Abstract
Physical vapor deposition (PVD) multilayered coatings with titanium nitride and chromium nitride top layers were deposited on UNS S17400 alloy in an attempt to improve the corrosion and corrosion-wear resistance of this stainless steel in corrosive environments. The coatings were produced in an industrial chamber by cathodic arc PVD on heat-treated and mechanically polished stainless steel specimens. The microstructures of the substrates and coatings were characterized by X-ray diffraction and scanning electron microscope equipped with an energy-dispersive X-ray spectroscopy system. To evaluate the corrosion and corrosion-wear resistance, reciprocating-sliding tribometer and electrochemical tests were conducted in 3.5% NaCl solution. The results showed that nitride coatings possess, in general, better corrosion and corrosion-wear resistance compared with bare S17400 substrates. Specimens with CrN top coating revealed a typical compact structure and superior corrosion resistance compared with substrate and TiN top coating. However, the sliding motion damaged the surface with some microcracks on the coating, which act as the diffusion channels for NaCl solution; both TiN and CrN top coats experienced approximately similar behavior in corrosion-wear open-circuit potential testing.
1 Introduction
Corrosion and corrosion-wear resistance are among the most important factors in choosing stainless steels for engineering applications. However, chemical composition is not the only effective factor in these properties; other factors such as coating and surface treatment, heat treatment, and manufacturing processes are also important (Feng et al., 2003).
UNS S17400 is a precipitation-hardening martensitic stainless steel that provides high strength, high toughness, and moderate corrosion resistance (Chen et al., 2012; Hu et al., 2017). This alloy has been used for a variety of applications such as heavy load components, valves, fasteners, gears, steam turbine shafts, and propeller shafts (Lin et al., 2012). However, wear and corrosion resistance of the steel in chloride environments is not sufficiently high to ensure prolonged performance. Stress corrosion cracking and pitting corrosion have been reported for this alloy (Mazur, 2008; Liu et al., 2013). Recently, researchers have studied the effect of plasma nitrocarburizing and HVOF spraying to improve hardness and wear resistance of UNS S17400. They reported that this method could improve wear resistance by reducing the coefficient of friction and increasing surface hardness (Liu et al., 2013; Park et al., 2013).
A thin layer of coating is considered effective for the improvement of material surface properties, and physical vapor deposition (PVD) coatings are widely used to deposit wear-resistant hard coatings for industrial applications. Deposition of nitride coatings such as TiN and CrN by PVD technology on steel and cemented carbide substrates has been well developed to improve surface properties (Yashar & Sproul, 1999; Surviliene, 2004; Stueber, 2009; Song et al., 2013: Ebrahimzadeh & Ashrafizadeh, 2014); the results of previous investigations revealed that chromium and titanium nitride coatings greatly enhance the corrosion and wear resistance of stainless steel (Rossia et al., 1999; Shan et al., 2013; Chen et al., 2017). In nitrides, nitrogen has a significant role in corrosion resistance; the presence of nitrogen helps stabilize corrosion potential, which leads to an increase in the ability of the surface for passivation. After the local failure of the passive layer, desorption of the attacking anions by segregated nitrogen provides a condition in which nitrogen favors repassivation (Grabke, 1996; Lavigne et al., 2011). Gilewicz et al. (2016) reported that multilayer structures, formed by alternating deposition, are more effective on corrosion and wear resistance of coating. Cegil et al. (2014) found that the incorporation of aluminum into the cubic crystalline structure greatly enhances the corrosion resistance and hardness of nitride coatings.
Sedimentation of nitride layers by PVD technology is one of the effective methods to increase the corrosion and wear resistance of UNS S17400 stainless steel, which has not been considered by previous researchers. In cases where this stainless steel is used in corrosive and abrasive conditions, such as typical shafts, coating can improve corrosion and wear resistance. In the present work, TiN/TiAlN and CrN/CrAlN multilayered coatings were deposited by an industrial cathodic arc PVD on UNS S17400 stainless steel. The corrosion and corrosion-wear behavior of the deposited coatings were tested in a 3.5% NaCl solution and compared with uncoated martensitic stainless steel.
2 Materials and methods
2.1 Test materials and coating technique
UNS S17400 martensitic stainless steel bar (0.02% C, 16.46% Cr, 3.93% Ni, 0.94% Mn, 0.2% Nb, 3.24% Cu, and balanced Fe in wt%) was machined to 12 mm diameter and 4 mm length specimens as the substrate in this study. Two coatings were deposited by cathodic arc physical vapor deposition in an industrial chamber. After cleaning in an ultrasonic bath in acetone, the specimens were fixed in the chamber, evacuated and exposed to argon ions bombardment at 2 kV for 30 min for sputter cleaning of the surface. This treatment preheated the samples to above 200°C and ensured the surface cleanliness necessary for vapor deposition. The deposition process was started according to the parameters shown in Table 1.
Operational parameters for PVD coatings.
Substrate temperature | Target bias | Coating time | Specimen bias | Target purity | Target | Gas composition |
---|---|---|---|---|---|---|
200°C | 30 V (150 A) | 80 min | 600 V | Purity >99.5 wt% | Ti, Al, Cr | Ar (99.99%), N2 (99.99%) |
On some specimens, at first, TiN with a thickness of approximately 0.6 μm was deposited as an intermediate layer, and then a TiAlN layer and another 0.6 μm TiN as a top coat was added to provide a total thickness of 2 μm. In other specimens, CrN with a thickness of approximately 0.3 μm was deposited at first, followed by the CrAlN layer and, finally 0.2 μm CrN top coats to provide a total thickness of 2 μm. Multilayer coating was performed by switching the cathodes of different targets on and off (Ti, Al, and Cr). After cooling in the chamber, the specimens were removed, cleaned by ethanol, and quickly dried. For metallography, the coated specimens were cut and mounted vertically, then ground progressively down to 2500 grade sand paper and ultrasonically cleaned in acetone.
2.2 Characterization
The phase structures were studied by an X-ray diffractometer using CuKα radiation (λ=0.154 nm), with grazing incidence angle of 2°. The scanning angle ranged from 15° to 90° at a time step of 1 s with 0.1° step size. The PVD process commonly leads to a preferred growth of crystallography planes due to selective nucleation of crystallites, normally referred to as crystallographic texture (Matei et al., 2015). Quantitative estimation of the texture coefficient [TC(hkl)] can be expressed by Equation 1:
where I(hkl) is the measured relative intensity of the (hkl) line, I0(hkl) is the intensity assigned to the same (hkl) planes of random orientation, or according to ICDD, and n is the number of peaks used to recognize the TC. A texture coefficient of greater than one indicates the abundance of crystallites oriented with (hkl) direction normal to the coating surface (Abadias, 2008).
Crystal lattice parameter described with the Nelson–Riley equation and the least squares method using all four fundamental diffraction data from each X-ray pattern as shown in Equation 2:
where θ is the diffraction angle and f(θ) is Nelson–Riley function. The lattice parameters obtained are plotted against cos2, cos2θ/sinθ and the Nelson–Riley function, and the extrapolation of the curves to cos2=0 will provide the true value of the lattice parameter.
Lattice spacing is obtained by the Bragg relationship according to Equations 3 and 4:
For the cubic system,
where λ is the wavelength of incident radiation, d is the lattice spacing, θ is the diffraction angle, a is the lattice parameter and h, k, l are the Miller indices of the crystal plane (Yang et al., 2017). The coating surface, wear track morphologies, and specimen cross-section were observed by a field emission scanning electron microscope and the surface roughness was measured according to DIN EN ISO 4287-2010.
2.3 Corrosion and corrosion-wear test
Before the tests, the specimens were ultrasonically cleaned for 20 min, and 3.5% NaCl was used as an electrolyte in corrosion-wear and corrosion tests. Corrosion-wear tests were carried out in a ball-on-disk tribometer with a three-electrode system. Ag/AgCl electrodes as the reference electrode, Pt as the auxiliary electrode and specimen (coated and uncoated UNS S17400 stainless steel) as the working electrode were used in the tests. All specimens were soaking in 3.5% NaCl for 1 h to fix the open-circuit potential (OCP) before the test was begun. To start the test, the samples were placed inside the tribometer containing the electrolyte (Figure 1) and OCP was recorded for 2000 s before the sliding. A normal load of 5 N was applied on the SiC ball with 7 mm stroke length and the wear cycles of 0.5 Hz. The OCP recording was continued for 2000 s after the sliding was ended.

Setup of corrosion-wear test.
In potentiodynamic polarization tests, the specimens were connected to a working electrode, and saturated calomel electrode (SCE) and pure graphite rods were used for reference and counter-electrodes, respectively. A total of 1 cm2 of the specimen surface was in contact with the electrolyte and the other directions were covered by an insulating layer. The potentiodynamic polarization test was carried out using PARSTAT model 2273 potentiostat with potentials in the range –250 mVOCP to 250 mVSCE with a scan rate of 0.1 mV/s at room temperature (25°C). All tests were carried out at ambient temperature. To obtain the polarization resistance, the Stern–Geary relationship near the OCP potential was used, as expressed by Equation 5.
where Rp (cm2 Ω) is the polarization resistance, βa is the anodic Tafel slope, βc is the cathodic Tafel slope, and icorr (A/cm2) is the corrosion current density (Stern, 1958; Poursaee, 2010). On the assumption that the coating is electrochemically inert at low anodic overpotential, the porosity of the coating is estimated using Equation 6 (Creus et al., 2000).
where P is the total coating porosity rate, RpS and Rp are the polarization resistance of the substrate and coated steel, respectively. ΔEcorr is the difference potential between the free corrosion potentials of the coated steel and uncoated substrate, and βa is the anodic Tafel slope for the substrate.
3 Results and discussion
3.1 Structure and analysis of the coatings
X-ray diffraction patterns of typical coatings, produced by grazing angle technique, are shown in Figure 2 and the parameters are provided in Table 2. Both TiN and CrN coatings have fcc crystal lattice with a strong (200) preferred orientation. Generally, by increasing the thickness of the coating, preferred orientation changes from (200) to (111); (200) texture is stabilized by a high ion to neutral ratio during vapor deposition. This orientation can improve tribological properties and wear resistance. Because the active sliding plates in a cubic crystal are (111), the change of preferential orientation to (200) improved tribological and wear properties (Paulitsch et al., 2008). Four crystalline planes were indexed on diffraction patterns including (111), (200), (220), and (311) peaks of CrN and TiN phases. Titanium atomic radius is greater than that of chromium, thus, it has a larger lattice parameter with all the diffraction peaks are located toward smaller angles as compared with chromium nitride. The lower lattice parameter of the chromium nitride coating indicates that the coating is more compact than titanium nitride. This can improve the corrosion resistance of the coating.

XRD patterns of the TiN and CrN top coatings.
Plane spacing, lattice parameter and texture coefficient of the coatings.
Parameter | TiN top coat | CrN top coat | ||||||
---|---|---|---|---|---|---|---|---|
(hkl) Miller indices | (111) | (200) | (220) | (311) | (111) | (200) | (220) | (311) |
Plane spacing (Å) | 2.46 | 2.13 | 1.52 | 1.28 | 2.37 | 2.07 | 1.45 | – |
Lattice parameter (Å) | 4.25 | 4.26 | 4.29 | 4.28 | 4.08 | 4.12 | 4.14 | – |
Texture coefficient | 1.53 | 2.17 | 0.83 | 0.35 | 0.65 | 1.76 | 0.59 | – |
SEM micrographs of the coated and uncoated specimens are shown in Figure 3. The substrate was composed of martensite and a small amount of ferrite as in Figure 3A; a typical inclusion is also observed in the micrograph, probably composed of MnS. Figure 3B and C show the surface topography of CrN and TiN top-coated specimens, respectively. There are some microparticles and pinholes on the surface of coatings that are common features of cathodic arc vapor deposition. Average surface roughness (Ra) of the coatings was estimated at 115 and 75 nm for TiN and CrN top coatings, respectively. SEM micrographs of the cross-sections of multilayer coatings are shown in Figure 3D and E. These coatings exhibit a typical compact structure without cracking and delamination that indicate, conforming a good adhesion between layers and the substrate. The total thickness of the coating was estimated at approximately 2±0.2 μm for TiN and CrN multilayers.

SEM micrographs of surface and cross-sections of PVD-coated UNS S17400 stainless steel: (A) substrate, (B, D) topography and cross-section images of CrN top coat, (C, E) topography and cross-section images of TiN top coat.
3.2 Evaluation of corrosion and corrosion-wear
Figure 4 shows the polarization curves of the substrate and multilayer coatings. As shown in Table 3, the corrosion potential of the uncoated UNS S17400 stainless steel was –456 mV. TiN coating exhibited a higher corrosion resistance with a corrosion potential of –462 mV, and CrN still had a better corrosion behavior with a corrosion potential of –485 mV. By increasing the corrosion resistance, the corrosion current density of the specimen was decreased; these indications confirm that transition metal nitrides are not affected by chemical attacks, thereby increasing the corrosion resistance (Grips et al., 2006). The passive film on the nitride coating is enhanced by the presence of nitrogen and has an effective role in increasing the corrosion resistance. Nitrogen in the film reinforced the formation of NH3 and further NH4+ compounds in addition to metal oxides on the surface, increasing the pH and decreasing the material oxidation driving force (Ningshen et al., 2007; Lavigne et al., 2011). Grabke (1996) has demonstrated that nitrogen in a negatively charged state Nδ− at the metal surface under the passive film will become rich. Some nitrogen may traverse the film as NHx and will be dissolved as NH4+. The presence of Nδ− may inhibit pit initiation by suppressing Cl−, therefore, nitride coating on the surface can increase the corrosion resistance.

Potentiodynamic polarization curves of the specimens.
Results of potentiodynamic polarization tests.
Specimen | R p (Ω·cm2) | βc (mV/dec) | βa (mV/dec) | I corr (μA/cm2) | E corr (mV vs. SCE) | Coating porosity |
---|---|---|---|---|---|---|
UNS S17400 | 1659 | 29.43 | 96.99 | 5.9 | –453 | – |
CrN top coat | 3736 | 25.74 | 57.92 | 2.2 | –484 | 0.443 |
TiN top coat | 2275 | 36.39 | 51.44 | 4.2 | –461 | 0.729 |
SEM micrographs of the specimens after corrosion test are shown in Figure 5. Figure 5A reveals bare martensitic stainless steel surface containing a 127 μm diameter cavity. Unlike the uncoated steel, the coated specimens (Figure 5B and C) showed few pores as a result of pitting corrosion. Further examination of the SEM micrographs confirmed that CrN coating exhibited excellent corrosion resistance with a relatively smooth morphology.

Corrosion morphologies of specimens: (A) UNS S17400, (B) CrN top coating, (C) TiN top coating.
The presence of inclusions in the stainless steel substrate could be the initial point of corrosion leading to large cavities due to different potentials of the passive stainless steel and the inclusion in chloride-containing solutions. When the uncoated steel is susceptible to free corrosion potential, the inclusion will be polarized and begins to dissolve. The inclusion has low electrical conductivity and caused localized dissolution at the interface of matrix-inclusion. The inclusion dissolution creates a gap between the matrix and inclusion, and thus, an increasing dissolution process (Eklund, 1974; Vuillemin et al., 2003; Nakhaie & Moayed, 2014). Baker and Castle (1993) have shown that the inclusions can either be dissolved partially or completely. When the inclusion is dissolved, the concentration of Cl− increases in the pores leading to the formation of a cavity. Conversely, in the coated specimens, the inclusions have no contact with the electrolyte and corrosion must be initiated from the nitride surface. Multilayer coating has high corrosion resistance that depends on the contact area with the electrolyte and nonhomogeneities on the surface. As shown in Table 3 and Figure 5, the specimen with the CrN top coating represented better corrosion resistance than the TiN top coat. This is because of the ability of CrN to create a passive layer on the surface (Liu et al., 2001). Moreover, TiN coatings normally have columnar structure with some defects such as pinholes and pores that affect the corrosion resistance of the coating; porosity content of TiN layer was more than that of CrN. Lower coating density, columnar structure, and higher porosity content of the TiN coating caused penetration of electrolyte into the intercolumnar paths and premature corrosion attack. On the contrary, the microstructure of CrN is dense and provides no direct diffusion path for the corrosive solution (Grips, 2006; Beibei et al., 2016).
The Nyquist and the electrochemical impedance spectroscopy plots of coated and uncoated samples are shown in Figures 6 and 7. The Nyquist diagram for coated and uncoated specimens show a semicircular graph with a large radius for multilayer coated samples. The largest semicircle is the chromium nitride coating and the smallest is the substrate. Also, the maximum capacitive resistance is related to the CrN top coating and the lowest value is for the substrate. In general, the larger diameter of the circles in the diagram represents the higher corrosion resistance of the coating. As shown in Figure 7, the highest impedance at low frequency is also related to the sample with chromium nitride top coating. This reveals that the multilayer PVD coatings improved the corrosion resistance of UNS S17400 martensitic stainless steel. An equivalent circuit model was used to describe the mechanism of electrochemical reactions as shown in Figure 8. In this model, The electrode characterization includes the determination of the polarization resistance (Rs), Ccoat, and Rcoat are ascribed the coating properties and the interaction between the solution-coating; Cdl is related to double-layer capacitance, the Rct that couples with Cdl corresponds to the charge transfer resistance due to the formation of a double layer of charge at the substrate/electrolyte interface. The values of fitted parameters of the equivalent circuit for all specimens, using the Zview software, are presented in Table 4.

Nyquist plot of coated and uncoated specimens.

Electrochemical impedance spectroscopy plot of coated and uncoated specimens.

The equivalent circuit model.
EIS parameters for multilayer coating and substrate in 3.5% NaCl solution.
Specimen | R ct (kΩ cm2) | R coat (Ω cm2) | R s (Ω cm2) | C dl (μF/cm2) | C coat (μF/cm2) |
---|---|---|---|---|---|
UNS S17400 | 1.723 | – | 13 | 0.1555 | – |
CrN top coat 3 | 3.821 | 463 | 8 | 0.1807 | 83.23 |
TiN top coat | 2.941 | 167 | 9 | 0.1145 | 86.33 |
OCP of uncoated stainless steel and coated specimens as a function of time during corrosion-wear test is shown in Figure 9. In the early stages of soaking, OCP was constant. The potential was suddenly reduced to negative values when the sliding motion was started; the reduction was due to the rupture or removal of the passive film that caused the surface became exposed to corrosive environment. An almost constant state of OCP, by establishing a balance between the recovery rate of electrochemical and mechanical damage to the passive layer, was experienced (Kok et al., 2005; Diomidis et al., 2009; Beibei, 2016). When the motion was stopped, the OCP gradually increased with time. Potential fluctuations were observed in all three specimens, uncoated stainless steel, and the coated specimens.

E corr for coated and uncoated stainless steel as a function of testing time.
Figure 10 presents SEM micrographs and energy-dispersive X-ray spectroscopy (EDS) microanalyses of the wear tracks after corrosion-wear tests in 3.5% NaCl solution. Figure 10A and C represent the delamination of layers in the track. TiN top coating was damaged more than CrN coating with the distribution of numerous cracks on the failure zone edges. Results of EDS analyses at zones A and B are seen in Figure 10B and D, respectively. Ti, Al, N, O, and small amounts of Cl elements were identified in zone A, whereas Cr, Al, N, and O elements were seen in zone B. Aluminum appeared in the EDS analysis from the intermediate layer because of low thickness of the top layer. The presence of oxygen confirmed the formation of oxide films on the surface; this oxidation caused the surface to become passive with increased corrosion resistance at the beginning of the test. As sliding continued, the oxide film broke and cracks appeared on the surface. Apparently, the CrN coating exhibited a better corrosion-wear performance than TiN top coating. The roughness and porosity of CrN top coat are less than those in TiN coating, leading to a more compact coating, thus, resisting against the formation of microcracks on the surface. It is clearly observed that microcracks on the CrN surface are less than those on the TiN surface. Such cracks act as an effective path for accelerated diffusion of NaCl solution inside the coating. Chlorine ions are considered an important factor in corrosion failure of materials (Bhandari et al., 2015; Wang et al., 2016). More cracks and porosity as well as the influence of Cl− ions greatly deteriorated the interfacial bonding strength within the coating material; as a result, the top surface of TiN coating was destroyed before CrN coating.

SEM images and EDS analyses of the wear track after corrosion-wear tests; (A, B) CrN top coating, (C, D) TiN top coating.
4 Conclusions
Corrosion and corrosion-wear tests were conducted on PVD multilayer nitride coatings deposited on UNS S17400 stainless steel in 3.5% NaCl solution. The coatings were produced using an industrial cathodic arc process and characterized in terms of structure and phase composition. Overall results show that the application of hard nitride coatings improves the resistance of this alloy against pitting corrosion and corrosion-wear in NaCl solution. Furthermore, the CrN top layer showed an increased improvement in corrosion resistance compared with the TiN coating, although the difference in corrosion-wear behavior was less significant. Failure mechanism during corrosion-wear testing of the TiN coating included extensive cracking inside the wear track followed by diffusion of NaCl solution into the holes and detachment of coating from the edges of the channel.
Acknowledgment
The authors would like to thank Najafabad Islamic Azad University and Isfahan University of Technology for provision of research facilities, and Iran National Science Foundation for financial support.
References
Abadias G. Stress and preferred orientation in nitride-based PVD coatings. Surf Coat Technol 2008; 202: 2223–2235.10.1016/j.surfcoat.2007.08.029Search in Google Scholar
Baker MA, Castle JE. The initiation of pitting corrosion at MnS inclusions. Corros Sci 1993; 34: 667–682.10.1016/0010-938X(93)90279-PSearch in Google Scholar
Beibei Z, Wang J, Zhang Y, Han G, Yan F. Comparison of tribocorrosion behavior between 304 austenitic and 410 martensitic stainless steels in artificial seawater. RSC Adv 2016; 6: 107933–107941.10.1039/C6RA18497ASearch in Google Scholar
Bhandari J, Khan F, Abbassi R, Garaniya V, Ojeda R. Modelling of pitting corrosion in marine and offshore steel structures – a technical review. J Loss Prev Process Ind 2015; 37: 39–62.10.1016/j.jlp.2015.06.008Search in Google Scholar
Cegil O, Kilinc B, Sen S, Sen U. Corrosion properties of CrAlN and TiAlN coatings deposited by thermoreactive deposition process. Acta Phys Pol A 2014; 125: 359–361.10.12693/APhysPolA.125.359Search in Google Scholar
Chen Z, Zhou G, Chen Z. Microstructure and hardness investigation of 17-4PH stainless steel by laser quenching. Mater Sci Eng A 2012; 534: 436–541.10.1016/j.msea.2011.12.004Search in Google Scholar
Chen Q, Cao Y, Xie Z, Chen T, Wan Y, Wang H, Gao X, Chen Y, Zhou Y, Guo Y. Tribocorrosion behaviors of CrN coating in 3.5 wt% NaCl solution. Thin Solid Films 2017; 622: 41–47.10.1016/j.tsf.2016.12.023Search in Google Scholar
Creus J, Mazille H, Idrissi H. Porosity evaluation of protective coatings onto steel, through electrochemical techniques. Surf Coat Technol 2000; 130: 224–232.10.1016/S0257-8972(99)00659-3Search in Google Scholar
Diomidis N, Gockan N, Ponthiaux P, Wenger F, Celis JP. Assessment of the surface state behaviour of Al71Cu10Fe9Cr10 and Al3Mg2 complex metallic alloys in sliding contacts. Intermetallics 2009; 17: 930–937.10.1016/j.intermet.2009.04.003Search in Google Scholar
Ebrahimzadeh I, Ashrafizadeh F. High temperature wear and frictional properties of duplex-treated tool steel sliding against a two phase brass. Ceram Int 2014; 40: 16429–16439.10.1016/j.ceramint.2014.07.151Search in Google Scholar
Eklund GS. Initiation of pitting at sulfide inclusions in stainless steel. Electrochem Soc 1974; 121: 467–473.10.1149/1.2401840Search in Google Scholar
Feng H, Hsu CH, Lu JK, Shy YH. Effects of PVD sputtered coatings on the corrosion resistance of AISI 304 stainless steel. Mater Sci Eng A 2003; 347: 123–129.10.1016/S0921-5093(02)00578-6Search in Google Scholar
Gilewicz A, Chmielewska P, Murzynski D, Dobruchowska E, Warcholinski B. Corrosion resistance of CrN and CrCN/CrN coatings deposited using cathodic arc evaporation in Ringer’s and Hank’s solutions. Surf Coat Technol 2016; 299: 7–14.10.1016/j.surfcoat.2016.04.069Search in Google Scholar
Grabke HJ. The role of nitrogen in the corrosion of iron and steels. ISIJ Int 1996; 36: 777–786.10.2355/isijinternational.36.777Search in Google Scholar
Grips VKW, Barshilia HC, Selvi V, Rajam KS. Electrochemical behavior of single layer CrN, TiN, TiAlN coatings and nanolayered TiAlN/CrN multilayer coatings prepared by reactive direct current magnetron sputtering. Thin Solid Films 2006; 514: 204–211.10.1016/j.tsf.2006.03.008Search in Google Scholar
Hu Z, Zhu H, Zhang H, Zeng X. Experimental investigation on selective laser melting of 17-4PH stainless steel. Opt Laser Technol 2017; 87: 17–25.10.1016/j.optlastec.2016.07.012Search in Google Scholar
Kok YN, Akid R, Hovsepian PE. Tribocorrosion testing of stainless steel (SS) and PVD coated SS using a modified scanning reference electrode technique. Wear 2005; 259: 1472–1481.10.1016/j.wear.2005.02.049Search in Google Scholar
Lavigne O, Dumont CA, Normand B, Berger MH, Duhamel C, Delichere P. The effect of nitrogen on the passivation mechanisms and electronic properties of chromium oxide layers. Corros Sci 2011; 53: 2087–2096.10.1016/j.corsci.2011.02.026Search in Google Scholar
Lin X, Cao Y, Wu X, Yang H, Chen J, Huang W. Microstructure and mechanical properties of laser forming repaired 17-4PH stainless steel. Mater Sci Eng A 2012; 553: 80–88.10.1016/j.msea.2012.05.095Search in Google Scholar
Liu C, Bi Q, Matthews A. EIS comparison on corrosion performance of PVD TiN and CrN coated mild steel in 0.5 N NaCl aqueous solution. Corros Sci 2001; 43: 1953–1961.10.1016/S0010-938X(00)00188-8Search in Google Scholar
Liu RL, Qiao YJ, Yan MF, Fu YD. Layer growth kinetics and wear resistance of martensitic precipitation hardening stainless steel plasma nitrocarburized at 460°C with rare earth addition. Met Mater Int 2013; 19: 1151–1157.10.1007/s12540-013-5038-7Search in Google Scholar
Matei AA, Penceab I, Branzei M, Tranc DE, Tepes G, Sfat CE, Ciovica E, Gherghilescu AI, Stanciu GA. Corrosion resistance appraisal of TiN, TiCN and TiAlN coatings deposited by CAE-PVD method on WC-Co cutting tools exposed to artificial sea water. Appl Surf Sci 2015; 358: 572–578.10.1016/j.apsusc.2015.08.041Search in Google Scholar
Mazur Z. Steam turbine blade failure analysis. Eng Fail Anal 2008; 15: 129–141.10.1016/j.engfailanal.2006.11.018Search in Google Scholar
Nakhaie D, Moayed MH. Pitting corrosion of cold rolled solution treated 17-4 PH stainless steel. Corros Sci 2014; 80: 290–298.10.1016/j.corsci.2013.11.039Search in Google Scholar
Ningshen S, Mudali UK, Mittal VK, Khatak HS. Semiconducting and passive film properties of nitrogen-containing type 316LN stainless steels. Corros Sci 2007; 49: 481–496.10.1016/j.corsci.2006.05.041Search in Google Scholar
Park G, Bae G, Moon K, Lee C. Effect of plasma nitriding and nitrocarburizing on HVOF-sprayed stainless steel coatings. J Therm Spray Technol 2013; 22: 1366–1373.10.1007/s11666-013-0035-4Search in Google Scholar
Paulitsch J, Mayrhoferb PH, Munz WD, Schenkel M. Structure and mechanical properties of CrN/TiN multilayer coatings prepared by a combined HIPIMS/UBMS deposition technique. Thin Solid Films 2008; 517: 1239–1244.10.1016/j.tsf.2008.06.080Search in Google Scholar
Poursaee A. Potentiostatic transient technique, a simple approach to estimate the corrosion current density and Stern–Geary constant of reinforcing steel in concrete. Cem Concr Res 2010; 40: 1451–1458.10.1016/j.cemconres.2010.04.006Search in Google Scholar
Rossia S, Fedrizzia L, Leoni M, Scardia P, Massianib Y. (Ti,Cr)N and Ti/TiN PVD coatings on 304 stainless steel substrates: wear-corrosion behavior. Thin Solid Films 1999; 350: 161–167.10.1016/S0040-6090(99)00235-7Search in Google Scholar
Shan L, Wang Y, Li J, Li H, Wu X, Chen J. Tribological behaviours of PVD TiN and TiCN coatings in artificial seawater. Surf Coat Technol 2013; 227: 40–50.10.1016/j.surfcoat.2013.03.034Search in Google Scholar
Song GH, Yang XP, Xiong GL, Lou Z, Chen LJ. The corrosive behavior of Cr/CrN multilayer coatings with different modulation periods. Vacuum 2013; 89: 136–141.10.1016/j.vacuum.2012.02.046Search in Google Scholar
Stern M. A method for determining corrosion rates from linear polarization data. Corrosion 1958; 14: 440–444.10.5006/0010-9312-14.9.60Search in Google Scholar
Stueber M, Holleck H, Leiste H, Seemann K, Ulrich S, Ziebert C. Concepts for the design of advanced nanoscale PVD multilayer protective thin films. J Alloy Compd 2009; 483: 321–333.10.1016/j.jallcom.2008.08.133Search in Google Scholar
Surviliene S, Bellozor S, Kurtinaitiene M, Safonov VA. Protective properties of the chromium–titanium carbonitride composite coatings. Surf Coat Technol 2004; 176: 193–201.10.1016/S0257-8972(03)00716-3Search in Google Scholar
Vuillemin B, Philippe X, Oltra R, Vignal V, Coudreuse L, Dufour LC, Finot E. AFM and AES study of pitting corrosion initiated on MnS inclusions by microinjection. Corros Sci 2003; 45: 1143–1159.10.1016/S0010-938X(02)00222-6Search in Google Scholar
Wang Y, Cheng G, Wu W, Qiao Q, Li Y, Li X. Effect of pH and chloride on the micro-mechanism of pitting corrosion for high strength pipeline steel in aerated NaCl solution. Appl Surf Sci 2016; 349: 746–756.10.1016/j.apsusc.2015.05.053Search in Google Scholar
Yang M, Guo Z, Xiong J, Liu F, Qi K. Microstructural changes of (Ti,W)C solid solution induced by ball milling. Int J Refract Metals Hard Mater 2017; 66: 83–87.10.1016/j.ijrmhm.2017.03.008Search in Google Scholar
Yashar PC, Sproul WD. Nanometer scales multilayered hard coatings. Vacuum 1999; 55: 179–190.10.1016/S0042-207X(99)00148-7Search in Google Scholar
©2018 Walter de Gruyter GmbH, Berlin/Boston
Articles in the same Issue
- Frontmatter
- In this issue
- Reviews
- Review of hydrogen-assisted cracking models for application to service lifetime prediction and challenges in the oil and gas industry
- Electrochemical and surface analytical techniques applied to microbiologically influenced corrosion investigation
- Mini review
- Recent reviews on quinoline derivatives as corrosion inhibitors
- Original articles
- Inhibitory effect of Atlas cedar essential oil on the corrosion of steel in 1 m HCl
- Effects of oxygen concentration on the passivation of Si-containing steel during high-temperature oxidation
- Effect of soft cations on carbon steel corrosion in chloride media
- Comparative study of corrosion and corrosion-wear behavior of TiN and CrN coatings on UNS S17400 stainless steel
Articles in the same Issue
- Frontmatter
- In this issue
- Reviews
- Review of hydrogen-assisted cracking models for application to service lifetime prediction and challenges in the oil and gas industry
- Electrochemical and surface analytical techniques applied to microbiologically influenced corrosion investigation
- Mini review
- Recent reviews on quinoline derivatives as corrosion inhibitors
- Original articles
- Inhibitory effect of Atlas cedar essential oil on the corrosion of steel in 1 m HCl
- Effects of oxygen concentration on the passivation of Si-containing steel during high-temperature oxidation
- Effect of soft cations on carbon steel corrosion in chloride media
- Comparative study of corrosion and corrosion-wear behavior of TiN and CrN coatings on UNS S17400 stainless steel