearing radii (m) t̄¯ stress tensor
Re Reynolds number Re ¼ r  o  Rj  C=m0 t0 yield stress (Pa)
S0 Sommerfeld number S0 ¼ m0  o  Rj  LðRj =CÞ2 =ðp  WÞ j attitude angle (1)
T0 dimensionless yield stress T 0 ¼ t0  C=ðm0  o  Rj Þ o rotational speed of journal (rad s1)
t core thickness (distance from the bearing surface)
(Fig. 8a) (m2)

for a Bingham material. Experimental results from various types behave like non-Newtonian fluids which are different from
of greases are presented more recently by Mutuli et al. [9]. Bingham fluids. The relationship between shear stress and shear
Tichy [10] obtained different modified Reynolds equations strain rate seems to follow the cubic shear stress law [14–17], the
which are dependent on the possible local formation of a rigid power law [18,19], or the Eyring model [20].
core, which may be either attached to the surface or floating Recently, following the progress in computer technology, many
between the surfaces. Results are presented for a squeeze film researchers began to use commercial computational fluid dy-
damper and journal bearing. He also mentions that the analysis namics (CFD) programs in their investigations. The main advan-
may be useful to those studying lubrication issues and trying to tage of CFD code is that it uses the full Navier–Stokes equations
predict the behavior of electro-rheological fluids (ERFs). Kim et al. and provides a solution to the flow problem, whereas finite
[11] developed a mathematical model to investigate the influence difference codes are based on the Reynolds equation. The results
of the lubricant rheological properties on the conditions that obtained by the two approaches are therefore likely to differ.
occur in slider and journal bearings. Peng [12] studied the Moreover, the CFD packages are applicable in very complex
hydrodynamic characteristics of a journal bearing with an ER geometries. Chen et al. [21] studied the influence of end seal
fluid and also used the modified Bingham plastic model to clearance and flow path length on the performance of a circular
describe the behavior of the ER fluid. Lu et al. [13] presented a orbiting squeeze film damper with a central circumferential feed
series of experimental results that explored the frictional groove, using the CFD package CFX4.2. Ranjan et al. [22] presented
characteristics of a grease-lubricated journal bearing. Load, grease a CFD approach, using FLUENT, to model fluid flow in a journal
type, and bushing material are varied to examine their effects on bearing with three equally spaced axial grooves which was
the friction coefficient. supplied with water from one end. They also calculated the
All of the above investigations are applied to lubricants which stiffness and damping coefficients.
behave like a Bingham fluid. However, many commercial In the present work, a three dimensional CFD model is
lubricants, due to the presence of different types of additives, developed, using the FLUENT 6.3 package, to study the behavior
Author's personal copy
ARTICLE IN PRESS

1192 K.P. Gertzos et al. / Tribology International 41 (2008) 1190–1204

and the performance characteristics of a journal bearing lubri- friction coefficient, the flow rate, the maximum pressure, and the
cated with a Bingham fluid and a Newtonian lubricant. The results angle of the maximum pressure versus Sommerfeld number for
of the CFD model with the Newtonian fluid are compared with the several length over diameter (L/D) bearing ratios and dimension-
theoretical ones of Raimondi–Boyd [23] and Wada et al. [8] and less shear numbers T0 of the Bingham solid. All of the diagrams
found to be in good agreement. The core formation, adhering to above are presented in the Raimondi–Boyd [23] form. The results
the bearing or journal surface or the floating core, and the of this work could be used to design journal bearings lubricated
performance characteristics predicted by the CFD model, are with ER or MR fluids.
found to be in very good agreement with the experimental work
of Wada et al. [8] and Milne [3]. The validated model is used to
extract diagrams of the relative eccentricity, the attitude locus, the 2. Analysis

The coordinate system and the geometry of a journal bearing


are shown in Fig. 1. The journal rotates with an angular velocity o
and is in an equilibrium position under the external vertical load
W as well as the pressure of the lubricant film. The journal axis Oj
is at distance e from the bearing axis Ob. The film thickness h(y)
varies from its maximum value h0 at bearing angle y ¼ 0 to its
minimum value h1 at y ¼ p.

2.1. Assumptions

The following assumptions for the bearing model are used in


this work:
A rigid aligned bearing with the geometry of Fig. 1 is
considered; a steady-state operation is assumed; the flow
is laminar and isothermal; a constant external vertical load W is
applied to the journal; the relation between the shear stress and
the shear rate of the lubricant is expressed by the Bingham model,
as shown in Fig. 2.

2.2. Equations to be solved

For all flows, FLUENT solves conservation equations for mass


and momentum. For flows involving heat transfer or compressi-
bility, an additional equation for energy conservation is solved. For
flows involving species mixing or reactions, species conservation
equations are solved or, if the non-premixed combustion model is
used, conservation equations for the mixture fraction and its
variance are solved. Additional transport equations are also solved
when the flow is turbulent. In this paper, the conservation
equations for laminar flow (in an inertial, non-accelerating,
Fig. 1. Schematic journal-bearing geometry. reference frame) are presented.

Fig. 2. Variation of shear stress with shear rate, according to various Newtonian or non-Newtonian models.
Author's personal copy
ARTICLE IN PRESS

K.P. Gertzos et al. / Tribology International 41 (2008) 1190–1204 1193

2.2.1. Mass conservation equation cavitations. The FLUENT can handle cavitation flows, but cannot
The equation for conservation of mass, or the continuity use Reynolds BC. This happens because the Navier Stokes
equation, can be written as equations are solved instead of Reynolds equation. It is very
qr
uÞ ¼ 0.
þ r  ðr~ (1)
qt
Eq. (1) is the general form of the mass conservation equation
and is valid for incompressible as well as compressible flows.

2.2.2. Momentum conservation equations


Conservation of momentum in an inertial (non-accelerating)
reference frame is described by
q
uÞ þ r  ðr~
ðr~ u~ g þ~
uÞ ¼ rp þ r  ðtÞ þ r~ F, (2)
qt
where r~ g and ~ F are the gravitational body force and external body
forces, respectively. The stress tensor t is given by

t ¼ ma ½ðr~ uÞT Þ  23r  ~


u þ ðr~ uI, (3)
where the second term on the right hand side is the effect of
volume dilation.
All investigations that are referenced in this work solve the
fundamental Reynolds Eq. (4), or a modified form of (4) that takes
into account the equations describing the non-Newtonian lubri- Fig. 4. Relative eccentricity vs. Sommerfeld number for Newtonian lubricant,
continuous line: present CFD model results, dashed line: Raimondi and Boyd [23].
cant,
!
3
h
r rp ¼ r  ðhUÞ  V, (4)
12m

where U is the velocity of journal surface parallel to the film and V


the squeeze-film velocity.
When the film pressure drops down to the atmospheric
pressure due to heavy external load or high operating rotational
speed, rupture of the film or cavitation occurs. Taking into account
this situation, the following boundary condition is used which
applies the non-sub-atmospheric pressure constraint:
p  pa X0 at ppyp2p; z ¼ 0; z ¼ L. (5)
The Reynolds boundary condition, that makes the pressure curve
to drop in parallel with y-axis just after 1801 could be used instead
of Half Sommerfeld BC. The Reynolds boundary condition gives in
some cases more accurate results than the half Sommerfeld BC.
Although it is relatively accurate, the Reynolds BC is still an Fig. 5. Pressure distribution for Newtonian lubricant, continuous line: present CFD
approximation to the transition from full fluid flow to flow with model results, dashed line [8].

4.50e+05 4.50e+05
4.15e+05
3.80e+05 4.15e+05
3.45e+05
3.10e+05 3.80e+05
2.75e+05
2.40e+05 3.45e+05
2.05e+05
1.70e+05 3.10e+05
1.35e+05
1.00e+05 2.75e+05
6.50e+04
3.00e+04 2.40e+05
-5.00e+03
-4.00e+04 2.05e+05
-7.50e+04
-1.10e+05 1.70e+05
-1.45e+05
-1.80e+05 Y 1.35e+05 Y
-2.15e+05 X X
-2.50e+05 Z 1.00e+05 Z
p (Pa)

Fig. 3. Pressure contour in Pa on the journal surface: (a) without the half-Sommerfeld boundary condition effect and (b) after applying the half-Sommerfeld boundary
condition.
Author's personal copy
ARTICLE IN PRESS

1194 K.P. Gertzos et al. / Tribology International 41 (2008) 1190–1204

difficult to modify the Navier Stokes equations so as to include the The friction coefficient is calculated by the relation:
Reynolds BC. More work has to be done in a following
F fri
investigation where the cavitation should be taken into con- fi ¼ . (14)
W
sideration. When cavitation is included, the vaporization pressure,
the surface tension coefficient and the non-condensable gas mass The load-carrying capacity that the bearing will support is
fraction are also needed. found by integrating the pressure around the journal. The
boundary condition defined by Eq. (5) is included in the
2.3. Viscosity for non-Newtonian fluids integration to prevent sub-atmospheric pressure. At equilibrium
position, the load capacity of the journal equals the external
vertical load.
For incompressible Newtonian fluids, the shear stress is
ZZ
proportional to the rate-of-deformation tensor D:
F py ¼ p dAj ¼ W; . . . ; F px ¼ 0. (15)
t ¼ mD, (6) Aj

where D is defined by The lubricant flow rates, Q0 at the maximum position and Q1 at
  the minimum position of the film thickness, are calculated by
quj qui
D¼ þ (7) integrating the tangential velocity of the lubricant:
qxi qxj
ZZ ZZ
and m is the viscosity, which is independent of D. For some non- Q0 ¼ ut0 dAh0 ; . . . ; Q 1 ¼ ut1 dAh1 . (16)
Newtonian fluids, the shear stress can similarly be written in Ah0 Ah1
terms of a non-Newtonian viscosity ma:
The rate at which the lubricant is lost due to side leakage is
t ¼ ma ðDÞD. (8)
Q s ¼ Q 0  Q 1. (17)
In general, ma is a function of all three invariants of the rate-of-
deformation tensor D. However, in the non-Newtonian models
available in FLUENT, ma is considered to be only a function of the
shear rate, g_ , that is related to the second invariant of D and is
defined as
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1
g_ ¼ D : D, (9)
2
where the operand: denotes the double dot (or inner–outer)
product.

2.3.1. A FLUENT Herschel–Bulkley model for Bingham plastics


Bingham plastics are characterized by a non-zero shear stress
when the strain rate is zero:
ma g_ ¼ t0 þ m_g. (10)
For tot0, the material remains rigid.
For t4t0, the material flows as a power-law fluid.
The standard Herschel–Bulkley model is described by
Fig. 6. Pressure distribution for Bingham lubricant with T0 ¼ 0.4, L/D ¼ 1,
t  ma g_ ¼ t0 þ k_gn . (11) continuous line: present CFD model results, dashed line: theoretical [8], points:
experimental [8].
In FLUENT [24], the extended Herschel–Bulkley model com-
bines the effects of Bingham and power-law behavior in a fluid.
For low strain rates ð_got0 =m0 Þ, the ‘‘rigid’’ material acts like a very
viscous fluid with viscosity m0. As the strain rate increases and the
yield stress threshold t0 is exceeded, the fluid behavior is
described by a power law.
t  ma g_ ¼ t0 þ k½_gn  ðt0 =m0 Þn . (12)
The FLUENT Herschel–Bulkley model becomes identical to the
Bingham model when k ¼ m, n ¼ 1, and ðt0 =m0 Þ ! 0. This is
possible when a large value is given to m0. The aforementioned
shear stress models are presented in Fig. 2.

2.4. Performance characteristics

The friction force is calculated in the equilibrium position by


integrating the shear stress t over the bearing (or journal) area:
ZZ
F fri ¼ ti dAi ; i ¼ j or b, (13)
Ai
Fig. 7. Pressure distribution for Bingham lubricant with T0 ¼ 0.8, L/D ¼ 1,
where i ¼ j refers to the journal and i ¼ b refers to the bearing and continuous line: present CFD model results, dashed line: theoretical [8], points:
Ai is the total area of journal or bearing. experimental [19].
Author's personal copy
ARTICLE IN PRESS

K.P. Gertzos et al. / Tribology International 41 (2008) 1190–1204 1195

According to the coordinate system of Fig. 1 the attitude angle is The Navier–Stokes equations are solved in steady state, taking
computed by into account gravity forces. The operating pressure is set to
101 325 Pa. To simplify the geometry, one side of the clearance is
j ¼ 90  tan1 ðex =ðey ÞÞ. (18) used as a lubricant inlet and the other as an outlet. The boundary
The angle of the maximum pressure is computed by conditions are: ‘‘pressure outlet’’ with gauge pressure at zero
Pascal and ‘‘pressure inlet’’ with an appropriate value leading to
ypmax ¼ 90  tan1 ðxpmax =ðypmax ÞÞ. (19) the right-side lubricant flow rate, according to Eq. (17).
The bearing shell is modeled as a ‘‘stationary wall’’. The journal is
modeled as a ‘‘moving wall’’ with an absolute rotational speed of
3. CFD model—analysis (simulation model) 120 rpm. The rotational axis origin is set to the value of eccentricity.
When the problem is solved for a constant eccentricity, this value is
A 3-D simulation model is developed using the CFD package known. When the problem is solved for a constant external force,
FLUENT 6.3. The pre-processor Gambit 2.3 is used for the grid the final position of the origin is computed automatically by the
generation. The individuality of the problem examined here (the ‘‘dynamic mesh’’ technique. The ‘‘dynamic mesh’’ model in FLUENT
clearance size is very small compared to journal diameter and is used to model flows where the shape of the fluid domain is
length) enforces the use of only hexahedral cells because the use changing due to motion on the domain boundaries.
of tetrahedral cells leads to an enormous number of cells. Eight Three compiled user defined functions (UDF) in C are used. The
divisions were used across the journal-bearing film, 360 divisions first UDF applies the half Sommerfeld boundary condition to the
were used in the circumferential direction, and 25 divisions were upper side of the journal, setting all values of negative gauge
used in the axial direction. The number of cells is 72000. One pressures to the value of the operating pressure. Thus, the gauge
hundred and twenty divisions were used in the circumferential pressures for the top of journal (between y ¼ p and 2p) are set to
direction initially, giving unacceptable results for the angle of the zero when negative values occur. The FLUENT DEFINE_ON_
maximum pressure. DEMAND macro is used for the definition of the first UDF, while
The dimensions of the journal bearing used in this simulation the thread_loop_c, begin_c_loop macros are used to loop over all
are: diameter 0.05 m and clearance 235 mm. The C/D value is cells, and the thread_loop_f, begin_f_loop macros to loop over all
4.7  103, denoting a clearance in the upper limit of acceptable faces of the fluid domain. The second UDF is used to compute the
values. The L/D ratios used are 1/4, 1/2, 1, 2, and 4 instead of N. All equilibrium position of the journal when a known external force is
of the results are repeated with a value C ¼ 50 mm for the applied to it. The UDF integrates all pressure forces at the journal,
clearance. The C/D value is 1 103, denoting a clearance in taking into account the half Sommerfeld boundary condition. The
the lower limit of acceptable values. The results of the two DEFINE_CG_MOTION macro is used for the definition of the second
above cases are almost identical, denoting that the results are UDF. The pressure on the journal surface cells is computed by the
independent of the value of clearance. The viscosity is 0.2 Pa s F_P macro and the area and direction of the cell surface by the
while the density is 960 kg m3. As the Re number is 0.35 (for F_AREA_CACHE macro.
rotational speed 120 rpm), the viscous model is set to laminar. The When the problem is solved for a known relative eccentricity, a
range of rotational speed used is 19–240 rps. For the Bingham model is generated with the pre-processor GAMBIT. The mesh is
solid (non-Newtonian) simulation, the Herschel–Bulkley viscosity transferred to FLUENT, where the boundary conditions, model
model is used with values for the consistency index k ¼ 0.2 Pa s, properties, etc. are set. Many positions of eccentricity are
Power-law index n ¼ 1, yield viscosity 500 Pa s, and yield stress necessary to complete the diagrams presented here. To save time,
213.8 Pa. These values have the same effect as the Bingham a third UDF is defined, taking the desirable eccentricity as input
model with yield stress 213.8 Pa and viscosity 0.2 Pa s, as shown in and moving the journal to this position using the ‘‘dynamic mesh’’
Fig. 2. technique.

Fig. 8. (a) Schematic representation of the core at a cross section of the journal bearing and (b) calculated area of floating and adhered, to bearing and to journal, surface
core.
Author's personal copy
ARTICLE IN PRESS

1196 K.P. Gertzos et al. / Tribology International 41 (2008) 1190–1204

Fig. 9. Core formation, dimensionless core thickness and ratio of core over bearing surface for: (a) L/D ¼ 0.5, (b) L/D ¼ 1 and (c) L/D ¼ 2. Dark parts: core adhering to the
bearing or to the journal surface. Hatched parts: core floating in the lubricant film.

The smoothing mesh method is used with a convergence value and the angle of the maximum pressure, and the attitude
tolerance of 106 and number of iterations equal to 200. As angle.
this method is not enabled by default for hexahedral cells, this The segregated solver is chosen for the present numerical analysis.
is done via the text interface inside FLUENT. The dynamic The velocity–pressure coupling is treated using the SIMPLE algorithm
mesh zones defined are: (a) bearing as ‘‘stationary’’, (b) the and the second-order upwind scheme is used for momentum. For
two sides and the lubricant volume set to ‘‘deforming’’ with greater accuracy, a value of 106 is used for all residual terms.
‘‘spring smoothing’’ method, and (c) the journal as ‘‘rigid
body’’. The third of the previous UDFs is applied to the journal
as a motion UDF/profile for a known eccentricity, while the 4. Results and discussion
second UDF is applied for a known external force. The center of
gravity is also computed by the UDF. Additional code is written to The majority of the extracted results in this work, for both the
display the components of the pressure force on the journal, the Newtonian and the Bingham model hydrodynamic lubrication, are
Author's personal copy
ARTICLE IN PRESS

K.P. Gertzos et al. / Tribology International 41 (2008) 1190–1204 1197

Fig. 10. Relative eccentricity vs. Sommerfeld number for: (a) T0 ¼ 0, (b) T0 ¼ 0.4, (c) T0 ¼ 0.8 and (d) T0 ¼ 5.

presented in diagrams similar to the diagrams of Raimondi and


Boyd [24]. The following dimensionless quantities are used:

Re ¼ r  o  Rj  C=m,
S0 ¼ m  o  R  LðR=CÞ2 =ðp  WÞ,
T 0 ¼ t0  C=ðm  o  Rj Þ,
p ¼ ðp  pa Þ  C 2 =ðm  o  R2j Þ,

f ¼ f ðR=CÞ; q ¼ Q =ðRj CNLÞ,
qs ¼ Q s =Q 0 ; w ¼ WC 2 =ðmoR3 LÞ,
F fr ¼ F fr C=ðmoR2j LÞ; a ¼ ðAb þ Af Þ=ð2pRb LÞ,
t  ¼ maxftðyÞ=hðyÞg.

Fig. 3a shows the pressure distribution on the journal surface


when the half Sommerfeld boundary condition defined by Eq. (5)
is omitted. It is clear that sub-atmospheric and negative pressures
are present. Fig. 3b depicts the pressure distribution when the
boundary condition of Eq. (5) is applied to the solution. Fig. 11. Dimensionless load carrying capacity vs. relative eccentricity for L/D ¼ 1.

4.1. Validation of the CFD model with a Newtonian lubricant 19a, 20a are compared with the ones of the investigation of
Raimondi and Boyd [23] for L=D ¼ 14; 12; 1; 1, while the results of
Results are extracted for a Newtonian lubricant with L=D ¼ Fig. 5 are compared to the ones of Wada et al. [8].
1 1
4; 2;1; 4 and presented in diagrams in Figs. 4, 12a, 14a, 15a, 17a, Fig. 4 shows comparative results of the relative eccentricity
18a, 19a, 20a and 5. The results of Figs. 4, 12a, 14a, 15a, 17a, 18a, while Fig. 12a for the attitude angle vs. Sommerfeld number. There
Author's personal copy
ARTICLE IN PRESS

1198 K.P. Gertzos et al. / Tribology International 41 (2008) 1190–1204

Fig. 12. Attitude angle vs. Sommerfeld number for: (a) T0 ¼ 0, (b) T0 ¼ 0.4, (c) T0 ¼ 0.8 and (d) T0 ¼ 5.

differences appear when compared to the results from this paper


computed on the journal surface. In this investigation [23], a
linear lubricant velocity distribution is assumed between the
journal and the bearing, leading to the same velocity slope and
consequently to the same friction force at two surfaces. Results
extracted in this paper show that the velocity distribution is
nonlinear and the friction coefficients for the journal and bearing
differ significantly.
Fig. 17a shows the maximum dimensionless flow rate. The
agreement between the results of the present analysis and
Raimondi–Boyd [23] investigation is good, especially for low
Sommerfeld numbers. Small differences appear for high Sommer-
feld numbers (low relative eccentricities). The computed ratio Qs/
Q0 in this paper, presented in Fig. 18a is in very good agreement
with the one of Raimondi and Boyd [23].
Fig. 19a shows the mean over maximum film pressure on the
journal surface. Again, a very good agreement is observed. Fig. 20a
Fig. 13. Attitude angle vs. relative eccentricity for L/D ¼ 1. shows the angle of the maximum pressure. The angles computed
in this paper are a little higher (up to 21) when compared to the
values of Raimondi and Boyd [23]. One reason is due to the grid
is very good agreement between the values predicted by density of the computational model. In the present work, 360
this analysis and [23]. Figs. 14a and 15a show the normalized divisions are used in the circumferential direction, while the
friction coefficient on the bearing and journal surface conse- respective number of Raimondi and Boyd [23] varies from 20 for
quently. low eccentricities to 60 for high eccentricities.
The friction coefficient computed by Raimondi and Boyd [23] Fig. 5 shows the pressure distribution vs. bearing angle for
on the journal surface is in excellent agreement with the one from relative eccentricities e ¼ 0.5 and 0.7. The results of the present
the present work computed on the bearing surface. However, work are compared with the numerical results of the Wada et al.
Author's personal copy
ARTICLE IN PRESS

K.P. Gertzos et al. / Tribology International 41 (2008) 1190–1204 1199

Fig. 14. Normalized friction coefficient on bearing surface vs. Sommerfeld number for: (a) T0 ¼ 0, (b) T0 ¼ 0.4, (c) T0 ¼ 0.8 and (d) T0 ¼ 5.

[8] investigation for a Newtonian lubricant (T0 ¼ 0) and for computational model seems to be valid and for a Bingham
L/D ¼ 1. The agreement is excellent. lubricant.
The computational model constructed gives satisfactory
results for a Newtonian lubricant and thus is considered to be 4.3. Core formation
valid for design purposes.
In this paragraph, the predicted region of the core formed in
4.2. Validation of the CFD model with a Bingham lubricant the Bingham lubricant is attempted. When the shear stress in the
film does not exceed the yield value, the shear deformation of the
Results are extracted for relative eccentricities e ¼ 0.30, 0.41, material remains zero, the velocity gradient will be zero over a
0.50, 0.60, and 0.71 for two non-Newtonian cases with T0 ¼ 0.4 finite part of the film thickness, and a Bingham solid is formed. For
and 0.8 and L/D ¼ 1. The pressure distribution for the above cases simplicity of description, the following symbols are defined: p0y ¼
is compared with the numerical and experimental data of Wada qp=qy and p0z ¼ qp=qz, where z is the coordinate along the bearing
et al. [8]. The boundary condition for the numerical results of length. From the equilibrium of forces on an element in film, it is
the above investigation is the half-Sommerfeld, in agreement with clear that the previous pressure gradients are closely related to
the B.C. by Eq. (5). Figs. 6 and 7 show the comparative results for the core formation, because the gradients of qtx =qr and qty =qr are
the above cases. The change in temperature, measured during the equal to them [5], where r is the coordinate normal to the bearing
experiment of Wada et al. [8] is within 2 1C. This is close to the surface. The pressure gradient p0z is equal to zero at the bearing
assumption of paragraph 2.1 for isothermal flow. The experi- center, as the pressure distribution is symmetrical to it. When the
mental data in Figs. 6 and 7 are indicated with different symbols. pressure gradient p0y ¼ 0, no core is formed. When the gradient p0 y
A negative pressure is measured slightly in diverging films but its is present the shear stress varies in r direction and its distribution
value is very small as compared with positive pressure in is equal to p0 y. When the resultant shear stress of the bearing is
converging films. These results may satisfy the half Sommerfeld less than the yield stress or the gradient p0 y is large enough and
boundary conditions. In every case presented in the above figures, greater than zero, a core is formed and adheres to the bearing
the results of this work are in excellent agreement with the surface. When the pressure gradient increases further and the
theoretical and experimental ones of Wada et al. [8]. The shear stress on the bearing surface exceeds the yield stress, then
Author's personal copy
ARTICLE IN PRESS

1200 K.P. Gertzos et al. / Tribology International 41 (2008) 1190–1204

Fig. 15. Normalized friction coefficient on journal surface vs. Sommerfeld number for: (a) T0 ¼ 0, (b) T0 ¼ 0.4, (c) T0 ¼ 0.8 and (d) T0 ¼ 5.

of the y and z component of the shear stress. The thickness


decreases as the gradient p0 z increases.
The extent of the core depends on the geometrical character-
istics of the journal bearing, i.e. relative eccentricity, L/D ratio, and
the value of the dimensionless yield stress T0. The volume that a
core occupies is greater for a larger value of L/D for a specific
relative eccentricity and T0. This can be explained because the
pressure distribution across the z direction is smoother for
bearing with large length than one with a small length and
becomes a flat line for an infinite bearing. As explained previously
an increment to the gradient p0 z leads to core reduction. Also, the
range of the core increases with T0, for a specific relative
eccentricity and L/D ratio. This is because T0 is proportional to
the yield stress t0. A core is formed at a specific relative
eccentricity, when the gradient p0 y is large enough and adheres
to the bearing surface at the inlet side (y ¼ 0). The core ‘‘moves’’ to
the outlet side (y ¼ p) for greater values of relative eccentricity,
Fig. 16. Dimensionless frictional forces vs. relative eccentricity for L/D ¼ 1.
because of the corresponding increase of the p0 y gradient. As the
value of eccentricity increases more, the solid on the bearing
separates into two or three parts and a floating core between
the core is detached from the surface and floats in the film. Under these parts is observed. At a high value of relative eccentricity, a
the floating core reverse flow occurs. The range of the core core is formed and adheres to a small region of the journal,
decreases as p0 y increases. When p0 yo0 the core adheres to the because the pressure gradient p0 yo0. Fig. 8a shows a schematic
journal surface. The thickness of the core has a maximum value at representation of the three types of core that occur on a journal-
the bearing center, where the pressure gradient p0 z ¼ 0. The bearing system. The core is symmetrical to the plane between the
gradient p0 z also affects the core formation, which is the resultant journal and bearing centers. Fig. 8b shows a 3D representation of
Author's personal copy
ARTICLE IN PRESS

K.P. Gertzos et al. / Tribology International 41 (2008) 1190–1204 1201

Fig. 17. Dimensionless maximum lubricant flow rate vs. Sommerfeld number for: (a) T0 ¼ 0, (b) T0 ¼ 0.4, (c) T0 ¼ 0.8 and (d) T0 ¼ 5.

the core numerically calculated (and presented in next Fig. 9a–c) floating core. The ratio of the core adhered to journal over the
and is explained in the next paragraph. Only the bottom part of journal surface is very small and has a maximum value of 0.02 for
the journal is presented (between y ¼ 0 and p) as the core is L/D ¼ 1 and 0.05 for L/D ¼ 2. The dimensionless core thickness for
symmetrical to the upper part. the adhered to journal core has a maximum value of 0.22 for
The range of the core adherent to the bearing or to the journal L/D ¼ 1 and 0.31 for L/D ¼ 2. Vector diagrams for the shear stress
is defined in FLUENT by plotting a contour of wall shear stress on on the bearing surface, for the cases of floating core, show reverse
these surfaces having as maximum value the yield stress of the flow under the range of the core.
lubricant. The region of the floating core is determined as follows: Fig. 9a–c are compared to the theoretical and experimental
FLUENT considers the ‘‘rigid’’ material as a very viscous fluid with results of Milne [3], Tichy [10], and specifically of Wada et al. [8].
large viscosity m0 (500 Pa s in the current simulation), while the The procedure described above accurately predicts the values of e
‘‘fluid’’ has a value 0.2 Pa s for viscosity. The transition from ‘‘rigid’’ and T0, and the core is formed for a specific ratio L/D. Also, the
state to ‘‘fluid’’ state is not instantaneous. In practical terms, when bearing angle at which the core appears is well predicted.
the molecular viscosity has a value greater than 10 Pa s, a ‘‘solid’’ is Fig. 9a–c look like the photographs taken during the experimental
formed. Applying this value to the bearing and journal surfaces, procedure of the previous mentioned investigations.
the same regions are obtained for the core adhered to the journal As a conclusion, the computational model is validated against
and the bearing surfaces. Creating a so-called ‘‘iso-surface’’ experimental data for a Bingham lubricant and is used to compute
(contour), a surface with constant lubricant molecular viscosity the journal-bearing performance characteristics.
10 Pa s, the volume of the floating solid is defined.
Fig. 9a–c depicts the core region for various eccentricity
ratios, dimensionless yield stress values T0 and values for 4.4. Performance characteristics of a journal-bearing for a Bingham
L=D ¼ 12; 1 and 2, respectively. The dark parts in the figures lubricant
indicate the range of the core adhering to the bearing or journal
surface and the hatched parts the range of the core floating in the Results for the journal-bearing performances are presented in
film. The dimensionless core thickness t* and the core area a have the following Figs. 10–20, which have the form of Raimondi
also been calculated, for both the core adhered to bearing and the and Boyd graphs (performance characteristic vs. Sommerfeld
Author's personal copy
ARTICLE IN PRESS

1202 K.P. Gertzos et al. / Tribology International 41 (2008) 1190–1204

Fig. 18. Side leakage over maximum lubricant flow rate ratio vs. Sommerfeld number for: (a) T0 ¼ 0, (b) T0 ¼ 0.4, (c) T0 ¼ 0.8 and (d) T0 ¼ 5.

number). The performances are calculated for length over In Figs. 14a to 15d, the relation between the normalized friction
diameter ratios L=D ¼ 14; 12; 1; 2; and 4 and for dimensionless shear coefficient and the Sommerfeld number is presented. The
stress T0 ¼ 0, 0.4, 0.8, and 5. The value T0 ¼ 0 indicates a frictional forces on both surfaces increase with the eccentricity
Newtonian lubricant. ratio. The forces increase rapidly for relative eccentricities greater
Fig. 10a–d shows the relative eccentricity. As the dimensionless than 0.8, and Fig. 16 shows the relationship of dimensionless
shear stress increases, the relative eccentricity is decreased for frictional forces with relative eccentricity. The frictional force on
constant Sommerfeld number. Thus the load capacity of a the journal surface is larger than that on the bearing surface under
Bingham solid is larger than of a Newtonian fluid and increases the same conditions. The frictional forces become larger as the
together with the yield stress T0, as shown in Fig. 11. The relative dimensionless yield stress T0 increases. The friction coefficient
eccentricity decreases as the L/D ratio increases, as it is expected. also increases with an increase in the dimensionless yield stress T0
Fig. 12a–d depicts the attitude angle versus Sommerfeld under the same Sommerfeld number. The coefficient of friction
number. The curves appear a distortion for values of Sommerfeld calculated on the journal surface is also larger than that on the
number less than 0.15 and L/D ratio 2 and 4 for the non- bearing surface. This tendency is more pronounced as the
Newtonian cases. To investigate the influence of the core presence Sommerfeld number decreases. The friction coefficient decreases
to this distortion, the area where the core occurs is marked. When with an increase in L/D ratio.
a large area of a core adhered to bearing is present (for large L/D In Fig. 17a–d, the relation between the maximum dimension-
and T0 numbers), the attitude angle is decreasing leading to the less flow rate and the Sommerfeld number are presented. The flow
curves distortion. For very small Sommerfeld numbers the core is rate decreases as expected with the dimensionless yield stress T0.
detached from the surface and floats in the film. The floating core The flow rate is larger for smaller length over diameter ratios.
does not affect the attitude angle as much as the adhered core and Fig. 18a–d shows the lubricant side leakage over the maximum
the curves go back to their initial shape. The attitude angle flow rate ratio vs. the Sommerfeld number. The side leakage
increases with the dimensionless yield stress T0. The locus of the decreases with large length to diameter ratios, as expected and
journal center in a Bingham solid is slightly increased compared also with large dimensionless yield stress T0.
to that in a Newtonian fluid, as shown in Fig. 13, and the attitude In Fig. 19a–d, the relationship between the mean over the
angle increases with the L/D ratio. maximum pressure ratio and the Sommerfeld number is shown.
Author's personal copy
ARTICLE IN PRESS

K.P. Gertzos et al. / Tribology International 41 (2008) 1190–1204 1203

Fig. 19. Mean over maximum pressure ratio vs. Sommerfeld number for: (a) T0 ¼ 0, (b) T0 ¼ 0.4, (c) T0 ¼ 0.8 and (d) T0 ¼ 5.

This ratio increases with the dimensionless yield stress T0. Both The volume that a core occupies is greater for a larger value of
the mean and the maximum film pressure increase with the L/D and T0 for a specific relative eccentricity. A core is formed at a
dimensionless yield stress number, but the numerator of this ratio specific relative eccentricity and adheres to the bearing surface at
increases more than the denominator. The range where the core is the inlet side. The core ‘‘moves’’ to the outlet side for greater
present is also marked and in this diagram. The presence of the values of relative eccentricity. At a high value of relative
core influences the shape of the curves less than the attitude angle eccentricity, a core is formed and adheres to a small region of
ones. The presence of the core leads to maximum pressure the journal. As the value of eccentricity increases, the solid on the
increase and consequently to decrease of the pm/pmax ratio, which bearing separates into two or three parts and a floating core
is represented by the distortion of the curves shape for L/D ratio 2 between these parts is observed.
and 4 and T0 ¼ 5 (cases with great presence of core). The angle of The load carrying capacity, the film pressure, and the
the maximum pressure is shown in Fig. 20a–d. frictional force of a Bingham solid are larger than those of
a Newtonian fluid and they increase as the yield stress T0
increases.
5. Conclusions For low of eccentricity ratios, the effect of yield stress T0 on the
journal behavior is small. In ER and MR fluids, the yield stress
The performance characteristics and the core formation in a increases by varying the electric and magnetic field consequently.
hydrodynamic journal bearing lubricated with a Bingham fluid Therefore, semi-active control of the journal is possible by varying
were examined. The Navier–Stokes equations were solved using these quantities. The above CFD-validated model should be used
the FLUENT package. The results of the developed 3-D CFD model for such an investigation.
were compared with theoretical and experimental results of The designer–engineer of smart journal bearings, who wants
previous investigations, for both Newtonian and Bingham lubri- to design such a smart bearing, can use Figs. 10a–d for a specific
cants, and found to be in very good agreement. The validated T0, to find the equilibrium position of the journal, and Figs. 14a
model is used to extract a series of diagrams in the form of the to 15d to find the corresponding friction coefficient, etc. Thus,
Raimondi and Boyd graphs [23]. The following results are the use of the present work in the smart bearing design is
obtained: demonstrated.
Author's personal copy
ARTICLE IN PRESS

1204 K.P. Gertzos et al. / Tribology International 41 (2008) 1190–1204

Fig. 20. Maximum pressure angle vs. Sommerfeld number for: (a) T0 ¼ 0, (b) T0 ¼ 0.4, (c) T0 ¼ 0.8 and (d) T0 ¼ 5.

References [14] Swamy STN, Brabhu BS, Rao BVA. Steady state and stability characteristics of a
hydrodynamic journal bearing with a non-Newtonian lubricant. Wear
1977;2(42):229–44.
[1] Lord Corporation. Designing with MR fluids. Engineering note /www.lord.- [15] Tayal SP, Sinhasan R, Singh DV. Analysis of hydrodynamic journal bearings
comS, Rev 12/99. with skewed axes using non-Newtonian lubricants. Wear 1982;3(82):
[2] Cohen G, Oren JW. Film pressure distribution in grease lubricated journal 291–307.
bearing. Trans ASME 1949;71:555. [16] Sinhasan R, Goyal KC. Transient response of circular journal bearing
[3] Milne AA. A theory of rheodynamic lubrication. Kolloid Z 1954;139:96–100. lubricated with non-Newtonian Lubricants. Wear 1992;2(156):385–99.
[4] Batra RL. Rheodynamic lubrication of a journal bearing. Appl Sci Res 1965;Sec [17] Sharma SC, Jain SC, Sah PL. Effect of non-Newtonian behavior of lubricant and
A 15:331–44. bearing flexibility on the performance of slot entry journal bearings. Tribol Int
[5] Wada S, Hayashi H, Haga K. Behavior of Bingham solid in hydrodynamic 2000;7(33):179–84.
lubrication. Part 1, general theory. Bull JSME 1973;92(16):422–31. [18] Tayal SP, Sinhasan R, Singh V. Analysis of hydrodynamic journal bearings
[6] Wada S, Hayashi H, Haga K. Behavior of Bingham solid in hydrodynamic having non-Newtonian power law lubricants by the finite element method.
lubrication. Part 2, application to step bearing. Bull JSME 1973;92(16):432–40. Wear 1981;1(71):15–27.
[7] Hayashi H, Wada S. Hydrodynamic lubrication of journal bearings by pseudo [19] Ranghunandana K, Majumdar BC. Stability of journal bearing systems using
plastic lubricants considering effects of correlation. Part 3, theoretical non-Newtonian lubricants: an non linear transient analysis. Tribol Int
analysis. Bull JSME 1974;109(17):432–40. 1999;4(32):179–84.
[8] Wada S, Hayashi H, Haga K. Behavior of Bingham solid in hydrodynamic [20] Tayal SP, Sinhasan R, Singh DV. Analysis of hydrodynamic journal bearings
lubrication. Part 3, application to journal bearing. Bull JSME 1974;111(17):1182–91. having non-Newtonian lubricants. Tribol Int 1982;3(25):17–21.
[9] Mutuli S, Bonneau D, Frene J. Velocity measurements in the grease- [21] Chen PYP, Hahn EJ. Side clearance effects on squeeze film damper
lubricating film of a sliding contact. ASLE Trans 1985;29(4):515–22. performance. Tribol Int 2000;3–4(33):161–5.
[10] Tichy JA. Hydrodynamic lubrication theory for the Bingham plastic flow [22] Ranjan V, Pai R, Hargreaves DJ. Stiffness and damping coefficients of
model. J Rheol 1991;35(4):477–96. 3-axial grooved water lubricated bearing using perturbation technique.
[11] Kim JH, Seireng AA. Thermodynamic lubrication analysis incorporating In: Proceedings of fifth EDF & LMS poitiers workshop: bearing behavior
Bingham rheological model. J Tribol 2000;1(122):137–46. under unusual operating conditions. Futuroscope 5 October 2006.
[12] Peng J, Zhu KQ. Hydrodynamic characteristics of ER journal bearing with p. Q.1–6.
external electric field imposed on the contractive part. J Intell Mater Systems [23] Raimondi AA, Boyd J. A solution for the finite journal bearing and its
Struct 2005(16):493–9. application to analysis and design: III. Trans ASLE 1958;1(1):194–209.
[13] Lu X, Khonsari MM. An experimental investigation of grease-lubricated [24] Fluent. Fluent 6.3 user’s guide. Lebanon: New Hampshire (USA) Fluent Inc.;
journal bearings. Trans ASME 2007(129):84–90. 2006.

You might also like