Sei sulla pagina 1di 10

Carbon 96 (2016) 411e420

Contents lists available at ScienceDirect

Carbon
journal homepage: www.elsevier.com/locate/carbon

Fluorine and sulfur co-doped amorphous carbon films to achieve


ultra-low friction under high vacuum
Fu Wang a, b, Liping Wang a, *, Qunji Xue a
a
State Key Laboratory of Solid Lubrication, Lanzhou Institute of Chemical Physics, Chinese Academy of Sciences, Lanzhou 730000, PR China
b
University of Chinese Academy of Sciences, Beijing 100039, PR China

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

Article history: Fluorine and sulfur co-doped amorphous carbon (a-C:S:F) films were synthesized from C2H2 and SF6
Received 3 June 2015 plasma mixture, and the bonding states of S and F in films were scrutinized. The a-C:S:F films exhibited a
Received in revised form more ordered carbon structure than hydrogenated films. The friction behavior of films was tested against
16 September 2015
GCr15 bearing steel balls under high vacuum (HV) conditions. Ultra-low steady-state friction coefficient
Accepted 22 September 2015
Available online 26 September 2015
(0.01e0.02) was achieved for a-C:S:F films containing about 2.0 at.% H, depending on the concentrations
of S and F in films. Graphite-like transfer layer on the ball surface provided a low-adhesive sliding surface
terminated by eCF2- and eCF3 groups. The electrostatic repulsive interaction between the F-terminated
transfer layer and the S- and F-terminated carbon structure was probably responsible for the ultra-low
friction behavior. Intriguingly, the sp2-C clusters containing ‘thiophene-S’ (eCeSeCe) structure in films
might contribute to the low friction when sliding against an highly fluorinated surface. These results
demonstrate that the lubricating properties of a-C films in vacuum can be achieved through the S and F
co-doping, which is of great significance for developing a-C films as vacuum lubricants.
© 2015 Elsevier Ltd. All rights reserved.

1. Introduction growth of C-sp2 phase are limited. Thus a large decrease in strong
adhesion bonding that causes high friction and the weak van der
Amorphous carbon-based (a-C) films are very valuable tribo- Waals force at sliding interfaces allow the super-low friction
logical materials, and have been employed in some fields, including behavior [11e16]. Limited shear occurring within the transfer layers
automotive components, magnetic storage disks and infrared op- that may be formed on the uncoated counterparts also contribute
tical windows [1,2]. Some impressive tribological behaviors can be to the low friction [16e18]. Experimental evidences remarkably
observed for various a-C films under certain conditions [3,4]. Under support the aforementioned passivation hypothesis [19e21]. The H,
ultra-high vacuum (UHV) environments, super-low friction coeffi- O and OH passivation of carbon materials can satisfactorily explain
cient (<0.01) is observed for highly hydrogenated amorphous car- the dependence of their friction on test environments and condi-
bon (a-C:H) films, which show great prospects as vacuum solid tions [20e22]. The H- or OH-terminated diamond model signifi-
lubricants. However, all the a-C films containing less than ~34 at.% cantly reduced adhesion and friction between contact surfaces
hydrogen usually exhibit rather high friction value (about 0.5e0.7) [23,24].
under the same condition [5e10], which remarkably retards the The low friction of a-C films can be achieved by controlling
development of a-C films as vacuum solid lubricants. Thus far, there chemistry of their bulk or near surface, or interface tribochemistry
still remain many challenges to develop vacuum-related applica- [4]. Friction properties of various elements (Ti, Cr, W, Si, N, F, etc.)
tions of a-C films, and more attempts will be expected. doped a-C films were widely investigated in air, but few experi-
The super-low friction behavior of a-C:H films under UHV is ments consider their friction behavior in vacuum because of their
attributed to their atomically smooth and chemically inert surfaces. infeasibility. And yet, theoretically, some non-metal elements (be-
The high levels of hydrogenation produces the high chemical sides H), O, F, S, N, etc., can also terminate dangling carbon s-bonds,
inertness in a-C:H film, where the dangling carbon s-bonds and the and consequently reduce adhesion and friction of a-C films.
Nevertheless a basic understanding about the vacuum friction
properties of these films is still lacking. Fluorinated a-C film ach-
* Corresponding author. ieved super-low friction value of 0.005 under UHV [25,26]. Multiple
E-mail address: lpwang@licp.cas.cn (L. Wang).

http://dx.doi.org/10.1016/j.carbon.2015.09.084
0008-6223/© 2015 Elsevier Ltd. All rights reserved.
412 F. Wang et al. / Carbon 96 (2016) 411e420

doped a-C:H:Si:O:F film exhibited low friction coefficient (0.08)


against aluminum ball under vacuum, and a transfer layer passiv-
ated by eH, eF and eOH groups was evidenced [27]. Compared
with the H-terminated DLC model, the F-terminated DLC model
showed lower friction because of stronger repulsion between F
atoms [28,29]. In particular, incorporating S and F into a-C:H film
considerably improved vacuum friction properties of a-C:H films,
and the formation of CeS bonds appeared very important [30,31].
These results imply increasing opportunities for further developing
a-C films as vacuum lubricants by non-metal doping. Detailed
experimental investigations on this subject become very necessary
to understand the structure, properties, and vacuum friction
mechanism of these films.
In this paper, we investigate the F and S passivation of a-C films
towards their vacuum lubricity. Films were deposited from the
plasma mixture of C2H2 and SF6. The structure and composition of
films were determined by Raman spectra, XPS and ToF-SIMS
techniques. Their friction properties were assessed under high
vacuum (HV) conditions. In order to understand their underlying Fig. 1. Cross-section SEM image of a-C:S:F-4 film deposited on Silicon substrate. (A
friction mechanism, the nature of the transfer layer on steel ball colour version of this figure can be viewed online.)
was scrutinized. Finally, it is emphasized that incorporating F and S
into carbon-based network is highly significant for achieving their
low friction under HV conditions. of-flight secondary-ion mass spectrometry (ToF-SIMS) measure-
ment was performed in a ToF-SIMS IV instruments using 30 keV Biþ
2. Experimental details primary ions. Cross-section TEM sample taken from steel ball sur-
face covered with the transfer layer was prepared using focussed-
2.1. Materials ion beam (FIB) milling technique, and was observed by FEI Tecnai
F300 high-resolution transmission electron microscopy (HRTEM).
The samples were deposited on the Si (100) and highly polished
304 stainless steel substrates by parallel-plate hollow cathode 2.3. Friction and wear
plasma-enhanced chemical vapor deposition (PH-PECVD) system
with a base pressure of better than 2.0  103 Pa. Acetylene (C2H2) Friction behavior was evaluated using a home-built vacuum
and sulfur hexafluoride (SF6) were used as precursors. The sub- tribometer with rotational ball-on-disc configuration, equipped
strates were first cleaned ultrasonically using acetone and ethanol with frictional force sensor with accuracy rating of 0.01%. The
for 20 min, respectively. Before film deposition, the substrates counterpart was GCr15 bearing steel ball with diameter of 3.17 mm
underwent Arþ sputter cleaning procedure and thin amorphous (Hardness ¼ ~8.0 GPa, announced mean roughness ¼ 0.02 mm),
silicon interlayer (~80 nm) was deposited to enhance film adhesion. sliding tests were performed under HV (from 1.0  104 to
Si-containing carbon layer (~500 nm) was derived from SiH4 and 4.0  104 Pa) at room temperature. The rotational speed was
C2H2 as gradient layer. Then films were produced from the mixture 300 revs/min and rotational radius was 6.0 mm (linear speed:
of Ar, C2H2 and SF6 at a working pressure of 12 Pa. The samples were 0.19 m/s). The normal load was 1.0 N, corresponding to the initial
denoted as a-C:S:F-1, a-C:S:F-2, a-C:S:F-3 and a-C:S:F-4, corre- mean contact pressure (calculating from a Hertzian model) of 214,
sponding to source gases flow ratios (C2H2:SF6/sccm) of 50:40, 284, 388 and 445 MPa for a-C:S:F-4, a-C:S:F-3, a-C:S:F-2, a-C:S:F-1,
50:60 50:80 and 50:100, respectively. A pulsed-DC source supplied respectively. Sliding tests were repeated three times with the same
the power to gases discharge (voltage, 1.0 kV; pulse frequency, results. The wear volume of film was calculated by multiplying the
1.5 kHz; pulse duty ratio, 30%). Besides, F-containing a-C film (a- cross-sectional area of the track measured using a profile-meter
C:F:H) was also deposited from C2H2/CF4 under similar conditions (KLA-Tencor, D-100) by the circumference of the track. The wear
and a-C:H film was produced from C2H2 utilizing lower voltage, e volume of steel ball side was calculated as follow:
0.8 kV. Fig. 1 shows a fundamental structure pattern of a-C:S:F
  2 
films. The total film thickness is about 2.0 mm according their cross- ph 3d
V¼ þ h2 (1)
section images. 6 4

2.2. Characterizations sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi


d2
h ¼ r  r2  (2)
Raman spectra were recorded with a LabRam HR800 Jobin-Yvon 4
spectrometer with an excitation wavelength of 532 nm. The X-ray
photoelectron spectroscopy (XPS) was performed using a Thermo where d is the wear scar diameter, r is the radius of the steel ball.
Scientific ESCALAB 250Xi instrument equipped with a mono-
chromatic Al Ka (1486.8 eV) X-ray source. Binding energies were 3. Results and discussion
calibrated using the C 1s peak (284.8 eV). The hardness (H) and
Young's modulus (E) of the films were determined by the TI-950 3.1. Bonding characteristics of films
Tribolndenter equipped with a Berkovoich shape indenter. The
scanning electron microscopy (SEM) (JSM-5600LV) equipped with XPS was employed to investigate the elemental bonding states
energy-dispersive X-ray spectroscopy (EDS) and the field-emission in film a-C:S:F-4. In order to determine accurately fluorine chemical
SEM (Hitachi FE-SEM S4800) were used to examine the contact states, an Arþ ion etching step was conducted to remove the oxide
surfaces and the cross-section images of films, respectively. Time- layer on top surface. Fig. 2a shows the full-range XPS spectra of a-
F. Wang et al. / Carbon 96 (2016) 411e420 413

C:S:F-4 film. The O 1s peak disappears after Arþ etching 2 min, 2 position of ‘thiophene-S’ respectively [39,40]. Consequently, sul-
which indicates that oxygen species existed only on film surface. fur in a-C:S:F film was primarily merged into the planar sp2-C
High resolution C 1s spectrum in Fig. 2b can be fitted with five clusters in form of ‘thiophene-S’, and could terminate the formed
component peaks at 284.8, 286.0, 287.5, 289.8 and 292.1 eV, typi- dangling C s-bonds at sp2 cluster boundaries in form of ring CeSeC
cally assigned for the eCeH/eCeCe, eCeSe, eCeCFe, eCFe and bond. With the incorporation of F and S, thus, a-C:S:F films pro-
eCF2- functional groups, respectively [32e34]. It is clear that the vided a terminated carbon structure, which was expected to reduce
major form of F in film was the eCFe group, mono-fluorinated C adhesion and friction in vacuum.
atom. A large proportion of eCeCFe group suggests that most The changes in film microstructures were recorded using Raman
mono-fluorinated C atoms were not directly connected to each spectra. The spectrum of a-C:H film was also collected as a refer-
other, namely, a partly fluorinated carbon-based network. The F 1s ence. All films show typical a-C feature (Fig. 3), an asymmetric peak
peak at 687.2 eV (Fig. 2c) further indicates the eCFe groups. between 1000 and 1700 cm1, which can be decomposed into two
With respect to sulfur element, we have little knowledge of its Gaussian peaks: D peak around 1360 cm1 corresponding to the
bonding configurations in amorphous carbon films. Sulfur and sp2-C organized in delocalized rings and G peak around 1560 cm1
carbon can form various possible structures, including thiols corresponding to the total sp2-C [41,42]. The fitting parameters of
(ReSH), aromatic and aliphatic sulfides (ReSxeR) and thiophenic experimental Raman spectra in Fig. 3 are listed in Table 1. Due to
structures. It is well known that thiophenic structures with a p- varied deposition conditions, the spectrum of a-C:H film is not
electron conjugated system are significantly more stable than thiols comparable with others. Incorporating F into carbon structure
and sulfides. This agrees with the bond dissociation energy (BDE) usually causes the D peak to increase [32], and here a high ID/IG
data of CeS bonds, 300e370 kJ/mol for thiols and sulfides, a ratio of 1.90 is observed for a-C:F:H film. The higher ID/IG values of
calculated value 560 kJ/mol for thiophene [35,36]. And when 2.27 and 2.39 are noted for a-C:S:F-1 and a-C:S:F-4, respectively.
merged into some aromatic systems, the reactivities of thiopenic High ID/IG value for a-C:S:F-1 film should largely be ascribed to the S
structures markedly decrease in the order of thiophenes > incorporation since it contained 2.5 at.% F (Table 2). The stable
benzothiophenes > dibenzothiophenes [37]. Therefore, the energy thiophenic structures mentioned above might allow such a large
ion bombardments and relatively high plasma temperature increase in ID/IG, namely, increasing number or size of sp2 phase
(200e250  C) would preferentially support the stable thiophenic clusters in carbon matrix. The high G peak position is related to
structures during film deposition. What's more, a-C:H film consists increasing sp2 phase and short, strained C]C configuration [41].
of sp2-bonded planar carbon clusters (aromatic ring) embedded in Sulfur doping slightly shifted down the G peak frequency compared
a randomly oriented sp3-bonding matrix, some of the cluster with fluorine doping, suggesting that a more relaxed carbon
boundaries are decorated by H [38]. Thus the growing planar sp2-C structure could be obtained by the addition of S. The FWHM (G)
cluster boundaries decorated by S would favorably form the thio- measures structural disorder arising from bond length and bond
phenic structures, which stabilized the growing sp2 clusters and angle distortions [41], the more perfect or less strained the sp2
prevent energy ions from further decomposing sp2 phases. High clusters, the smaller the FWHM (G) value. The largest FWHM (G)
resolution S 2p spectrum in Fig. 2d can be resolved into two peaks value was observed for a-C:F:H film (172 cm1), conversely, the
at 163.7 and 164.9 eV, which were assigned to the S 2p3/2 and S 2p1/ smallest FWHM (G) value originated from a-C:S:F-1 film

Fig. 2. (a) Full-range XPS spectra of film a-C:S:F-4 before and after Arþ etching, high resolution XPS spectra of (b) C 1s, (c) F 1s, (d) S 2p after Arþ etching 2 min. (A colour version of
this figure can be viewed online.)
414 F. Wang et al. / Carbon 96 (2016) 411e420

Fig. 2a, approximately 8.0 at.% O was detected on the as-deposited


film surfaces, but the O content was less than 1.0 at.% in the bulk of
film. Hydrogen is a decisive factor that affects vacuum friction
properties of a-C:H films, but lack of a direct measurement. Based
on our previous result, a-C:H film containing 9.8 at.% H was syn-
thesized from a C2H2 plasma (Ar: 100 sccm, C2H2: 150 sccm) uti-
lizing the same PECVD systems (voltage: 0.8 kV, pressure: 3.4 Pa)
[44]. In investigated films, the a-C:Si:H layer was produced under
quite similar conditions (Ar: 100 sccm, C2H2: 150 sccm, SiH4:
5 sccm, voltage: e 0.8 kV, Pressure: 3.8 Pa), thereby about 10.0 at%
H. Comparing the intensity of H, CH and C2H ions from a-C:S:F
film with that from a-C:Si:H layer in Fig. 4, it is rationally speculated
that the H in a-C:S:F-4 film is about 2.0 at.% or less. For a-C:S:F films,
it is understandable due to lower proportion of H in precursor,
complete ionization of C2H2, Arþ bombardment and especially
chemical etching of F (removing H in form of HF [45,46]). Thus, H
exerted a negligible influence on friction properties of a-C:S:F-4
film.
ToF-SIMS depth profile analysis in Fig. 4 shows the bulk nature
Fig. 3. Raman spectra of a-C:H, a-C:F:H, a-C:S:F-1 and a-C:S:F-4 films. (A colour of a-C:S:F-4 film that is crucial for its friction performances under
version of this figure can be viewed online.)
HV conditions. The almost stable F intensity represents saturation
in the detector signal, rather than a constant F content. We can not
accurately speculate the F distribution in film as the depth, but high
Table 1 F concentration in the bulk is definite. The high F/H (6/1) atomic
Peak parameters of films obtained through fitting experimental Raman spectra in
ratio (6SF6/2C2H2) in source gases also permitted incorporating F
Fig. 3.
into the bulk [47]. The intensity of S, CH and C ions changes with
Sample G position/cm1 Area ratio Full width at half maximum (G) the depth. This could be associated with changed plasma nature
ID/IG FWHM (G)/cm1
(monitoring current varied from 0.64 A to 0.46 A) due to the growth
a-C:H 1528 1.32 159 of insulator film and therefore microstructural changes. The CH
a-C:F:H 1547 1.90 172 ion in a-C:S:F film layer, only H-containing signal, also indicates low
a-C:S:F-1 1543 2.27 146
a-C:S:F-4 1540 2.39 152
H content. The trace oxygen was also discovered throughout the
whole film. Therefore, the F and S co-doped carbon structure
allowed us to consider merely the influence of F and S on friction
behavior under HV conditions.
Table 2
The surface compositions of samples obtained from XPS measurements.

Sample C (at.%) F (at.%) S (at.%) 3.3. Frictional behavior under high vacuum conditions
a-C:F:H 83.2 16.8 e
a-C:S:F-1 88.9 2.5 8.6 Fig. 5 shows the representative friction coefficient curves, and
a-C:S:F-2 83.5 7.9 8.6 quite different friction behaviors are observed relying on the F and S
a-C:S:F-3 80.5 11.5 8.0 contents in films. The a-C:F:H film exhibited a high friction value
a-C:S:F-4 74.8 15.8 9.3 ~0.38 and failed after ~3500 sliding cycles. Similar friction perfor-
mances were observed for a-C:S:F-1 and a-C:S:F-2, both of which
were worn out after transient sliding, only more sliding cycles for a-
(146 cm1). This suggests that incorporating F into carbon network
led to more distorted or more strained sp2-C clusters, which could
be associated with the larger electron-negativity and atomic radii
(58 pm) of F atom in comparison with H atom (32 pm). However,
the incorporation of S caused structural ordering, relived the stress
felt by C]C bonds and promoted the formation of C]C bonds
organized in delocalized rings. In brief, amorphous carbon-based
network became more ordered and relaxed with introduction of
sulfur.

3.2. Chemical composition of films

The surface compositions of samples were determined by XPS,


and are summarized in Table 2. Accidentally, the a-C:S:F films
showed similar S contents and varied F contents from 2.5 to 15.8
at.%, which is conducive to the subsequent discussion on friction
mechanism. The a-C:F:H film contained 16.8 at.% F, slightly higher
than that of a-C:S:F-4 film. This study aimed to investigate the in-
fluence of F and S on vacuum friction behavior of a-C film and
therefore the existence of O and H in film was undesirable. Surface Fig. 4. Depth profile of secondary ion intensities versus Biþ sputter time for the a-
O contamination is usually observed for a-C films [43]. According to C:S:F-4 film. (A colour version of this figure can be viewed online.)
F. Wang et al. / Carbon 96 (2016) 411e420 415

properties of a-C film strongly depend on the energy (E) of


impacting ions, which is controlled by deposition parameters of
PECVD system: E f VB/P1/2, where VB is the self-bias voltage, P is
the working pressure [1]. The substrate voltage below e 1.0 kV did
not obtain desirable film due to inefficient ionization of SF6. Owing
to strong ions etching, F content in films decreased by 2.0 at.% e 4.0
at.% with the pressure decreasing. Low pressure (high ion energy)
simultaneously induces high film compression stress that affects
the friction performances of films [49], but it is difficult to comment
the film stress precisely. Local exfoliation resulting from high
compression stress was observed for the films deposited under
lower pressures. High compression stress and the decrease in F
content could therefore result in the deterioration of film proper-
ties, as films of 6, 8 and10 Pa.

3.4. Analyses of sliding interface

Fig. 5. Friction curves for films a-C:F:H, a-C:S:F-1, a-C:S:F-2, a-C:S:F-3 and a-C:S:F-4 3.4.1. Structure and chemistry of transfer layer
against GCr15 balls under HV. (A colour version of this figure can be viewed online.) Based on the chemical nature and the easy shear ability of
transfer layer, the friction responses of a-C film may vary a great
deal [16]. After 60 min of sliding against a-C:S:F-4 film, Raman
C:S:F-2. Both a-C:S:F-3 and a-C:S:F-4, containing higher F contents, spectra were collected from transfer layer, wear track and wear
achieved ultra-low steady-state friction coefficient 0.01e0.02, and debris (Fig. 7a). Raman spectrum of transfer layer clearly indicates
meanwhile a super-low friction value (~0.005) was observed at the presence of a carbonaceous layer, and partly separated G peak
initial sliding stages. The super-low friction was ascribed to the fully and D peak reveal a graphite-like characteristic. Raman spectrum of
terminated surface covered by H, F, S and some O-containing wear track resembles that of as-deposited film (Fig. 3), only a
groups. Before reaching the ultra-low friction level, an increasing slightly enhanced D peak in wear track. This could be associated
friction value could be associated with the formation of the transfer with the thin transformation layer (~10 nm) on the surface of wear
layer responsible for ultra-low friction. The tribochemical reactions track, which was beyond Raman detectability (probing depth:
and the formation of wear particles at sliding interface led to minor 50e100 nm for visible excitation). The spectrum in track thereby
friction fluctuation. As shown in Fig. 8a, a considerable amount of reflects film bulk nature and it is difficult to recognize this minor
wear debris was piled around wear scar. change on track surface. This transformed layer might be squeezed
Under HV environments, sliding resistance in solid contact is out the track region and became part of wear debris. The enhanced
primarily controlled by the adhesion [48], which is closely related D peak in wear debris clearly indicates structural changes induced
to the physical and chemical natures of sliding couples. Table 3 by friction. The ultra-low friction behavior of a-C:S:F films could be
summarizes the fundamental physical properties of films. Films partly attributed to the easy shear ability of the transfer layer and
showed decreasing hardness (H) and elastic modulus (E) as the transformed layer. But it is too complicated to quantify their con-
incorporation of F and S, and surface roughness Ra < 10 nm. tributions to low friction.
Apparently, the mechanical properties of films were not essential Chemical states of the transfer layer were determined by XPS.
for achieving vacuum low friction. As seen in Fig. 5, friction High-resolution spectra of C 1s, F 1s, and Fe 2p are shown in
behavior depended strongly on the chemical composition of films. Fig. 7bed. Three new peaks appear at C 1s binding energies of
For a-C:S:F films containing ~8.5 at.% S, the achievement of ultra- 292.5, 293.8 and 295.3 eV, corresponding to the eCF2-, eCF3 and
low friction required higher F content (>11.0 at.%), just as a-C:S:F- eCF2eCF3 functional groups, respectively [33,34]. The peaks at
3 and a-C:S:F-4. However, a-C:F:H film with high F content
exhibited high friction value, suggesting that S doping played a
crucial role in reducing friction to ultra-low level and led to more
enduring sliding. Based on these results, we recognize that a judi-
cious combination of F and S within films becomes particularly vital
to obtain intrinsic lubricity of a-C films. Depending on the varied S
content in film, different threshold in F content may be required to
achieve ultra-low friction under HV.
Preliminary process optimization was conducted to ameliorate
the film properties. The working pressure, substrate voltage or Ar
flow was regulated to alter film natures. Friction properties of as-
deposited films are presented in Fig. 6. Film deposited from work-
ing pressure of 12 Pa and Ar flow of 100 sccm provided the longest
wear life, exceeding 80000 sliding cycles. The structures and

Table 3
Fundamental physical properties of a-C:S:F films and fluorinated film.

Sample a-C:S:F-1 a-C:S:F-2 a-C:S:F-3 a-C:S:F-4 a-C:F:H

H/GPa 8.6 6.8 3.7 2.6 11.9


Fig. 6. Friction curves for films deposited at working pressure 6 Pa, 8 Pa, 10 Pa, 12 Pa
E/GPa 77.9 59.4 33.6 20.8 122.7
(Ar:100 sccm), and Ar flow 150 sccm (12 Pa) against GCr15 balls under HV. (A colour
Ra/nm 8.01 6.98 6.97 5.37 7.23
version of this figure can be viewed online.)
416 F. Wang et al. / Carbon 96 (2016) 411e420

Fig. 7. (a) Raman spectra of transfer layer, wear track and wear debris formed after 60 min of sliding against a-C:S:F-4 film, and corresponding high resolution XPS spectra of (b) C
1s, (c) F 1s and (d) Fe 2p on the transfer layer. (A colour version of this figure can be viewed online.)

286.5 and 288.9 eV are indefinite, and can be partly assigned to the active F radicals or ions.
b position (first neighbors) of eCF2/CF3 groups [34]. But we can Fig. 8 presents the SEM image of wear scar and corresponding
not exclude the O-containing groups (CeOeC/CeOH: 286.4 eV, O] EDS mappings after 60 min of sliding over a-C:S:F-4 film. A great
CeOH: 288.9 eV [50]), since both the residual H2O in vacuum of amount of wear debris is observed around the wear scar of about
about 104 Pa and exposure to air during transferring sample to XPS 200 mm diameter (Fig. 8a). This could be mainly attributed to severe
analysis chamber would cause impossible oxidation. The F 1s mechanical wear due to high film softness (2.6 GPa, Table 3). The
spectrum (Fig. 7c) can be decomposed into three peaks at 684.8, large-size wear particles further account for the mechanical wear.
687.1 and 689.6 eV, assigned to the FeF2, eCFe and eCF2/CF3 The EDS mappings (Fig. 8bed) confirmed the overall variation of
groups, respectively. The major peak at 689.6 eV suggests that F on elements induced by rubbing. The C and S signals were detected
the surface of transfer layer mainly existed in the eCF2- and eCF3 uniformly over the whole wear debris, but a significantly weak F
groups. The weak Fe 2p peak (Fig. 7d) at 711.6 eV confirms the signal in wear debris. This illustrates that F participated in tri-
presence of FeF2. There is no obvious S observed on the surface of bochemical reactions at sliding interface and was partly removed
transfer layer (Fig. 12). Based on the XPS analysis, the surface of the from the carbonaceous structure with the generation of wear
transfer layer composed of 68.3 at.% C, 27.1 at.% F, 2.5 at.% S and 2.0 debris. The CeF bonds could be broken more easily under rubbing
at.% Fe, a significant F accumulation. This means that the transfer conditions because of higher reactivity of F atom. This also hints the
layer provided an highly fluorinated carbon surface with eCF2/ formation of the fluorinated transfer layer (Fig. 7). However, owing
CF3 groups, which possessed low surface energy, favoring low to their stability, the sp2-C clusters containing thiophenic struc-
adhesion. tures might be removed as a whole though breaking the relatively
The accumulated F atoms on the transfer layer surface is very weak CeC bonds (BDE: 400e460 kJ/mol [36]) linking sp2-C clusters
important for achieving ultra-low friction of a-C:S:F film, and it was to sp3-C matrix. This may be the reason why S was remained in
also observed for a-C:F:H film sliding against aluminum or mag- wear debris, instead of transfer layer.
nesium alloy [51]. Although it is quite difficult to break CeF bonds
(BDE: 450e550 kJ/mol [36]) under thermodynamic conditions, the 3.4.2. Wear tracks and wear behavior
rubbing can facilitate the CeF bond cleavage due to mechanical The a-C:S:F-1 and a-C:S:F-2 films were completely removed
strains, triboemission and friction heat [52]. Furthermore, it had from steel substrates, accompanying by a sharp friction increase
been found that triboemission intensity increased with increasing (Fig. 5). After 60 min of sliding, SEM images of wear tracks for a-
electric resistance of solid [53]. The insulator properties of a-C film C:S:F-3 and a-C:S:F-4 are showed in Fig. 9. Local exfoliation (Fig. 9a)
can thereby promote this process and the CeF bond breaking. The and clear cracks (Fig. 9b) are seen on the track regions. This could be
cleavage of CeF bond was noticed in tribochemical decomposition associated with unreliable adhesion of film to substrates. High-
of ionic liquids, where the formation of CFþ and FeF2/FeF3 origi- resolution XPS spectra in Fig. 9c indicate the chemical states of
nating from the cleavage of CeF bond in (SO2CF3)2N [54,55]. More track surface for a-C:S:F-4 film. The C 1s (292.8 eV), F 1s (685.0 and
importantly, high reactivity of F atom may be responsible for their 689.4 eV) and Fe 2p peaks show the existence of small qualities of
accumulation on the transfer layer because the newly-formed multi-fluorinated C atoms and iron fluoride, which detached from
reactive sites at sliding interface may display strong affinity for the fluorinated transfer layer under rubbing actions. However, the
F. Wang et al. / Carbon 96 (2016) 411e420 417

Fig. 8. (a) SEM image of wear scar formed after 60 min of sliding contact with a-C:S:F-4, and corresponding EDS mappings for (b) C, (c) F, and (d) S. (A colour version of this figure
can be viewed online.)

wear track mainly reflected the intrinsic nature of film. The area
ratio of fitting peaks in F 1s spectrum (Fig. 9c) shows 12.0 at.% F in
eCFx-groups, suggesting a reduced F content in the bulk of film. In
contrast with a-C:S:F-4 (2.6 GPa), the wider track (270 mm) for a-
C:S:F-3 (3.7 GPa) illustrates that the wear resistance of films under
HV strongly depends on their compositions, and increasing F con-
tent in a-C:S:F-4 reduced wear. Owing to the build-up and repair of
the fluorinated transfer layer, film with relatively low F content
would thereby sacrifice larger volume to fulfill F demand, and
hence higher wear for a-C:S:F-3 film.
Considering local exfoliation of film layer in Fig. 9a, quantifying
wear of film was performed after 30 min of sliding test. The cor-
responding optical images and cross-sectional profiles of wear
tracks are shown in Fig. 10. The a-C:S:F-3 film showed higher wear
rate (1.24  106 mm3/Nm) than a-C:S:F-4 film (4.80  107 mm3/
Nm). According to equations (1) and (2), the wear rate of the steel
ball side was estimated to be 8.10  108 mm3/Nm from the wear
scar in Fig. 8a. This indicates that a-C:S:F film with high F and S
content provided low wear rate and the formation of fluorinated
transfer layer drastically reduced the wear of steel ball.

3.5. Roles of sulfur and fluorine in friction

Differing from H element, both F and S show high reactivity with


iron (Fe), signifying complex tribochemical reactions. A rough
judgment from the standard enthalpy of formation of iron (II)
fluoride and sulfide (△Hf0 (FeF2): 705.84 kJ/mol, △Hf0
(FeS): 100.42 kJ/mol, △Hf0 (FeS2): 171.54 kJ/mol) suggests that
steel counterpart would preferentially react with F within film. In
addition, lower BDE value for CeF bonds in comparison with ring
Fig. 9. SEM images of wear tracks formed after 60 min of sliding over (a) a-C:S:F-3 and
CeS bonds and higher F concentration in film also approves the
(b) a-C:S:F-4, (c) high-resolution XPS C 1s, F 1s and Fe 2p spectra taken from the area formation of FeF2. The F 1s and Fe 2p spectra in Fig. 7c and
marked by red dotted line in (b). (A colour version of this figure can be viewed online.) d confirm the presence of FeF2 on the transfer layer. The high-
418 F. Wang et al. / Carbon 96 (2016) 411e420

Fig. 10. Optical images and cross-sectional profiles of the wear tracks produced after 30 min of sliding over (a) a-C:S:F-3 and (b) a-C:S:F-4. (A colour version of this figure can be
viewed online.)

resolution TEM image in Fig. 11 further provides the evidence for films to achieve ultra-low friction under HV. Considering the
the presence of FeF2 in the transfer layer. A clear FeF2 layer is passivation, the F atom in carbon structure is similar to H, but the
observed and clings to the surface of GCr15 ball. It was proposed friction behavior of fluorinated a-C films under HV has been rarely
that the formation of AlF3 at Al/transfer layer interface promoted reported. Fluorinated carbon surfaces are considered to be greater
the build-up of carbon transfer layer [27]. We consider that the FeF2 potential as low friction surface than hydrogenated ones because of
at Fe/transfer layer also conduced to the formation of the transfer stronger repulsion between F atoms [28,29]. A fully F-terminated
layer and could reduce subsequent adhesion between film and steel diamond surface was entirely covered by charges (electron den-
counterpart. sity > 0.1) and more effectively blocked strong CeC covalent in-
Friction curves in Fig. 5 show that the F was essential for a-C:S:F teractions across sliding interface [28,56]. The larger electro-
negativity and size of F atom may therefore enhance its passiv-
ation ability. As reported by Fontaine et al., the super-low friction
(0.005) was achieved by fluorinated a-C film with 18 at.% F þ 5 at.%

Fig. 12. High resolution S 2p spectra of film, wear track, wear debris and transfer layer
Fig. 11. High-resolution transmission electron microscopy (HRTEM) image of transfer after 60 min of sliding against a-C:S:F-4 film. (A colour version of this figure can be
layer/GCr15 ball interface formed after sliding against a-C:S:F-4 film. viewed online.)
F. Wang et al. / Carbon 96 (2016) 411e420 419

Fig. 13. Schematic description of the sliding interface for achieving ultra-low friction of a-C:S:F films under HV. (A colour version of this figure can be viewed online.)

H [26] against a steel surface under UHV, a reduced demand for ultra-low friction. Consequently, based on current understanding,
passivation atoms (23 at.%, markedly less than 34 at.% for a-C:H fluorine and sulfur co-doping may be of great significance for
films). Besides the tribochemical reactions, stronger passivation developing vacuum-related applications of a-C films. And fortu-
ability of F atom might reasonably explain such a distinction. nately, the deterioration of film mechanical properties can be
Fluorinated carbon network is thereby more beneficial to provide a improved by the further design of film structure and composition
fully passivated surface. In this paper, 16.8 at.% F in a-C:F:H film modification.
could not completely shield strong CeC bond adhesion, which
hindered the formation of the fluorinated transfer layer, and hence 4. Conclusions
high friction value. But the formation of FeF2 at sliding interface
might retard the film failure due to impeding the strong adhesion In summary, we have synthesized and investigated fluorine and
to some extent. Accordingly, regarding fluorinated a-C films, sulfur co-doped a-C films. It is demonstrated that the ultra-low
despite possible favorable tribochemical reactions, a sufficiently friction behavior of a-C films under HV can be achieved by the S
passivated carbon network is still crucial for achieving their low and F co-passivation, not confined to the H passivation. The stable
friction under HV. For the investigated a-C:S:F films, however, more thiophenic structures merged into the sp2-C clusters have been
than 11 at.% F was required to produce a desirable carbon structure. identified as the likeliest form for S atom in films, which promoted
Sulfur is another contributor to the ultra-low friction in the a- the formation of more ordered carbon structure and contributed to
C:S:F films. Compared with the a-C:F:H film, the existence of S in a- the low friction. Both the delocalized p-system containing ‘thio-
C:S:F films directly reduced friction value to ultra-low level and phene-S’ and CeF bonds can allow forming negatively charged
reduced the demand for F atoms. Meanwhile, the S-doping a-C:H surface, and thereby the electrostatic repulsion between sliding
film with less than 30 at.% H also obtained low friction under HV surfaces achieve low adhesion and ultra-low friction. Considering
and allowed more enduring sliding due to the strong CeS bond the tribochemical reactions, stronger passivation ability, and the
(BDE in diatomic molecules: CeS: 713, CeC: 618, CeF: 514, CeH: thiophenic structures, we believe that a judicious combination of F
338 kJ/mol) [30,31]. Therefore, the low friction under HV was and S elements in film, in particular, a series of a-C:S:F films with
likewise achieved by S-doped a-C film containing less F or H atoms. slightly lower than 8 at.% S will be expected to produce more
In order to understand underlying friction mechanism, high- valuable films for possible applications. More importantly, this
resolution XPS S 2p spectra are shown in Fig. 12. The S 2p peaks study explores new pathways for ameliorating the intrinsic lu-
in wear track and wear debris are almost identical to that of as- bricity of a-C films. Multi-nonmetal doping is a promising approach
deposited film. As discussed above, these peaks can be assigned to design and develop a-C films as vacuum lubricants.
to the ring CeS bonds in thiophenic structures. The transfer layer
shows a very feeble S 2p peak (Fig. 12). This suggests that there Acknowledgments
were no obvious changes in CeS bonds during rubbing. This can be
explained by high reactivity of F atom and stability of thiophenic This work was supported by the State Key Project of Funda-
structures. Especially, with the formation of the fluorinated transfer mental Research of China (2013 CB 632302) and the National
layer, ‘thiophenic-S’ would reveal lower reaction tendency towards Natural Science Foundation of China (Grant No. 51322508).
fluorinated surface. The sp2-C clusters containing ‘thiophene-S’, as
a whole, were separated from the sp3-C matrix and agglomerated References
into wear debris due to lower CeC bond strength. Now the ‘thio-
phene-S’ in films, on the one hand, terminated many sp2-C clusters [1] J. Robertson, Diamond-like amorphous carbon, Mater Sci. Eng. R. 37 (4) (2002)
129e281.
and thereby decreased cross-link density of carbon-based network. [2] A.H. Lettington, Applications of diamond-like carbon thin films, Carbon 36 (5)
The hardness and wear resistance of films deteriorated. On the (1998) 555e560.
other hand, the ‘thiophene-S’ intrinsically eliminated C s-bonds at [3] K. Al Mahmud, M. Kalam, H. Masjuki, H. Mobarak, N. Zulkifli, An updated
overview of diamond-like carbon coating in tribology, Crit. Rev. Solid State
sp2 cluster edges and stabilized the clusters. These stable clusters Mater Sci. 40 (2) (2015) 90e118.
would reduce the production of the carbon dangling C s-bonds [4] A. Erdemir, O. Eryilmaz, Achieving superlubricity in DLC films by controlling
during friction. And numerous stabilized planar sp2 clusters hold- bulk, surface, and tribochemistry, Friction 2 (2) (2014) 140e155.
[5] C. Donnet, M. Belin, J. Auge, J. Martin, A. Grill, V. Patel, Tribochemistry of
ing delocalized p-electron systems would show low adhesion to- diamond-like carbon coatings in various environments, Surf. Coat. Technol. 68
ward highly fluorinated transfer layer. (1994) 626e631.
According to the above statements, a proposed interface model [6] C. Donnet, A. Grill, Friction control of diamond-like carbon coatings, Surf. Coat.
Technol. 94 (1997) 456e462.
is shown in Fig. 13. Owing to the elemental difference in electro-
[7] A. Erdemir, O. Eryilmaz, G. Fenske, Synthesis of diamond like carbon films
negativity (F-3.98, S-2.58, C-2.55, H-2.2), both CeF bonds and with super low friction and wear properties, J. Vac. Sci. Technol. A 18 (4)
delocalized p-systems in sp2-C clusters containing ‘thiophenic-S’ (2000) 1987e1992.
can allow the negatively charged surface. Therefore, the electro- [8] J. Fontaine, T. Le Mogne, J. Loubet, M. Belin, Achieving super low friction with
hydrogenated amorphous carbon: some key requirements, Thin Solid Films
static repulsion between highly fluorinated transfer layer and S- 482 (1) (2005) 99e108.
and F-terminated carbon structure may be a major reason for the au, J. Terrat, et al., Diamond-
[9] C. Donnet, J. Fontaine, T. Le Mogne, M. Belin, C. He
420 F. Wang et al. / Carbon 96 (2016) 411e420

like carbon-based functionally gradient coatings for space tribology, Surf. [33] G. Nanse, E. Papirer, P. Fioux, F. Moguet, A. Tressaud, Fluorination of carbon
Coat. Technol. 120 (1999) 548e554. blacks: an X-ray photoelectron spectroscopy study: I. A literature review of
[10] J. Fontaine, Towards the use of diamond-like carbon solid lubricant coatings in XPS studies of fluorinated carbons. XPS investigation of some reference
vacuum and space environments, Proc. IMechE Part J J. Eng. Tribol. 222 (8) compounds, Carbon 35 (2) (1997) 175e194.
(2008) 1015e1029. [34] T. Shirasaki, F. Moguet, L. Lozano, A. Tressaud, G. Nanse, E. Papirer, Fluorina-
[11] C. Donnet, J. Fontaine, A. Grill, T. Le Mogne, The role of hydrogen on the tion of carbon blacks: an X-ray photoelectron spectroscopy study: IV. Reac-
friction mechanism of diamond-like carbon films, Tribol. Lett. 9 (3e4) (2001) tivity of different carbon blacks in CF4 radiofrequency plasma, Carbon 37 (12)
137e142. (1999) 1891e1900.
[12] A. Erdemir, The role of hydrogen in tribological properties of diamond-like [35] Y.R. Luo, Comprehensive Handbook of Chemical Bond Energies, CRC press,
carbon films, Surf. Coat. Technol. 146 (2001) 292e297. 2007.
[13] J. Fontaine, C. Donnet, A. Grill, T. LeMogne, Tribochemistry between hydrogen [36] N. Hore, D.K. Russell, The thermal decomposition of 5-membered rings: a laser
and diamond-like carbon films, Surf. Coat. Technol. 146 (2001) 286e291. pyrolysis study, New J. Chem. 28 (5) (2004) 606e613.
[14] A. Erdemir, C. Donnet, Tribology of diamond-like carbon films: recent prog- [37] C. Song, An overview of new approaches to deep desulfurization for ultra-
ress and future prospects, J. Phys. D Appl. Phys. 39 (18) (2006) R311. clean gasoline, diesel fuel and jet fuel, Catal. Today 86 (1) (2003) 211e263.
[15] A. Erdemir, Design criteria for superlubricity in carbon films and related mi- [38] J. Robertson, E. Oreilly, Electronic and atomic structure of amorphous carbon,
crostructures, Tribol. Int. 37 (7) (2004) 577e583. Phys. Rev. B 35 (6) (1987) 2946.
[16] A. Erdemir, Genesis of superlow friction and wear in diamond like carbon [39] W. Kicin  ski, A. Dziura, Heteroatom-doped carbon gels from phenols and
films, Tribol. Lett. 37 (11) (2004) 1005e1012. heterocyclic aldehydes: sulfur-doped carbon xerogels, Carbon 75 (2014)
[17] J. Fontaine, T. Le Mogne, J. Loubet, M. Belin, Achieving superlow friction with 56e67.
hydrogenated amorphous carbon: some key requirements, Thin Solid Films [40] M. Li, C. Liu, H. Zhao, H. An, H. Cao, Y. Zhang, et al., Tuning sulfur doping in
482 (1) (2005) 99e108. graphene for highly sensitive dopamine biosensors, Carbon 86 (2015)
[18] Y. Liu, A. Erdemir, E. Meletis, A study of the wear mechanism of diamond-like 197e206.
carbon films, Surf. Coat. Technol. 82 (1) (1996) 48e56. [41] C. Casiraghi, A. Ferrari, J. Robertson, Raman spectroscopy of hydrogenated
[19] A. Konicek, D. Grierson, P. Gilbert, W. Sawyer, A. Sumant, R.W. Carpick, Origin amorphous carbons, Phys. Rev. B 72 (8) (2005) 085401.
of ultralow friction and wear in ultrananocrystalline diamond, Phys. Rev. Lett. [42] A. Ferrari, J. Robertson, Interpretation of Raman spectra of disordered and
100 (23) (2008) 235502. amorphous carbon, Phys. Rev. B 61 (20) (2000) 14095.
[20] A. Konicek, D. Grierson, A. Sumant, T. Friedmann, J. Sullivan, P. Gilbert, et al., [43] A. Al-Azizi, O.L. Eryilmaz, A. Erdemir, S.H. Kim, Surface structure of hydro-
Influence of surface passivation on the friction and wear behavior of ultra- genated diamond-like carbon-Origin of run-in behavior prior to super-
nanocrystalline diamond and tetrahedral amorphous carbon thin films, Phys. lubricious interfacial shear, Langmuir 31 (5) (2015) 1711e1721.
Rev. B 85 (15) (2012) 155448. [44] R. Zhang, L. Wang, Synergistic improving of tribological properties of amor-
[21] L. Cui, Z. Lu, L. Wang, Probing the low-friction mechanism of diamond-like phous carbon film enhanced by FeSi-doped multilayer structure under cor-
carbon by varying of sliding velocity and vacuum pressure, Carbon 66 rosive environment, Surf. Coat. Technol. 276 (2015) 626e635.
(2014) 259e266. [45] C. Bottani, A. Lamperti, L. Nobili, P. Ossi, Structure and mechanical properties
[22] S. Bhowmick, A. Banerji, A.T. Alpas, Role of humidity in reducing sliding of PACVD fluorinated amorphous carbon films, Thin Solid Films 433 (1) (2003)
friction of multilayered graphene, Carbon 87 (2015) 374e384. 149e154.
[23] J.D. Schall, G. Gao, J.A. Harrison, Effects of adhesion and transfer film formation [46] R. Prioli, L. Jacobsohn, M.M. da Costa, F. Freire Jr., Nanotribiological properties
on the tribology of self-mated dlc contactsy, J. Phys. Chem. C 114 (12) (2009) of amorphous carbon-fluorine films, Tribol. Lett. 15 (3) (2003) 177e180.
5321e5330. [47] M. Matsushita, M.Z.B. Harum, A. Kashem, S. Morita, Chemistry of Sulfur
[24] M.-I. De Barros Bouchet, G. Zilibotti, C. Matta, M.C. Righi, L. Vandenbulcke, Doping Mechanism in Amorphous Carbon by Plasma CVD of Methane,
B. Vacher, et al., Friction of diamond in the presence of water vapor and J. Photopolym. Sci. Technol. 12 (1) (1999) 11e13.
hydrogen gas. coupling gas-phase lubrication and first-principles studies, [48] K. Miyoshi, Considerations in vacuum tribology (adhesion, friction, wear, and
J. Phys. Chem. C 116 (12) (2012) 6966e6972. solid lubrication in vacuum), Tribol. Int. 32 (11) (1999) 605e616.
[25] C. Donnet, J. Fontaine, A. Grill, V. Patel, C. Jahnes, M. Belin, Wear-resistant [49] Y. Wu, H. Li, L. Ji, Y. Ye, J. Chen, H. Zhou, Vacuum tribological properties of a-C:
fluorinated diamond like carbon films, Surf. Coat. Technol. 94 (1997) H film in relation to internal stress and applied load, Tribol. Int. 71 (2014)
531e536. 82e87.
[26] J. Fontaine, J. Loubet, T.L. Mogne, A. Grill, Superlow friction of diamond-like [50] D. Yang, A. Velamakanni, G. Bozoklu, S. Park, M. Stoller, R.D. Piner, et al.,
carbon films: a relation to viscoplastic properties, Tribol. Lett. 17 (4) (2004) Chemical analysis of graphene oxide films after heat and chemical treatments
709e714. by X-ray photoelectron and Micro-Raman spectroscopy, Carbon 47 (1) (2009)
[27] F. Sen, X. Meng-Burany, M. Lukitsch, Y. Qi, A. Alpas, Low friction and envi- 145e152.
ronmentally stable diamond-like carbon (DLC) coatings incorporating silicon, [51] F. Sen, Y. Qi, A. Alpas, Material transfer mechanisms between aluminum and
oxygen and fluorine sliding against aluminum, Surf. Coat. Technol. 215 (2013) fluorinated carbon interfaces, Acta Mater 59 (7) (2011) 2601e2614.
340e349. [52] C.K. Kajdas, Importance of the triboemission process for tribochemical reac-
[28] F. Sen, Y. Qi, A.T. Alpas, Surface stability and electronic structure of hydrogen- tion, Tribol. Int. 38 (3) (2005) 337e353.
and fluorine-terminated diamond surfaces: A first-principles investigation, [53] K. Nakayama, Triboemission of charged particles and resistivity of solids,
J. Mater Res. 24 (08) (2009) 2461e2470. Tribol. Lett. 6 (1) (1999) 37e40.
[29] S. Bai, T. Onodera, R. Nagumo, R. Miura, A. Suzuki, H. Tsuboi, et al., Friction [54] R. Lu, S. Mori, K. Kobayashi, H. Nanao, Study of tribochemical decomposition
reduction mechanism of hydrogen-and fluorine-terminated diamond-like of ionic liquids on a nascent steel surface, Appl. Surf. Sci. 255 (22) (2009)
carbon films investigated by molecular dynamics and quantum chemical 8965e8971.
calculation, J. Phys. Chem. C 116 (23) (2012) 12559e12565. [55] M. Mahrova, M. Conte, E. Roman, R. Nevshupa, Critical insight into mecha-
[30] N. Moolsradoo, S. Watanabe, Modification of tribological performance of DLC nochemical and thermal degradation of imidazolium-based ionic liquids with
films by means of some elements addition, Diam. Relat. Mater 19 (5) (2010) alkyl and monomethoxypoly (ethylene glycol) side chains, J. Phys. Chem. C
525e529. 118 (39) (2014) 22544e22552.
[31] N. Moolsradoo, S. Watanabe, Deposition and tribological properties of sulfur- [56] S. Bai, H. Murabayashi, Y. Kobayashi, Y. Higuchi, N. Ozawa, K. Adachi, et al.,
doped DLC films deposited by PBII method, Adv. Mater Sci. Eng. 2010 (2010). Tight-binding quantum chemical molecular dynamics simulations of the low
[32] G. Yu, B. Tay, Z. Sun, L. Pan, Properties of fluorinated amorphous diamond like friction mechanism of fluorine-terminated diamond-like carbon films, RSC
carbon films by PECVD, Appl. Surf. Sci. 219 (3) (2003) 228e237. Adv. 4 (64) (2014) 33739e33748.

Potrebbero piacerti anche