Anda di halaman 1dari 37

Simulation Based Study of the System Piston-Ring-Cylinder

of a Marine Two-Stroke Engine

ANDRZEJ WOLFF
Warsaw University of Technology
Faculty of Transport, Koszykowa Str. 75
PL-00-662 Warsaw, Poland

ABSTRACT

A comprehensive model of the system piston-ring-cylinder (PRC) of a marine two-stroke


engine is presented. Among sub-models developed by the author concerning: gas flow through
the labyrinth seal of piston rings, oil flow, ring twist and axial ring motion in piston grooves, the
model includes sub-models of mixed lubrication formulated by Patir, Cheng and Greenwood,
Tripp. The main parts of developed model and software have been experimentally verified during
a research period of the author at a marine engine designing centre. A relatively good qualitative
and quantitative compatibility between the experimental measurements and calculated results has
been achieved.
In contrast to the previous papers of the author concerning mostly automobile engines, new
calculation results for a marine two-stroke engine are presented. These results include basic
physical quantities associated with gas and oil flow processes in the PRC system for the analyzed
engine. The developed model and software can be utilized for: evaluation of gas leakage through
the sealing ring set, prediction of lubrication conditions of piston rings and oil consumption,
defining areas of the possible cylinder liner wear and profile changes of piston rings sliding
surfaces, and thus can be useful for optimization of the PRC system design.

KEY WORDS

Marine Diesel Engines, Piston Rings, Labyrinth Seals, Hydrodynamic Lubrication, Surface Roughness

INTRODUCTION

Two-stroke marine engine


Low-speed marine diesel engines are tailored for the economic propulsion of container feeder
vessels or crude oil tankers (1). A general construction of such an engine can be seen in Fig. 1.

1
The analyzed two-stroke marine engine is equipped with a pressure accumulation oil supply
system (2,3). This is an external system that feeds the oil through the cylinder liner onto the
running surface (Fig. 2a). The volumetric lube oil feed pump is driven by a frequency controlled
electric motor. A frequency converter controls the motor speed and thus the oil feed volume. The
frequency adjustment is dependent on engine load. The oil is fed to the lube oil distributor blocks
and from there to the lube oil storage accumulators (2,3).
There are two lubrication levels: the upper at approximately 10% and the lower at 20%
engine stroke from piston top dead centre (TDC). Usually the oil quantity distribution between
the both levels is adjusted as follows: 30% for upper and 70% for lower lubrication level. Each
level is equipped with 8 accumulator quills placed along cylinder circumference (Fig. 2a). More
detailed description, scheme and photographs of the whole lubrication system can be found in (3).
It should be noted that a new oil supply system called pulse lubricating system has been
developed lately. This system allows precise lube oil feed timing, a better circumferential oil
distribution and therefore a safe lowering of the feed rate (1). High speed, but not atomized jets of
lube oil are generated in the quills placed at one lubrication level. The impact, when the jets hit
the cylinder wall under controlled angles, is a spread of oil reaching to the adjacent quill, which
ensures that the entire cylinder liner circumference is covered (Fig. 2b).

Piston-rings-cylinder system
Studies of piston ring pack operation have drawn attention of many researchers because of
their importance in determining the mechanical efficiency, wear, fuel economy and exhaust
emissions of an internal combustion (IC) engine. The system piston rings – cylinder is generally
the greatest contributor to engine frictional losses - about 20 – 40% (4).
Piston rings are the most complicated tribological components in the IC engine to analyze
because of large variations of speed, load, temperature and lubricant availability. Sliding surfaces
of piston rings and cylinder liner may experience boundary, mixed and hydrodynamic lubrication
in one single stroke of the piston.

Gas flow
Many researchers investigated and analyzed the gas flow in the system piston-rings-cylinder
(PRC) – for example: Tian et al. (5), Koszałka (6), Ruddy et al. (7). The PRC labyrinth seal was
considered as a series of interconnected spaces by orifices (ring end gaps and additional channels
in piston grooves). The gas flow through these gaps was treated as isentropic flow corrected by a

2
flow coefficient. Due to sudden changes in pressure in the cylinder the gas flow through the PRC
system was not steady and the pressure in different volume regions was variable in time. In
addition, there were short-term return flows. The developed gas flow models contained a
complete set of leak sources, i.e. ring end gaps, including their changes due to variation of
cylinder bore caused by wear and thermal expansion. These models also considered rapid
changes of the blow-out due to axial ring movements in piston grooves caused by gas, friction
and inertial forces. Ruddy et al. (7) presented a model of the labyrinth seal of piston rings
including a gas flow in the circumferential direction. Moreover, taking into account the cylinder
ovalization, the flow area of the ring end gap was dependent on its position on the circumference
of the cylinder. The author stated that the tangential flow was essential for engines with large
cylinder diameter (over 0.5 m) and for flow rates exceeding 0.01 kg/s.

Oil flow
Detailed oil flow analyses in the PRC system were shown by Dowson (8), Priest et al. (9),
Tian et al. (10) and Jeng (4). In the majority of publications one-dimensional hydrodynamic
lubrication models of piston rings were presented. It was assumed that the size of the oil gap
between the ring and cylinder liner surface was constant on the entire circumference of the
cylinder. In order to find a distribution of the hydrodynamic pressure in the oil gap of piston ring,
the well known Reynolds equation was solved with proper boundary conditions on the inlet and
outlet side of the ring. Studies on lubrication oil consumption were performed by Gulwadi (11),
Herbst and Priebst (12) and Tian (13). The following main oil consumption mechanisms have
been described: oil evaporation from cylinder liner wall, throw-off of accumulated oil above the
top ring, oil blow through the top ring end gap into the combustion chamber due to a reverse gas
flow and oil scraping of the piston top land edge. Cavitation phenomena in oil gaps of piston
rings were investigated by Jeng (4) and later by Sawicki and Yu (14).
Two-dimensional oil flow models were developed and characterized by Bolander et al. (15),
Ma et al. (16), Yang and Keith (17, 18). In this case also the variability of hydrodynamic pressure
and oil gap on the circumference of piston rings was taken into account.

Mixed lubrication
Many researchers investigated and analyzed the mixed lubrication phenomena concerning the
operation of piston rings. Two basic papers of Patir and Cheng (19,20) included the formulation
of a fluid flow model between rough surfaces. These authors proposed a modification to the

3
Reynolds equation containing additional coefficients and terms correcting pressure distribution
and flow rates caused by the surface roughness. A complete model of elastic contact of two rough
surfaces was proposed by Greenwood and Tripp (21). The authors presented a basic theory of
interaction of randomly distributed asperities on contacting surfaces, defining relations between
statistic roughness parameters and normal and tangential forces of interaction. In most cases
asperity contact interaction were handled using stochastic model of Greenwood and Tripp (21)
with an assumption that the dominant surface roughness was near Gaussian distribution. But this
was not always the case, because the surface texture created by the honing process of cylinder
liner should be characterized by groove-like features that were much deeper than the surrounding
base surface roughness. For that reason Bolander and Sadeghi (22) developed a new model to
characterize the cross-hatch pattern of cylinder surface from the measured three-dimensional
profile. Data from this model was then used to generate numerically equivalent surface, which
was included in the deterministic mixed lubrication model. Investigations concerning the effect of
the anisotropic nature of the cylinder liner hone on friction were also conducted by Michail and
Barber (23,24).

Friction and wear


Studies on friction and wear phenomena concerning piston rings were described in lots of
papers. Effects of modern low-phosphorus engine oils on the friction and wear characteristics of
typical cylinder of a car engine were experimentally investigated by Lin, Barber et al. (25).
Molybdenum-coated piston rings were oscillated against segments of production cast iron
cylinder bore. The friction coefficients were continuously recorded, and the wear depths on the
cylinder bore segments were measured. Arumugam and Sriram (26) investigated friction and
wear characteristics of diesel engine cylinder liner–piston ring combinations under different
lubricating conditions applying a pin-on-disc wear tribometer. Experimental tests were made
using bio-lubricant and biodiesel-contaminated lubricant and commercial synthetic lubrication
oil. It was concluded that usage of newly formulated bio-lubricant and biodiesel in the long run
might have a positive impact on engine life with respect to wear and friction. An engine cylinder
liner–piston ring tribotester was used by Yu Fu Xu et al. (27) to investigate the effects of
different diesel fuels on tribological properties of a system piston ring–cylinder liner. The results
proved that this system lubricated with bio-oil had a lower friction coefficient but higher wear
mass and greater surface roughness than one lubricated with diesel oil. Moreover, in case of using

4
emulsified bio-oil a certain decrease of wear and surface roughness of the friction pairs was
observed. The authors stated that bio-oil showed promising potential because its use resulted in
lower corrosion and lower wear.
A friction model for the piston assembly of a marine four-stroke engine was developed by
Livanos and Kyrtatos (28). Calculation results of total mechanical power loss of the complete
piston assembly (rings, skirt and gudgeon pin) were presented. The effect of engine speed and
load on friction was analyzed. An influence of cylinder liner surface texturing on friction between
piston rings and cylinder in large-bore IC engines was investigated by Takata, Li and Wong (29).
A comprehensive friction power loss analysis considering all oil film–lubricated contacts in
radial and axial slider bearings as well as piston–liner contacts for a car diesel engine was
performed by Offner (30). Due to the highly nonlinear interactions, the applied mathematical
models had to consider the dynamics of the overall flexible body of the engine and also detailed
properties of lubricated contacts. The multibody software AVL-EXCITE was applied for that
purpose. Mild and severe wear mechanisms in the cylinder liner–piston ring system were
experimentally investigated by Papadopoulos, Priest and Rainforth (31). In addition to wear, the
coefficient of friction was measured in order to observe the transitions between mixed and
boundary lubrication. Priest, Dowson and Taylor (32) developed a numerical model concerning
dynamics, lubrication and wear of piston rings. It was stated that both lubrication and wear
should be considered in combination. An evolution of piston ring profile with time was presented
and its dependence on complex interactions between lubrication and wear was described.
Ma et al. (33) developed a one-dimensional wear model of the PRC system and investigated the
influence of cylinder wall temperature and surface roughness on friction and wear.
Zhang at al. (34) analyzed the hot scuffing phenomenon of a piston ring in a ring groove with
different coating materials. The experimental results indicated that the anodized aluminum piston
groove had the best scuffing resistance when a stainless steel ring was applied.

Present study
Quite often the results concerning piston rings operation of automotive engines are analyzed.
In this paper, relevant experimental and computational results for a large bore marine Diesel
engine have been presented.
Reliable operation of the system piston-ring-cylinder (PRC) is one of the key functions of a
two-stroke marine engine. The demand for improved component lifetime and predictable

5
maintenance is high. The current targets are three years time between overhauls (about 20,000
hours’ operation) as standard (2). In addition a reduction of engine operating costs is needed, i.e.
safe lowering of the lubricating oil feed rate, reduction of friction losses in order to improve
engine fuel economy. However, friction reduction in the PRC system sometimes causes an
increase in lubricating oil consumption. Therefore an optimization of the oil film thickness
between piston rings and cylinder liner should always be performed. Lowering of oil feed rates is
also required to satisfy forthcoming particulate emissions legislation for marine diesel engines.
For all these reasons an improvement of the tribological performance of the PRC system
becomes an important subject. A thorough analysis of the following phenomena is needed:
• formation and development of the oil film between piston rings and cylinder liner,
• oil losses in the combustion chamber and in the volume of piston underside,
• distribution of the oil film thickness along cylinder liner surface,
• wear of piston rings (relative to each other) that can be predicted based on the surface
asperity contact pressure.
It is believed that existing tribological problems in the marine engine industry could be solved
(among other things) by optimization of:
• piston rings geometry, particularly sliding surface profiles and also ring end gaps
influencing gas pressure differences between the rings,
• location of oil supply quills,
• shape and location of oil grooves in the cylinder liner.
All the problems presented above can be analyzed by experimental and numerical methods.
Required numerical methods are very specific, and lubrication problems can not be solved by all
purpose commercial software. A development of specified numerical codes dedicated for
application in marine engine industry is one of the key problems. Such codes have been
developed only by few scientific groups and are not widely available. Due to that fact, the
development of the own specialized program seems to be a significant achievement.
The purpose of a research presented in this paper was to predict lubrication conditions, define
areas of the possible cylinder liner wear, determine changes of the shape of piston ring surface
due to deformation and due to wear, and finally determine the gas leakage through the sealing
ring set.

6
As a future result a more reliable PRC system should be achieved with less cylinder oil
consumption for the benefit of ship owners and operators as well as for the environment.

MODELING OF PISTON RING PACK OPERATION

Developed sub-models
A combined model of piston rings operation consists of two main models: a) model of gas
flow through the labyrinth seal: piston-rings-cylinder, b) model of oil flow in the lubrication gap
between the ring and the cylinder liner. The two mentioned models are coupled. In addition, sub-
models of the following mechanical phenomena have been used: a contact of rough surfaces, an
axial movement of rings within piston grooves and an elastic torsional deformation of piston
rings (35,36,37).

Model of gas flow through the labyrinth seal of piston rings


The gas flow model consists of several volume regions V1, V2, …, V9 , which are connected
by orifices with cross-section areas A1, A2, …, A12 ( Fig. 3 ). The volumes V3, V5 and V7
correspond to volumes between the piston rings, while volumes V2, V4, V6 and V8 correspond to
groove volumes behind the rings. Orifices with cross-section areas A1, A4, A7 and A10 correspond
to the ring end gaps, whereas orifices with cross-sections A2, A3, A5, A6, A8, A9, A11 and A12
correspond to ring-side crevices.
It was also assumed that the gas flow through orifices is isentropic (depending on pressure
ratio – subsonic or sonic). The heat transfer between gas volume regions and surrounding walls
was taken into account.
Thermal expansion of the piston and the cylinder liner and wear of the cylinder liner were
also taken into account. Leaks between piston rings and cylinder liner were defined by flow areas
of ring end gaps, which depend on the position of the piston in the cylinder.
In addition, mathematical description takes into account changes of gas volume regions and
cross-section areas between the rings and the piston grooves (due to axial movement of the rings)
(2,3,4,38).
In the mathematical model of these phenomena equations of the following physical laws are
utilized (here given for a gas volume region number k):
Equation of mass balance:
dmk = ∑ dm Ini − ∑ dmOut j [1]
i j

7
Equation of energy balance:

∑ dm
i
Ini ⋅ hIni − ∑ dmOut j ⋅ hOut j − δQWall − δE fri =d (mk ⋅ u k ) + pk ⋅ dVk
j
[2]

Gas state equation in differential form:


⎛ dp dV dm k ⎞
dTk =Tk ⋅ ⎜⎜ k + k − ⎟⎟ [3]
⎝ pk Vk mk ⎠

Then the final form of these equations was formulated.


Pressure variation over the first sealing ring is commonly determined by pressure
measurements inside the cylinder during a cycle of its operation. In the presented work, variation
of the pressure inside the cylinder is obtained by numerical simulation of the physical processes
taking place inside the cylinder during a single cycle. For this purpose an in-house computer
program DIESUL (developed and verified at the engine R&D centre) was applied. The pressure
between all the piston rings was calculated using the above presented gas flow model. Due to the
relatively high pressure differences between these volumes, the gas dynamic equation including
both subsonic and sonic flows in orifices connecting volumes had to be used.
For both mentioned cases the mass flow rate through an orifice can be calculated from the
following expression:

⎧ ⎡ 2 k +1
⎤ k
⎪ p0 2k 2 ⋅ k ⎢⎛ p ⎞ k ⎛ p ⎞ k ⎥ p ⎛ 2 ⎞ k −1
⎪C D A ⋅ ⎜ ⎟ −⎜ ⎟ for >⎜ ⎟
• dm ⎪ T0 R(k − 1) k − 1 ⎢⎜⎝ p0 ⎟⎠ ⎜⎝ p0 ⎟⎠ ⎥ p0 ⎝ k + 1 ⎠
Q= =⎨ ⎣⎢ ⎦⎥ [4]
dt ⎪ k +1 k
⎪ p0 k⎛ 2 ⎞ k −1 p ⎛ 2 ⎞ k −1
⎪C D A ⎜ ⎟ for ≤⎜ ⎟
⎩ T0 R ⎝ k +1⎠ p0 ⎝ k + 1 ⎠

where: CD = Cc⋅Cv

It should be noted that the pressure losses during gas-blow process have been indirectly taken
into account by using the velocity loss coefficient Cv in Equation [4]. Only energy losses Efri
(Eq. [2]) due to wall friction in the crevices (ducts formed by the walls of piston, cylinder and
neighboring rings) have been neglected. Such a simplification was justified by very small ring
end gaps used in the engine model under consideration. According to the publication (7)
of Ruddy et al., a significant influence of circumferential gas flow (friction flow) appears for the

8
ring gap size higher than 5 mm for similar type of engine. In the current analysis the maximal
ring end gap size only temporarily reached this limit.
Model of oil flow
Oil supply system of two-stroke engine
The model of an oil supply system concerning two-stroke marine engines takes into account
an inflow of lubricating oil through the cylinder liner onto the running surface. The oil is
fed at defined heights of the cylinder liner: the upper level at approximately 10% and the lower
level at 20% engine stroke from piston top dead centre (TDC).
An increase of the oil film thickness Δhlev (μm) due to oil supply at the certain level of
cylinder liner can be calculated in every two-stroke cycle as follows:

xlev ⋅ g oil ⋅ Pcyl


Δhlev = [5]
6 ⋅π ⋅ n ⋅ D ⋅ ρ ⋅ dq

where: xlev.= 30% for upper or xlev.= 70% for lower lubrication level.

It should be noted that oil distribution xlev to upper and lower level can be selected depending on
need.
Oil flow in a gap (with rough surfaces) between the ring and cylinder

Two main cases of oil flow in the system piston ring – cylinder liner are presented in Fig. 4.

A one-dimensional form (axi-symmetric oil flow model) of the average Reynolds equation
developed by Patir and Cheng (19,20) has been used to calculate hydrodynamic forces in the case
of rough gap surfaces. This equation is applicable to any general roughness structure and takes
the following form:
⎛ _
⎞ _ _
∂ ⎜ h 3 d p ⎟ u d h T u dφ S d h T
φx = + σ + [6]
∂ x ⎜⎜ 12μ d x ⎟⎟ 2 dx 2 dx dt
⎝ ⎠
The significance and mathematical description of empirical coefficients φx , φS and boundary
conditions of equation [6] are presented in (19,20).

Model of elastic contact between ring and liner

The effects of interacting asperities of piston ring and cylinder liner surfaces were modeled
using the mathematical model developed by Greenwood and Tripp (21). The normal contact
force is described by the following relation:

9
2 σ xr ⎛h⎞
Fc = 16 π (ηβσ ) 2 E ' ⋅ ∫ F5 / 2 ⎜ ⎟dx [7]
15 β xl ⎝σ ⎠
Integral boundaries xl and xr (Fig. 4b) define the possible contact area xl ≤ x ≤ xr , in which the

following relation is fulfilled: h/σ ≤ 4. The form of function F5/2 and all the components of the
presented equations can be found in works by Greenwood and Tripp (21). The friction force due
to asperity interaction was calculated by the following equation:
Fcx = τ 0 Ac + α 2 Fc [8]
where parameters: τ0 = 2·106 Pa and α2 = 0.08 (Gui & Liu (39)).
A calculation method of various surface roughness related parameters (used in Equation [7])
can be found in the first part of the Appendix.
Asperity contact interactions were handled using stochastic model of Greenwood and Tripp
(21) with an assumption that the dominant surface roughness was near Gaussian distribution.
Until recently, it was the most common method of modeling mixed and boundary lubrication
regimes in the system piston-ring-cylinder. Presently used flow and stress factors in Eq. [6] and
[7] are not most suitable for the plateau surface finish. However, at the time of development of
the mathematical model, the only available experimental data concerning surface roughness was
that described by the Gaussian distribution. It is believed that this approach is acceptable for the
purpose of this study. A new method of three-dimensional surface roughness modeling such as
presented in (22, 40) seems to be more appropriate for future studies.

Model of ring torsional deformation and ring axial movement in the piston groove

A scheme of forces acting on a piston ring, action lines and distances between these forces
and the centre of gravity S of the ring cross-section are shown in Fig. 5. All the forces are
referenced to unit circumference of the piston ring (unit forces (N/m)).
Typical set of equations for the piston ring has the following form:

a) in radial direction
ΣFr =Fh +Fc +Fygi +Fygi +2 −Fspr−Fgas=0 [9]
b) in axial direction
m
ΣFx = Rx − Ffri − Fcx + Fgi+1 + Fgi+2 − Fgi − (g + ap ) = 0 [10]
cimc

10
Using these equations, one can calculate the reaction force Rx between the ring and piston groove
in every time step. If the sign of this force changes, the axial movement of the ring in the piston
groove begins. At this point, the value of the reaction force Rx = 0 and the axial movement of the
ring relative to the piston groove can be described by the following differential equation:
m d 2xr mg
2
= −Ffri − Fcx + Fgi+1 + Fgi+2 − Fgi − [11]
cimc dt cimc

The ring movement xr completes when the ring reaches the opposite side of the piston groove.
Similarly like in articles of Tian et al. (5), Dowson (8) and Ma et al. (16), the twist around the
centre of gravity of the ring cross-section (point S in Fig. 5) can be described by the following
equation of equilibrium of acting moments:

( ) ( )
ΣM S = Fh xS − xFh + Fc xS − xFc − (F fri + Fcx )
Ar
2
− Fg i ( y2 + y sc − y1 ) − Fg i+1 ( y3 − y sc ) +
[12]
⎛y ⎞
+ Fg i+2 ⎜ 2 + y sc ⎟ + Rx ⋅ ysc − K ⋅θ = 0
⎝ 2 ⎠
Estimating the ring torsional stiffness K as described in (35) and using the equation [12]
of moment equilibrium, one can calculate the ring twist angle θ. All the piston grooves are
assumed to have curved surfaces in the radial direction (Fig. 5) due to their typical wear profiles.

NUMERICAL METHOD
The numerical solution of the Patir&Cheng equation [6] is based on the implicit finite
difference scheme. The applied method is described in the second part of the Appendix.

EXPERIMENTAL VERIFICATION OF THE MODEL

Preliminary evaluation of the model and computer code

The general properties of the developed model and computer program concerning
hydrodynamics and oil flow were verified for chosen calculation examples. It was carried out by
a comparison of calculation results from an in-house computer program of an engine R&D centre
with the results of a program developed by the author. This was only preliminary evaluation of
the developed model. It was quite natural to compare the simulation results of both mentioned
programs, even though that a new program version had better features than an old one. More
detailed description of this comparison can be found in the article (41).

11
Examples of the hydrodynamic system: scraper rings – piston rod were analyzed. A very
good agreement between both calculation results was achieved. The maximal relative differences
did not exceed several percent (41).
The main parts of the verification of the developed model and computer code are presented in
the subsections entitled “Experimental verification of computational results”.

Determination of model parameters


Basic engine technical data for this study were provided by engine development group. These
data includes: major dimensions, rotational speed at maximum power and geometric parameters
of the system piston-rings-cylinder, from dimensions and masses to required microprofiles and
roughnesses (see Tables 4 and 5, section “Calculation results”).
Major parameters of the rough structure of the liner and ring surface are presented in Table 1.
A calculation method of combined parameters is shown in the Appendix.
The roughness parameters of sliding surfaces of piston rings were determined using
mechanical contact measurement method, namely a standard surface roughness tester. The
roughness of cylinder liner was non-destructively evaluated and in detail analyzed by the
application of new optical measurement methods. It was possible at any stage of the life-time of
cylinder liner by the use of rubber compound replicas of the liner surface (2).
Table 1. Surface roughness parameters
Surface data parameters Cylinder liner Piston rings
RMS roughness σ1 = 0.22 μm σ2 = 0.044 μm
Elastic modulus E1 = 1.13⋅1011 N/m2 E2 = 1.5⋅1011 N/m2
Poisson’s ratio νP1 = 0.26 νP2 = 0.25
Combined parameters Cylinder liner and piston rings
RMS roughness σ = 0.224⋅10-6 m
Asperity density η = 1.114⋅1012 m-2
Asperity radius of curvature β = 0.2⋅10-6 m

The engine development group provided also data on: thermomechanical deformations of
piston and cylinder liner; wear of cylinder liner after 1000 hours of operation; parameters of
engine thermodynamic cycle such as gas pressure and temperature in the cylinder and in
crankcase, average temperatures of piston rings, and thermal state of piston and cylinder surface.
However, detailed data concerning thermal state of cylinder liner and piston and their
thermomechanical deformations was proprietary information of the engine development company
(see Acknowledgments). Therefore only dimensionless relations could be presented in Fig 6.

12
In addition, the engine development group provided estimation of physical properties of
engine oil, including viscosity, and quantity of oil lubricating the cylinder liner.
The temperature-dependent oil viscosity ν(TC) was evaluated by the use of following equation:
log10 [log10 (ν + 0.8)] = Av − Bv log10 TC [13]
where parameters: Av = 8.56 and Bv = 3.28
The assumed oil density ρ = 940 kg/m3 and the lubricating oil feed rate at full engine load
goil = 1 g/kWh.
Substantial information has been obtained from literature (6, 38). In particular, it concerned
thermodynamic and flow parameters, such as gas flow coefficients through the canals of the
labyrinth piston – rings – cylinder, heat transfer coefficients between gas and walls of this
labyrinth, etc. The ultimate tuning of these parameters was done by utilizing compatibility criteria
between numerical calculations and experimental results presented in the next Section.
Established values of gas flow coefficients are presented in Subsection “Variations of gas
pressure between piston rings”. The assumed value of heat transfer coefficients between gas and
walls of the PRC system equaled 400 W/(m2K).

Compatibility criteria between simulation and experimental results

As the criteria of compatibility between numerical simulation and experimental results the
following parameters were assumed. They mainly concern function variations versus crank angle
of an internal combustion engine.
a) Mean square deviation of measured and calculated function:

1 n
σ= ∑[ f meas ( xi ) − fcalc ( xi )]2
n i =1
[14]

where:
f meas – value of compared function for i – th value of argument x (for instance: time t, crank
angle α, x - coordinate along cylinder liner) obtained from measurements,
f calc – value of compared function for i – th value of argument x obtained from numerical
calculation,
n – number of points of analyzed function fmeas/calc(x).

13
b) Maximum value of differences between measured and calculated functions:
n
Δf max = max [ f meas ( xi ) − f calc ( xi )] [15]
i =1

c) Pearson’s linear correlation coefficient of measured and calculated function:


n

∑[ f meas ( xi ) − f meas
av
] ⋅[ f calc ( xi ) − f calc
av
]
rmeas / calc = i =1
[16]
n n

∑[ f
i =1
meas ( xi ) − f meas] ⋅ ∑ [ f calc ( xi ) − f
av 2

i =1
av 2
calc ]

where:
av av
f meas – average value of measured function; f calc – average value of calculated function.

Experimental verification of computational results

The engine used for the verification was a turbocharged two-stroke marine Diesel engine
(1,2,42). The engine was equipped in two seal systems that were included for experimental
verification:
a) labyrinth seal of piston rings,
b) pack of seal and scraper rings making a gland-box that separates the piston underside
from crankcase of the engine.

Variations of gas pressure between piston rings

Measurements of unsteady gas pressure in the cylinder, between the piston rings and under
piston were carried out at the R&D engine center using piezoelectric sensors mounted in the
piston. Then the experimental results were made available to the author of the manuscript, so that
the verification of the developed model could be performed. It should be noted that similar
measurements performed at the same global company for a marine four-stroke engine have been
thoroughly described by Tamminen et al. (43).
The piston of the engine (bore D=580 mm, stroke S=2416 mm, rotational speed n=105 rpm)
has four piston rings (Fig. 3). In Fig. 7 a comparison of measured and calculated gas pressures pi
(i = 1, 3, 5, 7) for the main volume regions of the labyrinth seal (Fig. 3) as a function of crank
angle is shown. The presented results correspond to full engine load.
Parameters that characterize this comparison are presented in Table 2. A satisfactory
qualitative and quantitative compatibility of the analyzed pressure variations has been achieved.

14
The relative differences between maximal values of gas pressures (measured and calculated) have
not exceeded 15%. Axial movements of rings in piston grooves have not been measured for this
engine.
Based on experimental verification of the gas flow model in the labyrinth seal of piston rings,
the value of the orifice discharge coefficient CD = Cc⋅Cv = 0.75 has been determined. Both flow
coefficients: Cc , Cv (of contraction and velocity loss) were selected, because their physical
meaning were precisely described in fluid mechanics. It was assumed that Cc = 0.83 and
Cv = 0.9. For the numerical simulation of gas flow the value of discharge coefficient CD
is important (Eq. [4]).

Table 2. Compatibility of gas pressure variations (measured and calculated)

Relative Ratio of average Ratio of mean Pearson’s corre-


Evaluated difference values square deviation to lation coefficient
variations between maximal (of calculated measured maximal (of measured
of gas values (measured and measured value and calculated
pressures and calculated) results) results)
(%) (–) (%) (–)
p1 -3.80 1.077 3.20 0.997
p3 7.20 1.012 1.94 0.998
p5 -12.13 1.118 5.85 0.994
p7 9.07 0.962 5.74 0.981

Volumes of scraped oil by gland-box of marine engine


Experimental verification of the hydrodynamic model of piston rings involved measurement
results of scraped oil volumes by a gland-box in the two-stroke marine engine (bore D=840 mm,
stroke S=2400 mm, rotational speed n=97.5 rpm). The purpose of a gland-box (Fig. 8) is to
separate the piston underside as tight as possible from the crankcase. Special oil tanks were
installed in the tested engine, where lubricating oil flew in. Then, oil volumes scraped up to
piston underside and down to crankcase during 24 hours of engine operation were measured. The
measurements were carried out at the R&D engine center. Then the experimental results were
made available to the author of the manuscript, so that the verification of the developed model
could be performed.
One of the input parameters for the simulation program is the rate of oil mist deposition on a
free surface of piston rod. Value of this parameter has been selected in such a way that a good

15
agreement between measured and calculated oil volumes scraped down into crankcase has been
achieved. In Table 3, a comparison of measured and calculated parameters and a percent scatter
of results are given. The maximal relative differences between measured and calculated values
have not exceeded 10%.

Table 3. Compatibility of the scraped oil quantities (measured and calculated) by the gland-box

Average value of Calculated Relative difference


Analyzed volumes measured parameter value between measured
of lubricating oil and calculated value
(dm3/24h) (dm3/24h) (%)
Quantity scraped to piston
8 7.252 9.35
underside
Quantity scraped to crankcase 12000 11020 8.17

CALCULATION RESULTS

The computer program incorporating the presented models has been used for simulation of
two-stroke Diesel engine (Tab. 4) operating at full load. The type of ring set considered is
common in marine engines. The piston ring pack consists of four rings (Figs. 3 and 9). The
package includes conventional straight ring end gaps.

Table 4. Main data of the marine engine under consideration

Cylinder bore (mm) 580


Piston stroke (mm) 2416
Engine rotational speed (rpm) 105

The surface geometry of the piston ring package, with vertical dimensions magnified by
factor of 1000 relative to the horizontal ones, is depicted in Fig. 9. All the rings have the same
asymmetrical barrel shape (Tab. 5).

Table 5. Basic geometric parameters of piston rings: 1, 2, 3, 4


Axial height of piston ring H = 16 mm
Radius of parabolic sliding surface R = 750 mm
Offset of parabolic sliding surface Of = 12 mm
Distance between piston rings Lp = 18 mm

In order to ensure very low wear of profiled surfaces, the piston rings are coated (for example
the top ring has chromium ceramic coating) (2,42). As a consequence, even hydrodynamic

16
conditions during a long period of piston rings operation can be ensured. It should be noted that
in new engine types another shape of piston ring surfaces (2, 3 and 4), namely symmetrical barrel
shape, is used (2).
Typically, the figures that will follow show variation of some physical parameters as a
function of the crankshaft rotation angle, beginning from the piston bottom dead centre (BDC) of
the two-stroke engine operation (0°). In this case the end of compression phase is at 180° of crank
angle (piston top dead centre - TDC).
Fig. 7 (shown in the last Section) presents the pressure variation between piston rings,
calculated on the basis of known cylinder pressure variation and simulated gas leakage through
the labyrinth sealing of the piston ring pack (orifices corresponding to ring end gaps and ring-side
crevices – Fig. 3). Generally, the gas pressure in the cylinder and all the inter-ring gas pressures
increase during the piston upstroke (compression phase) and decrease during a certain part of the
piston downstroke (expansion phase). In Fig. 7 the following maximum gas pressure values can
be seen: nearly 160 bar in cylinder, over 80 bar between the 1st and 2nd piston ring, about 50 bar
between the 2nd and 3rd ring and approximately 30 bar between the 3rd and 4th ring. In some cases
during the expansion phase the gas pressure under the first (or second) piston ring can reach a
higher value than the gas pressure over this ring. For this reason axial ring lifts in piston grooves
are anticipated (Fig. 10).
On the piston underside a relatively low scavenging air pressure is noticed. At the piston ring
pack location near scavenging ports (i.e. between 300° and 310° of crank angle) all inter-ring gas
pressures are visibly reduced (Figs. 7 and 11).
The higher is the gas pressure the stronger is the radial gas force acting to increase the ring
diameter. It means that the radial gas force can be many times greater than the natural force due
to ring stiffness acting in the same direction. First of all, the first ring (top ring) is strongly
pressed against the cylinder liner surface. The higher is the ring number (Figs. 3 and 9), meaning
more distant from the top, the less is the ring loaded (compare Figs. 7 and 11).
In Fig. 10 ring lifts in piston grooves as a function of crank angle are shown. Gas pressure
variations at the location of scavenging air ports are responsible for short lasting lifts of piston
rings 1, 2 and 3 during the piston upstroke. Much more important are ring lifts during the piston
downstroke (expansion period). Due to gas pressure variations between the piston rings, axial
movements of piston ring 1 and 2 in piston grooves can be observed, respectively in the range

17
between 220° and 290° of crank angle for the first one and between 245° and 280° of crank angle
for the second one. The other piston rings (number 3 and 4) are strongly pressed against the lower
flank of piston groove and do not move. Axial inertial forces acting on the piston rings are
significantly less important for low-speed marine engines than for automobile engines due to
much lower piston acceleration.
The nominal axial clearance of each ring in piston groove equals 0.35 mm. Each short-lasting
ring movement in the piston groove is followed by a change of the acting point of the reaction
force Rx to the other flank of piston groove and also a sign change of this force
(Fig. 5). Obviously, Rx=0 during the ring movement between the two piston groove flanks.
Rapid ring lifts are accompanied by short-lasting, but strong changes of gas flow areas
between rings and piston grooves. They are much stronger than flow areas of ring end gaps,
because the opening ranges include the whole piston circumference. In this case temporary rapid
changes of gas blow-out should be expected.
The hydrodynamic force acts in the radial direction on the ring and is counteracted by the
spring force, gas force and friction force in the piston groove. The inertia force in the radial
direction has been neglected due to very small values of the radial ring acceleration. Fig. 11
presents the hydrodynamic force for each piston ring necessary for compensating both the gas
pressure and radial forces resulting from the ring stiffness. The variations of hydrodynamic forces
look similar to variations of inter-ring gas pressures (Fig. 7).
Additionally, in the mixed lubrication cases the elastic radial contact forces are anticipated. In
Fig. 12 variations in radial components of elastic contact forces are shown. These forces occur in
the case of a high gas pressure and low oil viscosity caused by high temperature near the top dead
centre. It should be noticed that the values of elastic contact forces are much lower than
hydrodynamic forces acting on rings (compare Figs. 11 and 12).
The hydrodynamic tangential forces as functions of the crankshaft rotation are presented in
Fig. 13. These forces significantly depend on piston velocity. For this reason the highest values of
hydrodynamic tangential forces occur at crank angles, where the maximum piston velocity is
developed. These forces could be neglected in the piston motion phases corresponding to low
velocity near the reverse points. Near the top dead centre (at high oil temperature) the additional
tangential components of elastic contact forces (friction forces) should be noticed (Fig. 14). The
sign change of these forces at TDC results from the sign change of piston velocity.

18
The oil film thickness was calculated from the pressure distribution in the gap between the
ring land and cylinder liner. In addition, the temperature variation along the cylinder liner results
in differences in local oil viscosity values. In order to take this effect into account, the measured
temperature distribution along the liner was applied for numerical calculation of hydrodynamic
lubrication of piston ring pack.
The motion of the ring pack scraping and distributing oil on the cylinder liner leaves the oil
film profile shown in Fig. 15. This profile is formed after a few cycles of operation.
An uneven oil film distribution along the cylinder liner can be clearly seen. Low film
thickness near TDC and in the other part of cylinder liner at the location of scavenging air ports
should be noticed. The minimum oil film thickness at TDC is about 0.2÷0.3 μm and is
comparable with root mean square (RMS) roughness of the cylinder liner that equals 0.22 μm.
The very low local film thickness values near TDC can be explained by occurrence of high gas
pressure and high temperature in this area during the compression and working phases of engine
operation. Due to high gas forces piston rings are strongly pressed against the cylinder surface.
On the other hand, high temperature reduces the oil viscosity.
There are two places of oil supply for the cylinder liner of long-stroke IC engine located
below TDC. Two peaks of oil film thickness at these places can be clearly seen in Fig. 15.
In a marine two-stroke engine the location of scavenging air ports is also important for
cylinder lubrication. Their presence simply reduces the area of the mating surface between piston
rings and cylinder liner. In this area a simplified approach was applied. The oil film thickness
was assumed to be reduced at this location reflecting the reduced sliding surface.
Due to low gas pressure and oil temperature the greatest oil film thickness can be seen
between scavenging air ports and the bottom dead centre.
One important computational problem is the definition of the boundaries x1 , x2 of the ring
wetted area (see Fig. 4). After a series of piston operation cycles, each ring scrapes and
accumulates excessive oil, leaving behind an oil film not sufficient for full lubrication. The
variation of wetted area has an essential influence on the hydrodynamic bearing force of the ring
and the resulting radial ring velocity. Oil film thickness, piston velocity, ring stiffness and ring
surface roughness parameters have significant influence on changes of wetted area boundaries
and areas of direct contact of the ring and cylinder liner surfaces in the case of mixed lubrication.

19
Most often rings are only partially wetted in phases of high piston velocity. Rings are fully wetted
in the piston motion phases corresponding to low velocity near the reverse points.
Fig. 16 depicts changes of the wetted area boundaries of the first piston ring. Oil is mostly
concentrated in the central part of the ring profile. In the rear part the hydrodynamic pressure
drops and this area is ventilated by the gas outside the ring. In the case of very thin oil film, the
wetted area is also reduced in the upstream part of the ring surface. The sudden rise in the oil
wetted area for the region from 150° to 165° of crank angle is due to oil supply (see also Fig. 15).
Boundaries of the direct contact of contributing surfaces xl , xr (see Fig. 4) are also depicted in
Fig. 16. The contact boundaries have been defined for h/σ = 4 (19,20,21). Characteristic feature
of the contact zone is a small area at TDC (at 180° of crank angle).
Zones of wetting and direct contact for the second ring are presented in Fig. 17. Differences
in comparison with the first ring can be recognized at the location of oil supply area. Wetted and
contact zones for the third and fourth piston ring (not shown) look similar.
Variation of the twist angle for each ring is shown in Fig. 18. Maximum values of the twist
angle appear in the phases of piston motion corresponding to high gas pressure in the combustion
chamber and reach the following values: -5.5’ (minutes) for the 1st (top) ring, -7’ for the 2nd ring,
-6’ for the 3rd ring and -4.5’ for the 4th ring. Each ring twists in such a way that the gap
deformation causes the decrease of hydrodynamic force during upstroke and a higher amount of
oil scraping to combustion chamber is observed (35).
Due to axial movements of the 1st and 2nd piston ring during the piston downstroke
(expansion period) the resulting significant variations of ring twists can be observed (compare
Figs. 10 and 18). In addition, sudden decrease of gas pressure at the location of scavenging air
ports (Fig. 7) causes further substantial variations of ring twists.
It should be noted that full elastohydrodynamic phenomenon, involving both local distortion
of the elastic solids and the influence of pressure upon viscosity has not been analyzed. However,
the developed software has an option concerning influence of pressure upon lubricant viscosity –
well-known relation of Barus (18). In addition, elastic angular deformations (twists) of piston
rings have been taken into account (see Fig. 18). Maximum values of the twist angles occur at
high gas pressure in the combustion chamber, i.e. some degrees after TDC.

20
CONCLUSIONS AND REMARKS

The major conclusions that may be drawn from the results are as follows:

1. The model and simulation program have been successfully experimentally verified for
marine engines. The model adequately represents the operation of the piston ring pack, in
particular in terms of variations of gas pressure between the rings as functions of crank
angle. The assumed compatibility criteria between simulation and experimental results
concerning gas pressure variations have been satisfied (Tab. 2);
2. Examination of scraped oil volumes by the ring pack (of the gland-box of marine internal
combustion engine) proves a satisfactory quantitative agreement between numerical and
experimental results. The maximal relative differences between measured and calculated
values have not exceeded 10% (Tab. 3);
3. Elastic radial contact forces due to surface roughness occur in the area of piston TDC.
These forces are relevant to high gas pressure and low oil viscosity caused by high
temperature. However, the values of elastic contact forces are much lower than
hydrodynamic forces acting on piston rings (compare Figs. 11 and 12);
4. The results (Fig. 11 and 12) indicate that hydrodynamic forces are generated by relatively
low pressure acting on a large surface in contrast to high local contact pressure concentrated
on a very small area of elastic contact. Due to that fact, the elastic contact seems to be
responsible for wear process despite the low speed of the piston close to TDC (see Figs 6,
16 and 17);
5. The higher is the ring number (Figs. 3 and 9), meaning more distant from the top, the less
radial and tangential contact force is observed (Figs. 12 and 14). Consequently the first
(top) piston ring can be expected to experience the most intensive wear of the sliding
surface. Wear phenomena of other rings are less profound;
6. The most severe lubrication conditions occur close to TDC towards the end of the
compression and beginning of expansion stroke, especially for the first (top) ring. Due to
high temperature, the oil viscosity is very low and, consequently, the oil film along the
cylinder liner is very thin: 0.2 ÷ 0.3 μm at TDC (Fig. 15). In this case it is essential that a
model of mixed lubrication accounting for the surface roughness is used (RMS = 0.22 μm
for the cylinder liner of tested engine);

21
7. The influence of the cylinder liner temperature on oil viscosity decreases towards BDC
(Fig. 6) and consequently the increase of oil film thickness on the lower part of the cylinder
liner can be noted (Fig. 15). The simulation results indicate more stable lubrication
conditions on the lower part of the cylinder liner than on the upper part. Only in the
scavenging air ports location and close to BDC significant decreases of the oil film
thickness can be seen – respectively to the values of about 2 μm and 3 μm. It suggests that
not only near TDC, but also in the mentioned areas a relatively intensive wear of cylinder
liner can be expected;
8. It is likely that the oil film is unstable over the circumference of cylinder liner in narrow
places between scavenging air ports. In addition, intensive air flow in this area takes place
periodically, which might destroy the oil film development in the vicinity of the air ports.
Consequently, wear modes such as scratched liner above the ports (42) (concerning
previous engine types) could be explained;
9. The calculation results indicate that the twist movement of the piston rings (Fig. 18) should
be taken into account if lubrication conditions of cylinder liner and optimization of sliding
surfaces of piston rings are analyzed;
10. Oil supply places (Figs. 2a and 15) seems to be located suitably.

ACKNOWLEDGEMENTS
The author expresses his gratitude to Wärtsilä’s R&D centre in Winterthur (Switzerland) for
having the opportunity to work on projects concerning mathematical modeling and numerical
simulation of tribological systems of piston rings and gland-box during several research periods
at this company.
REFERENCES
(1) “Wärtsilä Technology Review”, information materials concerning IC engines designed at Wärtsilä
company.
(2) Räss, K., Amoser, M. (2007), “Progressive development of two-stroke engine tribology”, Paper No.
83, CIMAC Congress, Vienna.
(3) “Cylinder Lubrication System CLU 3 for 2-Stroke Crosshead Large Diesel Engines”, Vogel
Company, http://www.grouphes.com/lubemec/Vogel/dateien/us/pdf/izs_pdf/201-US.pdf accessed
August 21, 2013.
(4) Jeng, Y. R. (1992): “Theoretical Analysis of Piston-Ring Lubrication. Part II – Starved Lubrication
and Its Application to a Complete Ring Pack”, Tribology Transactions, Vol. 35, Iss. 4, pp. 707-714.

22
(5) Tian, T., Nordzij, L. B., Wong, V. W., Heywood, J. B. (1998), “Modeling Piston-Ring Dynamics,
Blowby, and Ring-Twist Effects”, Transactions of ASME, Journal of Engineering for Gas Turbines
and Power, Vol. 120, pp. 843-854.
(6) Koszałka, G. (2010), “Application of the piston-rings-cylinder kit model in the evaluation of
operational changes in blowby flow rate”, Eksploatacja i Niezawodnosc – Maintenance and
Reliability 4 (48), pp. 72-81.
(7) Ruddy, B. L., Dowson, D., Economous, P. N. (1981), “The Prediction of Gas Pressures within the
Ring Packs of Large Bore Diesel Engines”, Jour. Mechan. Eng. Science, Vol. 23, No. 6, pp. 295-
304.
(8) Dowson, D. (1993), “Piston Assemblies; Background and Lubrication analysis”, Engine Tribology,
pp. 213-240, Taylor C. M. (editor), Elsevier Science.
(9) Priest, M., Taylor, R. I., Dowson, D., Taylor, C. M. (1996), “Boundary Conditions for Reynolds
Equation with Particular Reference to Piston Ring Lubrication”, Tribology Series, vol. 31,
pp. 441-452.
(10) Tian, T., Wong, V. W., Heywood, B. (1996), “A Piston Ring Pack Film Thickness and Friction
Model for Multigrade Oils and Rough Surfaces”, SAE Paper, No. 962032, pp. 27-39.
(11) Gulwadi, S. D. (2000), “Analysis of Tribological Performance of a Piston Ring Pack”, Tribology
Transactions, Vol. 43, Issue 2, pp. 151-162.
(12) Herbst H., Priebsch, H. (2000), “Simulation of Piston Ring Dynamics and Their Effect on Oil
Consumption”, SAE Paper No. 2000-01-0919.
(13) Tian, T. (2002), “Dynamic Behaviors of Piston Rings and Their Practical Impact - Part II: Oil
Transport, Friction, and Wear of Ring/Liner Interface and the Effects of Piston and Ring Dynamics”,
Proc. I. Mech. E, Part J: Journal of Engineering Tribology, Vol. 216, pp. 229-247.
(14) Sawicki, J. T., Yu, B. (2000): “Analytical Solution of Piston Ring Lubrication Using Mass
Conserving Cavitation Algorithm”, Tribology Transactions, Vol. 43, Issue 3, pp. 419-426.
(15) Bolander, N. W., Steenwyk, B. D., Sadeghi, F., Gerber, G. R. (2005), “Lubrication Regime
Transitions at the Piston Ring - Cylinder Liner Interface”, Proceedings of the Institution of
Mechanical Engineers, Part J: Journal of Engineering Tribology, Vol. 219, pp. 19-31.
(16) Ma, M.-T., Sherrington, I., Smith, E. H., Grice, N. (1997), “Development of a detailed model for
piston-ring lubrication in IC engines with circular and non-circular cylinder bores”, Tribology
International, Vol. 30, No. 11, pp. 779-788.

(17) Yang, Q., Keith, T. G. (1996): “Two-Dimensional Piston Ring Lubrication – Part II: Elastic Ring
Consideration”, Tribology Transactions, Vol. 39, Iss. 4, pp. 870-880.

23
(18) Yang, Q., Keith, T. G. (1996): “Two-Dimensional Piston Ring Lubrication – Part I: Rigid Ring and
Liner Solution”, Tribology Transactions, Vol. 39, Iss. 4, pp. 757-768.
(19) Patir, N., Cheng, H. S. (1978), “An Average Flow Model for Determining Effects of Three-
Dimensional Roughness on Partial Hydrodynamic Lubrication”, Transactions of ASME, Vol. 100.
(20) Patir, N., Cheng, H. S. (1979), “Application of Average Flow Model to Lubrication between Rough
Sliding Surfaces”, Transactions of ASME, Vol. 101.
(21) Greenwood, J., Tripp, J. H. (1971) “The contact of Two Nominally Flat Rough Surfaces”, Proc. Inst.
Mech. Eng., Vol. 185, pp. 625-633.
(22) Bolander, N. W., Sadeghi, F. (2007): “Deterministic Modeling of Honed Cylinder Liner Friction”,
Tribology Transactions, Vol. 50, Issue 2, pp. 248-256.
(23) Michail, S. K., Barber, G. C. (1995): “The Effect of Roughness on Piston Ring Lubrication – Part I:
Model Developement”, Tribology Transactions, Vol. 38, Iss. 1, pp. 19-26.
(24) Michail, S. K., Barber, G. C. (1995): “The Effect of Roughness on Piston Ring Lubrication – Part II:
The Relationship between Cylinder Wall Surface Topography and Oil Film Thickness”, Tribology
Transactions, Vol. 38, Iss. 1, pp. 173-177.
(25) Lin, P., Barber, G., Zou, Q., Anderson, A. H., Tung, S., Quintana, A. (2008), “Friction and Wear of
Low-Phosphorus Engine Oils with Additional Molybdenum and Boron Compounds, Measured on a
Reciprocating Lubricant Tester”, Tribology Transactions, Vol. 51, Issue 5, pp. 659-672.
(26) Arumugam, S., Sriram, G. (2012): “Effect of Bio-Lubricant and Biodiesel-Contaminated Lubricant
on Tribological Behavior of Cylinder Liner–Piston Ring Combination”, Tribology Transactions,
Vol. 55, Issue 4, pp. 438-445.
(27) Yu Fu Xu , Hui Qiang Yu , Xiao Yang Wei , Zheng Cui , Xian Guo Hu , Teng Xue, Dan Yang
Zhang (2013): “Friction and Wear Behaviors of a Cylinder Liner–Piston Ring with Emulsified Bio-
Oil as Fuel”, Tribology Transactions, Vol. 56, Issue 3, pp. 359-365.
(28) Livanos G.A., Kyrtatos N. P. (2007): “Friction model of a marine diesel engine piston assembly”,
Tribology International, Vol. 40, pp. 1441–1453.
(29) Takata, R., Li, Y., Wong, V. W. (2006), “Effect of Liner Surface Texturing on Ring/Liner Friction
in Large-Bore IC Engines”, Paper ICEF2006-1525, Proc. of ICEF06 ASME Internal Combustion
Engine Division 2006 Fall Technical Conference, Sacramento, California, November 5-8, 2006.
(30) Offner, G. (2013): “Friction Power Loss Simulation of Internal Combustion Engines Considering
Mixed Lubricated Radial Slider, Axial Slider and Piston to Liner Contacts”, Tribology Transactions,
Vol. 56, Issue 3, pp. 503-515.
(31) Papadopoulos, P., Priest, M., Rainforth, W. M. (2007), “Investigation of fundamental wear
mechanisms at the piston ring and cylinder wall interface in internal combustion engines”,

24
Proceedings of the Institution of Mechanical Engineers, Part J: Journal of Engineering Tribology,
vol. 221, pp. 333-343.
(32) Priest, M., Dowson, D., Taylor, C. M. (1999), “Predictive Wear Modelling of Lubricated Piston
Rings in a Diesel Engine”, Wear, vol. 231, pp.89-101.

(33) Ma, Z., Henein, N. A., Bryzik, W. (2006), “A Model for Wear and Friction in Cylinder Liners and
Piston Rings”, Tribology Transactions, Vol. 49, Issue 3, pp. 315-327.
(34) Zhang, W., Becker, E., Wang, Y., Zou, Q., Zhou, B., Barber, G. (2008): “Investigation of Scuffing
Resistance of Piston Rings Run against Piston Ring Grooves”, Tribology Transactions, Vol. 51,
Issue 5, pp. 621-626.
(35) Wolff, A., Piechna, J. (2007), “Numerical simulation of piston ring pack operation with regard to
ring twist effects”, The Archive of Mechanical Engineering, Vol. LIV, No. 1, pp. 65-99.
(36) Wolff, A. (2009), “Numerical analysis of piston ring pack operation”, Combustion Engines / Silniki
spalinowe, No. 2, p. 128-141.
(37) Wolff, A. (2012), “Influence of engine load on piston ring pack operation of a marine two-stroke
engine”, Journal of KONES Powertrain and Transport, Vol. 19, No. 2, pp. 557-569.
(38) Niewczas, A., Koszałka, G., Guzik, M. (2006), “Modeling of collaboration between the piston ring
and the piston groove shelf in an internal combustion engine”, Eksploatacja i Niezawodnosc –
Maintenance and Reliability, No. 4 (32), pp. 82-86.
(39) Gui, C.L., Liu, K. (1992), “Effect of surface roughness on the lubrication properties of the piston
ring and cylinder of an engine, and calculation of lubrication and power loss analysis of piston-ring
pack of a S195 diesel engine”, Lubrication Science 4-4 (4) 263 0954-0075.
(40) Jocsak, J., Tian, T., Li, Y., Wong, V., W. (2006), “Modeling and Optimizing Honing Texture for
Reduced Friction in Internal Combustion Engines”, SAE Paper No. 2005-01-1641.
(41) Wolff, A. (2009), “Experimental verification of the model of piston ring pack operation of an
internal combustion engine”, The Archive of Mechanical Engineering, Vol. LVI, No. 1,
pp. 73-90.
(42) Demmerle, R., Barrow, S., Terrettaz, F., Jaquet, D. (2001), “New Insights into the Piston Running
Behaviour of “Sulzer” Large Bore Diesel Engines”, CIMAC Congress, Hamburg.
(43) Tamminen, J., Sandström, C.-E., Andersson, P (2006), “Influence of load on the tribological
conditions in piston ring and cylinder liner contacts in a medium-speed diesel engine”, Tribology
International, Vol. 39, pp. 1643–1652.
(44) Gelinck, E. (1999), “Mixed Lubrication of Line Contacts”, Doctoral Thesis, University of Twente,
Enschede, The Netherlands.

25
APPENDIX
Combining roughness of two surfaces
The parameters concerning surface roughness of the sliding pair piston ring – cylinder liner
have been calculated based on publications of Greenwood and Tripp (21) and Gelinck (44). The
standard deviations of the summits are combined as follows:

σ = σ 12 + σ 22 [A1]
with σ1 and σ2 the standard deviations of the summits of, respectively, surface 1 and 2, and σ the
equivalent standard deviations of the summits.
Then the radii of summits of the two surfaces are combined by:
1 1 1
= + [A2]
β β1 β2
with β1 and β2 the average radii of the summits of, respectively, surface 1 and surface 2.
Finally a value of the equivalent asperity density η should be determined. For surfaces of
equal roughness the value of a single surface can be taken. When one surface is much rougher,
then the density of this surface can be assumed. It should be also noted that the value of the
product:
ηβσ ≈ 0.03 ÷ 0.05 [A3]
for most surfaces (21).

Numerical solution of the Patir&Cheng equation


The numerical solution is based on the implicit finite difference scheme. The Patir&Cheng
equation [6] is discretized to form a set of linear equations (Eq. [A4]) which have a tri-diagonal
matrix form (Eq. [A5]), and can be effectively solved in each time step.

φxn, i −1 / 2 ⋅ (hin−1 / 2 )3 pin−1 − [φxn, i +1 / 2 ⋅ (hin+1 / 2 )3 + φxn, i −1 / 2 ⋅ (hin−1 / 2 )3 ] pin + φxn, i +1 / 2 ⋅ (hin+1 / 2 )3 pin+1
=
(Δx) 2
[A4]
hn − hn φn −φn h n − hin −1
= 6 μu n i +1 / 2 i −1 / 2 + 6μσu n S , i +1 / 2 S , i −1 / 2 + 12μ i
Δx Δx Δt
where: Δx – increment of x-coordinate, Δt – time step.
The numerical scheme [A4] includes subscripts i-1/2 and i+1/2 that signify numbers of auxiliary
nodes placed in the middle between the node numbers i-1, i and i, i+1 respectively. Superscripts
n-1 and n mean the previous and the current time level respectively.
The tri-diagonal matrix form of equation [A4] can be written as follows:

26
⎡ b1 c1 ⎤ ⎧ p1 ⎫ ⎧ d1 ⎫
⎢a b2 c2 ⎥⎪ p ⎪ ⎪ d ⎪
⎢ 2 ⎥ ⎪⎨ 2 ⎪⎬ = ⎪⎨ 2 ⎪⎬ [A5]
⎢ a3 ... ... ⎥ ⎪ ... ⎪ ⎪ ... ⎪
⎢ ⎥
⎣ ... bm ⎦ ⎪⎩ pm ⎪⎭ ⎪⎩d m ⎪⎭

For node numbers (i=1,2, … ,m), the following form of algebraic equations can be written:

ai pin−1 + bi pin + ci pin+1 = d i [A6]


where coefficients (ai, bi, ci) of trigonal matrix and of the right side vector di are as follows:
⎧a = φ n
⎪ i ( )
x , i −1 / 2 ⋅ hi −1 / 2
n 3

⎪b = −[φ n
⎪ i ( )
x , i −1 / 2 ⋅ hi −1 / 2
n 3
( ) 3
+ φ xn, i +1/ 2 ⋅ hin+1/ 2 ]
⎨ [A7]
(
⎪ci = φ xn, i +1/ 2 ⋅ hin+1/ 2 ) 3


( ) ( )
⎪⎩d i = 6μ u n Δx hin+1 / 2 − hin−1 / 2 + 6 μ σ u n Δx φSn, i +1/ 2 − φ Sn, i −1 / 2 + 12 μ (Δx ) ⋅ vsq
2

and: a1=0, cm=0

By calculating the coefficient di the necessary piston ring vertical (or squeeze) velocity can be
found in the following way:
hin − hin −1
vsq = [A8]
Δt
A general solution of a piston ring radial motion is obtained for every time step in the
following iterative way. Assuming that the minimal gap between the ring and cylinder liner hmin
(Fig. 4) is known, the squeeze velocity vsq is predicted, the hydrodynamic and roughness contact
pressure distributions are calculated, and the hydrodynamic force Fh (Fig. 5) and contact force Fc
estimated and compared with the sum of the gas force Fgas and ring elastic tension force Fspr :

Fh (vsq , hmin ) + Fc (hmin ) ≈ Fspr + Fgas [A9]

Depending on the difference between the force sum required and the force sum calculated,
a correction of the velocity vsq (and connected change of minimal gap hmin) is prescribed and
a new calculation performed. The iterations are continued until the desired force equilibrium
accuracy is reached. Radial forces are balanced in this way for each piston ring.

27
NOMENCLATURE

A – crevice cross-section area (m2)


Ac – real area of contact (per unit circumference) (m2/m)
ap – piston acceleration (m/s2)
cimc – ring circumference (m)
Cc – coefficient of contraction
CD – orifice discharge coefficient
Cv – coefficient of velocity loss
D – cylinder diameter (m)
dq – diameter of lubricating quill (m)
E’ – composite elastic modulus (for cylinder liner and ring) (Pa)
Efri – energy loss due to wall friction (J)
Fc – radial force (per unit circumference) due to asperity contact (N/m)
Fcx – friction force (per unit circumference) due to asperity contact (N/m)
Ffri – friction force (per unit circumference) due to viscous shear (N/m)
Fgas – gas force (per unit circumference) (N/m)
Fgi – leading side gas force (per unit circumference) (N/m)
Fgi+1(2) – trailing side gas forces (per unit circumference) (N/m)
Fh – radial hydrodynamic force (per unit circumference) (N/m)
Fspr – ring elastic tension force (per unit circumference) (N/m)
Fygi – leading edge gas force (per unit circumference) (N/m)
Fygi+2 – trailing edge gas force (per unit circumference) (N/m)
g – gravitational acceleration (m/s2)
goil – lubricating oil feed rate per cylinder (g/kWh)
h – nominal oil film thickness (m), specific enthalpy of gas (J/kg)
hmin – minimal gap between the ring and cylinder liner (m)
_
hT – average separation (gap between the ring face and cylinder liner) (m)
k – specific heat ratio
K – ring torsional stiffness (per unit circumf.) (N/rad)
m – mass (kg)
n – engine rotational speed (rpm)
p, po – pressure, stagnation pressure (Pa)
_
p – mean hydrodynamic pressure (Pa)
Pcyl – effective engine power per cylinder (kW)
Q – heat (J)
R – gas constant (J/(kg⋅K))
Rx – axial reaction force (per unit circumf.) between the ring and piston groove (N/m)
t – time (s)
T, To – temperature, stagnation temperature (K)
TC – temperature (°C)
u – piston ring axial velocity (m/s), specific internal energy of gas (J/kg)
vsq – piston ring vertical (or squeeze) velocity (m/s)
V – volume (m3)
x – coordinate along cylinder liner (m)

28
xlev – oil distribution for upper or lower lubrication level of cylinder liner (%)
xr – axial movement of ring in piston groove (m)
α2 – mixed friction force coefficient
β – asperity radius of curvature (m)
η – asperity density (m-2)
ρ – density of lubricant (kg/m3)
μ – oil dynamic viscosity (Pa⋅s)
ν – oil kinematic viscosity (cSt)
σ – composite root mean square roughness of both mating surfaces (m)
τA – shear stress in surface film arising from asperity interactions (Pa)
θ – twist angle of piston ring (rad)
Subscripts
i – number of inflow channel
In – gas inflow
j – number of outflow channel
k – number of analyzed gas volume
Out – gas outflow

29
Fig. 1. Scheme of a two-stroke marine engine (1)

a) b)

Fig. 2. Schemes of main lubricating oil supply systems of two-stroke marine engines: a) pressure
accumulation system, b) pulse lubricating system (1, 3)

30
p5, T5, V5, m5
Q8 A8
Q7 A7 p6, T6, V6, m6
Q9 A9
p7, T7, V7, m7
Q11 A11
p8, T8, V8, m8

A10
Q10
4
Q12 A12
p9, T9, V9

Fig. 3. Scheme of gas flow through the labyrinth seal: piston – rings – cylinder liner and the applied
physical model for the ring pack of four piston rings

Fig. 4. Flow parameters in the gap between the ring face and cylinder liner in the case of: a) full,
b) mixed lubrication

31
Fig. 5. Scheme and definitions of forces acting on a piston ring. Nomenclature: Fh – hydrodynamic
normal force, Fc – contact normal force, Fspr – ring spring force, Fgas – back ring gas force, Fygi
– leading edge gas force, Fygi+1 – trailing edge gas force, Rx – groove reaction force,
Ffri – viscous friction force, Fcx – contact friction force, Fgi – leading side gas force,
Fgi+1 , Fgi+2 – trailing side gas forces, m – ring mass, g – gravitational acceleration,
ap – piston acceleration

100
Cylinder liner
Relative parameter value y/y max [%]

90
80
70
60
50
Deform d/dmax [%]
40
Wear w/wmax [%]
30
Temp T/Tmax [%]
20
10
0
-0,25 0 0,25 0,5 0,75 1 1,25 1,5 1,75 2 2,25 2,5 2,75
X - coordinate along cylinder wall Xcyl [m]

Fig. 6. Relative variations of chosen parameters concerning cylinder liner (at full engine load): T/Tmax
[%] – temperature of cylinder surface, d/dmax [%] – thermal and mechanical deformation of
cylinder liner, w/wmax [%] – wear of cylinder surface per 1000 hours

32
160
Calculated curves - solid p1
lines
140
Experimental curves - dotted
lines
120
Gas pressure p [bar]

100
p3
80

60
p5
40 4

20 p7
p9
0
0 30 60 90 120 150 180 210 240 270 300 330 360
Crank angle [ ° ]

Fig. 7. Comparison of gas pressure variations pi (measured and calculated) as a function


of crankshaft rotation. Gas pressure nomenclature: p1 – in combustion chamber,
p3 – between the 1st and 2nd piston ring, p5 – between the 2nd and 3rd piston ring,
p7 – between the 3rd and 4th piston ring, p9 – under the piston

Fig. 8. Scheme of a gland-box of two-stroke marine engine (1)

Fig. 9. Ring pack geometry under consideration

33
0,4 Xr1 Xr2 Xr3 Xr4

Ring lifts in piston grooves Xr [mm] 0,35

0,3

0,25

0,2

0,15

0,1

0,05

0
0 30 60 90 120 150 180 210 240 270 300 330 360
Crank angle [ ° ]

Fig. 10. Variations of ring lift Xr_i in piston groove for each piston ring (i – ring number) versus crank
angle

250000
Fh1
Hydr. force (per unit circumf.) Fh [N/m]

225000 Fh2
200000 Fh3
Fh4
175000

150000

125000

100000

75000

50000

25000

0
0 30 60 90 120 150 180 210 240 270 300 330 360
Crank angle [ ° ]

Fig. 11. Variation of hydrodynamic force Fh_i for each piston ring (i – ring number) versus crank angle

34
70000
Fc1

Radial contact force / unit circumf. Fc [N/m]


Fc2
60000
Fc3
Fc4
50000

40000

30000

20000

10000

0
0 30 60 90 120 150 180 210 240 270 300 330 360
Crank angle [ ° ]

Fig. 12. Variation of radial component of contact force Fc_i for each piston ring (i – ring number) versus
crank angle

75
Hydr. friction force (per unit circumf.) Fri [N/m]

50

25

-25
Fri1
Fri2
-50
Fri3
Fri4
-75
0 30 60 90 120 150 180 210 240 270 300 330 360
Crank angle [ ° ]

Fig. 13. Variation of hydrodynamic tangential force Ffri_i for each piston ring (i – ring number) versus
crank angle

35
4000
Fcx1

Axial contact force / unit circumf. Fcx [N/m]


3000 Fcx2
2000 Fcx3
Fcx4
1000

-1000

-2000

-3000

-4000

-5000

-6000
0 30 60 90 120 150 180 210 240 270 300 330 360
Crank angle [ ° ]

Fig. 14. Variation of tangential component of contact force Fcx_i for each piston ring (i – ring number)
versus crank angle

8
End of downstroke
7
Oil supply places
Oil film thickness hoil [μm]

4
hmean
3
hoil
2

1
Scavenging
air ports
0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8 2 2,2 2,4 2,6
TDC Distance from top of cylinder wall x[m] BDC

Fig. 15. Variation of the oil film thickness hoil left by the ring pack along cylinder wall and mean value of
the oil film thickness hmean

36
Fig. 16. Oil wetted area (\\\\\\) and contact zone (//////) of the 1st ring as a function of crankshaft rotation

Fig. 17. Oil wetted area (\\\\\\) and contact zone (//////) of the 2nd ring as a function of crankshaft rotation

3
2
1
0
Twist angle [ ' ]

-1
-2
-3
-4
-5
Theta1
-6 Theta2
-7 Theta3
Theta4
-8
0 30 60 90 120 150 180 210 240 270 300 330 360
Crank angle [ ° ]

Fig. 18. Variation of twist angle Thetai for each piston ring (i – ring number) versus crank angle

37

Anda mungkin juga menyukai