You are on page 1of 33

Journal Pre-proofs

Review of engine journal bearing tribology in start-stop applications

Nathália Duarte Souza Alvarenga Santos, Vínicius Rückert Roso, Marco


Tulio C. Faria

PII: S1350-6307(19)31240-3
DOI: https://doi.org/10.1016/j.engfailanal.2019.104344
Reference: EFA 104344

To appear in: Engineering Failure Analysis

Received Date: 22 August 2019


Accepted Date: 25 November 2019

Please cite this article as: Duarte Souza Alvarenga Santos, N., Rückert Roso, V., Tulio C. Faria, M., Review of
engine journal bearing tribology in start-stop applications, Engineering Failure Analysis (2019), doi: https://
doi.org/10.1016/j.engfailanal.2019.104344

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version
will undergo additional copyediting, typesetting and review before it is published in its final form, but we are
providing this version to give early visibility of the article. Please note that, during the production process, errors
may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2019 Published by Elsevier Ltd.


REVIEW OF ENGINE JOURNAL BEARING TRIBOLOGY IN
START-STOP APPLICATIONS
Nathália Duarte Souza Alvarenga Santosa, Vínicius Rückert Rosoa, Marco Tulio C.
Fariaa
aUniversidade Federal de Minas Gerais, Belo Horizonte, MG, Brazil

ABSTRACT

The environmental impacts of the air pollution from motorized


transportation in the public health has driven new environmental regulations to reduce
the gas emission produced by combustion engines. To address this issue, the automotive
industry has developed strategies to meet these new requirements, and the start-stop
systems emerge as a promising technology to reduce exhaust emissions and fuel
consumption. However, the start-stop systems lead to high loads on the main engine
components, such as piston, piston rings and bearings, since the temperature variation
and lubrication supply are intermittent with the frequent engine start up. Thereby, the
present study aims at presenting a literature review of the tribological aspects of car start-
stop systems. A discussion of the operating parameters associated with the engine
tribological behavior, mainly focused on the bearing crankshaft, is the core of this study.
The review encompasses the recent bearing developments to meet the new load
requirements, as well as the latest bearing researches regarding friction, lubrication and
wear.

Keywords: Start-stop systems; internal combustion engines; lubrication; bearings

INTRODUCTION

The health effects associated with transport-related air pollution has pushed
for stricter environmental regulations for vehicle exhaust gases. In 1970 a pioneer
emission standard for passenger cars was introduced by an European council [1]. In 1992,
the European regulatory pathway introduced a new norm to decrease car emissions,
containing six stages of progressively stringent emission control requirements. This
European regulation has reached 17 of 20 state-members from the G20 countries, which
are responsible for 90 percent of global vehicle sales. Moreover, several of Asian and
Latin American countries have been currently applying, in some extent, the European
emission requirements [2].
In order to meet this new emission standards and the market demand for lower
fuel consumption, the automotive industry has implemented strategies such as engine
downsizing, lean combustion [3, 4], weight reduction and low-rolling-resistance tires,
improved vehicle aerodynamics and also implemented vehicle hybridization and
electrification [5]. One of the simplest measures found by vehicle manufactures is the
stop-start technology, a low-cost solution in which the internal combustion engine is
automatically switched off when the vehicle is stationary and restarted upon driver's
demand or when needed. By this measure, especially in city traffic, in which the engines
spend a great amount of time on idle, the wasted energy from idling is reduced, decreasing
considerably fuel consumption, and consequently emission rates [6]. The start-stop
system can be also seen as a step to introduce hybrid systems [6], in which an electric

1
engine actuates at low speed and idle conditions, whereas the internal combustion engine
(ICE) operates at high speed and longer distances. The switch from an engine to the other
also demands quick start and stop of the internal combustion engine.
However, despite the advantages of the start/stop and hybrid systems, the
increased number of engine starts exposes the engine to more load transients and modifies
the engine operation temperature, influencing the tribology aspects of several core engine
parts such as the crankshaft, connecting rods and bearings. One of the reasons is that,
during a start cycle, the low engine rotation speed is insufficient to generate a full
hydrodynamic oil film. Moreover, to avoid delays during engine restart, the start cycle
has to be quicker and more aggressive than the conventional starts. In this conditions the
oil pressure probably is not fully stabilized, affecting the hydrodynamic film and
generating direct loads on the bearings [7].
Although the system is already implemented in several commercial vehicles,
it is known that automakers are concerned with the influences of the system
implementation on the engine durability [8, 9], as commercial vehicles are often expected
to operate for a prolonged durability over the life of application [10]. Hence, for the
system implementation, design parameters such as component materials, coatings, and
lubrication should be accounted for [8]. Thus, this work presents a literature review on
some tribological features of an internal combustion engine operating with start-stop
systems. The influence of the start-stop operation on the engine bearing behavior is the
main focus of this review.

LITERATURE REVIEW

This literature review aims at presenting some relevant technical publications


to the development of the start-stop system and the current challenges for its
implementation in the automotive sector, mainly those related to the bearing design.
Hence, this section is divided into the following subsections: start-stop operation,
tribological features of the internal combustion engines (ICEs), bearing characteristics
and the most recent technical contributions to the state-of-the-art.

START STOP OPERATION

The engine stop-start is a simple and low-cost strategy to reduce the energy
loss waste when vehicle is stationary, what has been increasing with the urban vehicle
fleet growth and leads to more time at traffic lights and in traffic jams. It is estimated that
during a vehicle operation, the energy spent in idle conditions can account for up to 10%
of total fuel consumption [11]. In well-developed idle stop systems, the fuel economy is
estimated to be between 7 and 9% [12].
The vehicle electrification, although it can be considered a feasible alternative
to meet the regulations for emission and fuel consumption reduction [11, 13], faces
several challenges due the lack of infrastructure and its high costs [14]. On the other hand,
hybrid systems, that combine the power source of the internal combustion and electric
engines [5, 15], have steadily become a popular design concept. The start-stop systems
can be basically divided into the three following developing stages [16-18]:

 Micro hybrid: The electric engine is only used to store energy


(Regenerative Braking) and to switch off the engine in idle conditions,
restarting it upon driver's demand or when needed (Start-Stop);

2
 Mild hybrid: The electric engine has more capacity and can help to
accelerate the engine or store energy from the engine break, and to work
as a start-stop, but cannot run the vehicle alone;
 Full hybrid: The electric engine can power the car alone for some time
at mild speed. In this configuration, the electric motor is used at low speed
city traffic and to recover kinetic energy of the ICE during braking. The
ICE is used only at high speeds [19], optimizing the operation of the
power sources [20, 21].

In this scenario, besides being widely used in vehicles with ICEs, the start-
stop system appears as an essential technology to hybrid vehicles, causing the combustion
engine to be switched on and off. To guarantee good drivability in hybrid vehicles, special
procedures should be implemented to perform a quick switch between the two engines
[22], and specially, a quick ICE start [23]. Naturally the start-stop operation can cause
additional vibration and noise [24, 25], affecting the passenger comfort and increasing
torque gradients during engine start [6], what can be extremely damaging for many engine
parts.
Two well-known observable tribological phenomena in car engines with
start-stop systems are the lubricant starvation among contacting surfaces of the engine
moving parts, such as cylinder and crankshaft, and piston and bearings, and the local
temperature rise [26, 27]. The lubricant film in between those contacting surfaces
experiences lubrication regime transition, from boundary lubrication at the stroke end,
when the velocity reversion inhibits the hydrodynamic action of the lubricant film,
passing through the elastohydrodynamic lubrication in the mid-stroke phase, to the
hydrodynamic regime during the continuous engine operation [28-31]. This regime
transition occurs during every engine start-up, and its frequency in start-stop systems
causes a regular rupture of the lubricant film, which can lead to severe surface asperity
contact and, hence, to high frictional energy loss and wear damage [32].
Advanced studies of the ICE moving part lubricated contact phenomena
under stringent conditions are extremely important to evaluate accurately if the fuel
economy achieved by idle stop of hybrid operations is not harmed by increasing the
friction losses or if it causes irreversible damage in several engine parts. In this context,
experimental investigations of engines under stop-start conditions and tribological
evaluation of the engine components to estimate the damage on the most affected parts,
such as piston, cylinder walls and bearings have been developed [32]. These tests are
important both for the engine design and component manufacturing, which can predict
accurately the operating life of various engine components and search for new
technologies to reduce friction and wear. Among the new technologies, low friction
coatings based on diamond-like carbon or molybdenum-based composites, surface
finishing to reduce asperity contact and surface texturing by laser techniques, and
modifying the lubricant flow in contact zones can be highlighted because they are able to
reduce the negative impact of the friction and wear [33].

3
TRIBOLOGICAL FUNDAMENTALS OF THE CRANKSHAFT-BEARING
SYSTEM

A tribological system of great interest for evaluating the impact of start-stop


systems on engine life is the crankshaft supporting system. An initial understanding of
the tribological aspects of the crankshaft-bearings systems is very important to study how
the friction, lubrication and wear behave under start-stop conditions.

Engine Friction

Friction is associated with the resistance to the motion between two


continuum media caused by the interaction between their contacting surfaces. The
greatest cause of mechanical friction in an engine is the sliding friction between piston
and cylinder wall, followed by the sliding friction between the journal bearings and the
crankshaft, and the auxiliary systems, such as fans, pumps and magnets [34].
According to [35], the sliding friction force arises from two sources: the
adhesion force developed at the asperity contacts and the deformation force needed to
plough the asperities of the harder surfaces into the others [36]. Even highly polished
mechanical components present, in some level, surface roughness. When two surfaces
interact, the normal load increases the area of contact due to elastic deformation. When
these surfaces slide over each other, the smaller contact the irregularities have, due to
good surface polishing or low pressure, the higher the elastic profile of the deformation
[36]. Hence, the type of deformation occurring during sliding has great influence on the
friction force[34].
When two unlubricated surfaces slide in relation to each other or are in
impending condition of relative motion, a friction force acts on both surfaces in the
opposite direction of the sliding direction. In the cases where no relative sliding occurs
(incipient sliding), the energy for the surface motion is not present to overcome the
contact of the irregularities from both surfaces. This phenomenon, opposed to the sliding
movement is then called static friction force, which can be greater or equal to the kinetic
friction force, which is opposed to the sliding movement when enough energy is provided
to the system.
For both cases, according to the Amontons´ law, the friction force, F, the load
acting normal to the surfaces influencing adhesion, W, and the coefficient of friction, f,
determined by the physical properties of the contacting surfaces, can be related by
Equation 1. The friction coefficient depends mainly on the contacting body material
properties and the mechanical roughness of the surfaces. It varies during the sliding
process, due to work-hardening and junction growth, effects caused by plastic
deformation, which can lead to damage processes and temperature rise that potentially
can modify the surface characteristics [34, 36]. For both dry and lubricated contacts of
solid bodies, the static and kinetic friction coefficients of many engineering materials are
dependent on the normal load, sliding velocity and apparent contact area [37, 38]. Hence,
the friction coefficient can be considered constant only for a given pair of sliding
materials under a given set of operating conditions (temperature, humidity, normal
pressure and sliding velocity)[39].
𝐹
𝑓=𝑊 (1)

The presence of the friction force in the contact surfaces of engine


components causes energy dissipation, which reduces the overall engine efficiency. In

4
addition, the heat produced at the contacting surfaces, when work is done against the
friction forces, may cause melting or seizure in some hot contacting points, damaging the
engine components [34, 40]. Some procedures to reduce the friction force are the
introduction of a lubricant film with low shear strength in between the two sliding
surfaces and the employment of bearings manufactured with special materials and
coatings [36].

Lubrication

The purpose of the lubrication is to reduce the friction force and energy loss
associated with the surface sliding by separating the moving surfaces with a layer of
material with low shear strength. Although in some cases the lubricant may not
completely separate the surfaces, lubrication can reduce the irregularity contact or the
strength of their junctions. Thereby, to a greater or lesser extent, lubrication can decrease
friction and wear [34, 36, 40]. According to this description and considering various
potential lubricant materials, four types of lubrication regime can be defined:

 Hydrodynamic: Condition in which the surfaces are separated by a


lubricant film with thickness greater than their asperity heights and the
pressure of the film causes negligible surface elastic distortion;
 Elastohydrodynamic: Condition in which the lubricant film is relatively
thin, local pressures are high and, therefore, elastic deformation of the
surfaces cannot be neglected;
 Boundary: Condition in which considerable irregularity contact occurs
and lubrication cannot totally prevent junction and wear;
 Solid: Condition in which solid materials of low shear strength are used
instead of fluid or semi-fluid films to separate sliding surfaces.

In the technical literature of Tribology, the term mixed lubrication is usually


used to describe the transition between hydrodynamic/elastohydrodynamic and boundary
regimes. In these cases, there is frequent solid contact, but at least a portion of the bearing
remains under hydrodynamic lubrication, which can still prevent adhesion between
surfaces during most asperity encounters. Hence, the transition between the
hydrodynamic/elastohydrodynamic and boundary lubrication in this study is referred to
as mixed lubrication [39].
During engine start, at extremely high loads or low sliding speeds, the fluid
film is either not built up or the fluid film forces are insufficient to keep a continuous
fluid film between surfaces, leading to film rupture and direct contact between asperities.
At this condition, high friction and wear will occur. For some authors, this condition is
called “inefficient lubrication” [41]. In further interaction stages between contacting
bodies, the conditions called “scuffing or pre-seizure”, “delayed seizure”, and “immediate
seizure” take place, due to the thoroughly incomplete lubrication [42, 43]. With the use
of start-stop systems, the frequency in which the engine stops increases drastically, with
crankshaft rotation being responsible for generating the lubrication pressure between
engine components. The lubricant film sustains the journal rotation, preventing the solid
contact to the bearing pads. Hence, the start-stop systems increase the number of times
that journal bearings operate at boundary lubrication regimes, increasing consequently
the friction and wear associated to this regime [9].
When an oil film is present between surfaces but the engine is operating under
severe load application, local fluid film rupture can occur, leading to line or point

5
contacts, causing conditions for elastohydrodynamic lubrication. This local contact
increases local pressure that will be generally much higher than the hydrodynamic
pressure. At this condition, the film thickness represents a small percentage of the bearing
clearance and can be smaller than the size of several particles transported by the lubricant
[44]. Under this condition, the lubricant viscosity under pressure and the surface elastic
deformation play important roles. The contact area and pressure distribution are modified
due to elastic deformation of the surfaces and the effect of film thickness on normal load
is not so strong, whereas sliding velocity and viscosity have higher effect. Numerical
procedures usually are employed to analyze the elastohydrodynamic lubrication regime
in engine bearings [36].
However, during most of the engine operation, its components should operate
at hydrodynamic lubrication regime. In this case, the friction does not depend on the
normal load and thus, is practically independent of the material roughness [34]. When a
shaft rotates on a journal bearing, some of the lubricant adheres to the shaft and is carried
around it making the fluid film thickness under the shaft smaller than on the lubricant
leading edge, indicating the presence of a journal eccentricity. Viscous shear stresses vary
along the convergent oil film thickness, generating a hydrodynamic pressure distribution,
which separates the contacting surfaces [36]. For small loads and high speeds, the
rotating shaft tends to operate concentrically. On the other hand, for high loads and low
speeds, the journal tends to operate eccentrically [36].
The viscous friction on hydrodynamic bearings depends on the oil viscosity,
engine rotation, and bearing geometry, but not on the surface mechanical properties. The
normal load can cause a squeeze effect on the lubricant film, reducing the film thickness,
what can be observed in the end connecting rod and main bearings during the engine start
up [36].

Bearing Lubrication Regime

The friction between the journal bearing and the crankshaft during the engine
start experiences lubrication regime changes, from the boundary lubrication conditions to
the hydrodynamic regime conditions, passing through the elastohydrodynamic regime
conditions. In the highest loaded points, the fluid film can collapse, promoting direct
contact between shaft and bearing, increasing wear rates and damaging both shaft and
bearing surfaces [8]. Several procedures have been used to identify the lubrication regime
between contacting solid bodies [9].
An important concept widely used in Tribology to characterize the friction
in fluid-lubricated contacts is the Stribeck curve [45]. In the original Stribeck curve, the
bearing friction coefficient is related to the bearing load and journal speed, resulting in
the curves shown in Figure 1. From those curves, it is possible to estimate the operating
point of minimum friction for lubricated applications. It can be observed from those
curves that the friction coefficient for sliding bearings reaches high values at low speeds,
decreases to a minimum value when the lubricant film separates completely the
contacting solid bodies, and then increases almost linearly as the rotating speed increases.
Another form of presenting the Stribeck curve is depicted in Figure 2, which
allows to classify the different lubrication regimes. In the boundary lubrication regime,
with low rotating speeds, large journal eccentricity occurs, leading to close contact
between surface asperities. At this regime, when relative contacting surface sliding takes
place, the friction coefficient is extremely high, and the wear and friction are substantial.
As the sliding speed increases, more lubricant enters into the contacting area, reducing
the friction due to the lubricant low shear stress. Consequently, the energy necessary for

6
the movement becomes smaller, making the relative sliding between the solid surfaces
easier. As the lubricant shearing velocity increases and the lubricant pressure decreases,
the contact among surface asperities ceases to exist and the friction and the surface wear
tend to be reduced. This condition of lowest friction is called “Release point” and is the
recommended operating condition. As the film thickness increases, the surfaces become
totally separated by the lubricant and the friction coefficient starts to increase with the
lubricant shear velocity, which can cause increased viscous dissipation due to the
lubricant shear resistance [46]. For most fluid film bearing applications, the supporting
systems should be designed to operate more often under the hydrodynamic regime. The
boundary and elastohydrodynamic regime conditions should be avoided, because the
potential solid contact between rough surfaces can cause wear and superficial damage
[36].

7
Figure 1 – Original Stribeck curves [45]

8
Figure 2 - Stribeck Curve, Credits: Adapted from [47].
Therefore, the engine start forces the lubrication regime variation in many
engine components. At the points of high load, the fluid film can collapse, which can
provoke the direct contact between the shaft and the bearing, increasing wear rates and
damaging the contacting surfaces [8]. The bearing lubrication regime can be
approximately determined by analytical calculation, accounting for the pressure field on
the bearing, the lubricant flow and the bearing geometry. Two important bearing
parameters in this calculation are the Sommerfeld number and the journal bearing
clearance. Basically, the bearing characteristics permit to estimate a limit engine speed
at which the fluid film can be formed to separate the contacting bodies. However, if the
engine speed is below this limit, the fluid film can be ruptured [46].
The bearing lubrication regime can also be approximately determined by
computing the limit sliding speed at which the lubricant film between bearing and shaft
can be maintained during engine operation. According to Williams [48], at the first cycles
of the shaft rotation, there is solid contact between the shaft and the bearing, and the
contacting area can be initially estimated from the bearing geometry and external load. If
the shaft rotates in the presence of a lubricant, the shaft will drag the lubricant into the
convergent gap at the film leading edge. If the shaft speed is adequate to generate a
hydrodynamic pressure capable of supporting the applied load, the two contacting
surfaces are separated, as shown in Figure 3. At high engine speeds the shaft tends to
operate concentrically [39].

9
Bearing
Lubricant
Shaft

Motion

Load

Figure 3 - Lubricant behavior during shaft rotation; Adapted from [34]


The shaft speed capable of generating a hydrodynamic pressure for shaft
levitation can be determined by equation 2. In this equation, the pressure distributed on
the lubricated contact area is Pb, the bearing Sommerfeld number is expressed by S, the
relative bearing clearance is given by , and the lubricant dynamic viscosity is
represented by  [46],

𝑃𝑏 𝜓2
𝜔𝑏 = (2)
𝜇𝑆

Hence, if the shaft rotation exceeds this speed, the bearing will operate under
hydrodynamic lubrication regime. Otherwise, the pair will be under mixed lubrication.
Hence, until the engine reaches the separating rotating speed during its start, the bearing
and journal surfaces can be under solid contact conditions, which cause wear on both
parts [49]. With the start-stop systems, the frequency of solid contact can increase, and,
consequently, the wear tends to increase [8].

Engine Oil

Engine lubricant oils have different functions, mainly related to frictional


resistance, temperature control, shock absorption and corrosion prevention [50]. The
lubricant viscosity is one of the most important oil characteristic in the bearing behavior.
The lubricant presents fundamental importance at both low and normal engine
temperatures, providing low resistance during the engine start and rapid distribution of
the oil while the engine is cold, also promoting sealing of the piston, acceptable oil
consumption and low hydrodynamic friction losses when the engine is fully warmed up.
This constant development of engines usually demands greater stresses for
engine oils, also boosting the development of lubricants. Besides of meeting efficiency
requirements and environmental regulations, engine durability, performance and
reliability are fundamental to vehicle owners [51]. According to Johnson [52], the
development of lubricants for modern engine systems, including engines for hybrid
application, must take into account the bearing frequent operation at boundary and mixed
lubrication regimes, that can be observed during the engine starts and stops. Liu [53]

10
states that hot start-stop conditions lead to more serious asperity contact friction in the
early stages of the engine start-up, while cold start-stop condition generates more friction
losses. Moreover, Liu observed that the wear process of bearing surface can be
accelerated when oils with low viscosity are used.
However, the use of low viscosity engine oils (LVEO) can be a cost-effective
way to increase engine efficiency, since they can reduce the friction losses in engine tribo-
contacts, which represent nearly 10% of the total engine losses [54]. Considering the
current technological trends of ICEs, for example the development of start-stop systems,
fluid film bearings will tend to run at boundary and mixed lubrication regimes [55]. By
this, new low oil viscosity grades were also introduced by the society of automotive
engineers (SAE) [56], including SAE 0W-16, SAE 0W-12 and SAE 0W-8 viscosity
grades. Japanese and European manufacturers are also using low viscosity lubricants, as
0W-20 and 0W-16, for fuel saving in advanced engines [57]. Following this trend, new
heavy-duty engine oil categories CK-4 and FA-4 have been proposed by API (American
Petroleum Institute), whose viscosity exceeds the historic High Temperature High Shear
viscosity (HTHS) limit of 3.5 cP. In Europe, the HTHS value remains at 3.5 cP [58]. In
this sense, the use of advanced friction modifiers and antiwear additives also presents
potential tribological solutions as well as the low viscosity oils. In general, the original
equipment manufacturers (OEMs) are deeply interested in the effects of surface coatings
on friction and wear of low viscosity lubricants [59].
Also in order to prevent severe asperity contact and high friction, the use of
friction modifiers (FM) and anti-wear additives in lubricant oils can present beneficial
results [60]. They could form surface films that will influence the friction coefficient
under mixed/boundary lubrication conditions. The friction additive modifiers (FM) can
be basically divided into four main classes: the organic friction modifiers (OFM), the
organo-molybdenum compounds, polymers and dispersed nanoparticles.
The OFMs are basically free fatty acids derived from fats and vegetable oils
and normally present in additives of modern engine oils, and also in fuels. Their
amphiphilic molecules can facilitate the movement of other lubricant particles. The
monolayer can also withstand high pressures, which is a very positive characteristic in
engine lubricated by oil streams. However, the friction behavior of these additives can be
quite different under different friction speeds [61, 62]. Furthermore, although the most
studies proving OFMS effectiveness are conducted in saturated chains, the majority of
commercial OFMs, aiming cost reduction and solubility increase, are composed mainly
of unsaturated alkyl chains. Even though OFMs can be effective to produce lubricants
able to reduce friction, they are corrosive to metallic surfaces and are not recommended
for bearings employed in engines [63].
The organo-molybdenum compounds, initially used just as an anti-wear
additive, are nowadays also very effective in reducing boundary friction. Under sliding
conditions, these compounds form nanosheets of MoS2 on surfaces. However, these
sheets are formed only in the load-bearing asperity tips. It is observed that molybdenum
dithiocarbamate (MoDTC) is able to form films to reduce friction on a very wide range
of surfaces, including diamond-like carbon (DLCs), the tribofilms formed by zinc
dialkyldithiophosphate (ZDDP) [64] and ceramic coatings [65].
ZDDP has been used as an effective antiwear and antioxidant additive, which
can generate a strong phosphate tribofilm [52]. However, despite its effectiveness, the use
of this additive has been discouraged, because phosphorus and sulfur from ZDDP
decomposition can cause the accumulation of catalysts, what can provoke permanent
damage [66]. Then, investigations have been performed to evaluate alternative lubricants

11
to improve lubrication performance against wear, using DLC [67], Al2O3 or SiO2
nanoparticles [68].
Polymers can be designed specifically to be absorbed on specific polar
tribological surfaces, presenting good performance in some contact conditions. The
viscous modifier (VM) reduces the friction by separating surfaces with a pressurized fluid
film at low sliding speeds using only the base oil. One hypothesis for this behavior is an
increase in the local viscosity close to the surface. However, this phenomenon is not yet
understood. Some evidence shows that the polymers do not reduce friction properly in
severe reciprocated contact conditions, what is also not yet proven. Different from OFMs,
polymers are not able to self-organize in dense packings along the surface, a problem that
can become very serious as the molecular weight increases. One solution is to grow
polymers on surfaces through initiators pre-attached the surface, enabling the growth of
very dense polymer films, which present low friction at high and low speed conditions.
Unfortunately, due to difficulties of lubricant replenishment, this solution is not practical
for ICEs. The use of Polyisoprene as FM has presented satisfactory results, due its
capability of forming thick boundary lubricating films [69, 70].
A novel technological trend in additives is the use of dispersed nanoparticles,
since some studies have shown their good capability to reduce boundary friction. The
performance of nanoparticles as additives is strongly related to a specific application,
since nanoparticles can present very different behavior in different tests. For example,
Jatti [71] evaluated experimentally the tribological behavior of copper oxide
nanoparticles as additives in mineral based multi-grade engine oil, performing tests under
different loads and concentrations of nanoparticles. Results indicated that there is an
improvement in the lubricant properties as a function of the additive potential of entering
into the friction zone along with the flow of lubricant, through the deposition of soft CuO
nanoparticles on the worn surface, which decreases the shearing resistance and improves
the tribological properties. With this, nanoparticles can convert sliding friction into
rolling friction, reducing the effective friction coefficient in up to 50% with the addition
of CuO nanoparticles to the base oil. Esfe et al. [72] observed an reduction in viscosity of
nano oil when compared to a base oil, pointing to a considerable effect on cold start
performance. Other aim of nanoparticles use is to provide lower dependency of viscosity
to the temperature when compared to pure engine oil, especially in higher temperatures.
However, the nanoparticles can increase wear by abrasion according to their hardness. It
has been shown that hard nanoparticles are capable of removing antiwear films, while
soft nanoparticles, such as CuO, are not [73]. Recently it has been reported that carbon
nanoparticles can prevent MoDTC from reducing friction depending on the particle size
[74]. Consequently, although several research on Nano additives has been successfully
conducted, this has not yet been accompanied by their use in commercial liquid lubricants
[63].
According to Barhnhill [75], the development of engine oils is a growing
challenge. On the one hand, a higher viscosity oil can reduce the possibility of lubricant
films to become deleteriously thinner as sliding speed decreases and temperature rises
near the top ring reversal region of a stroke. On the other hand, a lower viscosity lubricant
would be preferable to reduce parasitic friction, however resulting in a challenge for wear
protection. By this, the development of lubricant additive formulations that provide
friction reduction while retaining anti-wear benefits, are a logical path for investigations
in order to resolve the lubrication problems.

12
Engine Wear

Wear, as well as friction, is a characteristic of engineering systems, and


represents material dissipation. It can be determined as the progressive loss of material
from a solid body surface due to mechanical action, normally related to relative motion
of surfaces or its contact with a solid, liquid or gaseous contacting body [76, 77]. It is
rarely catastrophic, but it not only may represent energy losses and reduction of operating
efficiency, but also it may produce dimensional changes of components or surface
damage, which can cause other mechanical problems, such as vibration or misalignment.
Moreover, the generation of wear debris, mainly in systems with small clearances, may
be more damaging than the component dimensional changes [78].
The type of mechanical contact is very important for evaluating wear losses,
mainly during instantaneous contact between solid surfaces, without any interfacial
medium, such as adsorbed layers, oxides, lubricants, dirt, etc. [79]. The normal contact
force between surfaces generates the contact among asperities and is directly related to
surface wear. The number of asperity contacts can increase as the normal load increases,
also modifying the contacting surface patterns. Also an increase in the normal load cause
an increase in the deformation of the asperities, that can lead to a shift from the elastic to
plastic contact [80].
Wear can be caused by several mechanisms of energetic and material
interactions among the elements of a tribosystem. The main wear mechanisms addressed
in this investigation will be the sliding and abrasive wear. The sliding wear is associated
with the surface damage caused by relative motion between two smooth solid surfaces in
contact under load. The load creates adhesive junctions among surface asperities and the
relative motion between surfaces lead these junctions to be torn away, forming new
irregularities and causing progressive loss of material. This wear mechanism is highly
influenced by the type of contact between solid surfaces, which can be elastic or plastic,
or totally separated by a lubricant [79].
The abrasive wear is the most rapid form of wear and it is basically the surface
loss of material by the passage of particles of equal or greater hardness. It normally
involves both plastic flow and brittle fracture and its prevention is a hard task, because
different wear mechanisms can be involved in the contact between the surface and the
particles. Normally, the application of a hard material or hard coating is used to solve this
problem. Ceramic and polymeric materials can be sufficiently hard and homogeneous,
being frequently used to coat materials with low hardness, as steels and non-ferrous
metals [81].
Sliding wear is a very usual phenomenon in the dry and lubricated contact of
solid bodies. Estimating the wear rate is very important in the design of durable and safe
mechanical components. Good maintenance practices can prevent unexpected component
failures and permit to perform substitution of defective parts with minimum time and
cost. Measuring the surface wear during some time of equipment operation and
extrapolating the measured wear rate can help to predict the component lifetime. This
basic procedure, however, may not be accurate as wear rates in some systems are not
constant throughout time. They can decrease after an initial operating period (normally
in lubricated systems where sliding wear is dominant) or they may increase after an initial
incubation period (if the material coating is removed due to wear and the wear behavior
of the lower layer is different)[8].
Wear can be also estimated by theoretical or empirical equations relating wear
rate with the operating parameters, such as load and speed. A simple model for wear
brought by Holm and Archard suggests that the amount of material removed from sliding

13
body is dependent on the sliding distance and the nominal pressure (load divided by the
nominal contacting area. The model can approximately estimate wear, as presented in
Equation 3, which relates wear rate, Q, which is the volume removed per unit of sliding
distance, to the normal load, W. In this equation, Hs is the hardness of the softer surface
in contact, and K is a dimensionless wear coefficient dependent on the state of lubrication
and material properties.

𝐾𝑊
𝑄= (3)
𝐻𝑠

Generally there is no direct correlation between wear rate and coefficient of


friction [81]. It is important to highlight that Equation 3 provides an expression for wear
estimation, but the parameters needed for the calculation can vary as the sliding
progresses. In order to obtain more accurate wear estimations, it is crucial to perform
analytical-numerical simulations or experimental tests. Equation 3, however, represents
a valuable engineering means of comparing the severity of wear processes in different
wear systems for different materials. Specifically, the quantity K/Hs or k (mm3/Nm) is
called the dimensional wear coefficient that represents the volume of material removed
by wear (mm3) per unit of normal load (N) per sliding distance (m), which can be very
useful [36].
The wear rate estimation can be also experimentally performed through
component tests, in which wear is evaluated at conditions expected in the component
service, this procedure has been frequently employed to predict wear in bearings, under
different loads, speeds, temperatures and states of lubrication. Tests with accelerated wear
evolution must be performed cautiously to guarantee that the wear mechanism is
compatible with the actual mechanism and that relevant operating conditions can
represent the real case [81].

BEARING CHARACTERISTICS

Bearings are used to allow the relative motion between moving mechanical
parts with minimum friction. Fluid film bearings have been widely employed to separate
surfaces with relative sliding motion to each other. One of the most common sliding
bearing type is the oil-lubricated cylindrical journal bearing, which is basically a metallic
sleeve with circular bore. The separation between the surfaces is provided mainly by the
flow of a viscous lubricating film, which generates a pressure field able to support the
external loads. The bearing geometric characteristics must be adequate to separate
efficiently surfaces in relative motion. When a continuous thin lubricant film is formed
between the surfaces, material properties do not play an important role in the process.
However, when the shaft is not rotating or the lubricant film is squeezed out from the
bearing contact area, which are frequent conditions during start stop operation, the
bearing system must be able to operate under poor lubrication [36]. Some fluid film
bearings in automotive engines are the connecting rod (conrod) bearings for the large
connecting rod eye, the connecting rod bushings for the small connecting rod eye, the
main bearings, the fitting or guide bearings, and thrust washers [82]. Among them, the
most affected by start-stop applications are the connecting rod bearings and the main
bearings. In this section, some design aspects as well as the material selection of these
two bearings will be discussed.

14
Some Design Aspects

For oil-lubricated bearings, the lubricant flow through the sliding surfaces is
associated with some bearing geometric characteristics, such as clearance, grooves and
holes. The peak of oil film pressure (POFP) and the minimum oil film thickness (MOFT)
are strongly associated with the bearing diameter (D) and length (L) or width. Larger
bearing length or larger bearing diameter can reduce the POFP and increase the MOFT.
In the design of journal bearings, the value of the bearing slenderness ratio L/D must
usually be maximized to obtain low values of oil film pressures and large values of oil
film thicknesses, since that the value of this ratio can guarantee the bearing mechanical
integrity at severe operating conditions.
The journal bearing clearance is the designed geometric difference between
the bearing and shaft radii, through which the lubricant flows. Small clearances can
provide higher values of hydrodynamic pressure, making the bearing able to support more
efficiently external loads. However, small clearances cause more viscous dissipation,
leading to an increase in the lubricant temperature, which reduces the oil viscosity. The
POFP normally increases and the MOFT decreases as the bearing clearance decreases
[82]. The bearing lower limits are usually equal to 0.6% of the shaft diameter for the
connecting rod bearings and to 0.75% for the main bearings [44].
The bearing grooves and holes allow the lubricating oil pass the channels until
the bearing surface. Usually grooves and holes are etched in the bearing pads or in the
shaft to improve the oil circulation. The groves can be machined in the partial or total
area of the bearing. Even though grooves and holes play important roles in lubricant flow,
they are undesirable on high-stress points, because they can cause an increase in the POFP
and a reduction in the MOFT. For high loads, the risk of solid contact between sliding
parts increases as well the possibility of damage caused by cavitation. Thereby, holes
and grooves are normally introduced only in the upper main and lower conrod bearings,
which are subjected to lower loads during the compression stroke [8, 82].

Material Selection

The selection of the material for all engine components subjected to sliding
contact is a very important task in the design of these components. Geometric and
operating characteristics of the bearings must be well known to permit the selection of
the most adequate material for durable, safe, and efficient designs. For convenience, this
section is divided into two subsections: General material selection and bearing material
selection.

General Material Selection

The material selection for mechanical components must account for several
design requirements and parameters, such as costs, corrosion resistance, mechanical
properties, and also tribological surface properties. Metals are the most common choice
of materials for the design of automotive mechanical components, because their
composition and microstructure are largely standardized and their overall properties are
usually well estimated. On the other hand, non-metallic materials are not so well
standardized and their properties tend to vary more widely.
For both metallic and non-metallic materials, tribological properties of
materials are more difficult to be quantified, different from the physical and mechanical
properties. Hence, for components such as bearings, in which the tribological properties

15
must be accurately determined, the selection of materials is generally based on analytical
or empirical expressions and on experimentation.
For sliding wear, the Archard equation can help to indicate how material
properties affect wear [80]. Even though the only explicitly property in the equation is
the hardness of the softer surface, Hs, some material properties are considered in the wear
coefficient, K. Especially when the contacting surfaces are not separated by a lubricant
film, the tribological compatibility of the two surfaces are of extreme importance.
Compatibility means the capability of the contacting surfaces to form an interfacial bond,
which would lead to high wear rates.
Solubility is one important factor impacting the tribological compatibility.
Surface films, such as oxides, can be very influential on the behavior of metallic surfaces.
This is observed when metals present normally low wear rates and the oxide film stability
can determine the dominant wear mechanism of the component. The hardness of metals,
which are capable of forming oxide films during sliding, is an important parameter to be
considered, because it can help to determine the stability of the film. If a metal is hard, it
is expected that its oxide layer is also hard, increasing the mild wear phase and decreasing
the wear rates. The presence of microstructural features, such as carbides and nitrides in
steels, the lubricants whose constituents are graphite and molybdenum, and the
microstructure formation of some metals, such as hexagonal metals, can have more
impact the wear rates than the material hardness.
Ceramic materials can exhibit very low wear coefficients, but present some
drawbacks. Their mechanical properties may not be adequate for the component
operating requirements, the manufacture of some special geometrical features can be
unpractical and severe wear can lead to surface fracture. On the other hand, ceramic can
be used efficiently in surface coatings. Ceramic and other type of hard coatings, surface
work hardening procedures, rough surfaces in general and ductile diffusion layers can
confer good surface wear resistance, since they present low solubility with other materials
and present limited junction growth. Polymers are normally used in bearing materials,
mainly under marginal or dry lubrication. The most common polymers used in
tribological applications are the nylons (polyamides), acetal, polyetheretherketone
(PEEK) and polyethersulphone (PES) [36].

Bearing Material Selection

The adequate choice of bearing material is based on the engine component


application and the material properties. The bearing load profile can describe the
mechanical and tribological requirements of the bearing. The material selection is always
the result of a compromise among all properties, which frequently counterpose each other.
The main properties desired in a bearing are listed as follows.

 Resilience: Ability to endure mechanical loads permanently without


developing fatigue cracks;
 Wear Resistance: mainly under the conditions of mixed lubrication
regime and in the presence of foreign particles circulating with the
lubricant;
 Compatibility (Seizure Resistance): Ability of the bearing material to
resist physical joining with the journal material at the lubrication area,
where bearing and journal materials can come to a direct contact;
 Embeddability: Ability of the material to absorb hard particles on the
sliding surface and circulating in the lubricant oil;

16
 Conformability: Ability to compensate and accommodate
misalignments prevenient of geometric variations in local contacts of the
journal, housing or bearing [8, 82];
 Corrosion Resistance: Ability to resist corrosion caused by weak organic
acids and strong mineral acids derived from fuel combustion products and
weak organic acids, formed as result of the oxidation of lubricating oil
[82];
 Cavitation Resistance: Ability to withstand impact stresses caused by
collapsing cavitation bubbles, which are formed as a result of sharp and
localized drops of pressure in the lubricant flow [82, 83];

Usually the design of a fluid-flow bearing assumes that the bearing and
journal surfaces will be fully separated by the oil hydrodynamic action. However, it is
important to account for the potential of solid contact during the shaft start, stop or
overload in the selection of the bearing material for poorly lubricated contacts. It is
important, then, that the bearing material exhibits compatibility with the journal material
to tolerate the occurrence of surface contact along with compatibility, embeddability,
fatigue resistance, conformability and corrosion resistance.
No single material offers an ideal combination of properties and, sometimes,
they are mutually incompatible. Generally a bearing is composed of material layers, and
each layer presents a specific function, so that all layers together provide the required
bearing properties [84]. Bearing materials can be practically divided into two major
categories: metals and non-metals. The metals most commonly used are tin- and lead-
based alloys (babbitts), copper-based alloys (brasses and bronzes), aluminum-based
alloys, cast iron, and porous metals. The non-metals are carbon-graphites, plastics,
elastomers, ceramics, cermets, among others. These materials can be either used as bulk
materials or as a lining on a bearing surface. In the case of lined bearings, the bearing
material is bonded to a stronger backing material such as steel. Many soft metals, such as
carbon-graphites, plastics, and elastomers, are used to slide against harder materials such
as stainless steel under unlubricated or poorly lubricated conditions. Also in many bearing
recent applications, materials are used as coatings by various deposition techniques, in
order to reduce friction and wear [39].
The modern bearing design has employed powerful computational resources
and sophisticated test rigs to evaluate the behavior of a specific bearing configuration for
different materials and under several load conditions. The operating and geometric
parameters for the main and conrod bearings are estimated with large design safety
coefficients, accounting for possible misalignments. Depending on the test results for
each material and application, it is possible to evaluate if the bearing with just a lining is
capable of withstanding the operating conditions, or other material layers are needed.
Thus, bearings can be distinguished into bimetallic and trimetallic bearings [82, 85]:

 Bimetallic bearings: They are normally used for bushings and thrust
washers, and usually consist of a supporting steel shell coated with a
lining of aluminum, bronze alloy or white metal. Their tribological
performance is determined by the bearing lining itself [84]. They are also
normally used in the low loaded conrod and main bearings shells, due to
their lower costs ;
 Trimetallic bearings: They have also a supporting steel shell coated with
an aluminum or bronze alloy lining. However, an intermediate layer with
thickness varying from 1 to 4µm is deposited on the top of the lining as a

17
diffusion barrier. Finally, an overlay with thickness of some tenths of
microns is electroplated or sputtered on the top, what improves the
required tribological properties, such as friction, seizure resistance,
conformability and embeddability [86]. Due to the material variety used,
the trimetallic bearings can provide more load capacity, achieving optimal
combination of individual properties. The sputter variables are under
development to endure the new engine concepts, with different load
capacities [83].

For the material selection it is also important to consider the type of


lubrication used during the bearing operation. Bearings have normally a steel backing and
a bearing lining of a few hundred micrometers thick applied on top. Considering that
bearings in start-stop operation will remain long with boundary lubrication or in direct
contact with the shaft, they can be considered marginally lubricated [36].
Among the various materials investigated for use in engine bearings, lead
(Pb) containing materials such as copper-lead (Cu–Pb) based linings and overlays have
been often used in the past due to the low friction characteristics and excellent
conformability and embeddability [87]. However, environmental concerns regarding the
widespread use of lead has entailed manufacturing restrictions and lead-free (Pb-free)
materials have being developed recently [88]. Various Pb-free bearing materials are
suggested by bearing manufacturers. For instance, Al–Sn based overlays have been
developed because they present better wear resistance when compared with lead based
(Pb-based) material [86]. These tin-based materials also have presented high resistance
to cavitation and corrosion damages, presenting excellent conformability and
embeddability for engine tribological applications [89]. Nowadays, for regular engine
applications, Al based crankshaft bearings are used for low to medium specific loads (≤70
MPa), while Cu based (CuNi2Si) bearings are used for medium to high specific loads.
A more recent concern of the automotive industry has been the selection of
bearing material capable of providing durable designs for severe lubrication conditions,
such as the engine start-stop operation. Passenger vehicle engines designed for start-stop
applications are required to withstand 250,000 to 300,000 starts. But, traditional bearing
shells with aluminum or copper lining show unacceptable wear after only 100,000 cycles.
Hence, new overlays on the bearing surface have been researched and developed [90].
In general, the bearing durability depends on the material top layer. Hence,
due to the new bearing design requirements, bearing manufacturers have been developing
intelligent materials for bearings coatings. For instance, materials capable of inducing
diffusion of a near layer material to the coating layer in order to thin the sliding layer and
increase the concentration of hard particles by load increase have been developed. This
technology can increase by 25% the bearing load resistance and also increase the bearing
resistance to wear [91]. Bearing coatings with concentration of graphite and
Molybdenumdisulfide (MoS2) as solid lubricant and mixed oxides for micro
reinforcements to improve seizure properties are also a solution for the latest start-stop
applications [9, 92].
Special attention has been given to Polyamide-Imide (PAI), an amorphous
thermoplastic with high mechanical strength, high resistant against diesel fuel, chemical
components and thermal degradation. This material is being extensively studied and it is
believed to be further improved by incorporating solid lubricants. In recent studies,
bearing materials with PAI based overlay containing graphite and MoS2 exhibited better
friction and wear properties than Pb-based and aluminum-tin (Al–Sn) based materials

18
[84], and similar seizure behavior as Pb-based coatings [93], indicating the potential of
this material.
Some results of other studies show that the coatings of PAI with several
additives, hard particles and oxide particles evenly dispersed presented lower friction and
were able to resist chemical attacks and wear. This overlay can be applied and bonded
equally to the typical bearing linings, such as aluminum or copper or bronze, and enables
the application of Al based bearing shells in situations that were possible only for Cu
based linings [90].

STATE OF THE ART

Regulatory agencies basically perform inspections to verify the engine


compliance with current emission and consumption standards, as well as oil quality and
mechanical requirements. However, engine and automotive mechanical component
manufacturers are responsible for developing technologies to improve the tribological
design of ICEs that can guarantee the integrity and quality of the equipment and offer
minimum risk to consumers and the environment [9, 94].
The start-stop systems are relatively recent in the automotive market, and so
is the development of parts for the engines with this technology. The wear estimation is
extremely important for the mechanical design of all engine components and can provide
a prediction of the component lifetime, which can prevent economic losses caused by
component failure or component reposition during the equipment operation. There are
basically 3 procedures to estimate the wear:

 Extrapolation from measurements on the actual system: the wear is


measured during some time during the equipment operation and
extrapolated to predict the component lifetime. This technique, however,
may be not accurate because wear rates in some systems are not constant
throughout time. The wear rate may decrease after a running-in period
(normally in lubricated systems where sliding wear is dominant) or may
increase after an initial incubation period (if the material coating is
removed due to wear and the wear behavior of the lower layer is
different).
 Wear estimation by equations: the use of theoretical or empirical
equations relating wear rate to the operating characteristics, such as load
and speed, can be useful, but provide only an estimation for initial design
calculations.
 Component tests: the wear is estimated by tests performed on the
component under conditions expected during the component service.
This method is particularly frequent to predict wear in bearings under
varying loads, speeds, temperatures and states of lubrication. Test
conditions used to accelerate wear production must be done with caution
to ensure that the wear mechanism is not changed and other relevant
operating conditions do not vary during that test [81].

Generally, new engine components are evaluated during special durability


tests developed to submit the part to extreme operation scenarios or during vehicle testing,
when the component is tested under conditions similar to real driving conditions. Despite
presenting extremely realistic and reliable results, the two types of tests are extremely

19
costly and time consuming, mainly for the engine parts presenting great compliance
issues and still in an early development stage.
When a core engine component presents unacceptable wear after the
implementation of start-stop system, special and simplified tests to simulate the load
pattern are developed, among them computer simulation and component test benches can
be highlighted. Even though the conditions considered in these simplified methods are
not capable of replacing the durability tests with the real engine, they can reduce the time
for the component development and the associated costs [95]. Thereby, the following
subsections present the recent research on bearing wear developed by using
computational and experimental procedures, which have been implemented for computer
simulation and component test bench.

Computational simulation

Several simulation models to predict friction and wear in automotive bearings


have been recently developed and relevant mechanical factors have to be considered in
order to obtain accurate predictions. Frequently, the computational models are developed
and validated using data from experimental tests. The majority of the numerical
investigation about wear and friction in bearings employs several simplifying
assumptions about the surface mechanics and operating conditions. The complexities
associated to wear and friction of moving surfaces, which usually present non-
deterministic asperity distributions, pose strong limitations on the analysis of several
contact geometries and loading. Sometimes, simple wear mechanism can be numerically
investigated and reliable predictions are possible only for wear of dry joints only [96-
98].
Aghdam et al [99] performed an analytical-experimental study on heavily
loaded lubricated journal bearings, in an attempt to establish correlations among wear
rate, power dissipation and temperature rise during operation. The correlations between
power dissipation and temperature rise were obtained using the finite element method.
The relationships brought by this work can be useful to predict wear, which is influenced
by the boundary and geometric effects. However, this study also demonstrates the need
of further investigations on the models for wear prediction, which are strongly associated
with the wear mechanism, lubrication regime and sliding configuration.
Noteworthy to say that the wear phenomenon represents a strong limiting
factor in the durability of many mechanical systems. Therefore, computational models
need to be capable of simulating not only specific engine operating conditions, but also
predicting wear in conditions of mixed lubrication [100]. Some studies have presented a
simplified analysis of the wear on lubricated contacts, assuming that the engines are
running at constant speed and disregarding combustion forces, external torque and other
actual engine parameters [101, 102]. Moreover, the majority of the computational studies
on engine component wear considers a direct transition from full film lubrication regime
to boundary lubrication regime, disregarding the elastohydrodynamic lubrication regime,
and consequently, not accounting for the asperity elasto-plastic interactions.
Some computational investigations have accounted for the complete
lubrication regime transition to evaluate the influence of the engine vibration on the wear,
using procedures to update the surface profile and wear rates [103]. For the development
of accurate computational procedures, experimental data from test rigs are extremely
important. Test rigs can provide basic data for simulation and measurements of wear
rates and other tribological characteristics for validation of computational procedures
[104]. The influence of surface adaptation caused by running-in wear on bearing friction

20
is considered an important aspect of the computational models [105]. The wear process
not only changes surface roughness, but also can modify the surface geometry, what
consequently changes the contact pressure between bearing and lubricant and may cause
misalignments [106, 107].
Moreover, wear profiles are important to friction simulation, since friction is
dependent on the nature of asperity contact between contacting bodies. Hence, the use of
actual surface pattern and its mechanical elastic deformation under load can lead to better
friction estimations. To simulate an accurate bearing wear profile during engine
operation, the wear rates of the bearing surface must be measured after test runs [104].
The measured surface height changes caused by wear can be averaged from repeated test
runs and used as database for several friction prediction simulation programs [108, 109]
Predictions rendered by a computational bearing model have been compared
to measurements performed on the slider bearing test rig LP06 of the MIBA-Bearing
Group. In this test rig, a static load is applied on a rotating bearing and its parameters are
employed to build a computational model to predict wear. The bearing simulation has
been performed using the test conditions of load, temperature, shaft speed and oil
properties provided by the test rig. This type of analytical-experimental wear analysis
can provide solid base to further simulation programs to predict the bearing friction [104,
110].
Reichart et al. [100] presented a computational study about the sliding wear
of two rough surface profiles, under non-lubricated and mixed-lubricated conditions,
including the measurements of the bearing and journal surface topology. In this study, the
analysis contemplated the microscopic description of the rough surfaces using the finite
elements method. The analysis employed a model of mixed-lubrication proposed by
Lorentz [111]. Based on the available literature [112], the wear estimated for non-
lubricated regimes presented good agreement with published results. However, the wear
estimated for lubricated systems did not agree well with results available, indicating the
need for further investigations.
Bergmann et al. [113] obtained experimental wear coefficients from journal
bearing test bench under start-stop conditions and implemented a computational
procedure for wear analysis. The computational simulation based on experimental
coefficients, using the finite element method, permitted the dynamic numerical evaluation
of wear processes. The predictions rendered by the computational procedure depict trends
similar to that shown by the results observed on the experiments. However, significant
deviation in the wear values was also noted. This deviation was attributed to the
conservative contact model chosen, what led to underestimates of the asperity contact and
wear. The study indicated the need of a more complete contact model, which could allow
the variation of the contact intensity between asperities, in order to enlarge the capability
of wear prediction through simulations.

Component test bench

Many tests have been used to study sliding wear either to examine the wear
mechanism or to simulate practical applications and provide useful data regarding friction
coefficients and wear rates. For both cases, it is very important to reproduce and measure
variables which can influence wear, since slight changes in conditions can lead to radical
changes in the dominant wear mechanism. Wear under sliding conditions depends on the
contacting distance, on the sliding velocity, and on the duration of the test. These three
variables affect the rate of frictional energy dissipation and temperature, and contact
pressure transitions can modify the lubricant regime. The testing temperature is also very

21
important to be controlled, because it influences the mechanical properties of the
materials and can affect the surface chemical processes. The atmospheric composition
should also be ideally controlled as it can contain reactive components such as water
vapor and oxygen, which can influence on wear rates.
Wear can be measured intermittently, by removing the specimen at certain
time intervals and weighing or measuring it, by controlling wear debris on lubrication oil,
or continuously, by measuring its position with a transducer and deducing the wear from
the dimensional changes. Several geometrical arrangements can be employed in wear
tests and the most common are asymmetric, in which the two sliding bodies will
experience different wear rates [36]. One of the most popular and simple test of wear
resistance is the "Pin-on-Disc" test, a non-standardized test that consists basically of a
specimen pressed against a rotating disc with fixed rotation and a pre-defined load. The
test can be run dry or lubricated contacts and, after a determined number of cycles, the
worn volume of the testing block is determined. Also, very common are the “Ring-in-
Disc” and “Block-on-Ring” tests, which intend basically to observe the behavior of the
contacting bodies during sliding contact. In the “Ring-on-Disc” approach, a polished ring
is mounted in the bottom part of the test cell. Then, a disc fixed to a rotating shaft is
loaded against the stationary ring. Similarly, in “Block-on-Ring” tests, a testing block is
pressed axially with a determined load on a rotating disc [82]. In all tests, wear
estimations can be conducted by posterior ring measurements or with a wear sensor [114].
The wear resistance depends, in particular, on the hardness of the material.
Therefore, trimetallic galvanized bearings cannot be compared to bimetallic bearings
because their overlays are very soft [82]. To better understand the wear mechanisms in
this test, torque, friction coefficient lubricant and specimen temperatures can be measured
[115].
The recent trend of decreasing the of film thickness and the use of lower
viscosity oils to decrease friction losses have led to an increase in frequency of solid
contact during the engine dynamic operation [116-118]. That situation concerns mainly
piston ring, liners and bearings. To evaluate journal bearing wear under these
triboconditions, a similar test was conducted by Bovington [119], in which a static load
was applied to a journal bearing and the friction torque were measured for a wide speed
range. In this study, different lubrication regimes were identified and tests were conducted
with several engine oils, with different formulations.
An interesting sliding wear study performed with a leaded brass specimen
sliding against a smooth hard and cetane-lubricated stellite ring was conducted. No
significant hydrodynamic film was formed and the wear behavior of the specimen
followed the Archard equation. However, at low loads, the wear rate presented a lower
wear coefficient than that at high loads, what can be associated with a drop-in contact
resistance, which becomes constant in a much lower value after the transition load is
reached. The surface roughness presented drastic change in these two regimes and is
much higher after transition. Analyzing the wear debris for both conditions, it was
possible to state that the mild wear is associated to the wear of the oxide layer covering
the specimen, and the severe wear is reached when the oxide layer is worn-out and
metallic particles are then removed. Normally, severe wear rates are so high that are
completely unacceptable in engineering applications [36].
Experimental studies have shown that the oxide layer is important for the
sliding wear rates. Therefore, an experimental research demonstrated that low sliding
speed and higher temperatures can increase the oxide layer formation and increase the
mild wear condition. On the contrary, at very high sliding speeds, the rates of oxide layer
development can also increase as the interface temperatures can also increase. The sliding

22
wear mechanism is very complex to estimate, mainly when it is unlubricated, because of
the dependence of the mechanical stresses, temperature and oxidation, factors that are
interrelated [120].
A special test rig for automotive bearings was developed with a connecting
rod connected to a hydraulic actuator and a shaft driven by an electric motor. This test
bench uses an oil conditioning system capable of regulating the oil temperature within
the range of 50-170°C. The maximal dynamic load applied in the bearings were of 500kN,
in a frequency range of 0–100Hz. The measurements collected in this test rig were the oil
temperature, the frictional torque and the electrical signal of the contact between the drive
shaft and the bearing, which can indicate the presence and intensity of the asperity
contact. The test runs lasted up to15 h and all measurements were acquired and averaged
over the whole run time for further analysis [104].
To evaluate the influence of low viscosity lubricants on the automotive oil-
lubricated journal bearings, the Idemitsu Kosan test rig was developed to study the
bearing reliability and to test a variety of commercial oils. In this test apparatus, oil was
supplied to the bearings and the load applied on the bearings fluctuated in the same way
of actual engines. Two tests were conducted, both with the shaft speeds of 500, 1000,
2000, and 3000 rpm and oil temperatures of 60, 80 and 100°C. In the first test, the static
load increased from 1 to 10 kN, by increments of 1 kN, and in the second, the 8 kN static
load varied ±3kN. The friction forces generated were measured. The apparatus was used
to evaluate the performance of the bearings with a variety of engine oils. In this study it
was shown that not only contact between the shaft and the journal bearing occurred more
readily when using lower viscosity oil, but that this contact was more likely to occur with
highly refined base oils than with low-refined base oils for the same low viscosity [121].
The manufacturer Federal Mogul (FM) also developed tribometers similar to
those from MIBA and Idemitsu to analyze bearing wear under different load and
lubrication conditions. The goal was to develop bearings adapted to the recent severe
engine operations, such as those observed in engines with start-stop systems. The seizure
test developed by FM, called “stress test”, has a first test phase, which is performed under
full lubrication (5h) with load and speed increasing stepwise to allow some initial surface
adaptation, and a second phase (30h), in which the oil supply is almost completely
interrupted. This test can evaluate the potential seizure and wear characteristics, such as
shaft roughness, bearing cover layer thickness and material variation. In order to simulate
the start-stop condition, the company developed the “Nautilus test”, in which a hydraulic
cylinder simulates the load application of 60MPa at shaft rotation of 1200rpm, repeating
the procedure 5 times during 1 min. Then at speed of 4200rpm, the procedure is
performed 5 times for 5 min and at a load of 127 MPa the tests are performed 6 times for
5min. The experimental results indicate that any bearing material that endures 30 minutes
of test presents excellent wear and seizure resistance [92]. However, although the tests
are performed at severe wear conditions for bearings, they do not reproduce the actual
engine conditions, that the bearing is required to withstand.
Gebretsadik et al. [84] used a “Block-on-Ring” test rig to evaluate the
performance of several bimetal and multi-layer Pb-free bearing materials with different
lining and overlay compositions. For friction and wear studies, test samples were cut of
the bearing shells and the ring was made of high-grade steel with known roughness. The
lower ring was driven by a shaft, on which the contacting surface specimen was mounted.
The lower part of the test ring was immersed in oil bath. In this test, significant results
were obtained indicating that bearing overlays, based on PAI and containing graphite and
MoS2, exhibit better friction and wear properties than the Pb-and Al-Sn based materials,
which has already been observed in other studies [90].

23
Zhang et al. [89] performed tests on material with a multilayer tin-copper
overlay and proposed a procedure to overcome some weaknesses of traditional monolayer
overlays. The deposit of the overlays was performed by electroplating on a range of
different test pieces, such as bronze bearings and bronze test pieces. Then, fatigue, wear
and seizure were monitored during the operation of the tribometer, by which a dynamic
load was applied to the test bearing by rotating an eccentric journal. Proper lubrication
the bearing was also provided. As a result, the tin-cooper multilayer overlay presented
increase seizure and wear resistance.
Summer et al. [122] studied two laboratory test benches to evaluate the
surface damage in journal bearings. Friction and wear performance, seizure events and
start-stop conditions were studied for automotive bearings. The first test rig was the “ring-
on-disc” model, in which the disc specimens were manufactured with the same material
used in the bearings and lubrication was provided. It should be however emphasized that
differences in lubrication type, area of contact, friction, and temperature characteristics
can affect the final results, when compared to bearings in service [114]. In the second test
rig, two fully lubricated bearings were tested in contact with a rotating shaft. In this test
bench, several sensors were employed, providing reliable measures of the bearing friction
characteristics. The results showed that the two experimental apparatuses can be used to
accurately describe friction, seizure and wear phenomena in bearing systems, rendering
important guidance for the lubricant and bearing material selection.

FINAL REMARKS

The present work is an attempt to discuss the tribological aspects associated


with the application of the start-stop systems in internal combustion engines used in
commercial vehicles. The search for more durable and reliable bearings that are
economically and technically feasible for use in the supporting systems of the engine
moving components is a continuous technological challenge for the automotive industry.
Stringent operating conditions for the engine bearings in vehicles with start-stop systems
require a much better understanding about the relevant bearing operating and geometric
characteristics, which can be used to design more efficient supporting systems to reduce
friction and wear.
This review describes some recent technical contributions to enlarge the
knowledge about friction, lubrication, and wear of automotive bearings that can be
employed in start-stop systems. The new operating conditions of these bearings require a
review of the engine tribological parameters, mainly regarding the quality of the
lubricating oil and the materials and coatings applied on its components, on which contact
is most critical. More specifically, this work summarizes some recent advancements
about the tribological aspects of the engine bearings, the parameters for their material
selection and the development of materials to meet the increasing severity in modern
engine operations. Furthermore, this review brings a brief description of the analytical-
experimental procedures widely used to evaluate the friction and wear in bearings, which
can bring some insights and guidance for engine manufacturers in the understanding of
the bearing characteristics for a safer design.

REFERENCES

[1] A. Faiz, C. S. Weaver, and M. P. Walsh, Air pollution from motor vehicles:
standards and technologies for controlling emissions. The World Bank, 1996.

24
[2] M. a. M. Williams, R, "A Technical Summary of Euro 6/VI Vehicle Emission
Standards," ed: ICCT International council on clean transportation, 2016.
[3] V. R. Roso, N. D. S. A. Santos, C. E. C. Alvarez, F. A. Rodrigues Filho, F. J. P.
Pujatti, and R. M. Valle, "Effects of mixture enleanment in combustion and
emission parameters using a flex-fuel engine with ethanol and gasoline," Applied
Thermal Engineering, 2019.
[4] N. D. S. A. Santos, C. E. C. Alvarez, V. R. Roso, J. G. C. Baeta, and R. M. Valle,
"Combustion analysis of a SI engine with stratified and homogeneous pre-
chamber ignition system using ethanol and hydrogen," Applied Thermal
Engineering, p. 113985, 2019.
[5] J. S. Martinez, D. Hissel, M.-C. Pera, and M. Amiet, "Practical control structure
and energy management of a testbed hybrid electric vehicle," IEEE Transactions
on Vehicular Technology, vol. 60, no. 9, pp. 4139-4152, 2011.
[6] N. Fonseca, J. Casanova, and M. Valdes, "Influence of the stop/start system on
CO2 emissions of a diesel vehicle in urban traffic," Transportation Research Part
D: Transport and Environment, vol. 16, no. 2, pp. 194-200, 2011.
[7] J. W. George and R. Brock, "Polymeric engine bearings for hybrid and start stop
applications," SAE Technical Paper, 0148-7191, 2012.
[8] S. Nebelsiek, "Auslegung einer Methodik zur Verschleißermittlung an
Triebwerkslagern," ed: Hochschule Esslingen - University of Applied Sciences,
2017.
[9] N. D. S. A. Santos, "STUDY OF A BEARING TEST BENCH CAPABLE TO
SIMULATE WEAR IN DIESEL START-STOP ENGINES," Master, UFMG,
2019.
[10] P. R. Windover, R. J. Owens, T. M. Levinson, and M. D. Laughlin, "Stop and
restart effects on modern vehicle starting system components: Longevity and
economic factors," 2015.
[11] J. Rueger, "Clean Diesel-Real Life Fuel Economy and Environmental
Performance," in SAE Government and Industry Meeting, Washington, DC, 2008.
[12] P. Floraday, "“Want stop-start technology on your next vehicle? Lobby the
EPA”," ed: Automobile Magazine, 2009.
[13] F. Canova and L. Sala, "Back to square one: Identification issues in DSGE
models," Journal of Monetary Economics, vol. 56, no. 4, pp. 431-449, 2009.
[14] A. Ozdemir and A. Mugan, "Stop/Start System Integration to Diesel Engine and
System Modelling & Validation," IFAC Proceedings Volumes, vol. 46, no. 25,
pp. 95-100, 2013.
[15] G. Rizzoni, L. Guzzella, and B. M. Baumann, "Unified modeling of hybrid
electric vehicle drivetrains," IEEE/ASME transactions on mechatronics, vol. 4,
no. 3, pp. 246-257, 1999.
[16] B. M. Baumann, G. Washington, B. C. Glenn, and G. Rizzoni, "Mechatronic
design and control of hybrid electric vehicles," IEEE/ASME Transactions On
Mechatronics, vol. 5, no. 1, pp. 58-72, 2000.
[17] K. Van Berkel, T. Hofman, B. Vroemen, and M. Steinbuch, "Optimal control of
a mechanical hybrid powertrain," IEEE Transactions on Vehicular Technology,
vol. 61, no. 2, pp. 485-497, 2012.
[18] F. R. Salmasi, "Control strategies for hybrid electric vehicles: Evolution,
classification, comparison, and future trends," IEEE Transactions on vehicular
technology, vol. 56, no. 5, pp. 2393-2404, 2007.
[19] C. C. Chan, "The state of the art of electric, hybrid, and fuel cell vehicles,"
Proceedings of the IEEE, vol. 95, no. 4, pp. 704-718, 2007.

25
[20] L. Fang and S. Qin, "Optimal control of parallel hybrid electric vehicles based on
theory of switched system," Asian Journal of Control, vol. 8, no. 3, pp. 274-280,
2006.
[21] Y. Gurkaynak, A. Khaligh, and A. Emadi, "State of the art power management
algorithms for hybrid electric vehicles," in Vehicle Power and Propulsion
Conference, 2009. VPPC'09. IEEE, 2009: IEEE, pp. 388-394.
[22] W. Shabbir and S. A. Evangelou, "Threshold-changing control strategy for series
hybrid electric vehicles," Applied energy, vol. 235, pp. 761-775, 2019.
[23] K. Kataoka and K. Tsuji, "Crankshaft positioning utilizing compression force and
fast starting with combustion assist for indirect injection engine," SAE Technical
Paper, 0148-7191, 2005.
[24] M. L. Kuang, "An investigation of engine start-stop NVH in a power split
powertrain hybrid electric vehicle," SAE Technical Paper, 0148-7191, 2006.
[25] H.-Y. Hwang, "Minimizing seat track vibration that is caused by the automatic
start/stop of an engine in a power-split hybrid electric vehicle," Journal of
Vibration and Acoustics, vol. 135, no. 6, p. 061007, 2013.
[26] G. Ryk, Y. Kligerman, and I. Etsion, "Experimental investigation of laser surface
texturing for reciprocating automotive components," Tribology Transactions, vol.
45, no. 4, pp. 444-449, 2002.
[27] Y. Kagohara, S. Takayanagi, S. Haneda, M. Fujita, and Y. Iwai, "Tribological
property of plain bearing with low frictional layer," Tribology International, vol.
42, no. 11-12, pp. 1800-1806, 2009.
[28] R. Taylor, "Transient effects in engines operating at steady speeds & loads," in
Tribology Series, vol. 43: Elsevier, 2003, pp. 123-131.
[29] B. J. Hamrock, S. R. Schmid, and B. O. Jacobson, Fundamentals of fluid film
lubrication. CRC press, 2004.
[30] D. Dowson, G. Higginson, and A. Whitaker, "Elasto-hydrodynamic lubrication: a
survey of isothermal solutions," Journal of Mechanical Engineering Science, vol.
4, no. 2, pp. 121-126, 1962.
[31] C. F. Taylor, The Internal-combustion Engine in Theory and Practice:
Combustion, fuels, materials, design. MIT press, 1985.
[32] J. Walker, T. Kamps, and R. Wood, "The influence of start–stop transient velocity
on the friction and wear behaviour of a hyper-eutectic Al–Si automotive alloy,"
Wear, vol. 306, no. 1-2, pp. 209-218, 2013.
[33] A. King, "Non friction," ed: Chemistry & Industry News, 2012.
[34] A. R. Rogowski, "Elements of Internal Combustion Engines," ed: McGraw-Hill,
1953.
[35] F. P. Bowden and D. Tabor, The friction and lubrication of solids. Oxford
university press, 2001.
[36] I. Hutchings and P. Shipway, Tribology: friction and wear of engineering
materials. Butterworth-Heinemann, 2017.
[37] H. Murthy and K. Vadivuchezhian, "Estimation of friction distribution in partial-
slip contacts from reciprocating full-sliding tests," Tribology International, vol.
108, pp. 164-173, 2017.
[38] P. Jabłoński et al., "Tribological properties evaluation of Ni-P coating
manufactured with electroless plating on aluminum alloy substrate," in MATEC
Web of Conferences, 2017, vol. 112: EDP Sciences, p. 04014.
[39] B. Bhushan, Introduction to tribology. John Wiley & Sons, 2013.
[40] H. Zumbühl, "Motoren - Ein Buch über Wärmekraftmaschinen und ihre
Brennstoffe," ed: SDV-Fachbücher, 1946.

26
[41] H. Czichos, "Failure criteria in thin film lubrication: investigation of the different
stages of film failure," Wear, vol. 36, no. 1, pp. 13-17, 1976.
[42] G. Salomon, "Failure criteria in thin film lubrication—the IRG program," Wear,
vol. 36, no. 1, pp. 1-6, 1976.
[43] T. Fowle, "A note on the four-ball machine," Tribology News, vol. 25, pp. 10-12,
1974.
[44] B. Manual, "Bearing Group Miba," Miba Gleitlager AG, 2000.
[45] R. Stribeck and M. Schröter, Die wesentlichen Eigenschaften der Gleit-und
Rollenlager: Untersuchung einer Tandem-Verbundmaschine von 1000 PS.
Springer, 1903.
[46] J. Affenzeller and H. Gläser, Lagerung und Schmierung von
Verbrennungsmotoren. Springer-Verlag, 2013.
[47] W. Steinhilper, B. Sauer, and J. Feldhusen, Konstruktionselemente des
Maschinenbaus 1: Grundlagen der Berechnung und Gestaltung von
Maschinenelementen. Springer-Verlag, 2008.
[48] J. Williams, Engineering tribology. Cambridge University Press, 2005.
[49] S. M. Chun and M. M. Khonsari, "Wear simulation for the journal bearings
operating under aligned shaft and steady load during start-up and coast-down
conditions," Tribology International, vol. 97, pp. 440-466, 2016.
[50] L. R. Rudnick, Lubricant additives: chemistry and applications. CRC press, 2017.
[51] H. Rahnejat, Tribology and dynamics of engine and powertrain: fundamentals,
applications and future trends. Elsevier, 2010.
[52] B. Johnson, H. Wu, M. Desanker, D. Pickens, Y.-W. Chung, and Q. J. Wang,
"Direct formation of lubricious and wear-protective carbon films from
phosphorus-and sulfur-free oil-soluble additives," Tribology Letters, vol. 66, no.
1, p. 2, 2018.
[53] R. Liu, X. Meng, and Y. Cui, "Influence of numerous start-ups and stops on
tribological performance evolution of engine main bearings," International
Journal of Engine Research, p. 1468087418810094, 2018.
[54] B. Tormos, L. Ramírez, J. Johansson, M. Björling, and R. Larsson, "Fuel
consumption and friction benefits of low viscosity engine oils for heavy duty
applications," Tribology International, vol. 110, pp. 23-34, 2017.
[55] F. Summer, F. Grün, M. Offenbecher, and S. Taylor, "Challenges of friction
reduction of engine plain bearings–Tackling the problem with novel bearing
materials," Tribology International, vol. 131, pp. 238-250, 2019.
[56] M. J. Covitch, M. Brown, C. May, T. Selby, I. Goldmints, and D. George,
"Extending SAE J300 to viscosity grades below SAE 20," SAE International
Journal of Fuels and Lubricants, vol. 3, no. 2, pp. 1030-1040, 2010.
[57] A. Gangopadhyay, "A review of automotive engine friction reduction
opportunities through technologies related to tribology," Transactions of the
Indian Institute of Metals, vol. 70, no. 2, pp. 527-535, 2017.
[58] ACEA, "ACEA European Oil Sequences," ed, 2016.
[59] V. W. Wong and S. C. Tung, "Overview of automotive engine friction and
reduction trends–Effects of surface, material, and lubricant-additive
technologies," Friction, vol. 4, no. 1, pp. 1-28, 2016.
[60] R. Cipollone, D. Di Battista, and M. Mauriello, "Effects of oil warm up
acceleration on the fuel consumption of reciprocating internal combustion
engines," Energy Procedia, vol. 82, pp. 1-8, 2015.

27
[61] B. Briscoe, D. Evans, and D. Tabor, "The influence of contact pressure and
saponification on the sliding behavior of stearic acid monolayers," Journal of
Colloid and Interface Science, vol. 61, no. 1, pp. 9-13, 1977.
[62] B. Briscoe and D. Tabor, "Rheology of thin organic films," ASLE transactions,
vol. 17, no. 3, pp. 158-165, 1974.
[63] H. Spikes, "Friction modifier additives," Tribology Letters, vol. 60, no. 1, p. 5,
2015.
[64] K. Topolovec-Miklozic, T. R. Forbus, and H. A. Spikes, "Performance of friction
modifiers on ZDDP-generated surfaces," in STLE/ASME 2006 International Joint
Tribology Conference, 2006: American Society of Mechanical Engineers, pp.
159-167.
[65] M. Yajun, Z. Wancheng, L. Shenghua, J. Yuansheng, W. Yucong, and T. Simon,
"Tribological performance of three advanced piston rings in the presence of
MoDTC-modified GF-3 oils," Tribology Letters, vol. 18, no. 1, pp. 75-83, 2005.
[66] R. Elo and S. Jacobson, "Formation and breakdown of oil residue tribofilms
protecting the valves of diesel engines," Wear, vol. 330, pp. 193-198, 2015.
[67] A. Kovalchenko, O. O. Ajayi, A. Erdemir, and G. R. Fenske, "Friction and wear
performance of low-friction carbon coatings under oil lubrication," SAE
Transactions, pp. 902-910, 2002.
[68] A. Kotia, S. Borkakoti, and S. K. Ghosh, "Wear and performance analysis of a 4-
stroke diesel engine employing nanolubricants," Particuology, vol. 37, pp. 54-63,
2018.
[69] P. Cann and H. Spikes, "The behavior of polymer solutions in concentrated
contacts: immobile surface layer formation," Tribology transactions, vol. 37, no.
3, pp. 580-586, 1994.
[70] H. Mitsui and H. Spikes, "Predicting EHD film thickness of lubricant polymer
solutions," Tribology transactions, vol. 41, no. 1, pp. 1-10, 1998.
[71] V. S. Jatti and T. Singh, "Copper oxide nano-particles as friction-reduction and
anti-wear additives in lubricating oil," Journal of Mechanical Science and
Technology, vol. 29, no. 2, pp. 793-798, 2015.
[72] M. H. Esfe, A. A. A. Arani, S. Esfandeh, and M. Afrand, "Proposing new hybrid
nano-engine oil for lubrication of internal combustion engines: Preventing cold
start engine damages and saving energy," Energy, vol. 170, pp. 228-238, 2019.
[73] Y. Olomolehin, R. Kapadia, and H. Spikes, "Antagonistic interaction of antiwear
additives and carbon black," Tribology Letters, vol. 37, no. 1, p. 49, 2010.
[74] K. Yamamoto, A. Kotaka, and K. Umehara, "Additives for improving the fuel
economy of diesel engine systems," Tribology Online, vol. 5, no. 4, pp. 195-198,
2010.
[75] W. C. Barnhill et al., "Tribological bench and engine dynamometer tests of a low
viscosity SAE 0W-16 engine oil using a combination of ionic liquid and ZDDP
as anti-wear additives," Frontiers in Mechanical Engineering, vol. 1, p. 12, 2015.
[76] O. R. G. o. W. o. E. Materials, "Friction, Wear and Lubrication - Tribology of
Terms and Definition," ed: OECD, 1969.
[77] D. I. f. Normung, "DIN 50320: Verscheiss - Begriff, Analyse von
Verschleissvorgängen, Gliederung des Verschleissgebietes," ed: Beuth Verlag,
1979.
[78] F. Barwell, "Theories of wear and their significance for engineering practice," in
Treatise on Materials Science & Technology, vol. 13: Elsevier, 1979, pp. 1-83.

28
[79] K. H. ZUM-GAHR, "Microstructure and wear of materials. Institute of Materials
Technology," University of Siegen, Siegen, Federal Republic of Germany, vol.
10, 1987.
[80] J. Archard, "Contact and rubbing of flat surfaces," Journal of applied physics, vol.
24, no. 8, pp. 981-988, 1953.
[81] G. Stachowiak, A. Batchelor, and T. Stolarski, "Engineering tribology: Elsevier,
1993, ISBN 0-444-89235-4, 960 pp,£ 156.00," ed: Elsevier, 1994.
[82] Mahle, "Gleitlager," Zylinderkomponenten: Eigenschaften, Anwendungen,
Werkstoffe, pp. 49-69, 2009.
[83] L. Tambellini, L. Oliveira, and M. Tsubouchi, "Bearing Fundamentals," ed:
MAHLE Metal Leve S.A., GEA-B, 2007.
[84] D. W. Gebretsadik, J. Hardell, and B. Prakash, "Friction and wear characteristics
of different Pb-free bearing materials in mixed and boundary lubrication
regimes," Wear, vol. 340, pp. 63-72, 2015.
[85] M. GmbH, "Gleitlager," Zylinderkomponenten: Eigenschaften, Anwendungen,
Werkstoffe, pp. 49-69, 2009.
[86] F. Grün, I. Gódor, W. Gärtner, and W. Eichlseder, "Tribological performance of
thin overlays for journal bearings," Tribology International, vol. 44, no. 11, pp.
1271-1280, 2011.
[87] G. C. Pratt, "Materials for plain bearings," International Metallurgical Reviews,
vol. 18, no. 2, pp. 62-88, 1973.
[88] E. P. Becker, "Trends in tribological materials and engine technology," Tribology
International, vol. 37, no. 7, pp. 569-575, 2004.
[89] Y. Zhang, I. Tudela, M. Pal, and I. Kerr, "High strength tin-based overlay for
medium and high speed diesel engine bearing tribological applications,"
Tribology International, vol. 93, pp. 687-695, 2016.
[90] A. Adam, M. Prefot, and M. Wilhelm, "Crankshaft bearings for engines with start-
stop systems," MTZ worldwide, vol. 71, no. 12, pp. 22-25, 2010.
[91] F. M. Powertrain, "Application Engineering FM Bearings," ed: Federal Mogul
Powertrain, 2010.
[92] J. Häring, "IROX II - Neue Hochleistungspollymerbeschichtung für Gleitlager,"
ed: VDI Wissenforum, 2018.
[93] D. W. Gebretsadik, J. Hardell, and B. Prakash, "Seizure behaviour of some
selected Pb-free engine bearing materials under lubricated condition," Tribology
International, vol. 111, pp. 265-275, 2017.
[94] ACEA, "ACEA 2016 European Oil for Service-Fill Oils," ed, 2016.
[95] S. Shanta, G. Molina, and V. Soloiu, "Tribological effects of mineral-oil lubricant
contamination with biofuels: a pin-on-disk tribometry and wear study," Advances
in Tribology, vol. 2011, 2011.
[96] A. A. Olyaei and M. R. Ghazavi, "Stabilizing slider-crank mechanism with
clearance joints," Mechanism and Machine Theory, vol. 53, pp. 17-29, 2012.
[97] L. X. Xu and Y. C. Han, "A method for contact analysis of revolute joints with
noncircular clearance in a planar multibody system," Proceedings of the
Institution of Mechanical Engineers, Part K: Journal of Multi-body Dynamics,
vol. 230, no. 4, pp. 589-605, 2016.
[98] N. Nikolic, T. Torovic, and Z. Antonic, "A procedure for constructing a
theoretical wear diagram of IC engine crankshaft main bearings," Mechanism and
Machine Theory, vol. 58, pp. 120-136, 2012.

29
[99] A. Aghdam and M. Khonsari, "Prediction of wear in grease-lubricated oscillatory
journal bearings via energy-based approach," Wear, vol. 318, no. 1-2, pp. 188-
201, 2014.
[100] S. Reichert, B. Lorentz, S. Heldmaier, and A. Albers, "Wear simulation in non-
lubricated and mixed lubricated contacts taking into account the microscale
roughness," Tribology International, vol. 100, pp. 272-279, 2016.
[101] G. B. Daniel and K. L. Cavalca, "Analysis of the dynamics of a slider–crank
mechanism with hydrodynamic lubrication in the connecting rod–slider joint
clearance," Mechanism and Machine Theory, vol. 46, no. 10, pp. 1434-1452,
2011.
[102] L. J. Wu, A. P. Li, and X. M. Liu, "Modeling and simulation of a slider-crank
mechanism with clearance and lubricant joint," in Advanced Materials Research,
2012, vol. 510: Trans Tech Publ, pp. 458-466.
[103] M. Haneef, R. Randall, W. Smith, and Z. Peng, "Vibration and wear prediction
analysis of IC engine bearings by numerical simulation," Wear, vol. 384, pp. 15-
27, 2017.
[104] C. Priestner, H. Allmaier, H.-H. Priebsch, and C. Forstner, "Refined simulation of
friction power loss in crank shaft slider bearings considering wear in the mixed
lubrication regime," Tribology International, vol. 46, no. 1, pp. 200-207, 2012.
[105] D. Bartel, L. Bobach, T. Illner, and L. Deters, "Simulating transient wear
characteristics of journal bearings subjected to mixed friction," Proceedings of the
Institution of Mechanical Engineers, Part J: Journal of Engineering Tribology,
vol. 226, no. 12, pp. 1095-1108, 2012.
[106] M. Fillon and J. Bouyer, "Thermohydrodynamic analysis of a worn plain journal
bearing," Tribology international, vol. 37, no. 2, pp. 129-136, 2004.
[107] J. Sun et al., "Effect of surface roughness, viscosity-pressure relationship and
elastic deformation on lubrication performance of misaligned journal bearings,"
Industrial Lubrication and Tribology, vol. 66, no. 3, pp. 337-345, 2014.
[108] H. Allmaier, C. Priestner, C. Six, H. Priebsch, C. Forstner, and F. Novotny-
Farkas, "Predicting friction reliably and accurately in journal bearings—A
systematic validation of simulation results with experimental measurements,"
Tribology International, vol. 44, no. 10, pp. 1151-1160, 2011.
[109] H. Allmaier, C. Priestner, F. Reich, H. Priebsch, C. Forstner, and F. Novotny-
Farkas, "Predicting friction reliably and accurately in journal bearings–The
importance of extensive oil-models," Tribology international, vol. 48, pp. 93-101,
2012.
[110] H. Allmaier, C. Priestner, F. M. Reich, H.-H. Priebsch, and F. Novotny-Farkas,
"Predicting friction reliably and accurately in journal bearings—extending the
EHD simulation model to TEHD," Tribology international, vol. 58, pp. 20-28,
2013.
[111] B. Lorentz and A. Albers, "A numerical model for mixed lubrication taking into
account surface topography, tangential adhesion effects and plastic
deformations," Tribology International, vol. 59, pp. 259-266, 2013.
[112] H. Czichos and K.-H. Habig, Tribologie-Handbuch: Tribometrie,
Tribomaterialien, Tribotechnik. Springer-Verlag, 2010.
[113] P. Bergmann, F. Grün, F. Summer, and I. Gódor, "Evaluation of Wear Phenomena
of Journal Bearings by Close to Component Testing and Application of a
Numerical Wear Assessment," Lubricants, vol. 6, no. 3, p. 65, 2018.
[114] K. S. Pondicherry, F. Grün, I. Gódor, R. Bertram, and M. Offenbecher,
"Applicability of ring‐on‐disc and pin‐on‐plate test methods for Cu–steel and Al–

30
steel systems for large area conformal contacts," Lubrication Science, vol. 25, no.
3, pp. 231-247, 2013.
[115] T. Clausthal, "Pin-on-disc Test," ed: TU Clausthal, 2018.
[116] C. Taylor, "Automobile engine tribology—design considerations for efficiency
and durability," Wear, vol. 221, no. 1, pp. 1-8, 1998.
[117] M. Priest and C. Taylor, "Automobile engine tribology—approaching the
surface," Wear, vol. 241, no. 2, pp. 193-203, 2000.
[118] S. C. Tung and M. L. McMillan, "Automotive tribology overview of current
advances and challenges for the future," Tribology International, vol. 37, no. 7,
pp. 517-536, 2004.
[119] C. Bovington, S. Korcek, and J. Sorab, "The importance of the Stribeck curve in
the minimisation of engine friction," Tribology Series, vol. 36, pp. 205-214, 1999.
[120] W. A. Glaeser, K. C. Ludema, and S. Rhee, "Wear of Materials 1977, Presented
at the International Conference on Wear of Materials, St. Louis, Missouri, April
25-28, 1977," in International Conference on Wear of Materials (1977: St. Louis),
1977: American Society of Mechanical Engineers.
[121] T. Katafuchi and M. Kasai, "Effect of base stocks on the automobile engine
bearing," Tribology International, vol. 42, no. 4, pp. 548-553, 2009.
[122] F. Summer, P. Bergmann, and F. Grün, "Damage Equivalent Test Methodologies
as Design Elements for Journal Bearing Systems," Lubricants, vol. 5, no. 4, p. 47,
2017.

31
Research highlights

- Start-stop controls are used in internal combustion engines to reduce exhaust


emissions and fuel consumption.
- This technology affects load, temperature variations, and lubricant delivery to key
interfaces.
- This review focuses on the impact of start-stop operation on ICE bearing tribology.

Yours sincerely,

Nathália Duarte Souza Alvarenga Santos

32

You might also like