You are on page 1of 16

ARTICLE IN PRESS

Tribology International 40 (2007) 16801695


www.elsevier.com/locate/triboint

Compatibility between tribological surfaces and lubricant


additivesHow friction and wear reduction can be controlled
by surface/lube synergies
A. Nevillea,, A. Morinaa, T. Haquea, M. Voonga,b
a

Corrosion and Surface Engineering Research Group, School of Mechanical Engineering, University of Leeds, Leeds LS 2 9JT, UK
b
Cummins Turbocharger Technologies Ltd, Huddersfield, UK
Received 1 October 2006; received in revised form 17 January 2007; accepted 18 January 2007
Available online 26 March 2007

Abstract
This paper reviews the recent trends in materials technology and lubricant additive technology in engines. The paper will review key
developments in surface engineering, application of nanocomposite materials and other advanced materials (including light alloys). It
will also assess the trends towards greener lubricant additives, driven by environmental legislation and will discuss the implications for
lubrication in the next decade. The key part of the paper will be to review the extent to which materials and lubricants are being used in
partnership in engineering systems to capitalize on the synergies, which can exist between surfaces and lubricants in boundary
lubrication. In a similar manner there are some important antagonisms that need to be identiedan appreciation of such compatibility
issues can assist engineers in selecting a lubrication system.
The paper will review existing literature from outside the work conducted by the authors and will substantiate some of the important
aspects of boundary lubrication surface/lubricant compatibility through reference to some recent work conducted by the authors.
r 2007 Elsevier Ltd. All rights reserved.
Keywords: Tribochemistry; Surface coatings; Lubrication

1. Introduction and background


Environmental protection, resource utilization and
customer satisfaction are the three main drivers for
technology development in the automotive sector. Advances in technology related to engines, fuels, materials
and engine oils, are all required if optimum green, fuel
efcient and durable systems are to be developed [1].
Introducing new materials in IC engines in isolation will
not itself yield the improvements requireda fact demonstrated by the introduction of ceramics in engine parts in
the 1980s which was not accompanied by optimized engine
designs or appropriate lubrication strategies. Focusing on
lubricants and their ability to efciently lubricate a contact,
it is now becoming increasingly apparent that lubricant
formulations currently used are designed and tailored to
Corresponding author. Tel.: +44 113 343 6812; fax: +44 113 242 4611.

E-mail address: A.Neville@leeds.ac.uk (A. Neville).


0301-679X/$ - see front matter r 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.triboint.2007.01.019

work mainly on Fe-based materials, used traditionally in


engines. Obtaining optimum durability (wear) and fuel
economy (friction) of tribological systems, especially where
new materials and surfaces or surface treatments are used
relies on compatibility between the surface and the
lubricant. In this paper, a review of the current situation
in terms of use of new materials and surfaces (treatments
and coatings) in the IC engine is reviewed. Results from an
experimental programme are reported which focus on two
main aspects:




How different tribocouples respond to single and binary


lubricant additives?
How the interactions between additives affects the
tribolm and the tribological responsehighlighting
synergies and antagonisms?

The experimental programme in this paper covers four


different tribocouples and four different lubricants and as

ARTICLE IN PRESS
A. Neville et al. / Tribology International 40 (2007) 16801695

such cannot offer a universal appraisal of surface/lubricant


compatibility. It does however, offer an illustration of the
importance of such reactions and an assessment of the
mechanisms by which surfaces and additives can interact
synergistically or antagonistically. A detailed appraisal of
the openly available literature dealing with (a) lubrication
of non-Fe surfaces and surface/additive interactions and
(b) of additive/additive interactions on Fe-based surfaces
sets the scene for the results presented later in the paper.
The literature dealing with lubricant action at non-Fe
based surfaces 5-years ago was sparse but is growing
rapidly as more researchers appreciate the importance of
optimizing surface/additive performance. In the latest
review by Erdemir [2] the importance of considering the
surface and lubricant as a single system was also highlighted. Nanostructured coatings and laser texturing of
surfaces were both reported to improve the lubrication.
Also, improvements in both friction and wear were seen to
occur when materials with high lubricity (solid lubricants)
are used in systems operating in the boundary lubrication
regime.
It is in the boundary lubrication regime (and mixed
regime) where friction and wear reducing additives
normally exhibit their functionality. The high load bearing
capacity and lubricity of solid lubricants in this regime
provide a back-up function to lubricants [2]. However, it is
not common for the surface and lubricant to purposely be
used in partnership to optimize performance. It is known
that wear and friction performance in the boundary
lubrication regime is controlled mainly from the lubricant
additives which form tribolms in the contacting surfaces
but surface treatments and coatings have an important role
to play in providing an improved performance or they can
in fact eliminate the benets of the additives. Knowing the
details of how surfaces and additives react is paramount in
understanding how to achieve optimal lubrication in the
boundary (and mixed) regimes.

1681

1.1. Materials and surfaces in IC engines


A wide range of tribo-materials, tribo-coatings and
surface treatment techniques for engine tribo-components
have been used in the last few decades, many of which
are being replaced by newly developed materials and
coatings to meet the increasing demands of increased fuel
economy. The most commonly used tribo-materials, wear
resistant coatings and running-in coatings are summarized
in Tables 13, respectively.
1.1.1. Piston ring
The components of the piston assembly are considered
to be the most complicated tribological components to
analyse because they experience large variations in load,
speed, temperature and lubrication regime (which extends
from boundary to hydrodynamic lubrication regime [3]).
Traditionally, low cost and readily available cast
irons namely grey CI, carbidic cast iron and malleable/
nodular CI were used as piston ring materials [47].
However, because of high strength and good fatigue
properties, recently nitrided stainless-steel [7] and tool steel
[4] are widely being used as piston ring materials. The
compression ring experiences extremely high temperatures resulting in severe wear. Arc-ion plating of CrN or
Cr on steel or cast iron compression rings has been
reported to give low friction loss [5,8] However, CrN
plating gives much lower friction compared to Cr plating
resulting in 90% reduction in ring wear and 15% reduction
in bore wear [4]. This is because the electrodeposited Cr
coatings are very densely packed and the low porosity
compared to CrN and give poor oil entrapment and
therefore, it is very difcult to maintain oil lm at the
sliding interface [6]. Piston rings made of cast iron or steel
can also be treated by gas or ion nitriding, Si3N2 dispersed
NiP plating, etc [4]. More recently, plasma sprayed
molybdenum, cermet, ceramics and diamond like carbon

Table 1
Commonly used tribo-materials for piston/cylinder and valve train assembly

Conventional
materials

Piston ring

Piston skirt

Cylinder bore/liner

Camshaft

Shim or follower

Grey cast iron (CI)

Grey cast iron

Monolithic grey CI

Grey CI

Nodular iron, high Crcontaining ferro-based


powdered sintered metal

Carbidic CI
Malleable/nodular
CI

Nodular CI
Chilled hardened
CI
High chromium cast iron
Silicon nitride ceramic

Recently used
materials

Nitrided steel

Tool steel

Copper and nickelbased aluminium


alloy (e.g. Al 336)

Si-containing Al
alloy (AA 390)

Forced steel

Low chromium steel

Cast Al-based metal


matrix composite
Steel liner
Compacted
graphite iron

Titanium

Carburized steel

Ceramic
Composite Material

Powder sintered alloy


(FeCrMoC)

ARTICLE IN PRESS
A. Neville et al. / Tribology International 40 (2007) 16801695

1682

Table 2
Commonly used wear resistant tribo-coatings/surface treatments for piston/cylinder and valve train assemblies
Piston ring

Piston skirt

Cylinder bore

Camshaft

Shim or Follower

Chromium plating

Cr plating on CI

Hard chromium coating on CI

Chromium plated steel

Plasma or ame
sprayed molybdenum
on CI
DLC coating on steel
piston ring

Ni-ceramic (SiCNiP)
coating

Chromium plated steel liner

Induction/ame of
medium carbon steel
Carburizing of low
carbon steel

PbSn plating

Gas/iron nitriding of
steel

Solid lubricants:
polytetrauoroethylene
graphite-based material
MoS2

Coating on Si-containing Al
alloys:
SiC dispersed nickel plating;
Plasma sprayed coating of hard
ferrous or non ferrous alloy
Coating on hypereutectic Al
alloy:
Laser hardening of cast iron;

Carbo-nitriding of nodular iron

Ion plating of Cr2N on steel

Ion plating of TiN/Ti on tool


alloy steel

Surface heat treatment of


AlMMC
Si3N2 dispersed NiP
plating
Plasma sprayed
molybdenum
Ceramic/cermet coating
DLC coating

DLC coating on steel shim

Table 3
Commonly used running-in coatings and surface treatment techniques for piston/cylinder and valve train assembly
Running-in coating/
surface treatment
techniques

Tribo-components

Phosphating

Phosphates of iron, zinc or


manganese is formed by immersion

Oxidizing

Iron oxide (with/without surfer


compound)
Sulphur, nitrogen and carbon
containing layer formed by
diffusion
Harder than sulnuz layer and
absence of sulphur into the treated
surface
Electroplated layer of cadmium, tin
or copper on cast iron piston ring

Sulnuz

Tufftriding

Electroplating

Piston ring

Cylinder bore

Camshaft

Follower/shim

Manganese phosphate
is formed by immersion
oxide coating treating
in steam

Same as camshaft

Same as piston ring

Same as piston ring

Induction/ame
hardening
Carburizing

(DLC) coatings are also becoming popular as piston ring


coatings [5]. The DLC coating on piston ring signicantly
improves the friction properties, engine reliability and
work life, however, high internal stress of DLC which
limits the thickness of DLC coatings further reduces the
longer working life [9].

Non-brittle surface layer of carbon


bearing epsilon FeN

Hard martensitic
surface layer with
tough core Diffusion of
carbon on the surface
of low carbon content
steel and case
hardening

1.1.2. Piston skirt


Grey cast iron has been widely used as a piston material
but recently, because of high thermal conductivity, cast
iron piston materials are being replaced by different kinds
of light weight and high strength aluminium alloys. The
aluminium alloys experience high thermal expansion,

ARTICLE IN PRESS
A. Neville et al. / Tribology International 40 (2007) 16801695

which requires maintaining high clearance between piston


skirt and cylinder bore. The copper and nickel-based
aluminium alloy (Al 336) provides 13% lower thermal
expansion coefcient than pure Al and thus helps to
maintain low clearance between piston skirt and cylinder
liner [7]. One major limitation of piston skirts made of
aluminium alloys sliding against cast iron cylinder liners is
high friction. Polytetrauoroethylene (PTFE), graphitebased materials, MoS2, etc., are used as coating materials
to reduce friction losses [4]. The nickelceramic composite
coatings or molybdenum coatings on aluminium alloys
have also been recommended for piston skirts because of
their improved thermal as well as friction behaviour [7].
1.1.3. Cylinder bore
Monolithic grey cast iron and cast aluminiumsilicon
alloys are used as cylinder bore materials [4,7,10]. However, the hard Si particles in the Al matrix can act as
abrasive particles and increase wear. Therefore, cast iron
cylinder liners are often used on aluminium alloy cylinder
boresstill retaining a 30% reduction of weight compared
to cast iron cylinder bore [4]. For further reduction of
weight of the engine blocks, more recently, cast aluminium
alloy is being used without any cast iron cylinder liner
where hypereutectic AlSi alloys, especially AA390 is
commonly used as cylinder bore material and the inner
surface of the cylinder bore is modied by SiC dispersed
NiP plating, plasma sprayed coating by hard ferrous or
non-ferrous alloys, etc. Conversely, scufng resistance of
hypereutectic alloys is very poor because the interfaces
between Al and Si-rich phases acts as the potential site for
fatigue crack initiation and propagation [9] giving rise to
severe wear in low lubrication conditions. Therefore, more
recently, hypereutectic AlSi alloys are being coated with
electrolytic iron coating, Nikasil (nickel and silicon carbide
matrix coating) plating or thermal sprayed coatings. Short
carbon ber or SiC particulates containing Cast Al-based
Metal Matrix Composite (MMC) is another kind of
material that facilitates signicant weight reduction.
Special honing techniques are used to etch the surface so
that hard particulates are exposed to the piston ring thus
reducing friction [10]. More recently, because of the
increasing demand of high cylinder pressure, especially
for diesel engines, the development of Compacted Graphitic Iron (CGI) has been evolved as an emerging material
that gives 75% higher tensile strength and double the
fatigue strength of unalloyed grey CI and aluminium
[11,12].
1.1.4. Valve train
Friction plays an important role in the selection of valve
train materials; efcient selection of cam/follower materials
and coatings can signicantly reduce the friction loss. It is
generally considered that the cam/follower is operated in
the boundary lubrication regime. However, according to
Taylor [13], the elasto-hydrodynamic lubrication regime
also signicantly affects the friction characteristics at the

1683

cam/follower interface. The failure mechanisms at the cam/


follower interface, which are strongly inuenced by the
regime of lubrication, are pitting, polishing and scufng.
Therefore, the selection of material combination, surface
property and lubrication packages for cam/follower conguration in relation to boundary lubrication and elastohydrodynamic lubrication is crucial.
Traditionally, different types of cast iron are used to
make camshafts followed by hardening the sliding surface
of the cam lobes by various hardening techniques
[4,5,7,14]. The most commonly used cast irons are grey
cast iron [7], nodular cast iron [4] and chilled hardened
alloy cast iron [4,14]. Recently there have been moves to
replace these with forced steel, composite materials, etc.
[14]. However, cast iron is still used for the large volume of
production because it is cheap and readily available
whereas steel is used for low volume production because
it is required to be machined to make camshafts resulting in
huge material loss. To improve the running-in performance
as well as to prevent the early failure of the cam, several
surface treatment techniques such as induction hardening
or ame hardening of cast iron, carburizing of low carbon
steel, induction hardening of medium carbon steel, etc., are
performed. In addition, phosphate coatings, oxide coatings, carbon bearing epsilon FeN layers, etc., can be
deposited on the cam surface to improve the wetting and
spreading of the oil and thus gives improved running-in
performance [15].
The cam/follower is mainly operated in the boundary
lubrication regime; due to asperity contact wear rates can
be high. As such, it is necessary to make the cam/follower
components from high wear resistant materials. Because of
their excellent wear resistance properties, ferro-based
powder sintered metal with high chromium, high chromium cast iron or silicon nitride ceramics are conventionally used as shim materials [16]. The use of ceramics as a
shim material has not proven to be economically viable to
date [7]. Recently, steel or light weight forged aluminium
have been used as shim materials [11]. Careful selection of
the surface topology of the shim is very important because
it greatly inuences the running-in property and the
transition between boundary lubrication to elasto-hydrodynamic lubrication and consequently, further increases of
the volume of wear. Ion plated CrN and Cr2N on shims
gives excellent scufng resistance under boundary lubrication but the smooth surface of Cr plating gives poor
lubrication characteristics [15]. Ion-plating of TiN/Ti of
several microns thickness [17] or the use of DLC coating
on steel shim can remarkably reduce the friction loss in
cam/follower arrangement [18].
Apart from the use of different kinds of wear-resistant
and friction reduction coatings, several expendable coatings or surface treatment techniques are also used to
improve the running-in performance of the contacting
surfaces of the engine components. Some of the wear
resistant coatings show good running-in properties while
others need to be deposited with running-in coatings.

ARTICLE IN PRESS
1684

A. Neville et al. / Tribology International 40 (2007) 16801695

Running-in performance is usually improved by modifying


the contacting surfaces by a chemical conversion method
such as phosphating, oxidizing, sulnuz, tufftriding, etc.,
where the immediate surface of metal is converted to metal
compound [4,13]. Such metal compounds provide soft and
porous surfaces and thus promote wettability and oil
entrainment. In addition, electroplating, induction/ame
hardening, carburizing, etc., are also used to improve the
running-in performance of the sliding surfaces. The most
commonly used techniques to modify the surface with a
view to improving running-in performance of the sliding
surfaces of the piston-cylinder and valve train assembly
have been given in Table 3. In recent years, because of the
excellent running-in properties of amorphous carbon
coatings [19,20], several attempts have been made to use
DLC coatings on the shims in automotive valve train
application [16,21,22].
1.2. Light alloy integration
The increasing demand for increased fuel-economy
requires reduction of energy consumption, which can
consequently reduce air pollution. The weight reduction
of a vehicle is directly related to the fuel and oil
consumption and to reduction in emissions. As mentioned
earlier, because of the excellent characteristics of aluminium alloys such as their light weight, high strength, high
thermal conductivity and corrosion resistance, the engine
block, which is one of the heavier parts of the engine, is
being manufactured using aluminium alloys. On the other
hand, because of excellent high strength to weight ratio,
good formability, good corrosion resistance and recycling
potential, existing steel or copper-based sheet metals can
potentially be replaced by sheet aluminium. In 2000,
reported by Miller et al. [23], aluminium casting was used
to manufacture 100% of pistons, about 75% of cylinder
heads and for many other power train and chassis
applications. However, poor wear resistance and low
seizure loads have restricted the direct use of aluminium
alloys for cylinder bores and piston skirts giving rise to
bore distortion and poor piston ring sealing and so causes a
huge oil consumption [9]. To improve the wear resistance
and to reduce friction, cast iron is used as liner material but
such use further increases the dimension and wear of
engines. Several approaches have been taken to overcome
the drawbacks of aluminium alloys. The use of Al metalmatrix composites (MMC) reinforced with solid lubricant
(graphite, MoS2, WS2, etc.), hard ceramic particles (e.g.
SiC, Si3N4, Al2O3, etc.) and short bres not only reduces
the friction and wear but also reduces the weight and
thereby reduces fuel consumption and vehicle emissions
[10]. Nanocoatings are developed by deposition of iron
oxide or a mixture of iron oxide and aluminium on the
aluminium alloy followed by fusion using a laser beam and
this approach, which is widely known as Laser Surface
Engineering (LSE), is being considered as a promising
surface treatment technique for the aluminium cylinder

bore [8]. Traditionally, a cast iron liner was used on the


bores of the aluminium alloy engine blocks and this was
later replaced by iron plating. The traditional iron plating
processes require the use of cyanides, which is further
eliminated owing to environmental restrictions [24]. Because of attractive tribological behaviour, several researchers tried to deposit DLC coatings on aluminium alloys
using different deposition techniques [2426]. The naturally
formed oxide layer on the aluminium alloys has been
identied as one of the main limitations, which causes poor
adhesion of the DLC coating to the substrate. Malaczynski
et al. [24] recommended that the aluminium alloy surface
should be cleaned by argon sputtering and then the
adhesion promoter Si:C bond layer of 50 nm can be
formed on the Al alloy by carbon implantation followed by
deposition of the DLC coating. Nie et al. [26] claimed that
the deposition of DLC on the alumina intermediate layer
by plasma-immersion ion implantation provides excellent
wear resistance and low and stable friction coefcient.
However, despite the excellent tribological behaviour of
DLC coatings, they cannot be used in automotive
industries until the mass production of such coatings is
economically feasible. Commercially available engine oils
are basically designed/optimized for the metallic surfaces,
and engine oils optimized for DLC or Al-MMC surfaces
are yet to be commercialized. Therefore, engine oils,
basically designed for metallic (Fe-based) surfaces, have
to be used for the components made of newly developed
composite materials or coated with DLC, and the
replacement of these oils is still posing a technological
challenge.
1.3. DLC as a low friction coating
DLC is a kind of amorphous carbon coating having a
network composed of sp2 (graphite like) and sp3 (diamond
like) bonds, and its physical as well as chemical property
greatly varies with the varying ratio of sp2 and sp3 bonded
atoms. DLC coatings usually show high hardness, excellent
wear resistance, high corrosion resistance, high thermal
and chemical stability, low friction property [27,28] and
excellent running-in property [19,20,27]. Because of the
excellent tribological performance and to meet the increasing demand of fuel economy and clean environment,
recently, DLC coatings are extensively being used for
automotive engine components as well as for other
technical and medical applications [29,30]. Gahlin et al.
[30] claimed that about 30 million coated parts are being
used in automotive industry with an annual increase of
around 50%. Since the tribo-performance of DLC is
greatly inuenced by the environmental parameters such as
temperature, relative humidity, etc. [31,32], the use of DLC
in automotive engine is very crucial because the automotive
components are operated at high temperature, high load,
partial lubrication and in an environment favourable for
oxidation. Therefore, modications of DLC by doping
with hydrogen, different kinds of metals, nitrides and

ARTICLE IN PRESS
A. Neville et al. / Tribology International 40 (2007) 16801695

carbides have been developed to improve the mechanical


and tribological properties and to enhance adhesion
strength of the coating [30,33].
With the increase of temperature the crystalline graphitic
transfer layer is formed by a sp3-sp2 phase transformation of the DLC coatings in dry sliding conditions [34]. The
steady-state low friction of DLC in ambient air is due to
the temperature increase at contact asperities facilitating
the formation of a graphitized tribolayer [35]. Thus DLC
provides self-lubrication in dry air or in inert atmospheric
conditions. But degradation of DLC increases with
temperature and the friction and wear performance
deteriorates because of the damage of DLC coating at
elevated temperature [36]. It has been reported that the
tribological behaviour of DLC starts to change around
100 1C [37] and it starts to delaminate and lose its
effectiveness above 300 1C [38]. High temperature also
facilitates the release of hydrogen from the DLC matrix.
The engine tribo-components, responsible for the major
portion of friction loss, especially piston/cylinder and valve
train assembly, are operated at relatively high temperature,
pressure and sliding velocity and the maximum operating
temperatures at the sliding interfaces of those components
are 300 and 150 1C, respectively [3]. Therefore, the use of
DLC coating in those tribo-components is quite challenging. On the other hand, the increase of humidity level in
the environment reduces the rate of graphitization
[27,31,39] and thereby, increases friction and wear, and
the friction coefcient may go as high as 0.1 in high
humidity conditions. Thus the temperature and environmental parameters greatly inuence the stability of friction
and wear.
1.4. Lubricated contactssteel/lubricant and coating/
lubricant interactions
1.4.1. Steel/lubricant interactionsfocus on ZDDP and
MoDTC
Interactions between the lubricant and the lubricated
component surface can be of physical in nature and/or
chemical in nature. In terms of physical interactions,
nanostructured coatings and laser texturing of surfaces

1685

were seen to improve lubrication [2]. However, wear and


friction performance in the boundary lubrication regime is
controlled mainly from the chemical reactivity of lubricant
additives, which form tribolms in the contacting surfaces.
In the following sections, interactions of the zinc dialkyl
dithiophosphate (ZDDP) and molybdenum dialkyl dithiocarbamate (MoDTC) additives with the lubricated steel
surface material in improving wear and friction are
reviewed.
1.4.1.1. ZDDP. Most common anti-wear additives used
in practice are organochlorine, organosulphur, organophosphorus (tricresyl phosphateTCP and dibutyl phosphiteDBP), organometallic (ZDDP, MoDTP, MoDTC)
and organic borate compounds [40,41]. To deliver their
anti-wear functionality, anti-wear additives, through tribochemical reactions, facilitate the formation of a very thin
lm (now commonly referred to as the tribolm) on the
lubricated surface.
The zinc dialkyl dithiophosphate (generally referred to
as ZDDP or ZDTP) additive is one of the most widely used
in engine lubricants. This is because it is shown to have
multifunctional properties, namely anti-wear and antioxidant action. ZDDP has been subject of several review
papers [4244], but in this review the focus will be on the
ZDDP/Fe-based surface interactions. The composition of
ZDDP tribolms in Fe-based materials is analysed with a
wide range of surface analytical techniques and shown to
be comprised of different layers. In general, it is agreed that
ZDDP form a glassy phosphate lm, with different chain
length as a function of depth, on top of an oxy/sulphide
layer. A schematic representation of a ZDDP tribolm is
given in Fig. 1.
Considering that rubbing is necessary to produce the
speciation of phosphorus and sulphur, Martin [46,47],
proposed a model for lm formation with metal dithiophosphates in which both chemical and mechanical aspects are
linked together. According to Martin [46,47], reduction of
polyphosphate chain length is a result of the chemical
process between the long-chain polyphosphate formed
from ZDDP decomposition with the iron oxide wear
particles. On the basis of the hard and soft acids and bases

Hydrocarbon-rich Layer
100 to
1000 nm
approx.

Glassy
Polyphosphate
+ ZnO, ZnS

Organic
Radicals
Increasing

Fe, FeO
FeS
Increasing

Inorganic Layer
Iron Sulphide
and/or Oxides
Iron/Steel Substrate

Fig. 1. ZDDP lm structure [45].

ARTICLE IN PRESS
A. Neville et al. / Tribology International 40 (2007) 16801695

1686

(HSAB) principle a set of tribochemical reactions between


the polyphosphates and the oxides were proposed [48].
According to this approach, a possible route for the
elimination of 1 mol Fe2O3, and formation of short-chain
polyphosphates, will be
5ZnPO3 2 Fe2 O3 ! Fe2 Zn3 P10 O31 2ZnO:
ZnO; P2 O5

(1)

3ZnO; 5P2 O5 ; Fe2 O3

By this, three-body abrasive wear is practically eliminated, suggesting that ZDDP effectiveness in wear reduction is a result of its interaction with the substrate debris, in
this case iron oxide. This reaction develops because the
Fe3+ is a harder Lewis acid than Zn2+ and the cation
exchange is energetically favourable from the point of view
of the HSAB principle. Phosphates are hard bases and will
specically react with hard acids, resulting in reaction
between iron and the phosphate.
According to some authors [49], the ZDDP tribolm
formation is not dependent on the nature of the surface.
From XANES spectra made on thermal lms and antiwear lms it was concluded that ZDDP-derived anti-wear
and thermal lms on steel are chemically similar [4951]. In
these studies, the presence of Fe2O3 is not seen as being a
requirement for ZDDP anti-wear lm formation. The
formation of short-chain polyphosphates, observed in
ZDDP tribolms, is proposed to be as a result of hydrolysis
of polyphosphates [49]. A linkage isomer of ZDDP (LIZDDP) was proposed as an important precursor for lm
formation after analysis of thermal lms and lubricant
insoluble ZDDP decomposition products. Thus a mechanism of lm formation from ZDDP is suggested and is
explained with the reactions below:
ZnRO2 PS2 2 solution ) ZnRO2 PS2 2
ZDDP adsorbed;
ZnRO2PS2 2 solution ) ZnO2 PSR2 2
LIZDDP in solution;
ZnO2 PSR2 2 solution ) ZnO2 PSR2 2
LIZDDP adsorbed;
ZnRO4 P2 S4 O2 or ROOH ) ZnPO3 2 polyphosphate
sulphur species;

7ZnPO3 2 6H2 O ) Zn7 P5 O16 2 4H3 PO4 ;


2ZnPO3 2 3H2 O ) Zn2 P2 O7 2H3 PO4 ;
shortchain polyphosphates:

(2)

During steady state, where there is reduced formation of


substrate debris, wear protection is suggested to be a result
of wear debris from ZDDP tribolm re-entering the
contact and by that reduce wear signicantly [46,52]. The
ZDDP debris is generated as a result of ZDDP tribolm

delamination. Another mechanisms by which the ZDDP


phosphate lms reduce wear are that the lm material can
behave as a viscous lubricant in boundary conditions [46].
The viscosity of a glass varies continuously with the
temperature, and can decrease appreciably under shear.
The mechanisms by which ZDDP tribolms are formed are
important if new surfaces are to be effectively protected by
existing lubricants. With there still being some doubts
about whether the production of Fe-containing species is a
necessary requirement for tribolm formation, it is not
certain whether non-Fe-based surfaces will be effectively
protected from wear by ZDDP.
Another interaction between the surface and ZDDP
additive is seen in the formation of iron sulphide. Runningin wear, when ZDDP was used, is believed to take place
principally by adhesive and oxidative processes [45].
Localized adhesion and metal transfer between contacting
asperities is prevented from growing into larger scale
adhesion, by reacted iron sulphide lms formed by the
sulphide products that occur early in the frictional
decomposition process of ZDDP.
Formation of iron sulphide and Zn/Fe phosphate shows
the interaction of ZDDP additive with the surface resulting
in formation of the tribolm.
1.4.1.2. MoDTC. Engine efciency can be increased by
reducing the mechanical losses, which are mostly caused
by friction. One of the ways to reduce mechanical losses is
by the use of friction reduction additives, which operate in
boundary and mixed lubrication conditions. These additives generally fall into two classes: the physically adsorbed
molecules such as fatty acids and amides, and the
chemically reactive species such as molybdenum dithiocarbamates [5355].
A very important class of friction-reducing additives,
which are used extensively in lubricant formulations are
molybdenumsulphur containing compounds. MoS compounds are introduced in late 1950s [56] and interests on
them increased since the 1970s. These compounds are
documented to reduce friction by forming a MoS2 lm on
metal surfaces [5766]. The layer-lattice structure of the
molybdenum disulphide makes it possible for these
compounds to facilitate low friction [57]. In molybdenum
disulphide there is powerful covalent bonding between
atomic species, but between lattice layers there is only very
weak Van der Waals attraction. The weak Van der
Waals forces between MoS2 layers maintain easy shear
within the molecule and are responsible for the low friction
properties.
Analyses of MoDTC lms formed on non-rubbing
surfaces [65,67], lms that are formed mainly due to
thermal decomposition of MoDTC, have shown that only
MoO3 is formed. Surface analyses with X-ray Photoelectron Spectroscope (XPS) [58,6769] have shown that only
MoO3 forms outside the wear scar, while in the wear scar
MoS2 and MoO3 are detected. The factors associated with
rubbing and which could stimulate MoS2 formation in the

ARTICLE IN PRESS
A. Neville et al. / Tribology International 40 (2007) 16801695

1687

wear scar are (a) contact temperature and (b) exposure of


the metal surface due to the removal of oxide lm. The
contact temperature and mechanical removal of the oxide
layer, which will form a nascent iron surface, will initiate
formation of MoS2 at the contact points. This mechanism
of MoS2 formation explains why no MoS2 is formed
outside the wear scar, since there is no a rubbing and hence
no high contact temperature or nascent iron surface
produced.
Grossiord et al. [65] in addition to nding MoS2 and
MoO3, also found some residual MoDTC in the tribolm.
From the chemical point of view, they suggested the
MoDTC decomposition occurs by a two-step process, as
shown in Fig. 2:

N-containing radical, formed from breakdown of MoDTC,


reacts with the nascent surface according to the HSAB
principle, since S2 and metal atoms are known to be soft
base and acid, respectively, forming FeSx and leaving the N
part to deposit on the surface [70]. The formation of iron
sulphide from MoDTC has been observed in other studies,
too [56,72], showing another interaction between the
MoDTC additive and the surface material. Formation of
iron sulphide will act as a protective layer reducing wear
[56,72,73] and by that allowing the formation of a frictionreducing layer of MoS2 from the other radical.
Interaction between MoDTC and surface material is
seen on formation of MoS2 linked with rubbing and
formation of iron sulphide species.

1.4.2. Coating/lubricant interactions


DLC is considered as a promising non-ferrous coating
because of its excellent tribological properties as mentioned
earlier. Since the temperature and environmental parameters greatly inuence the stability of friction and wear of
DLC coatings, the use of lubricating oil will isolate the
coating from the surrounding hostile environment and act
as a coolant to keep the temperature in the allowable limit.
In addition, the lubricant additives may interact with the
DLC coating and produce low friction and wear resistant
tribolms. Most of the lubricants developed so far are
customized to form the tribolm that will adhere to ferrous
materials and until now no lubricants have been designed
for the non-ferrous coatings. To design new lubricants for
non-ferrous coatings, especially for the DLC coating, it is
essentially important to understand how this coating
interacts with the existing lubricant additives. Knowledge
of how DLC interacts with existing lubricant additives is
scarce and contradictory. For example, Kano and Yasuda
[74] showed that no stable tribolm was found on the DLC
surface, when additive containing oils were used, while
other research groups observed tribolm formation on
DLC at boundary lubrication condition [75,76]. The
physical properties of DLC coatings will signicantly vary
with deposition techniques, type of dopant used and the
hydrogen content in the DLC matrix. Hence, any research
on the interactions between DLC and lubricant additives
should take into account the intrinsic properties of DLC.

MoS chemical bond breakage in MoDTC additive due


to the electron transfer, leading to three free radicals:
one corresponding to the core of the molecule and the
other two to the chain ends.
Recombination of chain end radicals to form thiuram
disulphide, whereas the core radical decomposes into
MoS2 and MoO2. MoO2 can oxidize in the presence of
air and form MoO3.

This is in agreement with the XPS detection of MoS2 and


MoO3 in the tribolm but the role of the two chain free Ncontaining radicals is not claried. A recent work, using
XPS, has shown the presence of N and S-containing species
in the MoDTC tribolms [70]. It is observed that these
species are formed before the formation of any Mo species
inuencing the evolution of the low friction lm formation
[70,71]. It is suggested that one possible mechanism
for nitrogen being at the surface is that the S in the

a
R

S
N

Mo

Mo

electron transfer
S

R
N

b
N

Mo*

Mo*

S*

c
S S

R
N
R

R
C

MoS2 + MoO2

S S
thiuram disulfide
MoS2 sheet

+ 12 O2

R
Mo

S
Mo
S

Mo
S

Fig. 2. MoDTC decomposition [65].

MoO3
molybdenum
oxide

1.4.2.1. Role of hydrogen in DLC/lubricant interactions. DLC coatings are generally described as hydrogenated amorphous carbon (a-C:H) and non-hydrogenated
amorphous carbon (a-C). Hydrogen stabilizes the random
covalent network of DLC and prevents its collapse into a
graphitic phase during deposition [77,78]. In the dry
atmospheric condition, release of hydrogen occurs due to
high contact temperature resulting in the conversion of sp3
to sp2 structure and provides high wear of a-C:H coating.
However, in oil lubricated condition, DLC-coated surfaces
are believed to be separated by thin oil lm and the
presence of oil at the sliding contact may act as coolant and
reduce contact temperature. Thus the formation of

ARTICLE IN PRESS
1688

A. Neville et al. / Tribology International 40 (2007) 16801695

graphitic layer is believed to be suppressed and the


tribochemical interactions of oil additives and DLC
surfaces become more dominant.
Yasuda et al. [79] observed that the a-C:H exhibited high
friction (0.11) and the a-C gave low friction coefcient
(0.07) while they slid against steel pin (AISI 52100) using
engine oil 5W-30 (SJ grade) to lubricate the surface.
Surface analyses performed with XPS on a-C did not show
any tribolm formed, although a tribolm was detected in
the steel counterbody [74]. The low friction observed in a-C
is assumed to be related to the low surface energy of a-C:H
as compared to a-C, effecting the adsorption of oil
additives. Now the question arises why the surface of aC:H is chemically inert and why the surface energy is low.
To explain such characteristics of a-C:H, it is necessary to
investigate the atomic structure of DLC rst. The structure
of DLC is inherently meta-stable and the heat generated
from the tribo-friction can easily transform the sp3
structure into sp2 structure (graphite-like). In sp2 conguration, three of the four valance electrons of a carbon
atom are arranged trigonally and forms strong covalent
bond with the nearest neighbouring atoms (Fig. 3b). The
fourth electron which is known as p electron perpendicularly lies on the sp2 hybrids as shown in Fig. 3b. Thus it
forms the perfect graphite lattice where the atoms in
hexagonal network are staked in the sequence of ABAB
(Fig. 3a). The dangling covalent bond of the sp2 bonded
cluster remains at high-energy state and can easily be
passivated by the adsorbed species from the surrounding
environment. The possible reason of low friction exhibited
by a-C, as observed in [74,79], is because the oil additives
were easily adsorbed and passivated the surface resulting in
low friction. In contrast, in case of a-C:H, the dangling
bonds are passivated by the hydrogen atoms within the
coating and thus the hydrogencarbon combination gives a
non-polar inert surface having low surface energy which
further causes poor wettability and probably little or no
adsorption of oil additives on the sliding surface. Mabachi

et al. [80] performed experiments in cylinder head cam/


follower rig using oil containing MoDTC and found that
a-C coatings provided low friction torque as compared to
a-C:H.
A completely opposite picture was portrayed by BarrosBouchet et al. [75] where they explained the positive role of
hydrogen in the formation of tribolm on the DLC
coating. In their observation, the a-C:H showed low
friction as compared to a-C. They found that the MoDTC
and ZDDP are more active on the a-C:H facilitating the
formation of MoS2. It is interesting to note that no P was
found in the tribolm and consequently no ZDDP-derived
anti-wear lm was noticed. However, it has been claimed
that ZDDP enhanced the formation of MoS2 on the
amorphous carbon surface by supplying S atoms. From
XPS analysis, they observed that the ratio of MoS2/MoO3
for hydrogenated DLC is ve times higher than that
observed on non-hydrogenated DLC. They argued that the
hydrogen-terminated surface was damaged by the sliding
action and gave nascent dangling bonds, which reacted
with lubricant additives. They explained such phenomenon
by making assumption based on chemical hardness
approach (HSAB). According to their argument, the soft
base hydrogenated carbon reacts with soft acid Mo+4 and
forms low friction MoS2 layer while hydrogen-free carbon
material which is considered as intermediate base reacts
with hard acid Mo+6 and forms high friction MoO3 layer.
Thus the a-C:H promotes the formation of MoS2 and
reduces friction efciently while the a-C gives detrimental
effect to the friction performance by forming large amount
of MoO3.
1.4.2.2. Role of ferrous and non-ferrous dopants in DLC/
lubricant interactions. Metal is usually doped in DLC to
improve durability, reliability [81] and to maintain low
friction and wear under severe operating conditions
[8284]. Metal doped in DLC (Me-C:H) forms metal
carbide and improves the strength of DLC lms, and it is

Fig. 3. (a) Atomic arrangement of lamellar graphite, (b) atomic structure of sp2 hybridized graphite layer transformed from sp3 hybridized DLC structure.

ARTICLE IN PRESS
A. Neville et al. / Tribology International 40 (2007) 16801695

reported that boundary lubrication performance can be


improved if a high strength metal carbide is formed in the
defective part of the cross-linked carbon network of DLC
[76]. In comparison to the tribological performance of nonmetal-doped DLC, metal (Ti, Mo and Fe)-doped DLC
outperformed at boundary lubrication condition using oil
containing both ZDDP and MoDTC, and Ti-DLC showed
superior performance over Mo-DLC and Fe-DLC because
it provided favourable surface to form low shear strength
friction layer from MoDTC and ZDDP additives.
Ban et al. [81] performed the XPS analysis of the
tribolms formed by the oil containing ZDDP on the Sidoped DLC and found the presence of ZnO, ZnS, FePO4,
FeS and FeS2 compounds, and it was believed that those
compounds contributed to achieve low friction and high
wear resistance. Although the presence of Si in DLC
reduces the friction property, they observed that the
friction coefcient linearly increased with Si percentage in
DLC matrix. On the other hand, Gangopadhyay et al. [21]
investigated the performance of Si-containing amorphous
hydrogenated DLC coating as the coating on shim sliding
against cast iron cam lobes. Manganese phosphate coating
was deposited on the top of DLC coating to improve the
braking-in performance. Interestingly, the DLC-coated
shim showed higher friction torque than that of production
steel shim. It was argued that the use of oil prevented the
formation of friction layer as found on DLC at dry sliding
condition but no elemental analysis was performed to
support their argument.
In case of W containing DLC, low friction tribo-layer of
WS2 nanocrystals on the steel counter surface is formed
when S containing EP additive was used in oil under
boundary lubrication condition [85,86]. It has been argued
that S of EP additives react with the nanocrystalline W and
WC, transferred from the W-DLC coating, and forms WS2
tribolm, and consequently provides low friction behaviour. Podgornik and Vizintin [87] showed that the increase
of the concentration of S containing EP additive has no
effect on the tribological behaviour of a-C:H while it gives
signicant improvement in friction performance of WDLC coating by forming low friction lamellar WS2
tribolm. The P containing anti-wear additive gave high
friction for both a-C:H and W-DLC but no tribochemical
analysis was performed to nd if any anti-wear lm wear
formed on the DLC coatings.
Kalin [88] observed that doped and undoped DLC
showed different tribological behaviour in presence of EP
and AW additives under boundary lubrication condition.
The steel/undoped DLC (a-C:H) combination showed that
the use of base oil provided low wear of a-C:H while EP/
AW containing oil gave high wear. In case of doped DLC
(TiC:H, SiC:H), the additive containing oils showed low
wear as compared to base oil.
Miyanaga et al. [89] used fully formulated oil and found
that Ti-DLC showed the highest wear resistance and the
lowest friction coefcient among all other coatings (a-C:H,
MoS2, TiN, Mn-phosphate). Using the Secondary Ion

1689

Mass Spectroscopic (SIMS) analysis, they claimed that


more adsorbed extreme pressure agent was observed on the
Ti-DLC than that of undoped-DLC.
1.4.2.3. Selection of the tribo-materials/coatings. The
selection of tribo-metal and/or tribo-coating combination
is a challenging issue because the reactivity of the additive
containing oil mainly varies with the property of friction
surfaces as summarized in Table 4. The most possible
combinations of coatings and substrate materials operated
under boundary lubrication condition using wide range of
oil formulations have been compiled from the published
papers, and based on the friction and wear performance,
suitable combinations have been sorted out. In two cases,
friction coefcient of steelsteel combination was found
higher than a-C/steel combination [27,90] while it was
lower than a-C/steel combination in one case [75]. On the
other hand, in two cases, a-C/steel showed lower friction
coefcient than a-C:H/steel [27,90] whereas in one case aC:H showed better friction performance than a-C/steel
combination [75]. Therefore, higher friction performance
of a-C/steel was supported by more researchers than those
of steel/steel and a-C:H /steel combinations.
No measurable wear of DLC coating was observed in the
study of Ronkainen et al. [27], and the possible reason for
such results might be because of the application of low
force and sliding velocity. Stallard and Teer [90], who
applied comparatively high force and sliding velocity,
noticed signicant wear of DLC coatings. A similar trend
of wear of DLC coatings sliding against cast iron counterbody has been reported by Haque et al. [91]. In terms of
wear performance, the wear of steel counterbody sliding
against a-C:H was found to be lower than in steel/steel and
a-C/steel combinations [27,75,90]. However, when the wear
of DLC coatings are compared, a-C sliding against steel
showed lower wear than those of a-C:H and steel [90]. This
could be because of the high hardness of a-C compared to
a-C:H which results in high plastic deformation in the
softer counterbody.
Results from a study by Barros-Bouchet et al. [75], as
shown in Table 4, clearly indicate that a-C:H/a-C:H and
TiC:H/ TiC:H combination exhibit lower wear performance than a-C:H/steel, TiC:H/steel and steel/steel
combination but the improvement of frictional performance using DLC/DLC combination is found to be low
compared with the DLC/steel combination. XPS analyses
on a-C:H and TiC:H sample showed that a MoDTCtribolm is formed [75], suggesting that formation of low
friction tribolm from MoDTC is not affected by the
presence of iron in the contact. The XPS results also
suggest MoDTC and ZDDP are more active on the a-C:H/
a-C:H contact than on TiC:H.
Conversely, Podgornik et al. [19,20] reported that, under
boundary lubrication conditions using additive-containing
oil, the running-in time of WC-C:H/WC-C:H combination
was much longer resulting in a higher wear rate than that
of WC-C:H /steel combination. This is to be suggested

1690

Table 4
Selection of tribo-metal and tribo-coating combination: role of hydrogenated and metal-doped DLC
Source

Experimental condition

Lubricant and
lubrication

Counterbody

Coatings on the plate/disc

Friction

Total wear
Counterbody
300  109 mm3/Nm
80  109 mm3/Nm
50  109 mm3/Nm
90  109 mm3/Nm

Test: pin-on-disc
Force: 10 N
Hz. Pr.:0398 Gpa
V: 0.004 m/s

Mineral base oil+EP


additive
Boundary lubrication
Cond.

AISI 52100 (100Cr6)

a-C
a-C:H
a-C:H (Ti)
AISI 52100

0.08
0.13
0.10
0.13

No
No
No
No

Stallard et al. [90]

Test: pin-on-disc
Force: 40 N
Hz. Pr.: 1.6 Gpa
V: 0.2 m/s

Semisynthetic oil
(10 W40)

AISI 52100

a-C
a-C:H
AISI 52100

0.07
0.1
0.09

0.03  109 mm3/Nm 4.3  109 mm3/Nm


0.12  109 mm3/Nm 0.6  109 mm3/Nm
0.35  109 mm3/Nm 7.4  109 mm3/Nm

Barros Bouchet et al.


[75]

Test: cylinder-on-at
Force: 50350 N
Hz. Pr.: 0.6 Gpa (Max)
V: 0.2 m/s

PAO+MoDTC+ZDDP AISI 52100

a-C:H (50 at.% H)


TiC:H (35 at.%H)

a-C
a-C:H (50 at.% H)
Ti-C:H (35 at.% H)
AISI 52100
a-C:H (50 at.% H)
TiC:H (35 at.%H)

0.08
0.05
0.05
0.06
0.04
0.06

Not
Not
Not
Not
Not
Not

WC-C:H

WC-C:H

0.07

0.98  103 mm3/min

Steel

WC-C:H

0.07

0.6  103 mm3/min

Podgornik et al. [19]

Tung and Gao. [92]

Haque et al. [91]

wear
wear
wear
wear

mentioned
mentioned
mentioned
mentioned
mentioned
mentioned

50  109 mm3/Nm
2  109 mm3/Nm
2  109 mm3/Nm
6  109 mm3/Nm
0.07  109 mm3/Nm
0.4  109 mm3/Nm

Test: two crossed


cylinder
Force: 1401700 N
Hz. Pr.: 2.45.6 GPa

GL-4 (fully formulated


oil)

Test: Modied high


frequency reciprocating
friction machine
Force: 80 N
Oscillation frequency:
10 Hz
V: 0.138 m/s Sliding
distance: 6.9 mm

PAO/engine oil

Steel piston ring

Cast iron cylinder bore

0.11

7.40.6  109 m3

3.7  109 m3

PAO+MoDTC

DLC-coated piston ring


Steel piston ring

Cast iron cylinder bore


Cast iron cylinder bore

0.1
0.04

7.4  109 m3
0.8  109 m3

0.8  109 m3
2.8  109 m3

DLC-coated piston ring

Cast iron cylinder bore

0.08

0.8  109 m3

0.4  109 m3

Cast iron

a-C:H

0.07

9.18  1019 m3/Nm

9.45  1019 m3/Nm

Test: pin-on-plate test


Force: 326 N
Hz. Pr.: 0.560.64 GPa

Base oil+ZDDP+Moly
Dimer (MoDTC)

ARTICLE IN PRESS

Ronkainen et al. [27]

A. Neville et al. / Tribology International 40 (2007) 16801695

Plate/disc

ARTICLE IN PRESS
A. Neville et al. / Tribology International 40 (2007) 16801695

because with the WC-C:H/WC-C:H combination, more


contact cycles were necessary to smoothen the surfaces and
provide conformity whereas, for the DLC/steel combination, smoothening of the steel with the harder DLC counter
surface facilitated good running-in performance. It was
suggested that the WC-C:H/WC-C:H combination showed
poor frictional performance because no tribologically
activated chemical reaction occurred on the WC-C:H. On
the other hand, formation of the synergistic tribolm on
steel at WC-C:H /steel was claimed and it was believed that
the tribolm facilitated low friction and wear.
As mentioned earlier, the use of steel as a piston ring or
shim material and cast iron as cylinder liner or cam
material is still popular in the automotive industries.
However, some research has been performed to assess the
feasibility of using DLC coatings on steel shims or piston
rings sliding against cast iron. Tung and Gao [92]
performed bench tests to investigate the possibility of
using DLC coating on steel instead of using bare steel ring
sliding against cast iron cylinder bore. They found that the
use of PAO gives higher friction coefcient than that of
using PAO with MoDTC for both DLC/cast iron and steel/
cast iron combination (Table 4). However, no signicant
improvement in friction behaviour was noticed by using
DLC coated steel ring over the uncoated steel ring. It is
interesting to note that the results of material wear clearly
indicate that the use of DLC can signicantly reduce the
wear of cast iron counter surface. In addition, MoDTC
further can reduce wear in both DLC/cast iron and steel/
cast iron combinations. Therefore, it is quite reasonable to
use DLC as a tribo-coating for steel piston ring in piston
cylinder assembly. Although the contact pressure at piston
ring and cylinder liner (10 MPa) is much lower than that
of cam/follower contact (0.6 GPa), similar approach of
the selection of material couple can be checked for cam/
follower application.
This review has shown that the range of surface coatings
and treatments used in automotive components is diverse
and opens up opportunities to improve friction and wear
performance. There is so far little published literature to
demonstrate how lubricant additive/surface combinations
can be designed to optimize performance. In this next part
of the paper, some experimental results using a limited subset of oils/additives and tribocouples are presented to
illustrate how synergistic or antagonistic effects can exist
and that there is real potential if the compatibility is
matched.
2. Experimental details
Two ferrous-based materials (BS EN1452 cast iron and
BS EN31 Cr-bearing steel) and an AlSi piston alloy were
studied. Tribological properties of these materials in
lubricating oils were examined using a pin-on-reciprocating
plate system. The pin was loaded using a static load applied
through a lever arm and the plate was reciprocated
underneath through a crank mechanism.

1691

The pins used were 20 mm in length; 6 mm in diameter


and the ends of the pins were machined to a 20 mm radius
of curvature. The rectangular plate measured 17  6 
3 mm3. The test-stroke and test frequency are 10 mm and
1 Hz, respectively. The load applied on the pin was 185 N,
which corresponds to a Hertzian pressure of between 0.5
and 0.9 GPa (depending on the material couple). A bidirectional load cell measured the friction force on the
static pin during the course of the experiment. The friction
force was recorded every 30 min for 3 cycles during the test
duration of 8 h. Tests were conducted at 100 1C.
Using the friction force data, the friction coefcient (m)
was calculated by rst calculating the average absolute
friction force and then dividing this by the applied load.
This is then plotted as a function of time and the average
friction coefcient is calculated for the 8-h period. Three
seconds of data (friction force versus time) were recorded
at 30-min intervals. A square wave trace was recorded as
expected for a reciprocating motion.
Typical friction coefcient values of 0.070.18 were
recorded suggesting that the contact regime is in boundary
lubrication and this is veried from the square wave traces.
The nature of the contact regime is further veried by the
calculation of the lambda ratio, L, calculated from the
expression by Dowson and Hamrock [93]. The lambda
ratio (based on the starting values of Rq) for all the
materials is below unity (0.055oLo0.176).
At the start of each experiment, the pin and plate were
cleaned using acetone to degrease and remove any
impurities from the surface. They were then weighed using
a Mettler AJ150 micro-balance with an accuracy of
70.1 mg. Lubricating oils supplied by Inneum UK Ltd.
were used in each test. The oil charge for each test was
2.5 ml.
After each test the pins and plates were dipped in
heptane for 12 s to remove excess oil from the surface and
then weighed to calculate the volume of material loss from
the plate, pin and the pin/plate system. The plates and pins
were then analysed using optical microscopy, Environmental Scanning Electron Microscopy (ESEM) with EDX
attachment for chemical analysis and XPS.
2.1. Material and lubricants characteristics
This study focuses on the combined effects of material
couple and the nature of the lubricant additives.
2.1.1. Materials
In this paper, current representative engine materials
were tested. The materials tested included plates of four
materials; BS EN31 Cr-bearing steel, AlSi alloy (supplied
by Federal Mogul Ltd.), hydrogenated DLC coating
(Diamonicproduced by Teer Coatings LTD) and pins
of EN 1452 cast iron and steel pins with DLC coating. The
EN31 has a measured hardness and density of 185 HV and
7.8 g/cm3, respectively while the EN1452 has a measured
hardness and density of 150 HV and 7 g/cm3, respectively.

ARTICLE IN PRESS
A. Neville et al. / Tribology International 40 (2007) 16801695

1692

The AlSi piston alloy has a hardness measuring 120 HV


and density of 2.6 g/cm3. The AlSi alloy contains 23 wt%
Si and trace amounts of Cu (1.1 wt%) and Mg (1.3 wt%) in
addition to C, O and Al. The DLC coating hardness is
specied by Teer Coatings to be 1400 HV and has a
thickness of approximately 5 mm. The material couples
tested (pin/plate) were cast iron/EN 31, cast iron/AlSi,
cast iron/DLC and DLC/DLC.
2.1.2. Lubricants
Test were run using four lubricants referred to as: base
oil, base oil+MoDTC, base oil+MoDTC+ZDDP, base
oil+ZDDP. The names are self-explanatory and the level
of additive was 500ppm in the case of the MoDTC additive
and 1.2 wt% of ZDDP (at 0.05%P). The base oil used was
a Group III base oil. The group designation relates to the
saturate level, the sulphur content and the viscosity index
of the crude oil. The base oil has a dynamic viscosity of
3  103 N-s/m2 at 100 1C.
3. Results and discussion

0.25

CI/Al-Si

0.2

CI/EN-31

0.15

CI/DLC
DLC/DLC

0.1
0.05
0

Tribocouple

Adding ZDDPMoDTC

Adding MoDTCZDDP

CI/EN 31
CI/Al-Si
CI/DLC
DLC/DLC

Reductionsynergy
Increaseantagonism
Small increaseantagonism
Increaseantagonism

Reductionsynergy
Increaseantagonism
Reductionsynergy
Reductionsynergy

couple. To assess this, the information in Table 5 sets out


whether by adding a second additive (here ZDDP or
MoDTC) to a single additive oil (again MoDTC or ZDDP)
increases or decreases friction. As an example, where it says
the effect of adding ZDDP to MoDTC then leads to a
comparison between the friction in the Oil+MoDTC and
Oil+MoDTC+ZDDP system. The main message conveyed from these results is that there is not a consistent
effect across all four tribocouples. The interactions between
the additives are also dependent on the nature of the
rubbing surfaces. The results for the ferrous system are
consistent with the literatureconrming the synergistic
effects between ZDDP and MoDTC in reducing friction
[9496].
Table 5 and Fig. 4 summarize how the additives and
their interactions affect friction. In this study, the oils
containing ZDDP and tribocouples were assessed in
relation to durability and tribolm formation from the
ZDDP additive. The characterization of the polyphosphate
glass, formed at the tribological interface during rubbing in
presence of ZDDP, has been by assessing the nature of the
oxygen detected by XPS. The oxygen peak at 132 eV was
analysed and the ratio of Bridging (POP) to non-bridging
(P 0 and POZn) oxygen, which is equal to BO/
NBO (n1)/2(n+1) [47]. From this the glass polymerization number (n) is calculated. In the cases where n 1,
the glass is an orthophosphate, n 2 is a pyrophosphate
and in the case that n is higher than 2, it is a metaphosphate
[97].
Fig. 5 shows the change in n as a function of material
tribocouple, oil and as a function of depth. The depth
proling is given as an etch time in minutes and no attempt
has been to quantify the exact depth given the uncertainties
about the nature of the tribolm and how the removal rate
by etching changes as a function of depth into the
tribolm. Hence the comparison at this stage is qualitative
only. However, some very clear trends are shown (and
these are also shown in Table 6):

D
+Z
TC

oD

oi
ba

se

oi

l+

se
ba

ba

se

oi

l+

l+

ZD

oD

l
oi
ba
se

TC

Final friction coefficient

The tests were run for 8 h and the friction was monitored
as a function of time during that period. In this paper, the
friction coefcient at the end of the test (assuming the trace
is stable at this point) is compared. In this way the transient
effects often seen in friction-modied oils as the additive is
activated are not considered.
Fig. 4 summarizes all of the nal friction coefcient data
for the four material couples lubricated by the four oils. It
is clearly seen that the friction coefcient depends both on
the material couple and the lubricant, as expected. DLC
provides inherently lower friction coefcients than the
other tribocouples using three of the four oils but with
Oil+MoDTC+ZDDP the lowest friction is observed by
the ferrous system.
What is of interest in the context of this current paper is
the interactions between lubricant additives and as importantly how these interactions depend on material

Table 5
Synergisticantagonistic effects of lubricant additives on friction on four
tribocouples

Fig. 4. Friction coefcients (nal values) measured for four tribocouples


with four oil types.

In all cases, the polyphosphate glass has a higher


polymerization number at the outer edge, consistent
with the theory that long-chain polyphosphates are
transformed during the rubbing process to shorter chain
polyphosphates
On addition of MoDTC to ZDDP a mixed effect on the
value of n is observed. For CI/EN 31, Ci/DLC there is a

ARTICLE IN PRESS
A. Neville et al. / Tribology International 40 (2007) 16801695

in term of providing optimum performance in terms of fuel


economy (reduced friction) and durability (reduced wear)
can be summarized as:

Base oil+ZDDP

1693

Base oil+MoDTC+ZDDP

4
3
2

1
0
5

10

15

Etching time (mins)

CI/EN31

CI/Al-Si

CI/DLC

DLC/DLC

Fig. 5. Polymerization number, n, as a function of the tribocouple


material, oil and etching time.

Table 6
Effect of MoDTC on the nature of the phosphate lm formed at the
tribocouple (plate) surface from the ZDDP additive
Tribocouple

Effect on n/wear

CI/EN 31
CI/Al-Si
CI/DLC
DLC/DLC

Reduction/decrease
Increase/increase
Reduction/decrease
Small reduction/no measurable change

To optimize synergies between additives for a specific


tribocouplein the small subset of materials and
lubricants investigated here it has been demonstrated
that synergies between MoDTC and ZDDP are maximized for the ferrous system.
To exploit the attributes of both the surface and lubricant
additives to optimize system performancethere are
several developments in lubricious coatings and these
can be used to provide a proportion of friction-reducing
capability of a system as oil formulations are changed.

signicant reduction in n, for CI/AlSi there is a


signicant increase in n. For these couples the higher n
is associated with higher wear and vice versa.
For the DLC/DLC couple the differences were smalla
small reduction in n with no measurable change in what
was a very small wear rate.

4. Conclusions
The rst part of this paper reviewed the current range of
materials that are employed in various components in the
internal combustion engine and from this review it is clear
that there is a general trend towards the use of surface
engineering to improve durability. This is in parallel with
increasingly stringent legislation, which is ensuring that oil
formulators must reduce (and eventually eliminate) the
level of P (and S) in engine oils. To provide the
functionality required without conventional additive systems is going to require that surfaces are going to have to
work in cooperation with oil additives and this work has
demonstrated that to predict these interactions, and also
interactions between additives, is not simple. The effectiveness of additives in reducing friction and wear and the
interactions between the two additives ZDDP and MoDTC
have been shown to depend on the nature of the
tribocouple. The major challenges, and also opportunities,

References
[1] Korcek S, Sorab J, Johnson MD, Jensen RK. Automotive lubricants
for the next millennium. Ind Lubr Tribol 2000;52(5):20920.
[2] Erdemir A. Review of engineered tribological interfaces for improved
boundary lubrication. Tribol Int 2005;38:24956.
[3] Priest M, Taylor CM. Automobile engine tribologyapproaching
the surface. Wear 2000;241(2):193203.
[4] Enomoto Y, Yamamoto T. New materials in automotive tribology.
Tribol Lett 1998;5:1324.
[5] Tung SC, McMillan ML. Automotive tribology overview of current
advances and challenges for the future. Tribol Int 2004;37(7):51736.
[6] Taylor BJ, Eyre TS. A review of piston ring and cylinder liner
materials. Tribol Int 1979;12(2):7989.
[7] Becker EP. Trends in tribological materials and engine technology.
Tribol Int 2004;37(7):56975.
[8] Yoshida H. Effects of surface treatments on piston ring friction force
and wear. SAE paper 900589, 1990.
[9] Dahotre NB, Nayak S. Nanocoatings for engine application. Surf
Coat Technol 2005;194(1):5867.
[10] Prasad SV, Asthana R. Aluminum metal-matrix composites for
automotive applications: tribological considerations. Tribol Lett
2004;17(3):44553.
[11] McCune GAW RC. Encyclopedia of materials: science and
technology. Automot Engine Mater 2004:42634.
[12] Willson D. SinterCast and the engine of the future. /http://www.
valuerichonlinecom/mag/04fall/storyphp?id=NOsintercastS 2004.
[13] Taylor RC. Valve train lubrication analysis, vehicle tribology. In:
Proceedings of the 17th LeedsLyon symposium on tribology, 1991,
p. 11931.
[14] Nakamura Y, Egami YS. Development of an assembled camshaft by
mechanical bonding. Special publication SP-1138, Society of Automotive Engineers, Warrendale, PA, 1996, SAE paper 960302.
[15] Eyre TS, Crawley B. Camshaft and cam follower materials. Tribol Int
1980;13(4):14752.
[16] Kano M, Tanimoto I. Wear mechanism of high wear-resistant
materials for automotive valve trains. Wear 1991;151(2):22943.
[17] Masuda M, Ujino M, Shimoda K, Nishida K, Marumoto I,
Moriyama Y. Development of titanium nitride coated shim for a
direct acting OHC engine. SAE paper 970002, 1997.
[18] Schamel AR, Grishje M, Bethke R. Amorphous carbon coatings for
low friction and wear and bucket tappet valvetrains. SAE paper
970004, 1997.
[19] Podgornik B, Jacobson S, Hogmark S. Inuence of EP additive
concentration on the tribological behaviour of DLC-coated steel
surfaces. Surf Coat Technol 2005;191(23):35766.
[20] Podgornik B, Jacobson S, Hogmark S. DLC coating of boundary
lubricated componentsadvantages of coating one of the contact
surfaces rather than both or none. Tribol Int 2003;36(11):8439.

ARTICLE IN PRESS
1694

A. Neville et al. / Tribology International 40 (2007) 16801695

[21] Gangopadhyay A, Soltis E, Johnson MD. Valvetrain friction and


wear: inuence of surface engineering and lubricants. Proc Instn
Mech Engrs, J Eng Tribol 2004;218(Part J):14756.
[22] Johnston SV, Hainsworth SV. Effect of DLC coatings on wear in
automotive applications. Surf Eng 2005;21(1):6771.
[23] Miller WS, Zhuang L, Bottema J, Wittebrood AJ, Smet PD, Haszler
A, et al. Recent development in aluminium alloys for the automotive
industry. Mater Sci Eng A 2000;280(1):3749.
[24] Malaczynski GW, Hamdi AH, Elmoursi AA, Qiu X. Diamond-like
carbon coating for aluminum 390 alloyautomotive applications.
Surf Coat Technol 1997;93(23):2806.
[25] Clyne TW, Peng XL. Residual stress and debonding of DLC
lms on metallic substrates. Diamond Relat Mater 1998;7(7):
94450.
[26] Nie X, Wilson A, Leyland A, Matthews A. Deposition of duplex
Al2O3/DLC coatings on Al alloys for tribological applications using
a combined micro-arc oxidation and plasma-immersion ioj implantation technique. Surf Coat Technol 2000;131(13):50613.
[27] Ronkainen H, Varjus S, Holmberg K. Friction and wear properties in
dry, water- and oil-lubricated DLC against alumina and DLC against
steel contacts. Wear 1998;222(2):1208.
[28] Kodali P, Walter KC, Nastasi M. Investigation of mechanical and
tribological properties of amorphous diamond-like carbon coatings.
Tribol Int 1997;30(8):5918.
[29] Hauert R. An overview on the tribological behavior of diamond-like
carbon in technical and medical applications. Tribol Int 2004;
37(1112):9911003.
[30] Gahlin R. ME-C:H coatings in motor vehicles. Wear 2001;249(34):
3029.
[31] Liu Y, Erdemir A, Meletis EI. Inuence of environmental parameters
on the frictional behavior of DLC coatings. Surf Coat Technol
1997:945 [4638].
[32] Enke K, Dimigen H, Hubsch H. Frictional properties of diamondlike
carbon layers. Appl Phys Lett 1980;36(4):2912.
[33] Hogmark S, Jacobson S, Larsson M. Design and evaluation of
tribological coatings. Wear 2000;246(12):2033.
[34] Voevodin AA, Zabinski JS. Nanocomposite and nanostructured
tribological materials for spacet. Compos Sci Technol 2005;65(5):
7418.
[35] Liu Y, Erdemir A, Meletis EI. An investigation of the relationship
between graphitization and frictional behavior of DLC coatings. Surf
Coat Technol 1996;8687(13 Part 2):5648.
[36] Bremond F, Fournier P, Platon F. Test temperature effect on the
tribological behavior of DLC-coated 100C6-steel couples in dry
friction. Wear 2003;254(78):77483.
[37] Vanhulsel A, Blanpain B, Celis JP, Roos J, Dekempeneer E, Smeets J.
Study of the wear behaviour of diamond-like coatings at elevated
temperatures. Surf Coat Technol 1998;98(13):104752.
[38] Erdemir A, Fenske GR. Tribological performance of diamond and
diamondlike carbon lms at evaluated temperatures. Tribol Trans
1996;39(4):78794.
[39] Ronkainen H, Varjus S, Holmberg K. Tribological performance of
different DLC coatings in water-lubricated conditions. Wear
2001;249(34):26771.
[40] Xue Q, Liu W. Tribochemistry and the development of AW and EP
oil additivesa review. Lubr Sci 1994;7(1):8192.
[41] Wei D. Future directions of fundamental research in additive
tribochemistry. Lubr Sci 1995;7(3):21132.
[42] Barnes AM, Bartle KD, Thibon VRA. A review of zinc dialkyldithiophosphates (ZDDPs): characterisation and role in the lubricating oil. Tribol Int 2001;24:38995.
[43] Spikes H. The history and mechanisms of ZDDP. Tribol Lett
2004;17(3):46989.
[44] Nicholls MA, Dao T, Norton PR, Kasrai M, Bancroft GM. Review
of the lubrication of metallic surfaces by zinc dialkyl-dithiophosphates. Tribol Int 2005;38(1):1539.
[45] Bell JC, Delargy KM, Seeney AM. The removal of substrate material
through thick zinc dithiophosphate antiwear lms. In: Dowson D,

[46]

[47]

[48]
[49]

[50]

[51]

[52]

[53]
[54]
[55]
[56]
[57]
[58]

[59]
[60]

[61]

[62]

[63]
[64]

[65]

[66]

[67]

[68]

[69]

et al., editors. Tribology series, 21, wear particles: from the cradle to
the grave. Amsterdam: Elsevier; 1992. p. 38796.
Martin JM. Lubricant additives and the chemistry of rubbing
surfaces: metal dithiophosphates triboreaction lms revisited. Jpn J
Tribol 1997;42(9).
Martin JM, Grossiord C, Mogne TL, Bec S, Tonck A. The two-layer
structure of Zndtp tribolms Part I. AES, XPS and XANES analyses.
Tribol Int 2001;34:52330.
Martin JM. Antiwear mechanisms of zinc dithiophosphate: a
chemical hardness approach. Tribol Lett 1999;6:18.
Fuller MLS, Kasrai M, Bancroft GM, Fyfe K, Tan KH. Solution
decomposition of zinc dialkyl dithiophosphate and its effect on
antiwear and thermal lm formation studied by X-ray absorption
spectroscopy. Tribol Int 1998;31(10):62744.
Bancroft GM, Kasrai M, Fuller M, Yin Z, Fyfe K, Tan KH.
Mechanisms of tribochemical lm formation: stability of tribo- and
thermally-generated ZDDP lms. Tribol Lett 1997;3:4751.
Varlot K, Kasrai M, Martin JM, Vacher B, Bancroft GM,
Yamaguchi ES, et al. Antiwear lm formation of neutral and basic
ZDDP: inuence of the reaction temperature and of the concentration. Tribol Lett 2000;8:916.
Belin M, Martin JM. Local order in surface ZDDP reaction lms
generated on engine parts, using EXAFS and related techniques.
Lubr Sci 1995;81.
Rudnick L. Lubricant additives: chemistry and applications. New
York: Marcel Dekker; 2003.
ATC. Lubricant additives and the environment. Document 49, 1993.
Mortier R. Gasoline engine lubricants, Chapter for institute of
petroleum book, 1999.
Mitchell PCH. Oil-soluble MoS compounds as lubricant additives.
Wear 1984;100:281300.
Lansdown AR. Molybdenum disulphide lubrication. Amsterdam:
Elsevier; 1999.
Yamamoto Y, Gondo S. Friction and wear characteristics of
molybdenum dithiocarbamate and molybdenum dithiophosphate.
Tribol Trans 1989;32(2):2517.
Davis FA, Eyre TS. The effect of a friction modier on piston ring
and cylinder bore friction and wear. Tribol Int 1990;23(3):16371.
Martin JM, Mogne TL, Grossiord C, Palermo T. Tribochemistry
of ZDDP and MoDDP chemisorbed lms. Tribol Lett 1996;2:
31326.
Graham J, Spikes H, Korcek S. The friction reducing properties of
molybdenum dialkyldithiocarbamate additives: Part Ifactors inuencing friction reduction. Tribol Trans 2001;44(4):62636.
Graham J, Spikes H, Jensen R. The friction reducing properties of
molybdenum dialkyldithiocarbamate additives: Part IIdurability of
friction reducing capability. Tribol Trans 2001;44(4):63747.
Martin JM, Mogne TL, Boehm M, Grossiord C. Tribochemistry in
the analytical UHV tribometer. Tribol Int 1999;32:61726.
Grossiord C, Martin JM, Mogne TL, Palermo T. In situ MoS2
formation and selective transfer from MoDTP lms. Surf Coat
Technol 1998;108109:3529.
Grossiord C, Varlot K, Martin JM, Mogne TL, Esnouf C, Inoue K.
MoS2 single sheet lubrication by molybdenum dithiocarbamate.
Tribol Int 1998;31(12):73743.
Miklozic KT, Graham J, Spikes H. Chemical and physical analysis of
reaction lms formed by molybdenum dialkyl-dithiocarbamate
friction modier additive using Raman and atomic force microscopy.
Tribol Lett 2001;11(2):7181.
Gondo S, Yamamoto Y. Mechanism of the surface lm formation of
molybdenum dithiocarbamate (MoDTC) and effect of rubbing
materials. Jpn J Tribol 1991;36(3):32333.
Morina A. Lubricant additive interactions, surface reactions and the
link to tribological performance in engines. Edinburgh: Heriot-Watt
University; 2005.
Yamamoto Y, Gondo S. On properties of surface lms formed with
molybdenum dithiocarbamate (MoDTC) under different conditions.
Jpn J Tribol 1991;36(3):30921.

ARTICLE IN PRESS
A. Neville et al. / Tribology International 40 (2007) 16801695
[70] Morina A, Neville A, Priest M, Green JH. ZDDP and MoDTC
interactions and their effect on tribological performance-tribolm
characteristics and its evolution. Tribol Lett 2006;24(3):24356.
[71] Morina A, Neville A, Green JH, Priest M. Additive/additive
interactions in boundary lubricationa study of lm formation
and tenacity. In: Proceedings of the 31st Leeds-Lyon symposium
on tribology, life cycle tribology: 2004, Leeds UK: Elsevier; 2004.
p. 75766.
[72] Isoyama H, Sakurai T. The lubricating mechanism of di-u-thiodithio-bis (diethyldithiocarbamate) dimolybdenum during extreme
pressure lubrication. Tribol Int 1974;7:15160.
[73] Forbes ES. The load-carrying action of organo-sulphur compoundsa review. Wear 1970;15:8796.
[74] Kano M, Yasuda Y. The effect of ZDDP and MoDTC additives on
friction properties of DLC-coated and steel cam follower in engine oil
/http://wwwoetgat/website/wtc2001cd/html/m-27-11-230-kanopdfS
(accessed on 22 June 2006) 2005.
[75] Barros-Bouchet MID, Martin JM, Le-Mogne T, Vacher B. Boundary
lubrication mechanisms of carbon coatings by MoDTC and ZDDP
additives. Tribol Int 2005;38(3):25764.
[76] Miyake S, Saito T, Yasuda Y, Okamoto Y, Kano M. Improvement
of boundary lubrication properties of diamond-like carbon (DLC)
lms due to metal addition. Tribol Int 2004;37(9):75161.
[77] Donnet C. Recent progress on the tribology of doped diamond-like
and carbon alloy coating. Surf Coat Technol 1998;100101:1806.
[78] Fontaine J, Donnet C, Grill A, LeMogne T. Tribochemistry between
hydrogen and diamond-like carbon lms. Surf Coat Technol
2001;146147:28691.
[79] Yasuda Y, Kano M, Mabuchi Y, Abou S. Diamond-like carbon
coatings for low-friction valve lifter. SAE paper 2003-01-1101, 2003.
[80] Mabuchi Y, Kano M, Hamada T, Yasuda Y. The development of
diamond-like carbon coated valvelifter for engine friction reduction.
In: Synopses of the international tribology conference, Kobe, 2005:A-5.
[81] Ban M, Ryoji M, Fuji S, Fujioka J. Tribological characteristics of Sicontaining diamond-like carbon lms under oil-lubrication. Wear
2002;253(34):3318.
[82] Rincon C, Zambrano G, Carvajal A, Prieto P, Galindo H, Mart nez
E, et al. Tungsten carbide/diamond-like carbon multilayer coatings
on steel for tribological applications. Surf Coat Technol 2001;
148(23):27783.
[83] Rosado L, Jain VK, Trivedi HK. The effect of diamond-like carbon
coatings on the rolling fatigue and wear of M50 steel. Wear
1997;212(1):16.
[84] Podgornik B. Coated machine elements-ction or reality? Surf Coat
Technol 2001;146147:31823.

1695

[85] Podgornik B, Hren D, Vizintin J. Low-friction behaviour of


boundary-lubricated diamond-like carbon coatings containing tungsten. Thin Solid Films 2005;476(1):92100.
[86] Podgornik B, Hren D, Vizintin J, Jacobson S, Stavlid N, Hogmark S.
Combination of DLC coatings and EP additives for improved
tribological behaviour of boundary lubricated surfaces. Wear 2006;
261(1):3240.
[87] Podgornik B, Vizintin J. Tribological reactions between oil additives
and DLC coatings for automotive applications. Surf Coat Technol
2005;200(56):19829.
[88] Kalin M, Vizintin J. Differences in the tribological mechanisms when
using non-doped, metal-doped (Ti, WC), and non-metal-doped (Si)
diamond-like carbon against steel under boundary lubrication, with
and without oil additives. Thin Solid Films 2006;515(4):273447.
[89] Miyanaga M, Oda K, Ohara H, Ikegaya A. Friction coefcient of Ticontaining DKC in engine oil. SEI technical review 2005 /http://
www.sei.co.jp/tr/t_technical_pdf/sei10370.pdfS (accessed online on
22 June 2006).
[90] Stallard J, Teer DG. A study of the tribological behaviour of CrN,
graphit-iC and dymon-iC coatings under oil lubrication. Surf Coat
Technol 2004:1889 (5259).
[91] Haque T, Morina A, Kapadia R, Arrowsmith S, Neville A. Nonferrous coating/lubricant interactions in tribological contacts: assessment of tribolms. Tribol Int 2006, accepted for publication.
[92] Tung SC, Gao H. Tribological characteristics and surface interaction
between piston ring coatings and a blend of energy-conserving oils
and ethanol fuels. Wear 2003;255(712):127685.
[93] Hamrock BJ, Dowson D. Ball bearing lubrication. In: Jacobson BO,
editor. The elastohydrodynamics of elliptical contacts. London:
Wiley; 1981.
[94] Kasrai M, Cutler JN, Gore K, Canning G, Bancroft GM, Tan KH.
The chemistry of antiwear lms generated by the combination of
ZDDP and MoDTC examined by X-ray absorption spectroscopy.
Tribol Trans 1998;41(1):6977.
[95] Muraki M, Yanagi Y, Sakaguchi K. Synergistic effect on frictional
characteristics under rollingsliding condition due to a combination
of molybdenum dialkyldithiocarbamate and zinc dialkyldithiophosphate. Tribol Int 1997;30(1):6975.
[96] Unnikrishnan R, Jain MC, Harinarayan AK, Mehta AK. Additiveadditive interaction: an XPS study of the effect of ZDDP on the
AW/EP characteristics of molybdenum based additives. Wear 2002;
252:2409.
[97] Minfray C, Martin JM, Esnouf C, Mogne TL, Kersting R,
Hagenhoff B. A multi-technique approach of tribolm characterisation. Thin Solid Films 2004;447:2727.

You might also like