You are on page 1of 14

See

discussions, stats, and author profiles for this publication at: http://www.researchgate.net/publication/238370060

Predictive wear modelling of lubricated piston


rings in a diesel engine
ARTICLE in WEAR JUNE 1999
Impact Factor: 1.86 DOI: 10.1016/S0043-1648(99)00125-8

CITATIONS

DOWNLOADS

VIEWS

42

162

190

3 AUTHORS, INCLUDING:
Martin Priest

Duncan Dowson

University of Bradford

University of Leeds

82 PUBLICATIONS 507 CITATIONS

380 PUBLICATIONS 7,384 CITATIONS

SEE PROFILE

SEE PROFILE

Available from: Duncan Dowson


Retrieved on: 15 September 2015

Wear 231 1999. 89101

Predictive wear modelling of lubricated piston rings in a diesel engine


M. Priest ) , D. Dowson, C.M. Taylor
School of Mechanical Engineering, The Uniersity of Leeds, Leeds, LS2 9JT, UK
Received 15 March 1999; accepted 18 March 1999

Abstract
The tribological performance of piston rings in reciprocating internal combustion engines can only be fully understood when both
lubrication and wear are considered in combination. To this end, a numerical model has been developed that predicts the dynamics,
lubrication and wear of piston rings interactively for the first time. This paper reports the application of this new model to the piston ring
pack of a diesel engine. With the overall aim of evaluating the correlation between theory and experiments, this analysis is divided into
two discrete parts. First, the model is used to predict the lubrication performance of measured ring packs before and after periods of
running, at constant speed and load, in a Caterpillar 1Y73 single-cylinder diesel engine: the objective being to establish the change in
tribological behaviour with observed wear in the engine. Secondly, the model is used interactively to predict the lubrication and wear of
the top compression ring from the same engine. This research advances the understanding of piston ring profile evolution with time and
its dependence on complex interactions between lubrication and wear. q 1999 Elsevier Science S.A. All rights reserved.
Keywords: Piston rings; Lubrication; Wear; Diesel engine

1. Introduction
The tribological behaviour of piston rings has long been
recognised as an important influence on the performance
of internal combustion engines in terms of power loss, fuel
consumption, oil consumption, blow-by and harmful exhaust emissions.
The primary role of the piston ring pack is to maintain
an effective gas seal between the combustion chamber and
the crankcase. The rings of the piston ring pack, which
together effectively form a labyrinth seal, achieve this by
closely conforming to their grooves in the piston and to the
cylinder wall. The small quantity of gas that does find its
way into the crankcase, blow-by, is normally piped back to
the inlet valve and fed back into the cylinder.
In addition to causing a dramatic increase in pressure,
the combustion event generates a large amount of heat.
Much of this thermal energy is convected into the piston
causing a marked increase in the temperature of the piston,
which is dissipated by heat transfer to adjacent components and the engine coolant. The secondary role of the
piston ring pack is to transfer this heat from the piston into
the cylinder wall and thence into the coolant.
)

Corresponding author. Tel.: q44-113-2332178; fax: q44-1132332150; E-mail: m.priest@leeds.ac.uk

The final function of the piston ring pack is to limit the


amount of oil that is transported from the crankcase to the
combustion chamber. This flow path is probably the largest
contributor to the oil consumption of an engine and leads
to an increase in harmful exhaust emissions as the oil
mixes and reacts with the other contents of the combustion
chamber w1x. The desire to extend service intervals of
engines and minimise harmful exhaust emissions to meet
ever more stringent legislative requirements, means that
the permissible oil-consumption levels of modern engines
are very low compared to their predecessors of 10 or 20
years ago w2x.
The piston ring pack must fulfil these three roles with a
minimum of frictional power loss, most notably at the
sliding interface with the cylinder wall, and a minimum of
wear in order to maximise component life. Unfortunately,
the piston ring pack is one of the largest sources of friction
in the internal combustion engine over the normal range of
engine speeds and loads encountered in service w35x.
Exact figures vary from engine to engine, but typically the
piston assembly, comprising both the piston rings and the
piston skirt, accounts for 4050% of total engine friction
w1x. Piston ring pack friction losses are greater than those
of the piston skirt at low to moderate engine speeds but the
situation may be reversed at high engine speeds due to the
large wetted area of the piston skirt contributing to viscous

0043-1648r99r$ - see front matter q 1999 Elsevier Science S.A. All rights reserved.
PII: S 0 0 4 3 - 1 6 4 8 9 9 . 0 0 1 2 5 - 8

90

M. Priest et al.r Wear 231 (1999) 89101

ture and lubricant availability. In one single stroke of the


piston, the piston ring may experience boundary, mixed
and full fluid film lubrication w7x. Elastohydrodynamic
lubrication of piston rings is also possible in both gasoline
and diesel engines on the highly loaded expansion stroke
after firing w15x.
Incorporating a consideration of wear into the dynamics
and lubrication analysis of piston rings adds a further layer
of complexity to the model. This is compounded by the
fact that wear is the least understood of the three main
processes in tribology: friction, lubrication and wear.
A piston ring tribology model incorporating prediction
of the change in ring face profile with wear in the engine
has recently been developed w16,17x. It assumes that wear
of the ring profile may be described by the Archard wear
equation, in the form proposed by Lancaster w18x:
V s kWx s
Fig. 1. Caterpillar 1Y73 piston and ring pack.

friction w6x. In terms of wear, there is insufficient understanding of the interaction with the lubrication process. So,
even though manufacturers can produce rings that have an
excellent life expectancy, these components may be far
from optimum from a lubrication and friction standpoint.
As a consequence of their importance to engine performance, the theoretical and experimental study of piston-ring
lubrication has received much attention in the literature,
leading to major advances in our understanding of their
behaviour w7x. The mathematical analysis of piston-ring
lubrication is complex and by necessity requires simplifying assumptions. However, rapid development in numerical methods over the last 25 years has resulted in sophisticated piston-ring lubrication models that are finding application in the design process w8x.
It is the proposition of the research presented in this
paper, however, that a complete understanding of the
tribological performance of piston rings in reciprocating
internal combustion engines can only be achieved when
both lubrication and wear are considered in combination.
The running profile of the piston ring that slides against
the cylinder wall wears significantly in service, even with
wear-resistant materials and coatings, such that the ring
profile after only a short period of running in the engine
differs greatly from that of the component as new w914x.
Modification of the ring profile by wear has a large effect
on lubrication, friction and oil transport at the interface
between the piston ring and cylinder wall which then in
turn modifies the wear conditions. This interaction between lubrication and wear has important implications for
the performance and life of the internal combustion engine.
The piston ring is perhaps the most complicated tribological component in the internal combustion engine. It is
subjected to large, rapid variations of load, speed, tempera-

1.

The wear factor, k, is a function of the interacting


materials, their surface topography, the lubricant and the
operating conditions. This can alternatively be expressed
as a variation of wear factor with specific film thickness
relative to the wear factor in the boundary lubrication
regime, k 0 . The wear factor in the boundary lubrication
regime is determined from bench test rig experiments
using actual components and lubricant at operating conditions of load, speed and temperature indicative of boundary lubrication. This empirical input to the model clearly
exposes our lack of fundamental understanding of the wear
processes taking place in such tribological interfaces. This
approach, however, has been applied successfully in automotive valve train wear modelling w19,20x.
With this relationship and the cyclic variation of minimum film thickness predicted by the lubrication model, the
wear factor can be determined at each crank angle in the
engine cycle. Thus it is possible to predict, interactively,
the changes in wear and lubrication of the piston ring that

Fig. 2. Piston ring measurement on the Form Talysurf profilometer.

M. Priest et al.r Wear 231 (1999) 89101

take place with running time in the engine. Full details of


the analysis may be found in Refs. w16,17x.
This paper considers the application of this new interactive dynamics, lubrication and wear model to the piston
ring pack of a Caterpillar 1Y73 single-cylinder diesel
engine. After describing piston-ring and cylinder-wall wear
experiments undertaken on the engine, the theoretical analysis is presented as two discrete parts. In the first, the

91

model is used to predict the lubrication performance of


measured ring packs before and after periods of running in
the engine, the objective being to establish the change in
tribological behaviour with observed wear in the engine.
Secondly, the model is used interactively to predict the
lubrication and wear of the top compression ring from the
same engine, with the objective of evaluating the correlation between the new model and experiments.

Fig. 3. Variation of measured ring profiles with time.

92

M. Priest et al.r Wear 231 (1999) 89101

2. Experimental results
Fig. 1 shows the geometry of the piston ring pack and
the piston from the Caterpillar 1Y73, single-cylinder, diesel
engine. The piston ring pack consists of three compression
rings and an oil-control ring mounted on a cast-iron piston
with a steel insert forming the top ring groove. The top
compression ring, Ring 1, is a barrel-faced, chromiumplated, cast-iron ring and the other two compression rings,
Rings 2 and 3, are of the same design and are taper-faced,
plain cast-iron, scraper rings with an internal step. This
step gives them a slight upward twist when fitted, giving
them a dished appearance, such that they offer their lower
edge to the cylinder wall. The oil-control ring, Ring 4, is a
twin-land, single-piece design with interland drainage slots
and a coil spring expander, the lands of which are
chromium plated. The cylinder is a wet liner design and is
manufactured from induction hardened cast iron.
Further physical data for the ring pack is given for
reference in Appendix A. The engine was run at a
crankshaft speed of 1200 rpm and a nominal brake load of
either 10 or 14 bar brake mean effective pressure BMEP..
Basic engine data and operating parameters relevant to the
piston assembly for these two conditions are also given in
Appendix A.
Extended duration tests, at constant speed and load,
were undertaken to study the changes occurring in the
piston ring profiles and the surface roughness of the rings
and the cylinder wall. Piston ring profiles and surface
roughness data were measured using a Rank Taylor Hobson Form Talysurf profilometer with a laser-referenced
stylus rather than the more traditional inductive system.
The arrangement used to measure the Caterpillar 1Y73
piston rings on the Form Talysurf is shown in Fig. 2. A jig
was devised to locate the piston rings, which consisted of
two vee block supports attached to tee slots in the machine
base with a solid steel cylinder suspended between them.
The cylinder had a flat machined along its length to which
a small vice was fixed which held the piston ring and a
laboratory slip gauge as shown in Fig. 2. The slip gauge
was used to establish a datum for the ring profiles to
enable comparison to be made between different measurements. Before each ring profile measurement was taken,
the relative heights of the vee block supports were adjusted
until a 2.0-mm traverse across the slip gauge gave a
maximum vertical displacement of the stylus of less than
1.0 mm, which was equivalent to a maximum slip gauge
surface inclination of 0.0288. As the slip gauge was
clamped against the ring flank, this gave a reference for all
the measurements, assuming that the ring flanks did not
wear greatly in service.
Ring profiles were measured at three circumferential
locations remote from the ring gap using a stylus with a
2-mm tip radius. It was decided at this stage in the research
not to measure ring profiles near the ring gap, where
unknown dynamic effects in service may lead to a wear

pattern that could not confidently be reproduced using


analytical wear models. A simple template was used to
mark the ring flanks prior to measurement.
Fig. 3 shows the measured ring profiles at a circumferential position 908 offset from the ring gap after 0, 120 and
628 h running.
The rings at 0 at 120 h are the same rings measured
before and after running at a constant speed and load of
1200 rpm and 14 bar BMEP. The data reported for 628 h,
however, is for a completely different ring pack that was
run for half the test at 1200 rpm and 10 bar BMEP and
half at 1200 rpm and 14 bar BMEP. Although not a true
comparison with the results at 0 and 120 h, the measurements at 628 h show an entirely consistent trend and are a
useful indication of the long-term wear of the rings in this
engine.
Wear of the piston rings at the other two circumferential
positions was similar for the compression rings, which is
not surprising given that all three measurement locations
are remote from the ring gap and the rings are free to
rotate around the piston. The oil-control ring exhibited
some circumferential variation but this was small compared to the resolution of the measurement technique. Full
details can be found in Ref. w16x.
The variation of root mean square surface roughness of
the components, in the direction of sliding, throughout the
test is given in Table 1. The surface roughness was
determined using the software provided with the Form
Talysurf which simulates the standard electronic filtering
techniques developed for inductive stylus profilometers
w21x. The data for the piston rings is for the full profile at 0
h and for the worn regions of the profile at 120 and 628 h.
The surface roughness values for the cylinder wall are the
mean of a series of measurements taken at top dead centre,
midstroke and bottom dead centre. This approach was
dictated by the use of these parameters in the mathematical
model.
The root mean square surface roughness of the cylinder,
measured using the Form Talysurf with standard electronic
filtering, is strongly affected by the deep honing marks
introduced into the surface during manufacture. This parameter is used in the mixed lubrication model of the ring
pack, in which the deep honing marks play no part, and it

Table 1
Variation of axial surface roughness parameters with time
Component

s mm.
Ring 1
Ring 2
Ring 3
Ring 4 upper land.
Ring 4 lower land.
Cylinder

Time h.
0

120

628

0.112
0.227
0.227
0.033
0.039
0.476

0.008
0.029
0.016
0.007
0.039
0.139

0.078
0.023
0.012
0.017
0.037
0.180

M. Priest et al.r Wear 231 (1999) 89101

was therefore considered unrepresentative to input the


measured values including the honing marks into the model.
A parametric study was undertaken of the surface roughness of small sections of the cylinder wall not containing
honing marks, manually chained together to give sufficient
data for statistically significant results. This suggested that
new cylinder roughness values should be reduced by a
factor of 0.50 and worn values by a factor of 0.32 before

93

input to the model. These scaling factors are included in


the data reported in Table 1.
3. Lubrication predictions for measured piston ring
profiles
Fig. 4 shows the variation of minimum lubricant film
thickness throughout the engine cycle for the new ring

Fig. 4. Predicted film thickness variation at 0 h.

94

M. Priest et al.r Wear 231 (1999) 89101

pack in both fully flooded and starved lubrication conditions, 08 crank angle being top dead centre firing. The fully
flooded model of lubricant flow in a ring pack assumes
that there is an unlimited supply of lubricant available to
each ring at all stages in the engine cycle such that the
inlet region of the ring profile is always full of lubricant.
In reality, however, the quantity of lubricant available to
each ring is the thin film smeared on the cylinder wall by

the preceding ring and consequently the inlet region of the


ring profile may be starved of lubricant. This more sophisticated approach is termed starved lubrication w7x.
Included on each graph in Fig. 4 is the upper limit of
surface contact, and hence wear, which is equal to four
times the composite root mean square roughness of the
piston ring and cylinder wall, 4s . This defines the transition between the mixed and full fluid film lubrication

Fig. 5. Predicted film thickness variation at 120 h.

M. Priest et al.r Wear 231 (1999) 89101

regimes. Hence, if the film thickness is less than 4s ,


surface contact and wear will occur. The transition between the mixed and boundary lubrication regimes is equal
to half the composite surface roughness, 0.5 s . The lubrication model uses this latter value as the minimum permitted film thickness and it can be observed as a distinct
minimum limit in Fig. 4. The upper limit of surface

95

contact for the oil-control ring was computed using the


largest of the two land roughness values as the minimum
film thickness can occur on either land depending on the
attitude of the ring to the cylinder wall.
In fully flooded conditions, all four rings experience
full fluid film, mixed and boundary lubrication. The very
thin film generated on the downstrokes by Ring 4, the

Fig. 6. Predicted film thickness variation at 628 h.

M. Priest et al.r Wear 231 (1999) 89101

96
Table 2
Predicted friction power loss
Time h. Friction power loss W .
Fully flooded lubrication

Starved lubrication

Ring 1 Ring 2 Ring 3 Ring 4 Ring 1 Ring 2 Ring 3


0
120
628

162.7
126.9
119.5

138.1
91.0
87.6

132.0
72.8
73.3

135.4
75.9
46.0

356.8
417.2
270.2

147.1
173.3
160.0

143.8
170.6
158.6

oil-control ring, is of great concern as this ring controls the


supply of lubricant to the compression rings under starved
lubrication conditions. Extreme starvation of the compression rings by the oil-control ring is indicated in the starved
results of Fig. 4. All the compression rings have much
reduced film thicknesses compared to their fully flooded
values, with no ring achieving a full fluid film at any stage
in the engine cycle. Wear would occur at all stages in the
engine cycle under these conditions.
A step change can be seen in film thickness in some of
the curves of Fig. 4, e.g., at 4708 for Ring 1 and 5008 for
Rings 2 and 3. The occurs when the ring moves axially
from the top to the bottom of the piston groove, or vice
versa, resulting in a change in radial gas load, a phenomenon often referred to as ring lift.
Fig. 5 shows the film thickness parameters for the same
ring pack after 120 h running in the engine. First, it should
be noted that the lubrication regime transition limits have
smaller absolute values due to reductions in composite
surface roughness of the rings and the cylinder as reported
in Table 1. Looking first at the fully flooded results it can
be seen that all the compression rings, Ring 1 to 3, develop
a full fluid film for much more of the engine. In particular,
Ring 1 no longer has the sustained period of damaging
boundary lubrication from 7038 to 568 crank angle observed with the new ring in Fig. 4. Ring 4 exhibits very
modest hydrodynamic film action on all strokes but with a
full fluid film only achieved on the downstrokes. Overall,
the performance of this ring has apparently deteriorated
from the new ring performance given in Fig. 4. This
consequently starves the compression rings of lubricant
even more severely than at 0 h with very low levels of
hydrodynamic film generation for all three compression
rings under starved lubrication conditions.
The results for 628 h running in the engine are given in
Fig. 6. Modest improvements in film thickness are observable for the compression rings in fully flooded conditions.
Ring 4 generates full fluid films on the downstrokes with a
boundary film on the upstrokes. This is the complete
opposite to the behaviour of the new ring and very different to the performance after 120 h. The existence of a full
fluid film under the oil-control ring on the downstrokes
lessens the severity of lubricant starvation of the compression rings. The performance of all three compression rings
under starved conditions has consequently improved

slightly. Ring 3 has improved the most such that on the


downstrokes it is predicted to have a full fluid film in the
midstroke region.
Friction power loss predictions for the complete ring
pack at 0, 120 and 628 h running in the engine are
summarised in Table 2. Looking first at the fully flooded
results, the compression rings show a pattern very much
akin to running in with significant reductions in friction in
the first 120 h and little change thereafter. The oil-control
ring, Ring 4, however, has similar reductions at the end of
each time interval. The starved lubrication results for the
compression rings give a much more confusing picture.
The lubricant starvation caused by the oil-control ring
apparently masks any friction changes due to profile modification, reductions in surface roughness and interaction
between the compression rings. The likelihood of this
extreme level of starvation occurring in practice is discussed later in this paper.

4. Piston ring profile wear predictions


To investigate predicted changes in piston ring geometry and performance due to wear, the piston ring lubrication and wear model w16,17x was applied to the top compression ring, Ring 1, of the Caterpillar 1Y73 engine over
the first 120 h of running. The ultimate objective was to
achieve good correlation between predicted and measured
ring profiles at 120 h.
The ring profile of Ring 1 at 0 and 120 h is shown in
Fig. 2. As each measurement of ring profile has a different
coordinate origin, the ring profile measured after 120 h
running cannot be simply plotted over the new profile to
evaluate wear. Efforts were therefore made to overlay the
profiles and Fig. 7 includes the results of two alternative
approaches. First there is what is referred to as a geometric
overlay where the worn ring profile was moved in x and z
until the new profile and the apparently unworn regions of
the worn profile appeared to correlate well. Second, there

Fig. 7. Measured profiles of the new and worn ring.

M. Priest et al.r Wear 231 (1999) 89101

is what has been labeled a mass overlay which was


determined from the weight loss of the ring over the 120-h
test period, 19.9 mg, assuming that equal mass was lost at
all locations around the circumference of the ring. The
worn profile was then moved in x and z until the area
between the new and worn profiles correlated with the
weight loss. The obvious difference between the geometric
and mass overlays in Fig. 7 suggests that the wear of the
ring was not equal at all locations around the ring circumference. It should be noted that no rotation of the worn
profiles was introduced as part of the overlay procedures.
The first stage in the solution process was to input the
initial topography and operating conditions to the lubrication model to determine the cyclic variation of lubricant
film thickness, applied radial load and axial velocity which
are required by the wear model to determine wear rates. As
this is a single-ring analysis and as such no lubrication
predictions are undertaken for the other rings in the pack,
fully flooded lubrication conditions were assumed throughout. Also, the lubricant separation boundary conditions for
Reynolds equation developed by Coyne and Elrod w22x
were adopted.
The lubrication model used in this analysis did not
determine the cyclic variation of torsional twist angle of
the piston ring, which is also needed by the wear model,
and so an assumed ring twist pattern was required. Fig. 8
shows the predicted cyclic variation of twist angle for the
top compression ring of a 203-mm bore, single-cylinder
diesel engine from the published literature w23x.
Based on this predicted behaviour, a sinusoidal cyclic
ring twist variation with an amplitude of 0.18, a period of
2408 crank angle and zero offset, 08 ring twist at 08 crank
angle, was assumed for this investigation. This was chosen
to give best agreement with the data in Fig. 8 during the
first half of the power stroke when the wear rates were
expected to be highest. A graphical representation of the
assumed ring twist characteristic is given in Fig. 9.

97

Fig. 9. Assumed ring twist variation.

Another important input parameter to the wear model is


the wear factor under boundary lubrication conditions, k 0 .
Reciprocating wear tests in boundary lubrication conditions were undertaken on samples of the piston ring and
cylinder liner from this engine. The tests were run with a
good quality diesel engine lubricant in both fresh and
degraded conditions and the results are summarised in
Table 3.
After some initial trials, a value of k 0 of 1.5 = 10y1 2
mm3 mmy1 Ny1 was adopted as most appropriate to the
top ring of this engine during the first 120 h of running.
During this time, the lubricant will go from fresh to a
degraded condition due, primarily, to exposure to heat. The
bulk temperature of the lubricant available to the piston
ring at top dead centre, where much of the wear appeared
to occur according to the initial trials, is assumed to be
equal to that of the cylinder wall at this position, which
was 1408C. It was also assumed that the surface roughness
of the piston ring and cylinder wall reduced linearly with
simulation time from the measured values at 0 h to the
measured values at 120 h as given in Table 1.
The wear model was then used to predict the development of the piston ring profile over a 2-h period assuming
the lubricant film thickness at any point in the engine cycle
remained constant during this time. The worn piston ring
topography was then fed back to the lubrication model and
the results fed forward into the wear model for another 2-h
period of simulated profile development. This interactive
process was continued until the total desired simulation
time of 120 h was achieved.
The minimum film thickness variation throughout the
engine cycle, as predicted by the lubrication model for the

Table 3
Piston ring wear factors from lubricated reciprocating wear tests

Fig. 8. Predicted ring twist angles from the literature w23x.

Bulk
temperature 8C.

Lubricant
condition

k0
mm3 mmy1 Ny1 .

200
300
300

Fresh
Fresh
Degraded

2=10y1 3
4=10y1 2
5=10y1 1

M. Priest et al.r Wear 231 (1999) 89101

98

Fig. 10. Cyclic variation of specific film thickness at 0 h.


Fig. 12. Predicted ring profile compared with the measured profiles.

initial profile at 0 h, is plotted in Fig. 10 as specific film


thickness, l, where

ls

h min

2.
s
Also marked on the graph are the transitions between
boundary and mixed lubrication and mixed and full fluid
film lubrication. The film-thickness predictions indicate
that at 0 h, wear occurs for the majority of the engine cycle
as a full fluid film is only achieved in midstroke regions
on the power, intake and compression strokes. Note, the
shape of the curve is slightly different to that presented in
Fig. 4 due to the use of the Coyne and Elrod cavitation
model rather than the more common Reynolds cavitation
and reformation model w22x.
The wear-rate variation with crank angle predicted by
the wear model at 0 h is shown in Fig. 11 and indicates
that, although wear takes place for a significant proportion
of the engine cycle, the largest values occur around top
dead centre firing where the applied load is greatest. The
wear rate tends to zero at the dead centres because the
sliding velocity falls to zero as the piston changes direction.
The predicted profile at 120 h is shown in Fig. 12 along
with the measured ring profiles. Good agreement is appar-

ent in both the quantity of material worn away and the


shape of the profile.

5. Discussion
5.1. Lubrication predictions with measured piston ring
profiles
The film thickness predictions from the fully flooded
lubrication analysis indicate that the mathematical model is
capable of evaluating the expected improvement in performance of these rings with running time in the engine. In an
attempt to quantify the general health of the interface
between the piston ring and cylinder wall, Table 4 presents
the proportion of the engine cycle for which the film
thickness is less than the upper limit of surface contact and
hence wear is likely to occur. The compression rings,
Rings 1 to 3, all greatly improve their performance between 0 and 120 h and then change little between 120 and
628 h indicating the completion of running-in by 120 h as
suggested by the friction results in Table 2. The same
cannot be said of the oil-control ring, Ring 4, which shows
no categorical improvement.
The oil-control ring operating conditions are crucial in
determining the predicted performance of the compression
rings through the starved lubrication analysis. The very
low film thicknesses predicted for the compression rings in
starved conditions in Figs. 46 and the high levels of wear

Table 4
Assessment of the fully flooded results
Time h.

0
120
628
Fig. 11. Cyclic variation of wear rate at 0 h.

Proportion of the engine cycle for which the


minimum film thickness is less than 4s %.
Ring 1

Ring 2

Ring 3

Ring 4

41
7
10

61
11
9

62
9
8

61
78
50

M. Priest et al.r Wear 231 (1999) 89101

99

Table 5
Laser-induced fluorescence LIF. film-thickness measurements
Stroke

Power
Exhaust
Induction
Compression

Minimum film thickness mm. wdegrees of crank angle after top dead centrex
Ring 1 w1038x

Ring 2 w938x

Ring 3 w888x

Ring 4 upper land. w848x

Ring 4 lower land. w818x

0.55
0.60
1.2
1.4

3.2
1.3
2.5
2.5

5.2
2.2
5.1
3.3

9.3
6.5
9.1
5.3

7.6
10.7
4.5
10.5

that would result seem rather unrealistic. It is proposed that


the oil-control film thickness predictions are too severe
and the oil-control ring is therefore excessively starving
the compression rings of lubricant to a degree that seems
unlikely in practice. A possible cause of these severe
predictions could be inaccuracies in the assumed oil-control ring profile. In measuring the profile of the oil-control
ring lands, in terms of inclination and relative height
difference, we have attempted to resolve to within 1 mm
over a Form Talysurf traverse of 6 mm. This corresponds
to a slope of 0.00958, which is less than the tolerance of
0.0288 allowed on setting the datum of the measurement
system, as discussed previously.
To investigate this further, Table 5 shows laser-induced
fluorescence LIF. film-thickness measurements under the
piston rings of this engine from a single transducer mounted
in the cylinder liner near midstroke. Details of the application of this technique to the Caterpillar 1Y73 engine can
be found in Ref. w24x.
The measured film thicknesses for the oil-control ring
lands are large when compared to the theoretical prediction
of oil-control ring minimum film thickness, which is output as the smallest of the two land film-thickness values at
any instant in time, in Figs. 46. Only on the downstrokes
at 628 h, Fig. 6, do the predictions come close to the
measurements with a predicted maximum film thickness of
approximately 3.2 mm. This adds considerable weight to
the conclusion that the mathematical model is predicting
film-thickness levels that are far too low for the oil-control
ring.
As a direct consequence of this argument, it is likely
that the true running condition of the compression rings
lies somewhere between the fully flooded and starved
predictions of the analysis. Comparison of the measured
film thicknesses in Table 5 and film-thickness predictions
after 120 h running in Fig. 5, probably the most comparable conditions to those of the measurements, supports this
view.
5.2. Piston-ring profile wear predictions
The prediction of the wear for the top compression ring
from the Caterpillar 1Y73 diesel engine over the first 120
h of running, reported in this paper, is the first application
of the piston-ring dynamics, lubrication and wear model

developed in Refs. w16,17x to fired engine test data. As


such, the results are most encouraging with good agreement between measurements and theory in terms of the
amount of wear taking place and the shape of the worn-ring
profile, as shown in Fig. 12. The most important aspects of
the piston ring profile, in terms of lubrication performance,
are its curvature and the axial position of the peak of the
profile. In both respects, the measured and predicted profiles show a high degree of correlation.
As noted previously, incorporating a consideration of
wear into the dynamics and lubrication analysis of piston
rings adds a further layer of complexity to the model.
Also, because of our basic lack of fundamental understanding of the wear process, it necessitates a degree of empiricism. Hence the need to input a wear factor for the
boundary lubrication regime to the model based on bench
test rig results such as those reported in Table 3. Closer
consideration of the initial trials of the model, undertaken
to determine the actual value of wear factor to be input
from the range reported in Table 3, reveals a self-calibration mechanism inherent in the model. More precisely,
given the worn ring profile and a detailed knowledge of
the operating history of the piston ring, it should be
possible to works backwards to determine the wear factor
in the boundary lubrication regime. This process, naturally,
depends upon the confidence that can be placed in the
accuracy of the model. It is envisaged, that further validation of the model using data from fired engine tests and
bench wear test rigs, will reduce the reliance of the model
on empirical input.
The assumption of fully flooded lubrication conditions
in these predictions is clearly an oversimplification but
was necessary as only one ring was analysed in this first
application of the model. However, the choice of wear
factor ameliorated the effect on the results.
Table 6
Piston-ring data for the Caterpillar 1Y73
Parameter

Mass g.
Elastic modulus GPa.
Fitted ring gap mm.
Tangential ring tension N.

Ring number
1

47.9
117
0.66
54

35.1
117
0.61
31

35.0
117
0.61
32

42.4
117
0.56
40

100

M. Priest et al.r Wear 231 (1999) 89101

Table 7
Basic engine data and operating conditions for the Caterpillar 1Y73
Parameter

Operating conditions
1200 rpm, 10 bar

Cylinder bore diameter mm.


Piston stroke mm.
Connecting rod length mm. wbetween bearing centresx
Gudgeon pinrcrank axis offset mm.
Lubricant dynamic viscosity at 408C mPa s.
Lubricant dynamic viscosity at 1308C mPa s.
Engine speed rpm.
Piston land temperature above ring 1 8C.
Piston land temperature below ring 1 8C.
Piston land temperature below ring 2 8C.
Piston land temperature below ring 3 8C.
Cylinder wall temperature for ring 1 at top dead centre 8C.
Cylinder wall temperature at midstroke 8C.
Cylinder wall temperature for ring 4 at bottom dead centre 8C.

One criticism that can be fairly made of the results is


that rather bold assumptions were made about the torsional
twisting behaviour of the piston ring based on very limited
studies available in the literature w23x. This arose because
the lubrication model used in this analysis did not determine the cyclic variation of torsional twist angle of the
piston ring, which is needed by the wear model. Subsequent development of the model has given it the capability
to predict torsional twisting of the ring, albeit at an increased cost in terms of numerical complexity. Somewhat
fortuitously, the predicted variation of torsional twist angle
around top dead centre firing, where much of the wear is
predicted to occur, as shown in Fig. 11, agrees well with
the assumed variation of twist angle given in Fig. 9.
Hence, a greater level of confidence can be placed in the
reported wear predictions than would have originally been
assumed. The application of the more advanced model,
incorporating the calculation of torsional twist angle, will
be reported in a future paper.

6. Conclusions
i. This paper reports the first application of a new
interactive dynamics, lubrication and wear model for piston rings to data from fired engine tests.
ii. The profile that the piston rings offer to the cylinder
wall was observed to wear significantly during the engine
tests, particularly for the top compression ring during the
first 120 h. The surface roughness of the piston rings and
the cylinder wall also reduced dramatically over this period.
iii. Lubrication predictions for the measured ring profiles have highlighted the sensitivity of lubricant film
thickness and friction, between the piston ring and cylinder
wall, to wear of the ring profiles.
iv. The lubrication predictions also demonstrate the
capability of the model to evaluate the expected improve-

1200 rpm, 14 bar


130.2
165.1
273.1
0
86.04
5.14
1200

244
198
184
165
130
100
90

270
218
200
170
140
105
90

ment in piston ring performance during the early life of the


components, often referred to as the running-in period.
v. Predicting the performance of the oil-control ring
has been shown to be very problematic, as the lubrication
analysis is very sensitive to measured profile parameters
such as the relative heights of the two lands.
vi. Interactive wear and lubrication predictions for the
top compression ring of the diesel engine studied show
encouraging correlation between measured and predicted
ring profiles after 120 h running. This is particularly
encouraging in the light of assumptions made in the analysis in terms of wear behaviour in the boundary lubrication
regime, availability of lubricant and torsional twisting of
the piston ring.
vii. Wear of the new top compression ring was predicted to occur for much of the engine cycle with the
largest wear rates around top dead centre firing, where the
applied loads are high and film thicknesses low.
viii. Strategies to overcome the limiting assumptions of
the model have been identified.
ix. This research advances the understanding of piston
ring profile evolution with time and its dependence on
complex interactions between lubrication and wear.
7. Nomenclature
h min
k
k0
V
W
xs
x
z
l
s

minimum film thickness between the piston


ring and cylinder wall
wear factor
wear factor in the boundary lubrication regime
worn volume
normal load
sliding distance
axial coordinate of piston ring profile
radial coordinate of piston ring profile
specific film thickness
composite root mean square surface roughness of the piston ring and cylinder wall

M. Priest et al.r Wear 231 (1999) 89101

Acknowledgements
The authors would like to thank Shell Research, Shell
Research and Technology Centre, Thornton, UK and the
Industrial Unit of Tribology, University of Leeds, UK for
their financial and technical support of this research. Particular thanks go to Shell Research for use of the Caterpillar 1Y73 engine and for supplying the wear factor data
reported in Table 3 and the LIF film thickness measurements summarised in Table 5. The advice of Dr. J.C. Bell
of Shell Research in relation to surface topography and
wear is greatly appreciated by the authors.

Appendix A
A.1. Caterpillar 1Y73 engine data

w9x

w10x

w11x

w12x

w13x

w14x
w15x

Further physical data for the ring pack is given for


reference in Table 6. Basic engine data and operating
parameters relevant to the piston assembly are given in
Table 7.

w17x

References

w18x

w1x S.T. Gazzard, D.R. Eastham, R.J. Jakobs, R.L. Lunsford, Piston
system design for low emissions, Leading Through Innovation, T&N
Symposium 1995, Paper 20, 1995.
w2x R. Munro, Emissions impossiblethe piston and ring support system, SAE Paper 900590, 1990.
w3x M.L. Monaghan, Engine frictiona change in emphasis, 2nd BP
Tribology Lecture, Inst. Mech. Eng., 1987.
w4x M.L. Monaghan, Putting friction in its place, 2nd Int. Conf.: Combustion EnginesReduction of Friction and Wear, Inst. Mech. Eng.
Conf. Pub. 1989-9, Paper C375rKN1, 1989, pp. 15.
w5x D.A. Parker, D.R. Adams, Friction losses in the reciprocating internal combustion engine, tribologyKey to the Efficient Engine, Inst.
Mech. Eng. Conf. Pub. 1982-1, Paper C5r82, 1982, pp. 3139.
w6x R.J. Chittenden, M. Priest, The theoretical analysis of piston ring
and piston skirt performance and its application in engine friction
modelling, submitted to J. Tribol., Trans. ASME, 1998.
w7x D. Dowson, Piston assemblies; background and lubrication analysis,
in: C.M. Taylor Ed.., Engine Tribology, Tribology Series, 26,
Chap. 9, Elsevier, 1993, pp. 213240.
w8x A. Mierbach, M.L. Hildyard, D.A. Parker, H. Xu, Piston ring

w16x

w19x

w20x
w21x
w22x

w23x

w24x

101

performance modelling, Leading Through Innovation, T and N


Symposium 1995, Paper 15, 1995, pp. 15.115.13.
L.L. Ting, J.E. Mayer Jr., Piston ring lubrication and cylinder bore
wear analysis: Part I. Theory, J. Lubr. Technol., Trans. ASME,
Paper 73-LUB-27, 1974, pp. 305313.
L.L. Ting, J.E. Mayer Jr., Piston ring lubrication and cylinder bore
wear analysis: Part II. Theory verification, J. Lubr. Technol., Trans.
ASME, Paper 73-LUB-25, 1974, pp. 258266.
S. Furuhama, M. Hiruma, The relationship between piston ring
scuffing and the formation of surface profile, Piston Ring Scuffing,
Inst. Mech. Eng. Conf., Paper C67r75, 1976, pp. 3543.
G.M. Hamilton, S.L. Moore, Measurement of piston ring profile
during running-in, Piston Ring Scuffing, Inst. Mech. Eng. Conf.,
Paper C69r75, 1976, pp. 6170.
A.V. Sreenath, N. Raman, Running-in wear of a compression ignition engine: factors influencing the conformance between cylinder
liner and piston rings, Wear 38 1976. 271289.
A.V. Sreenath, N. Raman, Mechanism of smoothing of cylinder
liner surface during running-in, Tribol. Int. 9 2. 1976. 5562.
J.E. Rycroft, R.I. Taylor, L.E. Scales, Elastohydrodynamic effects in
piston ring lubrication in modern gasoline and diesel engines, in:
Dowson et al. Ed.., ElastohydrodynamicsFundamentals and Applications in Lubrication and Traction, Proc. 23rd Leeds-Lyon Symposium on Tribology, September 1996, Elsevier, 1997, pp. 4954.
M. Priest, The wear and lubrication of piston rings, PhD thesis, The
University of Leeds, 1996, 246 pp.
M. Priest, D. Dowson, C.M. Taylor, Prediction of the lubrication
and wear of piston ringstheoretical model, World Tribology
Congress, Abstracts of Papers, September, 1997, p. 82, full paper in
preparation for J. Eng. Tribol. Proc.Inst. Mech. Eng., Part J, 1999.
J.K. Lancaster, Dry bearings: a survey of materials and factors
affecting their performance, Tribology, December, 1973, pp. 219
251.
J.C. Bell, P.J. Davies, W.B. Fu, Prediction of automotive valve train
wear patterns with simple mathematical models, Mechanisms and
Surface Distress, Proc. 12th Leeds-Lyon Symposium on Tribology,
Paper XI iii., 1985, pp. 323333.
J.C. Bell, T. Colgan, Pivoted follower valve train wear: criteria and
modelling, Lubr. Eng., J. STLE, February 1991. 114121.
T.R. Thomas, Rough Surfaces, 1st edn., Longman Group, Harlow,
UK, 1982.
M. Priest, D. Dowson, C.M. Taylor, Theoretical modelling of cavitation in piston ring lubrication, accepted for J. Mech. Eng. Sci, Proc.
Inst. Mech. Eng., Part C, 1999.
B.L. Ruddy, D. Dowson, P.N. Economou, A.J.S. Baker, Piston ring
lubrication: Part III. The influence of ring dynamics and ring twist,
in: S.M. Rohde, D.F. Wilcock, H.S. Cheng Eds.., Energy Conservation Through Fluid Film Lubrication Technology: Frontiers in Research and Design, ASME pub., 1979, pp. 191215.
M.A. Brown, H. McCann, D.M. Thompson, Characterization of the
oil film behaviour between the liner and piston of a heavy-duty
diesel engine, SAE Paper 932784, 1993.

You might also like