Acessibilidade / Reportar erro

The effect of chloride, sulfate, and ammonium ions on the semiconducting behavior and corrosion resistance of AISI 304 stainless steel passive film

ABSTRACT

Water cooling systems usually receive additional chemical treatment and different features to prevent pitting corrosion on stainless steel which is dependent on medium factors such as species concentration. Chloride is an aggressive ion with which sulfate can act as an inhibitor. By applying cyclic voltammetry, electrochemical impedance spectroscopy (EIS), and Mott-Schottky plots, the behavior of 304 stainless steel (304 SS) was evaluated in solutions containing chloride, sulfate and ammonium ions. The polarization curves indicated the inhibiting effect of sulfate ions against pitting corrosion and the increase of the repassivation potential with increasing ammonium concentration. The most pronounced inhibiting effect occurred with a smaller ratio between chloride and sulfate ions (1:2). By EIS, the values obtained for charge transfer resistance of 304 SS sample corroborated the results of the cyclic voltammetry, ranging between 0.4 and 0.7 MΩcm2. The passive film showed semiconductor behavior, in agreement with the model of the chromium oxide inner layer and iron oxide outer layer. The passive films formed were highly doped, with doping density values on the order of 1020-21 cm–1. The difference between the corrosion potential and the flat band potential was found in the solutions where the material was less susceptible to corrosion.

Keywords
Stainless steel; Passive film; Semiconducting behavior; Aggressive ions

1. INTRODUCTION

The electricity generation in Brazil is dominated by hydropower plants, but thermoelectrics represent 13% of the energy power sector which reached 621.2 TWh in 2020 [1[1] EMPRESA DE PESQUISA ENERGÉTICA, Balanço Energético Nacional - BEN 2021, Brasília, EPE, 2021.]. Due to the need for cooling to remove waste heat, thermoelectric power plants require large amounts of water and usually draw water from rivers or lakes to cool down their units [2[2] BRAVO, H.R., “Assessing the effects of thermal and hydro energy production on water systems”, Journal of Environmental Studies and Sciences, v. 6, n. 1, pp. 140–148, 2016. doi: http://dx.doi.org/10.1007/s13412-016-0382-9.
https://doi.org/10.1007/s13412-016-0382-...
, 3[3] USINA ELÉTRICA A GÁS DE ARAUCÁRIA, Gestão de água em complexo de geração termelétrica, 1 ed., Curitiba, Grafo Estúdio, 2020.]. Water reuse practices consist of the use of alternative sources and the reduction of volumes captured [3[3] USINA ELÉTRICA A GÁS DE ARAUCÁRIA, Gestão de água em complexo de geração termelétrica, 1 ed., Curitiba, Grafo Estúdio, 2020., 4[4] STILLWELL, A.S., MROUE, A.M., RHODES, J.D., et al., “Water for energy: systems integration and analysis to address resource challenges, current sustainable/renewable”, Energy Reports, v. 4, pp. 90–98, 2017. doi: http://dx.doi.org/10.1007/s40518-017-0081-5
https://doi.org/10.1007/s40518-017-0081-...
].

In a recirculating cooling system, water is kept in a closed loop with a cooling tower and uses ambient air to cool its temperature. This kind of system draws much less water from the environment, compared to the open cycle, but consumes more than the closed cycle due to water evaporation [2[2] BRAVO, H.R., “Assessing the effects of thermal and hydro energy production on water systems”, Journal of Environmental Studies and Sciences, v. 6, n. 1, pp. 140–148, 2016. doi: http://dx.doi.org/10.1007/s13412-016-0382-9.
https://doi.org/10.1007/s13412-016-0382-...
4[4] STILLWELL, A.S., MROUE, A.M., RHODES, J.D., et al., “Water for energy: systems integration and analysis to address resource challenges, current sustainable/renewable”, Energy Reports, v. 4, pp. 90–98, 2017. doi: http://dx.doi.org/10.1007/s40518-017-0081-5
https://doi.org/10.1007/s40518-017-0081-...
]. Blowing down is removing part of the circulating water and replacing it with fresh water to decrease the concentration of chemical substances. This process can help to slow down system corrosion. In industrial terms, the main requirements imposed on the quality of cooling water are low temperature, less formation of mineral deposits and biofouling, and prevention of corrosion processes in the equipment and pipelines [2[2] BRAVO, H.R., “Assessing the effects of thermal and hydro energy production on water systems”, Journal of Environmental Studies and Sciences, v. 6, n. 1, pp. 140–148, 2016. doi: http://dx.doi.org/10.1007/s13412-016-0382-9.
https://doi.org/10.1007/s13412-016-0382-...
, 3[3] USINA ELÉTRICA A GÁS DE ARAUCÁRIA, Gestão de água em complexo de geração termelétrica, 1 ed., Curitiba, Grafo Estúdio, 2020., 5[5] OCHKOV, V.F., ORLOV, K.A., IVANOV, E.N., et al., “Calculation and visual displaying of the water chemistry conditions in return cooling systems at thermal power stations”, Thermal Engineering, v. 60, n. 7, pp. 465–471, 2013. doi: http://dx.doi.org/10.1134/S0040601513070070
https://doi.org/10.1134/S004060151307007...
]. The cooling water from rivers or alternative sources usually receives additional treatment inside the plant such as biocide applications or acid addition for pH control [6[6] OCHOA, N., BARIL, G., MORAN, F., et al., “Study of the properties of a multi-component inhibitor used for water treatment in cooling circuits”, Journal of Applied Electrochemistry, v. 32, n. 5, pp. 497–504, 2002. doi: http://dx.doi.org/10.1023/A:1016500722497
https://doi.org/10.1023/A:1016500722497...
, 7[7] MOUDGIL, H.K., YADAV, S., CHAUDHARY, R.S., et al., “Synergistic effect of some antiscalants as corrosion inhibitor for industrial cooling water system”, Journal of Applied Electrochemistry, v. 39, n. 8, pp. 1339–1347, 2009. doi: http://dx.doi.org/10.1007/s10800-009-9807-4
https://doi.org/10.1007/s10800-009-9807-...
]. Thereafter, the ions resulting from the treatment as chloride and sulfate or ions from the source as ammonium still feature in the water [3[3] USINA ELÉTRICA A GÁS DE ARAUCÁRIA, Gestão de água em complexo de geração termelétrica, 1 ed., Curitiba, Grafo Estúdio, 2020., 5[5] OCHKOV, V.F., ORLOV, K.A., IVANOV, E.N., et al., “Calculation and visual displaying of the water chemistry conditions in return cooling systems at thermal power stations”, Thermal Engineering, v. 60, n. 7, pp. 465–471, 2013. doi: http://dx.doi.org/10.1134/S0040601513070070
https://doi.org/10.1134/S004060151307007...
].

Austenitic stainless steels (SS) are useful materials with many industrial applications as they are extensively used in cooling water systems [8[8] GUPTA, P., GUPTA, S., “Investigation of pin-hole leaks in stainless steel water pipes”, International Journal of Systems Assurance Engineering and Management., v. 6, n. 3, pp. 342–349, 2015. doi: http://dx.doi.org/10.1007/s13198-014-0272-z
https://doi.org/10.1007/s13198-014-0272-...
, 9[9] RIOS, J., SHOR, W.W., “Performance of stainless steel tubes in condensers associated with cooling towers”, Journal of Materials for Energy Systems., v. 4, n. 3, pp. 131–135, 1982. doi: http://dx.doi.org/10.1007/BF02833404
https://doi.org/10.1007/BF02833404...
] especially when ammonium is present, and the utilization of copper is not advisable [10[10] LI, K., CHEN, Z., LI, J., et al., “Corrosion mechanism of copper immersed in ammonium sulfate solution”, Materials and Corrosion., v. 69, n. 11, pp. 1597–1608, 2018. doi: http://dx.doi.org/10.1002/maco.201810222
https://doi.org/10.1002/maco.201810222...
]. Their corrosion behavior is dependent on several medium factors such as ion concentration, temperature, and water velocity. Thus, the composition changes in the cooling water promoted by evaporation and blow down process can result in different corrosivity.

The influence of sulfate and chloride on the passive film properties and corrosion resistance of SS is associated with the ion concentration in the solution. Chloride is an aggressive ion since it has a critical level to promote pitting, which is very difficult to prevent [9[9] RIOS, J., SHOR, W.W., “Performance of stainless steel tubes in condensers associated with cooling towers”, Journal of Materials for Energy Systems., v. 4, n. 3, pp. 131–135, 1982. doi: http://dx.doi.org/10.1007/BF02833404
https://doi.org/10.1007/BF02833404...
, 11[11] JAVIDI, M., NEMATOLLAHI, M.R., LALEHPARVAR, M.M., et al., “Failure analysis of AISI 321 Austenitic stainless steel water piping in a power plant”, Journal of Failure Analysis and Prevention, v. 16, n. 2, pp. 209–215, 2016. doi: http://dx.doi.org/10.1007/s11668-016-0070-9
https://doi.org/10.1007/s11668-016-0070-...
]. On the other hand, the corrosion risk decreases with increasing sulfate concentration which acts as a pitting inhibitor [12[12] PYUN, S.I., MOON, S.M., PYUN, S., “ The inhibition mechanism of pitting corrosion of pure aluminum by nitrate and sulfate ions in neutral chloride solution”, Journal of Solid State Electrochemistry, v. 3, n. 6, pp. 331–336, 1999. doi: http://dx.doi.org/10.1007/s100080050163
https://doi.org/10.1007/s100080050163...
, 13[13] POHJANNE, P., CARPÉN, L., KINNUNEN, P., et al., Stainless steel pitting in chloride sulfate solutions: the role of cations, Nashville, Tennessee, United States of America, NACE International, 2007.]. The effect of ammonium still is not totally known, and fewer experimental studies are conducted in an environment containing ammonium, chloride, and sulfate. YANG et al. [14[14] YANG, J., WANG, S., XU, D., et al., “Effect of ammonium chloride on corrosion behavior of Ni-based alloys and stainless steel in supercritical water gasification process”, International Journal of Hydrogen Energy, v. 42, n. 31, pp. 19788–19797, 2017. doi: http://dx.doi.org/10.1016/j.ijhydene.2017.05.078
https://doi.org/10.1016/j.ijhydene.2017....
] reported that the presence of ammonium and chloride ions in supercritical water provided a severe corrosion environment for 316 SS promoting pitting and intergranular corrosion. According to TIAN et al. [15[15] TIAN, H., SUN, F., CHU, F., et al., “Passivation behavior and surface chemistry of 316 SS in the environment containing Cl- and NH4+”, Journal of Electroanalytical Chemistry, v. 886, pp. 115138, 2021. doi: http://dx.doi.org/10.1016/j.jelechem.2021.115138
https://doi.org/10.1016/j.jelechem.2021....
] the passive film formed on 316 SS in ammonium chloride (NH4Cl) solution is more compact and has a better barrier effect on the aggressive ions, compared with the sodium chloride (NaCl) solution with the same concentration. Conversely, the corrosion resistance of 316 SS is minor with the increase in the concentration of chloride and ammonium ions.

The corrosion resistance of SS is also dependent on the composition and structure of the passive film formed on the metal surface. SS passive films are often reported as having a duplex structure, consisting of an inner barrier of the oxide film and outer oxide or hydroxide film, acting as p–n heterojunction [16[16] LI, D.G., WANG, J.D., CHEN, D.R., et al., “Influences of pH value, temperature, chloride ions and sulfide ions on the corrosion behaviors of 316L stainless steel in the simulated cathodic environment of proton exchange membrane fuel cell”, Journal of Power Sources, v. 272, pp. 448–456, 2014. doi: http://dx.doi.org/10.1016/j.jpowsour.2014.06.121
https://doi.org/10.1016/j.jpowsour.2014....
, 17[17] FERREIRA, M.G.S., DA CUNHA BELO, M., HAKIKI, N.E., et al., “Semiconducting properties of oxide and passive films formed on AISI 304 stainless steel and Alloy 600”, Journal of the Brazilian Chemical Society, v. 13, n. 4, pp. 433–440, 2002. doi: http://dx.doi.org/10.1590/S0103-50532002000400005
https://doi.org/10.1590/S0103-5053200200...
] When these oxides have perfect crystalline structure and stoichiometry, they act as insulators, but the presence of defects in the structure provides characteristics of extrinsic semiconductors when exposed to an aqueous solution [18[18] FERNÁNDEZ-DOMENE, R.M., BLASCO-TAMARIT, E., GARCÍA-GARCÍA, D.M., et al., “Effect of alloying elements on the electronic properties of thin passive films formed on carbon steel, ferritic and austenitic stainless steels in a highly concentrated LiBr solution”, Thin Solid Films, v. 558, pp. 252–258, 2014. doi: http://dx.doi.org/10.1016/j.tsf.2014.03.042
https://doi.org/10.1016/j.tsf.2014.03.04...
]. In this work, the Mott-Schottky equation (Equation 1) was applied to estimate the reciprocal of the square of the capacitance of the oxides formed on SS in different solutions.

C2=2kBTNqqεε0EEfbkBTq.(1)

Where: Nq is the doping density, ε is the dielectric constant, ε0 denotes the vacuum permittivity, q is the elementary charge, kB is the Boltzmann constant, T denotes temperature, Efb is the flat band potential, and E is the applied potential.

It was possible to study the properties and fundamental parameters of the semiconductor behavior of the SS passive films and relate them to the corrosion resistance of SS [19[19] FATTAH-ALHOSSEINI, A., VAFAEIAN, S., “Comparison of electrochemical behavior between coarse-grained and fine-grained AISI 430 ferritic stainless steel by Mott-Schottky analysis and EIS measurements”, Journal of Alloys and Compounds, v. 639, pp. 301–307, 2015. doi: http://dx.doi.org/10.1016/j.jallcom.2015.03.142
https://doi.org/10.1016/j.jallcom.2015.0...
,20[20] FATTAH-ALHOSSEINI, A., “Passivity of AISI 321 stainless steel in 0.5 M H2SO4 solution studied by Mott-Schottky analysis in conjunction with the point defect model”, Arabian Journal of Chemistry, v. 9, pp. S1342-S1348, 2016. doi: http://dx.doi.org/10.1016/j.arabjc.2012.02.015
https://doi.org/10.1016/j.arabjc.2012.02...
,21[21] MANDEL, M., KIETOV, V., HORNIG, R., et al., “On the polarisation and Mott-Schottky characteristics of an Fe-Mn-Al-Ni shape-memory alloy and pure Fe in NaCl-free and NaCl-contaminated Ca(OH)2, at solution: a comparative study”, Corrosion Science, v. 179, pp. 109172, 2021. doi: http://dx.doi.org/10.1016/j.corsci.2020.109172
https://doi.org/10.1016/j.corsci.2020.10...
,22[22] OJE, A.M., OGWU, A.A., RAHMAN, S.U., et al., “Effect of temperature variation on the corrosion behaviour and semiconducting properties of the passive film formed on chromium oxide coatings exposed to saline solution”, Corrosion Science, v. 154, pp. 28–35, 2019. doi: http://dx.doi.org/10.1016/j.corsci.2019.04.004
https://doi.org/10.1016/j.corsci.2019.04...
,23[23] TRANCHIDA, G., DI FRANCO, F., MEGNA, B., et al., “Semiconducting properties of passive films and corrosion layers on weathering steel”, Electrochimica Acta, v. 354, pp. 136697, 2020. doi: http://dx.doi.org/10.1016/j.electacta.2020.136697
https://doi.org/10.1016/j.electacta.2020...
,24[24] DAROWICKI, K., KRAKOWIAK, S., ŚLEPSKI, P., “Selection of measurement frequency in Mott-Schottky analysis of passive layer on nickel”, Electrochimica Acta, v. 51, n. 11, pp. 2204–2208, 2006. doi: http://dx.doi.org/10.1016/j.electacta.2005.04.079
https://doi.org/10.1016/j.electacta.2005...
]. In the present work, measurements of Mott–Schottky plots, electrochemical impedance spectroscopy (EIS), and polarization curves were used to investigate the influence of chloride, sulfate, and ammonium ions on the semiconducting properties of passive films and the corrosion resistance of AISI 304 SS.

2. MATERIALS AND METHODS

2.1. Chemical and materials

All solutions were prepared with deionized water and the following reagents were analytical grade and used as received: sodium chloride - NaCl (Synth), sodium sulfate - Na2SO4 (Synth), ammonium hydroxide - NH4OH (Synth), potassium chloride - KCl (Aldrich), and sodium borate - Na2B4O7·10H2O (Aldrich).

The solutions were prepared to simulate the cooling water conditions in a sodium borate buffer at pH 8.0 and varying concentrations of the reagents: NaCl, Na2SO4, and NH4OH, whose concentrations were reported in Table 1.

Table 1
Solution compositions to study the effect of chloride, sulfate, and ammonium ions on the semiconducting behavior of SS passive film.

The electrochemical measurements were performed in a setup employing a three-electrode cell using a Gamry 3000 potentiostat. AISI 304 SS in a cylindrical shape with a 0.6 cm diameter, cold mounted in epoxy resin was featured as the working electrode, a platinum foil as the counter electrode, and a saturated calomel electrode (SCE) as the reference electrode.

2.2. Characterization of the SS

The SS samples were procured in a cylindrical format, possessing a diameter of 0.6 cm, and analyzed as received.

Characterization of the SS microstructure was carried out through metallography. The samples were wet ground with silica papers having granulometry of #180 – #240 – #400 – #600 – #1500. Afterward, the samples were polished with alumina 1.0 µm and finely polished with alumina 0.3 µm. Electrolytic etching was performed with a solution of 10.0% (w/w) oxalic acid with an applied potential of 6 V for 60 s. A Zeiss Axio Vert. A1 inverted optical microscope was used to obtain the micrographs.

Phase identification through X-ray diffractometry was also employed using a Bruker D8 Advance Eco diffractometer with Bragg Brentano geometry. The scanning was done in the 2θ range of 40° to 100°, at a scan rate of 6° min–1 using a Cu-Kα radiation source (λ = 1.5406 Å).

2.3. Electrochemical measurements

Before all the electrochemical tests, the samples were wet ground with silica papers with a granulometry of #180 – #240 – #400 – #600 – #1500 – #2000. Afterward, an electrochemical pre-treatment was applied using a 5 mA.cm–2 cathodic current for 10 minutes.

The potentiodynamic polarization curves were obtained at a potential scan rate of 5 mV.s–1 with starting potential at –2.00 V, and when it reached +1.60 V, the direction was reversed until it reached the repassivation potential. For the EIS and capacitance measurements by Mott-Schottky plots, the passive film was formed by applying a constant potential of –0.10 V for 2 h. The EIS measurements were performed at the same film formation potential, applying an excitation voltage of 10 mV (amplitude) with a frequency ranging from 100 kHz to 10–2 Hz. Mott-Shottky plots were obtained by starting the potential scan at –0.10 V and concluding the experiment at –1.50 V, with a 50 mV step at 2000 Hz.

All the experiments were carried out at room temperature and to ensure the repeatability of the experiment, the curves were performed leastwise 3 times.

3. RESULTS AND DISCUSSION

3.1. Characterization

Figure 1 exhibits the optical micrography of the SS. Austenite grains containing shear bands (arrows with designation A) and twins (arrows with designation B) mainly constitute the microstructure of the AISI 304 SS [25[25] HE, W., LI, F., ZHANG, H., et al., “The influence of cold rolling deformation on tensile properties and microstructures of Mn18Cr18 N austenitic stainless steel”, Materials Science and Engineering A, v. 764, pp. 138245, 2019. doi: http://dx.doi.org/10.1016/j.msea.2019.138245
https://doi.org/10.1016/j.msea.2019.1382...
,26[26] MARQUES, A.V.M., DO CARMO, K.M., LAGE, W.C., et al., “Evaluation of the effect of plastic deformation on the microstructure, hardness, and magnetic properties of AISI type 316L stainless steel”, Revista Materia., v. 25, pp. 1–10, 2020. doi: http://dx.doi.org/10.1590/S1517-707620200002.1011
https://doi.org/10.1590/S1517-7076202000...
,27[27] RAMIREZ, A.H., RAMIREZ, C.H., COSTA, I., “Influence of cold deformation on pitting corrosion resistance of ISO NBR 5832-1 austenitic stainless steel used for orthopedic implants”, Journal of the Brazilian Chemical Society, v. 25, pp. 1270–1274, 2014. doi: http://dx.doi.org/10.5935/0103-5053.20140105
https://doi.org/10.5935/0103-5053.201401...
]. The presence of shear bands indicates that this material bar was strengthened through cold work. As is widely known, this can induce the formation of martensite and defects such as deformation bands through plastic deformation of the austenite grain, which can decrease its corrosion resistance [27[27] RAMIREZ, A.H., RAMIREZ, C.H., COSTA, I., “Influence of cold deformation on pitting corrosion resistance of ISO NBR 5832-1 austenitic stainless steel used for orthopedic implants”, Journal of the Brazilian Chemical Society, v. 25, pp. 1270–1274, 2014. doi: http://dx.doi.org/10.5935/0103-5053.20140105
https://doi.org/10.5935/0103-5053.201401...
, 28[28] SOLOMON, N., SOLOMON, I., “Effect of deformation-induced phase transformation on AISI 316 stainless steel corrosion resistance”, Engineering Failure Analysis, v. 79, pp. 865–875, 2017. doi: http://dx.doi.org/10.1016/j.engfailanal.2017.05.031
https://doi.org/10.1016/j.engfailanal.20...
]. Moreover, the microstructure was also composed of rounded particles uniformly distributed, probably inclusions resulting from the SS bar manufacturing process [29[29] GHAZANI, M.S., EGHBALI, B., “Characterization of the hot deformation microstructure of AISI 321 austenitic stainless steel”, Materials Science and Engineering A, v. 730, pp. 380–390, 2018. doi: http://dx.doi.org/10.1016/j.msea.2018.06.025
https://doi.org/10.1016/j.msea.2018.06.0...
, 30[30] BEZERRA, E.D.O.T., NASCIMENTO, J.J.D.S., LUNA, C.B.B., et al., “Avaliação de não conformidades de próteses de quadril fabricadas com ligas de titânio e aço inox”, Revista Materia., v. 22, n. 1, pp. e11782, 2017. doi: http://dx.doi.org/10.1590/s1517-707620170001.0114
https://doi.org/10.1590/s1517-7076201700...
]. Nevertheless, the material exhibited a step structure in the austenite grain boundaries, indicating that the material was not sensitized.

Figure 1
Optical micrographs of the SS, electrolytically etched with 10% oxalic acid at 6 V for 60 s. Magnification of 200×. Austenite grains with shear bands (A) and twins (B).

The XRD analysis shown in Figure 2 indicated the material was mainly composed of austenite (γ) since only its main peaks were identified in the diffractogram. This confirms that the cold deformed state of the sample did not produce any martensite that could affect the corrosion resistance of the SS and that no significant deleterious phase content, such as intermetallic (σ phase) and carbide content, was identified. Therefore, the particles exhibited in the micrograph (Figure 1) were probably inclusions resulting from the manufacturing process. Moreover, the microstructure and phases of the studied material were in the as-received state, and hence faithfully represented the structure of the commercial SS widely applied in many industrial sectors.

Figure 2
X-Ray diffractogram of the SS in the 2θ range of 40° to 100°. The γ symbol represents the austenite phase.

3.2. Pitting corrosion

To understand the effects of the ions on the SS corrosion resistance, the results were separated into two different ammonium concentrations and ratios between the concentration of chloride and sulfate. Figure 3a shows the comparison between solutions C and E (Table 1), with the same concentrations of chloride and sulfate but different ammonium concentrations. The electrochemical processes were not significantly different and both ammonium concentrations showed similar corrosion potential (Ecorr) and passive potential range. However, the repassivation potential (Erepass) values were different. The highest concentration had more difficulty in stopping the pitting processes and returning to the passivated state.

Figure 3
Polarization curves (5 mV·s–1) of SS at (a) different ammonium ion concentrations and (b) different chloride and sulfate ion concentrations, where C, E, B, D, and A were the concentrations of the solutions used (Table 1).

The substantial difference brought about by the variation in ammonia concentration was evident in the highlighted anodic peak in the figure. In the polarization curves an anodic peak was detected at 1.10 V and was attributed to a passive film change, resulting from the oxidation reaction from Cr (III) to Cr (VI), which occurred close to the pitting potential [31[31] KOCIJAN, A., DONIK, Č., JENKO, M., “Electrochemical and XPS studies of the passive film formed on stainless steels in borate buffer and chloride solutions”, Corrosion Science, v. 49, n. 5, pp. 2083–2098, 2007. doi: http://dx.doi.org/10.1016/j.corsci.2006.11.001
https://doi.org/10.1016/j.corsci.2006.11...
]. The increase in the concentration of ammonium promoted a localized increase in the pH value, favoring the oxidation reaction of chromium and enhancing the peak current. According to TIAN et al. [15[15] TIAN, H., SUN, F., CHU, F., et al., “Passivation behavior and surface chemistry of 316 SS in the environment containing Cl- and NH4+”, Journal of Electroanalytical Chemistry, v. 886, pp. 115138, 2021. doi: http://dx.doi.org/10.1016/j.jelechem.2021.115138
https://doi.org/10.1016/j.jelechem.2021....
] initially, the ammonium ion contributes to the formation of the passive film on the SS surface, however, these cations hydrolyze and produce excess H+, resulting in the value of pH reduction which will promote the growth and propagation process of pitting. As depicted in Figure 3b, following this peak, a subsequent rise in current was observed as the scan reaches 1.37 V, coinciding with the initiation of pit corrosion.

The parameters obtained from polarization curves for SS in all solutions investigated were summarized in Table 2. The pitting potential (Epit) is the potential value that stable pits start to grow, and repassivation (Erepass) which is reached after reversal of the potential sweep direction is the potential value below which the already growing pits are repassivated and the growth is stopped. The maximum current is linked to the breakdown of the passive film and the highest level of corrosion. For Solution B, this value was attained at the scan’s limit potential. In solutions with a higher concentration of sulfate ions, the breakdown of the passive film is hindered, resulting in less pitting corrosion on the material. Consequently, the measured current is also lower.

Table 2
Corrosion parameters obtained from polarization curves for SS in solutions A-E with different chloride, sulfate, and ammonium concentrations.

With constant ammonium concentration, the corrosion resistance of SS was dependent on the proportion of chloride and sulfate ions (Cl:SO42–). The passive range was smaller for solution B, with a higher ratio between chloride and sulfate ions (2:1). The pitting potential (Epit in red - Figure 3b) value was lower, about 0.7 V and the current density reached 9 mAcm–2, indicating intense corrosion pitting.

Previous studies have reported that chloride and sulfate ions participate in the pitting corrosion process since chloride is an aggressive ion and sulfate is an inhibitor. For the sulfate inhibitor effect to occur, a lower chloride/sulfate ionic ratio is needed [13[13] POHJANNE, P., CARPÉN, L., KINNUNEN, P., et al., Stainless steel pitting in chloride sulfate solutions: the role of cations, Nashville, Tennessee, United States of America, NACE International, 2007., 16[16] LI, D.G., WANG, J.D., CHEN, D.R., et al., “Influences of pH value, temperature, chloride ions and sulfide ions on the corrosion behaviors of 316L stainless steel in the simulated cathodic environment of proton exchange membrane fuel cell”, Journal of Power Sources, v. 272, pp. 448–456, 2014. doi: http://dx.doi.org/10.1016/j.jpowsour.2014.06.121
https://doi.org/10.1016/j.jpowsour.2014....
] The inhibitor action of the sulfate ions can be observed by increasing the value of the pitting potential to more positive values [13[13] POHJANNE, P., CARPÉN, L., KINNUNEN, P., et al., Stainless steel pitting in chloride sulfate solutions: the role of cations, Nashville, Tennessee, United States of America, NACE International, 2007., 32[32] LAITINEN, T., “Localized corrosion of stainless steel in chloride, sulfate and thiosulfate containing environments”, Corrosion Science, v. 42, n. 3, pp. 421–441, 2000. doi: http://dx.doi.org/10.1016/S0010-938X(99)00072-4
https://doi.org/10.1016/S0010-938X(99)00...
, 33[33] EL-EGAMY, S.S., BADAWAY, W.A., “Passivity and passivity breakdown of 304 stainless steel in alkaline sodium sulphate solutions”, Journal of Applied Electrochemistry, v. 34, n. 11, pp. 1153–1158, 2004. doi: http://dx.doi.org/10.1007/s10800-004-1709-x
https://doi.org/10.1007/s10800-004-1709-...
].

According to EL-EGAMY and BADAWAY [33[33] EL-EGAMY, S.S., BADAWAY, W.A., “Passivity and passivity breakdown of 304 stainless steel in alkaline sodium sulphate solutions”, Journal of Applied Electrochemistry, v. 34, n. 11, pp. 1153–1158, 2004. doi: http://dx.doi.org/10.1007/s10800-004-1709-x
https://doi.org/10.1007/s10800-004-1709-...
], the inhibitor effect of sulfate does not include a stage of propagation of pitting; it only hinders the initiation process. This inhibitor effect on the pitting initiation could be observed by the Epit, changing the ratio Cl:SO42– 2:1 to 1:1 or 1:2. However, in the solutions where the ratios between chloride and sulfate ions were 1:1 and 1:2, respectively, the pitting potential (Epit in black - Figure 3b) was the same as for 1.30 V. At this potential, the electrolyte reaction also took place and the local acidification from H+ formation promoted the breakdown of passivation, so pitting corrosion occurred at the electrode.

Variations in the maximum current density and repassivation potential could be observed, indicating different intensities of pitting corrosion. At the 1:2 ratio in solutions C and E, the highest sulfate concentration promoted lower current density and more positive repassivation potential values. In the case of the 1:1 ratio, the corrosion increased with the total ion concentration, thus, in the more concentrated solution D the current reached a bigger value.

The pits on the surfaces of 304 SS generated at different chloride and sulfate ion concentrations with greater (solution B) and lower corrosion (solution C) were shown in Figure 4. The results were consistent with the potentiodynamic polarization curves. Solution B with ratio Cl:SO42– 2:1 showed a greater number of pits as shown in Figure 4a and the width of the pinholes reached 68 µm (Figure 4b). On the other hand, solution C with ratio Cl:SO42– 1:2 showed less and thinner (28 µm) pits as shown in Figures 4c and 4d.

Figure 4
Optical micrographs of the 304 SS after potentiodynamic polarization, electrolytically etched with 10% oxalic acid at 6 V for 60 s. Solution B: (a) magnification of 50×; (b) magnification of 500× and Solution C: (c) magnification of 50×; (d) magnification of 500×.

The pitting number density difference was not obvious with the increase in ammonium concentration, indicating that the addition of NH4+ had little influence on the transition of metastable to stable pitting corrosion [34[34] TIAN, H., FAN, L., LI, Y., et al., “Effect of NH4+ on the pitting corrosion behavior of 316 stainless steel in the chloride environment”, Journal of Electroanalytical Chemistry, v. 894, pp. 115368, 2021. doi: http://dx.doi.org/10.1016/j.jelechem.2021.115368
https://doi.org/10.1016/j.jelechem.2021....
]. The growth rate of pits was controlled by diffusion, independent of the ions in the solution, thus in solution B the pitting corrosion starts before, with a smaller Epit resulting in greater pinholes. Pitting corrosion extended along the depth direction and along the circular direction. The presented morphology showed a single initiation pit surrounded by a few other small holes that assemble to form a larger hole [35[35] ERNST, P., NEWMAN, R.C., “Pit growth studies in stainless steel foils. I. Introduction and pit growth kinetics”, Corrosion Science, v. 44, n. 5, pp. 927–941, 2002. doi: http://dx.doi.org/10.1016/S0010-938X(01)00133-0
https://doi.org/10.1016/S0010-938X(01)00...
].

3.3. Passive film properties

The passive film was potentiostatically formed at –0.10 V, which was below the repassivation for all solutions. The EIS measurements were performed at the passive film formation potential and the Nyquist plots are shown in Figure 5a. The equivalent circuit used to adjust the EIS data was shown in Figure 5b. This circuit was also simulated in previous works for SS in different medium [19[19] FATTAH-ALHOSSEINI, A., VAFAEIAN, S., “Comparison of electrochemical behavior between coarse-grained and fine-grained AISI 430 ferritic stainless steel by Mott-Schottky analysis and EIS measurements”, Journal of Alloys and Compounds, v. 639, pp. 301–307, 2015. doi: http://dx.doi.org/10.1016/j.jallcom.2015.03.142
https://doi.org/10.1016/j.jallcom.2015.0...
, 23[23] TRANCHIDA, G., DI FRANCO, F., MEGNA, B., et al., “Semiconducting properties of passive films and corrosion layers on weathering steel”, Electrochimica Acta, v. 354, pp. 136697, 2020. doi: http://dx.doi.org/10.1016/j.electacta.2020.136697
https://doi.org/10.1016/j.electacta.2020...
, 31[31] KOCIJAN, A., DONIK, Č., JENKO, M., “Electrochemical and XPS studies of the passive film formed on stainless steels in borate buffer and chloride solutions”, Corrosion Science, v. 49, n. 5, pp. 2083–2098, 2007. doi: http://dx.doi.org/10.1016/j.corsci.2006.11.001
https://doi.org/10.1016/j.corsci.2006.11...
].

Figure 5
(a) Nyquist plots and (b) equivalent circuit for adjustment of the EIS measurements performed in different ammonium, chloride, and sulfate concentrations, where A, B, C, D, and E were the concentration of the solutions used (Table 1).

The series resistance (R1) accounts for the resistance of the electrolyte, cables, and current collectors, while the charge transfer resistance (R2) refers to the charge transfer at the electrode/electrolyte interface, and the constant phase element (CPE) relates to the double-layer capacitance at the electrode/electrolyte interface (QDL). The model assumes that the passive film is composed of a non-homogeneous layer with the presence of defects. Together with the electrode’s rough surface and some adsorbed species, the behavior can deviate from an ideal capacitor. The parameter α can range from 0 to 1, where unity represents a perfectly flat electrode. The parameters obtained through fitting EIS data are shown in Table 3.

Table 3
Parameters obtained through fitting EIS data for 304 SS in solutions A-E with different chloride, sulfate, and ammonium concentrations.

R1 variations were related to solution concentration. R2 values, which were related to the resistance to 304 SS oxidation and its dissolution through the passive film, were about 0.4–0.7 MΩ.cm2, indicating greater corrosion resistance at the potential applied. The results for R2 in the different solutions agreed with the polarization curves. The charge transfer resistance was lower for solution B with Cl:SO42– 2:1 and greater for Cl:SO42– 1:2 ratio. The increase in the ammonium concentration produced a less resistant passive film. The capacitances were calculated with Brug’s Equation [36[36] BRUG, G.J., VAN DEN EEDEN, A.L.G., SLUYTERS-REHBACH, M., et al., “The analysis of electrode impedances complicated by the presence of a constant phase element”, Journal of Electroanalytical Chemistry, v. 176, n. 1-2, pp. 275–295, 1984. doi: http://dx.doi.org/10.1016/S0022-0728(84)80324-1
https://doi.org/10.1016/S0022-0728(84)80...
] and the obtained values remained virtually unchanged.

The Mott-Schottky plots (Figure 6) showed the variation of capacitance values with the potential applied, which was similar to the behavior reported in the literature for SS [17[17] FERREIRA, M.G.S., DA CUNHA BELO, M., HAKIKI, N.E., et al., “Semiconducting properties of oxide and passive films formed on AISI 304 stainless steel and Alloy 600”, Journal of the Brazilian Chemical Society, v. 13, n. 4, pp. 433–440, 2002. doi: http://dx.doi.org/10.1590/S0103-50532002000400005
https://doi.org/10.1590/S0103-5053200200...
, 18[18] FERNÁNDEZ-DOMENE, R.M., BLASCO-TAMARIT, E., GARCÍA-GARCÍA, D.M., et al., “Effect of alloying elements on the electronic properties of thin passive films formed on carbon steel, ferritic and austenitic stainless steels in a highly concentrated LiBr solution”, Thin Solid Films, v. 558, pp. 252–258, 2014. doi: http://dx.doi.org/10.1016/j.tsf.2014.03.042
https://doi.org/10.1016/j.tsf.2014.03.04...
, 37[37] GE, H.H., XU, X.M., ZHAO, L., et al., “Semiconducting behavior of passive film formed on stainless steel in borate buffer solution containing sulfide”, Journal of Applied Electrochemistry, v. 41, n. 5, pp. 519–525, 2011. doi: http://dx.doi.org/10.1007/s10800-011-0272-5
https://doi.org/10.1007/s10800-011-0272-...
, 38[38] HAKIKI, N.E., “Influence of surface roughness on the semiconducting properties of oxide films formed on 304 stainless steel”, Journal of Applied Electrochemistry, v. 38, n. 5, pp. 679–687, 2008. doi: http://dx.doi.org/10.1007/s10800-008-9487-5
https://doi.org/10.1007/s10800-008-9487-...
]. Region I, from –0.10 V to –0.40 V, presented a positive slope, indicating an n-type semiconductor behavior of the iron oxide (γ-Fe2O3). In Region (III), between –0.70 V and –1.05 V, a negative slope was noted, consistent with the p-type semiconductor behavior of the chromium oxide (Cr2O3) inner layer. Region II was characterized by the flat band potential region and the capacitance was related to the Helmholtz layer.

Figure 6
Mott-Schottky plots for 304 SS in different ammonium, chloride, and sulfate concentrations, where A, B, C, D, and E were the concentrations of the solutions used (Table 1).

The densities of dopants and the flat band potential were shown in Table 4. The dielectric constant (ε) used was 15.6 which was like the reported [15[15] TIAN, H., SUN, F., CHU, F., et al., “Passivation behavior and surface chemistry of 316 SS in the environment containing Cl- and NH4+”, Journal of Electroanalytical Chemistry, v. 886, pp. 115138, 2021. doi: http://dx.doi.org/10.1016/j.jelechem.2021.115138
https://doi.org/10.1016/j.jelechem.2021....
, 20[20] FATTAH-ALHOSSEINI, A., “Passivity of AISI 321 stainless steel in 0.5 M H2SO4 solution studied by Mott-Schottky analysis in conjunction with the point defect model”, Arabian Journal of Chemistry, v. 9, pp. S1342-S1348, 2016. doi: http://dx.doi.org/10.1016/j.arabjc.2012.02.015
https://doi.org/10.1016/j.arabjc.2012.02...
, 39[39] KIM, J.J., YOUNG, Y.M., “Study on the passive film of type 316 stainless steel”, International Journal of Electrochemical Science, v. 8, n. 10, pp. 11847–11859, Oct. 2013. doi: http://dx.doi.org/10.1016/S1452-3981(23)13227-5
https://doi.org/10.1016/S1452-3981(23)13...
]. The measured capacitance was associated with the space charge layer formed in the passive layer of the 304 SS on contact with the different solutions [40[40] VERBRUGGEN, F., FISET, E., BONIN, L., et al., “Stainless steel substrate pretreatment effects on copper nucleation and stripping during copper electrowinning”, Journal of Applied Electrochemistry, v. 51, n. 2, pp. 219–233, 2021. doi: http://dx.doi.org/10.1007/s10800-020-01485-2
https://doi.org/10.1007/s10800-020-01485...
, 41[41] FERREIRA, M.G.S., HAKIKI, N.E., GOODLET, G., et al., “Influence of the temperature of film formation on the electronic structure of oxide films formed on 304 stainless steel”, Electrochimica Acta, v. 46, n. 24-25, pp. 3767–3776, 2001. doi: http://dx.doi.org/10.1016/S0013-4686(01)00658-2
https://doi.org/10.1016/S0013-4686(01)00...
]. The values of the donor densities were lower than those of the acceptor densities due to the difference in space charge layer thickness developed by the iron oxide and the chromium oxide layers [38[38] HAKIKI, N.E., “Influence of surface roughness on the semiconducting properties of oxide films formed on 304 stainless steel”, Journal of Applied Electrochemistry, v. 38, n. 5, pp. 679–687, 2008. doi: http://dx.doi.org/10.1007/s10800-008-9487-5
https://doi.org/10.1007/s10800-008-9487-...
]. The higher concentration of species in the solution favored the formation of defects in the oxide structure and increased the number of dopants, aligning with the findings of [42[42] FATTAH-ALHOSSEINI, A., GOLOZAR, M.A., SAATCHI, A., et al., “Effect of solution concentration on semiconducting properties of passive films formed on austenitic stainless steels”, Corrosion Science, v. 52, n. 1, pp. 205–209, 2010. doi: http://dx.doi.org/10.1016/j.corsci.2009.09.003
https://doi.org/10.1016/j.corsci.2009.09...
]. The small increase in the ammonium concentration almost did not change the dopant densities. The change in the number of dopants was more pronounced in the p-type chromium oxides as can be visualized by the slope of the linear region. Even though the densities of dopants did not exhibit an evident correlation with the corrosion resistance [15[15] TIAN, H., SUN, F., CHU, F., et al., “Passivation behavior and surface chemistry of 316 SS in the environment containing Cl- and NH4+”, Journal of Electroanalytical Chemistry, v. 886, pp. 115138, 2021. doi: http://dx.doi.org/10.1016/j.jelechem.2021.115138
https://doi.org/10.1016/j.jelechem.2021....
]. Since the protection performance of SS passive film was closely related to chromium oxides, these changes could affect the corrosion.

Table 4
Semiconducting parameters from Mott-Schottky plots for 304 SS in solutions A-E with different chloride, sulfate, and ammonium concentrations.

Solutions containing higher concentrations of sulfate had more negative flat-band potential values. Anions adsorbed on the passive film cause an increase in negative surface charge, which results in a decrease in the potential drop in the Helmholtz layer, resulting in a negative change in the flat band potential [43[43] KONG, D.S., CHEN, S.H., WANG, C., et al., “A study of the passive films on chromium by capacitance measurement”, Corrosion Science, v. 45, n. 4, pp. 747–758, Apr. 2003. doi: http://dx.doi.org/10.1016/S0010-938X(02)00148-8
https://doi.org/10.1016/S0010-938X(02)00...
].

Solutions C and E showed less pitting corrosion, and the flat band potentials from p-type were more negative. When the corrosion potential is more positive than the flat band potential, the p-type semiconductor will be in an accumulation situation and the n-type semiconductor will form a depletion layer that will inhibit electron transfer, which agrees with the EIS measurements [44[44] HARRINGTON, S.P., DEVINE, T.M., “The influence of the semiconducting properties of passive films on localized corrosion rates”, ECS Transactions, v. 16, n. 52, pp. 117–123, Aug. 2009. doi: http://dx.doi.org/10.1149/1.3229960.
https://doi.org/10.1149/1.3229960...
]. The corrosion resistance was more subject to this effect than the dopant densities for the evaluated concentrations.

4. CONCLUSIONS

This study was conducted in solutions containing ammonium, chloride, and sulfate to evaluate the effect of these ions on the semiconducting behavior and corrosion resistance of 304 stainless steel passive film. From the polarization curves, it was possible to observe the inhibiting effect of sulfate ions on SS. This effect was more pronounced in the solution where the ratio between the molar concentration of the aggressive chloride ions and the sulfate inhibitor ions was 1:2. On the other hand when the proportion was 2:1, there was a virtual absence of inhibiting effect. The corrosion behavior was not significantly different with the increase in the ammonium concentration, however, the repassivation potential (Erepass) presented a less positive value indicating more difficulty in stopping the pitting processes and returning to the passivated state. The EIS results agreed with the polarization curves, and the charge transfer resistance values were lower in the solutions with a higher ratio between chloride and sulfate ions (2:1).

The Mott-Schottky plots showed different semiconductor properties: an n-type behavior of the iron oxide (γ-Fe2O3) outer layer and a p-type behavior of the chromium oxide (Cr2O3) inner layer. The presence of sulfate ions in greater molar proportion concerning chloride ions promoted the formation of a more resistive passive film, regardless of the defect density in the crystal structure. The higher concentration of sulfate promotes a pitting inhibitory effect, due to preferential sulfate adsorption, but it also made the passive film more resistant due to the more negative flat band potentials.

Therefore, the control of species in recirculating water needs to be done in terms of concentration and in proportion between them.

5. ACKNOWLEDGMENTS

The authors are grateful for the infrastructure and support regarding both human and financial resources from Usina Elétrica a Gás de Araucária – UEGA and ANEEL (PD 0539-0004/2015); CNPq process 308777/2020-4, DT 1C, the Lactec; and UFPR/PIPE.

6. BIBLIOGRAFIA

  • [1]
    EMPRESA DE PESQUISA ENERGÉTICA, Balanço Energético Nacional - BEN 2021, Brasília, EPE, 2021.
  • [2]
    BRAVO, H.R., “Assessing the effects of thermal and hydro energy production on water systems”, Journal of Environmental Studies and Sciences, v. 6, n. 1, pp. 140–148, 2016. doi: http://dx.doi.org/10.1007/s13412-016-0382-9.
    » https://doi.org/10.1007/s13412-016-0382-9
  • [3]
    USINA ELÉTRICA A GÁS DE ARAUCÁRIA, Gestão de água em complexo de geração termelétrica, 1 ed., Curitiba, Grafo Estúdio, 2020.
  • [4]
    STILLWELL, A.S., MROUE, A.M., RHODES, J.D., et al, “Water for energy: systems integration and analysis to address resource challenges, current sustainable/renewable”, Energy Reports, v. 4, pp. 90–98, 2017. doi: http://dx.doi.org/10.1007/s40518-017-0081-5
    » https://doi.org/10.1007/s40518-017-0081-5
  • [5]
    OCHKOV, V.F., ORLOV, K.A., IVANOV, E.N., et al, “Calculation and visual displaying of the water chemistry conditions in return cooling systems at thermal power stations”, Thermal Engineering, v. 60, n. 7, pp. 465–471, 2013. doi: http://dx.doi.org/10.1134/S0040601513070070
    » https://doi.org/10.1134/S0040601513070070
  • [6]
    OCHOA, N., BARIL, G., MORAN, F., et al, “Study of the properties of a multi-component inhibitor used for water treatment in cooling circuits”, Journal of Applied Electrochemistry, v. 32, n. 5, pp. 497–504, 2002. doi: http://dx.doi.org/10.1023/A:1016500722497
    » https://doi.org/10.1023/A:1016500722497
  • [7]
    MOUDGIL, H.K., YADAV, S., CHAUDHARY, R.S., et al, “Synergistic effect of some antiscalants as corrosion inhibitor for industrial cooling water system”, Journal of Applied Electrochemistry, v. 39, n. 8, pp. 1339–1347, 2009. doi: http://dx.doi.org/10.1007/s10800-009-9807-4
    » https://doi.org/10.1007/s10800-009-9807-4
  • [8]
    GUPTA, P., GUPTA, S., “Investigation of pin-hole leaks in stainless steel water pipes”, International Journal of Systems Assurance Engineering and Management., v. 6, n. 3, pp. 342–349, 2015. doi: http://dx.doi.org/10.1007/s13198-014-0272-z
    » https://doi.org/10.1007/s13198-014-0272-z
  • [9]
    RIOS, J., SHOR, W.W., “Performance of stainless steel tubes in condensers associated with cooling towers”, Journal of Materials for Energy Systems., v. 4, n. 3, pp. 131–135, 1982. doi: http://dx.doi.org/10.1007/BF02833404
    » https://doi.org/10.1007/BF02833404
  • [10]
    LI, K., CHEN, Z., LI, J., et al, “Corrosion mechanism of copper immersed in ammonium sulfate solution”, Materials and Corrosion., v. 69, n. 11, pp. 1597–1608, 2018. doi: http://dx.doi.org/10.1002/maco.201810222
    » https://doi.org/10.1002/maco.201810222
  • [11]
    JAVIDI, M., NEMATOLLAHI, M.R., LALEHPARVAR, M.M., et al, “Failure analysis of AISI 321 Austenitic stainless steel water piping in a power plant”, Journal of Failure Analysis and Prevention, v. 16, n. 2, pp. 209–215, 2016. doi: http://dx.doi.org/10.1007/s11668-016-0070-9
    » https://doi.org/10.1007/s11668-016-0070-9
  • [12]
    PYUN, S.I., MOON, S.M., PYUN, S., “ The inhibition mechanism of pitting corrosion of pure aluminum by nitrate and sulfate ions in neutral chloride solution”, Journal of Solid State Electrochemistry, v. 3, n. 6, pp. 331–336, 1999. doi: http://dx.doi.org/10.1007/s100080050163
    » https://doi.org/10.1007/s100080050163
  • [13]
    POHJANNE, P., CARPÉN, L., KINNUNEN, P., et al, Stainless steel pitting in chloride sulfate solutions: the role of cations, Nashville, Tennessee, United States of America, NACE International, 2007.
  • [14]
    YANG, J., WANG, S., XU, D., et al, “Effect of ammonium chloride on corrosion behavior of Ni-based alloys and stainless steel in supercritical water gasification process”, International Journal of Hydrogen Energy, v. 42, n. 31, pp. 19788–19797, 2017. doi: http://dx.doi.org/10.1016/j.ijhydene.2017.05.078
    » https://doi.org/10.1016/j.ijhydene.2017.05.078
  • [15]
    TIAN, H., SUN, F., CHU, F., et al, “Passivation behavior and surface chemistry of 316 SS in the environment containing Cl- and NH4+”, Journal of Electroanalytical Chemistry, v. 886, pp. 115138, 2021. doi: http://dx.doi.org/10.1016/j.jelechem.2021.115138
    » https://doi.org/10.1016/j.jelechem.2021.115138
  • [16]
    LI, D.G., WANG, J.D., CHEN, D.R., et al, “Influences of pH value, temperature, chloride ions and sulfide ions on the corrosion behaviors of 316L stainless steel in the simulated cathodic environment of proton exchange membrane fuel cell”, Journal of Power Sources, v. 272, pp. 448–456, 2014. doi: http://dx.doi.org/10.1016/j.jpowsour.2014.06.121
    » https://doi.org/10.1016/j.jpowsour.2014.06.121
  • [17]
    FERREIRA, M.G.S., DA CUNHA BELO, M., HAKIKI, N.E., et al, “Semiconducting properties of oxide and passive films formed on AISI 304 stainless steel and Alloy 600”, Journal of the Brazilian Chemical Society, v. 13, n. 4, pp. 433–440, 2002. doi: http://dx.doi.org/10.1590/S0103-50532002000400005
    » https://doi.org/10.1590/S0103-50532002000400005
  • [18]
    FERNÁNDEZ-DOMENE, R.M., BLASCO-TAMARIT, E., GARCÍA-GARCÍA, D.M., et al, “Effect of alloying elements on the electronic properties of thin passive films formed on carbon steel, ferritic and austenitic stainless steels in a highly concentrated LiBr solution”, Thin Solid Films, v. 558, pp. 252–258, 2014. doi: http://dx.doi.org/10.1016/j.tsf.2014.03.042
    » https://doi.org/10.1016/j.tsf.2014.03.042
  • [19]
    FATTAH-ALHOSSEINI, A., VAFAEIAN, S., “Comparison of electrochemical behavior between coarse-grained and fine-grained AISI 430 ferritic stainless steel by Mott-Schottky analysis and EIS measurements”, Journal of Alloys and Compounds, v. 639, pp. 301–307, 2015. doi: http://dx.doi.org/10.1016/j.jallcom.2015.03.142
    » https://doi.org/10.1016/j.jallcom.2015.03.142
  • [20]
    FATTAH-ALHOSSEINI, A., “Passivity of AISI 321 stainless steel in 0.5 M H2SO4 solution studied by Mott-Schottky analysis in conjunction with the point defect model”, Arabian Journal of Chemistry, v. 9, pp. S1342-S1348, 2016. doi: http://dx.doi.org/10.1016/j.arabjc.2012.02.015
    » https://doi.org/10.1016/j.arabjc.2012.02.015
  • [21]
    MANDEL, M., KIETOV, V., HORNIG, R., et al, “On the polarisation and Mott-Schottky characteristics of an Fe-Mn-Al-Ni shape-memory alloy and pure Fe in NaCl-free and NaCl-contaminated Ca(OH)2, at solution: a comparative study”, Corrosion Science, v. 179, pp. 109172, 2021. doi: http://dx.doi.org/10.1016/j.corsci.2020.109172
    » https://doi.org/10.1016/j.corsci.2020.109172
  • [22]
    OJE, A.M., OGWU, A.A., RAHMAN, S.U., et al, “Effect of temperature variation on the corrosion behaviour and semiconducting properties of the passive film formed on chromium oxide coatings exposed to saline solution”, Corrosion Science, v. 154, pp. 28–35, 2019. doi: http://dx.doi.org/10.1016/j.corsci.2019.04.004
    » https://doi.org/10.1016/j.corsci.2019.04.004
  • [23]
    TRANCHIDA, G., DI FRANCO, F., MEGNA, B., et al, “Semiconducting properties of passive films and corrosion layers on weathering steel”, Electrochimica Acta, v. 354, pp. 136697, 2020. doi: http://dx.doi.org/10.1016/j.electacta.2020.136697
    » https://doi.org/10.1016/j.electacta.2020.136697
  • [24]
    DAROWICKI, K., KRAKOWIAK, S., ŚLEPSKI, P., “Selection of measurement frequency in Mott-Schottky analysis of passive layer on nickel”, Electrochimica Acta, v. 51, n. 11, pp. 2204–2208, 2006. doi: http://dx.doi.org/10.1016/j.electacta.2005.04.079
    » https://doi.org/10.1016/j.electacta.2005.04.079
  • [25]
    HE, W., LI, F., ZHANG, H., et al, “The influence of cold rolling deformation on tensile properties and microstructures of Mn18Cr18 N austenitic stainless steel”, Materials Science and Engineering A, v. 764, pp. 138245, 2019. doi: http://dx.doi.org/10.1016/j.msea.2019.138245
    » https://doi.org/10.1016/j.msea.2019.138245
  • [26]
    MARQUES, A.V.M., DO CARMO, K.M., LAGE, W.C., et al, “Evaluation of the effect of plastic deformation on the microstructure, hardness, and magnetic properties of AISI type 316L stainless steel”, Revista Materia., v. 25, pp. 1–10, 2020. doi: http://dx.doi.org/10.1590/S1517-707620200002.1011
    » https://doi.org/10.1590/S1517-707620200002.1011
  • [27]
    RAMIREZ, A.H., RAMIREZ, C.H., COSTA, I., “Influence of cold deformation on pitting corrosion resistance of ISO NBR 5832-1 austenitic stainless steel used for orthopedic implants”, Journal of the Brazilian Chemical Society, v. 25, pp. 1270–1274, 2014. doi: http://dx.doi.org/10.5935/0103-5053.20140105
    » https://doi.org/10.5935/0103-5053.20140105
  • [28]
    SOLOMON, N., SOLOMON, I., “Effect of deformation-induced phase transformation on AISI 316 stainless steel corrosion resistance”, Engineering Failure Analysis, v. 79, pp. 865–875, 2017. doi: http://dx.doi.org/10.1016/j.engfailanal.2017.05.031
    » https://doi.org/10.1016/j.engfailanal.2017.05.031
  • [29]
    GHAZANI, M.S., EGHBALI, B., “Characterization of the hot deformation microstructure of AISI 321 austenitic stainless steel”, Materials Science and Engineering A, v. 730, pp. 380–390, 2018. doi: http://dx.doi.org/10.1016/j.msea.2018.06.025
    » https://doi.org/10.1016/j.msea.2018.06.025
  • [30]
    BEZERRA, E.D.O.T., NASCIMENTO, J.J.D.S., LUNA, C.B.B., et al, “Avaliação de não conformidades de próteses de quadril fabricadas com ligas de titânio e aço inox”, Revista Materia., v. 22, n. 1, pp. e11782, 2017. doi: http://dx.doi.org/10.1590/s1517-707620170001.0114
    » https://doi.org/10.1590/s1517-707620170001.0114
  • [31]
    KOCIJAN, A., DONIK, Č., JENKO, M., “Electrochemical and XPS studies of the passive film formed on stainless steels in borate buffer and chloride solutions”, Corrosion Science, v. 49, n. 5, pp. 2083–2098, 2007. doi: http://dx.doi.org/10.1016/j.corsci.2006.11.001
    » https://doi.org/10.1016/j.corsci.2006.11.001
  • [32]
    LAITINEN, T., “Localized corrosion of stainless steel in chloride, sulfate and thiosulfate containing environments”, Corrosion Science, v. 42, n. 3, pp. 421–441, 2000. doi: http://dx.doi.org/10.1016/S0010-938X(99)00072-4
    » https://doi.org/10.1016/S0010-938X(99)00072-4
  • [33]
    EL-EGAMY, S.S., BADAWAY, W.A., “Passivity and passivity breakdown of 304 stainless steel in alkaline sodium sulphate solutions”, Journal of Applied Electrochemistry, v. 34, n. 11, pp. 1153–1158, 2004. doi: http://dx.doi.org/10.1007/s10800-004-1709-x
    » https://doi.org/10.1007/s10800-004-1709-x
  • [34]
    TIAN, H., FAN, L., LI, Y., et al, “Effect of NH4+ on the pitting corrosion behavior of 316 stainless steel in the chloride environment”, Journal of Electroanalytical Chemistry, v. 894, pp. 115368, 2021. doi: http://dx.doi.org/10.1016/j.jelechem.2021.115368
    » https://doi.org/10.1016/j.jelechem.2021.115368
  • [35]
    ERNST, P., NEWMAN, R.C., “Pit growth studies in stainless steel foils. I. Introduction and pit growth kinetics”, Corrosion Science, v. 44, n. 5, pp. 927–941, 2002. doi: http://dx.doi.org/10.1016/S0010-938X(01)00133-0
    » https://doi.org/10.1016/S0010-938X(01)00133-0
  • [36]
    BRUG, G.J., VAN DEN EEDEN, A.L.G., SLUYTERS-REHBACH, M., et al, “The analysis of electrode impedances complicated by the presence of a constant phase element”, Journal of Electroanalytical Chemistry, v. 176, n. 1-2, pp. 275–295, 1984. doi: http://dx.doi.org/10.1016/S0022-0728(84)80324-1
    » https://doi.org/10.1016/S0022-0728(84)80324-1
  • [37]
    GE, H.H., XU, X.M., ZHAO, L., et al, “Semiconducting behavior of passive film formed on stainless steel in borate buffer solution containing sulfide”, Journal of Applied Electrochemistry, v. 41, n. 5, pp. 519–525, 2011. doi: http://dx.doi.org/10.1007/s10800-011-0272-5
    » https://doi.org/10.1007/s10800-011-0272-5
  • [38]
    HAKIKI, N.E., “Influence of surface roughness on the semiconducting properties of oxide films formed on 304 stainless steel”, Journal of Applied Electrochemistry, v. 38, n. 5, pp. 679–687, 2008. doi: http://dx.doi.org/10.1007/s10800-008-9487-5
    » https://doi.org/10.1007/s10800-008-9487-5
  • [39]
    KIM, J.J., YOUNG, Y.M., “Study on the passive film of type 316 stainless steel”, International Journal of Electrochemical Science, v. 8, n. 10, pp. 11847–11859, Oct. 2013. doi: http://dx.doi.org/10.1016/S1452-3981(23)13227-5
    » https://doi.org/10.1016/S1452-3981(23)13227-5
  • [40]
    VERBRUGGEN, F., FISET, E., BONIN, L., et al, “Stainless steel substrate pretreatment effects on copper nucleation and stripping during copper electrowinning”, Journal of Applied Electrochemistry, v. 51, n. 2, pp. 219–233, 2021. doi: http://dx.doi.org/10.1007/s10800-020-01485-2
    » https://doi.org/10.1007/s10800-020-01485-2
  • [41]
    FERREIRA, M.G.S., HAKIKI, N.E., GOODLET, G., et al, “Influence of the temperature of film formation on the electronic structure of oxide films formed on 304 stainless steel”, Electrochimica Acta, v. 46, n. 24-25, pp. 3767–3776, 2001. doi: http://dx.doi.org/10.1016/S0013-4686(01)00658-2
    » https://doi.org/10.1016/S0013-4686(01)00658-2
  • [42]
    FATTAH-ALHOSSEINI, A., GOLOZAR, M.A., SAATCHI, A., et al, “Effect of solution concentration on semiconducting properties of passive films formed on austenitic stainless steels”, Corrosion Science, v. 52, n. 1, pp. 205–209, 2010. doi: http://dx.doi.org/10.1016/j.corsci.2009.09.003
    » https://doi.org/10.1016/j.corsci.2009.09.003
  • [43]
    KONG, D.S., CHEN, S.H., WANG, C., et al, “A study of the passive films on chromium by capacitance measurement”, Corrosion Science, v. 45, n. 4, pp. 747–758, Apr. 2003. doi: http://dx.doi.org/10.1016/S0010-938X(02)00148-8
    » https://doi.org/10.1016/S0010-938X(02)00148-8
  • [44]
    HARRINGTON, S.P., DEVINE, T.M., “The influence of the semiconducting properties of passive films on localized corrosion rates”, ECS Transactions, v. 16, n. 52, pp. 117–123, Aug. 2009. doi: http://dx.doi.org/10.1149/1.3229960.
    » https://doi.org/10.1149/1.3229960

Publication Dates

  • Publication in this collection
    23 Oct 2023
  • Date of issue
    2023

History

  • Received
    02 May 2023
  • Accepted
    01 Sept 2023
Laboratório de Hidrogênio, Coppe - Universidade Federal do Rio de Janeiro, em cooperação com a Associação Brasileira do Hidrogênio, ABH2 Av. Moniz Aragão, 207, 21941-594, Rio de Janeiro, RJ, Brasil, Tel: +55 (21) 3938-8791 - Rio de Janeiro - RJ - Brazil
E-mail: revmateria@gmail.com