Next Article in Journal
Superlubricity of Materials: Progress, Potential, and Challenges
Previous Article in Journal
Silicon Nitride Ceramics: Structure, Synthesis, Properties, and Biomedical Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Corrosion Properties of Bronze Alloys in NaCl Solutions

1
Institute for the History of Science and Technology, Inner Mongolia Normal University, 81 Zhaowuda Road, Hohhot 010022, China
2
College of Physics and Electronic Information, Inner Mongolia Normal University, 81 Zhaowuda Road, Hohhot 010022, China
*
Author to whom correspondence should be addressed.
Materials 2023, 16(14), 5144; https://doi.org/10.3390/ma16145144
Submission received: 29 June 2023 / Revised: 18 July 2023 / Accepted: 19 July 2023 / Published: 21 July 2023

Abstract

:
Chloride ions play an important role in the corrosion of bronze through their active reactivity to copper alloys. The corrosion behavior of bronze alloys in NaCl solution was investigated by using X-ray diffraction (XRD), a scanning electron microscope (SEM), and electrochemical tests, with a special emphasis on the corrosion resistance of the α and δ phases in Cu-20 wt%Sn bronze alloys. The experimental results show that the corrosion current density of Cu-20 wt%Sn bronze alloys increases from 1.1 × 10−7 A/cm2 to 2.7 × 10−6 A/cm2 with the increase in the chloride ion concentration from 10−3 mol/L to 1 mol/L. After a soaking duration of 30 days, the matrix corrosion depth reaches 50 μm. The α phase of the alloys is easily corroded in NaCl solution, while the δ phase with high Sn content has strong corrosion resistance. This study provides relevant data for the analysis and protection of ancient bronze alloys.

1. Introduction

Bronze is mainly composed of elements such as copper and tin and exhibits excellent mechanical properties [1]. Although it is one of the earliest alloys used by humans, it still has widespread application in modern industry [2]. Prior to excavation, bronze artifacts unavoidably react with salts present in water or soil, causing oxidation–reduction reactions, hydrolysis reactions, and precipitation reactions between the bronze matrix and its corrosion products. Over a long period under earth, both the matrix and the surface of ancient bronzes undergo different degrees of corrosion [3].
Archaeologists have conducted extensive research on the corrosion and protection of bronzes [4,5,6,7,8,9,10,11]. There are some models that could explain the corrosion mechanism of bronze alloys. Some studies suggest that corrosion occurs preferentially in grains with impurities or inclusions [12], while others propose grain boundary corrosion as a prominent mode of degradation in bronze alloys [13]. Additionally, pitting corrosion is recognized as a common form of bronze corrosion [14,15]. When cooled from the liquid phase to the solid phase, Sn bronze alloys tend to segregate, forming Cu-rich α solid solution dendrites and a Sn-rich δ phase (Cu41Sn11). Despite the inherent corrosion resistance of bronze, selective corrosion of Cu-rich or Sn-rich phases results from alloy segregation [16]. Some studies indicate that the α phase in the alloy is preferentially corroded [17], while others report preferential corrosion of the δ phase [18]. Moreover, the corrosion state of bronze is closely related to the environmental conditions, such as humidity and various anions [19].
Chloride ions play a crucial role in the formation and transformation of bronze patina by increasing the solubility of the patina layer and enhancing the conductivity of the matrix [20]. Chloride ions readily penetrate defects and grain boundaries in the alloy matrix, leading to the formation of cuprous oxide and chlorides [21]. The initial formation of cuprous oxide occurs on the surface of the matrix, gradually passivating the bronze surface, while the presence of chloride ions promotes the transformation of the passivated layer into loosely packed copper chlorides [22]. The chloride ions can cause preferential corrosion of the Cu-rich phase in bronze alloys in high-oxygen seawater conditions, while the α + δ eutectic phase is susceptible to corrosion under low-oxygen conditions. Currently, there is extensive research on the corrosion behavior of copper alloys in seawater environments [23]. The occurrence and development of corrosion in bronze alloys are closely related to the presence of moisture and chloride ions, constituting a self-catalytic corrosion process [24]. When chloride ions are present, they react with the bronze matrix, forming a CuCl layer adjacent to the alloy matrix [25], and the surface morphology and elemental composition of corrosion products largely depend on the chemical composition of the alloy [26].
The above work mainly focuses on studying the corrosion laws of tin bronze alloys in different environments, most of which are about the effect of chloride ions on the corrosion of copper alloys. It has been proven that in chloride aqueous solutions, the passivation layer on copper alloys consists of copper oxides and copper chlorides [27]. However, there is relatively limited research on the influence of chloride ion concentration on the corrosion of copper-tin alloys, although higher chloride ion concentrations favor the occurrence of pitting corrosion [28]. Investigating the corrosion rate of bronze alloys in water environments with different chloride ion concentrations and the corrosion resistance of different phases within the alloy is needed for understanding the corrosion and protection of Cu-Sn alloys and bronzes.
This article reports the corrosion behavior of synthetic bronze alloys in NaCl solutions (1 mol/L, 0.1 mol/L, 0.01 mol/L, and 0.001 mol/L) [29] of different concentrations investigated by using XRD, SEM, the polarization curve method [30], and electrochemical impedance spectroscopy [31]. This article selects Cu-20 wt%Sn bronze for corrosion resistance research. The Cu-rich α phase and Sn-rich δ phase in Cu-20 wt%Sn bronze account for almost half each, making it more intuitive to compare the corrosion situation of the α phase and δ phase.

2. Materials and Methods

2.1. Materials

Copper and tin with 99.9% purity were used as raw materials. First, 100 g of copper was heated to 1200 °C in an induction furnace, and then 20 wt%Sn was added, at which point the temperature was kept constant for 5 min. In order to imitate the preparation process of ancient bronzes, the sample is prepared in air and naturally cooled. Finally, the ingot sample was cut into a 10 mm × 10 mm × 10 mm cube. Other materials used in the experiment included NaCl (from Fu Chen Chemical Reagents Co., Ltd., Tianjin, China), a AgCl electrode (from Shanghai Xianren Chemical Reagents Co., Ltd., Shanghai, China), a Pt electrode (self-made), and deionized water (self-made).

2.2. Methods

The measurements were conducted using a CHI660E electrochemical workstation (Shanghai Chenhua, Shanghai, China). Electrochemical impedance spectroscopy and polarization curve analysis were employed to investigate the effects of different immersion times and chloride ion concentrations on the corrosion of bronze alloys. Firstly, under stable open-circuit potential, electrochemical impedance spectroscopy was performed over a frequency range from 100 kHz to 0.1 Hz with a perturbation signal amplitude of 10 mV. Secondly, polarization curve testing was conducted at a scan rate of 20 mV/min within a scanning range of ±400 mV relative to the open-circuit potential. The sample’s structure was analyzed using XRD (PANalytical Empyrean, Almelo, The Netherlands). The metallic structures were observed and photographed under a metallographic microscope (ZEIZZ Observer A1m, Oberkochen, Germany). Scanning electron microscope (Hitachi TM3030, Tokyo, Japan) was used to observe the sample’s morphology, and EDS (Bruker Q70, Billerica, MA, USA) was used to analyze the elemental composition.

3. Results and Discussion

3.1. Structure and Composition of the Bronze Alloys

Figure 1 presents the XRD patterns of Cu-20 wt%Sn bronze alloys without soaking and when soaked for 30 days in 0.1 mol/L NaCl solution. The analysis of the XRD patterns showed that the structure of the sample prior to soaking exhibited the coexistence of an α phase with a face-centered cubic structure and a δ phase with a complex cubic structure. In contrast to the sample soaked in NaCl solution, there were some diffraction peaks corresponding to the CuCl and Cu2O phases in addition to the α and δ phases.
Figure 2a shows the metallographic structure of the Cu-20 wt%Sn bronze. The morphology of the alloy indicates the presence of an α phase and an (α + δ) eutectic microstructure. It also can be seen that the sample contains defects and shrinkage voids.
Figure 2b displays the backscattered electrons (BSE) SEM image of the Cu-20 wt%Sn bronze. The BSE image contains information about the sample’s composition in the form of image contrast. The difference in contrast between the α and δ phases is clearly visible. The dark gray regions correspond to the α phase, while the light gray regions represent the δ phase.
Table 1 lists the compositions of the α and δ phases in the Cu-20 wt%Sn bronze. The main chemical composition of the sample consists of Cu and Sn elements, and the mass ratio of the elements in the sample is generally consistent with the alloy’s nominal composition. The α solid solution phase contains 13.50% Sn. The δ phase contains 23.82% Sn, which is slightly lower than the theoretical value of 32.6% in Cu41Sn11. The reason is that the actual measurement area for characteristic X-rays is larger than the selected area; therefore, it includes some of the α phase components.

3.2. Microscopic Corrosion Morphology and Products

To directly compare the corrosion resistance of the α and δ phases, we selected a Cu-20 wt%Sn bronze with a near-equal proportion of both phases for in-depth analysis. Figure 3 shows the SEM images of the sample surface soaked in NaCl solution for 30 days. Table 2 presents the EDS compositional analysis of the corroded sample. According to the results, the Cu-rich α-phase contains higher oxygen content than the Sn-rich δ-phase, and Cu is more depleted in the α-phase.
Figure 4 and Figure 5 respectively show the cross-sectional SEM image and the corresponding elemental distribution map of the Cu-20 wt%Sn bronze soaked in a 0.1 mol/L NaCl solution for 30 days. The cross-sectional SEM image reveals that the Cu-rich phase near the matrix surface is more susceptible to corrosion than the Sn-rich phase and that corrosion occurs along the grain boundaries. The element distribution map clearly shows severe copper depletion and an increase in oxygen content in the Cu-rich phase near the matrix surface. The surface (Figure 3) composition of the alloy soaked in NaCl solution for 30 days indicated higher oxygen content in the Cu-rich phase than in the Sn-rich phase, which is consistent with the elemental distribution along the cross-section (Figure 5). The corrosion depth near the matrix surface is approximately 50 μm.
Sn is more active in NaCl solution than copper, so Sn is initially oxidized to form a protective Sn-rich passivation film on the surface, preventing further corrosion. Only when the passivation film is continuously eroded or breached does the copper start to corrode. Due to the Sn-rich δ-phase in the sample, the passivation film has strong protective properties and high corrosion resistance. According to Wang’s study [32], the corrosion product of tin is mainly SnO2, and the corrosion behavior of the alloy is influenced by the tin content. Alloys with higher tin content exhibit a lower corrosion current density and stronger corrosion resistance [33].
The elemental analysis results show that copper and chloride ions increase near the surface of the alloy, while in the alloy matrix, the Cu content of the α phase decreases. The Cu lost from the matrix and Cl ions in the solution form CuCl products on the surface of the sample. The XRD results (Figure 1) show the presence of Cu2O and CuCl phases in the post-immersion sample.
CuCl is an unstable intermediate product because it releases chloride ions during hydrolysis and oxidation processes [34], leading to the continuous dissolution of the bronze alloy matrix and the formation of Cu2O on the surface [35,36,37]. The Cu-rich α-phase in the matrix is more reactive with chloride ions to form CuCl. The combination of O2 generated after corrosion with the crevices and the Cu-rich phases in the matrix further contributes to the formation of metal oxides, resulting in accelerated corrosion of the Cu-rich phases in the alloy. Therefore, the copper loss is faster in the α-phase, while the δ-phase exhibits stronger corrosion resistance.
Figure 6 shows the SEM images and corresponding EDS spectra of Cu-20 wt%Sn during the initial corrosion stage (1 h) in different concentrations of NaCl solution. The composition of the samples was analyzed using EDS. The results indicate that the samples exhibited lower levels of oxygen (O) and chlorine (Cl) at the beginning of immersion in NaCl solutions of varying concentrations. As the solution concentration increased from 0.001 mol/L to 0.01 mol/L, the oxygen content in the samples increased. However, when the solution concentration increased from 0.01 mol/L to 1 mol/L, the oxygen content in the samples decreased. The chlorine content in the samples, on the other hand, increased with the increase in the solution concentration.

3.3. Electrochemical Corrosion Characteristics

Open circuit potential (OCP) is an important parameter for evaluating the corrosion behavior of alloys and is used to analyze the corrosion mechanism of bronzes. Figure 7 shows the OCP curves of the Cu-20 wt%Sn bronze in NaCl solutions of different concentrations. It can be seen that the concentration in NaCl solution has an impact on the corrosion potential of the bronze alloys once the self-corrosion potential reaches a stable state. With the increase in concentration, the self-corrosion potential of the sample increases and the OCP changes from −0.225 V to −0.02 V.
Figure 8 illustrates Tafel curves for the sample in NaCl solutions of different concentrations. The polarization curve was fitted using CHI660E software (CHI Version 14.05); the results are shown in Table 3. It can be seen that the corrosion current density of the sample increases with the increasing concentration of NaCl. The fitting results indicate that the corrosion current density of the bronze increases from 1.1 × 10−7 A/cm2 to 2.7 × 10−6 A/cm2 with a change in concentration from 10−3 mol/L to 1 mol/L. Through Stern’s formula Rp = Ba × Bc/[2.3 (Ba + Bc) × Iccor] (Ba and Bc are the Tafel slopes of the anode and cathode, respectively, and Rp is the polarization resistance). To calculate the polarization resistance Rp of the bronze alloys at different concentrations, Rp (10−3 mol/L) > Rp (10−2 mol/L) > Rp (10−1 mol/L) > Rp (1 mol/L). The higher the polarization resistance, the stronger the corrosion resistance of the alloy, indicating that bronze alloys have strong corrosion resistance in low-concentration NaCl solutions.
Figure 9a illustrates the Nyquist plot of the electrochemical impedance spectroscopy for the sample in NaCl solutions of different concentrations. In the Nyquist plot, the horizontal and vertical coordinates correspond to the real part (Z′) and imaginary part (Z″) of the complex transfer function. The capacitance arc of the alloy decreases with the increase in concentration. The size of the capacitance arc can be used to determine the charge transfer resistance and, consequently, the corrosion rate. Figure 9b shows the Bode plot, the horizontal coordinate represents the logarithmic frequency scale (Hz), and the vertical coordinate represents the phase shift of the transfer function of the system. From the Bode plot, it can be seen that the electrochemical impedance spectroscopy in 10−3, 10−2, 10−1, and 1 mol/L NaCl solutions exhibit only one capacitance arc, indicating the presence of a single time constant. This suggests that an oxide film does not form on the surface of the alloy during the initial 1 h immersion in NaCl solutions of different concentrations. The phase angle values for the different concentrations are 54.3, 53.5, 58.28, and 70.73 degrees. When the solution concentration is 10−3 and 10−2 mol/L, the change is not significant, and the concentration further increases, and the phase angle also increases.
The equivalent circuit diagram of the bronze in NaCl solution is shown in Figure 11a. In this circuit, Rs represents the resistance of the NaCl solution, Rct corresponds to the charge transfer resistance, and Qdl is a constant phase element (CPE), which is the double-layer capacitance between the NaCl solution and the sample, as well as the non-ideal behavior of the double-layer capacitance. Table 4 provides the fitting results of the electrochemical components of the equivalent circuit for the Cu-20 wt%Sn bronze in NaCl solutions. The fitting results indicate that Rs decreases gradually with the increasing concentration. This leads to a decrease in resistance because the conductivity of the solution increases with the higher NaCl concentration. Rct is an important indicator of corrosion rate. Higher charge transfer resistance indicates better corrosion resistance of the alloy. The fitting results show that within the range of 10−3 to 1 mol/L NaCl concentration, Rct decreases with the increasing chloride ion concentration. The accelerated dissolution of Cu elements in the alloy with increasing solution concentration was confirmed by subsequent immersion experiments. This suggests that the increase in chloride ion concentration contributes to the corrosion rate of the bronze substrate.
Figure 10a shows the Nyquist plot; it demonstrates another set of parallel experiments where the Cu-20 wt%Sn bronze is soaked in a 0.1 mol/L NaCl solution for different immersion periods: 0 days, 3 days, and 30 days. Figure 10b shows the Bode plot of the sample; the phase angles at different times are 58.28, 52.58, and 22.37 degrees. As the soaking time increases, the phase angle becomes smaller; it shows only one time constant in the high-frequency region, indicating activation of the metal substrate and uniform corrosion on the electrode alloy surface without the formation of an oxide film. With an increase in the soaking time span to 3 days, the Bode plot of the alloy still exhibits only one time constant. However, when the soaking time is extended to 30 days, a double time constant appears in the Bode plot. The time constant in the high-frequency region corresponds to the response of the passive film interface, while the time constant in the medium–low-frequency region corresponds to the corrosion process of the metallic matrix. This suggests the presence of an oxide film on the electrode’s surface, which partially inhibits the corrosion process of the matrix. Figure 10c shows the Bode plot (magnitude) of the sample; as the soaking time increases, Z gradually decreases.
From the EIS tests and fitting results, it can be seen that the charge transfer resistance (Rct) decreases gradually with the increase in the chloride ion concentration in the NaCl solution. In chloride ion solutions, the Sn-rich δ-phase in the alloy demonstrates better corrosion resistance than the α-phase, indicating that the corrosion of the α-phase is the main factor affecting impedance changes. The increase in the solution concentration accelerates the dissolution of Cu elements. The results from the OCP, Tafel, and EIS tests are consistent and confirm that the increase in the chloride ion concentration contributes to the corrosion rate of the bronze matrix.
The impedance analysis from the soaking tests (Figure 10) reveals that the bronze initially undergoes uniform corrosion during immersion, and after 30 days, a protective film forms on the alloy surface. Based on the analysis of the experimental results mentioned above, the corrosion behavior and equivalent circuit diagram of the bronze in NaCl solution at different time intervals can be depicted, as shown in Figure 11. The schematic of the corrosion mechanism of the Cu-20 wt%Sn bronze after immersion in 0.1 mol/L NaCl solution is shown in Figure 12.

4. Conclusions

The present work aimed to describe the complex corrosion chemistry of a binary Cu-Sn alloy, understand the corrosion process of Cu-20 wt%Sn bronze in NaCl solution, and elaborate on the corrosion resistance of different phases in the alloy. From the experimental results, the main conclusions are as follows. Within the NaCl concentration range of 10−3 to 1 mol/L, the corrosion current density of the bronze increases from 1.1 × 10−7 A/cm2 to 2.7 × 10−6 A/cm2 as the chloride ion concentration increases, indicating that chloride ions accelerate corrosion of the bronze. In the immersion experiment with a 0.1% NaCl solution, after 30 days of immersion, significant copper loss occurs in the α-phase in the alloy, resulting in the formation of a film composed of CuCl and Cu2O on the surface, with a maximum depth of 50 µm. The Sn-rich δ-phase in the bronze exhibits stronger corrosion resistance in chloride ion solutions.

Author Contributions

Conceptualization, O.T. and Z.S.; methodology, Z.S.; formal analysis, Z.S.; resources, Z.S.; data curation, Z.S.; writing—original draft preparation, Z.S.; writing—review and editing, O.T.; visualization, Z.S.; supervision, O.T.; funding acquisition, O.T. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the Natural Science Foundation of Inner Mongolia Autonomous Region (2022QN05002).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon request.

Acknowledgments

The authors appreciate the help of Wei Wei and Zhang Linrui for his advice on the data analysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kohler, F.; Campanella, T.; Nakanishi, S.; Rappaz, M. Application of single pan thermal a nalysis to Cu–Sn peritectic alloys. Acta Mater. 2008, 56, 1519–1528. [Google Scholar] [CrossRef] [Green Version]
  2. Sürme, Y.; Gürten, A.A.; Bayol, E.; Ersoy, E. Systematic corrosion investigation of various Cu–Sn alloys electrodeposited on mild steel in acidic solution: Dependence of alloy composition. J. Alloys Compd. 2009, 485, 98–103. [Google Scholar] [CrossRef]
  3. Wang, Z.; Li, Y.; Jiang, X.; Pan, C. Research progress on ancient bronze corrosion in different environments and using different conservation techniques: A review. MRS Adv. 2017, 2, 2033–2041. [Google Scholar] [CrossRef]
  4. Liu, L.; Zhong, Q.; Jiang, L.; Li, P.; Xiao, L.; Gong, Y.; Zhu, Z.; Yang, J. Metallurgical and corrosion characterization of warring states period bronzes excavated from Pujiang, Chengdu, China. Heritage Sci. 2022, 10, 36. [Google Scholar] [CrossRef]
  5. Privitera, A.; Corbascio, A.; Calcani, G.; Della Ventura, G.; Ricci, M.A.; Sodo, A. Raman approach to the forensic study of bronze patinas. J. Archaeol. Sci. Rep. 2021, 39, 103115. [Google Scholar] [CrossRef]
  6. Li, H.; Zuo, Z.; Cui, J.; Tian, J.; Yang, Y.; Yi, L.; Zhou, Z.; Fan, J. Copper alloy production in the Warring States period (475-221 BCE) of the Shu state: A metallurgical study on copper alloy objects of the Baishoulu cemetery in Chengdu, China. Heritage Sci. 2020, 8, 67. [Google Scholar] [CrossRef]
  7. Mu, D.; Luo, W.; Song, G.; Qiao, B.; Wang, F. The features as a county of Chu State: Chemical and metallurgical characteristics of the bronze artifacts from the Bayilu site. Archaeol. Anthropol. Sci. 2019, 11, 1123–1129. [Google Scholar] [CrossRef]
  8. Li, B.; Jiang, X.; Wu, R.; Wei, B.; Hu, T.; Pan, C. Formation of black patina on an ancient Chinese bronze sword of the Warring States Period. Appl. Surf. Sci. 2018, 455, 724–728. [Google Scholar] [CrossRef]
  9. Armetta, F.; Saladino, M.L.; Scherillo, A.; Caponetti, E. Microstructure and phase composition of bronze Montefortino helmets discovered Mediterranean seabed to explain an unusual corrosion. Sci. Rep. 2021, 11, 23022. [Google Scholar] [CrossRef]
  10. Scott, D.A. Copper and Bronze in Art: Corrosion, Colorants, Conservation; Getty Publications: Los Angeles, CA, USA, 2002; pp. 125–128. Available online: https://www.getty.edu/publications/virtuallibrary/temp/9780892366385.pdf (accessed on 25 May 2023).
  11. Buccolieri, G.; Buccolieri, A.; Donati, P.; Marabelli, M.; Castellano, A. Portable EDXRF investigation of the patinas on the Riace Bronzes. Nucl. Instrum. Methods Phys. Res. Sect. B 2015, 343, 101–109. [Google Scholar] [CrossRef]
  12. Petitmangin, A.; Guillot, I.; Chabas, A.; Nowak, S.; Saheb, M.; Alfaro, S.C.; Blanc, C.; Fourdrin, C.; Ausset, P. The complex atmospheric corrosion of α/δ bronze bells in a marine environment. J. Cult. Herit. 2021, 52, 153–163. [Google Scholar] [CrossRef]
  13. Piccardo, P.; Mödlinger, M.; Ghiara, G.; Campodonico, S.; Bongiorno, V. Investigation on a “tentacle-like” corrosion feature on Bronze Age tin-bronze objects. Appl. Phys. A 2013, 113, 1039–1047. [Google Scholar] [CrossRef]
  14. Osório, W.R.; Freire, C.M.; Caram, R.; Garcia, A. The role of Cu-based intermetallics on the pitting corrosion behavior of Sn–Cu, Ti–Cu and Al–Cu alloys—ScienceDirect. Electrochim. Acta 2012, 77, 189–197. [Google Scholar] [CrossRef]
  15. Saud, S.; Hamzah, E.; Abubakar, T.; Bakhsheshi-Rad, H. Correlation of microstructural and corrosion characteristics of quaternary shape memory alloys Cu–Al–Ni–X (X = Mn or Ti). Trans. Nonferrous Met. Soc. China 2015, 25, 1158–1170. [Google Scholar] [CrossRef]
  16. Robbiola, L.; Blengino, J.-M.; Fiaud, C. Morphology and mechanisms of formation of natural patinas on archaeological Cu–Sn alloys. Corros. Sci. 1998, 40, 2083–2111. [Google Scholar] [CrossRef]
  17. Masi, G.; Esvan, J.; Josse, C.; Chiavari, C.; Bernardi, E.; Martini, C.; Bignozzi, M.C.; Gartner, N.; Kosec, T.; Robbiola, L. Characterization of typical patinas simulating bronze corrosion in outdoor conditions. Mater. Chem. Phys. 2017, 200, 308–321. [Google Scholar] [CrossRef] [Green Version]
  18. Luo, W.; Jin, R.; Qin, Y.; Huang, F.; Wang, C. Analysis of the corrosion products of the ancient bronzes excavated from Qiaojiayuan tombs. Appl. Phys. Res. 2010, 2, 156. [Google Scholar] [CrossRef]
  19. Oudbashi, O. A methodological approach to estimate soil corrosivity for archaeological copper alloy artefacts. Heritage Sci. 2018, 6, 2. [Google Scholar] [CrossRef] [Green Version]
  20. Arroyave, C.; Lopez, F.; Morcillo, M. The early atmospheric corrosion stages of carbon steel in acidic fogs. Corros. Sci. 1995, 37, 1751–1761. [Google Scholar] [CrossRef] [Green Version]
  21. Chang, T.; Herting, G.; Goidanich, S.; Amaya, J.S.; Arenas, M.; Le Bozec, N.; Jin, Y.; Leygraf, C.; Wallinder, I.O. The role of Sn on the long-term atmospheric corrosion of binary Cu-Sn bronze alloys in architecture. Corros. Sci. 2019, 149, 54–67. [Google Scholar] [CrossRef]
  22. De Oliveira, F.; Lago, D.; Senna, L.; De Miranda, L.; D’elia, E. Study of patina formation on bronze specimens. Mater. Chem. Phys. 2009, 115, 761–770. [Google Scholar] [CrossRef]
  23. Walker, R. Aqueous Corrosion of Tin-Bronze and Inhibition by Benzotriazole. Corrosion 2000, 56, 1211–1219. [Google Scholar] [CrossRef]
  24. Hu, Y.; Wei, Y.; Li, L.; Zhang, J.; Chen, J. Same site, different corrosion phenomena caused by chloride: The effect of the archaeological context on bronzes from Sujialong Cemetery, China. J. Cult. Herit. 2021, 52, 23–30. [Google Scholar] [CrossRef]
  25. Alfantazi, A.; Ahmed, T.; Tromans, D. Corrosion behavior of copper alloys in chloride media. Mater. Des. 2009, 30, 2425–2430. [Google Scholar] [CrossRef]
  26. Constantinides, I.; Adriaens, A.; Adams, F. Surface Characterization of Artificial Corrosion Layers on Copper Alloy Reference Materials. Appl. Surf. Sci. 2002, 189, 90–101. [Google Scholar] [CrossRef]
  27. Liang, Z.; Jiang, K.; Zhang, T. Corrosion behaviour of lead bronze from the Western Zhou Dynasty in an archaeological-soil medium. Corros. Sci. 2021, 191, 109721. [Google Scholar] [CrossRef]
  28. Wallinder, I.O.; Zhang, X.; Goidanich, S.; Le Bozec, N.; Herting, G.; Leygraf, C. Corrosion and runoff rates of Cu and three Cu-alloys in marine environments with increasing chloride deposition rate. Sci. Total Environ. 2014, 472, 681–694. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Zhao, J.; Qiu, R.; Chai, F.; Chen, L.; Li, T.; Zhu, H. A Study on Corrosion Behavior of Low-Alloy Steel for Ships in Simulated Seawater. J. Zhejiang Sci.-Tech. Univ. 2014, 31, 487–490. [Google Scholar] [CrossRef]
  30. González-Parra, R.; Covelo, A.; Barba, A.; Hernández, M. Electrochemical Polarization as a Sustainable Method for the Formation of Bronze Patina Layers on a Quaternary Copper Alloy: Insight into Patina Morphology and Corrosion Behaviour. Sustainability 2023, 15, 1899. [Google Scholar] [CrossRef]
  31. González-Parra, R.; Covelo, A.; Hernández, M. Determination of optimal electrochemical parameters to reproduce copper artistic patina on quaternary alloys. Mater. Lett. 2022, 309, 131414. [Google Scholar] [CrossRef]
  32. Wang, T.; Wang, J.; Wu, Y. The inhibition effect and mechanism of L-cysteine on the corrosion of bronze covered with a CuCl patina. Corros. Sci. 2015, 97, 89–99. [Google Scholar] [CrossRef]
  33. Liao, X.N.; Cao, F.H.; Chen, A.N.; Liu, W.J.; Zhang, J.Q.; Cao, C.N. In-situ investigation of atmospheric corrosion behavior of bronze under thin electrolyte layers using electrochemical technique. Trans. Nonferrous Met. Soc. China 2012, 22, 1239–1249. [Google Scholar] [CrossRef]
  34. Yang, X.; Wu, W.; Chen, K. Investigation on the electrochemical evolution of the Cu-sn-Pb ternary alloy covered with CuCl in a simulated atmospheric environment. J. Electroanal. Chem. 2022, 921, 116636. [Google Scholar] [CrossRef]
  35. Kear, G.; Barker, B.; Walsh, F. Electrochemical corrosion of unalloyed copper in chloride media—A critical review. Corros. Sci. 2004, 46, 109–135. [Google Scholar] [CrossRef]
  36. Chmielová, M.; Seidlerová, J.; Weiss, Z. X-ray diffraction phase analysis of crystalline copper corrosion products after treatment in different chloride solutions. Corros. Sci. 2003, 45, 883–889. [Google Scholar] [CrossRef]
  37. Grayburn, R.; Dowsett, M.; Hand, M.; Sabbe, P.J.; Thompson, P.; Adriaens, A. Tracking the progression of bronze disease—A synchrotron X-ray diffraction study of nantokite hydrolysis. Corros. Sci. 2015, 91, 220–223. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the Cu-20 wt%Sn bronze in 0.1 mol/L NaCl solution. (a) Without soaking and (b) when soaked for 30 days.
Figure 1. XRD patterns of the Cu-20 wt%Sn bronze in 0.1 mol/L NaCl solution. (a) Without soaking and (b) when soaked for 30 days.
Materials 16 05144 g001
Figure 2. (a) Optical micrograph and (b) SEM image of the Cu-20 wt%Sn bronze (the red circle area indicated by the arrow is α or δ phase).
Figure 2. (a) Optical micrograph and (b) SEM image of the Cu-20 wt%Sn bronze (the red circle area indicated by the arrow is α or δ phase).
Materials 16 05144 g002
Figure 3. SEM-EDS image of the Cu-20 wt%Sn bronze soaked in 0.1 mol/L NaCl solution, without corrosion (a,c,d) and with corrosion 30 days (b,e,f), the red circle area indicated by the arrow is α or δ phase.
Figure 3. SEM-EDS image of the Cu-20 wt%Sn bronze soaked in 0.1 mol/L NaCl solution, without corrosion (a,c,d) and with corrosion 30 days (b,e,f), the red circle area indicated by the arrow is α or δ phase.
Materials 16 05144 g003
Figure 4. SEM-EDS image of the cross-section of the Cu-20 wt%Sn bronze soaked in 0.1 mol/L NaCl solution for 30 days: SEM (a) and EDS (bf). Areas (1–5) represent different areas of the sample, arrows indicate α or δ phase, the area above the purple line shows the corrosion area.
Figure 4. SEM-EDS image of the cross-section of the Cu-20 wt%Sn bronze soaked in 0.1 mol/L NaCl solution for 30 days: SEM (a) and EDS (bf). Areas (1–5) represent different areas of the sample, arrows indicate α or δ phase, the area above the purple line shows the corrosion area.
Materials 16 05144 g004
Figure 5. Element distribution in the cross-section of the Cu-20 wt%Sn bronze soaked in 0.1 mol/L NaCl solution for 30 days: (a) Cu, (b) Sn, (c) O, and (d) Cl.
Figure 5. Element distribution in the cross-section of the Cu-20 wt%Sn bronze soaked in 0.1 mol/L NaCl solution for 30 days: (a) Cu, (b) Sn, (c) O, and (d) Cl.
Materials 16 05144 g005
Figure 6. SEM images and EDS spectra of Cu-20 wt%Sn in NaCl solution (1 h): (a,e) for 1 mol/L, (b,f) for 0.1 mol/L, (c,g) for 0.01 mol/L, and (d,h) for 0.001 mol/L.
Figure 6. SEM images and EDS spectra of Cu-20 wt%Sn in NaCl solution (1 h): (a,e) for 1 mol/L, (b,f) for 0.1 mol/L, (c,g) for 0.01 mol/L, and (d,h) for 0.001 mol/L.
Materials 16 05144 g006aMaterials 16 05144 g006b
Figure 7. OCP curves of the Cu-20 wt%Sn bronze in NaCl solutions.
Figure 7. OCP curves of the Cu-20 wt%Sn bronze in NaCl solutions.
Materials 16 05144 g007
Figure 8. Tafel curves of the Cu-20 wt%Sn bronze in NaCl solutions.
Figure 8. Tafel curves of the Cu-20 wt%Sn bronze in NaCl solutions.
Materials 16 05144 g008
Figure 9. (a) Nyquist plot and (b) Bode plot of the Cu-20 wt%Sn bronze in NaCl solutions.
Figure 9. (a) Nyquist plot and (b) Bode plot of the Cu-20 wt%Sn bronze in NaCl solutions.
Materials 16 05144 g009
Figure 10. (a) Nyquist plot and (b,c) Bode plots of the Cu-20 wt%Sn bronze soaked in 0.1 mol/L NaCl solution for different time spans.
Figure 10. (a) Nyquist plot and (b,c) Bode plots of the Cu-20 wt%Sn bronze soaked in 0.1 mol/L NaCl solution for different time spans.
Materials 16 05144 g010
Figure 11. Equivalent circuits show the Cu-20 wt%Sn bronze soaked in 0.1 mol/L NaCl solution initially (a) and for 30 days (b).
Figure 11. Equivalent circuits show the Cu-20 wt%Sn bronze soaked in 0.1 mol/L NaCl solution initially (a) and for 30 days (b).
Materials 16 05144 g011
Figure 12. Schematic of the corrosion mechanism of the Cu-20 wt%Sn bronze after immersion in 0.1 mol/L NaCl solution (the arrow direction indicates that the soaking time of the sample increases sequentially).
Figure 12. Schematic of the corrosion mechanism of the Cu-20 wt%Sn bronze after immersion in 0.1 mol/L NaCl solution (the arrow direction indicates that the soaking time of the sample increases sequentially).
Materials 16 05144 g012
Table 1. The compositions of different phases in the Cu-20 wt%Sn bronze.
Table 1. The compositions of different phases in the Cu-20 wt%Sn bronze.
PositionSn (wt%)Cu (wt%)
average value20.1279.88
α-phase13.5086.50
(α + δ) phase23.8276.18
Table 2. EDS analysis of the Cu-20 wt%Sn bronze with and without corrosion.
Table 2. EDS analysis of the Cu-20 wt%Sn bronze with and without corrosion.
PositionSn (wt%)Cu (wt%)O (wt%)
Without corrosionα-phase13.5086.50<D.L.
δ-phase23.8276.18<D.L.
With corrosionα-phase48.4129.5322.06
δ-phase37.3259.523.16
Table 3. Fitting parameters of the polarization curves of the Cu-20 wt%Sn bronze in NaCl solutions.
Table 3. Fitting parameters of the polarization curves of the Cu-20 wt%Sn bronze in NaCl solutions.
SolutionEcorr/VIcorr/A·cm2Ba/V−1Ba/V−1Rp/KΩ
1 mol/L−0.2232.769 × 10−68.80110.9137.9648
10−1 mol/L−0.1874.578 × 10−72.74511.16068.301
10−2 mol/L−0.1523.032 × 10−73.92113.84380.723
10−3 mol/L−0.1231.101 × 10−731.9449.34895.635
Table 4. EIS fitting results of the Cu-20 wt%Sn bronze in NaCl solutions.
Table 4. EIS fitting results of the Cu-20 wt%Sn bronze in NaCl solutions.
SolutionRs/(Ω·cm2)Qdl/(Ω−1·cm2·s n)nRct/(Ω·cm2)
1 mol/L6.446.148 × 10−50.81859406
10−1 mol/L54.927.777 × 10−50.769116,306
10−2 mol/L478.51.333 × 10−40.667237,271
10−3 mol/L58,5561.168 × 10−40.691858,556
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

Song, Z.; Tegus, O. The Corrosion Properties of Bronze Alloys in NaCl Solutions. Materials 2023, 16, 5144. https://doi.org/10.3390/ma16145144

AMA Style

Song Z, Tegus O. The Corrosion Properties of Bronze Alloys in NaCl Solutions. Materials. 2023; 16(14):5144. https://doi.org/10.3390/ma16145144

Chicago/Turabian Style

Song, Zhiqiang, and Ojiyed Tegus. 2023. "The Corrosion Properties of Bronze Alloys in NaCl Solutions" Materials 16, no. 14: 5144. https://doi.org/10.3390/ma16145144

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