Next Article in Journal
Study on the Tribological Properties of DIN 16MnCr5 Steel after Duplex Gas-Nitriding and Pack Boriding
Previous Article in Journal
Tribological Properties of Composites Based on Single-Component Powdered Epoxy Matrix Filled with Graphite
Previous Article in Special Issue
Laser Shock Peening: Fundamentals and Mechanisms of Metallic Material Wear Resistance Improvement
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Effect of Nitriding Temperature of AISI 316L Steel on Sub-Zero Corrosion Resistance in C2H5OH

by
Beata Kucharska
*,
Janusz Kamiński
,
Krzysztof Kulikowski
,
Tomasz Borowski
,
Jerzy Robert Sobiecki
and
Tadeusz Wierzchoń
Faculty of Materials Science and Engineering, Warsaw University of Technology, ul. Wołoska 141, 02-507 Warsaw, Poland
*
Author to whom correspondence should be addressed.
Submission received: 28 April 2024 / Revised: 11 June 2024 / Accepted: 18 June 2024 / Published: 21 June 2024

Abstract

:
In this paper, glow nitriding processes at cathode potential are used at various temperatures to investigate how they affect the corrosion resistance of 316L steel in ethanol at temperatures of 22 °C and −30 °C. Lowering the test temperature reduces the corrosion rate of the nitrided layers. Conversely, glow nitriding at 450 °C improves the corrosion resistance of the tested steel. Increasing the nitriding temperature to 520 °C increases the corrosion rate. It should be noted that the ethyl alcohol solution, due to the lack of aggressive ions, does not cause significant changes in the corrosion rate of the steel. The value of the corrosion current varies in the range of 10−2–10−3 µA/cm2. Nitrided layers increase the contact angle measured for water and are entirely wettable for ethanol. The objective of this study is to evaluate the effect of the nitriding temperature of AISI 316L steel on its corrosion resistance in an ethanol solution at room temperature and at −30 °C.

1. Introduction

Austenitic stainless steels (ASSs) are of great interest due to their excellent formability, corrosion resistance, and biocompatibility [1,2]. ASSs are widely used in the chemical, oil and gas, and orthopedic industries due to their high resistance to corrosion processes, which is mainly due to their high chromium content, sometimes with the addition of molybdenum in solid solution, which is responsible for the formation of a natural self-healing passive layer [3]. However, their low hardness and low resistance to frictional wear severely limit their industrial application. An increasing number of thermochemical methods have been used to improve the surface of ASSs, including nitriding [4,5,6,7,8,9,10], carburizing [9,11,12,13,14], nitrocarburizing [13,15,16], oxynitriding [17], and boriding [18]. Of the above technologies, the most widely used and researched is glow discharge nitriding, which, at a low nitriding temperature (below 450 °C) and a short process time (less than 6 h), produces a precipitation-free nitrogen austenite (γN) on the surface, the so-called S phase (expanded austenite), with high corrosion resistance and good wear resistance [15,16,19]. Exceeding this nitriding temperature can lead to the formation of CrN precipitates, resulting in chromium-depleted zones [6,20]. Higher process temperatures, due to the additional precipitation of cubic (γ′-Fe4N) and hexagonal (ε-Fe2–3N) nitrides, improve wear resistance but significantly reduce the corrosion resistance of ASSs in the presence of chlorides [21,22]. Therefore, the nitriding temperature is a critical factor affecting the corrosion resistance of the nitrided layer. The centrifugal medical cannabis extractor developed in this project, which is mainly made of ASSs, operates at −30 °C in an ethanol solution. There is a large body of work investigating the corrosion resistance of nitrided ASSs, particularly in chloride-containing environments [23,24], which have the greatest influence on the pitting of the ASS surface. Tests are mostly conducted at room temperature, but also at 37 °C, a temperature comparable to that of the human body, when testing in solutions simulating human body fluids [25]. Fewer studies are conducted in oxidizing solutions, such as acids [26,27] and alkalis [28]. However, it has been noted that there is an absolute lack of literature data on the corrosion resistance of ASSs in alcohols, and in particular in C2H5OH ethanol.
The aim of this study was to evaluate the effect of the nitriding temperature of AISI 316L steel on its corrosion resistance, measured by potentiodynamic method and impedance spectroscopy in an ethanol solution at room temperature and at −30 °C, as well as on the contact angle of the tested surfaces, also measured using ethanol.

2. Materials and Methods

2.1. Substrate and Layer Deposition

AISI 316L type 17/10/2 alloy steel (Cr/Ni/Mo) was used as the substrate. The specimens were cut from bars with a diameter of Φ 14 and then ground on sandpaper up to 1200′. The glow nitriding processes were carried out at the cathode potential (the specimen was negatively polarized) for 6 h under the 2 mbar pressure in N2/H2 atmosphere (1/1 ratio). The glow nitriding processes were carried out at variable temperatures (450 °C, 480 °C, and 520 °C) with the other technological parameters remaining constant.
In the case of each process, cathodic sputtering was carried out at a temperature of 430 °C in an Ar/H2/N2 = 2/2/1 atmosphere for 30 min while the processed specimens were being heated, in order to clean and activate the specimen surfaces and increase the number of structural defects in the near-surface zone, leading to the formation of a nitrided layer of greater thickness as a result of accelerated diffusion processes.

2.2. Microstructure, Morphology, and Chemical Composition Studies

The specimens were cut on a precision saw and then hot-incorporated in conductive phenolic resin with graphite filler. Their cross-sections were then ground with 240–1200 grit sandpaper and polished with an Al2O3 suspension. The layer structure was revealed by etching the specimens with a reagent consisting of 50% HCl + 25% HNO3 + 25% H2O. The microstructure, surface morphology, and chemical composition of the layers were examined using a scanning electron microscope (S3500N, HITACHI, Tokyo, Japan) equipped with the energy-dispersive spectroscopy (EDS) microanalysis system.

2.3. XRD Studies

The XRD studies of the nitrided layers were performed on an X-ray diffractometer (D8 ADVANCE, Bruker, Billerica, MA, USA) using filtered Cu Kα radiation (λ = 0.154056 nm) at room temperature. The conditions were as follows: voltage—40 kV, current—40 mA, angular range 2θ—from 15° to 120°, step Δ2Θ—0.05°, and counting time—3 s.

2.4. Microhardness Tests

Microhardness tests were carried out on a hardness tester (HMV-G, Shimadzu, Kyoto, Japan) using the Vickers method using loads of 0.2 and 0.5 kg.

2.5. Corrosion Tests

Corrosion resistance tests were carried out using the electrochemical impedance spectroscopy (EIS) and potentiodynamic methods in 99% C2H5OH water (1%) solution using a potentiostat (AutoLab PGSTAT100, Atlas-Sollich, Kartuzy County, Poland) at −30 °C and 22 °C. The choice of a low test temperature and ethanol for a corrosive environment was due to the potential use of this material to construct an extractor using ethanol operating at low temperatures. The tests were carried out in a three-electrode system: test electrode, reference electrode (the silver/silver chloride SSC electrode), and auxiliary electrode (platinum mesh: total height 150 mm, cylinder diameter 35 mm) (Figure 1a). The application of silver electrodes (118 mV vs. RHEK—Reversible Hydrogen Electrode kit (25 °C)) made it possible to conduct electrochemical tests both in anhydrous environments and at reduced temperatures.

2.5.1. Electrochemical Impedance Spectroscopy Method

Impedance tests were carried out in the frequency range 105–10−3 Hz, with a sinusoidal signal amplitude of 20 mV, at an open-circuit potential (OCP). The impedance spectra were analyzed using Baukamp’s EQUIVCRT program (version 4.9.007). Prior to the tests, the specimens were stabilized in a currentless condition to determine the open system potential (OCP) for a period of 10,000 s. An electrical equivalent system (Figure 1b) with two time constants was used to analyze the obtained spectra, allowing the characterization of both the high resistance of the electrolyte (99% C2H5OH) and the properties of the double layer formed at the metal–electrolyte interface. Rs denotes the electrolyte resistance, Y0-CPE is the constant phase element (CPE) at the interface between the electrolyte and specimen, and Rct is the charge transfer resistance of the specimens. The selection of the electrical equivalent system for the obtained impedance spectra was conditional on both the image of the corrosion damage and on the smallest matching errors of the system elements (determined using the least-square method). The obtained spectra were presented in the form of Bode and Nyquist plots.

2.5.2. Polarization Method

Polarization resistance was tested using the Stern method by polarizing the test material from a potential 10 mV below to 10 mV above the determined open-circuit potential at a sweep rate of 0.2 mV/s. The polarization resistance (Rpol) was determined from the E = f(i) dependence. Potentiodynamic tests were performed at a potential of 3000 mV. The materials were polarized at a potential change rate of 0.2 mV/s. The corrosion current density (Icorr) and corrosion potential (Ecorr) were determined using the Tafel extrapolation method.

2.6. Contact Angle Measurements

Contact angle tests were carried out using a goniometer (90-U3-PRO, RAME-HART, Ledgewood, NJ, USA) equipped with a CCD (charge-coupled device) camera, an adjustable stage, a syringe for the manual dosing of drops, and DROPImage software (ramé-hart DROPimage Pro). The liquid used in the study was distilled water and a 99% C2H5OH solution. A total of five measurements were performed on each specimen.

3. Results and Discussion

3.1. Effect of Nitriding Temperature on the Surface Morphology, Microstructure, and Chemical Composition

Small deposits on the surface of the layers (Figure 2) were formed as a result of the cathodic sputtering of the steel table on which the specimens were placed (Table 1). No other elements indicating contamination were observed. Deformations were identified near the grain boundaries, which were caused by significant stresses generated during the diffusion supersaturation of grains with different crystallographic orientations. The nitriding process led to a varying degree of expansion of adjacent grains and created a relief visible on the steel surface. The higher nitriding temperature resulted in the greatest saturation of the layer and relief effect. The surface of the layer produced in these conditions was the most cracked (Figure 2) [29,30].
The layers were characterized by different thicknesses depending on the process temperature (Figure 3). The nitrided layer produced at 450 °C was about 8 µm thick, while at 480 °C, it was about 28 µm thick, and at 520 °C, it was about 60 µm thick. The layers nitrided at 480 and 520 °C show a double structure with a thinner sublayer closer to the core. This is a layer of carbon austenite formed as a result of the movement of carbon contained in the steel into deeper zones of the layer through the action of nitrogen (Figure 3).

3.2. XRD Studies

Figure 4 shows the phase diffractograms of the tested specimens with nitrided layers. The 316L steel substrate was composed of the austenitic phase, and no other phases were detected in the substrate (Figure 4). In the case of the nitrided layer at a temperature of 450 °C, the S phase (nitrogen austenite), iron nitride Fe4N, and trace amounts of chromium nitride CrN were identified. The S phase and the Fe4N nitride showed the highest peak intensity. In turn, in the layer formed at a temperature of 480 °C, nitrogen austenite and iron nitrides Fe4N and FeN were also observed. Signals originating from CrN were characterized by the lowest intensity. In the case of the layer produced at 520 °C, the peaks from the S phase and the Fe4N nitride showed the highest intensity. The peaks from chromium nitride CrN were also the most intense compared to the layers produced at a lower temperature.

3.3. Microhardness

The microhardness of 316L steel in its initial state was approximately 336 HV0.2 (Table 2). As a result of the glow nitriding process, the surface microhardness increased to approximately 796 HV0.2 and 714 HV0.2 for the 450 °C and 480 °C processes, and to approximately 1412 HV0.2 for the 520 °C process. The increase in hardness was the result of the formation of an “s” zone on the surface, approximately 4 μm thick, and an outer zone of FeN and Fe4N iron nitrides (Figure 3), the thickness of which increased with the process temperature. In the case of the layer produced at a temperature of 520 °C, the surface hardness obtained during measurements under a load of 0.2 kg is related to its large thickness (approximately 60 μm), which allows the unhardened substrate to participate to a small extent in the plastic deformation process. The layers produced at temperatures of 450 °C and 480 °C are characterized by a much smaller thickness of the surface nitride zone (Figure 3), and when tested under high loads (0.2 kg and 0.5 kg), the influence of the plastic deformation of the substrate was clearly visible. A significant influence of the load size on the microhardness value was also observed. This is due to the fact that in the case of lower loads, the substrate had no influence on the microhardness values. In the case of increasing the load by 2.5 times, a clear influence of the substrate on the hardness was observed, which is related to the depth of penetration of the indenter during the test.

3.4. Corrosion Studies

Figure 5 shows the polarization curves of AISI 316L steel in its initial state, exposed to temperatures of −30 °C and 22 °C. The analysis of the curves shows an increase in the corrosion resistance of the steel as the exposure temperature decreases, as well as a slight shift in the corrosion potentials toward the cathodic side as the temperature decreases. The reduction in the electrochemical activity of electrode processes is determined using the Arrhenius equation ( k = A e E a R T , k—rate constant, A—pre-expotential factor, Ea—activation energy, R—universal gas constant, and T—absolute temperature), which expresses the dependence of the reaction rate constant on temperature.
Table 3 summarizes the characteristic electrochemical values obtained from potentiodynamic tests.
The applied glow nitriding processes at the cathode potential, due to the high electrochemical resistance (~1 M Ωcm2) and the low aggressiveness of the corrosive environment (lack of chloride ions), did not significantly affect the changes in the corrosion resistance and the durability of the nitrided layers produced (Figure 6). However, the influence of increasing the nitriding temperature on the corrosion current density was observed (correlation coefficient = 0.92). No distortions were observed during potentiodynamic studies despite using a potential scan rate of 0.2 mV/s [31].
It can be noticed that the produced layers are anodic in relation to the substrate (Figure 6 and Figure 7; Table 4), which could favor the etching of the layers as a result of long-term exposure. However, due to the high solubility of oxygen in ethyl alcohol ( C O 2   7.5–11.6 mM (25 °C)) [32], any defects in the layer (as well as nondefected areas) have the potential to auto-oxidize and form oxy-nitrides, which, in the absence of chlorides in the solution, would have a positive effect on both the corrosion resistance and the durability of the surface layers.
The shape of the potentiodynamic curves indicates the lack of pitting corrosion in the entire range of the tested potentials. The observed constant increase in the current density value with increasing polarization intensity indicates the low-intensity uniform corrosion of the nitrided layer; current densities in the surface layer (at 2000 mV) were approximately 1 µA/cm2 in the case of low-temperature glow nitriding and approximately 5 µA/cm2 in the case of the layer produced at 520 °C. This result indicates a slightly increased intensity of electrode processes taking place on the tested surface, probably due to the increased roughness of the substrate.
A similar tendency was also observed in the case of potentiodynamic tests at a lower (−30 °C) temperature (correlation coefficient = 0.93). However, due to the reduction in the intensity of electrochemical processes (according to the Arrhenius equation), the corrosion current densities were an order of magnitude lower than those of the tests at room temperature.
Additionally, an additional electrochemical process was observed (at a potential of approx. 400 mV (Figure 8)), increasing the intensity of anodic processes. It can be assumed that the additional electrochemical process is related to the oxidation of the nitrided layer with the simultaneous formation of oxygen and nitrides on the surface. However, due to the lack of chloride ions in the solution, pitting corrosion initiation was not observed in these areas. The best corrosion resistance occurred in the case of layer nitride at 450 °C (Figure 9, Table 5).
Figure 10 shows the surfaces of layers nitrided at temperatures of 480 °C and 520 °C after corrosion tests in an ethyl alcohol (22 °C). In both cases, the local etching of the nitrided layer was observed, producing dielectric oxides.
The impedance test results (Figure 11 and Figure 12) confirmed the positive effect of low-temperature (450–480 °C) glow nitriding on the corrosion resistance of 316L steel. Increasing the glow treatment temperature (520 °C) resulted in a significant reduction in corrosion resistance (Table 6) due to both a significant increase in surface development and the possibility of chromium nitride (CrN) formation at the grain boundaries, which on the one hand increases hardness but may also initiate intergranular corrosion. This is confirmed by the work of Skolek et al., which shows the possibility of chromium nitride formation at temperatures above 440 °C [33]. The observed increase in layer hardness also led to stress corrosion cracking in the deflection zone of the specimen as a result of external stresses and long-term exposure at −30 °C (Figure 13).
Figure 14 and Figure 15 show the impedance spectra (Bode and Nyquist, respectively) of the tested layers compared to 316L exposed at lower (−30 °C) temperature, exposed in 99% C2H5OH, while Table 7 lists the characteristic electrochemical values of the double layer present at the nitrided layer–electrolyte phase boundary.
The observed high-frequency peaks in the impedance spectra characterize the resistivity of the environment. In the case of the 316L steel, the resistance of the environment at room temperature was significant and amounted to approximately 0.79 × 106 Ωcm2. When the temperature decreased to around −30 °C, the value further increased to 2.01 × 106 Ωcm2. The increase in solution resistance as a result of lowering the temperature is one of the reasons for the observed decrease in corrosion current density in the case of potentiodynamic tests (Table 5). In the case of impedance tests, where it is possible to eliminate the influence of the resistance of the environment on the remaining electrochemical results (i.e., the resistance of charge through the double layer (resistance Rct) and the capacity of the double layer or the value of the ‘n’ parameter), we can precisely determine the influence of the glow treatment temperature on the corrosion resistance of nitrided layers exposed in ethyl alcohol. The analysis of the change in the resistance value Rct indicates a decrease in its value in proportion to the increase in the glow treatment temperature. This trend was observed regardless of the exposure temperature (i.e., 22 °C and −30 °C). Additionally, in the impedance spectra of nitrided layers produced at elevated temperature (520 °C), a slightly marked additional capacitive peak was observed (in the frequency range 10−3–105 Hz), which was caused by an increase in the roughness of the surface layer.
The constant ‘n’ parameter of the tested layers, amounting to approximately 0.85, indicates the passive–active nature of the surface layers exposed at a temperature of 22 °C. Slight differences in the values of the ‘n’ parameter were observed in the case of tests at a reduced temperature. The obtained variable values did not significantly change the nature of the layer; only in the case of the layer produced at 520 °C does the reduced value of the ‘n’ parameter (n = 0.67) indicate a significant influence of diffusion factors determining the electrochemical resistance of the layer. It can be assumed that the obtained result is caused by the significant development of the surface.
Figure 12 and Figure 15 show the Nyquist plots of the 316L substrate and the nitrided layers. As can be seen in the figures, in the higher temperature, there are two semicircles for each test specimen. The first semicircle is connected with electrolyte resistance. In the case of a lower temperature, only the first semicircle is clearly visible, which indicates a different corrosion mechanism. The electrolyte resistance is about twice that of a higher test temperature (Table 6 and Table 7). The higher slopes of the second semicircle at −30 degrees indicate significantly higher charge transfer resistance between the electrolyte and the tested materials under these test conditions.

3.5. Contact Angle

The nitrided layers were characterized by a higher water contact angle than the initial state (316L) (Table 8). The highest value of the contact angle in water was obtained for the specimen nitrided at 520 °C. Nitrided specimens were characterized by complete wettability by ethanol.

4. Conclusions

  • Notably, the 99% ethyl alcohol solution was characterized by low aggressiveness due to the lack of chloride ions in the solution and low conductivity, which was further diminished by the reduced temperature. Nyquist spectra suggest different corrosion mechanisms of tested corrosion temperatures.
  • Low-temperature glow nitriding of 316L steel slightly increased the material corrosion resistance.
  • Glow nitriding at 520 °C significantly reduced the corrosion resistance of 316L steel. Additionally, due to the significant thickness and hardness of the surface layer, the material was sensitive to stress corrosion cracking.
  • The high solubility of oxygen in ethyl alcohol facilitates the oxidation of the nitrided layer.
  • Nitrided layers were characterized by a higher water contact angle than the initial state. The highest value of the contact angle in water was obtained for the specimen nitrided at 520 °C.
  • The nitrided specimens were completely wettable using ethanol.

Author Contributions

Conceptualization, J.R.S.; methodology, B.K., J.K., K.K. and T.B.; validation, B.K. and J.R.S.; formal analysis, T.W.; investigation, B.K., J.K., K.K. and T.B.; writing—original draft preparation, B.K. and T.B.; writing—review and editing, B.K.; supervision, J.R.S. and T.W.; funding acquisition, J.R.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Center for Research and Development, as part of the “Path for Mazovia” competition, grant number MAZOWSZE/0070/19-00.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Abedi, H.R.; Salehi, M. Effect of post-oxidizing temperature on tribological and corrosion behavior of plasma nitrided austenitic stainless steel. Mater. Des. 2011, 32, 2100–2106. [Google Scholar] [CrossRef]
  2. Jinlong, L.; Hongyun, L. Effects of strain and strain-induced α′-martensite on passive films in AISI 304 austenitic stainless steel. Mater. Sci. Eng. C 2014, 34, 484–490. [Google Scholar]
  3. Khatak, H.S.; Raj, B. Woodhead Publishing Series in Metals and Surface Engineering, Corrosion of Austenitic Stainless Steels; Woodhead Publishing: Cambridge, UK, 2002. [Google Scholar]
  4. Manova, D.; Mändl, S. Initial phase formation during nitriding of austenitic stainless steel. Surf. Coat. Technol. 2023, 456, 129258. [Google Scholar] [CrossRef]
  5. Lu, Y.; Li, D.; Ma, H.; Liu, X.; Wu, M.; Hu, J. Enhanced plasma nitriding efficiency and properties by severe plastic deformation pretreatment for 316L austenitic stainless steel. J. Mater. Res. Technol. 2021, 15, 1742–1746. [Google Scholar] [CrossRef]
  6. Huang, Z.; Guo, Z.; Liu, L.; Guo, Y.; Chen, Y.; Zhang, Z.; Li, Y.; Li, Y.; Zhou, Y.; Liang, Y. Structure and corrosion behavior of ultra-thick nitrided layer produced by plasma nitriding of austenitic stainless steel. Surf. Coatings Technol. 2021, 405, 126689. [Google Scholar] [CrossRef]
  7. Lin, K.; Li, X.; Dong, H.; Guo, P.; Gu, D. Nitrogen mass transfer and surface layer formation during the active screen plasma nitriding of austenitic stainless steels. Vacuum 2018, 148, 224–229. [Google Scholar] [CrossRef]
  8. Mali, A.S.; Vagge, S.T.; Rathod, M.J. Mapping the Accouterment Effects of Plasma Nitriding on AISI 316L in Biomedical Applications. Coatings 2023, 13, 839. [Google Scholar] [CrossRef]
  9. Borgioli, F. The Corrosion Behavior in Different Environments of Austenitic Stainless Steels Subjected to Thermochemical Surface Treatments at Low Temperatures: An Overview. Metals 2023, 13, 776. [Google Scholar] [CrossRef]
  10. Moskaliovienė, T.; Andriūnas, P.; Galdikas, A. Coupled Modeling of Anisotropic Stress-Induced Diffusion and Trapping of Nitrogen in Austenitic Stainless Steel during Nitriding and Thermal Annealing. Coatings 2023, 13, 415. [Google Scholar] [CrossRef]
  11. Maniscalco, S.; Caligari Conti, M.; Cassar, J.; Grima, C.; Karl, A.; Schembri Wismayer, P.; Mallia, B.; Buhagiar, J. Low temperature carburised austenitic stainless steel for metal-on-metal tribological contact. Thin Solid Film. 2016, 620, 103–113. [Google Scholar] [CrossRef]
  12. Chiavari, C.; Bernardi, E.; Balijepalli, S.K.; Kaciulis, S.; Ceschini, L.; Martini, C. Influence of low-temperature carburising on metal release from AISI316L austenitic stainless steel in acetic acid. J. Food Eng. 2014, 137, 7–15. [Google Scholar] [CrossRef]
  13. Spies, H.-J. 6–Corrosion behaviour of nitrided, nitrocarburised and carburised steels. In Thermochemical Surface Engineering of Steels; Mittemeijer, E.J., Somers, M.A.J., Eds.; Woodhead Publishing: Cambridge, UK, 2015; pp. 267–309. [Google Scholar]
  14. Illing, C.; Ren, Z.; Ernst, F. Low-Temperature Carburization: Ex Situ Activation of Austenitic Stainless Steel. Metals 2023, 13, 335. [Google Scholar] [CrossRef]
  15. Borowski, T.; Kulikowski, K.; Adamczyk-Cieślak, B.; Rożniatowski, K.; Spychalski, M.; Tarnowski, M. Influence of nitrided and nitrocarburised layers on the functional properties of nitrogen-doped soft carbon-based coatings deposited on 316L steel under DC glow-discharge conditions. Surf. Coat. Technol. 2020, 392, 125705. [Google Scholar] [CrossRef]
  16. Kajzer, A.; Ceglarska, M.; Sura, N.; Kajzer, W.; Borowski, T.; Tarnowski, M.; Pilecki, Z. Effect of Nitrided and Nitrocarburised Austenite on Pitting and Crevice Corrosion Resistance of 316 LVM Steel Implants. Materials 2020, 13, 5484. [Google Scholar] [CrossRef] [PubMed]
  17. Chen, G.; Wang, J.; Zhang, H.; Li, L.; Fan, H. Low-temperature oxy-nitriding of 316 L austenitic stainless steel for improved corrosion resistance in liquid lead-bismuth eutectic. Scr. Mater. 2021, 202, 114014. [Google Scholar] [CrossRef]
  18. Akkurt, I.; Calik, A.; Akyıldırım, H. The boronizing effect on the radiation shielding and magnetization properties of AISI 316L austenitic stainless steel. Nucl. Eng. Des. 2011, 241, 55–58. [Google Scholar] [CrossRef]
  19. Czerwiec, T.; Renevier, N.; Michel, H. Low-temperature plasma-assisted nitriding. Surf. Coat. Technol. 2000, 131, 267–277. [Google Scholar] [CrossRef]
  20. Murali, A.P.; Alphonse, M.; Ganesan, D.; Salunkhe, S.; Hussein, H.M.A.M. Sliding wear behaviour of salt bath nitrided 316LN austenitic stainless steel. Appl. Surf. Sci. Adv. 2023, 15, 100401. [Google Scholar] [CrossRef]
  21. Fossati, A.; Borgioli, F.; Galvanetto, E.; Bacci, T. Corrosion resistance properties of glow-discharge nitrided AISI 316L austenitic stainless steel in NaCl solutions. Corros. Sci. 2006, 48, 1513–1527. [Google Scholar] [CrossRef]
  22. Baranowska, J.; Arnold, B. Corrosion resistance of nitrided layers on austenitic steel. Surf. Coat. Technol. 2006, 200, 6623–6628. [Google Scholar] [CrossRef]
  23. Cheng, H.; Luo, H.; Wang, X.; Pan, Z.; Jiang, Y.; Li, X. Electrochemical corrosion and passive behavior of a new high-nitrogen austenitic stainless steel in chloride environment. Mater. Chem. Phys. 2022, 292, 126837. [Google Scholar] [CrossRef]
  24. Yang, W.J.; Zhang, M.; Zhao, Y.H.; Shen, M.L.; Lei, H.; Xu, L.; Xiao, J.Q.; Gong, J.; Yu, B.H.; Sun, C. Enhancement of mechanical property and corrosion resistance of 316L stainless steels by low temperature arc plasma nitriding. Surf. Coat. Technol. 2016, 298, 64–72. [Google Scholar] [CrossRef]
  25. De Las Heras, E.; Ybarra, G.; Lamas, D.; Cabo, A.; Dalibon, E.L.; Brühl, S.P. Plasma nitriding of 316L stainless steel in two different N2-H2 atmospheres—Influence on microstructure and corrosion resistance. Surf. Coat. Technol. 2017, 313, 47–54. [Google Scholar] [CrossRef]
  26. Köster, K.; Kaestner, P.; Bräuer, G.; Hoche, H.; Troßmann, T.; Oechsner, M. Material condition tailored to plasma nitriding process for ensuring corrosion and wear resistance of austenitic stainless steel. Surf. Coat. Technol. 2013, 228 (Suppl. S1), S615–S618. [Google Scholar] [CrossRef]
  27. Gnanarathinam, A.; Palanisamy, D.; Manikandan, N.; Devaraju, A.; Arulkirubakaran, D. Comparison of corrosion behavior on laser welded austenitic stainless steel. Mater. Today Proc. 2021, 39 Pt 1, 649–653. [Google Scholar] [CrossRef]
  28. Ramdane, H.; Benaioun, N.; Guezzoul, M.; Moulayat, N.; Craciun, V. Investigation of the corrosion-preventing properties of nickel-coated austenitic stainless steel (AISI 304) in NaOH and NaOH + 10% NaCl solutions. Inorg. Chem. Commun. 2023, 152, 110671. [Google Scholar] [CrossRef]
  29. Borowski, T. Enhancing the Corrosion Resistance of Austenitic Steel Using Active Screen Plasma Nitriding and Nitrocarburising. Materials 2021, 14, 3320. [Google Scholar] [CrossRef] [PubMed]
  30. Corujeira Gallo, S.; Dong, H. New insights into the mechanism of low-temperature active-screen plasma nitriding of austenitic stainless steel. Scr. Mater. 2012, 67, 89–91. [Google Scholar] [CrossRef]
  31. Duarte, T.; Meyer, Y.A.; Osório, W.R. The Holes of Zn Phosphate and Hot Dip Galvanizing on Electrochemical Behaviors of Multicoatings on Steel Substrates. Metals 2022, 12, 863. [Google Scholar] [CrossRef]
  32. Dvoranová, D.; Barbieriková, Z.; Brezová, V. Radical Intermediates in Photoinduced Reactions on TiO2. Molecules 2014, 19, 17279–17304. [Google Scholar] [CrossRef] [PubMed]
  33. Skolek-Stefaniszyn, E.; Kaminski, J.; Sobczak, J.; Wierzchon, T. Modifying the properties of AISI 316L steel by glow discharge assisted low-temperature nitriding and oxynitriding. Vacuum 2010, 85, 164–169. [Google Scholar] [CrossRef]
Figure 1. (a) Scheme of an electrochemical cell for corrosion resistance studies: 1—reference electrode, 2—auxiliary electrode (platinum electrode), 3—test electrode (specimen), 4—99% ethanol environment, P—potentiostat, 5—insulator with a spring and a screw pressing the sample to the corrosion cell, and 6—hole in the vessel (20 mm); (b) model of the equivalent circuit proposed for curve fitting of the EIS data.
Figure 1. (a) Scheme of an electrochemical cell for corrosion resistance studies: 1—reference electrode, 2—auxiliary electrode (platinum electrode), 3—test electrode (specimen), 4—99% ethanol environment, P—potentiostat, 5—insulator with a spring and a screw pressing the sample to the corrosion cell, and 6—hole in the vessel (20 mm); (b) model of the equivalent circuit proposed for curve fitting of the EIS data.
Materials 17 03056 g001
Figure 2. Morphology and topography of the surface of nitrided layers at temperatures of 450 °C, 480 °C, and 520 °C.
Figure 2. Morphology and topography of the surface of nitrided layers at temperatures of 450 °C, 480 °C, and 520 °C.
Materials 17 03056 g002
Figure 3. Cross-sections of the layers nitrided at temperatures of 450 °C, 480 °C, and 520 °C.
Figure 3. Cross-sections of the layers nitrided at temperatures of 450 °C, 480 °C, and 520 °C.
Materials 17 03056 g003
Figure 4. XRD results of (a) 316L substrate and the layers nitrided at temperatures of (b) 450 °C, (c) 480 °C, and (d) 520 °C studied on the surface.
Figure 4. XRD results of (a) 316L substrate and the layers nitrided at temperatures of (b) 450 °C, (c) 480 °C, and (d) 520 °C studied on the surface.
Materials 17 03056 g004
Figure 5. Polarization curves of 316L steel in its initial state exposed to 99% C2H5OH at temperatures of 22° and −30 °C.
Figure 5. Polarization curves of 316L steel in its initial state exposed to 99% C2H5OH at temperatures of 22° and −30 °C.
Materials 17 03056 g005
Figure 6. Potentiodynamic curves of 316L steel and nitrided 316L steel, exposed in ethyl alcohol at 22 °C.
Figure 6. Potentiodynamic curves of 316L steel and nitrided 316L steel, exposed in ethyl alcohol at 22 °C.
Materials 17 03056 g006
Figure 7. Characteristic electrochemical values of 316L steel and nitrided 316L steel exposed in ethyl alcohol at 22 °C.
Figure 7. Characteristic electrochemical values of 316L steel and nitrided 316L steel exposed in ethyl alcohol at 22 °C.
Materials 17 03056 g007
Figure 8. Potentiodynamic curves of 316L steel and nitrided 316L steel (A450, A480, and A520), exposed in ethyl alcohol at −30 °C.
Figure 8. Potentiodynamic curves of 316L steel and nitrided 316L steel (A450, A480, and A520), exposed in ethyl alcohol at −30 °C.
Materials 17 03056 g008
Figure 9. Characteristic electrochemical values of 316L steel and nitrided 316L steel exposed in ethyl alcohol at −30 °C.
Figure 9. Characteristic electrochemical values of 316L steel and nitrided 316L steel exposed in ethyl alcohol at −30 °C.
Materials 17 03056 g009
Figure 10. Surfaces of layers nitrided at temperatures of 480 °C and 520 °C after corrosion tests in ethyl alcohol (22 °C) with corrosion products (red circles) and chemical composition of corrosion products.
Figure 10. Surfaces of layers nitrided at temperatures of 480 °C and 520 °C after corrosion tests in ethyl alcohol (22 °C) with corrosion products (red circles) and chemical composition of corrosion products.
Materials 17 03056 g010
Figure 11. EIS spectra with fitted data of 316L steel with nitrided layers exposed in 99% C2H5OH at 22 °C.
Figure 11. EIS spectra with fitted data of 316L steel with nitrided layers exposed in 99% C2H5OH at 22 °C.
Materials 17 03056 g011
Figure 12. Nyquist spectra with fitted data of 316L steel with nitrided layers exposed in 99% C2H5OH at 22 °C.
Figure 12. Nyquist spectra with fitted data of 316L steel with nitrided layers exposed in 99% C2H5OH at 22 °C.
Materials 17 03056 g012
Figure 13. SEM studies of the surface of stress corrosion of layer nitrided at 520 °C as a result of long-term (3 days) exposure at −30 °C (on bent specimen).
Figure 13. SEM studies of the surface of stress corrosion of layer nitrided at 520 °C as a result of long-term (3 days) exposure at −30 °C (on bent specimen).
Materials 17 03056 g013
Figure 14. EIS spectra with fitted data of 316L steel with nitrided layers exposed in 99% C2H5OH at −30 °C s.
Figure 14. EIS spectra with fitted data of 316L steel with nitrided layers exposed in 99% C2H5OH at −30 °C s.
Materials 17 03056 g014
Figure 15. Nyquist spectra with fitted data of 316L steel with nitrided layers exposed in 99% C2H5OH at −30 °C.
Figure 15. Nyquist spectra with fitted data of 316L steel with nitrided layers exposed in 99% C2H5OH at −30 °C.
Materials 17 03056 g015
Table 1. Chemical composition of surface layer nitride at 480 °C (measurement points 1–4 marked in Figure 2).
Table 1. Chemical composition of surface layer nitride at 480 °C (measurement points 1–4 marked in Figure 2).
ElementAtom [%]SD
Fe66.563.56
Cr19.731.07
Ni7.940.93
N2.961.94
Mn1.440.22
Mo0.730.24
Si0.640.27
Table 2. Microhardness of 316L steel and layers nitrided at temperatures of 450 °C, 480 °C, and 520 °C.
Table 2. Microhardness of 316L steel and layers nitrided at temperatures of 450 °C, 480 °C, and 520 °C.
SpecimenHV0.2SDHV0.5SD
316L steel3366.893223.18
450 °C7969.1548619.54
480 °C7144.7648922.58
520 °C14122.05126617.39
Table 3. Characteristic electrochemical values of 316L steel.
Table 3. Characteristic electrochemical values of 316L steel.
Test TemperatureEcorr [mV] vs. SSCIcorr [µA/cm2]
22 °C−1059.28 × 10−3
−30 °C−1802.15 × 10−3
Table 4. Characteristic electrochemical values of 316L steel and nitrided 316L steel exposed in ethyl alcohol at 22 °C.
Table 4. Characteristic electrochemical values of 316L steel and nitrided 316L steel exposed in ethyl alcohol at 22 °C.
SpecimenEcorr [mV] vs. SSCIcorr [µA/cm2]
316L steel−1059.28 × 10−3
450 °C−1556.43 × 10−3
480 °C−16511.10 × 10−3
520 °C−20512.48 × 10−3
Table 5. Characteristic electrochemical values of 316L steel and nitrided 316L steel exposed in ethyl alcohol at −30 °C.
Table 5. Characteristic electrochemical values of 316L steel and nitrided 316L steel exposed in ethyl alcohol at −30 °C.
SpecimenEcorr [mV] vs. SSCIcorr [µA/cm2]
316L steel−1802.15 × 10−3
450 °C−2152.09 × 10−3
480 °C−2903.60 × 10−3
520 °C−3904.10 × 10−3
Table 6. Characteristic electrochemical values of the metal–electrolyte double layer (99% C2H5OH, room temperature) after glow nitriding processes on 316L steel.
Table 6. Characteristic electrochemical values of the metal–electrolyte double layer (99% C2H5OH, room temperature) after glow nitriding processes on 316L steel.
SpecimenRs
[MΩcm2]
Rct
[MΩcm2]
Error
%
Y0-CPE
[μF/cm2 s(n−1)]
Error
%
nError
%
316L steel0.891.673.911.75.50.762.3
450 °C1.122.548.715.212.20.834.2
480°C0.882.2512.623.65.40.851.8
520 °C1.010.755.922.312.80.844.5
Rs—environmental resistance; Rct—charge transfer resistance through the double layer; Y0-CPE—capacitance of the solid phase element; n—imperfection coefficient of the solid phase element (CPE); error (%)—sum of squares of deviations between measured data points and the calculation line.
Table 7. Characteristic electrochemical values of the metal–electrolyte double layer (99% C2H5OH, room temperature) after glow nitriding processes on 316L steel (−30 °C).
Table 7. Characteristic electrochemical values of the metal–electrolyte double layer (99% C2H5OH, room temperature) after glow nitriding processes on 316L steel (−30 °C).
SpecimenRs
[MΩcm2]
Rct
[MΩcm2]
Error
%
Y0-CPE
[μF/cm2 s(n−1)]
Error
%
nError
%
316L steel1.8019.623.57.327.40.822.4
450 °C1.7333.115.48.555.10.801.7
480 °C1.6323.519.11.475.80.821.8
520 °C1.814.178.65.418.80.712.8
Rs—environmental resistance; Rct—charge transfer resistance through the double layer; Y0-CPE—capacitance of the solid phase element; n—imperfection coefficient of the solid phase element (CPE); error (%)—sum of squares of deviations between measured data points and the calculation line.
Table 8. The contact angle of 316L steel and coatings nitrided at 450 °C, 480 °C, and 520 °C temperature.
Table 8. The contact angle of 316L steel and coatings nitrided at 450 °C, 480 °C, and 520 °C temperature.
SpecimenContact Angle
Distilled Water [°]
SDContact Angle
99% Ethanol [°]
SD
316L steel65.42.32121.11
450 °C85.02.63Total wettability
480 °C84.22.85Total wettability
520 °C96.31.02Total wettability
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kucharska, B.; Kamiński, J.; Kulikowski, K.; Borowski, T.; Sobiecki, J.R.; Wierzchoń, T. The Effect of Nitriding Temperature of AISI 316L Steel on Sub-Zero Corrosion Resistance in C2H5OH. Materials 2024, 17, 3056. https://0-doi-org.brum.beds.ac.uk/10.3390/ma17133056

AMA Style

Kucharska B, Kamiński J, Kulikowski K, Borowski T, Sobiecki JR, Wierzchoń T. The Effect of Nitriding Temperature of AISI 316L Steel on Sub-Zero Corrosion Resistance in C2H5OH. Materials. 2024; 17(13):3056. https://0-doi-org.brum.beds.ac.uk/10.3390/ma17133056

Chicago/Turabian Style

Kucharska, Beata, Janusz Kamiński, Krzysztof Kulikowski, Tomasz Borowski, Jerzy Robert Sobiecki, and Tadeusz Wierzchoń. 2024. "The Effect of Nitriding Temperature of AISI 316L Steel on Sub-Zero Corrosion Resistance in C2H5OH" Materials 17, no. 13: 3056. https://0-doi-org.brum.beds.ac.uk/10.3390/ma17133056

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop