You are on page 1of 32

INTERNATIONAL JOURNAL FOR NUMERICAL METHODS IN ENGINEERING

Int. J. Numer. Meth. Engng 2004; 60:1759–1790 (DOI: 10.1002/nme.1022)

Finite element method for mixed elastohydrodynamic lubrication


of journal-bearing systems

Wing Kam Liu1,∗,†,‡ , Shangwu Xiong1,§ , Yong Guo1,¶ , Q. Jane Wang1, ,


Yansong Wang1,∗∗ , Qingmin Yang2,†† and Kumar Vaidyanathan2,‡‡
1 Department of Mechanical Engineering, Northwestern University, 2145 Sheridan Road,
Evanston, IL 60208, U.S.A.
2 Federal-Mogul Technical Center, 47001 Port Street, Plymouth, MI 48170, U.S.A.

SUMMARY
A finite element model is presented for mixed lubrication of journal-bearing systems operating in
adverse conditions. The asperity effects on contact and lubrication at large eccentricity ratios are
modelled. The elastic deformation due to both hydrodynamic and contact pressure, and the cavitation
of the lubricant film are considered in the model system. Two verification problems with both
theoretical and experimental comparisons are given to show the effectiveness of this model. Finally,
a new example is presented which discusses the influence of waviness depth, secondary roughness,
external force and shaft speed on the mixed lubrication. Copyright 䉷 2004 John Wiley & Sons, Ltd.

KEY WORDS: mixed elastohydrodynamic lubrication; journal-bearing system; finite element method

0. INTRODUCTION

Mixed lubrication consists of a partial lubrication region where both hydrodynamic lubrication
and boundary lubrication coexist. Intermittent material contacts may appear when the oil film

∗ Correspondence to: W. K. Liu, Department of Mechanical Engineering, Northwestern University, 2145 Sheridan
Road, Evanston, IL 60208, U.S.A.
† E-mail: w-liu@northwestern.edu
‡ Walter P. Murphy Professor
§ Senior Research Associate
¶ Formerly Research Assistant, Mechanical Engineering, Northwestern University. Currently Senior Scientist,
Livermore Software Technology Corp., 7374 Las Positas Road, Livermore, CA94550, U.S.A.
Associate Professor
∗∗ Post Doctoral Fellow
†† Senior Technical Analyst
‡‡ Chief Technical Analyst

Contract/grant sponsor: NSF


Contract/grant sponsor: NSF-IGERT
Contract/grant sponsor: Federal Mogul Corporation
Contract/grant sponsor: CSET
Received 1 November 2001
Revised 15 June 2003
Copyright 䉷 2004 John Wiley & Sons, Ltd. Accepted 3 October 2003
1760 W. K. LIU ET AL.

breakdowns, which can induce surface damages such as massive wear and scuffing. Therefore,
a comprehensive understanding of the fundamental phenomena of mixed lubrication is very
important for journal bearing design.
Since Peppler [1] firstly tried to study the lubricated contact between gear teeth in 1936,
a lot of valuable work has been done in numerical computations of hydrodynamic pressures
and journal-bearing elastic deformations [2, 3]. It seems that the finite element method (FEM)
[4] and the FEM-based influence-function method [5] have been the most popular techniques
for calculating the journal-bearing elastic deformation under hydrodynamic pressures, while the
finite difference method (FDM) [6] and FEM [7] have been widely applied to obtain hydro-
dynamic pressures. Since the 1970s, several typical modellings considering surface roughness
effect for mixed lubrication problems have been developed, such as the longitudinal-roughness
flow model by Christensen [8], the average Reynolds equation by Patir and Cheng [9], and
the asperity contact model by Greenwood and Tripp [10]. Zhu et al. [11] studied the mixed
lubrication problem of a piston-cylinder contact, focusing on the convex piston. Wang et al.
[12] analysed the journal-bearing conformal contacts under steady-state conditions, considering
the combined effects of bearing deformation, surface roughness and asperity interaction. With
the increasing computational power of the modern computer, the deterministic approaches have
been proposed using artificial surfaces to simulate surface irregularities since 1980s [13]. How-
ever, these deterministic models require large computational time and sophisticated numerical
skills to obtain a converged solution [14].
Today, many thermodynamic (THD) models are available, which successfully study the
thermal effects on journal bearing lubrication [15, 16]. These models are based on ideally
smooth bearing surfaces operating with relatively low eccentricity ratios. Recently Shi and
Wang [17] developed a steady-state mixed thermo-elasto-hydrodynamic (TEHD) model for
journal-bearing conformal contacts working at large eccentricity ratios, in which FDM was
used to solve the hydrodynamic pressures.
For dynamically loaded bearings, several efficient approaches have been utilized thus far,
including the short-bearing theory [18], the mobility method [19] and the semianalytical method
[20]. Two review papers by Campbell et al. [21] and Martin [22] presented comparisons of
various analytical methods and some experimental results. Mainly since the 1980s, a more
robust approach, finite element method has found applications in journal-bearing modelling
[4, 23–26]. This paper presents a mixed lubrication study of finite journal bearings operating
under steady and dynamic loads conditions with large eccentricity ratios. The average Reynolds
equation derived by Patir and Cheng [9] is used to consider the effects of roughness on lubricant
flow, and an FEM model for solving complicated mixed lubrication problems is developed.

1. MIXED LUBRICATION MODEL FOR JOURNAL BEARINGS

1.1. Geometry of the journal bearings modelling


For continuous-sleeve bearings, the simplest class of journal bearings, a lubrication model is
illustrated in Figure 1. A lubricant forms a thin film between the outer surface (bearing housing)
and the inner (journal) surface of the pair of sliding elements. O and O  represent the position
of the axes of the housing and journal respectively, R1 and R2 are the radii of the surface
curvatures of the journal and bearing housing respectively, and e = OO  is the eccentricity

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
FINITE ELEMENT METHOD FOR MIXED ELASTOHYDRODYNAMIC LUBRICATION 1761

Figure 1. Geometry of the journal bearing model.

Figure 2. Texture of the bearing surface.

of the journal. The bearing housing is of a finite length. In the absence of the eccentricity,
the initial radial clearance between the journal and bearing housing is denoted by c, then
R2 − R1 = c. Note that c>R1 , R2 , and R1 ≈ R2 ≈ R. Two co-ordinate systems are used for
the journal-bearing model. O − XY Z is a global co-ordinate system while xz is the co-ordinate
system for the lubricant film (x = R).
For the flow of a thin lubricant film in journal bearings, the surface roughness may have
remarkable effects. The bored bearings, for example, may have a surface that can be idealized
with the texture shown in Figure 2, in which AA is a section cut along the axis of the journal.
The asperity of the surface is dominated by longitudinal wavenesses along the sliding direction
(x-direction), which is periodic in the z-direction. Finer secondary roughness may exist on the

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
1762 W. K. LIU ET AL.

Figure 3. Thickness of the lubricant film: (a) cross-section; and (b) enlargement of the circled area.

surface of the waviness. In Figure 2, the periodic length of the waviness is designated by
L = L1 + L2 + L3 , where L1 is the width of the plateau, L2 and L3 are the widths of the
valley slopes, and D is the depth of the valley. The composite surface roughness of the journal
and housing surfaces, , is defined as a root mean square of the standard deviations of the
two secondary roughness amplitudes

 = 21 + 22 (1)
where 1 and 2 are the standard deviations of the secondary roughness of the two sliding
surfaces.

1.2. Film thickness


The interspace for the lubricant is determined by the surface profiles of the journal and
housing. Figure 3 shows the film thickness distribution due to the eccentricity of the journal
(the dashed lines represent the original positions). If the deformation of the housing is taken
into consideration, while the journal is assumed to be rigid, the film thickness at an arbitrary
point (x, z) on the journal surface can be approximated by
h̄(x, z, t) ≈ c − e(t) cos( − ) + d(x, z, t) (2)
where e is the eccentricity of the journal,  is the orientation of the eccentricity, and d is the
surface distortion of the housing (positive if the surface expands outwards) and can be found
from the structural response to the fluid pressure p̄ by
d = Ep̄ (3)
where E is the compliance operator, which will be given later with a finite element approxi-
mation. If we let
eX = e sin (t), eY = e cos (t) (4)

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
FINITE ELEMENT METHOD FOR MIXED ELASTOHYDRODYNAMIC LUBRICATION 1763

then the thickness of the lubricant film can be rewritten as


h̄(x, z, t) = c − eX (t) sin  − eY (t) cos  + d(x, z, t) (5)
Here and in the following sections, h̄ indicates the mean film thickness of the lubricant.
Referring to Figure 2, the real film thickness hT has a complicated profile due to the asperities
(longitudinal wavinesses and secondary roughness). However, the characteristic length of these
asperities is far less than that of the lubrication region. Thus, we can smooth the film thickness
by averaging it over a small area. This area is of small size compared to the journal surface but
includes a large number of asperities. In the case of longitudinal roughness, the averaging area
should contain at least one waviness. The average film thickness over an individual waviness
can be derived as
h̄ = E[hT ] = E[h̄T + ] = E[h̄T ] (6)
where h̄T is the film thickness measured from the mean lever of the waviness surface and the
operator E[·] indicates an average over the period of the waviness. If the film thickness at the
plateau is designated by hm , from geometry (see Figure 2), we have

1 L
h̄ = h̄T dz
L 0
 L1  L2      
1 D D
= hm dz + hm + z dz + hm + z dz
L 0 L1 L2 L3

L2 + L3 D D
= hm + = hm +  (7)
L1 + L2 + L3 2 2

where  is a constant given by


 = 1 − L1 /L (8)
Note that the value of the mean film thickness, h̄, is larger than the film thickness measured
from the plateau of the waviness, hm . Thus, if h̄  D/2, there would be solid to solid contact
between the journal and housing. In the macroscopic sense, the mean film thickness, h̄, is a
continuous function of spatial coordinates.

1.3. Load equations


In mixed lubrication, the total load is balanced by two components: the integral of pressure
within the lubricant film and the asperity to asperity contact load. The cross-section of a journal,
which is rotating in the clockwise direction and carrying a load transverse to the axis O  , is
shown in Figure 4(a). The loads applied to the journal are listed as follows:
1. External forces (including inertial forces). The projected forces in the global X- and
Y -directions are designated by FX and FY .
2. Lubricant pressure which results in a supporting force P . On a slice sector  (see
Figure 4(b)), the normal pressure is equal to p̄R and its projected forces in the
X- and Y -directions are, respectively, dPX = −p̄R sin  and dPY = −p̄R cos .

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
1764 W. K. LIU ET AL.

Figure 4. Load equilibrium for a journal: (a) load balance; and (b) summation of pressure.

3. Contact loads. When there exists asperity to asperity contact between the journal and
housing, there are contact loads distributed on the contacted area. These loads can result
in a normal force, wa , and a friction, a . The resulting contact normal pressure is against
the eccentricity of the journal but the direction of the friction depends on the relative
sliding between the journal and housing. Since the housing is usually fixed, the load
directions are shown as in Figure 4(a) for a journal rotating clockwise. The projected
forces in the X-direction are wa sin( − ) and a cos( − ) while the projected forces
in Y -direction are wa cos( − ) and −a sin( − ).
By summing up the projected forces in the X- and Y -directions of the above loads, we can
write the load balance equations for the journal as
 
fX = − p̄(x, z, t) sin  d − wa (t) sin  − a (t) cos  + FX (t) = 0, and (9)

 
fY = − p̄(x, z, t) cos  d − wa (t) cos  + a (t) sin  + FY (t) = 0 (10)


where d = R d dz = dx dz and  is the surface of the journal bearing. In the above equations,
the angle , which indicates the orientation of the eccentricity OO  , is also a function of time
t. In the calculation of the resultant force of the pressure distribution, we have shifted the origin
of axis of the journal from O  to O. Because the eccentricity is within the range of the radial
clearance, it is small when compared with the radius of the journal surface. Therefore, the error
due to this approximation is negligible. The pressure distribution is governed by an average
Reynolds equation, which will be described in detail in the next section. The calculation of
the contact loads is given in Section 1.5.

1.4. Average Reynolds equation


Under isothermal conditions, the lubricant pressure is governed by an average Reynolds
equation [9]
 3 
h̄ p  *(h̄)
∇·  · ∇ p̄ = ∇ · h̄U + s · V + (11)
12 *t

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
FINITE ELEMENT METHOD FOR MIXED ELASTOHYDRODYNAMIC LUBRICATION 1765

Figure 5. Local coordinate system for the lubricant.

and a unit flow


h̄3 p
q = h̄U + s · V −  · ∇ p̄ (12)
12
where p̄ and h̄ are the mean pressure and mean film thickness of the lubricant respectively, U
and V are the average velocity and relative velocity of the two sliding surfaces respectively,
p and s are flow factors reflecting the effects of the surface roughness, and ,  and  are
material parameters.
The average Reynolds equation is derived from the Navier–Stokes equations by considering
the effects of the surface roughness and by assuming the followings:
• the flow is laminar and incompressible,
• the body forces and inertial forces are negligible compared with the viscous forces,
• the film thickness is small compared with the other dimensions (i.e. h̄>R1 ),
• the pressure and viscosity are constant across the thickness of the lubricant film, and no
slip exists at the fluid–solid interfaces.
The average Reynolds equation was used in the numerical simulations of sheet metal forming
by Liu et al. [27–29].
In the model of the journal bearings, the effect of the curvature is negligible because the
lubricant film is very thin. Thus, the circumferential dimension can be unfolded to a flat plane
and represented by xz (see Figure 5), where conditions should be the same at  = 0 and
2. We assumed that the motions of the shaft and housing are steady (usually the housing is
stationary) and that the rotational velocity of the shaft is chosen to be positive when it rotates
clockwise. Then the average Reynolds equation can be written as
   
* p h̄ *p̄
3 * p h̄ *p̄
3 u1 + u2 *(h̄) u1 − u2 *s *(h̄)
x + z = +  +
*x 12 *x *z 12 *z 2 *x 2 *x *t
(13)

where p̄(x, z, t) is the mean pressure within the lubricant, h̄(x, z, t) the mean lubricant film
thickness,  density of the lubricant,  viscosity of the lubricant,  r.m.s. of secondary roughness,
u1 velocity of the journal and positive for clockwise motion, u2 velocity of the bearing housing
and positive for clockwise motion, and p (h̄) pressure flow factor matrix and s (h̄) shear flow
factor matrix.
The flow factors reflect the effects of surface roughness and are therefore dependent on
the roughness of the sliding surfaces. In Equation (13), the two terms on the left-hand side

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
1766 W. K. LIU ET AL.

represent the pressure-driven flows, the first term on the right-hand side is due to the fluid
entraining and the second term indicates the effect of the surface roughness on shear and the
third term accounts for the normal motion (squeezing) of the two sliding surfaces.
The flow factor matrices can be obtained by comparing the stochastic Reynolds equation
with the average Reynolds equation. For the surface texture of the bearings shown in Figure
2, the flow factors given by Ai and Cheng [26] are (also see Appendix A):
 2  2  3
p  3 D 1 2 D
x = 1 + 3 + (1 − ) + ( − 1) (14)
h̄ 4 h̄ 4 h̄
 
D 2
1−
p 2h̄
z = , and (15)
D
2 + (1 − )
1−  h̄ 
+ 

D 2  D
1− 1+ 1−
2h̄ 2 h̄

s = 0 (16)

The above flow factors depend on the surface roughness parameters, which, in our case, are
D and  for the primary roughness (waviness) and  for the secondary roughness. However,
the effects of the surface roughness vary with the film thickness because the amplitudes of the
surface roughness enter the above equations as ratios to the mean film thickness. When >h̄
p p
and D>h̄, x → 1 and z → 1, the flow of the lubricant is like that between two smooth
surfaces. On the other hand, when the lubricant film is thin, the surface roughness influences
the flow substantially. A matrix H (h̄) can be defined as
 3 
p h̄
 x 12 0 
 
H (h̄) =   , and (17)
 p h̄
3 
0 z
12
u1 + u2
u= (18)
2
and g(x, z) is a scalar function and v(x, z) is a vector function with
 

 *g 
 
 
*x  *vx *vz
∇g = , and ∇ · v = + (19)

 *g 
 *x *z

  
*z
then the average Reynolds equation (Equation (13)) can be rewritten as
*h̄ ˙
∇ · (H · ∇ p̄) = u + h̄ (20)
*x

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
FINITE ELEMENT METHOD FOR MIXED ELASTOHYDRODYNAMIC LUBRICATION 1767

Figure 6. Asperity contact between the journal and the housing: (a) cross-section;
and (b) enlargement in axial directions.

1.5. Asperity contact load


Bearings are designed to achieve full film lubrication, and contact between the bearing and
journal surfaces should be avoided. However, when the rotational velocity of the journal di-
minishes or the transverse load increases, the eccentricity of the journal may be large enough
to trigger the solid contact between the two surfaces. Also, at the start and stop periods there
exists contact between the journal and housing.
Consider an asperity ridge in the bearing waviness (see Figure 6). The width L of the asperity
ridge in the axial direction is much smaller than the contact length in the circumferential
direction. Therefore, it can be assumed that the contact of the asperity ridge with the housing
surface is in a state of plane stress [26]. From the cross-sectional view, the contact pressure
is distributed with a parabolic-shape profile, but in the axial direction can be assumed to be
uniform. Thus, the value of the contact pressure, pw , is given as pressure per unit axial length.
For an individual contact asperity ridge, the relationship between the nominal overlap ov
and the contact pressure pw may be given by the Hertzian contact theory for the plane stress
state as (see Appendix B)
   
2pw Ee0
ov = ln −
(21)
E 2pw
where

ov = e − e0 , and (22)
D
e0 = c −  +d (23)
2
In Figure 6(b), e0 indicates the difference between the real radii of the two contacted surfaces
and therefore, it includes the effects of the surface waviness and the deformation of the housing.
E,
are, respectively, the equivalent Young’s modulus and Poisson’s ratio defined by
   
1 1 1 1 1 1 1 1
= + , and = + (24)
E 2 E1 E2
2
1
2

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
1768 W. K. LIU ET AL.

where E1 ,
1 , E2 , and
2 are the material property parameters for the journal and housing
respectively, d is the housing distortion at the center of the contact due to the combustion
and/or inertial forces, and pw the contact normal load per unit axial length. In Figure 6, the
solid lines indicate the current positions of the journal and housing while the dashed lines
represent the initial positions.
Assuming that the pressure is uniformly distributed along the axial direction in each individual
asperity ridge, the total contact pressure, wa , results from the contributions of all the contacted
asperity ridges in the axial direction according to


N
wa = L (1 − ∗ )pw
i
(25)
i=1

where L is the periodic length of the asperity waviness in the bearing, and ∗ a coefficient
affected by the deformation of the asperity, which is given by
 
∗ ov
1− =1−+ (26)
D

The asperity friction force a is then given by

a = fa wa (27)

where fa is the friction coefficient of the solid to solid contact.

1.6. Film cavitation


In practice, the lubricant film can only support a very slight negative pressure before film
rupture occurs. The inadmissibility of negative pressures of the fluid film is another important
source of non-linearity in the problem. It is therefore necessary to find the region where the
film is cavitated. This can be achieved by combining the complementarity problem with the
Murty algorithm [25, 30] and requires an initial estimation of the extent of the cavitated region.
For the nodes in the cavitated region it is assumed that p = 0, while in the non-cavitated
region, the pressure, p, is found by solving the Reynolds equation. Thus, the complementarity
conditions can be listed as

O(p) = 0, p0, and (28)

p = 0, O(p) < 0 (29)

where O(p) represents the average Reynolds equation (Equation (20)). Once the pressures
are calculated, the extent of the cavitated region is checked. If all the nodes satisfy the
complementarity conditions, the designations of the cavitated and non-cavitated regions are
satisfactory and the computation is advanced to the next increment.

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
FINITE ELEMENT METHOD FOR MIXED ELASTOHYDRODYNAMIC LUBRICATION 1769

2. COMPUTATIONAL SIMULATION BY FEM

2.1. Governing equations


The governing equations for the journal bearing lubrication model, as described in the previous
section, are the two load equations for the journal (Equations (9) and (10)), the average Reynolds
equation for the lubricant (Equation (20)), and the asperity contact load equations (Equations
(25) and (27)). By applying the film thickness relation (Equation (5)), the final unknowns are
the mean pressure field, p̄, and the two components of the eccentricity, eX and eY (or e and
). For example, from the film thickness relation and noting  = x/R, we can derive
*h̄ cos  sin  *d
=− eX + eY + , and (30)
*x R R *x

h̄˙ = −ėX sin  − ėY cos  + d˙ (31)

where a superposed dot means the derivative with respect to time t. The housing distortion, d,
depends on the pressure distribution in the lubricant. Substituting the above relationships into
the average Reynolds equation eliminates variable h̄. Therefore, the final governing equations
for the bearing problem can be stated as

p̄ sin  d + wa (p̄, eX , eY )(sin  + fa cos ) − FX (t) = 0 (32)


p̄ cos  d + wa (p̄, eX , eY )(cos  + fa sin ) − FY (t) = 0, and (33)

 
cos  sin  *d(p̄) ˙ p̄)
∇ · (H · ∇ p̄) = u − eX + eY + − ėX sin  − ėY cos  + d( (34)
R R *x
In the above equations,  is determined by the eccentricity of the journal (tan  = eX /eY ).
The boundary conditions for the journal-bearing problem consist of two parts. One is for the
prescribed lubricant pressure. For example, the bearing may have an oil supply groove and the
magnitude of the supply pressure is negligible, thus the lubricant pressure at the groove should
be zero. Another boundary condition occurs when there is film cavitation in the lubricated area.
At the interface between the lubricated area and the cavitated region, the fluid pressure must be
the cavitation pressure, which may be zero. This condition makes the solution to the problem
more difficult because it changes with time and is highly non-linear.
The strong form for the journal bearing simulation can be found by solving for the pressure
field in lubricant, p̄(x, z, t), and the eccentricity of the journal, eX (t) and eY (t), such that they
satisfy Equations (32)–(34), and by applying the following conditions in addition to appropriate
initial conditions:

p̄ = p∗ , on  = p , and (35)

p̄ = 0, on  = f (t) (36)

where p∗ is the prescribed pressure on boundary p , and f indicates the free interface between
the cavitated region and the lubricated region.

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
1770 W. K. LIU ET AL.

2.2. Weak form equations


All the equations in the previous section, except for the average Reynolds equation and the
deformation of the housing, should be satisfied exactly. Because the deformation of the housing
is determined by the structural response, the formulation can be found in a textbook on
computational structural mechanics. Here, we need to write the weak form only for the average
Reynolds equation.
The weak form of the average Reynolds equation can be obtained by multiplying Equation
(20) by a variation of pressure, p, and integrating over the whole domain of the problem (the
surface of the journal bearing). After using the divergence theorem we have
  
*h̄
∇ p · (H · ∇ p̄) d + pu d + p h̄˙ d = 0 (37)
  *x 
Substituting Equations (30) and (31) into the above equation yields

   
*d ˙
0 = ∇ p · H · ∇ p̄ d + pu d + pd d − p cos  deX
  *x  
  
− p sin  dėX + p sin  deY − p cos  dėY (38)
  

where = u/R.
Thus, the weak form of the journal bearing simulation can be stated as: Solve for p̄, eX and
eY , such that for any variation, p, Equations (32), (33) and (38), and the boundary conditions
are satisfied.

2.3. The galerkin form of finite element discretization


The model of the journal bearings consists of three objects: the journal, the housing and the
lubricant. Because we assume that the journal is rigid, its motion can be determined by the
eccentricity and the rotational velocity. The domain of the lubricant fluid and the bearing
housing is discretized using the finite element method [4, 31].

2.3.1. Deformation of the bearing housing. If the thermal effect and the asperity contact load are
not considered for the distortion of the bearing housing, only the lubricant pressure contributes
to the deformation of the housing. The lubricant pressure and its variation can be interpolated
as follows
p̄ = NP, p = N P (39)
where N are the shape functions, and P and P are nodal pressure parameters and variations,
respectively. We use continuous finite elements to discretize the whole structure of the bearing
housing, which leads to the following set of linear algebraic equations
Kh d = fhext (40)
where d represents displacements at all the nodes in the housing model and Kh is the stiffness
matrix. The external nodal forces, fhext , can be calculated from the pressure applied by the

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
FINITE ELEMENT METHOD FOR MIXED ELASTOHYDRODYNAMIC LUBRICATION 1771

lubricant on the inner surface of the housing by


 
fhext = Nht p̄ d = Fh P, and Fh = Nht N d (41)
 

where Nh are the shape functions on the inner surface of the housing. The housing distortion
on the inner surface is then determined by

d(x, z, t) = Nh d = Nh Kh−1 Fh P = E(x)P(t) (42)

The deformation matrix, E(x), and its derivative with respect to x, Ex , are given by

E(x) = Nh (x)Kh−1 Fh , and (43)

*Nh (x) −1
Ex (x) = Kh Fh (44)
*x

2.3.2. Discretization of the average Reynolds equation. Since the lubricant is a thin film,
we use two-dimensional elements with neq nodes to discretize the fluid domain. Substituting
Equations (39) and (42) into the weak form of the average Reynolds equation (Equation (38)),
we have
   
R= ∇Nt H∇N d + Nt uEx d P + Nt E dṖ
  
 
− Nt cos  deX − Nt sin  dėX (45)
 
 
+ Nt sin  deY − Nt cos  dėY = 0
 

2.3.3. Discretization of load equations. The discretization of the two load equations for the
journal-bearing (Equations (32) and (33)), leads to

Rneq +1 = N sin  dP + WX − FX (t) = 0, and (46)


Rneq +2 = N cos  dP + WY − FY (t) = 0 (47)


where

WX = wa sin  + a cos , and (48)

WY = wa cos  + a sin  (49)

Note that wa is a function of eccentricity and lubricant pressure (given by Equation (25)).

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
1772 W. K. LIU ET AL.

2.3.4. Linearized equations. The above discretized equations, Equations (45)–(49), are the
finite element equations for the unknowns, P, eX and eY . Since they are highly non-linear, full
implicit integration must be used to solve these equations. Therefore, we have to derive the
linearized system of equations.
For the discretized average Reynolds equation (Equation (45)), we have
   
*R 1 *H
= ∇Nt H∇N d + Nt uEx d + Nt E d + ∇Nt ∇N dP (50)
*P   t  *P
  
*R 1 *H
= − Nt cos  d − Nt sin  d + ∇Nt ∇N dP, and (51)
*eX  t   *eX
  
*R 1 *H
= Nt sin  d − Nt cos  d + ∇Nt ∇N dP (52)
*eY  t   *eY

where
*Ṗ *ėX *ėY 1
= = = (53)
*P *eX *eY t
Since
 
p h̄3
 x (h̄) 0   
 12  H11 0
 
H (h̄) =  = (54)
 p h̄ 
3
0 H22
 0 z (h̄) 
12

and recalling the film thickness relationship, (Equation (5)), we can write

*H11 *H11 *h̄ *H11


= = E, and (55)
*P *h̄ *P *h̄
*H22 *H22
= E (56)
*P *h̄
Similarly, we have

*H11 *H11 *H22 *H22


= (− sin ), = (− sin ) (57)
*eX *h̄ *eX *h̄
*H11 *H11 *H22 *H22
= (− cos ), and = (− cos ) (58)
*eY *h̄ *eY *h̄

For the discretized load equations (Equations (46) and (47)), we have

*Rneq +1 *WX
= N sin  d + (59)
*P  *P

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
FINITE ELEMENT METHOD FOR MIXED ELASTOHYDRODYNAMIC LUBRICATION 1773

*Rneq +1 *WX *Rneq +1 *WX


= , = (60)
*eX *eX *eY *eY

*Rneq +2 WY
= N cos  d + (61)
*P  *P

*Rneq +2 *WY *Rneq +2 *WY


= , and = (62)
*eX *eX *eY *eY

Note that eX = e sin  and eY = e cos , so

*e * cos 
= sin , =
*eX *eX e
(63)
*e * sin 
= cos , and =−
*eY *eY e

then from Equations (48) and (49), which define WX and WY , we obtain

WX *wa
= (sin  + fa cos ) (64)
*P *P
*WX *wa wa
= sin (sin  + fa cos ) + cos (cos  − fa sin ) (65)
*eX *e e

*WX *wa wa
= cos (sin  + fa cos ) − sin (cos  − fa sin ) (66)
*eY *e e

WY *wa
= (cos  − fa sin ) (67)
*P *P
*WY *wa wa
= sin (cos  − fa sin ) − cos (sin  + fa cos ), and (68)
*eX *e e

*WY *wa wa
= cos (cos  − fa sin ) + sin (sin  + fa s cos ) (69)
*eY *e e

Using the contact load equation (Equation (25)), we have

*wa N *pi N *∗



= L (1 − ∗ ) w − L i
pw
*P i=1 *P i=1 *P


N −E − 2pw i /e  N
(1 − ∗ ) 
0
=L  E − EL pw i
, and (70)
Ee 0 D
i=1
2 ln i

− 1 i=1
2pw

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
1774 W. K. LIU ET AL.

*wa N *pi N *∗



= L (1 − ∗ ) w − L i
pw
*e i=1 *e i=1 *e


N E  N
=L (1 − ∗ )   + L pw i
(71)
Ee 0 D
i=1
2 ln i

−1 i=1
2pw

2.4. Matrix form of equations


Discretization of the Galerkin weak formulation of the average Reynolds equation and the two
load equations for the journal leads to the following system of equations
M(D)Ḋ + K(D)D = F(t) (72)
where D are neq + 2 unknown variables and F(t) the external load vector. The mass matrix,
M, and the stiffness matrix, K, are defined as
 
M Mneq +1 Mneq +2
M= , and (73)
0 0 0
 
K Kneq +1 Kneq +2
 
K =  Kneq +1 WX  (74)
Kneq +2 WY

where

MIJ = NI EJ d (75)


neq +1
MI =− NI sin  d (76)


neq +2
MI =− NI cos  d (77)

 
KIJ = (∇N )Ii Hij (∇N)jJ d + uNI (Ex )J d (78)
 

neq +1
KI =− NI cos  d (79)


neq +2
KI = NI sin  d (80)


KJneq +1 = NJ sin  d, and (81)


Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
FINITE ELEMENT METHOD FOR MIXED ELASTOHYDRODYNAMIC LUBRICATION 1775


KJneq +2 = NJ cos  d (82)


and WX , WY can be considered as a functional giving the contact loads from the eccentricity
and the deformation of the housing. In the above expressions, I and J are in the range between
1 and neq . We can see that matrices M and K are non-symmetric.
Because the problem is highly non-linear, full implicit integration should be used according
to
 
1 1
R = K + M Dn+1 − MDn − F(tn+1 ) = 0 (83)
t t
For the Newton-Raphson iterative scheme, we have
 (v)
*R
D(v+1) = −R(v) , and (84)
*D
(v+1) (v)
Dn+1 = Dn+1 + D(v+1) = Dn + Dn+1 (85)

where
 (v)
Dn+1 = Dn+1 (86)
v

The Jacobian matrix is defined as


 neq +1 neq +2 
K̄ K̄ K̄
 
*R  neq +1 neq +2 
K̄ equ
= =  K̄neq +1 K̄neq +1 K̄neq +1  (87)
*D 


neq +1 neq +2
K̄neq +2 K̄neq +2 K̄neq +2

where

1 *Hij
K̄IJ = KIJ + MIJ + (∇N)Ii (∇N)jK EJ dPK (88)
t  *h̄

neq +1 n +1 1 neq +1 *Hij
K̄I = KI eq + M − (∇N)Ii (∇N)jK sin  dPK (89)
t I  *h̄

neq +2 n +2 1 neq +2 *Hij
K̄I = KI eq + M − (∇N)Ii (∇N)jK cos  dPK (90)
t I  *h̄

J *WX
K̄neq +1 = KJneq +1 + (91)
*PJ

J *WY
K̄neq +2 = KJneq +1 + (92)
*PJ

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
1776 W. K. LIU ET AL.

neq +1 *WX
K̄neq +1 = (93)
*eX

neq +2 *WX
K̄neq +1 = (94)
*eY

neq +1 *WY
K̄neq +2 = , and (95)
*eX

neq +2 *WY
K̄neq +2 = (96)
*eY

3. NUMERICAL EXAMPLES

In order to test the performance of the finite element model, we present the following three
numerical examples:
(1) Without asperity contact and the roughness effect, we compare the results with those
of by Wang et al. [12] where combined FDM and FEM were used for a finite journal
bearing under constant loads with large eccentricity.
(2) Without asperity contact and the roughness effect, we compare the numerical results with
the experimental ones by Patrick [32] under dynamic loads.
(3) Considering asperity contact and the roughness effect, a continuous-sleeve journal-bearing
problem is solved. The influence of waviness depth, secondary roughness, and external
force and shaft speed on the mixed lubrication are discussed.

3.1. Comparisons under constant loads


A schematic representation of the journal bearing geometry for this problem is shown in
Figure 7(a). Both the journal and the bearing are assumed ideally smooth, and the journal is
assumed rigid because it has a higher Young’s modulus (the elastic modulus (E) of the bearing
and the journal is 73 and 210 GPa, respectively). An engine oil is used at 100◦ C (viscosity:
11.36cSt or 0.01 PaS) [12]. Due to symmetry, only half bearings are used.
A good agreement is found when compared with the results given by Wang et al. [12],
which used FDM for solving Reynolds equation and FEM for the elastic deformation of the
housing. The comparisons are shown in Figures 7(b) and (c), where hT is Dyson’s average
gap–compliance relationship for positive compliance values [33] according

hT / = h/ + e−(0.599+0.9936(h/)
2 +0.22(h/)3 +0.04545(h/)4 )

and the Sommerfeld number is S = uLR 2 /(P c2 ). Moreover, the simulations show that for
rigid cases, the convergence is easily obtained when the eccentricity ratio c/e is less than 0.99,
but the solution finally fails to converge as c/e tends towards 1.

3.2. Comparisons under Dynamical loading


The performance of sleeve bearings subjected to dynamic loads is simulated to match one of the
experiments conducted by Patrick [32]. The diametrical clearance is 0.0015 in (or 0.0381 mm),

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
FINITE ELEMENT METHOD FOR MIXED ELASTOHYDRODYNAMIC LUBRICATION 1777

Figure 7. A comparison of the total bearing load (L/D = 2/3, 500 rpm,  = 0.5 m): (a) the journal
bearing structure. Bearing ID = 24.4 mm, bearing OD = 28.4 mm, total height = 30.5 mm, radial
clearance = 0.0254 mm; (b) as a function of the eccentricity ratio; and (c) as a function of the
dimensionless minimum film thickness, m = (hT /)m .

the alternating load is sinusoidal with a maximum load of 4.33 tonf (or 42434 N), both the
journal speed and load frequency are 320 rpm. Two diametrically opposed axial grooves, 1/4 in
(or 6.35 mm) broad, were cut in the bearings. The grooves were set in a plane normal to the
load plane and were supplied with oil from a 1/4-in (or 6.35 mm) diameter hole situated at
the mid-point of each groove, and an SAE 40 mineral oil was used [32]. In this paper, both
the journal and the bearing are assumed ideally smooth and rigid, and we take the oil viscosity
as 10−5 reyns (or 0.068948 PaS).

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
1778 W. K. LIU ET AL.

Figure 8. Numerical predictions of the minimum film thickness and the oil film pressure: (a) history
of hmin ; and (b) a comparison of the pressure distribution at different load phase angles.

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
FINITE ELEMENT METHOD FOR MIXED ELASTOHYDRODYNAMIC LUBRICATION 1779

Figure 9. Continuous-sleeve journal bearings.

Table I. Main conditions for a continuous-sleeve journal-bearing problem.


Radius of shaft, 12.192 mm Initial clearance, 0.00254 mm
External radius of housing, 14.224 mm Waviness depth, Dwav = 0.882, 3, 4, 5, 6 m
Length of housing, 8.218 mm Waviness length: 0.3385 mm
Initial radial clearance: 0.00254 mm Waviness parameter L1 : 0.1077 mm
Young’s modulus for journal, 205 GPa Sigma of secondary roughness,  = 0.5, 1.0, 1.5 m
Young’s modulus for housing, 73 GPa Poisson’s ratio for journal and housing, 0.29
Viscosity of lubricant, 59.9 N∗ S/mm2 Angular speed of shaft, us = 2000, 4000, 8000 rpm
Friction coefficient, 0.1 Angular speed of housing, 0.0
Frequency of external force, Amplitude of external force, F̄ext = 10, 20, 30, 40 kN
uext = 0.5us

The predicted minimum film thickness vs crank angle, illustrated Figure 8(a), shows that
the steady solution is obtained after only two cycles of loading. Figure 8(b) shows that the
generation of load-bearing pressures in the oil film is well known as a predominantly squeeze
film action, with the maximum oil film pressures occurring at the instant of the maximum of
the applied load. The numerical results present a similar trend.

3.3. Bearing performance under dynamic loads and major factors to be considered
The continuous-sleeve journal-bearing problem, shown in Figure 9, is solved next. For simplicity,
the housing is assumed as a cylindrical tube of a finite length operating under a horizontal
periodic external force F̄ext (1 − cos ext ). The main parameters are listed in Table I. Again, only
a half journal-bearing system is modelled because of symmetry. For the lubricant film, 30 × 4
eight-node 2D quadratic elements are used, while 30 × 4 × 2 eight-node 3D brick elements

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
1780 W. K. LIU ET AL.

Figure 10. The relationship between the minimum film thickness and the crank angle: (a) the influence
of  at us = 2000 rpm, Dwav = 5 m, Fext = 10 kN; (b) the influence of us at  = 0.0015 mm,
Dwav = 5 m, Fext = 10 kN; (c) the influence of Fext at us = 2000 rpm, Dwav = 5 m,  = 0.0005 mm;
and (d) the influence of Dwav at us = 2000 rpm,  = 0.001 mm, Fext = 10 kN.

are used to model the deformation of the housing. All displacements components in x and y
directions at the nodes on the external cylindrical surface are prescribed to be zero.
The minimum film thickness, hm , takes an instant minimum value h̄m as the external force
Fext almost reaches a peak value F̄ext , and yields a maximum, h̄M , when Fext is around
(1/4 ∼ 2/5)F̄ext at the increasing stage of the loading curve, as plotted as in Figure 10. It
shows a smaller value with the increment of F̄ext and a reduction of the waviness depth Dwav .
The minimum film thickness is slight smaller if the secondary roughness, , is larger in the
range between the (i − 1)th h̄m and the ith h̄M . Moreover, the value of h̄m increases while
h̄M decreases as the angular speed of shaft us is increased.
The orbit of the shaft center is illustrated in Figure 11. As the system solution becomes
stable, the orbit forms a single connecting domain for a smaller F̄ext (10 and 20 kN), but
becomes a double connecting one for a larger F̄ext (30 and 40 kN). The single connecting
domain spreads slightly with the increase of , and its shape becomes more round when us
and Dwav are increased.
Figure 12 shows the maximum film pressure pm , which presents a similar shape as the
external loading curve but with a slight delay. Also, pm increases with F̄ext rapidly and the
influence of  can be negligible. It can also be found that a larger minimum value of pm is
related to the higher us and larger Dwav .

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
FINITE ELEMENT METHOD FOR MIXED ELASTOHYDRODYNAMIC LUBRICATION 1781

Figure 11. The orbits of journal centre: (a) the influence of  at us = 2000 rpm, Dwav = 5 m,
Fext = 10 kN; (b) the influence of us at  = 0.0015 mm, Dwav = 5 m, Fext = 10 kN; (c) the
influence of Fext at us = 2000 rpm, Dwav = 5 m,  = 0.0005 mm; and (d) the influence of Dwav at
us = 2000 rpm,  = 0.001 mm, Fext = 10 kN.

Figure 12. The relationship between the maximum film pressure and the crank angle: (a) the influence
of  at us = 2000 rpm, Dwav = 5 m, Fext = 10 kN; (b) the influence of us at  = 0.0015 mm,
Dwav = 5m, Fext = 10 kN; (c) the influence of Fext at us = 2000 rpm, Dwav = 5m,  = 0.0005mm;
and (d) the influence of Dwav at us = 2000 rpm,  = 0.001 mm, Fext = 10 kN.

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
1782 W. K. LIU ET AL.

Figure 13. The relationship between the total contact pressure and the crank angle: (a) the influence
of  at us = 2000 rpm, Dwav = 5 m, Fext = 10 kN; (b) the influence of us at  = 0.0015 mm,
Dwav = 5 m, Fext = 10 kN; (c) the influence of Fext at us = 2000 rpm, Dwav = 5 m,  = 0.0005 mm;
and (d) the influence of Dwav at us = 2000 rpm,  = 0.001 mm, Fext = 10 kN.

The total contact pressure (the sum of all asperity contact load), wa , increases with F̄ext , 
and Dwav while decreases with us , as shown in Figure 13. It should be mentioned that wa is
zero when Dwav is very small (e.g. 0.882 m), and double peaks can be also seen for a lower
us and larger Dwav .
Because the problem is highly non-linear, the time integration of the solution converges
slowly when the fully Newton–Raphson method is used and the time increments are small,
especially for the cases with heavier load, higher waviness depth, smaller viscosity of lubricant
and lower shaft velocity. The simulations show that the modified Newton–Raphson method
works better than the full Newton–Raphson method, and the convergent solutions can be
obtained generally after 2 or 3 cycles of load (as shown in Figures 10–13). However, it is also
found that when Dwav is larger, for example, above 6 m, convergence becomes very difficult.
Therefore, new methods for faster convergence need be developed to improve the efficiency of
the finite element analysis.

4. CONCLUSIONS

A finite element model for mixed lubrication of journal-bearing systems working under
adverse conditions is developed. With large eccentricity ratios, the asperity effects on con-
tact and lubrication are considered. The elastic deformations due to both hydrodynamic and
contact pressure, as well as the cavitation of the lubricant film, are included in the model.

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
FINITE ELEMENT METHOD FOR MIXED ELASTOHYDRODYNAMIC LUBRICATION 1783

Without considering roughness effects, good agreement is obtained with the experimental re-
sults by Patrick [32] and previous work by Wang et al. [12]. Finally, the influence of waviness
depth, secondary roughness, and external force and shaft speed on the mixed lubrication were
discussed. It is found that for journal bearings under dynamic loads, the convergent solutions
can be obtained generally after only 2 or 3 cycles of loading using the modified Newton–
Raphson method. However, the convergence becomes difficult when the waviness depth Dwav
is larger than 6 m.
In engineering analysis and design, major concerns are placed on the effectiveness and
efficiency of the specific methodology. The finite element modelling of a journal bearing may
involve a large amount of computation because of the high non-linearity of the problem and the
large number of finite element nodes needed to obtain more accurate solutions. Further research
is needed to improve the computational efficiency and robustness. Meshfree methods [34, 35]
and multiscale methods [36, 37] may be good approaches to use for the in-depth evaluation of
roughness effects.

APPENDIX A: FLOW FACTORS IN AVERAGE FLOW MODEL

The flow factors in the average Reynolds equation reflect the influence of the surface roughness
on the lubricant flow. These factors can be thought of as correction factors, comparing the mean
flow between rough surfaces to that between smooth surfaces with the same nominal geometries
[28].
p p
According to Patir and Cheng’s average flow theory [9, 38], the flow factors, x , z and
 , can be obtained by comparing the average Reynolds equation with the stochastic Reynolds
s

equation. For our journal-bearing model, the average Reynolds equation can be written as
   
* p h̄
3*p̄ * p h̄ *p̄
3 u1 + u2 *(h̄) u1 − u2 *s *(h̄)
x + z = +  +
*x 12 *x *z 12 *z 2 *x 2 *x *t
(A1)

In the case of pure longitudinal roughness (see Figure 2), the flow factors are found to be [26]

p E[h3T ]
x = (A2)
h̄3
p 1
z = (A3)
h̄3 E[1/ h3T ]

s = 0 (A4)

where hT is the true film thickness and h̄ is the average film thickness. The operator, E[·],
denotes the average over a smaller area when compared to the lubricant region and includes
a large number of asperities. In our case, this area should contain at least one longitudinal
waviness.

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
1784 W. K. LIU ET AL.

p
A.1. Circumferential pressure flow factor x
Let  represent the amplitude of the secondary roughness, D the amplitude of the primary
roughness (depth of the waviness) and h̄T the film thickness measured from the mean lever of
the surface of the wavinesses (see Figure 2), we have

hT = h̄T +  , and (A5)

h3T = h̄3T + 3h̄2T  + 3h̄T 2 + 3 (A6)

Then
E[h3T ] = E[h̄3T ] + 3E[h̄2T  ] + 3E[h̄T 2 ] + E[3 ] (A7)
Because the characteristic length of the secondary roughness is far less than that of the primary
roughness, we have
E[h̄2T  ] = E[h̄2T ]E[ ] and E[h̄T 2 ] = E[h̄T ]E[2 ] (A8)
Note that
E[ ] = E 3
= 0, E[2 ] = 2 , and E[h̄T ] = h̄ (A9)
Then we get
E[h3T ] = E[h̄3T ] + 32 h̄ (A10)
As the overall surface topography is assumed to be periodic, E[h̄3T ] can be calculated over a
periodic length as
 L   L2  3  L3  3 
1 1 L1 D D
E[h̄3T ] = h̄3 dz = h3m dz + hm + z dz + hm + z dz
L 0 T L 0 L1 L2 L2 L3
 
L2 + L3 3 2 1
= h3m + hm D + hm D 2 + D 3 (A11)
L1 + L2 + L3 2 4

Since
L2 + L3 D D
h̄ = E[hT ] = E[h̄T ] = hm + = hm +  (A12)
L1 + L2 + L3 2 2
then E[h̄3T ] (Equation (A11)) can be rewritten as

E[h̄3T ] = h̄3 + h̄( − 43 2 )D 2 + 41 ( − 1)2 D 3 (A13)


where  = 1 − L1 /L.
p
The flow factor x is hence given by [26]
 2    2  3
p E[h̄3T ] + 32 h̄  3 D 1 2 D
x = =1+3 + 1−  + ( − 1) (A14)
h̄3 h̄ 4 h̄ 4 h̄

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
FINITE ELEMENT METHOD FOR MIXED ELASTOHYDRODYNAMIC LUBRICATION 1785

and the first derivative with respect to the mean film thickness is
p
      2  3 
*x 1  2 3 D 3 2 D
=− 6 + 2 1 −  + ( − 1) (A15)
*h̄ h̄ h̄ 4 h̄ 4 h̄

p
A.2. Axial pressure flow factor z
Because the flow of the lubricant is dominated by the flow in the x direction, the secondary
p
roughness can be neglected for the flow factor z as a first-order approximation, i.e.,
   
1 1
E 3
≈E (A16)
hT h̄3T

For a periodic surface topography, we have


  
1 1 L 1
E = dz
h̄3T L 0 h̄3T
 
 L1  L2  L3
1 1 1 1 
=  dz +
3 dz +
3 dz
L h3m
0 L1 hm + D
L2 z
L2 hm + D
L2 z
"   #
1 L1 1 1 1
= + − (L2 + L3 )
L 3
hm 2D 2
hm (hm + D)2

1 −   2hm + D
= + (A17)
h3m 2 h2m (hm + D)2

Since
hm  D
hm = h̄ − D/2, and =1− (A18)
h̄ 2 h̄
then
 
  D
2 1− +D
1 1−  2h̄
h̄3 E = 3 +  2  2 =   (A19)
h̄3T D D D D D 2
1− 2 1− 1− + 1−
2h̄ 2h̄ 2h̄ h̄ 2h̄
where

 2 + (1 − ) h̄
D
1−
= D
+
 2
1− 2h̄
2
1 + 1 − 2 D

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
1786 W. K. LIU ET AL.

p
The flow factor z is derived as [26]
 
p 1 1 D 2
z =  = 1− (A20)
1 2h̄
3
h̄ E 3
h̄T

and the first derivative with respect to the mean film thickness is
 
 
 D D
   − 3 − (1 − ) 1 − 1−
D  
p
*z D 1 (1 − ) 2 h̄ 2h̄ 
= 2 1−  +  −   
*h̄ h̄ 2h̄  2 D
 D 3 
2 1 − 2
2 1 + 1 −
2h̄ 2 h̄
(A21)

APPENDIX B: ASPERITY CONTACT LOAD

As mentioned before, the solid contact between the journal and housing is usually minimal,
and the contact area is small compared to the circumferential dimension of the journal surface.
Thus, the contact between the journal and housing can be approximated by the solution of the
half-space contact.
The deformation of a half-space relative to a point A which is at a depth l below the center
of the Hertzian contact pressure distribution (see Figure B1) is given by Johnson [39] as
 
p 2l
= 2 ln −
¯ (B1)
Ē a

where is the displacement at the center of the Hertzian contact pressure, a the half contact
width, and p the contact load per unit length (along the direction perpendicular to the paper).

Figure B1. A half space under a contact pressure, p.

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
FINITE ELEMENT METHOD FOR MIXED ELASTOHYDRODYNAMIC LUBRICATION 1787

Figure B2. Deformation of the journal and the housing due to contact.

The Young’s modulus E and Poisson’s ratio


are given by
 
 E 

  for plane strain


Ē = 1 −

¯ = 1 −
(B2)

 

E for plane stress


Because the contact of the journal and housing is actually between the ridge of the wavinesses
in the journal and the housing surface, the compression is assumed to be in the state of plane
stress [26]. The contact load causes the journal surface to deform inwards and the housing
surface to deform outwards (see Figure B2). The displacements at the center of the contact
pressure are given by
 
p 2R1
1 = 2 ln −
1 , and (B3)
E1 a
 
p 2R2
2 = 2 ln −
2 (B4)
E2 a
where E1 ,
1 , E2 and
2 are material parameters for the journal and housing, and R1 and
R2 are the distances from the center of the contact pressure to the centers of the journal and
housing respectively. The total overlap between the journal and the housing is given by
  
p 2 2R1 2 2R2
1
2
ov = 1 + 2 = ln + ln − + (B5)
 E1 a E2 a E1 E2
Since R1 ≈ R2 ≈ R, then
      
p 1 1 2R
1
2p 2R
ov = 2 + ln − + 2 = 2 ln −
(B6)
 E1 E2 a E1 E2 E a
where
   
1 1 1 1 1 1 1 1
= + , and = + (B7)
E 2 E1 E2
2
1
2
From the Hertzian’s contact theory, the half width of the contact area is given by
$  
4pR1 R2 1 1
a= + (B8)
(R2 − R1 ) Ē1 Ē2

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
1788 W. K. LIU ET AL.

Figure B3. Compression of the asperity ridge due to contact.

For our case of the asperity contact, we have


$   $
4pR1 R2 1 1 2p
a= + ≈ 2R (B9)
(R2 − R1 ) E1 E2 E(R2 − R1 )

Hence
 
2p E(R2 − R1 )
ov = ln −
(B10)
E 2p
Considering the distortion of the housing (see Figure 6), we have
D
e0 = R2 − R1 = c + d −  , and (B11)
2
ov = R1 − R2 + e = e − e0 (B12)

Therefore, for an individual contact ridge, the contact pressure is determined by


 
2p Ee0
ov = ln −
(B13)
E 2p
The total asperity contact load is then given by
 B
wa = p dz (B14)
0

where B is the length of the bearing housing. Assuming that the contact pressure is uniformly
distributed in the axial direction on each ridge of the waviness, we have


N
wa = L(1 − ) pn (B15)
n=1

where L is the periodic length of the waviness. Because the deformation of the housing due
to the contact is very small, the compression of the asperity ridge is almost equal to the total
overlap (see Figure B3). Because of the compression of the asperity ridge, the width of the
plateau increases from L1 to L∗ by
ov ov ov
L∗1 − L1 ≈ L2 + L3 = (L − L1 ) (B16)
D D D

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
FINITE ELEMENT METHOD FOR MIXED ELASTOHYDRODYNAMIC LUBRICATION 1789

Thus, the total contact load can be written as



N
wa = L (1 − ∗ )pn (B17)
n=1

L∗1 ov
where ∗ = 1 − =− .
L D

ACKNOWLEDGEMENTS
The authors would like to acknowledge NSF, NSF-IGERT, Federal Mogul Corporation, and CSET for
financial support, and also wish to thank Dr Chao Zhang, Tongji University in China, for his kind
help in testing a part of the computation programs.

REFERENCES
1. Peppler W. Untersuchungen über die Druckükertragung bei belasteten und geschmierter umlaufenden a
chsparallelen zylindern. Maschinenelemente-Tagung Aachen, 1935; 42; V.D.I. Verlag: Berlin, 1936.
2. Fuller DD. A survey of journal bearing literature. Am. Soc. Lub. Engrs., 1958.
3. Dowson D, Higginson GR. Elastohydrodynamic Lubrication. Pergaman Press: Oxford, 1966.
4. Rohde SM. Thick film and transient elastohydrodynamic lubrication problems. In Fundamentals of Tribology,
Proceedings of the International Conference on the Fundamentals of Tribology, Suh NP, Saka N (eds). The
MIT Press: Cambridge, MA. 1978.
5. Woodward W, Paul B. Contact stress for closely conforming bodies—application to cylinders and spheres.
DOT-TST-77-8.
6. Venner CH, Lubrecht AA. Multilevel Methods in Lubrication. Elsevier Science: Amsterdam, 2000.
7. Booker JF, Huebner KH. Application of finite element methods to lubrication: an engineering approach.
Journal of Lubrication Technology (ASME) 1972; 94:313–323.
8. Christensen H. A theory of mixed lubrication. Proceedings of IME 1972; 186:41–72.
9. Patir N, Cheng HS. An average flow model for determining effects of three-dimensional roughness on partial
hydrodynamic lubrication. Journal of Lubrication Technology (ASME) 1978; 100:12–17.
10. Greenwood JH, Tripp JH. The contact of two nominally flat rough surfaces. Proceedings of IME 1971;
185:625–633.
11. Zhu D, Cheng HS, Arai T, Hamai K. A numerical analysis for piston skirt in mixed lubrication—part II:
deformation considerations. Journal of Tribology (ASME) 1993; 115:115–125.
12. Wang Q, Shi F, Lee SC. A mixed-lubrication study of journal bearing conformal contacts. Journal of
Tribology (ASME) 1997; 119:456–461.
13. Goglia PR, Cusano C, Conry TF. The effects of surface irregularities on the elastohydrodynamic lubrication
of sliding line contacts: part I—single irregularities; part II—wavy surfaces. Journal of Tribology (ASME)
1984; 106:104–119.
14. Jiang X, Cheng HS, Hua DY. A theoretical analysis of mixed lubrication by macro micro approach:
part I—results in a gear surface contact. Tribology Transactions 2000; 43:689–699.
15. Winer WO. Temperature effects in elastohydrodynamically lubricated contacts. In New Directions in
Lubrication, Materials, Wear and Surface Interactions, Tribology in the 80’s, Loomis WR (ed). Noyes
Publications: New Jersey, 1985.
16. Pinkus O. Thermal Aspects of Fluid Film Tribology. ASME Press: New York, 1990.
17. Shi F, Wang Q. A mixed-TEHD model for journal-bearing conformal contacts—part 1: model formulation and
approximation of heat transfer considering asperity contact. Journal of Tribology (ASME) 1998; 120:198–205.
18. DuBois GB, Ocvirk F. Analytical derivation and experimental evaluation of short-bearing approximation for
full journal bearings. NACA Report 1157, 1953.
19. Booker JF. Dynamically loaded journal bearings: mobility method of solution. Journal of Lubrication
Technology 1965; 87:537–546.

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790
1790 W. K. LIU ET AL.

20. Ritchie GS. The prediction of journal loci in dynamically loaded internal combustion engine bearings. Wear
1975; 35:291–297.
21. Campbell J, Love PP, Martin FA, Rafique SO. Bearings for reciprocating machinery: a review of the present
state of theoretical, experimental and service knowledge. Proceedings of IME vol. 182, part 3A, 1967–1968.
22. Martin FA. Developments in engine bearings. 9th Leeds–Lyon Symposium on Tribology, Tribology of
Reciprocating Engines, Sept. 1982; Tribology International 1983; 16:147–164.
23. Goenka PK. Dynamically loaded journal bearings: finite element analysis. Journal of Tribology (ASME)
1984; 106:429–439.
24. Fantino B, et al. Dynamic behavior of an elastic connecting-rod-bearing-theory study. SAE International
Congress, Detroit, February 1983, Paper 830307, SAE/SP-359: studies of engine bearings and lubrication,
1983; 23–32.
25. Oh KP. The numerical solution of dynamically loaded elastohydrodynamic contact as a nonlinear
complementarity problem. Journal of Tribology (ASME) 1984; 106:88–95.
26. Ai X-A, Cheng HS. A finite element analysis of dynamically loaded journal bearings in mixed lubrication.
Report to Engine and Powertrain Research Laboratory, Nissan Research Center, Nissan Motor Co., Ltd,
March 1995.
27. Hu Y-K, Liu WK. An ALE hydrodynamic lubrication finite element method with application to strip rolling.
International Journal for Numerical Methods in Engineering 1993; 36:885–880.
28. Liu WK, Hu Y-K. Finite element hydrodynamic friction model for metal forming. International Journal for
Numerical Methods in Engineering 1994; 37:4015–4037.
29. Liu WK, Hu Y-K, Belytschko T. ALE finite elements with hydrodynamic lubrication for metal forming.
Nuclear Engineering and Design 1992; 130:1–126.
30. Murty KG. Note on a bard type scheme for solving the complementarity problem. Opsearch 1974; 11:
123–130.
31. Belytschko T, Liu WK, Moran B. Nonlinear Finite Elements for Continua and Structures. Wiley; New York,
2000.
32. Patrick JK. An experimental investigation into the performance of sleeve bearings subjected to a range of
alternating loads. Proceedings of the Institution of Mechanical Engineers 1967–68; 182:77–87.
33. Dyson A. A failure of EHD lubrication of circumferentially ground disk. Proceedings of IME 1976; 190:52–76.
34. Li S, Liu WK. Meshfree and Particle Methods and their Applications. Applied Mechanics Review 2002;
55:1–34.
35. Liu WK, Jun S, Zhang YF. Reproducing kernel particle methods. International Journal for Numerical Methods
in Fluids 1995; 20:1081–1106.
36. Liu WK, Zhang YF, Ramirez MR. Multi-Scale Finite Element Methods. International Journal for Numerical
Methods in Engineering 1991; 32:969–990.
37. Liu WK, Chen Y. Wavelet and multiple scale reproducing kernel method, International Journal for Numerical
Methods in Fluids 1995; 21:901–933.
38. Patrir N, Cheng HS. An application of average flow model to lubrication between rough sliding surfaces.
ASME Journal of Lubrication Technology 1979; 101:220–230.
39. Johnson KL. Contact Mechanics. Cambridge University Press: Cambridge, New York, 1985.

Copyright 䉷 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 60:1759–1790

You might also like