You are on page 1of 8

Materials and Design 190 (2020) 108519

Contents lists available at ScienceDirect

Materials and Design

journal homepage: www.elsevier.com/locate/matdes

How to get noWear? – A new take on the design of in-situ formed high
performing low-friction tribofilms
B. Kohlhauser a,b,⁎, M.R. Ripoll b, H. Riedl a, C.M. Koller a, N. Koutna a, A. Amsüss c, H. Hutter c, G. Ramirez d,
C. Gachot b,e, A. Erdemir d, P.H. Mayrhofer a
a
Institute of Materials Science and Technology, TU Wien, A-1060 Vienna, Austria
b
AC2T research GmbH, 2700 Wiener Neustadt, Austria
c
Institute of Chemical Technologies and Analytics, TU Wien, A-1060 Vienna, Austria
d
Argonne National Laboratory, 60439, IL, United States of America
e
Institute for Engineering Design and Logistics Engineering, TU Wien, A-1060 Vienna, Austria

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• In-situ formation of WS2 tribofilms on


WN based coatings with S-containing
additives
• Influence of p, T and S-content on the
formation of WS2 via an intermediate
oxide step during the in-situ lubrication
• In-situ formed WS2 based tribofilms
lead to coefficients of friction of 0.05
with no noticeable wear after 100 h

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

Article history: Despite growing environmental concern, combustion engines are projected to stay the main technology of trans-
Received 29 July 2019 portation for many years, so will unpopular S- and P- containing additives in lubricating oils. Therefore, it is es-
Received in revised form 21 January 2020 sential to maximize efficiency and minimize the harmful impacts of these additives in the meantime. Coatings
Accepted 22 January 2020
based on Mo and W hold great promise as they can chemically interact with S-bearing additives and form low
Available online 23 January 2020
friction tribofilms like MoS2 and in particular WS2. Here we highlight the reaction pathways involving first the
Keywords:
formation of oxide intermediates, followed by the generation of a highly protective and low-shear transition
In-situ WS2 metal disulfide tribofilm on a tailored WN-based coating, affording coefficients of friction as low as 0.05 com-
Coating-lubricant interaction bined with an extraordinary wear protection (i.e., no measurable wear on sliding surfaces even after 100 h of test-
Tungsten nitride ing). High resolution TEM and Raman spectroscopy confirm the formation of WS2 sheets and hence the huge
Thin film potential of these coatings for lubricated surfaces.
Tribofilm © 2020 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://
creativecommons.org/licenses/by-nc-nd/4.0/).

⁎ Corresponding author at: Technische Universität Wien, Institute of Material Science and Technology, Getreidemarkt 9/E308, A-1060 Vienna, Austria.
E-mail address: bernhard.kohlhauser@tuwien.ac.at (B. Kohlhauser).

https://doi.org/10.1016/j.matdes.2020.108519
0264-1275/© 2020 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
2 B. Kohlhauser et al. / Materials and Design 190 (2020) 108519

1. Introduction the absence of dispersants (which, if applied, reduce the lubricating


effect of the nanoparticles [15,22]). On top of that, nanoparticles pose
Twenty-two percent, that is the shockingly low energy fraction of an potential health risks, leading us to our goal of generating TMD, partic-
internal combustion engine that actually moves our cars in relation to ularly WS2 in-situ (due to its better thermal stability and load bearing
the chemical energy we put into them in the form of fuel [1]. The energy ability compared to MoS2) [23]. Conventional lubricants have been
necessary to simply overcome the friction of engine and transmission developed to operate on ferrous surfaces and cannot reach their full po-
alone is roughly 17% [1]. Environmental concerns and the resulting leg- tential in combination with the increased application of coatings.
islative pressure (to reduce emissions) are mounting and pushing man- Pioneering studies demonstrated that a reaction between S from lubri-
ufacturers to offer more efficient engine designs. While it is unlikely, cant additives or H2S [24,25] and W or Mo containing surfaces is in prin-
that we will soon fully substitute internal combustion engines with ciple possible [26–28], but so far only modest improvements were
electric engines, the necessity for increased mobility is unabated. Even achieved or the tested surfaces were of lower technological relevance.
a complete substitution with electric engines is predicted to only cut In order to enable this in-situ formation on all types of materials, and
the losses due to friction approximately in half [2,3]. Two strategies at the same time protect the surfaces from wear and excessive con-
have been pursued in the control and reduction of friction and wear sumption, our approach relies on W containing hard coatings. W
on machine elements: (i) the optimization of lubricants and (ii) the im- containing coatings deposited by physical vapor deposition are a robust
provement of the contact interface by increasing the surface quality or solution that can be applied for the protection of heavy loaded
functionality, i.e. by surface texturing or by the application of coatings components.
like diamond like carbon (DLC) on selected component surfaces [4]. We envision a new take on much needed [29] coating–lubricant syn-
On the lubrication side, low viscosity and ultra-low viscosity lubri- ergies, tailoring the coatings to boost their interaction with commer-
cants (like 0W-16 or 0W-8) can reduce viscous losses, but at the ex- cially available additives rather than developing new additives at great
pense of wear protection due to much increased metal-to-metal expenses. The friction reduction and wear protection are to be achieved
contacts. To compensate for the reduced surface separation by such via the in-operando tribochemical reactions as presented in Fig. 1,
oils, and increase service life in general, additives have to provide which illustrates our concept of targeted tribofilm formation based on
wear protection. Early discoveries highlight that lubrication and wear most desirable solid lubricants like MoS2 or WS2. A benefit of this ap-
protection are not solely reliant on a hydrodynamic oil film between proach is that the formation of MoS2 or WS2 does not rely on the use
two counteracting surfaces. Under mixed friction and particularly of lubricant additives containing heavy metals, since the source of Mo
under boundary lubrication, thin reaction layers, called tribofilms, or W is a hard coating. To achieve these goals, we investigate the funda-
which originate from interactions between exposed surface material mental mechanisms of the in-situ formation of these solid lubricants in
and lubricant molecules or additives, avert any direct contact between rotating ball-on-disk tribological tests, as the point contact allows to
the components' surfaces [5]. In engines (with constant changes of the reach high contact pressures as they often arise in non-conformal con-
sliding speed and direction), the whole range of the Stribeck-curve is tacts, e.g. in geared wheels and rolling contact bearings. Applying this
utilized and modern cars' automatic start-stop mechanisms highlight new knowledge enables us to tailor a Mo or W containing coating for
the necessity to protect surfaces with a tribofilm in boundary lubrication an optimal in-situ lubrication. The S delivery system is based on an ex-
and even in the initial motion, when the stiction of direct surface contact treme pressure additive (40 at.-% S) in a PAO 8 baseoil (polyalphaolefin,
has to be overcome. a synthetic oil for engines, transmissions, gears and pulleys). As a reac-
The most common additive in lubrication is zinc tion partner, transition metals are popular dopants for commercial coat-
dialkyldithiophosphate (ZDDP) [6], which under the influence of ings, but especially Mo and W also form interesting hard nitride
shear stresses decomposes [7,8] and (due to the entropy-gain driven
tribochemical reaction) [9] forms protective phosphate based
tribofilms. With its high S and P content, ZDDP is under scrutiny for
environmental concerns and negative effects on a cars' after-
treatment catalysts and exhaust filters. Current oils are highly re-
fined to provide not only low friction and wear, but also to maintain
good viscosity over broad temperature ranges, as well as to inhibit
oxidation and corrosion at the same time [6]. A combination of
ZDDP and Mo and S containing additives, i.e. molybdenum dithiocar-
bamate (MoDTC), reduce friction based on the formation of MoS2
[10]. So far it is not possible to completely eliminate S and P from
our lubricants, as no direct substitutes for those additives are avail-
able. Wear protection of the future engines will rely on a complex
range of additive technologies [11]. Some of the most promising can-
didates include layered transition metal dichalcogenides (TMD) like
MoS2 or WS2. Their unique 2D structure consists of a layer of metal
atoms that is covalently sandwiched between two layers of chalco-
gen atoms, which can provide exceptional mechanical and tribolog-
ical performances in addition to impressive optical [12] and
electrical [13,14] properties. In bulk TMD structures, the individual
sandwich sheets are stacked on top of each other while being only
weakly bound by Van der Waals interaction, creating an easy shear
plane, ideal for solid lubrication. TMDs, in the form of nanoparticles
[15–17] (in lubricants) or coatings (unlubricated) have attracted
particular attention due to their chemical stability, which in compar-
ison to ZDDP reduces the negative effects of S and does not pose en-
vironmental concerns [18–21]. Fig. 1. The proposed tribochemical reaction is separated into three steps (and works for W
Nanoparticles based on TMDs are troubled with high production and Mo analogously). In the 3rd step, the initially amorphous WS2 undergoes a shear
costs, and once blended with lubricants, can agglomerate especially in stress induced self-organization [27,30,31].
B. Kohlhauser et al. / Materials and Design 190 (2020) 108519 3

coatings on their own. While Mo based systems proved crucial for surveyed with optical microscopy (Olympus STM6) and an optical
uncovering the mechanism behind the reaction, due to the superior tri- profilometer (Bruker, Contour GT).
bological properties of WS2 in applications, our efforts in tailoring coat-
ings are concentrated on tungsten nitride based hard coatings. 2.3. Surface analyses of the tribofilm

2. Methods The n-heptane-cleaned samples were investigated, with a confocal


Raman microscope (inVia Reflex, Renishaw) at a 50-fold magnification,
2.1. Deposition of the WNx coating directly after the tribological tests. The 633 nm laser power was reduced
to 6 mW, to avoid any heat induced reactions during the measurement.
In preparation for the coating process, polished AISI 5200 steel flats The Raman instrument was calibrated with an internal silicon standard,
were cleaned in an ultrasonic bath with Acetone and subsequently with and commercial MoS2 and WS2 powders (≥99%, Sigma-Aldrich) were
Ethanol. All coatings were prepared in a CemeCon coating chamber used as a reference. The spectra were averaged over three individual
with a twofold rotating sample holder (giving the coating a multilay- measurements and recorded in a range of 150–2200 cm−1. In order to
ered look due to shadowing effects). To eliminate surface contamina- investigate the formation of lubricating carbon films, some measure-
tion an Ar ion etching step was implemented directly before coating. ments reached up to 3200 cm−1.
The coatings were deposited from a W target (N99.5%) in a high- The SIMS measurements were performed on a TOF-SIMS5 (ION-TOF
power impulse magnetron sputtering (HIPIMS) process by applying GmbH) equipped with a bismuth liquid metal ion gun (LMIG, primary
4000 W (9 W/cm2). The deposition was carried out at constant 270 °C gun), a dual source column (DSC, secondary gun) and a low energy
and a total pressure of 0.4 Pa in an Ar/N2 mixture (130/55 scccm, flood gun.
respectively). The samples were measured in the high current bunched mode
The coating hardness was determined with a UMIS Nanoindenter, (HCBU) with approximately 16 nA target current in the non-pulsed
equipped with a Berkovich diamond tip. The high surface quality of state. This measurement mode provides the best mass resolution and
the coating allowed for direct measurements on the untreated as depos- maximum count rate. Bi+ 1 was used as the primary projectile. The region
ited coating. Indents were placed along a line with a distance of 20 μm of interest (ROI) was set to 500 ∗ 500 μm2 at a resolution of 512 ∗ 512 Px
and a load range of 3–45 mN. The indentation depths of individual in- for the overview measurements.
dents for evaluation were kept below 10% of the overall film thickness The measurements in positive and negative ion mode were realized
to avoid substrate interference [32]. The load-displacement curves with two different projectiles, dependent on the enhancement effect. A
were evaluated according to Oliver and Pharr [33]. To achieve a repre- 2 kV O+ 2
2 beam with a sputter crater of 1000 ∗ 1000 μm was utilized for
sentative hardness, three individual measurements with minimum 40 the detection of positive ions. For negative measurements, a 1 kV Cs
indents each were consolidated and evaluated. beam with a sputter crater of the same size was set.
The phase structures of the coating were investigated by x-ray dif- The measurements were performed in the non-interlaced mode
fraction with a Panalytical Empyrean diffractometer in the Bragg- with 1 acquisition frame, 1 s of sputtering and 0.5 s pause time. This sep-
Brentano configuration. The measurements were carried out with aration of the analysis and the sputter cycle leads to an improved S/N
Bragg-Brentano HD mirror, a 0.04 rad soller slit, a fixed 2° Ta anti scatter ratio and generally speaking often to a higher data quality.
slit and a divergence slit of ½°. The diffractograms where evaluated with The floodgun provided about 15 μA of low energy electrons to effec-
HighScore (Panalytical) and a PDF-4+ 2018 database (ICDD). tively prevent charging of the samples. This is highly recommended for
the used materials.
2.2. Ball on disk tribological test The pressure in the main chamber was approximately 3 ∗ 10−8 mbar
for the negative measurements. Due to the EI gas source for O+ 2
The tests were performed in a unidirectional rotating ball on disk sputtering, also the pressure of the main chamber was slightly increased
setup and were carried out with a Nanovea T50 Tribometer. All speci- to around 7 ∗ 10−8 mbar.
mens were ultrasonically cleaned in two steps with acetone and n- The lamellas for TEM investigations were prepared using a FEI
heptane. They were tested against a 9.5 mm diameter Si3N4 ball, to Quanta 200 3D DualBeam- focused ion beam (FIB). Specimens were
avoid any chemical contamination with other transition metals or any coated with 4 nm gold and subsequently protected with a thin Pt
catalytic effects of additional metal surfaces. The surface roughness layer in order to conserve the fragile tribofilm on the surface. Compared
(Ra) of the Si3N4 balls was 0.015 μm. The test temperature was 90 °C, to the W based coating, the tribofilm is very fragile and was difficult to
which was controlled with a radiant heater and the relative humidity conserve. Therefore, the final thinning of the lamella was achieved by
was measured at ≈40%. The contact was lubricated with a PAO 8 mix- a 20 min ion polishing procedure using a Gentlemill (TechnoorgLinda).
ture, whose S-content was adjusted between 0 and 2 at.-% with the ex- Transmission electron microscopy (TEM) investigations were per-
treme pressure additive Lubrizol LZ 5340 MW. 0.05 ml of oil were formed with FEI TECNAI F20 operated with an acceleration voltage of
sufficient to cover the surface of the samples. At a constant sliding 200 kV, which was equipped with an Apollo XLTW SDD EDAX detector
speed of 0.1 ms−1 (boundary lubrication), the elevated temperature for chemical analyses.
and a contact pressures of 0.5 GPa, the parameters mimic conditions
in common applications. The tests were carried out with 3 repetitions 2.4. DFT calculations on the in-situ formation
at different radii (0.5 mm apart), while the revolutions per minute
were adapted to allow for a constant sample speed and test distance The calculations were performed using the Density Functional The-
(360 m). ory (DFT) as implemented in the Vienna Ab initio Simulation package
The Mo and W samples were polished to a roughness of Ra = 20 nm, (VASP) [34,35] together with the projector augmented plane wave
while the coated samples were tested as deposited. The annealing of (PAW) pseudopotentials under the generalized gradient approximation
tungsten samples for the testing of a pre-oxidized surface was carried (GGA) [36] and the Perdew-Burke-Ernzerhof (PBE) exchange correla-
out in ambient air at 600 °C for 15 min and yielded an even WO3 ther- tion functional [37]. The plane-wave cutoff energy was always set to
mally grown oxide layer. 600 eV and the reciprocal space was sampled with Γ-centered
After the test, ball and disk were separated immediately and allowed Monkhorst-Pack meshes [38] equivalent to at least 25,000 k-points
to cool before the oil was removed with n-heptane. In order to conserve atoms. All systems were converged to about 10−3 eV/at.
the tribofilm, the samples were cleaned only by submersing without Table 1 provides structural description of the phases of WN, WC,
any additional agitating or ultrasonic procedures. All samples were WS2, and WO3 considered in our study. Equilibrium lattice parameters
4 B. Kohlhauser et al. / Materials and Design 190 (2020) 108519

of all systems were obtained by fitting the minimum of the energy vs. The addition of 2 at.-% S in the form of the sulphurized additive, drasti-
volume curve with the Murnaghan equation of state [39]. cally changes the tribological behavior, lowering the COF to just below
! 0.05 after a run-in period and contributing to a smoother and more sta-
0
B0 V ðV 0 =V ÞB0 B0 V 0 ble run. The corresponding Raman spectroscopy in Fig. 2b can link the
EðV Þ ¼ E0 þ 0 þ1 − 0 ð1Þ
B0 B00 −1 B0 −1 highly improved performance on Mo to the in-situ formation of the
solid lubricant MoS2 (see red line for 2 at.-% S).
In order to determine whether we can translate the results of a Mo
where the parameter E0 denotes the total energy of the system at
surface to a chemically similar W surface we conducted similar tests
equilibrium volume V0, while B0and B0′ are the corresponding bulk
with stepwise increased S content in the oil. While the friction is
modulus and its pressure derivative at the reference state of zero pres-
reduced to 0.1 by adding 0.6 to 2.0 at.-% S additive (no trend within
sure. In order to analyze the pressure effect on the stabilities of WN, WC,
this concentration-window), it does not reach the low friction level
WS2, and WO3 polymorphs, we calculate their formation enthalpies
we would expect from a WS2 lubricated contact. As evidenced by
! Raman spectroscopy (Fig. 2b) the W surface with its low chemical reac-
1 X
H f ðpÞ ¼ H tot ðpÞ− ns μ s ð2Þ tivity (compared to Mo) does not facilitate the in-situ formation of
∑s ns x dichalcogenides. Interestingly, X-ray photoelectron spectroscopy re-
vealed that the Mo contact contained 37 at.-% oxygen from metal oxides
where Htot(p) is the enthalpy of the system, ns and μs are the number of in the contact zone. For W, in comparison, the test conditions are not se-
atoms and the chemical potential, respectively, of a species s. The refer- vere enough for oxidative attack on the metallic surface (1 at.-% oxygen
ence chemical potentials for C, S, and W are conventionally set to the en- from metal oxides). The absence of oxides in the W contact (while an
thalpy per atom of graphite-C, μC, orth-S, μS, and bcc-W, μW, while the N2 oxide layer is present on the undisturbed sample surface) lead us to be-
and O2 molecules, represent the chemical potential of N, μN and O, μO. lieve that the in-situ formation of both, MoS2 and WS2, is dominated by
The total enthalpy in the above Eq. (2) reads an oxidation pathway, in which the transition metal is first oxidized in a
first step (activation, Fig. 1) and the oxide subsequently reacts with the
H tot ðpÞ ¼ EðpÞ þ pV ðpÞ ð3Þ
S of the additive to form the respective disulfide (step 2, Fig. 1). A repe-
tition of the tests on Mo surfaces in an O2-free dry nitrogen atmosphere
Using the definition of pressure, p = ∂E/∂V, formulas for V(p) and E (Fig. 1a) supports this suggested formation pathway, as the friction is
(p) can be derrived from Eq. (1) and plugged in Eq. (4), therefore, significantly higher.
0 1 For verification of the reaction pathway, another set of W samples
B0 V 0 B ðp=B0 Þ þ 1 C V0 have been thermally annealed in air prior to testing. They provide a hex-
H tot ðpÞ ¼ E0 þ 0 @ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi −1A þ p qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð4Þ
B0 −1 B0′ B0 =B p þ 1 B0′
 0
B =B p þ 1
 agonal WO3 (P6/mmm) surface with an up to 500 nm thick oxide scale
0 0 0 0
and were tested under the same conditions as the metallic samples. The
COF for the oxidized W surface (Fig. 2a) slightly decreases as the S con-
Since the reference chemical potentials for C, S, and W are related to tent in the oil increases and shows a drastic drop to 0.05 as a certain
the enthalpies of their. threshold concentration is exceeded and in-situ WS2 formation was
crystalline phases which do depend on pressure through the enabled (confirmed by Raman spectroscopy in Fig. 2b). The Raman in-
Murnaghan Eq. (1), we also consider μC, μS, and μW in Eq. (2) to vary vestigations in both cases (Mo and W) also demonstrate that the lubri-
with pressure, analogically to the binary systems. Following the argu- cating effect is solely based on the formation of TMD and no oil derived
mentation of Reuter and Scheffler [40], the μN and μO chemical poten- lubricating carbon structures [47] are present. While the COF on the Mo
tials are kept constant irrespectively of the pressure, due to the fact sample experiences an exponential decrease due to an initial shortage
that our DFT calculations are performed at 0 K. of oxides, the pre-applied oxide educts are at instantaneous disposal
on the annealed W surfaces, reducing the running in period to only a
2.5. Following breadcrumbs towards in-situ solid lubrication few cycles. The relationship between the availability of oxides and S is
also highlighted by the run in period with 0.6 at.-% S and the significant
Our investigations into the mechanism behind the in-situ formation rise in friction at the end of the test with 2 at.-% S. In both cases WS2 was
started with unidirectional ball-on-disk tests (contact pressure: initially formed, but the unbalanced availability of educts (low S content
0.5 GPa) of polished commercial Mo samples at 100 °C. To avoid any in the oil or thin oxide layer on the oxidized W samples) lead to depri-
chemical interactions or analytical interferences of the counter body, vation of one reaction partner and resulted in the termination of a
all tests were performed with inert Si3N4 balls. Without the S- sustained in-situ lubrication. Therefore, the in-situ lubrication relies
containing additive (pure PAO 8), the coefficient of friction (COF) on on a competing process whereby the oxidation rate of the transition
the Mo sample in Fig. 2a, starts at around 0.18, slightly decreases to metal source and the availability of S need to be well balanced.
0.14 in the run-in phase but stays well above the steel flat (AISI For any application, low friction on its own is not the whole success
52100) vs. Si3N4 ball reference (same COF for all S concentrations). story and might be easily achieved otherwise. The true value of WS2 lies

Table 1
Crystal symmetry classes and space groups of the here considered phases of WN, WC, WS2, and WO3 together with the calculated lattice parameters, a and c.

System Symmetry Space group a(Å) aref(Å) c (Å) cref(Å)

WN Cubic Fm3m 4.369 4.357 [41], 4.362 [42]


WN Cubic Pm3m 4.133 4.120 [43], 4.125 [41]
WN Hexagonal P6m2 2.877 2.865 [42] 2.912 2.912 [41]
WC Cubic Fm3m 4.394 4.398 [42]
WC Hexagonal P6m2 2.931 2.926 [42] 2.853 2.849 [42]
WS2 Hexagonal P63/mmc 3.201 3.153 [44] 14.349 12.323 [44]
WS2 Trigonal R3m 3.197 3.158 [44] 21.880 18.49 [44]
WO3 Cubic Pm3m 3.840 3.84 [45]
WO3 Hexagonal P6/mmm 7.475 7.298 [46] 3.840 3.899 [46]
WO3 Tetragonal P6/nmm 5.368 5.36 [45] 3.996 4.01 [45]
B. Kohlhauser et al. / Materials and Design 190 (2020) 108519 5

Fig. 3. (a) The in-situ formation of MoS2 offers excellent wear protection (e). (b) Tests on
W surfaces lead to the formation of abrasive particles and (f) result in high wear on the
ball. (c) A similar wear mechanism can be noticed on oxidized W surfaces without the
presence of S. (d) The addition of 1.2 at.-% S to the contact sustains the formation of
WS2, leading to the same polishing as with Mo, averting any measurable wear (g).
Fig. 2. (a) Interactions of the Mo surface with 2 at.-% S in the oil lead to a drastic reduction
of the COF. While metallic W surfaces do not follow this low friction behavior, a thermally
pressure initially demonstrates similar low friction behavior, but after
annealed tungsten oxide surface demonstrate an immediate drop in friction at sufficient S
concentrations in the lubricant. (b) The low COF for S-containing tests on Mo and oxidized the run in period the COF starts rising again until it levels out at 0.08.
W can be directly linked to the formation of MoS2 and WS2 respectively, via Raman
spectroscopy of the wear scar on the counter body while on W no sulfide is formed.

in its combination with excellent wear protection. As the tested bulk


materials possess different mechanical properties, compatibility of the
wear behavior is limited to investigations of the ball as counter-body.
Tests without the S containing additive (as well as tests in which no
TMD formation was sustained due to low S or O content), present
high abrasive wear on the ball (Fig. 3b,c,f). The systems involving a
Mo surface (Fig. 3a,e) or an oxidized W surface (Fig. 3d,g) in combina-
tion with the S additive on the other hand only experienced slight
changes in the surface morphology (noticeable shining finish), without
any measurable wear. This demonstrates the excellent wear protection
properties of the in-situ formed MoS2 and WS2 tribofilm.

2.6. Tungsten nitride coatings – the key to in-situ WS2 formation

While MoS2 was easily formed on metallic Mo surfaces, the generally


more stable and better performing WS2 needs additional modification
of the W-delivery system to work. Therefore, tungsten nitride based
coatings were applied to modify the surface oxidation rate (as the
Gibbs free energy is less negative) and at the same time to benefit
from synergetic protection of the surface by enhanced mechanical prop-
erties of the coating. This coating was produced with high power im-
pulse magnetron sputtering, leading to a nano-crystalline morphology.
It obtains a hexagonal sub-stoichiometric δ-WN structure (P63/mmc)
and exhibits a hardness of 31 GPa (compared to 22 GPa of a metallic
Fig. 4. (a) Low friction via the in-situ generation of WS2 can also be achieved by applying
W coating). Further coating details are given in the supplemental coatings in the presence of S containing additives. The evolution of the COF depends on the
information. contact pressure, as the tribofilm suffers from higher loads. Raman spectroscopy
Putting the coating to the (tribological) test, Fig. 4a shows that, un- (b) demonstrates that WS2 has formed in higher amounts as the contact pressure
like metallic W, the coating supports the in-situ formation of WS2 increases. (c) Without any additive to react and form the WS2 based tribofilm, the wear
is quite severe and increases proportionally. (d) Although, the COF increases for higher
(under the same conditions). Compared to a COF of 0.1 without the contact pressures (despite the presence of the S containing additive), the wear
presence of S the COF with 2 at.-% S in the lubricant is lowered to a stable protection capabilities are still excellent pushing the amount of wear below detectable
level between 0.04 and 0.05. Increasing the load and thereby the contact levels.
6 B. Kohlhauser et al. / Materials and Design 190 (2020) 108519

Raman spectroscopy of the contact area, Fig. 4b, suggests that this is the required tungsten oxides and the conditions are severe enough to facil-
result of the WS2 tribofilm being worn away at higher contact pressures. itate the subsequent WS2 formation.
The tribofilm is increasingly mechanically removed from the contact The lubricating and protective effect of WS2 requires the formation
while the reaction constantly replenishes the WS2 on the exposed sur- of easy slide planes, more precisely crystalline WS2 sheets, emerging
face. This leads to an accumulation of WS2 based debris around the from the initially unaligned or amorphous reaction products. The
wear track, delivering a stronger Raman signal. major fraction of the tribofilm has an amorphous appearance in TEM
Without the S-additive, the contact is subjected to high wear of the (Fig. 6a) and selected area diffraction (SAED). Detailed high-resolution
ball (increasing with the contact pressure, Fig. 4c), whereas any coating (HR) studies (Fig. 6b) reveal that the lubrication, as envisioned, is in
wear is below the detection limit. In presence of 2 at.-% S in the contact, fact supported by the formation of crystalline WS2 regions. These
tests that exhibit low friction also experience no wear. Unexpectedly, nanocrystallites are only few nanometers in diameter and are roughly
despite the elevated COF of 0.08 (at this level tests with metallic W ex- aligned parallel to the interface. These regions are too small for SAED,
perienced high wear on the ball), the increased contact pressure does but the spacing between the planes can be calculated (by Fourier trans-
not lead to any measurable wear on coating or ball. The 3D images of formation of HR image sections) and matches the distance between the
the associated balls, Fig. 4d, prove that although the removal of the easy shear planes of WS2. Although the major part of the tribofilm is not
tribofilm from the contact might hinder excellent lubricating properties crystalline, it still contains W and S in the ratio 1:2 according to EDX line
at high loads, the wear protection of the in-situ formed WS2 is still in- scans (Fig. 6c), ready to self-organize under shear forces in order to form
tact, preventing any measurable wear. The wear protection of the a few nm thick easy-glide top-layer on the tribofilm.
tribofilm and the coating at a contact pressure of 0.5 GPa also proves The in-situ reaction undoubtedly produces WS2, but it is unexpected
to be very reliable and extremely efficient in long term tests, exhibiting that oxides are the key trigger to the formation. Considering normal at-
no measurable wear on either ball or coating after 100 h (or 36 km) of mospheric conditions, we would imagine the metal oxide to be the most
testing (see supplemental data). stable compound. However, in tribological contacts the conditions tend
The Si3N4 ball (Fig. 5a) presents a small contact interface with a sur-
rounding rim of relatively thick tribofilm at areas of lower contact pres-
sure. Around the contact, accumulated debris (containing WS2 and oil
derived by-products) adheres to the ball surface. These deposits of de-
bris also form outside of the wear track on the coated flat sample
(Fig. 5b), whereas the tribofilm itself is predominantly developing on
the ball. Although no wear was measurable, the increased load leads
to a rising buildup of debris as more tribofilm is worn away and contin-
uously regenerated. This build-up also supports the concept of the in-
situ formation of WS2 as secondary ion mass spectroscopy investiga-
tions reveal that the tribofilm debris is rich in S (Fig. 5c) while being
oxygen-depleted (in respect of the surrounding surface oxidation,
Fig. 5d). The investigations also uncovered the WS2 generation area in
the center of the wear track where local oxidation generates the

Fig. 6. (a) The tribofilm (protected by an Au and Pt layer), that formed on the Si3N4 ball,
Fig. 5. (a) The contact area of a ball at 0.9 GPa with 2 at.-% S exhibits no measurable wear was investigated with TEM and exhibits a mostly amorphous appearance. (b) In the
and presents areas of accumulated debris (investigated by Raman spectroscopy, contains tribofilm crystalline WS2 nanoparticles are embedded. (c) The chemical composition
WS2) and tribofilm, that has been investigated with TEM. (b) The coated flat does not (determined by EDX) across the tribofilm is dominated by amorphous carbon (from
show any measurable wear as well, but illustrates the formation of a debris zone around decomposed oil) and W as well as S in a 1:2 ratio. (d) The in-situ formation is also
the wear track (intensifying with the contact pressure). This area was investigated with suggested by density functional theory calculations, which suggest that at increasing
secondary ion mass spectroscopy (c) revealing a high concentration of S in the debris, contact pressure the WS2 becomes energetically more favorable than WO3 which has
while (d) oxygen species are absent from the debris. formed in the contact.
B. Kohlhauser et al. / Materials and Design 190 (2020) 108519 7

to be more extreme and in order to truly understand a tribochemical lead the simulation approach and performed all the first-principles cal-
reaction we need to investigate behavior of materials under these culations. B.K. composed the manuscript with input from all authors.
borderline conditions. From an energetic point of view, the formation
of WS2, with a standard Gibbs energy ΔfGom(293.15 K) of −232.1 ± CRediT authorship contribution statement
3.1 kJ·mol−1, is feasible [48] but in experiments it seems to be kineti-
cally inhibited. While an oxidative environment in tribocontacts is B. Kohlhauser: Project administration, Investigation, Data curation,
common, it is surprising that stable oxides in combination with S com- Visualization, Writing - original draft. M.R. Ripoll: Conceptualization,
pounds react further to form a less stable disulfide. First of all, the energy Investigation, Writing - review & editing. H. Riedl: Investigation, Data
of formation for WO3 increases with temperature until it becomes pos- curation, Writing - review & editing. C.M. Koller: Investigation, Data
itive and decomposes. The free Gibbs energies for potential variations of curation. N. Koutna: Investigation. A. Amsüss: Investigation, Data
in-situ reaction mechanism on the other hand decrease with rising tem- curation. H. Hutter: Investigation, Data curation. G. Ramirez: Investiga-
perature. Only the thermal decomposition of WS2 hinders a pure ther- tion, Formal analysis. C. Gachot: Supervision, Formal analysis, Writing -
modynamic explanation. Neglecting these phase transformations, high review & editing. A. Erdemir: Supervision, Formal analysis, Writing -
temperatures would energetically favor the WS2 formation. As reported review & editing. P.H. Mayrhofer: Supervision, Writing - review &
by Polcar et al., our experimental data confirms that a sufficiently high editing.
contact pressure is a major condition for the formation of crystalline
WS2 [30,31]. Therefore, we studied the influence of pressure on the for-
mation enthalpy with ab-initio density functional theory calculations Declaration of competing interest
(Fig. 6d, details in the supplemental information). Much like high tem-
peratures, the high pressure in single asperity contacts destabilizes The authors declare that they have no known competing financial
tungsten oxides and in combination with local temperature bursts tips interests or personal relationships that could have appeared to influ-
the scale in favor of the WS2 formation. ence the work reported in this paper.

3. Conclusion Acknowledgement

The formation of WS2 is a delicately balanced system between the This work was funded by the Austrian COMET-Program (Project K2
X-Tribology, Grant no. 849109) and has been carried out within the
availability of S and the oxidation rate of the W-supply, as the reaction
predominantly takes the path of an initial oxidation and subsequent sul- TU Wien (Austria), Argonne National Laboratory (USA) and the Excel-
lence Centre for Tribology AC2T (Austria). SEM and TEM investigations
fide generation. Both, temperature and pressure, in the contact region of
single asperities are significantly higher than in the rest of the contact. were carried out using the USTEM facilities at the TU Wien, Austria.
Thanks is also due to the X-ray center (XRC) and the Institute of Chem-
These conditions enable the formation of WS2 in the center of the
wear track, which under these conditions becomes energetically more ical Technologies and Analytics at the TU Wien, Austria, for evaluating
structure and chemical compositions. The authors acknowledge the
favorable than the oxide.
WN coatings provide an important benefit to loaded surfaces, as TU Wien Bibliothek for financial support through its Open Access
Funding Program.
they are applicable on all technologically relevant materials, add in-
creased wear protection, and unlike metallic W facilitate the formation
of WS2 in presence of S-containing lubricant additives. Our approach is Data availability
compatible with current lubricant formulations since they readily con-
tain additives with S and avoids the need of organic lubricant additives The data that support the findings of this study are available from
containing heavy metals. the corresponding author upon reasonable request.
The WS2 is incorporated in the tribofilm on the counter face (ball) of
the coating and self organizes under shear stress, forming lubricating Appendix A. Supplementary data
nanocrystals. These nanocrystalline regions are reducing the COF signif-
icantly and offer excellent wear protection, with no measurable wear on Supplementary data to this article can be found online at https://doi.
either ball or disk after 100 h. org/10.1016/j.matdes.2020.108519.
While the tribofilm is continuously worn away under higher loads,
leading to a COF only slightly below the reference, the formation of References
new WS2 in the generation area is sufficient enough to sustain the
[1] K. Holmberg, P. Andersson, A. Erdemir, Global energy consumption due to friction in
wear protection. The in-situ formation of WS2 allows the harmonization passenger cars, Tribol. Int. 47 (2012) 221–234.
of coatings and lubricants, and offers a synergetic and novel approach [2] World Energy Council, Global Transport Scenarios 2050, 2011.
on long term wear protection. [3] S. Chu, A. Majumdar, Opportunities and challenges for a sustainable energy future,
Nature 488 (2012) 294.
[4] A. Erdemir, C. Donnet, Tribology of diamond-like carbon films: recent progress and
Contributions future prospects, J. Phys. D. Appl. Phys. 39 (18) (2006) R311.
[5] F.P. Bowden, J.N. Gregory, D. Tabor, Lubrication of metal surfaces by fatty acids, Na-
ture 156 (1945) 97.
M.R.R. conceived the idea. A.E., C.G. and P.H.M. directed the research [6] H. Spikes, The history and mechanisms of ZDDP, Tribol. Lett. 17 (3) (2004) 469–489.
project. B.K. is the principal investigator and together with G.R. pro- [7] E.S. Ferrari, K.J. Roberts, D. Adams, A multi-edge X-ray absorption spectroscopy
duced and optimized the coatings. B.K. and H.R. carried out the XRD as study of the reactivity of zinc di-alkyl-di-thiophosphates (ZDDPS) anti-wear addi-
tives:: 1. An examination of representative model compounds, Wear 236 (1)
well as nanoindentation measurements and analyzed the data resulting (1999) 246–258.
from those techniques. M.R.R. and B.K. carried out XPS analyses. G.R. and [8] M.L.S. Fuller, M. Kasrai, G.M. Bancroft, K. Fyfe, K.H. Tan, Solution decomposition of
B.K. carried out the tribological experiments, wear investigations and zinc dialkyl dithiophosphate and its effect on antiwear and thermal film formation
studied by X-ray absorption spectroscopy, Tribol. Int. 31 (10) (1998) 627–644.
Raman spectroscopy measurements. G.R., A.E., C.G. and B.K. analyzed
[9] J.M. Martin, T. Onodera, C. Minfray, F. Dassenoy, A. Miyamoto, The origin of anti-
the data obtained in the tribological characterization. H.H. and A.A. car- wear chemistry of ZDDP, Faraday Discuss. 156 (2012) 311–323.
ried out all SIMS measurements and the subsequent evaluation of re- [10] A. Morina, A. Neville, Understanding the composition and low friction tribofilm for-
sults. B.K. prepared the FIB lamellae. C.M.K. carried out the HRTEM as mation/removal in boundary lubrication, Tribol. Int. 40 (10–12 SPEC. ISS) (2007)
1696–1704.
well as the analysis of the EDX data. C.M.K, B.K. and H.R. performed [11] H. Spikes, Low- and zero-sulphated ash, phosphorus and sulphur anti-wear addi-
the phase analyses of HRTEM images and diffraction patterns. N.K. tives for engine oils, Lubr. Sci. 20 (2) (2008) 103–136.
8 B. Kohlhauser et al. / Materials and Design 190 (2020) 108519

[12] L.C. Flatten, Z. He, D.M. Coles, A.A.P. Trichet, A.W. Powell, R.A. Taylor, J.H. Warner, [30] T. Polcar, A. Cavaleiro, Self-adaptive low friction coatings based on transition metal
J.M. Smith, Room-temperature exciton-polaritons with two-dimensional WS2, Sci. dichalcogenides, Thin Solid Films 519 (12) (2011) 4037–4044.
Rep. 6 (2016), 33134. [31] P. Nicolini, R. Capozza, P. Restuccia, T. Polcar, Structural ordering of molybdenum di-
[13] X. Hong, J. Kim, S.-F. Shi, Y. Zhang, C. Jin, Y. Sun, S. Tongay, J. Wu, Y. Zhang, F. Wang, sulfide studied via reactive molecular dynamics simulations, ACS Appl. Mater. Inter-
Ultrafast charge transfer in atomically thin MoS2/WS2 heterostructures, Nat. faces 10 (10) (2018) 8937–8946.
Nanotechnol. 9 (2014) 682. [32] A.C. Fischer-Cripps, Critical review of analysis and interpretation of nanoindentation
[14] A. Raja, A. Chaves, J. Yu, G. Arefe, H.M. Hill, A.F. Rigosi, T.C. Berkelbach, P. Nagler, C. test data, Surf. Coat. Technol. 200 (14–15) (2006) 4153–4165.
Schüller, T. Korn, C. Nuckolls, J. Hone, L.E. Brus, T.F. Heinz, D.R. Reichman, A. [33] G.M. Pharr, W.C. Oliver, Measurement of thin film mechanical properties using
Chernikov, Coulomb engineering of the bandgap and excitons in two-dimensional nanoindentation, MRS Bull. 17 (7) (1992) 28–33.
materials, Nat. Commun. 8 (2017), 15251. [34] G. Kresse, D. Joubert, From ultrasoft pseudopotentials to the projector augmented-
[15] J. Kogovšek, M. Kalin, Various MoS2-, WS2- and C-based micro- and nanoparticles in wave method, Phys. Rev. B 59 (3) (1999) 1758–1775.
boundary lubrication, Tribol. Lett. 53 (3) (2014) 585–597. [35] G. Kresse, J. Furthmüller, Efficient iterative schemes for ab initio total-energy
[16] A. Tomala, B. Vengudusamy, M. Rodríguez Ripoll, A. Naveira Suarez, M. Remškar, R. calculations using a plane-wave basis set, Phys. Rev. B 54 (16) (1996)
Rosentsveig, Interaction between selected MoS2 nanoparticles and ZDDP tribofilms, 11169–11186.
Tribol. Lett. 59 (1) (2015) 26. [36] W. Kohn, L.J. Sham, Self-consistent equations including exchange and correlation ef-
[17] A. Tomala, M.R. Ripoll, C. Gabler, M. Remškar, M. Kalin, Interactions between MoS2 fects, Phys. Rev. 140 (4A) (1965) A1133–A1138.
nanotubes and conventional additives in model oils, Tribol. Int. 110 (2017) 140–150. [37] J.P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made sim-
[18] V.B. Niste, M. Ratoi, Tungsten dichalcogenide lubricant nanoadditives for demanding ple, Phys. Rev. Lett. 77 (18) (1996) 3865–3868.
applications, Materials Today Communications 8 (2016) 1–11. [38] H.J. Monkhorst, J.D. Pack, Special points for Brillouin-zone integrations, Phys. Rev. B
[19] M. Ratoi, V.B. Niste, J. Zekonyte, WS2 nanoparticles - potential replacement for ZDDP 13 (12) (1976) 5188–5192.
and friction modifier additives, RSC Adv. 4 (41) (2014) 21238–21245. [39] F.D. Murnaghan, The compressibility of media under extreme pressures, Proc. Natl.
[20] L. Rapoport, Y. Bilik, Y. Feldman, M. Homyonfer, S.R. Cohen, R. Tenne, Hollow nano- Acad. Sci. 30 (9) (1944) 244.
particles of WS2 as potential solid-state lubricants, Nature 387 (1997) 791. [40] K. Reuter, M. Scheffler, Composition, structure, and stability of RuO2 (110) as a func-
[21] A.A. Voevodin, C. Muratore, S.M. Aouadi, Hard coatings with high temperature adap- tion of oxygen pressure, Phys. Rev. B 65 (3) (2001), 035406.
tive lubrication and contact thermal management: review, Surf. Coat. Technol. 257 [41] K. Balasubramanian, S. Khare, D. Gall, Vacancy-induced mechanical stabilization of
(2014) 247–265. cubic tungsten nitride, Phys. Rev. B 94 (17) (2016), 174111.
[22] P. Rabaso, F. Dassenoy, F. Ville, M. Diaby, B. Vacher, T. Le Mogne, M. Belin, J. Cavoret, [42] D.V. Suetin, I.R. Shein, A.L. Ivanovskii, Elastic and electronic properties of hex-
An investigation on the reduced ability of IF-MoS2 nanoparticles to reduce friction agonal and cubic polymorphs of tungsten monocarbide WC and mononitride
and wear in the presence of dispersants, Tribol. Lett. 55 (3) (2014) 503–516. WN from first-principles calculations, Phys. Status Solidi B 245 (8) (2008)
[23] W.A. Brainard, The Thermal Stability and Friction of the Disulfides, Diselenides, and 1590–1597.
Ditellurides of Molybdenum and Tungsten in Vacuum, NASA Technical Note TN D- [43] Z.T.Y. Liu, X. Zhou, D. Gall, S.V. Khare, First-principles investigation of the structural,
5141, NASA, 1969. mechanical and electronic properties of the NbO-structured 3d, 4d and 5d transition
[24] W.G. Sawyer, T.A. Blanchet, Lubrication of Mo, W, and their alloys with H2S gas ad- metal nitrides, Comput. Mater. Sci. 84 (2014) 365–373.
mixtures to room temperature air, Wear 225–229 (I) (1999) 581–586. [44] W.J. Schutte, J.L. De Boer, F. Jellinek, Crystal structures of tungsten disulfide and
[25] I.L. Singer, T.L. Mogne, C. Donnet, J.M. Martin, Friction behavior and wear analysis of diselenide, J. Solid State Chem. 70 (2) (1987) 207–209.
sic sliding against mo in so2, o2 and h2s at gas pressures between 4 and 40 pa, [45] F. Wang, C. Di Valentin, G. Pacchioni, Electronic and structural properties of
Tribol. Trans. 39 (4) (1996) 950–956. WO3: a systematic hybrid DFT study, J. Phys. Chem. C 115 (16) (2011)
[26] V. Totolin, M. Rodríguez Ripoll, M. Jech, B. Podgornik, Enhanced tribological perfor- 8345–8353.
mance of tungsten carbide functionalized surfaces via in-situ formation of low- [46] D.B. Migas, V.L. Shaposhnikov, V.N. Rodin, V.E. Borisenko, Tungsten oxides. I. Effects
friction tribofilms, Tribol. Int. 94 (2016) 269–278. of oxygen vacancies and doping on electronic and optical properties of different
[27] F. Gustavsson, S. Jacobson, Diverse mechanisms of friction induced self-organisation phases of WO3, J. Appl. Phys. 108 (9) (2010), 093713.
into a low-friction material – an overview of WS2 tribofilm formation, Tribol. Int. [47] A. Erdemir, G. Ramirez, O.L. Eryilmaz, B. Narayanan, Y. Liao, G. Kamath, S.K.R.S.
101 (2016) 340–347. Sankaranarayanan, Carbon-based tribofilms from lubricating oils, Nature 536
[28] B. Podgornik, D. Hren, J. Vižintin, Low-friction behaviour of boundary-lubricated (2016) 67.
diamond-like carbon coatings containing tungsten, Thin Solid Films 476 (1) [48] P.A.G. O’Hare, W.N. Hubbard, G.K. Johnson, H.E. Flotow, Calorimetric measurements
(2005) 92–100. of the low-temperature heat capacity, standard molar enthalpy of formation at
[29] A. Neville, A. Morina, T. Haque, M. Voong, Compatibility between tribological sur- 298.15 K, and high-temperature molar enthalpy increments relative to 298.15 K
faces and lubricant additives-how friction and wear reduction can be controlled of tungsten disulfide (WS2), and the thermodynamic properties to 1500 K, J.
by surface/lube synergies, Tribol. Int. 40 (10–12 SPEC. ISS) (2007) 1680–1695. Chem. Thermodyn. 16 (1) (1984) 45–59.

You might also like