Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Surface & Coatings Technology 445 (2022) 128716

Contents lists available at ScienceDirect

Surface & Coatings Technology


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

Effect of low-pressure deposition on the mechanical and tribological


properties of a-C:H films deposited via modified pulsed-DC PECVD with
active screen as an additional cathode
A. Capote a, *, G. Capote b, E.J. Corat a, V.J. Trava-Airoldi a
a
Instituto Nacional de Pesquisas Espaciais (INPE), Laboratório Associado de Sensores e Materiais (LABAS), Avenida dos Astronautas 1758 - São José dos Campos,
12227-010, SP, Brazil
b
Universidad Nacional de Colombia, Facultad de Ciencias, Departamento de Física, Carrera 45, N. 26-85-Edificio Uriel Gutiérrez, Bogotá D.C., Colombia

A R T I C L E I N F O A B S T R A C T

Keywords: Due to their attractive mechanical, chemical, tribological, and biological properties, diamond-like carbon (DLC)
Diamond-like carbon coatings are widely used in many industrial applications. Normally, the deposition of amorphous hydrogenated
Deposition pressure carbon (a-C:H) films is obtained via the plasma-enhanced chemical vapor deposition (PECVD) technique.
PECVD
Recently, a modified pulsed-DC PECVD technique has attracted attention in this field due to its making it possible
Additional cathode
Tribology
to grow a-C:H films at low-pressures (up to 0.1 Pa) in a collision less regime, leading to the achievement of
Mechanical properties improved mechanical and tribological properties. The incorporation of an additional cathode allows the
confinement of electrons and ions, stabilizing and densifying the cold plasma at much lower pressure than
conventional PECVD systems. This method is distinguished by higher energy and acceleration of the C+ ions,
simpler reactor system equipment, and lower operating cost. In the present investigation, hydrogenated amor­
phous carbon films were grown on AISI 316 stainless steel using a thin amorphous silicon interlayer to improve
interface adherence. The characterization of the coatings allowed identifying the influence of the deposition
pressure on the microstructural, mechanical, and tribological properties and the film's adherence. The main
results indicate that obtaining films with high hardness, low compressive stress, low coefficient of friction, high
wear resistance, and appropriate adhesion is possible at deposition pressures almost 100 times lower than
conventional PECVD techniques.

1. Introduction phase [4,5]. Amorphous carbon hard films can be grown using physical
vapor deposition (PVD) processes such as sputtering, pulsed filtered
Hydrogenated amorphous carbon (a-C:H) films are part of the cathodic vacuum arc deposition (PFCVAD), and many others [1,6].
Diamond-Like Carbon (DLC) coatings family. They have interesting Among the chemical vapor deposition (CVD) techniques, plasma-
mechanical, tribological, and biological properties, among other char­ enhanced chemical vapor deposition (PECVD) allows better tempera­
acteristics [1]. These films can be used in several applications due to ture control of the substrates [3,7]. Normally, the properties of a-C:H
their high degree of hardness and wear resistance and low coefficient of films grown by means of the PECVD technique strongly depend on the
friction, in addition to their chemical inertness and biological compat­ process parameters, such as the pressure, bias-voltage, temperature,
ibility [2,3]. The allotropic nature of the carbon element allows the dilution composition, and chemical nature of the precursor gases
formation of many multi-atomic crystalline and amorphous structures [7–11].
with different contents of all three types of hybridization (sp3, sp2, and In the modified PECVD technique, when an additional cathode or
sp) [1,2]. It is possible to classify DLC films according to the hydrogen active screen is incorporated, it is possible to achieve lower deposition
content and the variation of the sp3 (diamond-like) and sp2 (graphite- pressures with a near collision less regime between plasma species. In
like) carbon bonding ratio [4]. In general, a-C:H films are those with addition, the cylinder shape dimension and grid transparency of the
hydrogen content between 20 and 60 % and up to 40 % of the C–C sp3 additional cathode produces an electric field that allows the

* Corresponding author.
E-mail address: ariel.sanchez@inpe.br (A. Capote).

https://doi.org/10.1016/j.surfcoat.2022.128716
Received 21 April 2022; Received in revised form 4 July 2022; Accepted 5 July 2022
Available online 12 July 2022
0257-8972/© 2022 Elsevier B.V. All rights reserved.
A. Capote et al. Surface & Coatings Technology 445 (2022) 128716

confinement of electrons and ions between the cathode surfaces [12,13], subjected to the same bias voltage. Due to the film's growth on both
thus increasing the ionization probabilities and producing a local in­ cathode surfaces, a sand-blasting cleaning processes must be performed
crease in the plasma density and therefore an increase in the tempera­ before each deposition in order to guarantee the uniformity of the
ture or energy of the species present in the plasma [14]. Under these electrode's electrical conductivity. A schematic representation of the
conditions, it is possible to obtain highly energetic ions with a narrower bipolar pulsed-DC PECVD system with an additional cathode is shown in
energy distribution [12]. The formation of a cold plasma with high- Fig. 1. A simplified piping and instrument diagram (P&ID) includes the
energy ions and high neutral species flux is important in a-C:H film pumping system, consisting of a vacuum pump, used to obtain the pri­
nucleation and densification [14–19]. Furthermore, this modified mary vacuum, and an oil diffusion pump, which together allow reaching
technique has allowed improving the uniformity of the DLC coatings' internal pressures of up to 0.01 Pa. The presence of two pressure gauges,
growth in complex three-dimensional geometries, obtaining high- one of the Pirani type (PT-1, for pressure transmitter) and the other of
quality films with enhanced hardness and adhesion [10,12,13,20]. the capacitance diaphragm type (PT-2), were used to monitor the
In order to obtain coatings that meet industrial specifications, it is pressure inside the reactor chamber. The precursor gases are inserted
necessary to control the deposition parameters of the PECVD technique, into the reactor with a mass flow control system. The system has several
which have a great influence on the properties of the DLC films. In valves along the reactor pipes, including manual actuation needle valves
general, predicting the influence of each parameter on the deposited (FCV-1, for the flow control valve) and pneumatic actuated valves (PCV-
coating's properties is challenging, because of the possible nonlinear 1 and PCV-2, for the pressure control valve). Additionally, a butterfly
behavior and the need to perform a large number of experiments valve (PCV-3) arranged in the diffusion pump is actuated by a servo­
[21,22]. Additionally, the material used as substrate has a considerable motor system.
influence on the adherence of these films, and limitations are usually First, an argon (Ar) plasma cleaning of the samples was performed at
due to the high residual compressive stress in the film's microstructure, low pressure (0.4 Pa), with 40 % duty cycle and bias voltage of − 600 V.
inherent in the deposition process [10]. Usually, the incorporation of an The Ar plasma acts on the surface, removing organic contaminants and
amorphous silicon interlayer improves adherence between the substrate other molecules remaining on the substrate. The argon flow is main­
surface and DLC coatings on metallic substrates [10,12,23,24]. tained throughout the deposition process because it serves as a plasma
In this study, DLC coatings were grown on AISI 316 stainless steel stabilizer for silane (SiH4) and acetylene (C2H2) [8]. The first deposition
substrates using the modified pulsed-DC PECVD technique with an step consists of the formation of a thin amorphous silicon (a-Si) inter­
additional cathode. The influence of the deposition pressure, especially layer on the substrate surface, leading to an increase in the interfacial
within the studied low range (0.1 to 1.3 Pa), on the film's adherence and adhesion between a-C:H films and metallic surfaces [10,24,25]. This a-Si
the microstructural, mechanical, and tribological properties was thin film was obtained with a silane (4 sccm) precursor gas mixed with
analyzed. The observed trends in the micro- and macro-scale properties argon (2 sccm) at low pressure (0.5 Pa), with a 40 % duty cycle and
of a-C:H films showed that it is possible to obtain hard, low-stressed a-C: applying a bias voltage of − 800 V. The second deposition step corre­
H films, with a low coefficient of friction, high wear resistance, and sponded to a transitional film layer between a-Si and a-C:H. This layer
appropriate adhesion. was grown using argon (2 sccm) and acetylene (30 sccm) and a gradu­
ally decreasing silane mass flow (0.5 sccm per minute) as precursor
2. Experimental details gases. This step was developed at a variable working pressure (P*) with a
40 % duty cycle and − 800 V bias voltage. For the final step, a-C:H film
2.1. Sample preparation deposition was performed at different working pressures within the 0.1
to 1.3 Pa range. The DLC film was deposited at a fixed 40 % duty cycle
AISI 316 stainless steel and silicon (100) were used as substrates for and applying a bias voltage of − 800 V. The a-C:H films were deposited
the deposition of a-C:H coatings. Samples with 15 mm diameter and 6 using an Ar / (Ar + C2H2) flow ratio near 6 %; this value was chosen due
mm average thickness were cut from a stainless steel bar. The upper to the reported influence of the low gas dilution at less compressive
surfaces of the samples were prepared in a wet sanding process with stress and higher ion penetration depth [8]. The main controllable
different grit sizes varying between 80 and 1200. Subsequently, the process parameters are shown in Table 1.
samples were polished with a 5-μm diamond powder paste, obtaining an
average roughness of 15 nm. Additionally, the substrates were cleaned 2.3. Characterization of coatings
in an ultrasound bath, using degreasing agents, deionized water, and
isopropyl alcohol before being placed in the reactor chamber. After the deposition of the DLC coatings on the substrates, tests were
performed in order to determine the microstructural, mechanical,
2.2. Film deposition tribological, and morphological properties. The microstructural prop­
erties were determined using the Raman spectroscopy and scanning
The DLC films were obtained via the modified PECVD technique electron microscopy (SEM) techniques. With the nano-indentation
combined with an additional cathode system, developed by DIMARE- technique, it was possible to obtain the mechanical properties such as
INPE (Brazil), for greater confinement of the electrons and ions hardness (H) and elastic modulus (E). The tribological characterization
[10–13,20,25,26]. The pulsed-DC power source used has a fixed fre­ allowed determining the coefficient of friction (COF) and wear rate.
quency of 20 kHz with a controllable variable pulse width between 5 and Finally, the scratch and HR-C indentation methods were used to evaluate
20 μs. Additionally, it is possible to choose the waveform amplitude in a the adhesion of the coatings to the AISI 316 substrates. A more detailed
low-voltage range between − 100 and − 1500 V or in a high-voltage description of the experimental techniques is presented below:
range between − 1500 and − 15,000 V. A short positive pulse of The microstructural characteristics of the coatings were evaluated
approximately 40 V is applied after the negative main pulse. This posi­ using the Raman shift spectrum, analyzed between 800 and 2000 cm− 1.
tive pulse favors the removal of molecules or charged particles that This technique is commonly used to study the microstructure of DLC
accumulate on the substrate's surface, modifying the relative potential of films [1–5,7]. A LabRAM Horiban Evolution spectrometer in back
the surface with respect to the plasma and affecting the flow of ions scattering geometry was used to obtain each film spectrum, and the
during the process [11,25,26]. The additional cathode used consists of a excitation was performed with an Ar+ ion laser with a wavelength of
cylindrical-shaped AISI 304 steel grid mesh. This grid can be designed to 514 nm and power of up to 400 mW. The data acquisition time was set at
act as an active screen in a fluctuating potential or as an extra cathode in 30 s, with 3 accumulations for each spectrum. The measurements were
order to increase plasma densification. In this modified PECVD system, performed under ambient temperature and humidity conditions. The
the additional cathode and the sample holder main cathode were spectral shape fitting was performed using two fully symmetrical

2
A. Capote et al. Surface & Coatings Technology 445 (2022) 128716

Fig. 1. P&ID and schematic representation of the pulsed-DC PECVD modified system with an additional cathode.

order to avoid the influence of the silicon (100) substrate or the a-Si
Table 1
interlayer on the measurement.
Parameters used in the cleaning and deposition stages.
The internal compressive stress was determined by measuring the
Process Pressure Bias Duty Time Precursor Mass silicon (100) substrate curvature before and after the a-C:H film depo­
stages (Pa) voltage cycle (min) gases flow
sition with a KLA-Tencor stylus profilometer model AlphaStep D-500.
(-V) (%) (sccm)
The test was performed by setting the stylus speed and force at 0.1 mm/s
Ar cleaning 0.4 600 40 30 Ar 10 and 1 mg (9.8 × 10− 6 N), respectively. A total length of 5 mm was used
a-Si 0.5 800 40 30 Ar + SiH4 2+4
interlayer
for all the samples. The stress values were obtained by calculating the
Transitional P* 800 40 8 Ar + SiH4 2 + (4 radius of curvature using the least square fit method. The substrate
layer + C2H2 → 0) curvature calculated after the coating deposition can be related to the
+ 30 total stress by determining the elastic properties of the substrate mate­
a-C:H P* 800 40 60 Ar + C2H2 2 + 30
rial and by using Stoney's equation [28].
To determine the coefficient of friction and the wear rate of the DLC
Gaussian lines. The data extracted from the Raman shift spectra allow coatings, a CETR UMT-2 tribometer was used, with 4 mm diameter
the determination of the corresponding band areas and peak intensities, zirconia spheres. Before each of the tests, both the samples and the
the band position, and the full width at half maximum of the G band. The spheres were cleaned for 5 min in an ultrasonic bath with acetone. A
thickness of each film was determined from images obtained with SEM- 100 N load cell was used, and the tests were performed in accordance
FEG TESCAN model MIRA3. The film's growth rate was calculated by with ASTM G133–05, under ambient controlled conditions (temperature
dividing the thickness by the a-C:H deposition time. of 293 K and relative humidity of 40 % RH) [29]. The test parameters
The hardness and the elastic modulus of the samples were deter­ were set with an action load of 5 N, displacement speed of 10 mm/s,
mined using the nanoindentation technique. An Anton Paar UNHT3- track size of 5 mm, and test time of 10 min (300 cycles), until obtaining a
STeP E400 nanoindentation tester was used. For the test, a force was stationary coefficient of friction. The volumetric wear rate was deter­
applied with a loading rate of 6 mN/min until reaching a maximum mined by analyzing the difference between the volumes of the zirconia
value of 3 mN, which was kept constant for 5 s, thus allowing the ac­ spheres before and after the test. The volume loss was evaluated using a
commodation of the material. Subsequently, the force was removed at a Veeco/Wyko NT 1100 optical profilometer, which provides high-
rate of 6 mN/min and the diamond probe tip was removed from the resolution three-dimensional surface measurements. The wear rate
sample. The hardness and the elastic modulus were determined from the was determined by dividing the volume worn by multiplying the load by
applied force, the shape of the tip, and the depth of penetration, in the distance traveled.
accordance with the Oliver and Pharr method [27]. The penetration Two experimental techniques were used for the adhesion evaluation,
depth of the indentation was set to be <10 % of each film's thickness, in a scratch test and HR-C indentation. In the scratch test, a CETR UMT-2

3
A. Capote et al. Surface & Coatings Technology 445 (2022) 128716

tribometer with a Rockwell C (HR-C) 120◦ angle diamond tip with a altering the microstructural characteristics of DLC films. Several expla­
radius of curvature of 200 μm was used, in accordance with ASTM nations have been suggested for the displacement of the D and G bands;
C1624–22 [30]. The load was applied progressively until reaching 60 N, these include changes in the size of the clusters, their atomic distribu­
and the indenter displacement rate was set to 0.1 mm/s until it reached a tion, the presence of internal compressive stress, and the chemical na­
5 mm long total displacement. The point where the first crack occurred ture of the bonding [5,33]. A phase transformation towards denser
was labeled critical load Lc1, and the value of the load where substrate atomic packing structure and some types of chemical recombination can
exposure was observed was labeled critical load Lc2. The second tech­ induce tensile stress [34–36]. This may explain the ~25 cm− 1 down­
nique used for the adhesion evaluation was based on the VDI 3198 shifted G band of the a-C:H films obtained, in comparison with the non-
indentation test [31]. The HR-C indentation test was performed in a stressed Raman shift G peak (at 1580 cm− 1) [34,36].
Reicherter Briviskop BVR 187.5H hardness tester, using a conical dia­ The Gaussian fitting of each spectrum allowed identifying the in­
mond tip and a load of 1471 N. The test was performed on the AISI 316 tensity and position of the D and G bands, as well as the full width at half
coated surface, producing a notable plastic deformation in the substrate. maximum of the G band peak (FWHMG). In the Raman spectra of DLC
Analysis of SEM/FEG micrographs of the produced failure, when films, changes in those parameters are related to possible microstruc­
compared with the image standard proposed in the VDI 3198 standard, tural variations of the coatings [32]. The intensity ratio for the D and G
is useful for classifying it as acceptable or unacceptable coating adhesion bands (ID/IG), the G band peak position (ωG), and the FWHM of the G
[31]. band (ΓG), obtained from the fitted parameters of DLC films growth with
variable deposition pressure, are plotted in Fig. 3. These results are
3. Results and discussion typical values for these parameters in high-quality a-C:H, and they are
similar to results obtained in previous studies [8,10,20,34]. In disor­
3.1. Microstructure of the a-C:H films dered a-C:H materials, it is known that the D band is affected by different
carbon structures and is formed whenever no C with sp3 hybridization
In the Raman spectra of DLC films, changes in the position, intensity, sites exist, confirming that with Raman spectroscopy it is generally not
and width of the D and G bands can be related to possible microstruc­ possible to determine the sp3/sp2 concentration ratio or the detailed
tural variations of the coatings [32]. In the Gaussian fitting of the Raman microstructural organization [34,35].
shift spectrum, the graphite band (G), usually with a peak position The ID/IG ratio shows a slight increase at lower deposition pressures.
around 1550 cm− 1, is related to the vibration modes of the atom bonds In amorphous carbon films, the ID/IG ratio is usually a parameter qual­
with sp2 hybridization in the aromatic rings or the olefin chains [7,33]. itatively related to the phase size with sp2 hybridization organized into
On the other hand, the disorder band (D), usually with a peak position aromatic rings, while the FWHMG value is mainly sensitive to structural
around 1350 cm− 1, is related to the interaction forces or breathing bond distortions [5]. The increase of the band's intensity ratio, together
modes of the aromatic rings present in the cluster agglomerations with the shift of the G band's position towards higher wave numbers,
[7,33]. accompanied by a reduction in the FWHM of the G band, is usually
The Raman shift spectra of the films showed that there is no signif­ interpreted in terms of an increase of the sp2 clusters' dominion, either in
icant difference in the spectral profiles of the samples deposited at number or in size [33,34].
different pressures when other deposition parameters are kept constant. On the other hand, the FWHMG values showed a tendency to remain
The typical Raman spectra of films grown with internal pressures be­
tween 0.1 and 1.3 Pa are shown in Fig. 2. The Raman spectra analyzed
for all the films showed a position of the D and G band centered around
1380 cm− 1 and 1555 cm− 1, respectively. In comparison with the normal
peak positions for DLC deposited using the conventional PECVD tech­
nique, in this modified system it's possible to identify that position of
both the D and the G bands are shifted, mainly attributable to the
incorporation of an additional cathode [12]. In this novel technique, the
confinement of ions and electrons increases the plasma density, thus

Fig. 2. Raman shift spectrum for the a-C:H film deposited on Si (100) at 0.4 Pa. Fig. 3. ID/IG intensity ratio, G peak position (ωG) and FWMHG (ΓG) for the
The Gaussian fitting lines show the D and G bands. films deposited on Si (100) as a function of the pressure.

4
A. Capote et al. Surface & Coatings Technology 445 (2022) 128716

constant for all pressures, with a value of around 165 cm− 1. Considering
the observed stability of the FWHMG value, it can be suggested that
increasing the internal pressure in the range evaluated is not associated
with structural disorder formation in the deposited DLC films' structures.
However, in Raman spectra the stress shift must be considered for the
correct assignment of the peaks [36]. Also, the results must be com­
plemented with other solid-state characterization techniques in order to
provide accurate interpretation of the material's structure, atomic
bonding, and network organization [35].
The modified PECVD technique with an additional cathode stands
out because of the possibility of obtaining DLC films with an improved
growth rate at low deposition pressure. The deposition rate was
measured by calculating the average thickness of the coatings deposited
on silicon (100). The cross-section micrographs of each film, such as the
one shown in Fig. 4, allowed differentiating the stages of the deposition
process and their respective thicknesses. Additionally, it was possible to
identify that the morphology of the DLC film growth was apparently a
dense, slightly columnar structure, a characteristic growth form for this
type of coating [37–39]. The growth rate of the a-C:H films as a function
Fig. 5. Growth rate of the a-C:H film deposited on Si (100) as a function of the
of the internal deposition pressure is shown in Fig. 5. A relatively high
deposition pressure.
growth rate was obtained (>17 nm/min), and a linear correlation with
the internal pressure was observed. At low pressures (up to 0.1 Pa), the
modified PECVD system allowed growing DLC films with relatively elastic deformation, high H/E values indicate high wear resistance.
higher deposition rates (about 1 μm/h), whereas a conventional PECVD Depending on the application of the DLC coatings, it is necessary to
system is not suitable for obtaining DLC films at similar pressures due to guarantee a high degree of hardness in order to improve the tribological
the high plasma instability. One possible explanation is that when using properties of the films and an adequate adhesion of the films to the
a conventional system, lower deposition pressure involves a lower deformations produced. A modified H/f.E (coefficient of friction, f)
probability of collision between electrons and molecules, thus reducing criterion can be used as a qualitative way to predict superior anti-wear
the level of dissociation and ionization of the precursor gases. Despite results of hard carbon coatings [34]. Variations in the surface micro-
that, the addition of an extra cathode increases the density of highly roughness and the coatings' surface chemistry have an effect on the
energetic species in the plasma phase (electrons, C+ ions, and radicals), coefficient of friction. The coating's roughness can be improved by
due to the partial confinement [12]. Therefore, in this near collision less reducing the surface diffusion and recrystallization effects during the
scenario, the lowest deposition pressures where it is possible to achieve film's nucleation and growth, when ions and radicals are able to develop
plasma stability allow increasing the ions' average energy while nar­ stronger atomic interlinking bonds with the film's surface, and when
rowing the ions' energy distribution. lower deposition temperature has been achieved [34,35]. Lower local
temperature at the film's atomic-scale surface may be achieved by using
3.2. Mechanical and tribological properties of the a-C:H films a pulsed deposition system, where the heat loss is more efficient and
there is a reduced probability of degrading the deposited carbon mate­
Some mechanical properties, such as hardness (H) and elastic rial [34]. Some transformations from diamond-like to graphite-like can
modulus (E), can be modified by controlling the deposition parameters be leaded due to friction heat; also, film oxidation and surface passiv­
[39,40]. In general, the plastic index H/E ratio serves as a parameter for ation should be considered in order to reach a more detailed wear
identifying the mechanical behavior of this type of material in wear behavior analysis. Moreover, the total compressive stress of a-C:H films
control [40]. While low H/E ratios usually indicate high resistance to is highly relevant, affecting the coating's adherence and anti-wear
behavior.
In the nanoindentation test, the penetration depth was monitored
during the loading and unloading stages. From the load-depth curves, as
shown in Fig. 6, it is possible to quantify the contact area at the
maximum load by using the contact depth (hc) and the material's contact
stiffness by using the initial unloading curve tangential slope (S) [27].
From the indenter geometry and the projected contact area, the hard­
ness, elastic modulus, and plastic index values for the films' growth at
different deposition pressures were obtained and are shown in Table 2.
The films reached their maximum H and E values at the lowest pressure,
27.8 ± 1.2 GPa and 207.3 ± 5.2 GPa, respectively. The hardness and the
elastic modulus of the DLC coatings exhibited a tendency to decrease
with an increase in the deposition pressure. It can be inferred that the
inferior mechanical properties obtained for films deposited at 0.7 Pa or
higher pressures is related to the amorphization trajectory described in
the Raman microstructure characterization. Also, an explanation is that
at higher pressures the neutral species in the plasma phase (radicals and
intact molecules) make an important contribution to the films' mass
growth. This phenomenon is governed by the chemical reactivity of the
radicals and molecules at the C–H or C–C sites [14,15,41]. The ob­
tained DLC films have a significantly higher degree of hardness than the
uncoated AISI 316 stainless steel substrates (~1.5 GPa), reinforcing the
Fig. 4. SEM micrograph 50 kx of the cross section for the a-C:H film deposited
on Si (100) at 0.5 Pa.
relevance of a-C:H as a protective coating for metallic surfaces. Also, it

5
A. Capote et al. Surface & Coatings Technology 445 (2022) 128716

Fig. 6. Penetration curves obtained from nanoindentation tests for a-C:H films
deposited on Si (100).

Table 2
Hardness (H), elastic modulus (E), plastic index parameter (H/E) and indenta­
tion depth of the DLC films deposited on Si (100) as a function of the deposition
pressure.
Pressure H (GPa) E (GPa) H/E Indentation depth
(Pa) (nm)

0.1 27.8 ± 207.3 ± 0.13 ± 92.8 ± 1.5 Fig. 7. Residual compressive stress and average roughness for the a-C:H films
1.2 5.2 0.01
deposited on Si (100) as a function of the pressure.
0.3 23.3 ± 195.3 ± 0.12 ± 94.4 ± 0.8
1.1 6.6 0.01
0.4 22.9 ± 171.6 ± 0.13 ± 101.5 ± 1.0 recrystallization of the graphite [16,34,35].
0.8 4.4 0.01 The coefficient of friction (COF) experienced some transitions
0.5 22.8 ± 179.4 0.13 ± 99.3 ± 1.6
throughout the tribological test. In general, these transitions are natural
±
1.0 4.6 0.01
0.7 15.5 ± 142.5 ± 0.11 ± 116.0 ± 3.7 or induced [43]. These transitions may involve changes in surface
1.3 9.0 0.01 roughness as well as variations in the surface composition or the for­
1.3 15.7 ± 128.4 ± 0.12 ± 124.8 ± 14.0 mation of a third body. Fig. 8 shows the behavior of the coefficient of
1.4 9.3 0.01
friction of the a-C:H films obtained at 0.1, 0.5, and 1.3 Pa, respectively.
For the film's growth at the lowest pressure, it is possible to identify a
was possible to identify high H/E ratios for the DLC coatings. The in­ running-in stage followed by a steady-state friction. This frictional
fluence of the internal pressure indicates that high H/E ratio, up to 0.13 behavior may be described as a case of boundary lubrication. An initial
± 0.01, could be reached in coatings obtained under 0.5 Pa. Therefore, it increase in the COF value due to the contact surface's asperities is
is possible to suggest that the a-C:H films grown at lower pressures tend
to have better wear resistance characteristics.
Determining the total compressive stress is important for improving
the practical adherence of the films, but it should also be considered in
order to achieve a more accurate understanding of the material char­
acterization. Whereas interfacial stress is related to a-C:H film's atomic
mismatch with the a-Si interface, the intrinsic stress and the coating's
thickness make a major contribution to high stress levels [34,35]. By
using the Pauleau formula [42], the intrinsic stress of the films obtained
at the highest pressure is expected to be up to 3 times higher than the
film's growth at 0.1 Pa. The total compressive stress and the surface
roughness of the a-C:H films as a function of the deposition pressure is
shown in Fig. 7. Relatively low compressive stress values, between 0.7
± 0.1 GPa and 1.8 ± 0.2 GPa, were calculated for the studied films. As
expected, higher stress values were observed in the coatings with greater
thickness. The average surface roughness of the a-C:H films deposited on
Si (100) showed that with higher deposition pressures, the roughness
tends to increase. Average surface roughness values between 4.1 ± 0.1
nm and 12.3 ± 1.2 nm were observed (see Fig. 7). The increase of sur­
face roughness could be associated with the larger number of energetic
positive ions impinging on the surface, which can lead to a preferential Fig. 8. Frictional behavior of the a-C:H film deposited on AISI 316 as a function
surface sputtering, higher surface temperature, and the effects of stress of the time.

6
A. Capote et al. Surface & Coatings Technology 445 (2022) 128716

gradually reduced as a result of the smoothening of the surfaces, mainly completely delaminated due to the considerably higher total compres­
due to wear and plastic deformation [43]. For the film's growth at 0.5 sive stress of the film, approximately − 3.5 GPa [10]. To overcome the
and 1.3 Pa, a different frictional behavior was observed. These curve presence of high residual stress on the deposited DLC films, a thin
shapes are usually identified in hard coatings on ceramics. This behavior amorphous silicon interlayer of about 250 nm in thickness has consis­
is related to roughness changes followed by debris layer formation tently been used as an interface between the a-C:H layer and the sub­
[44,45]. The coefficient of friction and the wear rate of the obtained strate surface. Strong chemical bonding between the Si 2p and C 1 s is
films as a function of the deposition pressure are shown in Fig. 9. Low probably responsible for decreasing the film's stress and improving the
COF values, between 0.087 ± 0.002 and 0.122 ± 0.002, were measured adherence of the a-C:H film to the a-Si interlayer [23].
for the coatings. These results showed that samples with higher values of A semi-quantitative measure of adherence of the DLC films was
average surface roughness consistently exhibited the highest COF. evaluated with a scratch test, the determination of the critical loads was
However, the coefficient of friction of the deposited films is relatively performed while considering the evolution of the coefficient of friction,
low and suggests the major advantage of using DLC coatings as solid the received signal measured by an acoustic emission sensor and the
lubricants when compared to the high coefficient of friction (approx. breakpoints of the track's micrographs. The critical load values (Lc1 and
0.6) of polished AISI 316 stainless steel substrates. Lc2) are shown in Fig. 10. The coatings obtained with pressures lower
A wear track was formed on the DLC coatings' surface due to the than 0.5 Pa have critical loads Lc1 and Lc2 >6.4 ± 1.2 N and 19.7 ± 1.4
typical abrasive mechanism at relatively high load conditions [39,46]. N, respectively. By comparison, the films deposited at higher pressures
Three-body abrasive wear can be verified from the presence of a transfer have critical loads lower than 3.0 ± 1.3 N and 2.7 ± 1.1 N. These values,
layer observed in the contact area on the zirconia sphere's surface. Lc1 and Lc2, were low in comparison with the values of the other samples
Larger zirconia volume loss and wear debris were noticed for the films' and were consecutive among themselves; once reaching critical load Lc1,
growth at internal pressures above 0.5 Pa. The highest wear rate was critical load Lc2 was immediately recorded. The lower critical load
observed for the film obtained at the highest pressure, and the volu­ values were observed in the samples obtained at pressures above 0.5 Pa
metric wear was (20.8 ± 1.4) x10− 8 mm3/Nm. This increased wear can were expected, due to the higher total compressive stress of the films, up
be explained by the film's lower hardness and lower H/E ratio, in to 2.6 times higher when compared with the lowest deposition pressure.
addition to the increased COF, which could lead to an insufficient It has previously been shown that the deposition of DLC films via the
transfer layer formation at the DLC coating and zirconia ball interface. modified PECVD technique with an additional cathode allows better
By contrast, the hardest DLC film exhibited lower missing net volume adhesion of the films to the substrate, due to the intense ionic action in a
and more shallow wear scar surface morphology. The lowest wear rate near collision less regime [13]. In this deposition technique measured
was observed for the film grown at 0.1 Pa: the volumetric material loss total compressive stress levels were relatively lower than the total stress
was estimated at around (4.3 ± 1.3) x10− 8 mm3/Nm. Thus it was (≥ − 2.5 GPa) achieved using the r.f. PECVD deposition technique
possible to determine that the coatings with higher hardness values [23,25]. Thus, this novel technique allows the growth of films at low
exhibited lower volumetric wear rates. Additionally, the H/E ratio, or pressures, increasing the mean free path, and as a consequence, a greater
plastic index, can be used as a comparative parameter between DLC ionic penetration depth is reached. In addition, low internal pressures
protective coatings in order to classify their resistance to wear. favor the reduction of residual stresses that end up affecting the adhe­
sion of the films to metallic substrates. The delamination of DLC films
when subjected to mechanical loads generally implies that the coatings
3.3. Adherence of the a-C:H films are not suitable for some industrial applications. The VDI 3198 inden­
tation test was used to qualitatively control the adhesion of the depos­
Obtaining adherent DLC films on metallic surfaces is influenced by ited coatings [30]. The HR-C macro-indentations performed on the
several variables of the deposition process, such as the presence and coated substrates are shown in Fig. 11. When compared to the VDI 3198
strength of compressive residual stresses, the film's microstructure standard, the coatings obtained at pressures lower than 0.5 Pa exhibited
composition, the interlayer bonding, and the physical and chemical an acceptable pattern of failure and therefore an adequate interfacial
properties of the substrate, such as roughness and type of material. adhesion between the films and the substrates, possibly due to the much
Extensive research has been carried out by DIMARE-INPE (Brazil) group lower measured total compressive stress of the films. However, the
[9–13,20,23–25,47]. In previous investigations, DLC deposition on coatings grown at higher pressures exhibited failure patterns that
metallic substrates without the incorporation of an a-Si interlayer

Fig. 9. Coefficient of friction (●) and wear rate (□) for the a-C:H films Fig. 10. Critical load values Lc1 and Lc2 for the a-C:H films deposited on AISI
deposited on AISI 316 as functions of the pressure. 316 as a function of the pressure.

7
A. Capote et al. Surface & Coatings Technology 445 (2022) 128716

Fig. 11. SEM micrographs of HR-C indentation test for the a-C:H films deposited on AISI 316 at variable pressure.

indicate low interfacial adhesion. The qualitative control results are layer was grown with approximately 500 nm in thickness [20].
congruent with the scratch test results, due to the decrease in the value
of critical loads for the coatings deposited at pressures >0.7 Pa. Thus, 4. Conclusions
suggesting a poor adhesion of the coatings to the substrate when sub­
jected to a force or work in the coated substrate surfaces. Thick DLC The influence of low deposition pressure on the microstructural,
coatings (>2.5 μm) with significantly lower stress (about 0.6 GPa) and mechanical, and tribological properties and the adherence of a-C:H
good adherence were obtained previously, using a multilayer structure coatings have not previously been systematically studied. By employing
adding thin a-Si interlayers between the a-C:H layers, where each DLC a modified pulsed-DC PECVD technique with an additional cathode, the

8
A. Capote et al. Surface & Coatings Technology 445 (2022) 128716

characterization of some a-C:H films' properties was reported. The re­ References
sults for coatings grown between 0.1 Pa and 1.3 Pa suggest that this
novel system might be a good alternative for the deposition of a-C:H [1] J. Robertson, Diamond-like amorphous carbon, Mater. Sci. Eng. R 37 (2002)
129–281.
films for mechanical and tribological applications, with relatively simple [2] J. Vetter, 60 years of DLC coatings: historical highlights and technical review of
reactor equipment, lower process operation cost, and easy scaling up. cathodic arc processes to synthesize various DLC types, and their evolution for
However, it should be taken into consideration that the characteristics of industrial applications, Surf. Coat. Technol. 257 (2014) 213–240.
[3] K.A.H. Al Mahmud, M.A. Kalam, H.M. Masjuki, H.M. Mobarak, N.W.M. Zulkifli, An
each reactor system, precursor atomic ratios, and level of purity of the update overview of diamond-like cabron coating in tribology, Crit. Rev. Solid State
precursor gas mixtures, as well as other controllable or non-controllable Mater. Sci. 40 (2) (2015) 90–118.
deposition parameters, exert a major influence on a-C:H characteriza­ [4] A. Ferrari, J. Robertson, Interpretation of Raman spectra of disordered and
amorphous carbon, Phys. Rev. B - Condens. Matter Mater. Phys. 61 (20) (2000)
tion results, especially when compared to other conditions or deposition 14095–14107.
techniques. [5] C. Casiraghi, A.C. Ferrari, J. Robertson, Raman spectroscopy of hydrogenated
The study of the microstructural composition suggests a similarity in amorphous carbons, Phys. Rev. B 72 (2005), 085401-1–085401-14.
[6] V.J. Trava-Airoldi, L.F. Bonetti, G. Capote, L.V. Santos, E.J. Corat, A comparison of
the Raman spectra for all the films, with slight variations in the ID/IG
DLC film properties obtained by r.F. PACVD, IBAD, and enhanced pulsed-DC
ratio, while the G band's position and FWHMG remain constant. Deter­ PACVD, Surf. Coat. Technol. 202 (2007) 549–554.
mining the influence of the internal pressure on the growth rate makes it [7] J. Robertson, Diamond-like carbon films, properties and applications, Compr. Hard
possible to predict the thickness of coatings obtained under equivalent Mater. 3 (2014) 101–139.
[8] G. Capote, G.C. Mastrapa, V.J. Trava-Airoldi, Influence of acetylene precursor
conditions. The characterization of the mechanical properties indicates diluted with argon on the microstructure and the mechanical and tribological
that at low pressure it is possible to obtain a-C:H films with high hard­ properties of a-C: H films deposited via the modified pulsed-DC PECVD method,
ness and elastic modulus, up to 28 GPa and 207 GPa, respectively. The Surf. Coat. Technol. 284 (2015) 145–152.
[9] G. Capote, G.F. Silva, V.J. Trava-Airoldi, Effect of hexane precursor diluted with
effect of the deposition pressure on the films' total compressive stress argon on the adherent diamond-like properties of carbon films on steel surfaces,
was observed by measuring the curvature of the deposited a-C:H coat­ Thin Solid Films 589 (2015) 286–291.
ings; at lower deposition pressures, reduced stress levels were achieved. [10] G. Capote, M.A. Ramírez, P.C.S. da Silva, D.C. Lugo, V.J. Trava-Airoldi,
Improvement of the properties and the adherence of DLC coatings deposited using
The characterization of the total stress of a-C:H films is relevant to the a modified pulsed-DC PECVD technique and an additional cathode, Surf. Coat.
understanding of the practical adherence and tribological behavior of Technol. 308 (2016) 70–79.
the coating material. At all studied deposition pressures, low coefficients [11] D.C. Lugo, P.C. Silva, M.A. Ramirez, E.J.D.M. Pillaca, C.L. Rodrigues, N.
K. Fukumasu, E.J. Corat, M.H. Tabacniks, V.J. Trava-Airoldi, Characterization and
of friction and wear rates were achieved, but the films grown at pres­ tribologic study in high vacuum of hydrogenated DLC films deposited using pulsed
sures lower than 0.7 Pa stand out, due to their greater hardness and DC PECVD system for space applications, Surf. Coat. Technol. 332 (2017) 135–141.
lower COF values. By correlating the plastic index ratios and the coef­ [12] P.C.S. da Silva, M.A.R. Ramos, E.J. Corat, V.J. Trava-Airoldi, DLC films grown on
steel using an innovator active screen system for PECVD technique, Mater.
ficient of friction results, it is possible to identify the influence of these a-
Research 19 (4) (2016) 882–888.
C:H films' properties on superior anti-wear behavior. Finally, the inter­ [13] M.A. Ramírez, P.C. Silva, E.J. Corat, V.J. Trava-Airoldi, An evaluation of the
facial adhesion of the films was determined to be acceptable, except for tribological characteristics of DLC films grown on Inconel alloy 718 using the
the films obtained at higher pressures, due to the lower critical loads and active screen plasma technique in a pulsed-DC PECVD system, Surf. Coat. Technol.
284 (2015) 235–239.
the unacceptable delamination of the films. [14] A. Michelmore, D.A. Steele, J.D. Whittle, J.W. Bradley, R.D. Short, Nanoscale
deposition of chemically functionalised films via plasma polymerisation, RSC Adv.
CRediT authorship contribution statement 3 (2013) 13540–13557.
[15] A. Michelmore, J.D. Whittle, R.D. Short, The importance of ions low pressure
PECVD plasmas, Front. Phys. 3 (3) (2015) 1–5.
A. Capote, and V.J. Trava-Airoldi: Conception and design of study. [16] Y. Lifshitz, G.D. Lempert, E. Grossman, I. Avigal, C. Uzan-Saguy, R. Kalish, J. Kulik,
A. Capote: Acquisition of data: D. Marton, J.W. Rabalais, Growth mechanisms of DLC films from C+ ions:
experimental studies, Diam. Relat. Mater. 4 (1995) 318–323.
A. Capote, G. Capote, and V.J. Trava-Airoldi: Analysis and/or [17] A. von Keudell, T. Schwarz-Selinger, M. Meier, W. Jacob, Direct identification of
interpretation of data. the synergism between methyl radicals and atomic hydrogen during growth of
A. Capote and G. Capote: Drafting the manuscript. amorphous hydrogenated carbon films, Appl. Phys. Lett. 76 (6) (2000) 676–678.
[18] K. Yamamoto, Chemical bond analysis of amorphous carbon films, Vacuum 84
A. Capote, G. Capote, E.J. Corat and V.J. Trava-Airoldi: Revising the (2010) 638–641.
manuscript critically for important intellectual content. [19] R. Hatada, S. Flege, M.N. Ashraf, A. Timmermann, C. Schmid, W. Ensinger, The
influence of preparation conditions on the structural properties and hardness of
diamond-like carbon films, prepared by plasma source ion implantation, Coatings
Declaration of competing interest
10 (4) (2020) 360.
[20] G. Capote, D.C. Lugo, J.M. Gutiérrez, G.C. Mastrapa, V.J. Trava-Airoldi, Effect of
The authors declare the following financial interests/personal re­ amorphous silicon interlayer on the adherence of amorphous hydrogenated carbon
lationships which may be considered as potential competing interests: coatings deposited on several metallic surfaces, Surf. Coat. Technol. 344 (2018)
644–655.
Ariel Capote reports financial support was provided by Conselho [21] R.K. Ghadai, K. Kalita, S.C. Mondal, B.P. Swain, PECVD process parameter
Nacional de Desenvolvimento Científico e Tecnológico (CNPq). optimization: towards increased hardness of diamond-like carbon thin films, Diam.
Relat. Mater. 33 (16) (2018) 1905–1913.
[22] R.K. Ghadai, K. Kalita, Accurate estimation of DLC thin film hardness using genetic
Data availability programming, Int. J. Mater. Res. 111 (6) (2020) 453–462.
[23] G. Capote, L.F. Bonetti, L.V. Santos, V.J. Trava-Airoldi, E.J. Corat, Adherent
Data will be made available on request. amorphous hydrogenated carbon films on metals deposited by plasma enhanced
chemical vapor deposition, Thin Solid Films 516 (12) (2008) 4011–4017.
[24] P.C.S. da Silva, G.V. Martins, L.A. Pereira, E.J. Corat, V.J. Trava-Airoldi, Adherence
Acknowledgments analysis of DLC films grown on AISI M2 steel substrates as a function of silicon
interlayer thickness, Mater. Sci. Forum 802 (2014) 388–391.
[25] G. Capote, E.J. Corat, V.J. Trava-Airoldi, Deposition of amorphous hydrogenated
The authors wish to thank the Fundação de Amparo à Pesquisa do carbon films on steel surfaces through the enhanced asymmetrical modified bipolar
Estado de São Paulo – FAPESP (grant numbers 2019/18572-7 and 2012/ pulsed-DC PECVD method, Surf Coat. Technol. 260 (2014) 133–138.
15857-1), Conselho Nacional de Desenvolvimento Científico e Tec­ [26] CLOROVALE DIAMANTE S/A (CVD VALE), Vladimir Jesus Trava-Airoldi; Evaldo
José Corat; Luís Francisco Bonetti. Processo de revestimentos de superfícies com
nológico – CNPq (grant number 132884/2020-8), Coordenação de
DLC (Diamond Like Carbon) via confinamento parcial de eletrons e ions. BR 10
Aperfeiçoamento de Pessoal de Nível Superior – CAPES, and Ph.D. Ca­ 2016 000262-1, 2016.
rina Barros Mello (INPE-Brazil) for the experimental support. [27] W.C. Oliver, G.M. Pharr, An improved technique for determining hardness and
elastic modulus using load and displacement sensing indentation experiments,
J. Mater. Res. 7 (1992) 1564–1583.
[28] G.G. Stoney, The tension of metallic films deposited by electrolysis, Proc. Nat.
Acad. Sci. United States Amer. 82 (1909) 172–175.

9
A. Capote et al. Surface & Coatings Technology 445 (2022) 128716

[29] American Society for Testing and Materials (ASTM), G133-05: Standard Test plasma enhanced chemical vapor deposition. An analogy with the structure zone
Method for Linearly Reciprocating Ball-on-flat Sliding Wear, ASTM International, model developed for metals, J. Appl. Phys. 92 (2002) 6572–6581.
West Conshohocken, 2016. [39] Xinyu Wang, Xudong Sui, Shuaituo Zhang, Mingming Yan, Jun Yang, Junying Hao,
[30] American Society for Testing and Materials (ASTM), C1624-22: Standard Test Weimin Liu, Effect of deposition pressures on uniformity, mechanical and
Method for Adhesion Strength and Mechanical Failure Modes of Ceramic Coatings tribological properties of thick DLC coatings inside of a long pipe prepared by
by Quantitative Single Point Scratch Testing, ASTM International, West PECVD method, Surf. Coat. Technol. 375 (2019) 150–157.
Conshohocken, 2022. [40] A. Leyland, A. Matthews, On the significance of the H/E ratio in wear control: a
[31] N. Vidakis, A. Antoniadis, N. Bilalis, The VDI 3198 indentation test evaluation of a nanocomposite coating approach to optimized tribological behaviour, Wear 246
reliable qualitative control for layered compounds, J. Mater. Process. Technol. (2000) 1–11.
143–144 (2003) 481–485. [41] A. Von Keudell, T. Schwartz-Selinger, W. Jacob, A. Stevens, Surface reactions of
[32] T. Kasiorowski, J. Linb, P. Soaresa, C.M. Lepienskic, C.A. Neitzkea, G.B. de Souzad, hydrocarbon radicals: suppression of the re-deposition in fusion experiments via a
R.D. Torres, Microstructural and tribological characterization of DLC coatings director liner, J. Nucl. Mater. 290 (2001) 231–237.
deposited by plasma enhanced techniques on steel substrates, Surf. Coat. Technol. [42] Y. Pauleau, Residual stresses in DLC films and adhesion to various substrates, in:
389 (2020), 125615. C. Donnet, A. Erdemir (Eds.), Tribology of Diamond-Like Carbon Films, Springer,
[33] A. Ferrari, J. Robertson, Raman spectroscopy of amorphous, nanostructured, New York, 2008, pp. 102–136.
diamond-like carbon, and nanodiamond, Philos. Trans. Royal Soc. A 362 (2004) [43] M.M. Khonsari, Sahar Ghatrehsamani, Saleh Akbarzadeh, On the running-in nature
2477–2512. of metallic tribo-components: a review, Wear 474-475 (2021) 203871.
[34] S. Neuville, Quantum electronic mechanisms of atomic rearrangements during [44] P.J. Blau, Running-in and other friction transitions, in: P.J. Blau (Ed.), Friction
growth of hard carbon films, Surf. Coat. Technol. 206 (2011) 703–726. Science and Technology: From Concepts to Applications, Boca Raton, CRC Press,
[35] S. Neuville, A. Matthews, A perspective on the optimization of hard carbon and 2008, pp. 315–343.
related coatings for engineering applications, Thin Solid Films 515 (2007) [45] P.J. Blau, On the nature of running-in, Tribol. Int. 38 (2005) 1007–1012.
6619–6653. [46] Yunhai Liu, Lei Chen, Bin Zhang, Zhongyue Cao, Pengfei Shi, Yong Peng,
[36] S. Neuville, Selective carbon material engineering for improved MEMS and NEMS, Ningning Zhou, Junyan Zhang, Linmao Qian, Key role of transfer layer in load
Micromachines 10 (2019) 539–582. dependence of friction on hydrogenated diamond-like carbon films in humid air
[37] J. Foo, T. Aizawa, Wettability control of Nano-columnar dlc thin films via eb- and vacuum, Materials 12 (2019) 1–12.
irradiation, IOP Conf. Ser.: Mater. Sci. Eng. 884 (2020), 012110. [47] G. Capote, J.J. Olaya, V.J. Trava-Airoldi, Adherent amorphous hydrogenated
[38] B. Dumay, E. Finot, M. Theobald, O. Legaie, J. Durand, P. Baclet, J.P. Goudonnet, carbon coatings on steel surfaces deposited by enhanced asymmetrical bipolar
Structure of amorphous hydrogenated carbon films prepared by radio frequency pulsed-DC PECVD method and hexane as precursor, Surf. Coat. Technol. 251
(2014) 276–282.

10

You might also like