Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/232380243

Experimental friction evaluation of cylinder liner/piston ring contact

Article  in  Wear · June 2011


DOI: 10.1016/j.wear.2010.08.028

CITATIONS READS

60 2,630

4 authors:

Staffan Johansson Per Henrik Nilsson


AB Volvo Volvo Technlogy AB
8 PUBLICATIONS   151 CITATIONS    275 PUBLICATIONS   10,866 CITATIONS   

SEE PROFILE SEE PROFILE

Robert Ohlsson Bengt-Göran Rosén


Independent Researcher Halmstad University
26 PUBLICATIONS   526 CITATIONS    121 PUBLICATIONS   1,171 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Business Model Innovation when adapting to Digital Production – opportunities and problems View project

Cylinder Liners View project

All content following this page was uploaded by Per Henrik Nilsson on 30 October 2017.

The user has requested enhancement of the downloaded file.


Wear 271 (2011) 625–633

Contents lists available at ScienceDirect

Wear
journal homepage: www.elsevier.com/locate/wear

Experimental friction evaluation of cylinder liner/piston ring contact


Staffan Johansson a,∗ , Per H. Nilsson a , Robert Ohlsson b , Bengt-Göran Rosén c
a
Volvo Technology AB, Chalmers Science Park, Sven Hultins Gata 9A, SE-412 88 Gothenburg, Sweden
b
Volvo Powertrain AB, Dept 91523, SE-405 08 Gothenburg, Sweden
c
Halmstad University, SET, Box 823, Halmstad University, SE-301 18 Halmstad, Sweden

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

Article history: Fuel consumption is an extremely important parameter for the automotive industry today. Anticipated
Received 23 January 2010 emission legislative demands in combination with a rising oil price are true motivators. In engines the
Received in revised form 27 August 2010 piston system is the largest source of frictional losses, accounting for about 50% of the total frictional
Accepted 31 August 2010
losses, thus it is important to optimize. Apart from frictional losses the piston system is a large con-
Available online 25 September 2010
sumer of lubricating oil, a considerable contributor to the total amount of particulate emissions (PM).
New materials, coatings and high-tech machining processes that previously were considered to be too
Keywords:
expensive and therefore only used in complex applications are today becoming more affordable. It is
Wear
Friction
important to develop reliable test methods to study these new concepts. The reciprocating tribometer at
Lubrication Volvo Technology has been updated to better evaluate the frictional difference between material combi-
Cylinder liner nations/surfaces; it is possible to evaluate a number of operational parameters in each experiment. The
Design of experiments components that were studied were a piston ring running against a cylinder liner. Friction, wear and
Tribometer change in surface morphology were studied in the experiments. It is shown that for the introduced DoE
based tribometer test the interaction of dynamic viscosity, velocity and contact pressure can be studied
within one experiment. The results show differences in friction which could be explained as the surface
creating beneficial contact conditions for oil film build-up. It is also apparent that surface roughness is
important regardless of material properties. To better understand the correlations between friction and
surface roughness a future study should include a study of similar materials with different roughness
values.
© 2010 Elsevier B.V. All rights reserved.

1. Introduction cylinder–piston rig interaction area, the properties of the lubricant


are dominant [2]. In mixed and boundary lubrication the surface
In heavy-duty diesel engine development there is a large focus texture plays a greater role [3]. A general trend in future heavy-
on decreasing fuel consumption and CO2 emissions. One part of duty engines is that the peak combustion pressure will increase
this focus can be accomplished by minimizing the frictional losses [4], this will result in increased frictional losses and wear of com-
of the engine. The single largest contributor to frictional losses ponents.
is the power cylinder unit, accounting for roughly half of the An alternative step to full-scale experiments is simplified rig
losses [1]. To measure the frictional losses of the piston system at component testing. In such tests it is possible to isolate the com-
full scale is complicated; cycle-to-cycle variation, vibrations, ther- ponents of interest and study a particular phenomenon and at the
mal effects and other noise contributing factors often overshadow same time minimize noise factors. In a typical engine a predeter-
frictional differences when examining different engine configura- mined set of different parameter settings is used to scan (map)
tions. the operational window. In typical tribometer experiments only
Minimizing friction in the power cylinder unit demands bet- one parameter setup is used throughout the experiment duration.
ter understanding for how the different parameters are linked. In The tribometer was further developed to incorporate a paramet-
hydrodynamic lubrication, which is present on the larger part of the ric design-of-experiments approach where the effects of speed,
dynamic viscosity and contact pressure could be evaluated during
the course of a single experiment.
The objective of this work was to develop and verify a method-
Abbreviations: CTDC, combustion top dead centre; DoE, design of experiments; ology capable of mimicking the real engine behaviour at boundary
LVDT, linear variable differential transformer; MVA, multivariate analysis; VIP, vari- and mixed lubrication regimes in order to minimize frictional losses
able importance in the projection (response in MVA).
∗ Corresponding author. Tel.: +46 31 322 99 66; fax: +46 31 82 08 87.
and wear. This work focused on the development of test rig method-
E-mail address: staffan.sj.johansson@volvo.com (S. Johansson).
ology and to investigate present and future candidate materials.

0043-1648/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.wear.2010.08.028
626 S. Johansson et al. / Wear 271 (2011) 625–633

Nomenclature

a0 correlation parameter of the Vogel equation


H Hersey parameter, H = N/p
Hv modified Hersey parameter, H v = v/p (m)
N rotational frequency (1/s)
p contact pressure (Pa)
T contact temperature (◦ C)
T1, T2 correction parameters of the Vogel equation (◦ C)
v sliding speed (m/s)

Greek letters
 shear rate (1/s)
c critical shear rate (1/s)
,  dynamic viscosity (Pa s)
∞ dynamic viscosity for high shear rates (Pa s)
0 dynamic viscosity for low shear rates (Pa s)
 oil density (kg/m3 )
 kinematic viscosity (m2 /s)

2. Materials and methods

2.1. Experimental setup


Fig. 2. Circuit diagram of electrical resistance measurement between piston ring
The schematics of the reciprocating eccentric tribometer are and cylinder liner.
shown in Fig. 1. The component used in the tribometer is a com-
plete piston ring and a segment of the cylinder liner. By using An unbalanced motion creates unwanted vibrations. The main
real production components (as opposed to simplified geometries) method of dealing with vibrations is to fix all components of the
that contain the correct surface morphology, coating thickness and rig onto a sturdy base plate which rests on rubber mountings. The
material composition the best component comparison to the real vibrations are further reduced by having a system of spur gears with
engine tests is made. adjustable weights that run in anti-phase with the sledge motion.
A three-phase electrical motor provides the motion in the tri- A central problem in tribometer experiments it is the geomet-
bometer. The motor is connected via an eccentric and a lever arm rical alignment. To minimize the linearity problem a number of
to a carbon fibre sledge. The sledge, which is the housing for the pre-samples were evaluated in order to minimize the linearity
cylinder liner segment moves in a reciprocating motion. The dis- problems before any experimental study is initiated.
placement of the sledge is measured with an LVDT. The temperature rise is generated from a heat gun, directed at
A plate fixture is pneumatically loaded onto the sled; this gen- the liner specimen. This resembles the heating of the piston system
erates normal force as the liner specimen is loaded onto the upper where combustion gases are the main contributors to heating.
specimen (fixed), namely the piston ring. To avoid frictional heating To ensure consistent input and output parameters, the displace-
and wear of the sled an air bearing is fitted onto the plate fixture; ment of the sled, the normal force acting between cylinder liner and
this gives the contact between plate fixture and sled little friction. piston ring and the temperature are controlled using a feedback
The friction force is measured with two piezo electric force loop.
transducers and the normal force is measured with a strain gage The main cooling mechanism of the piston system in the engine
transducer. The sample rate is set to be dependent on the recip- is the water circulation on the wet side of the cylinder liner. The
rocating frequency, each stroke is sampled at 2000 positions (a cooling is dealt with in the same fashion in the tribometer; a water
reciprocating frequency of e.g. 15 Hz will accordingly be sampled coolant channel runs through a channel in the sled. The main advan-
at 30 kHz). tage of having two systems that counteract the temperature change
is that the parameter stabilizes without difficulty. Since running
parameters settings are changed frequently in the rig it is important
to have a swift convergence of the feedback loop.
Oil is supplied directly to the ring sample holder via a peristaltic
pump. Because of the pump being volumetric it will always pump
the same amount of oil per time segment. For this testing the pump
delivers a constant amount of 0.4 ml/min throughout the experi-
ment; this mimics the starved lubrication conditions at combustion
top dead centre (CTDC).
The tribometer acts as a circuit where the contact between the
piston ring and the cylinder liner acts as a potentiometer. The pur-
pose of the electrical resistance measurement is mainly to quantify
the oil film thickness although oxide formation and oil additive
deposits also alter the resistive signal.
The circuit is fed with a current Uf of 50 mV DC (see Fig. 2). Rcontact
Fig. 1. Overview of the reciprocating eccentric tribometer. symbolizes the electrical resistance through the asperity contact of
S. Johansson et al. / Wear 271 (2011) 625–633 627

tions within the piston system [5,6]. Reynolds average equation


was used to determine the oil film thickness.
The aim of this study was to minimize the frictional losses. By
studying the frictional power losses using piston simulation it was
found that the highest loss due to mechanical contact is related to
the top ring and occurs close to CTDC, at around 15–30 crank angle
degrees (see Fig. 3).
The purpose of the tribometer friction study was to mimic the
piston ring/cylinder liner contact in the engine, it was however not
possible in a simplified tribometer experiment to utilize the same
speeds, loads and temperatures as can be found in the engine. The
objective was thus to find a parameter that can compare the contact
of the top piston ring against the cylinder liner in the engine and in
the tribometer. The parameter that is used for this comparison is
the Hersey parameter [7]; a combination of global contact pressure,
dynamic viscosity and velocity. The Hersey parameter was defined
with rotational frequency as velocity parameter, since a reciprocat-
ing motion is studied a modified version of the Hersey Parameter
Fig. 3. Engine friction power loss of top piston ring for 1200 rpm, 50%, 75% and full is used incorporating point speed as velocity parameter.
load. The dashed box indicates the CTDC area where the largest friction power loss The face profile of the piston ring had a shape of an unsymmetri-
is found.
cal barrel. This meant that normal Herzian theory based on simpler
geometry was not applicable; instead a linear elastic method [8]
the piston ring and the cylinder liner and/or the oil film. The volt- was used to solve the pressure for the unsymmetrical geometry.
age Ucontact is linearly correlated with Rcontact , this voltage is used The correlation between contact load and contact pressure is not
to represent the electrical resistance in further analysis, it is later linear (see Fig. 4) so it was needed to determine the contact pressure
sections referred to as resistive coefficient. R1 is a potentiometer for each specific contact load. Using the radial load, which was cal-
with fixed discrete levels. For this arrangement to work soundly the culated with piston simulation it was possible to solve the contact
value of Rcontact and R1 should be of the same magnitude, even if pressure at all crank angle degrees for the top piston ring/cylinder
Rcontact fluctuates on each stroke as transitions of lubrication regime liner contact.
occurs. The value of R1 is determined by running an initial exper- Using the oil film thickness calculated with piston simulation it
iment (using the same setup of parameters proposed for the test was possible to calculate the shear ratio. To retrieve the dynamic
series) to confirm that the signal can be quantified for all running viscosity the Vogel (Eq. (1)) and Cross (Eq. (2)) equation were solved
conditions during the duration of the experiment. [9]. Since the oil film thickness in the tribometer experiments was
not measured the dynamic viscosity could not be calculated using
2.2. DoE experimental setup and tribometer parametric the Cross equation, instead the relation between kinematic viscos-
description ity and density had to be used (Eq. (3)) [10].

The experimental plan used in the tribometer study is  T1 


shown in Table 1. This table shows the DoE approach with 0 = a0 exp (1)
T2 + T
high and low level of input parameters; temperature, fre-
quency and load. The parameters dynamic viscosity, sliding  0 − ∞
velocity and pressure are then calculated from the original  = ∞ + (2)
1 + |/c |
parameters (for details of calculation ahead in this section).
The test plan incorporates a parametric design-of-experiments
=× (3)
approach were the effects of sliding velocity, dynamic viscosity
and contact pressure can be evaluated during the experi-
ments. A good comparison between the calculated Hersey parameter
To ensure an accurate comparison between the engine and the of the engine and the tribometer was found, as can be seen in Fig. 5.
tribometer the first step was to find the engine parameters for the The rig settings used (red dots) correspond to the crank angle inter-
tribological contact; these parameters were retrieved from sim- val of 3–28◦ for full load and 3–22◦ for 50% load (blue curve). It is
ulated values. “Piston Simulation” was used for this task, it is a approximately at this crank angle interval that the frictional power
software developed dedicated to solve the force and contact equa- loss is highest (see Fig. 3).

Table 1
Experimental cycle used in tribometer experiments. Running-in stage is defined as the first stage (duration 180 min).

Step no. Time (min) Oil temperature (◦ C) (dynamic viscosity (Pa s)) Reciprocating frequency (Hz) (sliding velocity (m/s)) Load (N) (contact pressure (Pa))

1 180 91 (1.47E−02) 10 (0.6) 259 (9.86E+07)


2 30 79 (2.01E−02) 5 (0.3) 150 (6.91E+07)
3 30 79 (2.01E−02) 15 (0.9) 150 (6.91E+07)
4 30 112 (9.23E−03) 5 (0.3) 150 (6.91E+07)
5 30 112 (9.23E−03) 15 (0.9) 150 (6.91E+07)
6 30 91 (1.47E−02) 10 (0.6) 259 (9.86E+07)
7 30 79 (2.01E−02) 5 (0.3) 473 (1.41E+08)
8 30 79 (2.01E−02) 15 (0.9) 473 (1.41E+08)
9 30 112 (9.23E−03) 5 (0.3) 473 (1.41E+08)
10 30 112 (9.23E−03) 15 (0.9) 473 (1.41E+08)
11 30 91 (1.47E−02) 10 (0.6) 259 (9.86E+07)
628 S. Johansson et al. / Wear 271 (2011) 625–633

Fig. 4. The contact pressure as a function of piston ring position and contact load.

2.3. Materials 2.4. Roughness characterisation

Three different cylinder liner materials were exam- To enable the link to function a characterization of the surface is
ined:Reference material; Gray cast iron (REF), plateau honed. needed. Surfaces were measured using a chromatic confocal probe
that scanned the surfaces in three dimensions [11]. The surfaces
used for evaluation had the dimension of 1.75 mm × 1.75 mm with
• Surface parameter Sa = 0.46 ± 0.04 ␮m.
a 2 ␮m spacing. There were four roughness measurements con-
ducted per tribometer experiment; two were located at the centre
Standard gray cast iron with additional heat treatment (MATE- of the stroke and two at one of the reversal zones. The value used
RIAL1), plateau honed. in the evaluation was the mean of these four measurements (two
measured before and two measured after the experiment). Surfaces
were characterized by using the complete set of 3D parameters
• Surface parameter Sa = 0.71 ± 0.04 ␮m. described in standards ISO 25178 and EUR 15178N [12]. All sur-
faces were filtered using a robust Gaussian filtering with a cut-off
length of 0.5 mm prior to the calculation of the surface roughness
Thermally sprayed material (MATERIAL2), single stage honed. parameters.
By using the complete sets of parameters numerous properties
• Surface parameter Sa = 0.28 ± 0.01 ␮m. of a measured surface was characterized such as spatial fea-
tures, texture direction, roughness at different surface amplitude
etc.
The roughness values of Sa (above) were based on the surface
measurement prior to the experiment. MATERIAL2 was different 2.5. Wear characterisation
compared to the other materials; the valley part of this material
consists of a porous isotropic structure (common for thermally Wear was evaluated by measuring the complete worn sur-
sprayed materials). REF and MATERIAL1 were plateau honed with face, before and after experiment. The surfaces measured for wear
a honing angle of 54◦ . The counterpart in these experiments was evaluation purposes were 48 mm (axial direction, direction of
a PVD (CrN) coated top piston ring. Fresh 15W40 oil was used in motion) × 10 mm (tangential direction) with a lateral resolution of
the experiments. Each combination was represented by the mean 10 ␮m. The surfaces of the cylinder liner sample are averaged in
of five experimental repetitions. the tangential direction; this returns a profile of the wear depth in

Fig. 5. The correlation between Hersey parameter from engine tribometer for the investigated top piston ring/cylinder liner contact. Line indicates Hersey number of engine,
circle indicates average value of tribometer test cycle.
S. Johansson et al. / Wear 271 (2011) 625–633 629

of friction could be found in the reversal points and lowest at mid


stroke (see Fig. 7, left). The frictional behaviour was expected since
the tribometer experiments were intended to analyze the boundary
and mixed lubrication regimes.
To obtain a representative value of friction coefficient the vari-
ations of friction coefficient over the stroke is averaged (see Fig. 8,
left). It was important to compare the frictional responses of cycle
steps regardless of the cycle step occurring in the beginning or
towards the end of the experiment, it is thus crucial to remove
running-in trends from experimental step 2–11. In order to com-
pensate for the running-in effects two linear functions (vectors) are
subtracted from the specific parts of original friction values (see
Fig. 8, right), one acting on the high-load cycle steps (DoE stage
7–11) and one working on the all steps, apart from the first 150 of
the running-in period. The reason as to why two functions were
subtracted from the original friction values was that the running-
in effects (commonly a decrease in friction) were more apparent in
the experimental steps with high contact pressure (step 7–10) in
comparison to the experimental steps with lower contact pressure
Fig. 6. Example of quantification of wear depth. Profiles of original and worn surface (step 2–5).
averaged in tangential direction as seen in this figure.

the axial direction. Wear volume was calculated by subtracting the 2.7. Multivariate analysis
original profile with the profile of the worn surface (see Fig. 6).
With multivariate analysis (MVA), one can investigate the rela-
2.6. Friction characterisation tions between all variables in a single context. In this MVA approach
four approaches are examined. The statistical analysis tool used in
Friction was continuously measured in the tribometer experi- this study was Simca-P 10.0. In this study a linear correlation is used
ments. During the stroke the friction varies, hawing highest value to study the responses between input and output variables. In all

Fig. 7. Measurement of friction coefficient in tribometer. Left: An example of one friction coefficient measurement over one stroke. The highest friction can be found at the
reversal points (position coordinate 400, 950, 1100 and 1800), the lowest friction can be found at mid stroke (position coordinates 600 and 1550). Right: Friction coefficient
measurement of one experiment showing both the variations over the stroke length as well as the variations over the experimental duration. After the running-in stage
(180 min) the friction coefficient changes for each experimental cycle step.

Fig. 8. Average friction coefficient over all experimental cycle steps. Left: Friction graph and the two levelling vectors plotted against time. Right. Original friction graph,
friction graph after first levelling and friction graph after second levelling plotted against time.
630 S. Johansson et al. / Wear 271 (2011) 625–633

Fig. 9. Left: Cylinder liner average wear depth of the studied materials. Right: Surface roughness parameter Rq [␮m] of piston ring, after experiment, measured inside of the
wear mark. The error bars represent the standard deviation of the five experiments.

Fig. 10. Average friction coefficient including running-in stage (left) and excluding running-in stage (right). The error bars represent the standard deviation of the five
experiments.

analyzes the coefficients representing input parameters are scaled binations; there was a significant difference in average liner wear
and centred. depth. Compared to the stabile reference material (REF) wear depth,
MATERIAL1 shows both a 100% increase in average wear depth as
3. Results well as 400% increase of variation width (the error bars in Fig. 9 rep-
resents ± one standard deviation of five experiments). MATERIAL2
3.1. Wear shows a more than halved average wear depth and also decreased
variation. From a previous study [13] it was shown that it was ben-
Initial “running-in” wear of the cylinder liner is concentrated eficial to quantify wear on the top piston ring using the profile
to the top (plateau) part of the liner surface. In Fig. 9, the average parameter Rq [␮m] (ISO 4287 [14]). The Rq [␮m] values (Fig. 9,
liner wear is visualized for the three investigated material com- right) are based on five profiles measured inside of the wear mark

Fig. 11. Loading scatter plot of MODEL1. It can be observed that the friction coefficient correlates to the experimental parameters Velocity.
S. Johansson et al. / Wear 271 (2011) 625–633 631

0,30

CoeffCS[3](Mean Resistance Coefficient)


0,20 0,40

CoeffCS[3](MeanFriction Coeff)
0,10
0,00 0,20

-0,10
0,00
-0,20
-0,30
-0,20
-0,40
-0,50
-0,40
-0,60
-0,70 -0,60
REF

MATERIAL1

MATERIAL2

Dynamic Viscosity

Velocity

Contact Pressure

REF

MATERIAL1

MATERIAL2

Dynamic Viscosity

Velocity

Contact Pressure
Fig. 12. MODEL1. Left: Coefficient plot for average friction coefficient. Right: Coefficient plot for average resistive coefficient. The error bars in the coefficient plots represents
95% confidence interval.

0,00 0,00
0,20
-0,10 -0,10
CoeffCS[1](Mean

CoeffCS[1](Mean

CoeffCS[1](Mean
-0,20 0,00
-0,20
Friction Coeff)

Friction Coeff)

Friction Coeff)
-0,30 -0,20 -0,30
-0,40
-0,50 -0,40 -0,40
-0,60 -0,60 -0,50
-0,70
-0,60
-0,80 -0,80
-0,90 -0,70

Dynamic Viscosity
Dynamic Viscosity

Velocity
Dynamic Viscosity

Velocity
Velocity

Contact Pressure
Contact Pressure
Contact Pressure

Fig. 13. Coefficient plots for describing the correlation between dynamic viscosity, velocity and contact pressure on mean friction coefficient for the three investigated
materials. Left: REF. Centre: MATERIAL1. Right: MATERIAL2. The error bars in the coefficient plots represents 95% confidence interval.

measured after the experiment. The difference of the piston ring Rq running-in stage does not alter the friction coefficient significantly,
is insignificant due to the overlap of the variations. only a small decrease can be observed.

3.2. Friction measurement


3.3. Multi variate analysis of the combined correlations between
The average friction coefficient over the entire test cycle for the friction, resistive coefficient and surface roughness parameters
three combinations is shown in Fig. 9, the error bar marks one
standard deviation. With this coarse approach differences between Using MVA four different models were analysed to study the
MATERIAL1, REF and MATERIAL2 can be clearly observed. The correlations between input and output parameters:

0,150
CoeffCS[1](Mean
Friction Coeff)

0,100

0,050

0,000
Vvv
Vmc

Sa

Sk
Sxp

Sdc

Fig. 14. Coefficient plot describing the correlation of surface roughness parameters and mean friction coefficient. The error bars in the coefficient plots represents 95%
confidence interval.
632 S. Johansson et al. / Wear 271 (2011) 625–633

• MODEL1: Input parameters: Material type, dynamic viscosity, period for all investigated materials (Fig. 10). A future tribometer
velocity and contact pressure in relation to output parameters: study should include experiments with representative wear levels
average resistive coefficient and average friction coefficient. as well as a quantification of friction that is beyond the running-in
• MODEL2: This model is similar to previous model but with the period.
difference that each material type is treated separately. Input Running-in causes a decrease in friction during the experiment
parameters: dynamic viscosity, velocity and contact pressure in duration; however, effects associated with running-in can also be
relation to output parameter: average friction coefficient. seen in the experimental cycle steps of 30 min (see Fig. 8). It is of
• MODEL3: Input parameters: All surface roughness parameters importance to measure a stable friction signal thus the cycle step
(see) in relation to output parameter: average friction coefficient duration should be increased in future experiments.
There is a discrepancy between the experimental results in
Fig. 11 shows the loading scatter plot for MODEL1. It can clearly Fig. 12 and the expected “Stribeck” [7] behaviour for the bound-
be seen that mean friction coefficient anti-correlates to all the input ary and mixed lubrication regimes. Typically friction coefficient
parameters. Fig. 12 shows the coefficient values for all parameters decreases with a decrease in contact pressure in said regimes. How-
for MODEL1 (the values are scaled and centred). It can be observed ever, support for this result can be found in [15] but further studies
that the frictional signal is dependent on both material type and are required to confirm these findings.
experimental settings but resistive coefficient is only dependent There is no correlation between resistive coefficient and the
on material type. This means that frictional changes cannot be parameters velocity, dynamic viscosity and contact pressure. How-
quantified with resistive signal when different material types are ever, the resistance coefficient shows an excellent correlation to
examined. material type (see Fig. 12). One reason for this might be that the
Fig. 13 shows the coefficients for the three models (MODEL2). conductivity of the different materials overshadows the frictional
MATERIAL2, which exhibit the lowest friction coefficient (see differences.
Fig. 10) also shows the clearest dependence of dynamic viscosity Since the total number of investigated surface roughness param-
and velocity. In other words, MATERIAL2 has the possibility to gen- eters is large (48 roughness parameter prior to the reduction) and
erate a friction reducing oil film were REF and MATERIAL1 fails. many of the parameter inter-correlates a future study to remove
In Fig. 12 left the pressure shows the same behaviour as dynamic the redundant parameters can be beneficial [16].
viscosity and velocity, as the pressure increases, friction decreases. From the analysis of the surface roughness parameters correla-
MODEL3 describes the correlations between surface roughness tion to friction it could be shown that:
parameters and mean friction coefficient. The total number of sur-
face roughness parameters in this model was large (48 in total),
• Small value of surface roughness parameter Sxp indicates sur-
since a multitude of these parameters showed a correlation to
faces with small sharp peaks removed i.e. pre-run-in surfaces are
mean friction coefficient the parameter set were reduced to include
beneficial.
only the parameters with the highest correlation to friction. Fig. 14
• A surface with a small plateau amplitude (surface roughness
shows the six parameters that both had the largest absolute coef-
parameter Sk, Sds and Vmc) decreases friction in boundary and
ficient value as well as the largest model importance (also known
mixed lubrication regime.
as VIP values1 ). These selected parameters have the highest cor-
• Parameters that describes the valley part of the surface (sur-
relation to mean friction coefficient in comparison to the total
face roughness parameter Vvv and Sxp) also shows a correlation
number of surface roughness parameters in MODEL3. The resulting
to friction. Before any conclusions can be drawn from this it is
six parameters were divided into three groups that each describe
important to verify that there is no internal correlation between
similar a characteristic of the surface that is necessary to decrease
the roughness parameters describing the plateau amplitude of
friction:
the surface and the roughness parameters that describes the val-
ley amplitude of the surface.
• Properties relating to the plateau part of the surface amplitude
• Small value of surface roughness parameter Sa indicates that
• A smooth plateau part of the surface amplitude (surface rough-
besides a low plateau roughness amplitude also the amount of
ness parameter Sk)
high peak areas and low valley areas is minimized. Especially high
• A small height of the core (in reference to the material ratio
peaks can cause oil film penetration and less favorable frictional
curve) of the surface (surface roughness parameter Sdc (limits
regimes.
p = 10%, q = 80%)
• A small material volume of the core (in reference to the material
ratio curve) of the surface (surface roughness parameter Vmc 5. Conclusions
(limits p = 10%, q = 80%))
• Properties relating to the valley part of the surface amplitude
The purpose of the tribometer experiment has been to mimic
• A small void volume of the valleys (surface roughness param-
the conditions at TDC in the contact between cylinder liner and
eter Vvv (limit p = 80%)) and a small amplitude of the valleys top piston ring. The following conclusions can be drawn from the
(surface roughness parameter Sxp (p = 50% q = 97.5%) experimental results:
• Small total surface amplitude (surface roughness parameter Sa).

4. Discussion • It is shown that for the introduced DoE based tribometer test the
interaction of dynamic viscosity, velocity and contact pressure
The wear measured on the cylinder liner surface is small (wear can be studied within one experiment.
depth of MATERIAL2 is<100 nm), both in comparison to real wear • For a material/surface with lower friction (MATERIAL2) the
and comparison to the total surface amplitude. On top of this the importance of dynamic viscosity and velocity increases. For a
friction coefficient shows only small change for the running-in material/surface with higher friction only contact pressure is of
importance. This means that a surface has to be able to generate
conditions for oil film build-up, if this is not accomplished the
1
The VIP values (variable importance in the projection) reflect the importance of
properties of oil and the velocity has little (REF) or no (MATE-
terms in the model both with respect to Y and with respect to X (the projection). RIAL1) significance.
S. Johansson et al. / Wear 271 (2011) 625–633 633

• For the materials studied in this work it is apparent that sur- with Fuel, 2007 JSAE/SAE International Fuels and Lubricants Meeting Kyoto,
face roughness is important in the mild wear situation regardless Japan.
[3] I.M. Hutchings, Tribology: Friction and Wear of Engineering Materials (Met-
of material properties, this is accurate for boundary and mixed allurgy & Materials Science), Butterworth-Heinemann Ltd., London, 1992, pp.
lubrication regime. It can be clearly seen that all part of the 70.
surface amplitude should be minimized to decrease friction, how- [4] C.D. Rakopoulos, D.T. Hountalas, A.P. Koutroubousis, T.C. Zannis, Application
and evaluation of a detailed friction model on a DI diesel engine with extremely
ever, due to the multitude of surface roughness parameters that high peak combustion pressures, in: SAE 2002 World Congress, Detroit, Michi-
shows a significant correlation to friction it is difficult to draw gan, March 4–7, 2002.
conclusions of what surface characteristic (e.g. plateau or valley [5] T. Tian, Dynamic behaviors of piston rings and their practical impact – part i:
ring flutter and ring collapse and their effects on gas flow and oil transport,
amplitude) that is of most importance to decrease friction.
Proc IMechE Part J: J. Eng. Tribol. 216 (2002) 209–227.
[6] T. Tian, Dynamic Behaviors of Piston Rings and Their Practical Impact – Part II:
6. Suggestions for further work Oil Transport, Friction, and Wear of Ring/Liner Interface and the Effects of Piston
and Ring Dynamics, Proc IMechE Part J: J. Eng. Tribol. 216 (2002) 229–247.
[7] M.D. Hersey, Theory and Research in Lubrication, John Wiley & Sons, New York,
• To better understand the correlations between friction and sur- 1966, pp. 123–158.
face roughness a future study must include a study of similar [8] Andreas Almqvist, Fredrik Sahlin, Roland Larsson, Sergei Glavatskikh, On the
dry elasto-plastic contact of nominally flat surfaces, Tribol. Int. 40 (April (4))
materials with different roughness values.
(2007) 574–579.
• The wear amounts reported in the study are rather small in com- [9] Akiko Shimada, Yasuo Harigaya, Michiyoshi Suzuki, An analysis of oil film
parison to the normal wear found in an operating engine. Also, temperature, oil film thickness and heat transfer on a piston ring of internal
the change in friction during the running-in experimental step is combustion engine: the effect of local lubricant viscosity, in: 2004 Small Engine
Technology Conference, Graz (Austria), September 27–30, 2004.
quite small for all investigated materials. A future study should [10] S. Jacobson, S. Hogmark. Tribologi. Berlings, Arlöv, 1996, p. 51.
also include an accelerated test procedure with higher liner wear [11] H.-J. Jordan, Optical chromatic confocal probes. XII. International Colloquium
levels and clearer differences in the piston ring surface rough- on Surfaces Chemnitz, January 28th and 29th 2008.
[12] L. Blunt, X. Jiang, Advanced Techniques for Assessment Surface Topography.
ness after experiment as well as a quantification of friction that Kogan Page Limited, United Kingdom, ISBN 1 9039 9611 2.
is beyond the running-in period. [13] S. Johansson, P.H. Nilsson, R. Ohlsson, C. Anderberg, B.-G. Rosén, New cylinder
• A future study should include a full-scale validation of tribometer liner surfaces for low oil consumption, Tribol. Internat. 41 (September–October
(9–10)) (2008) 854–859.
test approach presented in this paper. [14] ISO 4287:1997, Geometrical Product Specifications (GPS) — Surface texture:
profile.
References [15] K. Xu, Effects of contact pressure on the coefficient of friction in friction
tests, in: Proc 2003 SAE World Congress, Detroit, Michigan, March 3-6,
2003.
[1] John B. Heywood, Internal Combustion Engine Fundamentals, McGraw-Hill
[16] B.G. Rosén, C. Anderberg, R. Ohlsson, Parameter correlation study of cylinder
Inc., United States of America, 1988, p. 730.
liner roughness for production and quality control, Proc IMechE Part B: J. Eng.
[2] Hiroshi Moritani, Hiroharu Tokoro, Mamoru Tohyama, Hiroyuki Mori, Toshi-
Manuf. 222 (2008) 1–13.
hide Ohmori, Motoichi Murakami, Challenge to the Diesel Engine Lubrication

View publication stats

You might also like