Next Article in Journal
Fiber-Reinforced Asphalt Mixture Design on Anti-Skid Surfacing for Field Testing High-Speed Vehicles on Pavements
Next Article in Special Issue
Some Brassicaceae Extracts as Potential Antioxidants and Green Corrosion Inhibitors
Previous Article in Journal
Effect of Nematogen Doping in Bent-Core Molecular Systems with a Helical Nanofilament and Dark Conglomerate
Previous Article in Special Issue
Comparing the Corrosion Resistance of 5083 Al and Al2O33D/5083 Al Composite in a Chloride Environment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electrochemical Impedance Analysis for Corrosion Rate Monitoring of Sol–Gel Protective Coatings in Contact with Nitrate Molten Salts for CSP Applications

by
V. Encinas-Sánchez
1,
A. Macías-García
2,
M. T. de Miguel
1,
F. J. Pérez
1 and
J. M. Rodríguez-Rego
2,*
1
Surface Engineering and Nanostructured Materials Research Group, Complutense University of Madrid, Complutense Avenue s/n, 28040 Madrid, Spain
2
Department of Mechanical, Energetic and Materials Engineering, School of Industrial Engineering, University of Extremadura, Avda. de Elvas, s/n, 06006 Badajoz, Spain
*
Author to whom correspondence should be addressed.
Materials 2023, 16(2), 546; https://doi.org/10.3390/ma16020546
Submission received: 9 November 2022 / Revised: 17 December 2022 / Accepted: 26 December 2022 / Published: 5 January 2023
(This article belongs to the Special Issue Corrosion Resistance Enhancement of the Materials Surface)

Abstract

:
The protective behaviour of ZrO2-3%molY2O3 sol–gel coatings, deposited with an immersion coating technique on 9Cr-1Mo P91 steel, was evaluated with corrosion monitoring sensors using the electrochemical impedance spectroscopy technique. The tests were carried out in contact with solar salt at 500 °C for a maximum of 2000 h. The results showed the highly protective behaviour of the coating, with the corrosion process in the coated system being controlled by the diffusion of charged particles through the protective layer. The coating acts by limiting the transport of ions and slowing down the corrosive process. The system allowed a reduction in the corrosion rate of uncoated P91 steel. The estimated corrosion rate of 22.62 μm·year−1 is lower than that accepted for in-service operations. The proposed ZrO2-3%molY2O3 sol–gel coatings are an option to mitigate the corrosion processes caused by the molten salts in concentrated solar power plants.

1. Introduction

Interest in Concentrated Solar Power (CSP) plants has increased in recent years [1]. This interest has grown steadily due to its high potential for improvements in efficiency and dispatchability when compared with other renewable energy technologies [2]. However, better dispatchability remains a crucial issue for increasing its competitiveness [3], and thermal energy storage could provide a realistic solution [4]. Although many studies are being performed with other salt mixtures [5,6,7], the most industrially used compound is 60 wt.% NaNO3/40 wt.% KNO3 (Solar Salt®) [8].
The degradation of materials in contact with molten salts in CSP plants has been studied by different researchers [9,10,11,12,13,14,15]. These studies recommend the use of carbon steel at temperatures below 300 °C, stainless steels between 300 °C and 550 °C and Ni-based alloys for temperatures above 550 °C [16]. CSP technologies are expensive and require the use of efficient and cheaper materials [17]. The replacement of stainless steels and Ni-based alloys (known for their high costs [18]) with low-alloy steels could be a solution [19]. However, these steels in contact with molten nitrate salts do not offer good corrosion behaviour [20]. Other alloys with low Cr content, such as P91 and X20CrMoV, also did not show good corrosion behaviour in studies carried out during 2500 h at 600 °C [18,21]. P91 steel in contact with molten nitrates for 1000 h at 580 °C showed less pronounced corrosion [21]. Other authors explained this behaviour by the formation of two layers, a superficial Fe2O3 layer and a protective interior layer, rich in chromium oxide, for a 1000 h test. At longer times, the protective chromium oxide layer decreases [22]. Therefore, the development of protective coatings for these alloys could be an economical alternative against corrosion for industrial applications in contact with molten salts [21]. Sol–gel coatings seem to be an interesting option because of their numerous advantages [23], including their low processing temperature and the ability to be deposited on complex shapes.
The sol–gel process uses a precursor solution as a protective coating for a certain substrate. This solution is transformed into a gel layer by the evaporation of the solvent and/or the chemical reactions that take place [24]. The use of yttria-stabilized zirconia (YSZ) sol–gel solutions is based on the properties provided by ZrO2 with good thermal shock resistance, mechanical and chemical properties on the one hand, and Y2O3 with thermal stability and anti-aging performance on the other hand [25,26]. Both components allow for good high-temperature stability of YSZ [27]. Previous studies have been promising [28]. P91 steel was dip-coated using a sol–gel ZrO2-3%molY2O3, and the results were comparable to those obtained with uncoated 304 steel. SEM micrographs corresponding to the above sol–gel solution gave a compact coating with a thickness ranging from 1 to 1.4 μm. Additionally, YSZ has been used as a protective coating in molten vanadates and sulphate media [29,30]). These media are well known for being highly corrosive, which suggests the great potential of YSZ as a protective material in molten Solar Salt environments.
Most molten salt corrosion analysis studies have been performed using techniques such as optical microscopy, scanning electron microscopy energy/scatter X-ray spectroscopy and X-ray diffraction [8,31,32]. However, these techniques used are laborious and complex [33]. Techniques that allow the register of corrosion rate and mechanism in real-time can help in gaining a better understanding of the corrosion processes. Electrochemical techniques are a good solution for the corrosion monitoring of materials, especially at high temperatures in the presence of molten salts [34].
Electrochemical Impedance Spectroscopy (EIS) is a technique that allows the recording of experimental data in real-time [33,35]. The main advantage of this technique is the low intensity of the excitation signal required and a reduction of the error rate associated with the measurement process [36,37]. This technique allows us to evaluate the corrosion process and to determine the corrosion rate [33]. For this purpose, the steel in contact with molten salt can be represented by circuits composed of resistance, capacitance and inductance elements under an alternating current [38].
The EIS technique has been previously used in many studies for evaluating the corrosion behaviour of different substrates in contact with molten salts (such as nitrates/nitrites [39,40], chlorides [41] and vanadates [42]). Corrosion investigations by EIS on different materials, such as 316 stainless steels in molten HITEC salts [39], Inconel 718 superalloy in molten Na2SO4, 80V2O5-20Na2SO4, NaVO3 [42], ferritic–martensitic steel with molten NaNO3/KNO3 [33] and 9Cr-1Mo steel in molten LiCl-KCl salt [34] showed that the corrosion processes exhibited different mechanisms. Zhu et al. [39] found that the corrosion of 316 stainless steels in molten salt HITEC was controlled by the outward diffusion of metal ions, while Jagadeeswara-Rao et al. [34] registered the formation of intermittent oxide films in 9Cr-1Mo steels in molten salt LiCl-KCl. Thus, the main purpose of this study is to assess the corrosion resistance of ZrO2- Y2O3 sol–gel coatings on 9Cr-1Mo P91 ferritic–martensitic steel in contact with Solar Salt at 500 °C for up to 2000 h by employing corrosion monitoring sensors that are based on the EIS technique.

2. Methodology

2.1. Materials

2.1.1. Preparation of the Nitrate Salt Mixture and Steel Samples

The nitrate mixture, 60 wt.% NaNO3/40 wt.% KNO3, was prepared using NaNO3 from BASF (Ludwigshafen, Germany) with 99% purity and KNO3 from Haifa (Madrid, Spain) with 98% purity. The required quantity of each compound was weighed and mixed in an alumina crucible. The impurity level present in both nitrates is gathered in Table 1.
The substrate used as the base material for CSP applications was 9Cr-1Mo P91 steel, (weight composition of 0.12% C, 0.21% Si, 0.49% Mn, 0.014% P, 0.002% S, 0.01% Al, 8.70% Cr, 0.85% Mo, 0.02% Ni, 0.18% V, 0.06% Nb and 0.053% N). The substrates were machined to a size of 20 × 10 × 2 mm3 and subsequently sanded and polished using 240, 600 and 800 grit sandpaper and 9 μm, 6 μm and 3 μm polishing cloths.

2.1.2. Coating Preparation and Deposition

The procedure of the sol–gel solution has been previously described in [27,43].
The coatings on the P91 substrates with the above solution were carried out using the immersion technique [27,43] at an extraction velocity of 25 mm-min–1 and subsequent heat treatment at 500 °C in a Hobersal® Furnace (Barcelona, Spain) for 2 h, at a heating/cooling velocity of 3 °C-min–1.
In order to reduce stresses in the coating, a drying process was performed on the coated samples before being heat-treated, thus minimizing the organic residue content in the coating [44]. This initial drying phase was carried out at 100 °C for 60 min, applying the previously mentioned heating/cooling rate. In the sintering process of the coatings, the formation of cracks occurs due to the existing tensions in the crystallisation process and/or thermal expansion; to inhibit them, slow heating/cooling ramps are carried out [45].
The structural quality and morphology of the coatings as deposited have been deeply analysed in previous research and can be consulted in [43].

2.2. Electrochemical Impedance Corrosion Study

Electrochemical impedance spectroscopy (EIS) allows observing the corrosion of the coated P91 samples in contact with the molten binary salt during 2000 h. For this purpose, points of the coated P91 sample were electrically connected by welding to a Kanthal wire. This joint was protected by the application of a ceramic slurry mixture. The wire was introduced in an alumina tube sealed with the same ceramic mixture avoiding contact between the sample–wire connection and the molten salts (Figure 1). The EIS uses the working electrode (WE), the auxiliary electrode (AE) and the reference electrode (RE) (patent reference code WO2017046427).
The crucible containing the molten salt mixture was placed in an electric camera furnace (Carbolite, Hope Valley, UK) with the electrodes immersed.
The surface of each electrode was in contact with the liquid salt (60 wt.% NaNO3/40 wt.% KNO3) at 500 °C, and to a depth of about 3.5 cm. EIS measurements were taken at 0, 24, 72, 500, 1000, 1500 and 2000 h.
It has been reported that this salt mixture starts to degrade at 535 °C [46]. The temperature selected for this study was 500 °C in order to assure the stability of the salt during the experiments. In addition, the parabolic shape of the reflecting mirrors used in the parabolic troughs makes the heat flux in the inferior part of the absorption tube greater than in the superior part, giving a non-uniform distribution of the temperature and thermal stresses that are accentuated above 500 °C. The presence of thermal stresses can deform the absorber tube and deviate the focal line, which affects its optical properties [47].
The EIS measurements were taken by means of VoltaLab 80 equipment (Radiometer Analytical SAS, Villeurbanne, France). The amplitude of the voltage perturbation was fixed at 10 mV with a sweep frequency from 50 kHz to 10 MHz. Impedance data fitting and simulations were performed with the software Zview (version 2018).

3. Results and Discussion

3.1. Electrochemical Impedance Corrosion Test

The 3% yttria-doped zirconia sol–gel was used to coat P91 steel by dip-coating and subsequent sintering heat treatment. Coated and uncoated samples are shown in Figure 2.
As expected, and in line with the previously reported study [48], coated samples showed good uniformity given the high homogeneity in colour, which was a methodology proposed by Morrow et al. [49].
Methods used to date for analysing the protective behaviour of the proposed coatings against molten salt corrosion are based on conventional techniques. This typical procedure is tedious and time-consuming, which makes it unsuitable for monitoring corrosion in real-time. EIS measurements represent an interesting solution for achieving rigorous and controlled degradation monitoring of the coated samples in molten salts. The corrosion monitoring system allowed the recording of information on the corrosive process, allowing, inter alia, the estimation of the corrosion rate.
Figure 3 shows the impedance spectra and the best-fitted equivalent electrical circuit of the coated P91 samples in molten binary salt at various times. The electrochemical data registered were adjusted to different models in order to assess the corrosion mechanism taking place. The results showed that the system is consistent with the theoretical model of the protective layer at all measuring times. The best fit was determined by a goodness-of-fit test, analysing the chi-square and relative error values. In line with this model, the impedance spectra show two semicircles, with the one at the high frequency being smaller. This loop at high frequencies refers to the charges that are circulating in the interface between the material and the electrolyte. Here, Re represents the molten-salt (electrolyte) resistance, and Rt is the electrochemical transfer resistance.
This transfer resistance is related by the simplified Butler–Volmer equation (Equation (1)), which reveals that it has an inverse relationship with the current density exchanged.
R t = R · T n · F · i o
where R is the gas constant, T is temperature, n is the number of electrons involved, F is the Faraday constant and io is the exchange current density. Cdl is the capacitance of the double layer that appears at the coating/salt interface, and ndl/ncp represent, respectively, the constant phase element coefficient of the first and second capacitance loops. It is common for the EIS experiments that capacitors do not behave ideally. There are several theories regarding the cause of this deviation, e.g., surface roughness, non-uniform current distribution or varying thickness or composition of the oxide scale [39,40]. To avoid this non-ideal behaviour, without dependence on its origin, constant phase elements (CPE) are used in the equivalent circuits instead of pure capacitors. In Table 2, Cdl and Ccp are the modulus values of the two constant phase elements (CPE) and ndl and ncp are their respective indices.
The protective layer model indicates that ion transport in the layer is the stage that limits the process and slows down the corrosion process. From a physicochemical point of view, the layer resembles a capacitor in series with the double-layer capacitance (see Figure 4). Thus, Ccp and Rcp respectively represent the protective layer capacitance and its resistance to the charged particles transfer. By way of example, Figure 5 shows the adjustment made after 2000 h of testing, including the models for a protective layer, localized corrosion and porous layer.
Therefore, given the good adjustment of the experimental data to the protective layer model, the good behaviour of the protective coatings throughout the test can be confirmed. However, it is worth noting that the behaviour of the system is unstable in the early stages of the test, with the formation of the two semicircles becoming clearer after 72 h of testing (see Figure 3b). This initial instability may be related to the weight loss that this type of coating suffers during the initial stages of the test due to the excess coating deposited [28], which implies that the system becomes more protective after this initial stage. Additionally, as previously reported in [33], a corrosion-monitoring system usually requires a few hours for stabilisation, which may lead to instability in the measurements taken during the initial stages.
Table 2 contains the values of each of the equivalent circuit elements in compliance with each testing time with the protective layer model. According to the obtained results and as shown in Figure 6, the resistance of the electrolyte (Re) (i.e., the electrical resistance of the molten salt) remains roughly constant in a range of between 5.051 Ω and 7.557 Ω. However, the resistance peak of 7.557 Ω obtained after 24 h of the test is attributed to the instability of the system during the initial stages. This instability is also visible when observing the remaining parameters. Thus, without considering the values obtained during the stability period (first 72 h of testing), it may be affirmed that the molten salt resistance has an average value of 5.914 ± 0.820 Ω. This value differs from that obtained in other studies [33,50,51], which may be due to the coating material that was detached during the initial stages of the test [28]. ZrO2 is a semiconductor material and leads to an increase in molten salt electrical resistance [50]. Additionally, the presence of coating components in the salt and the formation of volatiles throughout the test led to a reduction in ionic species concentration and, therefore, a reduction in electrolyte conductivity. The effect of the system on the Re value seems to be evident according to the dissimilar values found in those works cited [33,50,51].
By comparing charge transfer resistance (Rt) and protective layer resistance (Rcp), it is observed that Rcp is higher than Rt at every testing time (see Figure 7).
This difference implies that the corrosion process is controlled by the transport of charged particles moving through the protective layer since it is the slower process [50]. This fact is indicative of the good behaviour of the coating. In addition, it is important to highlight the significant reduction in time in both parameters. The decrease in both resistances reveals that the transference of charges and charged particles is eased so the kinetic of the degradation processes can be accelerated. This, together with the increase in Ccp (see Figure 7), may be due to the degradation of the coating over time. However, after 1000 h of testing, the system seems to be more stable, leading to slower degradation.
Finally, with regard to the exponents ndl and ncp, they remained constant during the entire test and remained below 1 (see Figure 8). These two exponents show a value of 0.79 ± 0.02 and 0.68 ± 0.06, respectively. According to Omar et al. [52], an n value equal to 1 indicates that the system acts as an ideal capacitor, which would be the closest to an ideal coating.

Corrosion Rate Estimation

As explained above, electrochemical impedance monitoring allows not only the evaluation of the controlling corrosion mechanism but also an estimation of the corrosion rate along the experiment. To this end, following the ASTM-G102 Standard Practice [53], Equation (2) was used for calculating the corrosion rate:
v corr = K · i corr · EW ρ
where vcorr is the corrosion rate in μm·cm−2, K is a system-type dependent constant, icorr is the corrosion current density, EW is the equivalent weight of the material and ρ is its density. For P91 steel, the parameters K, EW and ρ are, respectively, 3.27·10−3 μg·μA−2·cm−1·year−1, 25.3, and 7.76 g·cm−3 [33]. The current density is given by the Stern–Geary equation [54]:
i corr = K · B R p
where B is the Stern–Geary constant (26 mV) and Rp is the polarisation resistance obtained from the experimental values of Re and Zreal:
Z real = R e + R p
Thus, Figure 9 summarises the corrosion rate determined from the EIS test results during the entire experiment. It is important to highlight that the method used considers generalised corrosion.
Several results can be highlighted. Firstly, the corrosion rate of 22.62 μm·year−1 of the coated steel, estimated after 2000 h of testing, falls well below those estimated in other studies for uncoated P91 steel (118 μm·year−1 [16] and 300 μm·year−1 [33]). Additionally, according to the guide for corrosion rates used in the industry, where recommendations concerning the use of materials in molten salts are considered (Table 3 [55]), it can be observed that the estimated corrosion rate is closer to the upper limit set for materials recommended for long-term services (0.4–13 μm·year−1) [55]. This fact suggests that the proposed coated system could be an interesting option from an industrial point of view.
Furthermore, and considering the corrosion rates estimated at each testing time (see Figure 9), the initial stabilisation period is clearly observed (0–72 h), as well as the good behaviour of the coating during the intermediate periods (72–1000 h). Within the intermediate times, the corrosion rate remains within the range recommended by the industry for this material in long-term tests (see Table 3 [55]). However, the increase in the corrosion rate after 1000 h of testing is worth noting. Even though the protective properties of the coating are still suitable after 2000 h of testing (according to the EIS spectra), the coating is steadily degraded.

4. Conclusions

The protective behaviour of sol–gel ZrO2–3%molY2O3 coatings deposited using a dip-coating technique on 9Cr-1Mo P91 steel was isothermally assessed at 500 °C for up to 2000 h in contact with Solar Salt by employing corrosion monitoring sensors that use the electrochemical impedance spectroscopy technique.
The obtained results showed the highly protective character of the proposed coating, which acts by limiting ion transport, thus slowing down the corrosive process. According to the values obtained in terms of charge transfer resistance and coating layer resistance, the corrosive process in the coated system is controlled by the transport of charged particles moving through the protective layer, which indicates that this is the stage that limits the process, confirming the protective behaviour. The results made it possible to estimate the corrosion rate of the coated system in contact with Solar Salt at 500 °C, this being 22.62 μm·year−1. The estimated corrosion rate falls well below those estimated for uncoated P91 steel and those accepted for in-service operations over 1 year, being even closer to the upper limit established for materials recommended for long-term services.
The results suggest that the proposed coating system could be an interesting option for industrial concentrated solar power plants as a potential solution to the severe corrosion issues that currently affect industrial tanks and pipes operating in contact with Solar Salt.

Author Contributions

V.E.-S.: Conceptualization, Investigation, Validation, Writing—Original Draft; A.M.-G.: Methodology, Visualization, Writing—Review & Editing; M.T.d.M.: Conceptualization, Validation, Investigation, Writing—Review & Editing; F.J.P.: Conceptualization, Supervision, Funding acquisition; J.M.R.-R.: Visualization, Writing—Review & Editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work received funding from the Agencia Estatal de Investigation in the frame of the “Proyectos de I+D+I «Programación Conjunta Internacional» 2020” program and under project reference number PCI2020-120703-2-EUROPATMOS.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, H.; Baeyens, J.; Degrève, J.; Cacères, G. Concentrated solar power plants: Review and design methodology. Renew. Sustain. Energy Rev. 2013, 22, 466–481. [Google Scholar] [CrossRef]
  2. Crespi, F.; Toscani, A.; Zani, P.; Sánchez, D.; Manzolini, G. Effect of passing clouds on the dynamic performance of a CSP tower receiver with molten salt heat storage. Appl. Energy 2018, 229, 224–235. [Google Scholar] [CrossRef]
  3. Grosu, Y.; Bondarchuk, O.; Faik, A. The effect of humidity, impurities and initial state on the corrosion of carbon and stainless steels in molten HitecXL salt for CSP application. Sol. Energy Mater. Sol. Cells 2017, 174, 34–41. [Google Scholar] [CrossRef]
  4. de Miguel, M.; Encinas-Sánchez, V.; Lasanta, M.; García-Martín, G.; Pérez, F. Corrosion resistance of HR3C to a carbonate molten salt for energy storage applications in CSP plants. Sol. Energy Mater. Sol. Cells 2016, 157, 966–972. [Google Scholar] [CrossRef]
  5. Grosu, Y.; Udayashankar, N.; Bondarchuk, O.; González-Fernández, L.; Faik, A. Unexpected effect of nanoparticles doping on the corrosivity of molten nitrate salt for thermal energy storage. Sol. Energy Mater. Sol. Cells 2018, 178, 91–97. [Google Scholar] [CrossRef]
  6. Encinas-Sánchez, V.; de Miguel, M.; García-Martín, G.; Lasanta, M.; Pérez, F. Corrosion resistance of Cr/Ni alloy to a molten carbonate salt at various temperatures for the next generation high-temperature CSP plants. Sol. Energy Mater. 2018, 171, 286–292. [Google Scholar] [CrossRef]
  7. Villada, C.; Bonk, A.; Bauer, T.; Bolívar, F. High-temperature stability of nitrate/nitrite molten salt mixtures under different atmospheres. Appl. Energy 2018, 226, 107–115. [Google Scholar] [CrossRef]
  8. García-Martín, G.; Lasanta, M.; Encinas-Sánchez, V.; de Miguel, M.; Pérez, F. Evaluation of corrosion resistance of A516 Steel in a molten nitrate salt mixture using a pilot plant facility for application in CSP plants. Sol. Energy Mater. Sol. Cells 2017, 161, 226–231. [Google Scholar] [CrossRef]
  9. Fernández, A.; Cortes, M.; Fuentealba, E.; Pérez, F. Corrosion properties of a ternary nitrate/nitrite molten salt in concentrated solar technology. Renew. Energy 2015, 80, 177–183. [Google Scholar] [CrossRef]
  10. Slusser, J.W.; Titcomb, J.B.; Heffelfinger, M.T.; Dunbobbin, B.R. Corrosion in Molten Nitrate-Nitrite Salts. J. Met. 1985, 37, 24–27. [Google Scholar] [CrossRef]
  11. Kong, Z.; Jin, Y.; Hossen, G.M.S.; Hong, S.; Wang, Y.; Vu, Q.-V.; Truong, V.-H.; Tao, Q.; Kim, S.-E. Experimental and theoretical study on mechanical properties of mild steel after corrosion. Ocean Eng. 2022, 246, 110652. [Google Scholar] [CrossRef]
  12. de Miguel, M.; Lasanta, M.; García-Martín, G.; Díaz, R.; Pérez, F. Temperature effect and alloying elements impact on the corrosion behaviour of the alloys exposed to molten carbonate environments for CSP application. Corros. Sci. 2022, 201, 110274. [Google Scholar] [CrossRef]
  13. García-Martin, G.; Lasanta, M.I.; de Miguel, M.T.; Sánchez, A.I.; Pérez-Trujillo, F.J. Corrosion Behavior of VM12-SHC Steel in Contact with Solar Salt and Ternary Molten Salt in Accelerated Fluid Conditions. Energies 2021, 14, 5903. [Google Scholar] [CrossRef]
  14. Goods, S.H.; Bradshaw, R.W. Corrosion of Stainless Steels and Carbon Steel by Molten Mixtures of Commercial Nitrate Salts. J. Mater. Eng. Perform. 2004, 13, 78–87. [Google Scholar] [CrossRef]
  15. Turchi, C.; Vidal, J.; Bauer, M. Molten salt power towers operating at 600–650 °C: Salt selection and cost benefits. Sol. Energy 2018, 164, 38–46. [Google Scholar] [CrossRef]
  16. Walczak, M.; Pineda, F.; Fernández, G.; Mata-Torres, C.; Escobar, R.A. Materials corrosion for thermal energy storage systems in concentrated solar power plants. Renew. Sustain. Energy Rev. 2018, 86, 22–44. [Google Scholar] [CrossRef]
  17. Pizzolato, A.; Donato, F.; Verda, V.; Santarelli, M. CFD-based reduced model for the simulation of thermocline thermal energy storage systems. Appl. Therm. Eng. 2015, 76, 391–399. [Google Scholar] [CrossRef] [Green Version]
  18. Dorcheh, A.S.; Durham, R.N.; Galetz, M.C. Corrosion behavior of stainless and low-chromium steels and IN625 in molten nitrate salts at 600 °C. Sol. Energy Mater. Sol. Cells 2016, 144, 109–116. [Google Scholar] [CrossRef]
  19. Agüero, A.; Audigié, P.; Rodríguez, S.; Encinas-Sánchez, V.; de Miguel, M.T.; Pérez, F.J. Protective coatings for high temperature molten salt heat storage systems in solar concentration power plants. AIP Conf. Proc. 2018, 2033, 090001. [Google Scholar] [CrossRef]
  20. Dorcheh, A.S.; Galetz, M. Slurry aluminizing: A solution for molten nitrate salt corrosion in concentrated solar power plants. Sol. Energy Mater. Sol. Cells 2016, 146, 8–15. [Google Scholar] [CrossRef]
  21. Audigié, P.; Encinas-Sánchez, V.; Juez-Lorenzo, M.; Rodríguez, S.; Gutiérrez, M.; Pérez, F.; Agüero, A. High temperature molten salt corrosion behavior of aluminide and nickel-aluminide coatings for heat storage in concentrated solar power plants. Surf. Coat. Technol. 2018, 349, 1148–1157. [Google Scholar] [CrossRef]
  22. Fähsing, D.; Oskay, C.; Meißner, T.; Galetz, M. Corrosion testing of diffusion-coated steel in molten salt for concentrated solar power tower systems. Surf. Coat. Technol. 2018, 354, 46–55. [Google Scholar] [CrossRef]
  23. Wang, D.; Bierwagen, G.P. Sol–gel coatings on metals for corrosion protection. Prog. Org. Coat. 2009, 64, 327–338. [Google Scholar] [CrossRef]
  24. Berlin, I.J.; Lekshmy, S.S.; Ganesan, V.; Thomas, P.; Joy, K. Effect of Mn doping on the structural and optical properties of ZrO2 thin films prepared by sol–gel method. Thin Solid Films 2014, 550, 199–205. [Google Scholar] [CrossRef]
  25. Díaz-Parralejo, A.; Ortiz, A.; Caruso, R. Effect of sintering temperature on the microstructure and mechanical properties of ZrO2-3mol%Y2O3 sol–gel films. Ceram. Int. 2010, 36, 2281–2286. [Google Scholar] [CrossRef]
  26. Encinas-Sánchez, V.; Macías-García, A.; Pérez, F. Effect of withdrawal rate on the evolution of optical properties of dip-coated yttria-doped zirconia thin films. Ceram. Int. 2017, 43, 13094–13100. [Google Scholar] [CrossRef]
  27. Kirubaharan, A.K.; Kuppusami, P.; Priya, R.; Divakar, R.; Gupta, M.; Pandit, D.; Ningshen, S. Synthesis, microstructure and corrosion behavior of compositionally graded Ni-YSZ diffusion barrier coatings on inconel-690 for applications in high temperature environments. Corros. Sci. 2018, 135, 243–254. [Google Scholar] [CrossRef]
  28. Encinas-Sánchez, V.; Batuecas, E.; Garcia, A.M.; Mayo, C.; Díaz, R.; Perez-Trujillo, F.J. Corrosion resistance of protective coatings against molten nitrate salts for thermal energy storage and their environmental impact in CSP technology. Sol. Energy 2018, 176, 688–697. [Google Scholar] [CrossRef]
  29. Loghman-Estarki, M.; Razavi, R.S.; Jamali, H. Effect of molten V2O5 salt on the corrosion behavior of micro- and nano-structured thermal sprayed SYSZ and YSZ coatings. Ceram. Int. 2016, 42, 12825–12837. [Google Scholar] [CrossRef]
  30. Hajizadeh-Oghaz, M.; Razavi, R.S.; Ghasemi, A.; Valefi, Z. Na2SO4 and V2O5 molten salts corrosion resistance of plasma-sprayed nanostructured ceria and yttria co-stabilized zirconia thermal barrier coatings. Ceram. Int. 2016, 42, 5433–5446. [Google Scholar] [CrossRef]
  31. Liu, M.; Bell, S.; Segarra, M.; Tay, N.S.; Will, G.; Saman, W.; Bruno, F. A eutectic salt high temperature phase change material: Thermal stability and corrosion of SS316 with respect to thermal cycling. Sol. Energy Mater. Sol. Cells 2017, 170, 1–7. [Google Scholar] [CrossRef]
  32. Sarvghad, M.; Will, G.; Steinberg, T.A. Corrosion of steel alloys in molten NaCl+Na2SO4 at 700 °C for thermal energy storage. Sol. Energy Mater. Sol. Cells 2018, 179, 207–216. [Google Scholar] [CrossRef]
  33. Encinas-Sánchez, V.; de Miguel, M.T.; Lasanta, M.I.; García-Martín, G.; Pérez, F.J. Electrochemical impedance spectroscopy (EIS): An efficient technique for monitoring corrosion processes in molten salt environments in CSP applications. Sol. Energy Mater. Sol. Cells 2019, 191, 157–163. [Google Scholar] [CrossRef]
  34. Jagadeeswara-Rao, C.; Venkatesh, P.; Ningshen, S. Corrosion assessment of 9Cr-1Mo steel in molten LiCl-KCl eutectic salt by electrochemical methods. J. Nucl. Mater. 2019, 514, 114–122. [Google Scholar]
  35. Han, Y.; Wang, J.; Zhang, H.; Zhao, S.; Ma, Q.; Wang, Z. Electrochemical impedance spectroscopy (EIS): An efficiency method to monitor resin curing processes. Sens. Actuators A Phys. 2016, 250, 78–86. [Google Scholar] [CrossRef]
  36. Macdonald, J.R. Impedance spectroscopy: Old problems and new developments. Electrochim. Acta 1990, 35, 1483–1492. [Google Scholar] [CrossRef]
  37. Macdonald, D.D. Reflections on the history of electrochemical impedance spectroscopy. Electrochim. Acta 2006, 51, 1376–1388. [Google Scholar] [CrossRef]
  38. Montemor, M.; Simões, A.; Ferreira, M. Chloride-induced corrosion on reinforcing steel: From the fundamentals to the monitoring techniques. Cem. Concr. Compos. 2003, 25, 491–502. [Google Scholar] [CrossRef] [Green Version]
  39. Zhu, M.; Zeng, S.; Zhang, H.; Li, J.; Cao, B. Electrochemical study on the corrosion behaviors of 316 SS in HITEC molten salt at different temperatures. Sol. Energy Mater. Sol. Cells 2018, 186, 200–207. [Google Scholar] [CrossRef]
  40. Encinas-Sánchez, V.; Lasanta, M.; de Miguel, M.; García-Martín, G.; Pérez, F. Corrosion monitoring of 321H in contact with a quaternary molten salt for parabolic trough CSP plants. Corros. Sci. 2021, 178, 109070. [Google Scholar] [CrossRef]
  41. Salinas-Solano, G.; Porcayo-Calderon, J.; Gonzalez-Rodriguez, J.; Salinas-Bravo, V.; Ascencio-Gutierrez, J.; Martinez-Gomez, L. High temperature corrosion of Inconel 600 in NaCl-KCl molten salts. Adv. Mater. Sci. Eng. 2014, 2014, 8. [Google Scholar] [CrossRef]
  42. Trinstancho-Reyes, J.; Sandoval-Jabalera, R.; Orozco-Carmona, V.; Almeraya-Calderón, F.; Chacón-Nava, J.; Gonzalez-Rodriguez, J.; Martínez-Villafañe, A. Electrochemical impedance spectroscopy investigation of alloy Inconel 718 in molten salts at high temperature. Int. J. Electrochem. Sci. 2011, 6, 419–431. [Google Scholar]
  43. Encinas-Sánchez, V.; Macías-García, A.; Díaz-Díez, M.; Brito, P.; Cardoso, D. Influence of the quality and uniformity of ceramic coatings on corrosion resistance. Ceram. Int. 2015, 41, 5138–5146. [Google Scholar] [CrossRef]
  44. Croll, S.G. The origin of residual internal stress in solvent-cast thermoplastic coatings. J. Appl. Polym. Sci. 1979, 23, 847–858. [Google Scholar] [CrossRef] [Green Version]
  45. Hamden, Z.; Boufi, S.; Conceição, D.; Ferraria, A.; Rego, A.B.D.; Ferreira, D.; Ferreira, L.V.; Bouattour, S. Li–N doped and codoped TiO2 thin films deposited by dip-coating: Characterization and photocatalytic activity under halogen lamp. Appl. Surf. Sci. 2014, 314, 910–918. [Google Scholar] [CrossRef]
  46. Fernández, A.; Lasanta, M.; Pérez, F. Molten salt corrosion of stainless steels and low-Cr steel in CSP plants. Oxid. Met. 2012, 78, 329–348. [Google Scholar] [CrossRef]
  47. Fuqiang, W.; Ziming, C.; Jianyu, T.; Yuan, Y.; Yong, S.; Linhua, L. Progress in concentrated solar power technology with parabolic trough collector system: A comprehensive review. Renew. Sustain. Energy Rev. 2017, 79, 1314–1328. [Google Scholar] [CrossRef]
  48. Encinas-Sánchez, V.; Macías-García, A.; Díaz-Díez, M.; Díaz-Parralejo, A. Characterization of sol–gel coatings deposited on a mechanically treated stainless steel by using a simple non-destructive electrical method. J. Ceram. Soc. Jpn. 2016, 124, 185–191. [Google Scholar] [CrossRef] [Green Version]
  49. Morrow, J.; Tejedor-Anderson, M.I.; Anderson, M.A.; Ruotolo, L.A.; Pfefferkorn, F.E. Sol-gel synthesis of ZrO2 coatings on micro end mills with sol-gel processing. J. Micro Nano-Manuf. 2014, 2, 041002–041012. [Google Scholar] [CrossRef]
  50. Ni, C.; Lu, L.; Zeng, C.; Niu, Y. Evaluation of corrosion resistance of aluminium coating with and without annealing against molten carbonate using electrochemical impedance spectroscopy. J. Power Sources 2014, 261, 162–169. [Google Scholar] [CrossRef]
  51. Fernandez, A.; Rey, A.; Lasanta, I.; Mato, S.; Brady, M.; Pérez, F.J. Corrosion of alumina-forming austenitic steel in molten nitrate salts by gravimetric analysis and impedance spectroscopy. Mater. Corros. 2014, 65, 267–275. [Google Scholar] [CrossRef]
  52. Omar, S.; Repp, F.; Desimone, P.M.; Weinkamer, R.; Wagermaier, W.; Ceré, S.; Ballarre, J. Sol–gel hybrid coatings with strontium-doped 45S5 glass particles for enhancing the performance of stainless steel implants: Electrochemical, bioactive and in vivo response. J. Non-Cryst. Solids 2015, 425, 1–10. [Google Scholar] [CrossRef]
  53. ASTM G102; Standard Practice for Calculation of Corrosion Rates and Related Information from Electrochemical Measurements. ASTM: West Conshohocken, PA, USA, 1994.
  54. Stern, M.; Geary, A.L. Electrochemical Polarization: I. A Theoretical Analysis of the Shape of Polarization Curves. J. Electrochem. Soc. 1957, 104, 56–63. [Google Scholar] [CrossRef]
  55. Ruiz-Cabañas, F.J.; Prieto, C.; Osuna, R.; Madina, V.; Fernández, A.I.; Cabeza, L.F. Corrosion testing device for in-situ corrosion characterization in operational molten salts storage tanks: A516 Gr70 carbon steel performance under molten salts exposure. Sol. Energy Mater. Sol. Cells 2016, 157, 383–392. [Google Scholar] [CrossRef]
Figure 1. Sketch of an electrode used in the electrochemical sensor.
Figure 1. Sketch of an electrode used in the electrochemical sensor.
Materials 16 00546 g001
Figure 2. Surface appearance macrographs of (a) uncoated P91; (b) coated P91.
Figure 2. Surface appearance macrographs of (a) uncoated P91; (b) coated P91.
Materials 16 00546 g002
Figure 3. Impedance spectra of the coated P91 samples tested at 500 °C in Solar Salt at (a) 24 h; (b) 72 h; (c) 500 h; (d) 1000 h; (e) 1500 h; and (f) 2000 h.
Figure 3. Impedance spectra of the coated P91 samples tested at 500 °C in Solar Salt at (a) 24 h; (b) 72 h; (c) 500 h; (d) 1000 h; (e) 1500 h; and (f) 2000 h.
Materials 16 00546 g003
Figure 4. Equivalent circuit to which experimental data are fitted.
Figure 4. Equivalent circuit to which experimental data are fitted.
Materials 16 00546 g004
Figure 5. Adjustment of experimental data obtained after 2000 h.
Figure 5. Adjustment of experimental data obtained after 2000 h.
Materials 16 00546 g005
Figure 6. Variation in the resistance of the electrolyte (Re) during testing.
Figure 6. Variation in the resistance of the electrolyte (Re) during testing.
Materials 16 00546 g006
Figure 7. Variation in charge transfer resistance (Rt), protective layer resistance (Rcp) and protective layer capacitance (Ccp) during testing.
Figure 7. Variation in charge transfer resistance (Rt), protective layer resistance (Rcp) and protective layer capacitance (Ccp) during testing.
Materials 16 00546 g007
Figure 8. Variation in ndl and ncp exponents during testing.
Figure 8. Variation in ndl and ncp exponents during testing.
Materials 16 00546 g008
Figure 9. Corrosion rates estimated by EIS in coated P91.
Figure 9. Corrosion rates estimated by EIS in coated P91.
Materials 16 00546 g009
Table 1. Impurities composition in the chemicals used.
Table 1. Impurities composition in the chemicals used.
ChemicalsClSO42−CO32−
NaNO30.020.0050.02
KNO30.015<0.0005<0.02
Table 2. Values of each of the equivalent circuit elements in compliance for each testing time with the protective layer model.
Table 2. Values of each of the equivalent circuit elements in compliance for each testing time with the protective layer model.
Time, hRe, ΩRt, ΩCdl, Ω−1·snndlRcp, ΩCcp, Ω−1·snncp
247.55713.871.506·10−30.777144881.483·10−30.731
727.09529.091.973·10−30.74825261.884·10−30.711
5006.17133.991.632·10−30.80416573.060·10−30.738
10005.06511.813.014·10−30.7945955.982·10−30.669
15006.3276.644.350·10−30.8082611.021·10−20.615
20005.0726.654.440·10−30.8091881.195·10−20.610
Table 3. Guide for corrosion rates used in the industry [55].
Table 3. Guide for corrosion rates used in the industry [55].
Corrosion Rate, mm·yr−1Recommendation
>1275Completely destroyed within days
127–1274Not recommended for service greater than 1 month
64–126Not recommended for service greater than 1 year
14–63Caution recommended, based on the specific application
0.4–13Recommended for long-term service
<0.3Recommended for long-term service; no corrosion, other than as a result of surface cleaning, was evidenced
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

Encinas-Sánchez, V.; Macías-García, A.; de Miguel, M.T.; Pérez, F.J.; Rodríguez-Rego, J.M. Electrochemical Impedance Analysis for Corrosion Rate Monitoring of Sol–Gel Protective Coatings in Contact with Nitrate Molten Salts for CSP Applications. Materials 2023, 16, 546. https://doi.org/10.3390/ma16020546

AMA Style

Encinas-Sánchez V, Macías-García A, de Miguel MT, Pérez FJ, Rodríguez-Rego JM. Electrochemical Impedance Analysis for Corrosion Rate Monitoring of Sol–Gel Protective Coatings in Contact with Nitrate Molten Salts for CSP Applications. Materials. 2023; 16(2):546. https://doi.org/10.3390/ma16020546

Chicago/Turabian Style

Encinas-Sánchez, V., A. Macías-García, M. T. de Miguel, F. J. Pérez, and J. M. Rodríguez-Rego. 2023. "Electrochemical Impedance Analysis for Corrosion Rate Monitoring of Sol–Gel Protective Coatings in Contact with Nitrate Molten Salts for CSP Applications" Materials 16, no. 2: 546. https://doi.org/10.3390/ma16020546

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