Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

19

Lubrication regime transitions at the piston


ring cylinder liner interface
N W Bolander1 , B D Steenwyk1, F Sadeghi1, and G R Gerber2
1
School of Mechanical Engineering, Purdue University, West Lafayette, IN, USA
2
Caterpillar Inc., Lafayatte, IN, USA
The manuscript was received on 11 March 2004 and was accepted after revision for publication on 24 September 2004.
DOI: 10.1243/135065005X9664

Abstract: An experimental apparatus and an analytical model have been developed to investigate and determine the lubrication condition and frictional losses at the interface between a
piston ring and cylinder liner. In order to obtain a solution for the lubrication condition between
the piston ring and cylinder liner, the system of Reynolds and film thickness equations subject to
boundary conditions were simultaneously solved. The effects of boundary and mixed lubrication conditions were implemented using the Greenwood Tripp stochastic approach. The
Elrod cavitation algorithm was used to investigate the effects of fluid rupture and reformation
at the top and bottom dead centres. The experimental results indicate that the piston ring
and liner experience all the different lubrication regimes (i.e. boundary, mixed, and hydrodynamic lubrication) during a stroke. A comparison between experimental and analytical
results indicated that they are in good agreement and the analytical model developed for this
study can capture the different lubrication regimes that the piston ring and liner experience.
Keywords:

piston ring, mixed lubrication, friction, lubrication

INTRODUCTION

Modern reciprocating engines are expected to conform to strict efficiency standards. A key factor in
achieving these standards is the minimization of
parasitic losses due to friction. The piston ring
assembly is one of the main sources of friction in
an internal combustion engine, which by some
estimates can account for 20 40 per cent of engine
frictional losses [1]. A thorough understanding of
the lubrication condition at the piston ring cylinder
liner interface is vital in determining the sources of
frictional loss. It is well known that the piston ring
encounters the entire range of lubrication regimes
through each stroke (i.e. boundary lubrication,
mixed lubrication, elastohydrodynamic lubrication
(EHL), and hydrodynamic lubrication). Thus, an
investigation into frictional losses at the piston
ring cylinder liner contact must take into account

Corresponding author: School of Mechanical Engineering,

Purdue University, West Lafayette, IN 47907-1288, USA.

J01504 # IMechE 2005

the transitions from a state of full-film lubrication


to boundary lubrication.
An early example of modelling and analysis of the
frictional losses at the piston ring cylinder liner
contact is attributed to Rohde et al. [2]. This mixedfriction model was based on the averaged flow
Reynolds equation with the half-Sommerfeld boundary condition obtained by Patir and Cheng [3].
Asperity interactions were included through the
Greenwood Tripp [4] model. Dowson et al. [5]
used a hydrodynamic lubrication model to study
film thickness, lubricant transport, and viscous friction at the piston ring cylinder liner (PRCL) contact.
They used the Reynolds boundary condition in their
analysis. Later investigations showed strong EHL
effects near top dead centre (TDC). Jeng [1] discussed
the fact that the Sommerfeld-type boundary conditions used in many of the previous investigations
are deficient because of the violation of continuity
at the cavitation boundary. He developed a onedimensional model using the Reynolds boundary
condition combined with a system of two nonlinear differential equations. This method improved
on the work of Dowson et al. [5] by relaxing the
Proc. IMechE Vol. 219 Part J: J. Engineering Tribology

20

N W Bolander, B D Steenwyk, F Sadeghi, and G R Gerber

assumption of a constant flowrate across the ring.


The model was used to study film thickness, friction,
and starvation conditions for piston ring lubrication
through the complete engine cycle. Hu et al. [6] applied a non-axisymmetrical quasi-two-dimensional
analysis to the PRCL contact. This model used the
average flow Reynolds equation and the Reynolds
boundary conditions and employed the Greenwood
Tripp [4] model for asperity contact to solve the
mixed lubrication problem. Flow in the circumferential direction was not permitted in order to
simplify the solution; however, a linear complimentary problem was performed in this direction to
determine the deflection of the ring. Yang and
Keith [7] indicated that the pressure reformation
cannot be accurately predicted by the Reynolds
boundary condition, especially when high pressure
exists at the trailing edge of the ring. Proper determination of the reformation boundary is critical
to accurately determining the load capacity of the
piston ring. For this purpose a previously developed
cavitation algorithm [8] was modified to include the
EHL effects that Dowson et al. [9] had found so
important in the region near TDC. Later, their
model was extended to include flow in the circumferential direction [10, 11]. These models assumed
smooth surfaces and were therefore unable to determine realistically the frictional losses in regions
where mixed lubrication is dominant. Sawicki and
Yu [12] applied the Jakobsson Floberg Olsson cavitation theory to piston ring lubrication in a onedimensional model similar to that of Jeng [1]. They
note that Jengs model is able to provide a good estimation of the film thickness using the Reynolds
boundary condition but will probably underestimate
friction and power loss and overestimate flowrates.
Akalin and Newaz [13] developed a one-dimensional
analytical model based on the average flow Reynolds
equation obtained by Patir and Cheng. Mixed
lubrication is enabled through the use of the
Greenwood Tripp asperity contact model in a
manner similar to that of Hu et al. [6]. Reynolds
(Swift Stieber) boundary conditions were used and
results from this model were corroborated to experimental results [14]. However, Priest et al. [15]
suggested that the lack of detailed experimental
data in the area of piston ring lubrication requires
that future progress be based upon combined
theoretical and experimental investigations.
Ting [16, 17] developed a reciprocating test rig to
measure the friction coefficient between rings and
liners taken from actual engines. Friction results
were obtained using a piezoelectric-type load cell.
The results were presented in a Stribeck-type relationship. Dearlove and Cheng [18] developed a reciprocating test rig based on an actual single-cylinder engine. A
floating liner segment was instrumented to provide
Proc. IMechE Vol. 219 Part J: J. Engineering Tribology

frictional data. The lubricant film thickness at midstroke was measured using the laser fluorescence
technique. Arcoumanis et al. [19] measured film thickness throughout the stroke using a purpose-made
capacitance transducer. Their reciprocating test rig
had a maximum stroke length of 50 mm. Frictional
data were obtained using a deflection-based measurement. Akalin and Newaz [14] employed a deflectionbased force approach to measure friction using
actual engine rings and a cylinder liner.
In this study a model was developed to solve for
the lubrication condition and frictional losses at the
piston ring cylinder liner interface through a complete engine cycle. The two-dimensional cavitationenabled Reynolds equation is coupled with the
stochastic asperity contact model of Greenwood
and Tripp [4] to allow calculation of lubricant film
thickness and friction for a piston ring operating in
the mixed lubrication regime. A test rig has also
been developed to correlate with the numerical
model and to explore further the lubrication condition at the piston ringcylinder liner interface. Actual
piston rings and cylinder liner segments were used
in the experimental test rig. A piezoelectric-type
force transducer was employed for fast frequency
response to capture the frictional phenomena at
the TDC and bottom dead centre (BDC).

2.1

TEST RIG DESCRIPTION AND


EXPERIMENTAL PROCEDURE
Piston ring reciprocating liner test rig

A computer-aided design (CAD) drawing of the


piston ring reciprocating liner test rig designed,
developed, and constructed for this study is illustrated in Fig. 1. Figure 2 depicts a close-up view of

Fig. 1

CAD drawing of the piston ring reciprocating


liner test rig

J01504 # IMechE 2005

Lubrication regime transitions at the piston ringcylinder liner interface

21

generated on the piston ring. The piezoelectric


force transducer was chosen to maximize rigidity
and natural frequency, thus allowing high sampling
rates and minimum ring motion. The frequency
response of this piezoelectric force sensor allows
the frictional behaviour to be resolved near the
ends of stroke where the frictional force switches
direction nearly instantaneously. A Dell workstation
with a National Instruments data acquisition board
and software is used to collect the sensor data and
to control the charge amplifier. Data collected
include the friction, normal, and tangential loads,
as well as the rotational speed of the motor.
2.2
Fig. 2 Completed bench-scale piston ring reciprocating
liner test rig

the completed reciprocating carriage assembly. The


test rig was designed to have the cylinder liner reciprocate while keeping the piston ring stationary
and can accommodate a wide range of loads,
speeds, and lubricant conditions at the PRCL interface. The apparatus can accommodate cylinder
bore diameters of 51 140 mm (2.0 5.5 in) with a
stroke in the range 38 152 mm (1.5 6 in). A 2.2 kW
(3 hp) variable-speed d.c. motor is used to turn the
crank between 15 and 300 r/min, which is measured
by a magnetic pick-up. The output shaft of the
1750 r/min motor is connected to a 14:3 toothed
belt speed reduction drive system. The crank is connected to a Thompson linear carriage, which houses
a 608 section of the 137.2 mm (5.402 in) bore cylinder
liner. The piston ring is held stationary under the
pivot arm above the reciprocating liner segment in
a specially designed ring holder (Fig. 3). Load is
applied to the PRCL interface through dead weights
at the end of the arm. A piezoelectric Kistler threeaxis force transducer and charge amplifier are used
to measure the normal, tangential, and side loads

Experimental procedure

For each test the following procedure is followed.


First, two rows of 15 drops of SAE 30 oil (1.0 ml) are
evenly distributed along the entire stroke of
66.7 mm (2 58 in) on the axis of the liner segment.
The ring is raised above the liner to a repeatable
and stable position by lifting the loaded arm up
against a hard stop. At this point the data acquisition
begins and the charge amplifier is switched to operate mode which defines the zero frictional force.
Immediately, the ring is lowered on to the liner
near mid-stroke (crank position of 2708), followed
by starting the motor. Friction and speed are
recorded for 16 cycles. The sampling frequency of
the friction measurement is set such that 1000 data
points are collected per crank revolution.
In order to study the effects of load and speed on
friction and the lubrication at the PRCL interface,
the load and crank rotational speed were varied
from 1 to 8 kgf and from 30 to 300 r/min respectively.
Table 1 contains the load and speed conditions
tested. In all cases, the same amount, type, and
arrangement of oil is used. The stroke was held
constant at 66.7 mm (2 58 in). Speeds were chosen to
illustrate the various lubrication regimes.
Table 2 contains the resulting pressure between
the piston ring and liner. This nominal contact pressure can be thought of as the sum of the ring elastic
pressure and cylinder gas pressure acting on the back
side of the ring. A typical installation pressure for
the original equipment manufacturer (OEM) ring
used in this investigation is 0.24 MPa, a typical
peak engine compression pressure is 1.0 Mpa, and
a typical peak combustion pressure is 2.0 4.0 MPa
for gasoline engines.
The ring and liner surfaces were characterized using
an optical surface profilometer. Table 3 contains the
Table 1 Load and speed conditions

Fig. 3

Ring holder and force sensor assembly

J01504 # IMechE 2005

Load (kgf)
Speed (r/min)

1, 2, 3, 4, 6, 8
30, 60, 90, 120, 180, 240, 300

Proc. IMechE Vol. 219 Part J: J. Engineering Tribology

22

N W Bolander, B D Steenwyk, F Sadeghi, and G R Gerber

Table

The function f and cavitation index F are defined by

2 Load and nominal pressure


conditions over the ring

Normal force (kgf)

Nominal pressure (MPa)

1
2
3
4
6
8

0.14
0.29
0.43
0.58
0.87
1.15

p  pc
Ff
in the full-film region
pa  pc
r
1 (1  F)f
in the cavitation zone
rc

1
for f 5 0
F(x, y)
0
for f , 0

(2)
(3)
(4)

The dimensionless boundary conditions are given by


Table 3 Surface roughnesses of the piston ring and
cylinder liner

Piston ring
Cylinder liner

Ra(mm)

Rq(mm)

Rsk

0.55
0.86

0.90
1.10

23.04
20.55

arithmetic mean roughness Ra, r.m.s roughness Rq,


and skewness Rsk for the ring and liner.

NUMERICAL MODEL DESCRIPTION

In this study, a model was developed to investigate


the lubrication condition between a piston ring
and cylinder liner. It is assumed that the contact
operates under fully flooded conditions throughout
the stroke and the mode of cavitation is closed
form [12]. The following assumptions were made
in the derivation of the fluid momentum (Reynolds)
equation.
1.
2.
3.
4.

The lubricant is Newtonian.


The pressure is constant across the film.
The flow is laminar and viscous dominant.
The body forces and inertial effects are
neglected.
5. The lubricant viscosity is constant.
6. The lubricant cavitates when the film pressure
falls below the lubricant vapour pressure.

3.1

( pL pc
Ff

pa pc
pT pc
pa pc

@(F f) 

@Y Y 1/2

at the leading edge of the ring


at the trailing edge of the ring
@(F f) 

0

@Y Y 1/2

(5)
(6)

The first boundary condition sets the pressure at the


leading and trailing edges of the ring equal to the local
ambient conditions in the cylinder or ring gap. The
second boundary condition enforces the axisymmetry
of the piston ring section. Although the model presented here assumes that the piston ring is operating
under axisymmetry conditions, the underlying discretization of the governing equations remains two
dimensional. Thus, extension and inclusion of nonaxisymmetric effects and/or circumferential flow can
easily be accommodated.

3.2

Film thickness equation

Figure 4 illustrates the geometry of the piston ring


and the cylinder wall lubricated contact as modelled
in this study. The lubricant film thickness for a
piston ring with a symmetric parabolic face can be
expressed in terms of the minimum film thickness
hmin(t) as
h(x, y) hmin (t)

d
x2
(b/2)2

(7)

Reynolds equation

The non-dimensional isothermal two-dimensional


time-dependent Reynolds equation including the
cavitation algorithm is given by [20]




@
@(F f)
@
@(F f)
H3
H3

@X
@X
@Y
@Y
@
{1 (1  F)fH}
g
@X
@
{(1 (1  F)fH}
s
@T
Proc. IMechE Vol. 219 Part J: J. Engineering Tribology

Fig. 4

(1)

The physical geometry of the axisymmetric ring


and cylinder liner section used in the numerical
model

J01504 # IMechE 2005

Lubrication regime transitions at the piston ringcylinder liner interface

where d is the crown height and b is the width of


the ring. For the OEM piston rings used in this
study, d 13 mm and b 3.79 mm.

asperity coefficient of friction ma. The resulting


expression for total frictional force on the segment is
a=2 b=2 
Ff

3.3

23

Force balance equation

a=2 b=2



h @p hU


ma pa dx dy
2 @x
h
(10)

At any instant the forces acting on the ring can be


broken down into four components:
(a) hydrodynamic pressure ph acting on the ring face;
(b) asperity contact pressure ps acting on the ring
face;
(c) ring elastic pressure pel for the segment;
(d) gas pressure pg acting on the back of the ring.
The velocity of the piston ring varies constantly
throughout the stroke, reaching zero momentarily
at TDC and BDC. The time-dependent form of the
Reynolds equation used in this study includes the
squeeze-film effect to handle the conditions at
the TDC and BDC, as well as accurately capturing
the cavitation rupture and reformation boundaries.
As described in reference [12] the pressure contribution from the reformed lubricant film and the
trailing-edge pressure may contribute substantially
to the overall load support.
The asperity contact pressure for the ring segment
can be calculated using the GreenwoodTripp [4]
asperity contact model. The form of the contact
model used in this study is based on a simplifying
curve fit employed by Hu et al. [6] and Akalin and
Newaz [13]. The same parameters used by these investigators have been maintained in the current work.
The ring elastic pressure can be calculated as
pel

2Tr
bB

3.5

Solution method

Integration of the modified non-dimensional Reynolds


equation (1) over a control volume [21] yields
e n
w

(11)
The discretized form of equation (11) takes on the
form









3 @(F f) 
3 @(F f) 
3 @(F f) 
H
DY  H
DY H
@X e
@X w
@Y n



3 @(F f) 
 DX  H
DX
@Y s
g{1 (1  F)fH}je DY

 g{1 (1  F)fH} DY
w

sDXDY ({1 (1  F)fH}n


 {1 (1  F)fH}n1 )

DT

(8)

where Tr is the ring tension and B is the cylinder bore


diameter [1]. The pressure acting on the back of
the ring is assumed to be the larger of pL and pT.
The force balance equation can then be expressed
as a summation of the forces acting on the ring
segment according to
a/2 b/2





@
@(F f)
@
@(F f)
H3
H3

dY dX
@X
@Y
@Y
s @X

e n
@

g {1 (1  F)fH} dY dX
@X
w s

e n
@

s {1 (1  F)fH} dY dX
@T
w s

(12)

The discretization shown in equation (12) results in


a system of linear equations that are solved using the
GaussSiedel relaxation scheme. In order to ensure a
stable convergence, both F and f are relaxed after
every GaussSeidel sweep according to

frelaxed af fnew (1  af )fold


Frelaxed aF Fnew (1  aF )Fold

(13)

(pel pg ) dx dy
a/2 b/2
a/2

where

b/2
(ph ps ) dx dy 0

3.4

Frictional force equations

The frictional force acting between the piston ring


segment and the cylinder wall consists of a viscous
shear force in the lubricant film and friction between the asperities in contact. An experimentally
measured value of 0.14 is used for the dynamic
J01504 # IMechE 2005

(9)

a/2 b/2

Fnew

1
0

if frelaxed 5 0
if frelaxed , 0

According to Payvar and Salant [20] the key to


stability is in the proper selection of the relaxation
coefficients. Following their recommendation, relaxation coefficients of af 0.2 and aF 0.01 were
chosen. A mesh size of 100  50 was found to provide
a grid-independent pressure solution. Iteration
Proc. IMechE Vol. 219 Part J: J. Engineering Tribology

24

N W Bolander, B D Steenwyk, F Sadeghi, and G R Gerber

continues through the grid points until the solution


reaches convergence, defined as


f  fold 
 , 105
max  new

f
new

(15)

i,j

Since the piston ring lubrication condition is


highly transient, the force balance equation (9)
must also be satisfied at each time step according to
a=2 b=2
(pel pg ) dx dy
a=2 b=2

a=2 b=2
(ph pa ) dx dy 4 0:001

(16)

a=2 b=2

pel) in equation (9) were set to the constant applied


normal load. Measured surface roughness values
were used in the asperity contact model. Experiments were conducted at room temperature using
SAE 30 oil. In the numerical modelling a viscosity
of 0.20 Pa s was chosen for all the numerical analysis.
Figure 5 shows a typical result for the measured and
predicted values of coefficient of friction Cf as the
piston ring travels from TDC (08) to BDC (1808) and
back again to TDC (3608). TDC and BDC refer to
the crank position and not to the position of the
ring on the liner. The ring position is opposite from
that of an engine because here it is the liner that is
moving and not the ring. The sign of Cf indicates
the direction of travel. In general, the measured
and predicted values correlate well throughout the
range of operating conditions tested.

RESULTS AND DISCUSSION

The experimental test rig was designed and developed to operate under ambient conditions and to
maintain a constant load throughout the stroke.
Note that in a real engine the combustion chamber
pressure would lead to much higher loading of the
piston ring at TDC than near BDC, with correspondingly greater frictional loss. In order to corroborate
the experimental and numerical model results the
leading- and trailing-edge pressures in equation (5)
were set to the ambient pressure and the dynamic
components of the force balance equation ( pg and

Fig. 5

4.1

Lubrication regime transition

For a full rotation of the crank, the piston ring


transitions through different regimes of lubrication.
Figure 6 illustrates the transitions over an expansion
stroke (0 1808). At the beginning of the stroke
(08, TDC) the contact is dominated by asperity
asperity interaction. As evidenced by the large spike
in Cf, this portion of the stroke is in the boundary
lubrication region, which is defined here as pure
asperity-to-asperity solid contact. Note that the
maximum value of the friction spike does not exceed

Friction results exhibiting all three lubrication regimes (120 r/min; 3 kgf)

Proc. IMechE Vol. 219 Part J: J. Engineering Tribology

J01504 # IMechE 2005

Lubrication regime transitions at the piston ringcylinder liner interface

Fig. 6

25

Different lubrication regimes encountered during an expansion stroke at 120 r/min


and 3 kgf

the experimentally measured dry coefficient of friction. As the piston ring accelerates away from TDC,
a lubricant film separating the two contacting
bodies begins to develop. Here asperity asperity
interaction remains significant; however, the load is
supported partly by the asperity-to-asperity solid
contact and partly by the lubricant present in the
contact. This portion of the stroke is considered to
be the mixed lubrication regime. As the velocity
increases, the load balance shifts further towards
the lubricant as the surfaces separate and asperity
interaction decreases. As illustrated in Fig. 7a the
frictional force at the contact is due to asperity contact and viscous losses. The frictional force due to
asperity interaction (when present) is generally
much larger than the viscous losses, resulting in a
point of minimum Cf as the ring makes a transition
from mixed lubrication to full-film lubrication.
Through the full-film region the lubricant film is
thick enough to prevent asperity interaction; thus
all frictional losses during this section of the stroke
are due to viscous drag. Since viscous shear is proportional to the relative velocity of the ring a local
maximum in the friction curve is observed near
mid-stroke. The presence of this viscous hump in
the friction trace is evidence that the ring has
reached a state of full-film lubrication. From near
mid-stroke to BDC the opposite chain of events transpires. As the ring decelerates, the hydrodynamic
J01504 # IMechE 2005

action of the ring is decreased, which in turn leads


to smaller film thickness. The velocity continues to
decrease until it is no longer sufficient to build a
thick film to separate the surfaces completely, and
the ring returns to a state of mixed lubrication.
Asperity interaction and frictional losses continue
to increase until the ring reaches BDC. Note that
the friction trace is not perfectly symmetric due to
variation in acceleration and deceleration near TDC
and BDC.
Figures 7a and b illustrate the relationship
between the onset of asperity contact at the ends of
stroke and the minimum film thickness. While the
minimum film thickness remains greater than the
effective surface roughness sc, the surfaces are completely separated and the only contribution to the
coefficient of friction is from the viscous shear.
When the minimum film thickness falls below sc,
the asperities begin to come into contact.
Figure 8 depicts the variation in pressure over the
width of the ring section (20.5 , X , 0.5 and
Y 0). Beginning at TDC (08) the load is almost
entirely supported by asperity contact pressure symmetric about the centre-line. Note that, since the
asperity contact occurs over a relatively small contact area, the pressures must be large in order to support the load. These high asperity contact pressures
are the cause of the friction spikes at the ends of
stroke as is shown in Fig. 7. As the ring accelerates
Proc. IMechE Vol. 219 Part J: J. Engineering Tribology

26

Fig. 7

N W Bolander, B D Steenwyk, F Sadeghi, and G R Gerber

(a) Relative contributions of the asperity contact and viscous shear components to the total coefficient of
friction (predicted) (120 r/min; 3 kgf ); (b) relationship between the onset of asperity contact, minimum film
thickness, and effective surface roughness sc

away from the TDC position, hydrodynamic pressure


begins to build in the converging section of the ring
face. The transition of the load support from asperity
contact to hydrodynamic pressures is seen as the

peak pressure reduces, spreads out, and shifts away


from the centre-line to the converging section of
the ring face. The ring again decelerates as it
approaches BDC (1808). At these lower velocities

Fig. 8 Pressure acting over the width of the ring face as a function of crank angle

Proc. IMechE Vol. 219 Part J: J. Engineering Tribology

J01504 # IMechE 2005

Lubrication regime transitions at the piston ringcylinder liner interface

the hydrodynamic action is not sufficient to sustain


the thick film; thus the lubricant begins to squeeze
out of the contact as the surfaces move closer
together. Pressure generated through this squeezing
motion shifts the profile towards the centre-line.
Note that the squeeze pressure is independent of
the asperity contact pressure that begins to increase
as the lubricant film continues to drop.
4.2

Effect of speed

Figure 9 presents the effect of operating speed


on the coefficient of friction under a constant load.

Fig. 9

27

Hydrodynamic action in the converging section of


the ring is responsible for generating the loadsupporting pressure in the lubricant film. An
increase in the sliding speed enhances the wedging
action of the converging ring profile, providing a
better load capacity with a larger lubricant film thickness. Hence, the increased surface separation
decreases the amount of asperity asperity contact,
which is manifest as a lowering of the friction spike
at the ends of stroke. In the extreme, the lubricant
film is sufficiently thick to maintain the surfaces in
the full-film regime through the entirety of the
stroke. Since the viscous shear is proportional to

Effect of speed on coefficient of friction: (a) measured using piston ring reciprocating liner test rig;
(b) predicted from analysis

J01504 # IMechE 2005

Proc. IMechE Vol. 219 Part J: J. Engineering Tribology

28

N W Bolander, B D Steenwyk, F Sadeghi, and G R Gerber

velocity, a more pronounced rise in the mid-stroke


hump is seen with increased speed.
Figure 10 shows the effect of speed on the minimum film thickness under a constant load as
predicted by the numerical model. The effective surface roughness sc is used in this figure to denote the
onset of asperity interaction. As noted previously, at
low speeds the hydrodynamic action of the converging ring profile is not as strong. Accordingly, in
the 60 r/min case the piston ring remains in mixed
lubrication throughout the entire cycle. However, as
the sliding speed increases, the minimum lubricant
film thickness also rises, until in the 300 r/min case,
only very light asperity contact occurs near the ends
of stroke. Note that the minimum film thickness
profile is not symmetric. The point of absolute minimum is shifted a few degrees from the TDC and BDC
due to the squeeze-film effect. This effect is also seen
in the friction traces of Fig. 9, causing an asymmetry
in the friction spikes at the ends of stroke. The point
of maximum solidsolid friction corresponds to the
point of absolute minimum film thickness along the
stroke.
4.3

Effect of load

Figure 11 illustrates the effect of load on the coefficient of friction at a constant speed (120 r/min).
Figure 12 shows the corresponding predicted minimum film thickness. As expected, increasing the

Fig. 10

load decreases the lubricant film thickness. Consequently, the regions where asperity interaction is significant are increased. This is easily seen in the
highest loading case (8 kgf ) where the piston ring
remains in the mixed-lubrication regime throughout
the cycle. Not only are the friction spikes at the ends
of stroke much wider than for the lower loading
cases, but the friction remains high through the
mid-stroke region as well. Recall that, if the ring
makes a transition from mixed lubrication to fullfilm lubrication, a viscous hump will be produced.
It is also noted that the minimum film thickness
remains below sc through virtually the entire
stroke, again indicating the mixed-lubrication condition through the entire cycle. Contrast this with
the lightest load case of 2 kgf. In this case, the hydrodynamic pressure generated in the converging
section of the ring profile is able to produce a
much thicker lubricant film. This results in smaller
narrower friction spikes near the ends of stroke.

CONCLUSIONS

A numerical model and experimental test apparatus


have been developed to investigate the lubrication
and friction condition at the PRCL interface. Good
correlation was found between the experimental
and numerical results. The results presented here
encompass the entire range of lubrication regimes

Effect of speed on minimum film thickness under a constant 3 kgf load (predicted)

Proc. IMechE Vol. 219 Part J: J. Engineering Tribology

J01504 # IMechE 2005

Lubrication regime transitions at the piston ringcylinder liner interface

29

Fig. 11 Effect of load on coefficient of friction at a constant speed of 120 r/min; (a) measured using piston ring
reciprocating liner test rig; (b) predicted from analysis

experienced by the piston ring, from boundary to


full-film hydrodynamic lubrication. Measured and
predicted values for coefficient of friction and lubricant film thickness are used to investigate the lubrication condition throughout the stroke. The effect
of operating speed and load are described. Depending on operating conditions, all three lubrication
regimes can occur at different points in the stroke.
As expected, friction is highest in the mixed lubrication and boundary lubrication regimes that occur
near TDC and BDC.
This study has provided the details of the
transition of the piston ring through the various
J01504 # IMechE 2005

lubrication regimes and has identified areas with


potential for improvement. The transitions and the
lubrication condition throughout the stroke have
been examined. The results presented here have
also shown that the numerical model is capable of
predicting the frictional loss and lubricant film thickness for a given piston ring geometry. In a design
setting, the numerical model could be used to evaluate potential piston ring behaviour by specifying
the piston velocity, ring tension, cylinder pressure
history, pressure history between piston rings, ring
crown height, ring width, effective surface roughness, and viscosity.
Proc. IMechE Vol. 219 Part J: J. Engineering Tribology

30

N W Bolander, B D Steenwyk, F Sadeghi, and G R Gerber

Fig. 12 Effect of load on minimum film thickness at a constant speed of 120 r/min (predicted)

ACKNOWLEDGEMENT
The authors would like to extend their deepest
appreciation to the US Department of Energy for
their support of this project.

REFERENCES
1 Jeng, Y. Theoretical analysis of piston-ring lubrication.
Part 1: fully flooded lubrication. STLE Tribology Trans.,
1992, 35, 696 706.
2 Rohde, S. M., Whitaker, K. W., and McAllister, G. T. A
mixed friction model for dynamically loaded contacts
with application to piston ring lubrication. In Surface
Roughness Effects in Hydrodynamic and Mixed Lubrication, Proceedings of the ASME Winter Annual Meeting,
1980, pp. 19 50.
3 Patir, N. and Cheng, H. S. An average flow model for
determining effects of three-dimensional roughness on
partial hydrodynamic lubrication. Trans. ASME,
J. Lubric. Technol., 1978, 100, 12 17.
4 Greenwood, J. A. and Tripp, J. H. The contact of two
nominally flat rough surfaces. Proc. Inst. Mech. Engrs
1971, 185, 625 633.
5 Dowson, D., Economou, P. N., Ruddy, B. L.,
Strachan, P. J., and Baker, A. J. Piston ring lubrication.
Part II: theoretical analysis of a single ring and a complete ring pack. In Energy Conservation Through Fluid
Film Lubrication Technology: Frontiers in Research and

Proc. IMechE Vol. 219 Part J: J. Engineering Tribology

Design, Proceedings of the ASME Winter Annual Meeting, 1979, pp. 23 52.
6 Hu, Y., Cheng, H. S., Arai, T., Kobayashi, Y., and
Aoyama, S. Numerical simulation of piston ring in
mixed lubrication a non-axisymmetrical analysis.
Trans. ASME, J. Tribology, 1994, 116, 470 478.
7 Yang, Q. and Keith, T. G., An elasto-hydrodynamic cavitation algorithm for piston ring lubrication. STLE Tribology Trans., 1995, 38, 97 107.
8 Vijayaraghavan, D. and Keith, T. G. Development and
evaluation of a cavitation algorithm. STLE Tribology
Trans., 1989, 32, 225 233.
9 Dowson, D., Ruddy, B. L., and Economou, P. N.
The elastohydrodynamic lubrication of piston rings.
Proc. R. Soc. Lond. A, 1983, 386, 409 430.
10 Yang, Q. and Keith, T. G. Two-dimensional piston ring
lubrication. Part 1: rigid ring and liner solution. STLE
Tribology Trans., 1996, 39, 757 768.
11 Yang, Q. and Keith, T. G. Two-dimensional piston ring
lubrication. Part 2: elastic ring consideration. STLE
Tribology Trans., 1996, 39, 870 880.
12 Sawicki, J. and Yu, B. Analytical solution of piston ring
lubrication using mass conserving cavitation algorithm. STLE Tribology Trans., 2000, 43, 587 594.
13 Akalin, O. and Newaz, G. M. Piston ring cylinder
bore friction modeling in mixed lubrication regime.
Part I: analytical results. Trans. ASME, J. Tribology,
2001, 123, 211 218.
14 Akalin, O. and Newaz, G. M. Piston ring cylinder
bore friction modeling in mixed lubrication
regime. Part II: correlation with bench test data.
Trans. ASME, J. Tribology, 2001, 123, 219 223.

J01504 # IMechE 2005

Lubrication regime transitions at the piston ringcylinder liner interface

15 Priest, M., Dowson, D., and Taylor, C. M. Theoretical


modeling of cavitation in piston ring lubrication. Proc.
Instn. Mech. Engrs, Part C: J. Mechanical Engineering
Science, 2000, 214, 435 447.
16 Ting, L. L. Development of a reciprocating test rig for
tribological studies of piston engine moving components. Part 1: rig design and piston ring friction
coefficients measuring method. SAE paper 930685,
1993.
17 Ting, L. L. Development of a reciprocating test rig
for tribological studies of piston engine moving components: Part 2: measurements of piston ring coefficients and rig test confirmation. SAE paper 930686,
1993.
18 Dearlove, J. and Cheng, W. K. Simultaneous piston ring
friction and oil film thickness measurements in a reciprocating test rig. SAE paper 952470, 1995.
19 Arcoumanis, C., Duszynski, M., Flora, H., and
Ostovar, P. Development of a piston-ring lubrication
test-rig and investigation of boundary conditions for
modeling lubrication film properties. SAE paper
952468, 1995.
20 Payvar, P. and Salant, R. F. A computational method
for cavitation in wavy mechanical seal. Trans. ASME,
J. Tribology, 1992, 114, 199 204.
21 Patankar, S. V. Numerical Heat Transfer and Fluid
Flow, 1980 (Hemisphere, Washington, DC).

APPENDIX
Notation
a
b
B
F
Ff
h
hmin
href
H
p
pa
pc
pel
pg

width of ring section across the sliding


direction (m)
ring thickness in the sliding direction (m)
cylinder bore diameter (m)
cavitation index
frictional force (N)
film thickness (m)
minimum film thickness (m)
reference film thickness (m)
dimensionless film thickness h/href
pressure (MPa)
ambient pressure (MPa)
cavitation pressure (MPa)
ring elastic pressure (MPa)
gas pressure acting on the back of the ring
(MPa)

J01504 # IMechE 2005

ph
pL
pT
ps
t
tref
T
Tr
U
x
X
y
Y

a
g
d
h
ma
r
rc
s
sc
f

31

hydrodynamic pressure (MPa)


pressure at the leading edge of the
ring (MPa)
pressure at the trailing edge of the
ring (MPa)
asperity contact pressure (MPa)
time (s)
reference time (s)
dimensionless time t/tref
ring tension (N)
sliding velocity (m/s)
coordinate in the sliding direction (m)
dimensionless coordinate in the sliding
direction x/b
coordinate across the sliding direction (m)
dimensionless coordinate across the
sliding direction y/a
relaxation factor for pressure
characteristic number 6hUb/h2ref 
( p a 2 p c)
crown height of the ring profile (m)
lubricant viscosity (Pa s)
dynamic asperity coefficient of friction
lubricant density (kg/m3)
lubricant density at cavitation pressure
(kg/m3)
squeeze number 12hb 2/tref(pa 2 pc)h2ref
effective surface roughness (mm)
dimensionless dependent variable defined
by equations (4) and (5)

Subscripts
e
i
j
n
s
w

east boundary surface of the control


volume
index for the grid number in the
x direction
index for the grid number in the y direction
north boundary surface of the control
volume
south boundary surface of the control
volume
west boundary surface of the control
volume

Proc. IMechE Vol. 219 Part J: J. Engineering Tribology

You might also like