1 s2.0 S0043164809001422 Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Wear 267 (2009) 882–889

Contents lists available at ScienceDirect

Wear
journal homepage: www.elsevier.com/locate/wear

Tribological properties of CrSiN-coated 301 stainless steel


under wet and dry conditions
M. Azzi a,b , M. Benkahoul b , J.A. Szpunar a , J.E. Klemberg-Sapieha b,∗ , L. Martinu b
a
Dept. of Mining and Materials Engineering, McGill University, Montreal, Qc, Canada, H3A 2B2
b
Dept. of Engineering Physics, Ecole Polytechnique, Montreal, Qc, Canada, H3C 3A7

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

Article history: The tribological properties of CrSiN coatings were investigated under dry and wet conditions. CrSiN
Received 11 September 2008 films (2 ␮m thick) were deposited on 301 stainless steel substrates by dual magnetron sputtering, using
Received in revised form 12 January 2009 a chromium interface layer (500 nm thick) on the top of plasma nitrided surface. The morphology and
Accepted 12 January 2009
microstructure of the CrSiN films were examined by SEM, EDS and X-ray diffractometry. Dry and wet wear
tests were performed using a linear reciprocating ball-on-flat tribometer. In wet condition, the contact
Keywords:
was immersed in NaCl 1 wt.%, and the sliding wear tests were performed at open circuit potential (OCP),
CrN
and under cathodic and anodic conditions.
CrSiN
Tribocorrosion
CrSiN films were found to exhibit excellent wear resistance under dry condition: the films had no
301 tendency to delaminate after 1800 cycles of sliding wear under 1.6 GPa Hertzian stress. However, under
Wear wet condition, the wear resistance was rather poor: the films detached from the substrate at smaller
Corrosion contact pressures (1.2 GPa). It was demonstrated that the corrosion reactions at the electrode/electrolyte
interface were the reason for the degradation of the tribological properties of these films; in the absence
of corrosion reactions (under cathodic polarization), CrSiN coatings were found to resist much better
sliding wear.
© 2009 Elsevier B.V. All rights reserved.

1. Introduction structure in which nano-crystallites of CrSiN (4–10 nm) are embed-


ded in an amorphous ultra thin layer of silicon nitride. This model
Hard transition metal nitrides, MeN, such as chromium nitride was first proposed by Veprek [11]. However, Benkahoul et al. [7,8]
(CrN) and titanium nitride (TiN), deposited by physical vapour associated this improvement to the solid solution effect of Si. Yoo et
deposition (PVD) and plasma enhanced chemical vapour deposition al. [12] studied the effect of Si addition to CrN coatings on the cor-
(PECVD) techniques, have been the subject of extensive investiga- rosion resistance of CrN/stainless steel coating/substrate system.
tions in the past few decades due to their excellent mechanical They reported higher corrosion resistance and lower porosity with
properties, high wear resistance, and chemical inertness [1–3]. increasing Si/(Cr + Si) ratio.
Navinsek et al. [4] showed that CrN films deposited by PVD Most of the tribological investigations found in the literature on
techniques are potential candidates to replace hard chrome and CrN coatings were performed under dry conditions. However, in
cadmium coatings prepared by electroplating, a process recognized many applications (e.g., aeronautic, automobile, biotechnology and
today as a significant source of environmental pollution worldwide. others) these films are subjected simultaneously to wear and cor-
The tribological properties of CrN films have been widely rosion phenomena, and therefore, it is very important to study the
reported in the literature [5–8]. CrN films exhibit high wear resis- tribological behaviour of these coatings under corrosive environ-
tance mainly due to their high hardness and low friction coefficient. ment. In this paper, the tribological properties of CrN coatings were
The effects of element incorporation (e.g., Si, Ti, Ni) on the proper- systematically investigated under dry and wet conditions, with an
ties of CrN films have also been extensively studied. Many authors emphasis on the effect of corrosion on the wear resistance.
[5,7–10] have reported a significant improvement in the mechani-
cal and tribological characteristics of CrN films with the addition of
2. Experimental
Si. This was mainly attributed to the formation of nano-composite
2.1. Coating deposition

∗ Corresponding author. Tel.: +1 514 340 5747; fax: +1 514 340 3218. CrSiN films were deposited using a pulsed dual magnetron sput-
E-mail address: jsapieha@polymtl.ca (J.E. Klemberg-Sapieha). tering (PDMS) system. The system is equipped with two targets

0043-1648/$ – see front matter © 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.wear.2009.01.027
M. Azzi et al. / Wear 267 (2009) 882–889 883

Table 1
Experimental conditions of the samples pre-treatment, nitriding, and Cr and CrSiN deposition.

Operation Pressure (Pa) Gaseous mixture (sccm) Bias voltage (V) Targets current (A) Process duration (min) Temperature (K)

Ar sputter-cleaning 1 Ar (15) −600 15 573


Nitriding 4.2 N2 (84) −600 240 573
Cr deposition 0.4 Ar (15) −200 ICr : 0.7, ISi : 0 30 523
CrSiN deposition 1.2 Ar (15), N2 (14) −200 ICr : 0.7, ISi : 0.4 160 523

CrSiN layer was deposited by powering the two targets (Cr and
Si) in the presence of Ar and N2 . The pulsing frequency on both
targets was 300 kHz with a duty cycle of 88%. The Cr target cur-
rent (ICr ) and Si target current (ISi ) were maintained at 0.7 A and
0.4 A, respectively, to produce CrSiN films with 2.3 at.% Si, 47.7 at.%
Cr, and 50 at.% N. The experimental conditions of the samples pre-
treatment, nitriding, Cr and CrSiN deposition are summarized in
Table 1.
Fig. 1 schematically shows the architecture of the coated speci-
mens. SS and SS/N4h notations will be used in this text to refer to
the non-treated (just polished) and 4-hour nitrided stainless steel
Fig. 1. Schematic illustration of the architecture of the coated samples. substrates. SS/N4h/CrSiN and SS/CrSiN notations will refer to the
CrSiN-coated stainless steel substrates with nitriding and without
nitriding.
(one of chromium and one of silicon), inclined 20◦ with respect to
the rotating substrate holder axis. The magnetron sources are con- 2.2. Wear tests
nected to a pulsed DC generator (Pinnacle Plus, AE). The distance
between the targets and the substrate is approximately 80 mm. Wear tests were performed using a linear reciprocating ball-on-
The heatable sample holder is connected to an RF power supply plate tribometer illustrated in Fig. 2 and described previously in Ref.
of 13.56 MHz excitation frequency, and the deposition conditions [13]. Using this tribometer, it is possible to perform wear analysis
were adjusted with respect to the self-induced bias voltage. The under dry or wet condition. For dry wear tests, the cell is not filled
base pressure in the deposition chamber was 1.2 × 10−4 Pa, while with electrolyte, while for wet wear tests the contact ball/specimen
during CrSiN deposition, the total pressure of N2 and Ar was is fully immersed in the test solution. A three-electrode type elec-
adjusted to 1.2 Pa. trochemical cell is then established by using the sample as working
Rectangular plates of 301 stainless steel substrates (25 mm × electrode, a coiled platinum wire as counter electrode, and the
25 mm × 1 mm) were first mechanically polished using 1 ␮m alu- standard calomel electrode (SCE) as reference electrode. The three
mina suspension, followed by ultrasonic cleaning in acetone electrodes are connected to a potentiostat (Autolab PGSTAT302,
(15 min) and isopropanol (15 min), before being introduced into the Ecochemie). In this way, it is possible to control the electrochemical
deposition chamber. conditions of the surface subject to wear.
Once placed on the substrate holder, the substrates were first Sliding wear experiments were performed under dry and wet
sputter-cleaned for 15 min in Ar plasma, specifically meant to conditions using an alumina ball of 4.75 mm diameter as counter-
remove the native oxide formed after polishing. Subsequent plasma face, at a normal load of 9 N, and for a sliding distance of 36 m
nitriding was performed for 4 h in pure N2 plasma. Then, Cr bond (9 N, 36 m). This load corresponds to a maximum Hertzian stress
layer was deposited in the presence of pure Ar plasma. Finally, the of 1.2 GPa. In addition, dry wear tests were also performed at (9 N,

Fig. 2. Schematic drawing of the experimental apparatus used for wear testing.
884 M. Azzi et al. / Wear 267 (2009) 882–889

Fig. 3. SEM images of (a) the surface; (b) the cross-section, of SS/N4h/CrSiN.

72 m), (9 N, 144 m), and (22 N, 36 m) in order to emphasize the sig- The surface roughness Ra of the nitrided surface was ∼80 nm
nificant difference in wear behaviour of CrSiN coatings under dry compared to ∼10 nm for the non-treated (polished) surface. The
and wet conditions. The normal load of 22 N corresponds to a max- cross-sectional image shows that CrSiN film has a columnar struc-
imum Hertzian stress of 1.6 GPa. The stroke length was 10 mm and ture. Many authors [14–16] have observed such surface morphology
the sliding frequency was 1 Hz. For wet tests, NaCl 1 wt.% solution after nitriding treatment of austenitic stainless steels. It was sug-
was used as corrosive medium. Dry wear tests were repeated three gested that the formation of a nitrided layer is accompanied by high
times and wet wear tests were repeated two times to validate the compressive stresses due to the increase of the lattice parameter of
results. the stainless steel, which caused a plastic deformation of the mate-
Wet wear tests were performed under open circuit potential rial at the surface. The increase of the lattice parameter is the result
(OCP) condition, and under anodic and cathodic polarization con- of the insertion of nitrogen atoms in the stainless steel matrix.
ditions. In the first case, the OCP of the surface is monitored before, The thickness of the nitrided layer was approximately 6 ␮m, as
during, and after wear testing. Under anodic (cathodic) condition, derived from EDS line-scan analysis of the cross-section.
wear tests are performed under constant anodic (cathodic) poten- XRD patterns acquired at grazing incidence angle from the CrSiN
tial, and the current is measured before, during and after wear film, the Cr bond layer, and the nitrided layer are shown in Fig. 4. The
testing. The difference between the anodic and cathodic conditions pattern of the bare substrate is also included for comparison. The
is that in the former one, the surface is subjected to corrosion during peaks of the deposited layers indicate that the CrSiN film consists
wear, while in the second one, corrosion reactions are suppressed mainly of f.c.c. CrN phase and that the b.c.c. Cr is the main phase in
because of the cathodic protection. In this way, it is possible to inves- the bond layer. For the nitrided layer, the peaks correspond to the
tigate the effect of corrosion reactions on the tribological properties so-called “Expanded austenite, ␥N ” or “S phase” [14–17], which is
of CrSiN coatings. the interstitial solid solution of the f.c.c. ␥-Fe and of nitrogen. It is

2.3. Surface characterization

Field emission scanning electron microscope (FESEM, Philips


XL30) equipped with an energy dispersive spectrometer (EDS)
was used to assess the surface morphology and to perform cross-
section analysis of the as-deposited coatings. A Rigaku Rotaflex
X-ray diffractometer (XRD) was used to characterize the crystal
structure of the nitrided surfaces and the deposited layers (Cr and
CrSiN). XRD spectra were acquired at a glancing angle of 1.5◦ , using
Cu K␣ radiation (1.54 nm wavelength), under 40 kV voltage, and
40 mA current. Surface roughness (Ra ) of the coatings and wear
track profiles were measured by the Sloan Dektak II profilometer.

3. Results and discussion

In this section, the characterization of the as-deposited coatings


is firstly presented, and then the performance of the coatings under
dry and wet conditions is analysed and discussed.

3.1. Characterization of the as-deposited coatings

SEM images of the surface and cross-section of SS/N4h/CrSiN


are shown in Fig. 3. The surface appears to be etched, and the grain
boundaries of the stainless steel substrate are revealed. This etch-
ing is due to the nitriding process used to increase the substrate
hardness prior to deposition. In fact, the surface of the stainless Fig. 4. XRD patterns of (a) CrSiN film; (b) Cr bond layer; (c) nitrided layer; (d) 301
steel substrates exhibited the same morphology after nitriding. substrate.
M. Azzi et al. / Wear 267 (2009) 882–889 885

Table 2
Material loss, w, and wear track depth, d, measured on SS, SS/N4h after 36 m (1800 cycles) sliding distance in dry condition. Stroke length: 10 mm, frequency: 1 Hz, ( is the
standard deviation).

Normal load Material loss, w, ×105 ␮m3 () Wear track depth, d, ␮m ()

SS SS/N4h SS SS/N4h

9N 930 (14.8) 6.6 (0.4) 16.2 (1.3) 0.82 (0.04)


22 N 1620 (49.5) 23.8 (2.1) 28.7 (1.6) 1.91 (0.15)

sented in this table. It is clearly seen that nitriding improves ∼140


times the dry wear compared to the bare substrate. Such large dif-
ference is coincident with the significant increase of the surface
hardness after nitriding; the surface hardness was 15 GPa for the
nitrided surface compared to 5 GPa for the bare surface.

3.2.2. CrSiN-coated substrates (SS/N4h/CrSiN)


The average values of material loss and wear track depth mea-
sured on SS/N4h/CrSiN at different sliding distances (36 m, 72 m,
and 144 m) and different normal loads (9 N and 22 N) are presented
in Fig. 5. The depth of the wear track after 144 m of sliding at 9 N was
0.65 ␮m indicating that the CrSiN film did not delaminate. Even at
very high load (22 N), the film had no tendency to delaminate and
the wear track depth was almost 1 ␮m after 36 m of sliding wear. On
the other hand, the wear rate (material loss per unit sliding distance,
␮m3 /m) significantly decreases with increasing sliding distance; at
36 m, the wear rate was 5830 ␮m3 /m compared to 3540 ␮m3 /m
after 144 m. This is mainly due to the existence of a running-in
Fig. 5. Wear track depth (␮m) and material loss (␮m3 ) measured on SS/N4h/CrSiN period during which a transition in surface roughness and wear
at different sliding distances and normal loads under dry condition. Stroke length: rate from initial state to those in a steady state take place [19].
10 mm, frequency: 1 Hz.
The SEM images of the wear track after 36 m sliding distance at
9 N and 22 N are shown in Fig. 6. The white particles inside the track
called expanded austenite since the presence of nitrogen expands are wear debris formed during sliding. EDS analysis showed the
the lattice of the f.c.c. ␥-Fe. Several authors [14–18] studied the presence of oxygen in these debris which indicates that the CrSiN
crystalline structure of this phase using XRD. They all reported a film is oxidized during sliding wear in agreement with Ref. [7].
significant decrease in the diffraction angles of the austenitic struc-
ture after nitriding: ␥(1 1 1) peak was observed to shift from 43.5◦ 3.3. Wet wear tests
to approximately 40◦ , and the ␥(2 0 0) peak from 50.8◦ to approxi-
mately 46◦ . 3.3.1. Wear tests at open circuit potential
In these experiments, the evolution of the OCP was monitored
3.2. Dry wear tests before, during, and after wear testing. Fig. 7-a shows the evolu-
tion of the OCP on SS/N4h/CrSiN and SS. Two coated samples from
3.2.1. Nitrided substrates (SS/N4h) different deposition batches (under the same conditions) were
The average values of material loss, w, (in ␮m3 ) and wear track tested to validate the results. For the bare substrate SS, the OCP
depth, d, (in ␮m), measured on SS and SS/N4h after 36 m (1800 dropped immediately to a value of −375 mV when the sliding wear
cycles) sliding distance at two different normal loads (9 N and 22 N), started. When rubbing ceased, the OCP increased and progressively
are summarized in Table 2. The standard deviation, , is also pre- returned to its original value. This corresponds to the depassivation

Fig. 6. Wear track of SS/N4h/CrSiN after 36 m (1800 cycles) dry sliding distance at (a) 9 N; (b) 22 N normal load.
886 M. Azzi et al. / Wear 267 (2009) 882–889

Fig. 7. (a) Evolution of the OCP before, during and after wear tests on SS and SS/N4h/CrSiN. Normal load: 9 N, frequency: 1 Hz; (b) SEM images of the wear track of SS/N4h/CrSiN
after the wear test.

and repassivation of the 301 stainless steel during and after sliding, effect of corrosion on the tribological properties of CrSiN films. The
respectively. evolution of the current during the anodic wear experiment (at
For SS/N4h/CrSiN, the OCP progressively dropped to reach 300 mV) is shown in Fig. 8-a. Before rubbing, the current exponen-
approximately −375 mV, namely the same potential measured on tially decayed due to the improvement in the corrosion properties
SS. This indicates that the alumina ball went through the CrSiN of the passive layer that is formed at the electrode/electrolyte inter-
film and was rubbing on the 301 substrate. When rubbing ceased, face. At the onset of rubbing, the current increased almost one order
the OCP first steeply increased and then returned to its original of magnitude. Then, at around 500 s, the current started to increase,
value. The SEM image of the wear track after the tribocorrosion and it reached approximately 4 × 10−5 A at 700 s. Subsequently,
test confirmed the complete detachment of the CrSiN film from the the current continued to increase and it reached a maximum of
substrate inside the wear track, as shown in Fig. 7-b. 2 × 10−4 A at approximately 1600 s. When rubbing ceased, the cur-
The average values of material loss, w, and wear track depth, d, rent remained high and it did not return to its original value before
measured on SS, SS/N4h, and SS/N4h/CrSiN after 36 m (1800 cycles) wear. The SEM image of the worn area clearly shows the delamina-
sliding distance in wet environment are summarized in Table 3. tion of the CrSiN film and the presence of pits inside the wear track,
The material loss of Material loss, w, ×105 ␮m3 () SS decreased as shown in Fig. 8-b.
significantly (∼26 times) compared to the result of the dry wear The current increase at the onset of rubbing is related to
test (see Table 2). This may be attributed to a lower temperature at the destruction of the passivation layer present at the spec-
the sliding contact in the presence of an electrolyte as a result of a imen/electrolyte interface. Mendibide et al. [20] reported the
faster dissipation of frictional heat. In contrary, the material loss of presence of a p-type semiconductor thin oxide film on CrN films.
SS/N4h/CrSiN significantly increased (∼63 times) in the presence The second increase is associated with the delamination of the
of electrolyte due to a synergistic effect of corrosion and wear as it CrSiN film and the exposure of the 301 substrate to the rubbing
will be demonstrated in the next section. In the case of the nitrided action of the ball. Further increase in the current during sliding wear
substrate SS/N4h, the material loss was slightly higher compared could be the results of the formation of pits. Indeed, the current did
to the dry wear results, possibly due to a minor synergistic effect of not decrease when the rubbing ceased, indicating the presence of
corrosion and wear. active sites (pits) on the surfaces which were responsible for the
high current even in the absence of wear.
3.3.2. Wear tests at controlled potentials The evolution of the current measured during the cathodic wear
In these experiments, the potential of the surface subject to wear experiment (at −800 mV) is shown in Fig. 9-a. Before rubbing, the
was held constant and the evolution of the current was monitored current was approximately 8 × 10−5 A. At the onset of rubbing, the
before, during, and after wear testing. At anodic potentials (higher current increased to reach 1.1 × 10−4 A. It remained relatively sta-
than the OCP), corrosion and wear processes are involved in the ble during the entire wear experiment, and then dropped when
degradation of the surface, whereas at cathodic potentials (lower the rubbing ceased. SEM image of the wear track shows that the
than the OCP), only mechanical wear is involved since the corrosion CrSiN film did not delaminate from the substrate (see Fig. 9-b).
reactions are suppressed by the cathodic potential. The increase in the cathodic current at the onset of rubbing could
Wear tests were performed on SS/N4h/CrSiN at 300 mV (anodic also be the result of the removal of the oxide layer formed on the
potential) and at −800 mV (cathodic potential) to investigate the surface.

Table 3
Material loss, w, and wear track depth, d, measured on SS, SS/N4h, and SS/N4h/CrSiN after 36 m (1800 cycles) sliding distance in NaCl 1 wt.% at 9 N normal load. Stroke length:
10 mm, frequency: 1 Hz, ( is the standard deviation).

Material loss, w, ×105 ␮m3 () Wear track depth, d, ␮m ()

SS SS/N4h SS/N4h/CrSiN SS SS/N4h SS/N4h/CrSiN

35.6 (4.5) 8.2 (0.5) 132 (11.6) 2.6 (0.08) 1.2 (0.1) 6.4 (0.5)
M. Azzi et al. / Wear 267 (2009) 882–889 887

Fig. 8. (a) Evolution of the anodic current on SS/N4h/CrSiN before, during and after wear experiment. Applied voltage: 300 mV, normal load: 9 N, frequency: 1 Hz; (b) SEM
image of the wear track of SS/N4h/CrSiN after the wear test.

The result of these two tests proved that the corrosion reactions surface oxide layer was continuously removed and, therefore, the
at the electrode/electrolyte interface were responsible for the dete- CrSiN film was continuously repassivating. This process generates
rioration of the tribological properties of CrSiN films; in the absence electrons and it consequently lowers the potential of the surface.
of such reactions (under cathodic polarization), these films resist The stabilization of the potential during the experiment indicates
much better sliding wear. an equilibrium condition between the oxidation of the CrSiN inside
the wear track and the reduction of the oxygen outside of the wear
3.3.3. Effect of surface nitriding track.
Wear tests at OCP were also performed on the same coating When rubbing ceased, the potential progressively increased
deposited without substrate nitriding (SS/CrSiN). The evolution of towards higher potentials due to the formation of the passive layer
the OCP during wear testing was different from the one obtained on the CrSiN surface. This is in good agreement with the results of
on SS/N4h/CrSiN, as shown in Fig. 10-a. The OCP gradually dropped Barril et al. [21] who conducted wear experiments at OCP on TiN-
when sliding wear started and reached a steady-state value of about coated stainless steel; they also observed a decrease of the potential
−250 mV, compared to −375 mV for SS/N4h/CrSiN. The SEM image at the onset of rubbing and an increase of the potential when the
of the wear track after 1800 wear cycles (shown in Fig. 10-b) clearly rubbing stopped. They explained these observations by the destruc-
shows that the CrSiN film was not detached from the substrate and tion and consecutive growth of the oxide layer formed on the TiN
it can thus explain the higher OCP measured during rubbing. film in aqueous solutions.
In the case of SS/CrSiN, the drop of the OCP during rubbing is The above results demonstrate that nitriding (prior to deposi-
due to the depassivation of the wear track area. During rubbing, the tion) significantly reduces the wear resistance of CrSiN coatings in

Fig. 9. (a) Evolution of the cathodic current of SS/N4h/CrSiN before, during and after wear experiment. Applied voltage: −800 mV, normal load: 9 N, frequency: 1 Hz; (b) SEM
image of the wear track of SS/N4h/CrSiN after the wear test.
888 M. Azzi et al. / Wear 267 (2009) 882–889

Fig. 10. (a) Evolution of the potential before, during, and after wear testing on SS, and SS/CrSiN. Normal load: 9 N, frequency: 1 Hz; (b) SEM image of the wear track of SS/CrSiN
after the wear test.

Fig. 11. Morphology of the CrSiN films; (a) and (b) with substrate nitriding; (c) and (d) without substrate nitriding; (a) and (c) at low magnification; (b) and (d) at high
magnification.

the NaCl 1 wt.% solution. Nitriding process etches the substrate and wear) of defects that possibly contributed to the infiltration of the
increases the surface roughness. This morphology is transferred liquid, and consequently led to the destruction of the CrSiN film.
to the deposited films and the grain boundaries of the substrate This explains why SS/CrSiN (without nitriding) exhibited better wet
also become boundaries in the CrSiN deposited film as shown in wear resistance compared to SS/N4h/CrSiN.
Fig. 11-a and -b. The morphology of the same film, without nitrid-
ing treatment of the substrate, is shown in Fig. 11-c and -d. It is 4. Conclusions
clearly seen that without nitriding the film is more uniform and
has lower surface roughness. The tribological properties of CrSiN coatings deposited on
For SS/N4h/CrSiN, the discontinuity in the film at the grain 301 stainless steel substrates were investigated under dry and
boundaries may be the reason for the generation (during sliding wet conditions. These coatings exhibited excellent dry wear
M. Azzi et al. / Wear 267 (2009) 882–889 889

resistance; wear rate values of 1.15 × 10−6 mm3 N−1 m−1 and References
2.94 × 10−6 mm3 N−1 m−1 are obtained for SS/N4h/CrSiN and
SS/N4h, respectively. [1] H.C. Barshilia, B. Deepthi, K.S. Rajam, Surf. Coat. Technol. 201 (2007)
9468.
Surface nitriding of 301 stainless steel substrates significantly [2] V.K.W. Grips, H.C. Barshilia, V.E. Selvi, K.S. Kalavati, Rajam, Thin Solid Films 514
increased the surface hardness from ∼5 GPa to ∼15 GPa. Further- (2006) 204.
more, nitriding was found to etch the stainless steel surface; this [3] R.F. Zhang, S. Veprek, Acta Mater. 55 (2007) 4615.
[4] B. Navinsek, P. Panjan, I. Milosev, Surf. Coat. Technol. 116–119 (1999) 476.
is due to the compressive stresses generated by the insertion of [5] D. Mercs, N. Bonasso, S. Naamane, J.M. Bordes, C. Coddet, Surf. Coat. Technol.
nitrogen inside the austenitic structure of the stainless steel. 200 (2005) 403.
It was demonstrated that the corrosion reactions at the elec- [6] W.H. Zhang, J.H. Hsieh, Surf. Coat. Technol. 130 (2000) 240.
[7] M. Benkahoul, P. Robin, S.C. Gujrathi, L. Martinu, J.E. Klemberg-Sapieha, Surf.
trode/electrolyte interface are responsible for the deterioration Coat. Technol. 202 (2008) 3975.
of the wear resistance of the CrSiN coatings. When wear was [8] M. Benkahoul, P. Robin, L. Martinu, J.E. Klemberg-Sapieha, Surf. Coat. Technol.
accompanied by corrosion (under anodic condition) the CrSiN film 203 (2009) 934.
[9] J.H. Park, W.S. Chung, Y.R. Cho, K.H. Kim, Surf. Coat. Technol. 188–189 (2004)
delaminated from the wear track, whereas when the corrosion reac- 425.
tions were suppressed (under cathodic condition) the CrSiN film did [10] I.W. Park, D.S. Kang, J.J. Moore, S.C. Kwon, J.J. Rh, K.H. Kim, Surf. Coat. Technol.
not detach from the wear track. 201 (2007) 5223.
[11] S. Veprek, J. Vac. Sci. Technol. A 17 (1999) 2401.
Surface nitriding decreased significantly the wear resistance of
[12] Y.H. Yoo, J.H. Hong, J.G. Kim, H.Y. Lee, J.G. Han, Surf. Coat. Technol. 210 (2007)
the CrSiN coatings under wet conditions. This results from rather 9518.
peculiar morphology of the nitrided surface prior to deposition. The [13] M. Azzi, J.A. Szpunar, Biomol. Eng. 24 (2007) 443–446.
boundaries created in the coating may be the reason for the gener- [14] F. Borgioli, A. Fossati, E. Galvanetto, T. Bacci, G. Pradelli, Surf. Coat. Technol. 200
(2006) 5505.
ation (during sliding wear) of defects in the film, which contribute [15] F. Borgioli, A. Fossati, E. Galvanetto, T. Bacci, Surf. Coat. Technol. 200 (2005)
to the infiltration of the liquid, leading to a higher corrosion rate, 2474.
and consequently to a reduced wear resistance of the CrSiN film. [16] A. Fossati, F. Borgioli, E. Galvanetto, T. Bacci, Surf. Coat. Technol. 200 (2006)
3511.
[17] N. Mingolo, A.P. Tschiptschin, C.E. Pinedo, Surf. Coat. Technol. 201 (2006) 4215.
Acknowledgements [18] V.G. Gavriljuk, H. Berns, High Nitrogen Steels—Structure, Properties, Manufac-
ture, Applications, Springer-Verlag, 1999.
[19] P.J. Blau, Friction Science and Technology, Decker, New York, 1996.
The authors wish to thank Drs. Philippe Robin and Srinivasan [20] C. Mendibide, P. Steyer, J.-P. Millet, Surf. Coat. Technol. 200 (2005) 109.
Guruvenket for helpful discussions. This work has been supported [21] S. Barril, S. Mischler, D. Landolt, Tribol. Int. 34 (2001) 599.
by NSERC and CRIAQ within the CRDPJ328038-05 project.

You might also like