You are on page 1of 6

Surface & Coatings Technology 201 (2006) 4424 – 4429

www.elsevier.com/locate/surfcoat

Corrosion performance of the plasma nitrided 316L stainless steel


Linda Gil a,⁎, Sonia Brühl c , Lorena Jiménez a , Ovídio Leon a , Rafael Guevara, Mariana H. Staia b
a
Departamento de Ingeniería Metalúrgica, Universidad Nacional Experimental Politécnica (UNEXPO), Puerto Ordaz, Venezuela
b
Escuela de Ingeniería Metalúrgica y Ciencia de los Materiales, Facultad de Ingeniería, Universidad Central de Venezuela,
Apartado 49141, Caracas 1042-A, Venezuela
c
Universidad Tecnológica Nacional, Facultad Regional Concepción del Uruguay, Argentina
Available online 22 September 2006

Abstract

The AISI 316L stainless steel has been widely used in artificial knee or hip joints, as well as internal fixation devices. It is well known that this
material has a good corrosion resistance and acceptable biocompatibility properties. Ion nitriding is a well established process for steel hardening
that can also be applied to this kind of steels with the aim of enhancing its hardness but without reducing its corrosion resistance. In this work, the
effects of ion nitriding on the corrosion performance of a 3l6L stainless steel was evaluated in a 0.9% sodium chloride solution by using
electrochemical tests such as potentiodynamic polarization and linear polarization in both nitrided and untreated AISI 316L steel condition.
Surface characterization before and after corrosion testing was performed using scanning electron microscopy (SEM) with an energy dispersive X-
ray analysis (EDS). It was shown that an ion nitriding treatment in a 25% N2–75% H2 atmosphere performed at a temperature of 410 °C improves
the surface hardness of the AISI 316L stainless steel. However, under the experimental conditions carried out in this research, the nitrided steel is
as prone to localized corrosion as the untreated one. It is considered that this behavior is mainly due to the presence of CrN, which precipitates
during processing, contributing to the depletion of chromium from the adjacent matrix and leading to a galvanic corrosion mechanism.
© 2006 Published by Elsevier B.V.

Keywords: Plasma nitriding; 316 L stainless steel; Corrosion resistance; Medical implants

1. Introduction stainless steel, AISI 316L is not ferromagnetic which permits the
implanted patients to be examined by using magnetic resonance
In recent years, many attempts have been made to improve the [2]. The austenitic chromium–nickel–molybdenum steel achieves
corrosion properties and biocompatibility of metals and alloys used its “stainless” characterization through the formation of thin and
in orthopedic surgery [1]. The most common materials suitable for adherent chromium rich oxide film [3]. However, in surface defects
orthopedic implants are stainless steel, cobalt and nickel-alloy like fissures or pits the local corrosion rate may be drastically
enriched with chromium or molybdenum, and titanium or titanium increased, due to an incomplete repassivation of the surface caused
alloys. These alloys were chosen because they possess good by lack of oxygen. For implants made of stainless steel in vivo
mechanical properties, sufficient corrosion resistance and biotoler- pitting corrosion has been observed [3]. Corrosion of metallic
ance in implanted environments [1]. There are three types of materials in implants may affect the body tissue by cell reaction to
materials that constitute a prosthesis: the support materials, insuring electrical current, change the pH and release of metallic ion from
mechanical fixation, the friction materials, insuring the sliding of the implant. For stainless steel, the biological environments reacts
articulating surfaces (joint coupling), and the anchorage materials by formation of connective tissue between metal surface and body
permitting the prosthesis to fix to the bone [2]. Support materials tissue [3]; besides that, allergic reaction to nickel and chromium are
are exclusively made of metals such as stainless steels or titanium reported. Although the metallic orthopedic implants may have
alloys [2]. Medical implants that required high strength up to the excellent bulk properties such as ideal strength and elasticity, it has
onset of substrate plastic deformation, such as bone plates or screw, relatively poor surface properties, e.g. poor wear resistance and
are made of materials as AISI 316L stainless steels. Unlike other limited biocompatibility. It is necessary to make a compromise
between bulk properties and surface properties. In the case of hip
⁎ Corresponding author. Tel.: +58 286 9618724; fax: +58 286 9617850. replacement, the wear debris from the implant is one of the essential
E-mail address: lindagil@cantv.net (L. Gil). factors for the aseptic loosening which is a frequent cause of failure
0257-8972/$ - see front matter © 2006 Published by Elsevier B.V.
doi:10.1016/j.surfcoat.2006.08.081
L. Gil et al. / Surface & Coatings Technology 201 (2006) 4424–4429 4425

Table 1 The ion nitriding process was performed using an industrial


Nitriding parameters equipment with a DC pulsed power supply, and a chamber of
Nitriding time 8h 600 mm diameter and 1200 mm high; the nitriding temperature
Temperature 410 °C was measured by means of a thermocouple in contact with the
Pressure 6.5 hPa
Atmosphere 75% H2 + 25% N2
samples and the processing conditions are presented in Table 1.
Voltage 680 V Before nitriding, in order to remove the oxide film from the
Duty cycle ton / toff 70–200 μs surface, the samples were subject to a sputtering process during
Current density ≈ 1 mA cm− 2 2 h using a gas mixture of 60% Ar and 40% H2. The thick
samples were ground and mechanically polished by different
grades of SiC emery papers (240, 600, 800, and 1000), and then
of the prosthetic implants [4]. Life expectancy of an orthopedic manually polished by 1.0 μm and 0.5 μm alumina suspension
implant with mobile parts as a hip implant for instance is and degreased finally. A glyceregia reagent was used for
approximately 10 years. Such a short utile life span is due mainly to revealing the microstructural features. The microstructure of the
the mechanical wear and corrosion caused by the organisms on the untreated and treated samples before and after corrosion testing
mobile parts of the prosthesis, which demand their replacement [5]. was examined by means of an optical microscope. Scanning
Considerable research has been carried out related to bio- electron microscopy (SEM) technique coupled with energy
materials in the natural tissue environments. It is at the bone/ dispersion spectroscopy (EDS) analysis techniques were also
biomaterials interfaces that the biocompatibility of the materials employed. The glancing-angle XRD measurements at 2° in-
becomes the determining factor for the success of the implant, coming angle to the surface were performed with CuKα radi-
both in terms of acceptance of the implant by the organism and its ation in order to identify the microstructure of the nitrided zone.
functional suitability. It is generally accepted that the improvement Knoop microhardness profile measurements (HK0.01) were
of wear, corrosion resistance and biocompatibility of the implants carried out on the modified layer and on the matrix, using a
by surface engineering is an optimal option [5]. Nitriding is, Micromet tester, series 2100 (Buhler Ltd, USA).
nowadays, a well-established industrial practice for improving The electrochemical techniques used in the present work
wear and corrosion resistance of steel components. However, if include the linear polarization and potentiodynamic polariza-
austenitic stainless steels are treated at temperatures generally used tion. The anodic potentiodynamic polarization measurements
for nitriding of low steels or tools steels (about 450 °C or higher),
they suffer a significant decrease of corrosion resistance due to the
precipitation of substantial amounts of chromium nitride, which
depletes chromium from solid solution prejudicing the formation
of the protective film [6]. Nitriding techniques are effective in
improving both surface hardness and corrosion resistance only
when they are performed at temperatures below 450 °C. This has
been correlated to the formation of a metastable single phase,
known as supersaturated or expanded austenite γN [5,6], S phase
[7–10], or m phase [11–13], which provided to a high hardness
and a good corrosion resistance. Many nitriding techniques were
used to produce this phase, from plasma nitriding [6,13–18], ion
been implantation [19,20], glow-discharge nitriding [6], plasma
assisted nitriding [21]. Even if the first report on the so-called S
phase by Ichii et al. [22] dates nearly two decades ago and the
working conditions to form this phase are now outlined, its
structure is still a matter of debate and has not completely clarified.
The phases present in the modified layer depend on both the
physical parameters and the used nitriding technique.
In this work, the effects of ion nitriding on the corrosion
performance of a 3l6L stainless steel was evaluated in a 0.9%
sodium chloride solution at pH 6.3, 37 °C, to simulate the
natural tissue environment, by using electrochemical tests such
as potentiodynamic polarization and linear polarization in both
nitrided and untreated AISI 316L steel conditions.

2. Experimental procedure

AISI 316L stainless steel disc samples were cut from an


annealed AISI 316L steel bar (diameter: 60 mm) and progres- Fig. 1. Optical micrographs showing the etched microstructure of: (a) nitrided
sively polished with SiC emery papers down to 1000 grit level. AISI 316L (b) untreated AISI 316 L (etchant: Glyceregia).
4426 L. Gil et al. / Surface & Coatings Technology 201 (2006) 4424–4429

Table 2 3. Results and discussion


EDS analysis of: (a) nitrided and (b) untreated AISI 316L steel
Element Weight% Atomic% 3.1. Microstructure and morphology
a) Plasma nitrided 316L stainless steel
NK 7.87 25.08 After nitriding treatment, the surface of the sample shows a
Si K 0.72 1.15 peculiar morphology, as shown in Fig. 1a. The surface appears
Cr K 17.33 14.87
to be plasma etched, and this etching (also observed by other
Mn K 1.74 1.41
Fe K 64.53 51.56 authors [5,13,23]) can be ascribed to both the cathodic sput-
Ni K 7.81 5.93 tering performed before the nitriding process and the treatment
itself. The surface of the treated sample, as delineated by this
b) Untreated 316L stainless steel plasma etching, still shows an austenitic structure with the
Si K 0.73 1.25
characteristic twins, as commonly revealed by chemical etching
Cr K 18.23 19.32
Mn K 1.42 1.43 with glyceregia on untreated samples (Fig. 1b). Moreover, as
Fe K 68.49 67.56 indicated by Borgioli et al. [6] the slip bands are delineated
Ni K 11.13 10.44 within the grains of the treated sample, which cannot be
ascribed to the plasma etching and, therefore, they suggested
that the formation of the nitrided layer is accompanied by high
were performed using Gamry PC4/750 Potentiostat/Galvanostat compressive stresses which cause a plastic deformation in the
Model DHC2 (Gamry Instruments, Inc. USA). An electrolyte material. These features of the nitrided surface were reported
solution of NaCl (9 g/l of H20) at pH = 6.3 and a temperature of previously in the literature by other authors [5,13,23,24].
37 °C, was used with the aim to simulate the natural tissue Table 2 shows the EDS average analysis performed on top of
environment. A stabilization period of 3600 s was employed both nitrided and unitrided sample, respectively where it could
before starting the measurement. The electrode potential was be observed that the nitrogen content on the surface of the
raised from − 0.25 mV to 1.6 mV with the scanning rate of nitrided sample is approximately 25 at.%.
0.5 mV/s, and the current that flowed through the coating- In the literature, different investigations have reported that
substrate system was recorded. An Ag/AgCl electrode was used when the nitriding process is carried out between 310 °C and
as the reference electrode and graphite as the counter electrode. 420 °C, the nitrided layer corresponds to a supersaturated solid
It has to be reminded that the experiment was carried in such a solution of nitrogen in iron γ (f.c.c.) also called expanded
way that the electrolyte was in contact only with the coated austenite (γN), which presents high hardness and very good
surface. The contact area in all cases was 0.54 cm2. The linear corrosion resistance [6,13–18].
polarization experiments were carried out by scanning the In Fig. 2, the SEM micrograph of the sample cross-section
potential ± 10 mV about to open circuit potential (Ecorr) at rate of shows the microstructure of the nitrided sample treated at
1 mV s− 1. 410 °C, where two layers are identified whose total thickness is

Fig. 2. SEM micrographs of the nitrided AISI 316L stainless steel cross-section sample and EDS microanalysis of outer and inner layers.
L. Gil et al. / Surface & Coatings Technology 201 (2006) 4424–4429 4427

of approximately 6 μm. In this figure, the average composition


corresponding to each layer of the nitrided zone is also
indicated. The EDS microanalysis of the two layers shows a
different nitrogen concentration between them, the higher
concentration of 22.5 at.% N corresponds to the outer layer
and only 9.3 at.% N was determined for the inner one.
Czerwiec et al. [21] have also reported a bi-layered com-
pound zone and have attributed its existence to the in situ
cleaning by ion bombardment in Ar–H2 plasma prior to the
nitriding process. Typical X-ray diffraction patterns of 410 °C
the plasma nitrided and untreated 316 stainless steel are shown
in Fig 3. Both γ(111) and γ(200) peaks of austenite were observed
in case of the untreated 316 stainless steel sample. However, for Fig. 4. Knoop microhardness profile of AISI 316l L nitrided sample at 410 °C.
the plasma nitrided sample, a set of peaks which do not match
any existing ASTM X-ray diffraction index was observed.
These peaks have been associated with a metastable phase ness test. Near the surface of the nitrided layer, at 1 μm from the
called “expanded austenite, γN” or “S phase”. Both peaks ap- surface, the value obtained which was approximately 17.7 GPa
peared at lower angles when compared to the austenite phase (1800 HKN0.01) corresponds to the average hardness values of
which corresponds to the untreated sample, indicating an the two layers, and could be attributed to both the formation of a
expansion of the lattice, due to the incorporation of nitrogen in hard γN phase, which is a supersaturated solid solution of
the interstitial position of the fcc austenite structure [6–13]. The nitrogen in the austenite, and to the presence of chromium
X-ray diffraction shows that besides the γN phase, a weak peak nitride.
of CrN nitride was detected in the layer.
The hardness depth profile of the sectioned nitrided sample 3.2. Corrosion behavior
shown in Fig. 4 was determined by carrying out Knoop hard-
Fig. 5 shows the potentiodynamic polarization curves of both
nitrided and untreated AISI 316L samples in an aerated solution
of NaCl 9 g /1 (physiological serum) at pH = 6.3 and a
temperature of 37 °C. The anodic polarization curve of the
nitrided AISI 316L stainless steel can be divided into two
regions. In the first region, the dissolution of the nitrided AISI
316L steel was kinetically limited and the anodic current in-
creased slowly with potential, showing a “passive-like” be-
havior. Finally, there is a transpassive region beginning at a
critical potential (Epit) where the rapid increased in the current
value is due to breakdown of the passive film. This phenom-
enon is commonly known as pitting corrosion [25] and the

Fig. 5. Potentiodynamic polarization curves of the nitrided 316L stainless steel


Fig. 3. XRD pattern of the (a) nitrided AISI 316L sample (b) untreated AISI (– · –) and of the untreated 316L stainless steel (–) in aerated neutral 0.9 wt.%
316L sample. NaCl solution.
4428 L. Gil et al. / Surface & Coatings Technology 201 (2006) 4424–4429

Table 3 been demonstrated that the nitriding treatment, carried out in the
Electrochemical parameters estimated from the polarization tests performed in experimental conditions described in the present investigation,
an aerated 0.9 wt.% NaCl solution
did not bring the expected improvement in the corrosion re-
Sample Ecorr Epit icorr Rp Epit − Ecorr sistance of the AISI 316L steel and this system could not be
(V) (V) (nA/cm2) (KΩ) (V)
used for biomedical applications.
Untreated 316L steel −1.65 1.78 49 1.52 3.43 This behavior was mainly due to the formation of CrN in the
Nitrided 316L steel −1.24 2.50 27 72.84 3.74
compound layer, which contributed to the depletion of
chromium from the adjacent matrix. If the chromium content
is below 11–12% in this area, the steel is said to be sensitized
potential at which a rapid increase of the current density occurs and then, when exposed to an aggressive medium, is attacked
is usually termed the “pitting potential” or “breakdown poten- mainly due to a galvanic corrosion mechanism.
tial” (Epit or Ebk). The anodic potentiodynamic polarization This behavior is corroborated by experimental findings
curves obtained in the present investigation were similar to reported previously [13,26] which indicated that the presence of
those previously published for the nitrided 316L steel in mixed phase (nitrides and γN phase) produces average corrosion
chloride-containing solutions [13,14,25,26]. potentials which are similar to those of the original stainless
Table 3 presents the corrosion results as determined from the steel.
above experiments. These results show that the nitrided AISI Pitting tendencies can also be assessed from the values
316L exhibited almost the same corrosion resistance as the corresponding to the “pitting potential” (Epit) indicated in
316L untreated stainless steel. As a result of the nitriding pro- Table 3. If the value of Epit is close to that of Ecorr, little polar-
cess, the corrosion potential of the untreated 316L stainless steel ization is required to initiate the pits formation. Thus, samples
(− 1.65 V) was shifted towards a slightly more noble value which are more prone to exhibit pitting will have a relatively low
(− 1.24 V), the corrosion current density was decreased from Epit − Ecorr difference values. Table 2 summarizes the Epit − Ecorr
49 nA/cm2 to 27 nA/cm2 and the polarization resistance, Rp, difference values for nitrided and untreated samples. The results
was increased from 1.52 to 72.84 (KΩ). Nevertheless, it has presented in this table indicate that the nitrided AISI 316L

Fig. 6. SEM micrographs showing: (a) the corroded surface morphology of the nitrided AISI 316L stainless steel; (b) EDS microanalysis inside a pit.
L. Gil et al. / Surface & Coatings Technology 201 (2006) 4424–4429 4429

sample is as prone to localized corrosion as the untreated sample References


under experimental conditions carried out in this research.
[1] K. Meinert, G. Wolf, Surf. Coat. Technol. 200 (1998) 1148.
3.3. Post-tests analysis of samples [2] Y. Khelfaoui, M. Kerkar, A. Bali, F. Dalar, Surf. Coat. Technol. 200 (2006)
4523.
[3] F. Macionczyk, B. Gerold, R. Thull, Surf. Coat. Technol. 142–144 (2001)
Superficial corroded samples were examined using a scan- 1084.
ning electron microscope and their morphologies are presented [4] R. Hüber, A. Cozza, T.L. Marcondes, R.B. Souza, F.F. Fiori, Surf. Coat.
in Fig. 6. Abundant pitting corrosion occurs for the nitrided Technol. 142–144 (2001) 1078.
316L stainless steel samples. This result is consistent with the [5] X. Xu, Z. Yu, L. Wang, J. Qiang, Z. Hei, Surf. Coat. Technol. 162 (2003)
242.
suggestion that CrN is precipitated in the matrix of γN expanded [6] F. Borgioli, A. Fossati, E. Galvanato, T. Bacci, Surf. Coat.Technol. 200
austenite. Inhomogeneous microstructures enhanced corrosion (2005) 2474.
due to electrolytic cell reaction between second phase particles [7] E. Menthe, K.-T. Rie, J.W. Schultze, S. Simson, Surf. Coat. Technol. 74–75
and the matrix, whereas homogeneous structures are relatively (1995)412.
free of internal cell reactions. Therefore, when CrN precipitates [8] K.L. Dahm, P.A. Dearnley, Surf. Eng. 12 (1996) 61.
[9] E. Menthe, A. Bulak, J. Olfe, A. Zimmermann, K.-T. Rie, Surf. Coat.
result in substantial chromium segregation in 316L stainless Technol. 133–134 (2000) 259.
steel, local galvanic cells will be set up which further enhance [10] T. Bacci, F. Borgioli, E. Galvanetto, G. Pradelli, Surf. Coat. Technol. 139
the pitting corrosion [13–18,26–29]. (2001) 251.
[11] K. Marchev, C.V. Cooper, J.T. Blucher, B.C. Giessen, Surf. Coat. Technol.
4. Conclusions 99 (1998) 225.
[12] K. Marchev, M. Landis, R. Vallerio, C.V. Cooper, B.C. Giessen, Surf.
Coat. Technol. 116–119 (1999) 184.
❖ Plasma source ion nitriding of 316L austenitic stainless steel [13] V. Singh, K. Marchev, C.V. Cooper, E.I. Meletis, Surf. Coat. Technol. 160
at a process temperature of 410 °C for a nitriding time of 8 h (2002) 249.
produced a nitrogen expanded austenite (γN) layer with a [14] M.K. Lei, X.M. Zhu, Surf. Coat. Technol. 193 (2005) 22.
[15] M. Rahman, J. Haisder, M. Hashmi. Surf. Coat. Technol. Article in press,
thickness of 6 μm, with a nitrogen content varying from
doi:10.1016/j.surfcoat.2006.07.179.
approximately 22 to 25 at.%. [16] L. Trabzon, M.C. Igdil, Surf. Coat. Technol. 200 (2006) 4195.
❖ Nevertheless, it has been demonstrated that the nitriding [17] C.K. Lee, H. Shih, J. Mater. Sci. 35 (2000) 2361.
treatment, carried out in the experimental conditions de- [18] S. Kumar, M.J. Baldwin, M. Fewell, et al., Surf. Coat. Technol. 123 (2000)
scribed in the present investigation, did not bring the ex- 29.
pected improvement in the corrosion resistance of the AISI [19] R. Wei, J.J. Vajo, J.N. Matossian, P.J. Wilbur, J.A. Davis, D.L. Williamson,
G.A. Collins, Surf. Coat. Technol. 83 (1996) 235.
steel. [20] G.A. Collins, R. Hutchings, K.T. Short, J. Tendys, X. Li, M. Samandi,
❖ Abundant pitting corrosion occurs for the nitrided 316L Surf. Coat. Technol. 74–75 (1995) 417.
stainless steel samples (Fig. 6). This result indicates that CrN [21] T. Czerwiec, H. He, S. Weber, C. Dong, H. Michek, Surf. Coat. Technol.
is precipitated in the matrix of the γN expanded austenite, 200 (2006) 5289.
contributing to substantial chromium segregation towards [22] K. Ichii, K. Fujimura, T. Takao, Technol. Rep. Kansai Univ. 27 (1986).
[23] W. Liang, X. Xiaolei, X. Jiujun, S. Yaqin, Thin Solid Films 391 (2001) 11.
the grain boundaries and producing, therefore, local galvanic [24] B. Larisch, U. Brusky, H. Spies, Surf. Coat. Technol. 116 (1999) 205.
cells which further enhances the pitting corrosion. [25] D.L. Williamson, P.J. Wilbur, F.R. Fickett, S. Parascandola, in: T. Bell,
K. Akamatsu (Eds.), Stainless Steel 2000:Thermochemical Surface Eng.
Acknowledgments of Stainless Steel, Maney Pub, 2001, p. 333.
[26] H. Dong, Y. Sun, T. Bell, Surf. Coat. Technol. 90 (1997) 91.
[27] M.K. Lei, Z.L. Zhang, J. Vac. Sci. Technol. 15 (1997) 421.
The authors wish to acknowledge the financial support [28] E. Menthe, A. Bulak, J. Olfe, A. Zimmermann, K. Rie, Surf. Coat.
received from Fondo Nacional de Ciencia, Tecnología e Technol. 133–134 (2000) 259.
Innovación (FONACIT) through the project S1-2001000759, [29] S. Mandl, R. Günzel, E. Richter, W. Möler, Surf. Coat. Technol. 100–101
project UCV F-2001000600 and to CDCH-UCV through the (1998) 372.
projects PI 08-17-5120-2003 and P.G. 08-00-5791-2005.

You might also like