Next Article in Journal
An Experimental Study on Blade Surface De-Icing Characteristics for Wind Turbines in Rime Ice Condition by Electro-Thermal Heating
Previous Article in Journal
Antimicrobial Activity of Photocatalytic Coatings on Surfaces: A Systematic Review and Meta-Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Analysis of Blackening Reaction of Zn-Mg-Al Alloy-Coated Steel Prepared by Water Vapor Treatment

1
Department of Coast Guard Studies, Korea Maritime and Ocean University, Busan 49112, Republic of Korea
2
Department of Maritime Safety, Korea Maritime Transportation Safety Authority, Sejong 30100, Republic of Korea
3
Steel Solution R&D Center, POSCO, Incheon 21985, Republic of Korea
4
Department of Marine Engineering, Korea Maritime and Ocean University, Busan 49112, Republic of Korea
*
Author to whom correspondence should be addressed.
Coatings 2024, 14(1), 93; https://doi.org/10.3390/coatings14010093
Submission received: 8 December 2023 / Revised: 5 January 2024 / Accepted: 8 January 2024 / Published: 10 January 2024
(This article belongs to the Section Surface Characterization, Deposition and Modification)

Abstract

:
In the context of high-temperature water vapor treatment, Zn-Mg-Al alloy-coated steel sheets exhibit the emergence of a black surface. This study aims to explore the factors and mechanisms contributing to surface blackening by inducing black surfaces on Zn-Mg-Al alloy-coated steel sheets, which were fabricated through molten coating subjected to water vapor treatment at 150 degrees Celsius. The surface composition was predominantly identified as zinc oxide (ZnO) film validated through X-ray diffraction (XRD) and X-ray photoelectron spectroscopy (XPS). Morphological analysis of the surface and cross-section post-water vapor treatment revealed a disrupted lamellar structure with diffused features, resulting from the formation of an oxide film. Optical properties analysis demonstrated an increased absorbance and a decreased bandgap energy after water vapor treatment, which is indicative of an augmented blackening effect. Consequently, the high-temperature water vapor treatment led to the formation of oxides on the surface with the highly reactive Mg and Al extracting oxygen from the predominantly present ZnO surface. This process resulted in the creation of an oxygen-deficient oxide, ultimately causing surface blackening.

1. Introduction

In the contemporary landscape of industries, where sectors such as automotive, architecture, and electronics thrive on the pursuit of distinctive designs and unique color schemes [1], the significance of black surfaces has emerged as a prevailing trend. Black products, often associated with a luxurious, urban, and modern image, have witnessed a notable surge in demand across diverse age demographics [2]. Beyond the aesthetic allure, black surfaces confer functional advantages, particularly in terms of superior heat absorption [3,4]. The conventional approach to achieving a black appearance involves coating surfaces with black pigments, which is a process frequently applied to steel following zinc (Zn) or zinc-alloy plating with magnesium (Mg) and aluminum (Al) [5,6]. However, this method is accompanied by challenges such as cost implications, limitations in additional forming processes like welding, and environmental concerns, including volatile organic compound (VOC) emissions [7]. Moreover, relying solely on coating for blackening imposes constraints on achieving performance improvements beyond aesthetics with potential issues arising from coating damages and subsequent corrosion resistance problems [8,9]. To address these challenges, recent research has explored alternative, environmentally friendly approaches such as water vapor treatment and anodizing for achieving blackening on steel surfaces, moving away from traditional coating methods applied to Zn-Mg-Al alloy-plated steel sheets [5,6,10,11]. This study focuses explicitly on the water vapor treatment method for blackening, which is applied to Zn-Mg-Al alloy-plated steel produced through the hot-dipping method. Distinguishing itself from ongoing research in the field of Zn-Mg-Al alloy-plated steel, this study places emphasis on two crucial aspects. Firstly, it accentuates the blackening effect, prioritizing the enhancement of surface characteristics by delving into the formation of the blackening oxide layer. While existing studies often concentrate on corrosion resistance or physical properties, this research uniquely prioritizes visual and optical effects through changes in the blackening layer [12,13,14]. Secondly, by unraveling the mechanism behind the blackening effect, this study aims to explore novel applications for Zn-Mg-Al alloy-plated steel, surpassing the realm of corrosion resistance. The objective is to innovate both aesthetically and functionally, extending the potential applications of blackening. In summary, this research, distinct in its approach from existing studies, seeks to innovatively enhance the surface characteristics of Zn-Mg-Al alloy-plated steel through the blackening effect and broaden the applicability of blackening effects. Deemed crucial, this study could pave the way for new horizons in future surface treatment technologies and material applications.

2. Materials and Methods

The test specimens of Zn-Mg-Al were obtained by purchasing PosMAC steel plates from the South Korean company POSCO, which were produced using a hot dipping method [15,16]. The nominal composition of the alloy coating is 94.5 wt.% zinc, 3.0 wt.% magnesium, and 2.5 wt.% aluminum. Prior to water vapor treatment, impurities on the surface were removed by degreasing using sodium hydroxide (NaOH) and ultrasonic cleaning in ethanol (C2H5OH). The specimen was then washed with ultrapure water with a resistance value of 18.2 MΩ·cm and dried completely in air. To form a black coating layer, the prepared coating was subjected to water vapor treatment for 1 h and 30 min in a saturated steam atmosphere at 150 °C, according to the process shown in Figure 1. This treatment involves the infiltration of water molecules into an extremely thin naturally formed oxide layer on the metal plating surface in a high-temperature steam environment. During this process, electrons released from the Zn-Mg-Al alloy plating combine with water molecules, forming hydroxide ions (OH). The interaction between hydroxide ions and metal ions leads to the continuous formation of oxides in the high-temperature steam environment. Based on these reactions, a black surface layer is formed on the surface [17]. The appearance of the test specimen with the black surface formed through water vapor treatment can be observed in Figure 2. In this paper, the specimen before water vapor treatment is referred to as Bare, and the specimen after water vapor treatment is referred to as Bare ST. To elucidate the reaction of black coating formation, the surface and cross-sectional morphology of the specimens were observed using a field emission-scanning electron microscope (FE-SEM, MIRA 3, Tescan, Brno, Czech Republic) with an accelerating voltage of 10 kV and transmission electron microscope (TEM, JEM-2100F, JEOL, Tokyo, Japan) with an accelerating voltage of 200 kV. Elemental compositions of iron (Fe), Zn, O, Al, and Mg in the depth direction were analyzed using a glow discharge mass spectrometer (GD-MS, Element GD, Thermo-fisher, Waltham, MA, USA), and surface elemental compositions were analyzed through energy-dispersive spectroscopy (EDS) attached to FE-SEM equipment. X-ray diffraction (XRD, SmartLab, Rigaku, Tokyo, Japan) using Cu kα radiation (λ = 1.5418 Å, 40 kV, 200 mA) and a scanning speed of 1° min−1 at a 2θ range of 10–90° was applied to characterize the crystalline composition of the coatings. The surface chemistry analyzed by X-ray photoelectron spectroscopy (XPS, AXIS SUPRA, KRATOS Analytical Ltd., Manchester, UK) using Al Kα radiation (1486.6 eV, 20 kV, 15 mA). Binding energies were calibrated by the C1s of 284.8 eV. A UV-Vis/NIR spectrometer (V-770, JASCO, Tokyo, Japan) was used for optical property analysis, including absorbance and bandgap. UV-Vis/NIR spectra were analyzed within the range of 200–2500 nm.

3. Results and Discussion

3.1. Material Characterization

Figure 3 illustrates the surface morphology of test specimens for both Bare and Bare ST. In Figure 3a, the surface SEM image of Bare was analyzed at three different points, each presumed to exhibit distinct metal structures, with confirmed elemental compositions. For Bare ST, the surface closely resembled point 3 in Figure 3b, and additional surfaces corresponding to points 1 and 2 were identified. Elemental composition analysis for these points was conducted using EDS. As observed in Figure 3a, a lamellar structure consistent with previous research on Zn-Mg-Al was noted, indicating the presence of the Zn-matrix phase, Zn-MgZn2, and Zn-Mg2Zn11 binaries as well as Zn-MgZn2-Al ternary phases [18]. In Figure 3a, point 1 lacked a lamellar structure in the SEM image, and elemental composition analysis revealed a predominant presence of Zn. This implies a structure of Zn-Al with a Zn-matrix phase or an Al eutectic. For points 2 and 3, a decrease in Zn elemental composition and an increase in the Mg and Al ratio were observed through Table 1. Consequently, it was inferred that these points might correspond to Zn-MgZn2 binary, Zn-Mg2Zn11 binary, or Zn-MgZn2-Al ternary phases, as suggested by the elements presented in the surface composition of the Bare specimen in Table 1. Specifically, the examination of the bare specimen’s surface in Figure 3a speculated the presence of the Zn-matrix phase and other binary, ternary phases.
In Figure 3b of Bare ST, the destruction of the lamellar structure is depicted, and it is replaced by a diffusion-like shape. Point 3, characterized by the highest oxygen (O) content and a rough surface, is presumed to have undergone the most significant oxide formation during water vapor treatment. This observation suggests that light scattering due to surface roughness may not be the primary factor in blackening, but it could contribute to the development of a blackened appearance [19,20]. On the other hand, points 1 and 2, retaining the lamellar structure while undergoing diffusion, exhibit relatively less oxide formation, as indicated by the elements present in the surface composition of the Bare ST specimen in Table 2. In the case of point 3, which occupies the largest surface area on the surface of Bare ST, the proportion of Al increases. This is attributed to the high reactivity of Al in the presence of high-temperature water vapor, leading to oxidation and the formation of a surface layer. It is presumed that this process may assist in facilitating the oxidation of Zn in MgZn2 in the depth direction, contributing to the enhancement of blackening [8]. Analysis of the O composition in Bare ST suggests an association between the advancement of blackening and an increase in the ratio of O elements.
Additionally, to examine the surface morphology in more detail, a higher magnification of ×35,000 was employed to confirm that in Bare test specimens, the lamellar structure was maintained, which is consistent with the observations at ×5000. However, in the case of Bare ST test specimens, the lamellar structure could not be observed. Hence, due to the high-temperature water vapor treatment in the Bare, a distinct surface was formed in Bare ST. Through surface morphology and elemental composition, this is presumed to be an oxide layer, and it is identified as the primary factor contributing to surface blackening.
Similarly to the surface morphology results observed in Figure 3, when examining the cross-sectional morphology of the Bare specimen in Figure 4, it was confirmed that a lamellar structure is formed between the intermetallic compounds. Additional TEM analysis with EDS mapping revealed predominantly formed Zn, and when confirming the point element composition in the region where Mg elements were mapped, it was determined to be 23.99 at% Mg and 71.75% Zn. This suggests the formation of the intermetallic compound MgZn2. Furthermore, Al was found to be distributed around Zn and Mg, which is similar to the results observed on the surface. The oxide layer, as seen in the mapping, was found to be extremely thin near the surface, indicating the presence of a naturally formed native oxide layer on the metal surface.
In contrast, Figure 5 shows the examination of cross-sectional morphology and EDS mapping of Bare ST, where it was observed that the lamellar structure, akin to the surface morphology, was disrupted and appeared to be diffusing. This phenomenon was attributed to surface treatment occurring due to high-temperature steam. Through EDS mapping to confirm the positions of each element, it was evident that the boundaries were clearly defined in Bare, while in Bare ST, all elements were diffused. Particularly, oxygen (O) was found to be formed approximately 5 μm deep from the surface. This led to the conclusion that an oxide layer formed through steam treatment on the surface.
Figure 6 presents the depth-wise elemental composition analysis results from the surface analyzed through GD-MS. When examining the depth-wise elemental composition of Bare, it was observed that Zn predominantly occupied the surface, which is consistent with the composition of the test specimen. Small amounts of Mg and Al were also detected, and Fe showed an increase from a depth of 10 μm, intersecting with the decreasing Zn, indicating the formation of an approximately 10 μm alloy coating on the specimen.
Unlike Bare, in the case of Bare ST, a higher ratio of O was observed on the surface, with the ratio gradually decreasing to a depth of about 5 μm. Additionally, the depth of the coating layer was approximately 14 μm, with a surface oxide layer of 5 μm thickness. This observation indicates an increase in the thickness of the oxide layer. This observation aligns with the results obtained through TEM’s oxygen mapping, supporting the hypothesis that the main factor contributing to blackening is the formation of an oxide layer on the surface.
Next, we analyzed the crystal structure and phase information of the thin film using XRD. As seen in Figure 7, peaks for mainly Zn and MgZn2 were detected in the Bare specimen. In contrast, in Bare ST, peaks for the oxide, ZnO, were additionally confirmed, and peaks related to reactions with high-temperature water vapor, such as Mg2Zn11, were observed [21]. The formation of oxides on the surface was further confirmed, which was consistent with elemental composition analysis through EDS. Based on the ZnO peak from the XRD analysis, the crystal size and interplanar spacing were analyzed using Bragg’s law in Equations (1) and (2). As shown in Figure 8, the interplanar spacing of ZnO increased in Bare ST [22]. This is presumed to result from a chemical state change or oxygen vacancy, as illustrated in Figure 9 [23].
L = 0.9 λ F W H M · cos θ
(L: lattice interplanar spacing of the crystal, λ: wavelength of the characteristic X-ray, ϴ: X-ray incidence angle).
D = n · λ sin θ
(D: spacing of the crystal layers, n = integer, λ: wavelength of the X-ray, θ: the angle between incident ray and the scatter plane).
In addition, the crystal structure was verified through XPS spectra analysis, as illustrated in Figure 10 and Figure 11. Following water vapor treatment, discernible peaks emerged in the binding energy of the oxide in the Zn 2p, Mg 1s, and Al 2p results, corroborating the formation of oxide, which is in line with the findings from XRD.
Upon examining the Zn LMM spectra for the oxygen Auger peak in the Bare specimen, a peak indicative of Zn2+ ion binding was observed. Conversely, in the Bare, a peak corresponding to Zn1+ binding was identified. This observation suggests a shift in the ZnO peak in Bare ST, signifying a transformation from ZnO to ZnO1−x. The alteration is likely attributed to the change in oxidation state during the transformation from ZnO to ZnO1−x [24]. Furthermore, as depicted in Figure 11, the O1s binding energy peak was subjected to fitting, revealing three distinct peaks with binding energies of approximately 530.2, 530, and 528 eV, respectively. The lowest binding energy is associated with the metal–oxygen bond (M-O), such as ZnO or MgO; the intermediate energy corresponds to oxygen vacancy, and the highest energy is linked to the bond with moisture on the metal surface (OH) [25]. Notably, oxygen vacancy was not detected in Bare, but its occurrence was confirmed in Bare ST. Therefore, the observed increase in the interplanar distance of ZnO, as evidenced in the XRD results, is attributed to the occurrence of oxygen vacancy in ZnO, leading to the formation of an oxygen-deficient oxide.

3.2. Optical Characterization

Figure 12 illustrates the analysis of optical properties using UV-Vis-nir to measure reflectance across the wavelength range of 200–2500 nm. Upon measuring reflectance, it was observed that the black specimen exhibited low reflectance, which is akin to the lightness depicted in Figure 2. For the conversion of reflectance to absorbance, representing the extent of light absorption, the Kubelka–Munk equation (Equation (3)) was applied [26]. Upon examining the absorbance results, a sharp increase was observed around 390 nm in the visible light range for the Bare ST, suggesting the absorption of visible light by a certain factor, as depicted in Figure 12.
F ( R ) = 1 R 2 2 R = a s
(F(R): absorbance, R: diffuse reflectance, a: absorption coefficient, s: scattering coefficient).
This is attributed to the absorption of visible light. The absorption is a result of ZnO defects caused by an increase in interfacial distance due to oxygen vacancies, as revealed in the preceding XRD and XPS analyses [27,28]. Furthermore, the bandgap was measured based on the obtained absorbance using the Tauc plot. The Tauc plot graph can be obtained by using photon energy and optical energy as the x- and y-axes, respectively, and the formulas for obtaining photon energy and optical energy are shown in Equations (4) and (5) [29].
E = h c λ
(E: photon energy, h: Planck constant, c: speed of light, λ: wavelength of light).
ahv 2 = 2.303 · F ( R )
E(F(R): absorbance, E: photon energy).
Utilizing the absorbance data, the calculated bandgaps for Bare and Bare ST were found to be approximately 4.2 eV and 2.9 eV, respectively, as illustrated in Figure 13. The optical bandgap energy of Bare, estimated at 4.2 eV, corresponds to the optical bandgap energy of MgO [30]. This led to the conjecture that MgO was naturally formed in small quantities. In the case of Bare ST, a noticeable decrease in the optical bandgap energy to approximately 2.9 eV was observed. This is attributed to the thickening of the oxide layer on the surface, specifically the formation of the oxide layer of ZnO, as confirmed by cross-sectional morphology and GD-MS [31]. Bare ST’s bandgap energy is lower than the typical optical bandgap of ZnO, which is 3.4 eV, indicating that the bandgap energy decreases with the ratio of oxygen in ZnO [32]. In another study, it was observed that by adjusting the oxygen deficiency in the ZnO coating, the bandgap energy decreased to approximately 2.5 eV at an oxygen vacancy concentration of 4.8% and ZnO composition of 0.952 [33,34]. Therefore, the bandgap measurement implies the presence of oxygen deficiency in ZnO. As observed in Figure 14, the reactive Mg and Al ions combine with oxygen to form ZnO1−x, an oxygen-deficient oxide, by extracting oxygen from ZnO. Based on these results, it is inferred that the ZnO1−x structure, formed through the absorption of visible light, contributes to the development of a blackened surface [35,36].

4. Conclusions

This study investigates the reaction of blackening on Zn-Mg-Al alloy-coated steel sheets resulting from high-temperature water vapor treatment via the hot-dip galvanizing method. The key findings can be summarized as follows: After water vapor treatment, the steel sheet exhibited increased surface roughness and slightly enlarged particle size, with cross-sectional analysis revealing the formation of an oxide layer, primarily comprising ZnO. Confirmed oxides in Bare ST indicated the creation of oxygen vacancies, leading to an increased lattice spacing of ZnO and the formation of an oxygen-deficient oxide. Bare ST’s optical bandgap measurement revealed a value of approximately 2.9 eV, significantly lower than ZnO’s typical 3.4 eV, suggesting the formation of a non-stoichiometric structure (ZnO1−x) due to defects, with visible light absorption identified as the cause of blackening. With a composition of Zn 94.5 wt.%, Mg 3.0 wt.%, Al 2.5 wt.%, future research should explore blackening under varied Mg and Al ratios and investigate factors influencing blackening at lower energy levels, contributing to a comprehensive understanding. In conclusion, this study provides valuable insights into the intricate process of blackening, paving the way for further exploration and the optimization of conditions in future research.

Author Contributions

Formal analysis, S.-H.K.; writing—original draft preparation, S.-H.K., Y.-J.K., K.-H.L., J.K., M.-H.L. and Y.-S.Y.; writing—review and editing, supervision, Y.-S.Y.; funding acquisition, Y.-S.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Korea Evaluation Institute of Industrial Technology (KEIT) grant, funded by the Korea Government (MOTIE), grant number 1415184337.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

This research was a part of the project titled “Development of Surface Blackening Technology for High Corrosion Resistance Galvanized Alloy Coating”.

Conflicts of Interest

Author Yu-Jin Kang was employed by the company KOMSA and Author Kyung-Hwang Lee was employed by the company POSCO. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Babolhavaeji, M.; Vakilian, M.A.; Slambolchi, A. The role of product color in consumer behavior. Adv. Soc. Humanit. Manag. 2015, 2, 9–15. [Google Scholar]
  2. Labrecque, L.I.; Patrick, V.M.; Milne, G.R. The marketers’ prismatic palette: A review of color research and future directions. Psychol. Mark. 2013, 30, 187–202. [Google Scholar] [CrossRef]
  3. Lee, J.; Kim, D.; Choi, C.; Chung, W. Nanoporous anodic alumina oxide layer and its sealing for the enhancement of radiative heat dissipation of aluminum alloy. Nano Energy 2017, 31, 504–513. [Google Scholar] [CrossRef]
  4. Park, J.S.; Yun, D.B.; Kim, S.H.; Kim, T.Y.; Kim, S.J. Effects of hairline treatment on surface blackening and thermal diffusion of Zn-Al-Mg alloy-coated steel sheet. J. Korean Soc. Surf. Sci. Eng. 2023, 56, 69–76. [Google Scholar]
  5. Nakano, T.; Yamamoto, M.; Taketsu, H. Method for Producing Black-Plated Steel Sheet, and Method for Producing Molded Article of Black-Plated Steel Sheet. U.S. Patent 9,598,759, 21 March 2017. [Google Scholar]
  6. Nakano, T.; Ueno, S.; Yamamoto, M. Method for Manufacturing Black Plated Steel Sheet, Apparatus for Manufacturing Black Plated Steel Sheet, and System for Manufacturing Black Plated Steel Sheet. U.S. Patent 10,697,053, 30 June 2020. [Google Scholar]
  7. Porwal, T. Paint pollution harmful effects on environment. Soc. Issues Environ. Probl. 2015, 3, 2394–3629. [Google Scholar] [CrossRef]
  8. Nakano, T.; Yamamoto, M.; Taketsu, H. Black-Plated Steel Sheet. U.S. Patent 9,863,027, 9 January 2018. [Google Scholar]
  9. Lee, K.H.; Jeong, J.I.; Kim, H.J.; Yang, J.H. Black Plated Steel Sheet and Manufacturing Method Thereof. U.S. Patent 11,555,240, 17 January 2023. [Google Scholar]
  10. Kim, S.J.; Kwak, Y.J.; Kim, T.Y.; Jung, W.S.; Kim, K.Y. Surface darkening phenomenon of Zn–Mg alloy coated steel exposed to aqueous environment at high temperature. J. Mater. Res. 2015, 30, 3605–3615. [Google Scholar] [CrossRef]
  11. Masuda, R.; Kowalski, D.; Kitano, S.; Aoki, Y.; Nozawa, T.; Habazaki, H. Characterization of dark-colored nanoporous anodic films on zinc. Coatings 2020, 10, 1014. [Google Scholar] [CrossRef]
  12. Rai, P.K.; Rout, D.; Satish Kumar, D.; Sharma, S.; Balachandran, G. Effect of magnesium on corrosion behavior of hot-dip Zn-Al-Mg coating. J. Mater. Eng. Perform. 2021, 30, 4138–4147. [Google Scholar] [CrossRef]
  13. Zou, Y.; Wu, X.; Tang, S.; Zhu, Q.; Song, H.; Guo, M.; Cao, L. Investigation on microstructure and mechanical properties of Al-Zn-Mg-Cu alloys with various Zn/Mg ratios. J. Mater. Sci. Technol. 2021, 85, 106–117. [Google Scholar] [CrossRef]
  14. Kim, K.; Grandhi, S.; Oh, M. Improving the Coatability of Zn–Mg–Al Alloy on Steel Substrate by the Surface Pretreatment of SnCl2-Added Zinc Ammonium Chloride. Appl. Sci. 2023, 13, 950. [Google Scholar] [CrossRef]
  15. Son, I.R.; Kim, T.C.; Ju, G.I.; Kim, M.S.; Kim, J.S. Development of PosMAC® Steel and Its Application Properties. Korean J. Met. Mater. 2021, 59, 613–623. [Google Scholar] [CrossRef]
  16. Kim, T.C.; Kim, S.H.; Kim, S.Y.; Oh, M.S.; Yu, B.H.; Kim, J.S. Hot dip Zn-Al-Mg Alloy Plated Steel Sheet Having Excellent Corrosion Resistance and Method for Manufacturing. K.R. Patent 101,568,509, 20 November 2015. [Google Scholar]
  17. Saadi, N.S.; Hassan, L.B.; Karabacak, T. Metal oxide nanostructures by a simple hot water treatment. Sci. Rep. 2017, 7, 7158. [Google Scholar] [CrossRef] [PubMed]
  18. Dutta, M.; Halder, A.K.; Singh, S.B. Morphology and properties of hot dip Zn–Mg and Zn–Mg–Al alloy coatings on steel sheet. Surf. Coat. Technol. 2010, 205, 2578–2584. [Google Scholar] [CrossRef]
  19. Hruška, P.; More-Chevalier, J.; Novotný, M.; Čížek, J.; Melikhova, O.; Fekete, L.; Poupon, M.; Bulíř, J.; Volfová, L.; Butterling, M. Effect of roughness and nanoporosity on optical properties of black and reflective Al films prepared by magnetron sputtering. J. Alloys Compd. 2021, 872, 159744. [Google Scholar] [CrossRef]
  20. Honson, V.; Huynh-Thu, Q.; Arnison, M.; Monaghan, D.; Isherwood, Z.J.; Kim, J. Effects of shape, roughness and gloss on the perceived reflectance of colored surfaces. Front. Psychol. 2020, 11, 485. [Google Scholar] [CrossRef] [PubMed]
  21. Zywitzki, O.; Modes, T.; Scheffel, B.; Metzner, C. Examination of the Mg-Zn Phase Formation in Hot-Dip Galvanized Steel Sheet. Pract. Metallogr. 2012, 49, 210–220. [Google Scholar] [CrossRef]
  22. Epp, J. X-ray diffraction (XRD) techniques for materials characterization. In Materials Characterization Using Nondestructive Evaluation (NDE) Methods; Elsevier: Amsterdam, The Netherlands, 2016; pp. 81–124. [Google Scholar]
  23. Li, X.; Wang, Y.; Liu, W.; Jiang, G.; Zhu, C. Study of oxygen vacancies′ influence on the lattice parameter in ZnO thin film. Mater. Lett. 2012, 85, 25–28. [Google Scholar] [CrossRef]
  24. Lauermann, I.; Bär, M.; Fischer, C. Synchrotron-based spectroscopy for the characterization of surfaces and interfaces in chalcopyrite thin-film solar cells. Sol. Energy Mater. Sol. Cells 2011, 95, 1495–1508. [Google Scholar] [CrossRef]
  25. Cho, S.W.; Kim, K.S.; Jung, S.H.; Cho, H.K. Towards environmentally stable solution-processed oxide thin-film transistors: A rare-metal-free oxide-based semiconductor/insulator heterostructure and chemically stable multi-stacking. J. Mater. Chem. C 2017, 5, 10498–10508. [Google Scholar] [CrossRef]
  26. Shen, J.; Li, Y.; He, J. On the Kubelka–Munk absorption coefficient. Dyes Pigment. 2016, 127, 187–188. [Google Scholar] [CrossRef]
  27. Wang, J.; Wang, Z.; Huang, B.; Ma, Y.; Liu, Y.; Qin, X.; Zhang, X.; Dai, Y. Oxygen vacancy induced band-gap narrowing and enhanced visible light photocatalytic activity of ZnO. ACS Appl. Mater. Interfaces 2012, 4, 4024–4030. [Google Scholar] [CrossRef] [PubMed]
  28. Chen, L.; Zhai, B.; Huang, Y.M. Rendering visible-light photocatalytic activity to undoped ZnO via intrinsic defects engineering. Catalysts 2020, 10, 1163. [Google Scholar] [CrossRef]
  29. Jubu, P.R.; Obaseki, O.S.; Nathan-Abutu, A.; Yam, F.K.; Yusof, Y.; Ochang, M.B. Dispensability of the conventional Tauc’s plot for accurate bandgap determination from UV–vis optical diffuse reflectance data. Results Opt. 2022, 9, 100273. [Google Scholar] [CrossRef]
  30. Das, O.P.; Pandey, S.K. Optical, Compositional and Electrical Properties of Transparent MgO Thin Film for ReRAM Devices. J. Phys. Conf. Ser. 2023, 2426, 012031. [Google Scholar] [CrossRef]
  31. Qasem, A.; Mostafa, M.S.; Yakout, H.A.; Mahmoud, M.; Shaaban, E.R. Determination of optical bandgap energy and optical characteristics of Cd30Se50S20 thin film at various thicknesses. Opt. Laser Technol. 2022, 148, 107770. [Google Scholar] [CrossRef]
  32. Dostanko, A.P.; Ageev, O.A.; Golosov, D.A.; Zavadski, S.M.; Zamburg, E.G.; Vakulov, D.E.; Vakulov, Z.E. Electrical and optical properties of zinc-oxide films deposited by the ion-beam sputtering of an oxide target. Semiconductors 2014, 48, 1242–1247. [Google Scholar] [CrossRef]
  33. Wang, J.; Chen, R.; Xiang, L.; Komarneni, S. Synthesis, properties and applications of ZnO nanomaterials with oxygen vacancies: A review. Ceram. Int. 2018, 44, 7357–7377. [Google Scholar] [CrossRef]
  34. Zhang, C.; Geng, X.; Liao, H.; Li, C.; Debliquy, M. Room-temperature nitrogen-dioxide sensors based on ZnO1− x coatings deposited by solution precursor plasma spray. Sens. Actuators B Chem 2017, 242, 102–111. [Google Scholar] [CrossRef]
  35. Kanakkillam, S.S.; Krishnan, B.; Guzman, S.S.; Martinez, J.A.A.; Avellaneda, D.A.; Shaji, S. Defects rich nanostructured black zinc oxide formed by nanosecond pulsed laser irradiation in liquid. Appl. Surf. Sci. 2021, 567, 150858. [Google Scholar] [CrossRef]
  36. Xia, T.; Wallenmeyer, P.; Anderson, A.; Murowchick, J.; Liu, L.; Chen, X. Hydrogenated black ZnO nanoparticles with enhanced photocatalytic performance. RSC Adv. 2014, 4, 41654–41658. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of water vapor treatment process.
Figure 1. Schematic diagram of water vapor treatment process.
Coatings 14 00093 g001
Figure 2. Surface observation images and whiteness of Zn-Mg-Al alloy-coated steel specimens by 150 °C water vapor treatment.
Figure 2. Surface observation images and whiteness of Zn-Mg-Al alloy-coated steel specimens by 150 °C water vapor treatment.
Coatings 14 00093 g002
Figure 3. SEM images of surface of Bare and Bare ST; (a) ×5000 Bare, (b) ×5000 Bare ST, (c) ×35,000 Bare, (d) ×35,000 Bare ST.
Figure 3. SEM images of surface of Bare and Bare ST; (a) ×5000 Bare, (b) ×5000 Bare ST, (c) ×35,000 Bare, (d) ×35,000 Bare ST.
Coatings 14 00093 g003
Figure 4. SEM and EDS results of Bare’s cross-section.
Figure 4. SEM and EDS results of Bare’s cross-section.
Coatings 14 00093 g004
Figure 5. SEM and EDS results of Bare ST’s cross section.
Figure 5. SEM and EDS results of Bare ST’s cross section.
Coatings 14 00093 g005
Figure 6. GD-MS depth profile for (a) Bare, (b) Bare ST.
Figure 6. GD-MS depth profile for (a) Bare, (b) Bare ST.
Coatings 14 00093 g006aCoatings 14 00093 g006b
Figure 7. XRD results of Bare and Bare ST.
Figure 7. XRD results of Bare and Bare ST.
Coatings 14 00093 g007
Figure 8. ZnO lattice spacing comparison of Bare and Bare ST.
Figure 8. ZnO lattice spacing comparison of Bare and Bare ST.
Coatings 14 00093 g008
Figure 9. Schematic diagram of metal-deficient oxide with oxygen vacancy.
Figure 9. Schematic diagram of metal-deficient oxide with oxygen vacancy.
Coatings 14 00093 g009
Figure 10. XPS spectra of Bare and Bare ST: (a) survey spectra, (b) Zn 2p, (c) Zn LMM, (d) Mg 1s, (e) Al 2p.
Figure 10. XPS spectra of Bare and Bare ST: (a) survey spectra, (b) Zn 2p, (c) Zn LMM, (d) Mg 1s, (e) Al 2p.
Coatings 14 00093 g010
Figure 11. Bare and Bare ST’s XPS spectra of O1 s fitting.
Figure 11. Bare and Bare ST’s XPS spectra of O1 s fitting.
Coatings 14 00093 g011
Figure 12. Bare and Bare ST’s (a) reflectance, (b) absorbance.
Figure 12. Bare and Bare ST’s (a) reflectance, (b) absorbance.
Coatings 14 00093 g012
Figure 13. Dotted lines indicate estimated bandgap energy of Bare and Bare ST.
Figure 13. Dotted lines indicate estimated bandgap energy of Bare and Bare ST.
Coatings 14 00093 g013
Figure 14. Schematic diagram of Zn-Mg-Al specimen’s blackening.
Figure 14. Schematic diagram of Zn-Mg-Al specimen’s blackening.
Coatings 14 00093 g014
Table 1. EDS quantitative analysis at each point marked on Figure 3a.
Table 1. EDS quantitative analysis at each point marked on Figure 3a.
PointZn (wt.%)Mg (wt.%)Al (wt.%)O (wt.%)
194.121.032.581.53
289.314.956.580.64
383.963.926.124.51
Table 2. EDS quantitative analysis at each point marked on Figure 3b.
Table 2. EDS quantitative analysis at each point marked on Figure 3b.
PointZn (wt.%)Mg (wt.%)Al (wt.%)O (wt.%)
171.455.898.4714.19
265.986.8310.1317.06
359.166.1513.1321.56
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

Kim, S.-H.; Kang, Y.-J.; Lee, K.-H.; Kang, J.; Lee, M.-H.; Yun, Y.-S. Analysis of Blackening Reaction of Zn-Mg-Al Alloy-Coated Steel Prepared by Water Vapor Treatment. Coatings 2024, 14, 93. https://doi.org/10.3390/coatings14010093

AMA Style

Kim S-H, Kang Y-J, Lee K-H, Kang J, Lee M-H, Yun Y-S. Analysis of Blackening Reaction of Zn-Mg-Al Alloy-Coated Steel Prepared by Water Vapor Treatment. Coatings. 2024; 14(1):93. https://doi.org/10.3390/coatings14010093

Chicago/Turabian Style

Kim, Sang-Hee, You-Jin Kang, Kyung-Hwang Lee, Jun Kang, Myeong-Hoon Lee, and Yong-Sup Yun. 2024. "Analysis of Blackening Reaction of Zn-Mg-Al Alloy-Coated Steel Prepared by Water Vapor Treatment" Coatings 14, no. 1: 93. https://doi.org/10.3390/coatings14010093

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