2006 SCT 200 DLCandPolaritySaturationOfOILS Kalin

download 2006 SCT 200 DLCandPolaritySaturationOfOILS Kalin

of 8

Transcript of 2006 SCT 200 DLCandPolaritySaturationOfOILS Kalin

  • 7/31/2019 2006 SCT 200 DLCandPolaritySaturationOfOILS Kalin

    1/8

    The lubrication of DLC coatings with mineral and biodegradable oils

    having different polar and saturation characteristics

    M. Kalina,*, J. Viintina, K. Vercammenb, J. Barrigac, A. Arnxekd

    aUniversity of Ljubljana, Center for Tribology and Technical Diagnotics, Ljubljana, SloveniabVito, Mol, Belgium

    cTekniker, Eibar, SpaindPetrol, Ljubljana, Slovenia

    Received 19 January 2005; accepted in revised form 15 March 2005

    Available online 10 May 2005

    Abstract

    Due to improved performance over the last decade, diamond-like carbon (DLC) coatings are more frequently used in highly loaded

    mechanical components that sometimes need to operate under boundary- or mixed-lubrication conditions. However, DLC coatings are

    considered as inert coatings with a low surface energy and their lubrication ability according to conventional metal-lubrication mechanisms

    is therefore questionable. In order to investigate whether the base oil polarity and saturation characteristics play a role in these processes, a

    tribological investigation of the a-C:H coating lubricated with natural (sunflower) and synthetic (saturated and un-saturated) biodegradable

    oils that posses different amount of polar components and double bonds was performed. For a comparison we also used mineral base oil with

    low polar component. The DLC/DLC and steel/steel contacts were tested with base oils without additives and in combination with AW and

    EP additives. Despite the higher wear compared to steel/steel contacts, the results suggest that the wear of DLC/DLC contacts can be

    importantly improved by using oils with more polar groups and non-saturated molecules. These findings, together with well-known and even

    proverbial low-friction properties, suggest a great potential for the use of DLC coatings in combination with biodegradable oils,particularly when additives are present.

    D 2005 Elsevier B.V. All rights reserved.

    Keywords: DLC coating; Oil; Biodegradable; Lubrication; Polarity; Saturation; Wear; Friction

    1. Introduction

    Various hard coatings, including DLC, are nowadays

    used in many industrial areas. Over the past decade DLC

    coatings have shown a great potential for a broader range of

    uses due to their excellent properties under a variety ofconditions [16]. This improved performance suggests that

    DLC coatings could be suitable for heavily loaded

    mechanical components that operate under boundary- or

    mixed-lubrication conditions.

    Literature data show [7] that more than 10% of all the

    lubricants used in Europe are exposed to natural surround-

    ings, which increases the already high level of pollution in

    these areas. This can occur through the leakage, main-

    tenance, filtration of the systems or through deliberate

    spillage. Biodegradable oils are one of the possibilities to

    partially solve this problem in mechanical systems that are

    used in environmentally sensitive places, e.g., machinery inforests, agriculture, mining, construction. The common

    biodegradable oils are natural, i.e., rapeseed oil or sunflower

    oil, and synthetic esters. Natural biodegradable oils possess

    good anti-wear properties and low friction. However, their

    oxidation and thermal stability is poor, and this is their

    major drawback [8]. On the other hand, although synthetic

    esters are more resistant to oxidation and thermal degrada-

    tion, their tribological properties are not as good as those of

    natural esters [9]. This is especially true for saturated

    synthetic esters, while the properties of unsaturated syn-

    0257-8972/$ - see front matterD 2005 Elsevier B.V. All rights reserved.

    doi:10.1016/j.surfcoat.2005.03.016

    * Corresponding author. Tel.: +386 1 4771 462; fax: +386 1 4771 469.

    E-mail address: [email protected] (M. Kalin).

    Surface & Coatings Technology 200 (2006) 4515 4522

    www.elsevier.com/locate/surfcoat

  • 7/31/2019 2006 SCT 200 DLCandPolaritySaturationOfOILS Kalin

    2/8

    thetic esters are in between the two extreme situations. It

    could be said that the better the oxidation stability, the

    poorer the tribological properties. The properties that affect

    the oxidation stability most importantly are the number of

    reactive groups and the level of saturation of fatty acids in

    these oils, Fig. 1a. More double (or triple) bonds suggest

    easier oxidation of the oil. On the other hand, oils with non-saturated molecules and with more polar groups (like

    -COOH) posses also more sites for reactions and/or

    adsorption with metal surfaces that can provide boundary

    lubrication effects [10,11]. Namely, some of the conven-

    tional (metal) boundary lubrication mechanisms are due to

    the adsorption of oil/additive polar groups at the oxidized

    metal surface, Fig. 1b. This property thus influences the

    improved tribological behaviour of biodegradable oils as

    explained earlier. Accordingly, it is expected that higher

    number of polar groups improves the tribological perform-

    ance of conventional metal tribo-systems. While the clean

    mineral base oils contain no (or little) such polar groups, the

    natural biodegradable oils contain them in very high

    amounts. The relative ranking of the amount of polar

    groups in base oils used in our study (see Experimental) and

    expected properties in conventional tribological metal

    systems are presented in Fig. 2.Most metal-lubrication mechanisms are based on the

    physical or chemical adsorption of polar groups from the oil

    and/or additives onto the oxidised metal surface or via a

    chemical reaction between the additives and the reactive,

    clean metal surfaces under high-temperature and high-stress

    conditions [10,11]. So, here lies the major problem with the

    lubrication of DLC coatings, i.e., DLC coatings are

    considered as inert coatings with a low surface energy

    and their lubrication ability is therefore questionable. Some

    previous studies suggested that the tribological performance

    of DLC cannot be improved by lubrication and additives.

    Many efforts have been made in recent years to understandthe lubrication of DLC coatings as part of the European

    union research funding scheme known as the XGrowth?

    Programme [12]. The effects of various base oils, additives,

    coating modifications and operational conditions have been

    investigated using various testing geometries and scales

    [1316].

    In this work we have investigated the effects of polar

    characteristics and saturation of molecules of different base

    oils on the tribological behaviour of a-C:H coating under

    boundary lubrication conditions. Three types of biodegrad-

    able oils, i.e., natural (sunflower) and synthetic (saturated

    and un-saturated) biodegradable oils that posses different

    polar and saturation characteristics were used. For acomparison we also used a mineral base oil with low polar

    component. All the oils were tested without additives and in

    combination with AW and EP additives. To observe the

    pure lubrication effect due only to the properties of the

    selected DLC coating, the oils and the additives, and thus to

    avoid the expected interactions of the oils and additives with

    the steel in steel/DLC contacts [17], only self-mated DLC/

    DLC contacts are presented and discussed. As a reference, a

    conventional tribological system consisting of steel/steel

    contacts was also used.

    O

    C

    OH2C

    CO

    O

    HC

    CO

    OH2C

    Linoleic acid

    Oleic acid

    Stearic acid

    C

    OO

    C

    C

    C

    Me - O

    H H

    HH

    H H

    Polar group

    Metal

    Non-Polar tail

    a)

    b)

    Fig. 1. (a) Typical triglyceride composition (vegetable oil) with three

    different types of fatty acids having different level of saturation (number of

    double bonds). (b) A schematic of boundary lubrication mechanism on

    metal surface via adsorption of the polar group from oil or additive.

    Fig. 2. Schematic presentation of the effect of polar groups and unsaturation of molecules on tribological properties in metal contacts and oxidation stability of

    selected oils (see Experimental).

    M. Kalin et al. / Surface & Coatings Technology 200 (2006) 4515 45224516

  • 7/31/2019 2006 SCT 200 DLCandPolaritySaturationOfOILS Kalin

    3/8

    2. Experimental

    A single-layer pure amorphous a-C:H coating (without

    any doping elements) prepared using the RF PACVD (13.56

    MHz) process with a thickness of 1.78T0.09 Am was used

    in all the tests. A Si-based interlayer was employed to

    improve the adhesion properties of the coating. Theadhesion of the coatings was investigated with a scratch

    tester (Revetest, CSM Instruments SA, Switzerland) and the

    average Lc1, Lc2 and Lc3 values [18] were determined as

    9.4 N, 11.5 N and 15.6 N. The hardness and the Youngs

    modulus of the coating were measured using the depth-

    sensing indentation (DSI) technique (NanoTest 600 instru-

    ment with Berkovich indenter, Micro Materials Limited,

    UK). The hardness of the coating was 21.9 Gpa, and the

    Youngs modulus was 157 GPa. The surface roughness Raof the samples, i.e., the balls and discs, was measured after

    the coating deposition using the stylus-tip profilometer

    (T8000, Hommelwerke GmbH, Schwenningen, Germany)and the average value was 0.06 Am for the discs and 0.02

    Am for the balls. The coatings were deposited on the

    substrates from DIN 100Cr6 steel. All the steel balls and

    discs had a hardness of 850 HV (Leitz Miniload, Wild Leitz

    GmbH, D-6330 Wetzlar, Germany), and their roughness

    was negligibly lower than that described above. The steel

    balls were commercially available, standard ball bearings

    with a diameter of 10 mm. The steel flat samples were f24

    mm7.9 mm discs. Some of the flat and ball samples wereused as reference steel specimens in the tribological tests,

    while the rest of the discs and balls were further coated

    using the deposition technique described above. Table 1

    summarizes some of the most important material properties.In order to avoid effects of different counter-bodies

    (steel) or doping/alloying elements [13,17], only self-mated

    contacts were investigated. In the case of coated samples

    they are denoted as DLC/DLC. For comparison, steel/steel

    contacts were also investigated.

    Three different biodegradable base oils were used for this

    investigation, i.e., saturated synthetic ester, unsaturated

    synthetic ester and high oleic sunflower base oil, and as a

    reference, a paraffinic mineral base oil. They all had the

    same viscosity, corresponding to grade ISO VG 46. Two

    different additives that are used for conventional steel

    surfaces were selected. One of the additives was a multi-functional anti-wear/extreme-pressure (AW/EP) additive, a

    mixture of amine phosphates, having about 4.8% and 2.7%

    of P and N, respectively. The second additive used was a

    strong extreme-pressure (EP) additive, dialkyl dithiophos-

    phate ester, containing 9.3% of P and 19.8% of S. The

    characteristics for the oils and the additives are presented inTable 2.

    The wear experiments were performed in a reciprocating

    sliding machine. The lower, flat samples were fixed in the

    base, while the upper specimens, i.e., the balls, were fixed in

    the oscillating holder. In all the experiments, 10 N of normal

    load was applied through the loading system, which resulted

    in an initial average Hertzian contact stress of about 700

    MPa (1 GPa max). A stroke of 1 mm and an oscillating

    frequency of 50 Hz were used, resulting in a relative contact

    velocity of 0.1 m/s. In each test, the total sliding distance

    was 100 m. To ensure high severity of the contact the

    temperature was pre-set to 80 -C. Accordingly, the lambda

    (k) value (the Tallian parameter [19]) in all the experiments

    was less than 0.06, confirming the severe boundary-

    lubrication conditions: k

  • 7/31/2019 2006 SCT 200 DLCandPolaritySaturationOfOILS Kalin

    4/8

    ern). In addition, a limited number of samples were analysed

    with a FT-IR spectrometer equipped with a microscope and

    an ATR (Germanium crystal) objective (Spectrum One,

    Perkin Elmer Instruments). Also, the XPS (Mg Ka)

    measurements were performed in an ultrahigh-vacuum

    chamber with a base pressure of 11010 mbar, using a

    Perkin-Elmer ESCA/Auger spectrometer with a double-passcylindrical mirror analyzer.

    3. Results

    As a reference for the contact conditions, steel steel was

    used in combination with all the oils. The highest wear

    among the biodegradable base oils was found for the

    saturated synthetic ester, while the unsaturated ester and the

    sunflower base oil performed almost the same, and only

    negligibly better, see Fig. 3a. A substantial more than

    50%decrease in wear was observed for all the biodegrad-able oils when the additives were used. Also, the relative

    ranking between the different biodegradable oils was almost

    the same as for the base oils used without additives. The use

    of the AW/EP and EP additives resulted in about the same

    performance; however, just a little bit lower wear was

    always found with the AW/EP additive. The wear loss in the

    case of the mineral base oil was slightly lower when

    compared to the biodegradable base oils, while with the use

    of additives, the wear was similar to that of the biodegrad-

    able oils. However, the differences between the various

    types of oil are small.

    The results of the wear experiment for the self-mated

    DLC coatings are presented in Fig. 3b. The results are

    similar to those with the steel surfaces; however, thedifferences between the base oils are much more pro-

    nounced. Again, the highest wear among the biodegradable

    oils was found with the saturated ester, which was also

    higher compared to the wear in the steel steel contacts. The

    DLC contacts lubricated with the sunflower base oil

    experienced about 50% and 40% lower wear levels than

    the saturated and unsaturated synthetic esters, respectively.

    This wear was even 15% lower than in the case of the steel

    contacts. Additives reduced the wear of the DLC coatings

    significantly when used with the synthetic esters (more than

    30%), but their effect was reduced with sunflower oil, which

    showed low wear even without the additives. The wear ofthe DLC contacts with biodegradable oils with additives

    was always higher than in the case of the steel contacts. The

    mineral oil, which was used as a reference, provided the

    poorest wear protection for the coated surfaces when used

    without additives and with the AW/EP additives (Fig. 3b).

    However, when the EP additives were used, its performance

    was comparable to the other results with the DLC-coated

    surfaces.

    Fig. 4a presents the coefficient of friction obtained in the

    steel steel contacts. The highest coefficient of friction

    among the biodegradable oils was observed for the saturated

    synthetic ester, and the lowest was observed for the

    vegetable sunflower oil. Additives did not decrease, buttypically even increased the coefficient of friction. In

    general, mineral oil showed the highest values of the

    coefficient of friction with or without the additives. It is

    also clear that the use of AW/EP additives resulted in higher

    values of friction compared to the EP additives, irrespective

    of the type of oil used.

    The results of the coefficient-of-friction measurements

    for self-mated DLC contacts are presented in Fig. 4b. The

    values are remarkably lower (20 40%) than the steel

    contacts for all the conditions investigated. The highest

    friction was measured for the saturated oil, and the lowest

    for the sunflower base oil; however, the differences betweenthe unsaturated ester and the sunflower oil are almost

    negligible. The use of additives slightly increased the

    friction, but in contrast to the steel, the oils with the EP

    additives, rather than the AW/EP, resulted in higher levels of

    friction. Mineral base oil showed the lowest coefficient of

    friction, while the use of the EP additive provided the

    highest friction among all the experiments with coated

    surfaces.

    Fig. 5 shows the SEM images of the steel/steel contacts.

    The surfaces that were tested with saturated ester (Fig. 5a)

    and mineral oil (Fig. 5c) without additives show clear

    evidence of adhesive wear, with a layer formed by smeared

    a)

    b)

    0,0E+00

    1,0E-05

    2,0E-05

    3,0E-05

    4,0E-05

    5,0E-05

    6,0E-05

    7,0E-05

    8,0E-05

    Saturated ester Unsaturated

    ester

    Sunflower oil Mineral oil

    Wearloss(mm

    3)

    Base oil

    Base oil+AW/EP

    Base oil+EP

    0,0E+00

    1,0E-05

    2,0E-05

    3,0E-05

    4,0E-05

    5,0E-05

    6,0E-05

    7,0E-05

    8,0E-05

    Saturated ester Unsaturated

    ester

    Sunflower oil Mineral oil

    Wearloss(mm

    3)

    Base oil

    Base oil+AW/EP

    Base oil+EP

    Fig. 3. Wear loss of the ball in the (a) steel/steel and (b) DLC/DLC contacts

    as a function of the type of oil. Error bars representTone standard deviation

    of the measurements.

    M. Kalin et al. / Surface & Coatings Technology 200 (2006) 4515 45224518

  • 7/31/2019 2006 SCT 200 DLCandPolaritySaturationOfOILS Kalin

    5/8

    wear debris that subsequently delaminated and caused rather

    high wear under these conditions, see Fig 3a. A substantial

    amount of deformation, ploughing and delamination can be

    seen, in particular with the saturated ester, Fig. 5a. On the

    other hand, sunflower base oil (Fig. 5b) results in much less

    surface damage, there is no evidence of adhesion and the

    original sample preparation scratches can still be clearlyobserved. Despite this, the wear was comparable to the

    saturated and mineral base oils, see Fig. 3a. When additives

    were added to the oil, the adhesive wear was eliminated in

    all contacts (Fig. 5df), in accordance with the much lower

    wear loss in these contacts, see Fig. 3a. The surfaces tested

    with biodegradable oils with additives are similar and many

    original scratches are still visible. In accordance with the

    EDS analyses, which revealed traces of phosphorous on the

    surfaces (see arrows in Fig. 5df), small islands of such

    additive-reacted surfaces can be observed. However, when

    mineral oils with additives were used, the surfaces were

    much smoother and covered with a pronounced softtribochemical layera result of the reactions of additives

    with the steel surfaces and subsequent deformation of the

    layer. Obviously, quite different wear mechanisms result

    from the use of mineral or present biodegradable oils, even

    when employing the same additives. This was also observed

    for other coatings and material combinations [17].

    Fig. 6 shows SEM images of the DLC worn surfaces

    tested with different base oils and with the same oils

    combined with the EP additive. The two surfaces tested with

    biodegradable base oils again have a similar appearance, see

    Fig. 6a and b. Despite clearly measurable wear, these two

    surfaces present apparently no-wear conditions with the

    Fig. 5. SEM images of the steel surfaces tested with: (a) saturated ester, (b) sunflower oil, and (c) mineral oil. Images (d), (e), and (f) are surfaces tested with

    corresponding oils using the AW/EP additive. Arrows indicate locations of the EDS analyses.

    a)

    b)

    0,00

    0,05

    0,10

    0,15

    0,20

    0,25

    0,30

    0,35

    0,40

    0,45

    0,50

    Saturated ester Unsaturated

    ester

    Sunflower oil Mineral oil

    Coefficient

    offriction

    Base oil

    Base oil+AW/EP

    Base oil+EP

    0,00

    0,05

    0,10

    0,15

    0,20

    0,25

    0,30

    0,35

    0,40

    0,45

    0,50

    Saturated ester Unsaturated

    ester

    Sunflower oil Mineral oil

    Coefficientoffriction

    Base oil

    Base oil+AW/EP

    Base oil+EP

    Fig. 4. Coefficient of friction in the (a) steel/steel and (b) DLC/DLC

    contacts as a function of the type of oil. Error bars representTone standard

    deviation of the measurements.

    M. Kalin et al. / Surface & Coatings Technology 200 (2006) 4515 4522 4519

  • 7/31/2019 2006 SCT 200 DLCandPolaritySaturationOfOILS Kalin

    6/8

    original scratches from the sample preparation still visible,

    which suggests rather low wear, as seen in Fig. 3b. On the

    other hand, mineral base oil resulted in much smoother

    surfaces, indicating more intensive run-in and smoothening

    of the surfaces, see Fig. 6c. The SEM images are therefore

    in agreement with the wear data, showing noticeably higher

    wear for the mineral base oil, see Fig. 3b. In the tests when

    using the additives, careful inspection of the worn surfaces

    showed a quite different appearance and morphology from

    the original DLC surface, or from the worn surfaces whenusing only base oils. The worn surfaces with biodegradable

    oils (Fig. 6d and e) appear rougher than with the base oils,

    and are covered with a very thin and soft layer with many

    scratches. Some of the scratches are from the sample

    preparation, in agreement with the small amount of wear

    (Fig. 3b), but some are due to the sliding action and show

    signs of ploughing, indicating the soft nature of the surface.

    On the other hand, a quite different appearance of the

    surface can be found with the mineral oil with EP additive,

    see Fig. 6f. The surface is covered with a clear tribochemical

    layer that is very smooth and has an amorphous-like

    appearance. This is a common observation with softtribochemical layers [20,21].

    4. Discussion

    According to conventional lubrication mechanisms with

    steel surfaces based on physical and chemical adsorption,

    where anti-friction additives and fatty acids with their polar

    groups (Fig. 1) play a key role in interactions with the metal

    surfaces [10,11], the best tribological performance is

    expected for vegetable sunflower base oil (Fig. 2), which

    consists of a considerable amount of fatty acids with

    unsaturated bonds. Unsaturated synthetic esters have similar

    tribological characteristics, but are less favourable due to

    their modification for better oxidation stability. In contrast,

    saturated synthetic esters (and mineral oils) are much less

    prone to these interactions. Our wear data from the tests

    with steel surfaces (Fig. 3a) indicate this trend; however, the

    differences between the base oils are rather low. This might

    suggest that the severity of the contact conditions was

    already at the limits of effective lubrication protection when

    using only base oils, resulting in a low differentiation of thewear results. In addition, the evidence for adhesion and the

    formation of transfer films on some surfaces (Fig. 5a and c)

    confirm the above consideration and the poor wear

    protection under these conditions. Moreover, when using

    oils with additives the wear was significantly lower and the

    adhesion was eliminated (Fig. 5df). This was true for all

    types of oil, which clearly indicates that additives were

    predominantly responsible for the wear protection. The

    same behaviour was also observed for reference mineral oil.

    In all cases, mild AW/EP additives were slightly more

    effective than strong EP additives, and reactions with

    phosphorus were always detected by EDS, even with EPadditives.

    The coefficient-of-friction results are more indicative and

    support the better lubricity of the unsaturated synthetic ester,

    and especially that of the sunflower base oil compared to the

    saturated synthetic esters or (even more) compared to the

    mineral oil, which resulted in a noticeably higher friction in

    the steel/steel contacts, see Fig. 4a. On the other hand, for all

    the oils used, the friction was higher when additives were

    used, which explains the higher wear protection due to the

    stronger surface interactions between the additives and the

    steel with the higher shear strength of these bonds

    [10,22,23]. Indeed, the higher coefficient of friction with

    Fig. 6. SEM images of the DLC surfaces tested with base oils: (a) saturated ester, (b) sunflower oil, and (c) mineral oil. Images (d), (e), and (f) are surfaces

    tested with corresponding oils using the EP additive.

    M. Kalin et al. / Surface & Coatings Technology 200 (2006) 4515 45224520

  • 7/31/2019 2006 SCT 200 DLCandPolaritySaturationOfOILS Kalin

    7/8

    AW/EP additives coincides well with the lowest wear for all

    the oils used, see Fig. 3a. Despite this, the quite different

    appearance of the surfaces tested with biodegradable and

    mineral oil (Fig. 5df) suggests that different lubricating

    and wear mechanisms are acting in the contact, as was also

    observed by [22].

    In DLC/DLC contacts tested with sunflower base oil, thewear was clearly the lowest, i.e., about 30% compared to

    the saturated and unsaturated synthetic esters without

    additives and more than 50% compared to the mineral

    base oil, see Fig. 3b. The difference between the base oils

    that have very different chemical compositions was there-

    fore much more pronounced with the DLC than the steel

    surfaces. This suggests that the substantially large amount

    of polar and unsaturated molecules that are readily

    available in the oil and could interact with the surfaces

    (like for high oleic sunflower oil) is even more important

    for the tribological behaviour for the non-reactive

    surfaces like DLC coatings than for reactive steelsurfaces. The present results suggest a quite simple linear

    proportion for this relationship. Nevertheless, the wear of

    the DLC surfaces was higher than the steel/steel contacts,

    indicating that the shear strength of the adsorbed oil

    molecules on the DLC surfaces was lower than on the

    steel, resulting in poorer wear protection. As described

    previously, the low coefficient of friction is in agreement

    with this suggestion due to low interface shear strength, see

    Fig. 4b. In agreement with the adsorption mechanism and

    shear effect in DLC contacts is also the fact that mineral

    base oil showed the lowest coefficient of friction (the least

    polar base oil), while the use of the EP additive provided

    much higher friction, i.e., the highest among base oils used,due to additive effect [13]. Since the adhesion or any other

    catastrophic type of wear of DLC coatings is not

    observed (Fig. 6a c), the friction was rather low, in

    agreement with the well-known, and even proverbial,

    low-friction properties of DLC coatings.

    By using the additives, the wear of the DLC coatings was

    reduced, and, like with steel, it was very similar for all three

    types of biodegradable oils, see Fig. 3b. This again confirms

    the critical importance of the additives for the wear

    protection, and also with the DLC coatings [13]. A specific

    layer has formed in all cases on the DLC surfaces in the

    presence of EP additives; see Fig. 6df. Their appearancesuggests that it is most likely that the layers are softer than

    the original DLC surfaces and, obviously, they have formed

    in the contact as a result of an influence of the additives

    used. However, by employing the micro-FTIR and XPS

    analyses, no evidence that would indicate a clear chemical

    change at the surfaces can be revealed. Thus, based on these

    analyses it is presently not possible to explain the

    mechanisms of the observed changes and tribological

    performance. Primarily, it is not clear whether the chemical

    changes as a result of interaction between the additives and

    DLC coatings are occurring at the interface, but are smaller

    of detection limit of the techniques used, or, for example,

    the cleaning process before the analyses was not adequate

    and affected the results. Other physical-based phenomena,

    like the change in local viscosity and/or the local hot-spot

    temperatures [24] that could cause structural changes (i.e.,

    graphitisation), should also be explored more in detail.

    Because of the clear, empirically determined effect of the

    additives, further studies that would reveal the mechanismsare in progress; however, because of the complexity and

    large extent they are not included within the scope of this

    work.

    As with the steel/steel contacts, also with DLC/DLC

    contacts different mechanisms can be suggested for the

    biodegradable oils and the mineral oil [17]. Namely, in the

    case of the biodegradable oils, the surfaces appear less

    changed and covered only with a soft and very thin layer,

    which is just slightly rougher than the original surface (Fig.

    6d and e), while in the case of the mineral oil, a well-

    developed amorphous-like layer, which is very different

    from the original surfaces is formed, see Fig. 6f. The highwear, in particular with the mineral oil, can thus be

    explained by the formation of tribo-layers, which are

    typically soft and wear-protective [20,21,25], but in this

    case resulted in a higher removal rate than those on steel.

    Nevertheless, this wear was still low, in the range of 108

    mm3/Nm, corresponding to mild wear.

    A general observation in this work was that the wear of

    the steel/steel contacts was lower than the wear of the DLC/

    DLC contacts (Fig. 3a and b), which indicates that steel

    surfaces are indeed better suited for lubrication with the oils

    and additives that were used in this investigation than are

    the DLC coatings. However, with some other coatings the

    results could be more favourable for DLC coatings than inthis study [13]. Moreover, a 3050% reduction of wear due

    to the use of additives was found here for the pure DLC/

    DLC contacts for all types of oils (Fig. 3b), which is in

    agreement with our other findings, showing that the

    performance of self-mated coated contacts became compa-

    rable to that of steel/steel, if additives are used [13,17]. On

    the other hand, the coefficient of friction for all types of base

    oil, with and without additives, was up to 30% lower with

    the DLC contacts than with the steel surfaces, Fig. 4. The

    use of additives slightly increased the coefficient of friction;

    however, bearing in mind the substantial reduction in wear,

    this seems an acceptable compromise for the DLCcoatings.

    The results confirm good tribological performance of

    the biodegradable oils in combination with the DLC

    surfaces, especially in terms of the low coefficient of

    friction. This effect is proportional to the amount of polar

    groups and double bonds in their fatty acids, related to

    increased adsorption and shear strength at the interface

    [10,22,23]. On the contrary, the non-polar mineral oil

    performed the worse and caused different types of surface

    damage compared to ester-type biodegradable base oils. At

    the same time, the wear was in the range of 108 mm3/

    Nm, which is reasonably low wear, still corresponding to

    M. Kalin et al. / Surface & Coatings Technology 200 (2006) 4515 4522 4521

  • 7/31/2019 2006 SCT 200 DLCandPolaritySaturationOfOILS Kalin

    8/8

    mild wear. This suggests a good potential for a selected

    combination of oils and DLC coatings to perform well in

    conditions of industrial applications under rather severe

    boundary-lubrication conditions. Further analyses and field

    tests with different industrial systems are in progress; the

    first results [26] are very promising and support the above

    conclusions.

    5. Conclusions

    The wear of the steel/steel contacts was lower than that of

    the DLC/DLC contacts, which indicates that steel

    surfaces are indeed better suited for lubrication with

    conventional oils and additives than are the DLC

    coatings.

    On the contrary, the coefficient of friction for all types of

    base oil, with and without additives, was up to 30%

    lower with the DLC contacts than with the steel surfaces,which is related to their lower adsorption compared to

    that on steel surfaces and to low friction properties of

    DLC coatings.

    Large amounts of polar groups and unsaturated fatty

    acids, which are readily available in the lubricant, like in

    high oleic sunflower oil, substantially improve the

    efficiency of the base-oil lubrication of inert DLC

    surfaces. The results suggest a simple linear propor-

    tion for this relationship. The non-polar mineral base oil

    showed the worse tribological performance and caused

    different DLC coating damage compared to ester-type

    biodegradable base oils.

    The 30 50% reduction in wear of the DLC/DLCcontacts due to the use of additives with all biodegrad-

    able oils suggests the predominant effect of additives on

    wear behaviourin agreement with our previous find-

    ings based on mineral oil. However, different wear

    mechanisms were observed for the biodegradable and the

    mineral oils. However, the actual function of additives in

    these contacts remains unclear.

    The results suggest a good potential for the use of DLC

    coatings in combination with biodegradable oils under

    boundary-lubrication conditions, in particular when using

    oils with additives (AW- and EP-formulated).

    Acknowledgements

    This work was in part funded by the European

    Commission, within the Growth programme of the 5th EU

    Framework Project LUBRICOAT (G5RD-CT-2000-00410).

    The authors wish to thank all co-workers within this project

    for help with various aspects of this work, especially B.

    Stenbom and P. Nilsson (Volvo Technology AB, Sweden),

    E. Roman (CSIC, Spain), D. Neerinck (Bekaert Dymonics,

    Belgium) and F. Kopae (Center for Tribology and Technical

    Diagnostics, Slovenia).

    References

    [1] A. Erdemir, G.R. Fenske, J. Terry, P. Wilbur, Surf. Coat. Technol.

    94-94 (1997) 525.

    [2] A. Grill, Diamond Relat. Mater. 8 (1999) 428.

    [3] A.A. Voevodin, M.S. Donley, J.S. Zabinski, Surf. Coat. Technol. 92

    (1997) 42.

    [4] Y. Liu, A. Erdemir, E.I. Meletis, Surf. Coat. Technol. 82 (1996) 48.

    [5] K. Holmberg, A. Matthews, H. Ronkainen, A. Leyland, Surf. Coat.

    Technol. 100101 (1998) 1.

    [6] J. Fontaine, C. Donnet, A. Grill, T. LeMogne, Surf. Coat. Technol.

    146147 (2001) 286.

    [7] W.J. Bartz, Tribol. Int. 31 (1 3) (1998) 35.

    [8] A. Igartua, J. Aranzabe, J. Barriga, B. Rodriguez, Cost 516

    Tribology Symposium Proceedings, VTT, Espoo, ISBN: 951-38-

    4573-7, 1998, p. 135.

    [9] B. Kran, J. Viintin, in: J. Viintin, M. Kalin, K. Dohda, S. Jahanmir

    (Eds.), Tribology of Mechanical Systems: A Guide to Present and

    Future Technologies, ASME press, New York, 2004, p. 107.

    [10] R.M. Mortier, S.T. Orszulik, Chemistry and Technology of Lubricants,

    2nd edR, Blackie Academic and Professional, Glasgow, 1993.

    [11] G.W. Stachowiak, A.W. Batchelor, Engineering Tribology, Tribology

    series 24, Elsevier, Amsterdam, 1993.

    [12] 5th FP growth project, Lubricoat, Environmentally friendly lubricants

    and low friction coatings, A route towards sustainable products and

    production processes, 2001 2004 (G5RD-CT-2000-00410).

    [13] M. Kalin, J. Viintin, J. Barriga, K. Vercammen, K. Van Acker,

    A. Arnxek, Tribol. Lett. 17 (4) (2004) 679.

    [14] J. Barriga, M. Kalin, K. Van Acker, K. Vercammen, A. Ortega, M. San

    Martn, in: Proceedings of the 11th Nordic Symposium on Tribology,

    Tromso, Harstad, Hurtigruten, Norway, June 2004, NORDTRIB 2004,

    p. 508.

    [15] K. Vercammen, K. Van Acker, A. Vanhulsel, J. Barriga, A. Arnxek,

    M. Kalin, J. Meneve, Tribol. Int. 37 (2004) 983.

    [16] M. Kalin, J. Vizintin, Wear 259 (112 July) (2005).

    [17] M. Kalin, J. Viintin, in: Proceedings of the 11th Nordic Symposium

    on Tribology, Tromso, Harstad, Hurtigruten, Norway, June 2004,

    NORDTRIB 2004, p. 549.

    [18] European Standard EN1071-3:2000:E, Advanced technical ceramics

    methods of test for ceramic coatingsPart 3, CEN Management

    Centre, Stassartstraat 36, Brussels, Belgium, 2000.

    [19] T.E. Tallian, ASLE Trans. 10 (1967) 418.

    [20] J. Takadoum, Wear 170 (1993) 285.

    [21] M.G. Gee, N.M. Jennett, Wear 193 (1996) 133.

    [22] P. Waara, J. Hannu, T. Norrby, A. Byheden, Tribol. Int. 34 (2001) 547.

    [23] W. Zhan, Y. Song, T. Ren, W. Liu, Wear 256 (2004) 268.

    [24] M. Kalin, Mater. Sci. Eng., A 374 (2004) 390.

    [25] M. Kalin, S. Jahanmir, G. Drazic, J. Am. Ceram. Soc. 88 (2)

    (2005) 346.

    [26] J. Pezdirnik, M. Kalin, J. Viintin, in: J. Viintin, J. Bedenk, M. Kalin

    (Eds.), SLOTRIB 04: Proc. of the Conference on Fuels, Tribology

    and Ecology, Radenci, Slovenia, November 11 12, Slovenian Society

    for Tribology, Ljubljana, 2004, p. 213.

    M. Kalin et al. / Surface & Coatings Technology 200 (2006) 4515 45224522