Next Article in Journal
Effect of the Working Pressure and Oxygen Gas Flow Rate on the Fabrication of Single-Phase Ag2O Thin Films
Next Article in Special Issue
Characterization and Corrosion Behavior of Zinc Coatings for Two Anti-Corrosive Protections: A Detailed Study
Previous Article in Journal
Correlation between the Rheological Properties of Asphalt Mortar and the High-Temperature Performance of Asphalt Mixture
Previous Article in Special Issue
Research on the Influence of Coating Technologies on Adhesion Anti-Corrosion Layers in the Case of Al7175 Aluminum Alloy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hybrid Coating of Polystyrene–ZrO2 for Corrosion Protection of AM Magnesium Alloys

by
Luis Chávez
1,
Lucien Veleva
1,*,
Diana Sánchez-Ahumada
2 and
Rafael Ramírez-Bon
3
1
Applied Physics Department, Center for Research and Advanced Studies (CINVESTAV-Merida), Merida 97310, Yucatan, Mexico
2
Facultad de Ingeniería Mochis, Universidad Autónoma de Sinaloa, Fuente de Poseidón y Prol. Angel Flores, S.N., Los Mochis 81223, Sinaloa, Mexico
3
Centro de Investigación y de Estudios Avanzados (CINVESTAV-Querétaro), Fracc. Real de Juruquilla, Querétaro 76230, Querétaro, Mexico
*
Author to whom correspondence should be addressed.
Coatings 2023, 13(6), 1059; https://doi.org/10.3390/coatings13061059
Submission received: 13 May 2023 / Revised: 25 May 2023 / Accepted: 5 June 2023 / Published: 7 June 2023
(This article belongs to the Special Issue Surface Modification of Magnesium, Aluminum Alloys, and Steel)

Abstract

:
A hybrid material of polystyrene (PS)–ZrO2 was developed by the sol–gel technique and deposited by spin-coating on AM60 and AM60–AlN nanocomposite surfaces to enhance corrosion resistance in marine environments.   PS–ZrO 2 with an average thickness of ≈ 305   ±   20   nm was dispersed homogeneously, presenting isolated micro–nano-structure defects with air trapped inside, which led to an increase in roughness (≈4 times). The wettability of the coated substrates was close to the hydrophobic border ( θ CA = 90 ° 94 ° ) . The coated samples were exposed for 30 days to SME solution, simulating the marine–coastal ambience. The initial pH = 7.94 of the SME shifted to more alkaline pH ≈ 8.54, suggesting the corrosion of the Mg matrix through the coating defects. In the meantime, the release of Mg 2 + from the PS–ZrO 2 -coated alloy surfaces was reduced by ≈90% compared to that of non-coated. Localized pitting attacks occurred in the vicinity of Al–Mn and β–Mg17Al12 cathodic particles characteristic of the Mg matrix. The depth of penetration (≈23 µm) was reduced by ≈85% compared to that of non-coated substrates. The protective effect against Cl ions, attributed to the hybrid PS–ZrO2-coated AM60 and AM60–AlN surfaces, was confirmed by the increase in their polarization resistance (Rp) in 37% and 22%, respectively, calculated from EIS data.

1. Introduction

Currently, there is a high demand for lightweight materials for the industrial manufacture of components for automobiles, airplanes, and other vehicles of transport, motivated by the needed reduction in fuel consumption and decrease in the emission of gases ( CO 2 and NO x ) that are harmful to human health and climate change [1,2]. Studies have reported that it can stop generating emissions between 4 and 12 g/km per each 100 kg of weight reduction [3,4]. In this aspect, magnesium (Mg) and its alloys may offer solutions to increase the efficiency of vehicles, reducing their weight and emission of the pollutants generated [2]. As structural materials, they have been present in the automotive industry as several interior components such as steering wheels, pedals, and seats; structural components such as interior doors and instrument panels; and chassis components such as wheels and suspension arms, among others [5]. Although Mg and its alloys have great potential for the transportation sector, they are susceptible to localized corrosion in the presence of impurities or corrosion-active intermetallic particles in the Mg matrix [6]. In the AZ (Mg–Zn–Al) and AM (Mg–Al) alloy series, used in the automotive industry, the secondary phase of β–Mg17Al12 and that of Al–Mn intermetallic particles are the most common having anodic or cathodic activity. To face this problem, the incorporation of additional alloying elements and nano-reinforcement particles in the Mg matrix has allowed improvements in the corrosion resistance and mechanical properties. In this aspect, AlN nanoparticles of 1 wt.% and an average diameter of ≈80 nm have been added to the AM60 matrix as reinforcement as an excellent choice for grain refinement benefiting its ductility [7,8,9] and lower roughness (≈15%) of the surface. The properties of the manufactured AM60–AlN nanocomposite have been previously described [10,11,12].
The initial stages of electrochemical corrosion activity of AM60 alloy and AM60–AlN nanocomposite have been compared during their exposure to solutions, which simulated the formation of an aqueous layer at the metal surface at 100% air humidity of industrial acid rain (SAR) and marine–coastal (SME) aggressive environments [13,14]. The AlN nanoparticles have been observed as “attached”, forming clusters to those of Al–Mn intermetallic particles, the local efficient cathodes [15,16,17,18], which subsisted on the Mg matrix after the removal of corrosion layers, inducing localized corrosion in their vicinity. During the exposure of the AM60 alloy and the AM60–AlN nanocomposite, the pH of the model solutions shifted to alkaline values (>9), and besides the release of Mg ions, de-alloying of Al was suggested because of the instability of AlMn [18] and AlN [19], which is attributed to the formation of Al ( OH ) 3 corrosion products, confirmed by XPS analysis [14].
Consequently, the presence of Cl ions led to stronger corrosion in both area and depth penetration on the nanocomposite surface of AM60–AlN during its exposure to the SME–marine environment [14]; however, in the acid rain industrial (SAR) ambience [13], a dense and more protective corrosion layer was formed on the AM60–AlN nanocomposite. In both environments, the corrosion process was considered weakly persistent and localized in time, dominated by the fractional Gaussian noise (fGn) according to the power spectral density of free corrosion current fluctuations, and classified as electrochemical noise. The reported results recommended that the surfaces of AM60 and AM60–AlN need a posterior modification to improve their corrosion resistance to chloride ions attacks, characteristics of the marine environment.
A promising method for increasing the corrosion resistance of Mg alloys is the application of coatings on their surfaces [1,20], which may generate a physical barrier against aggressive corrosive substances present in the environment, diminishing abrasion damages, in addition to esthetic functions [21,22]. Chrome-free surface treatments and non-chromate conversion coating have been proposed for corrosion protection of magnesium and Mg alloys [23,24,25], as well as organic coatings [21,26], superhydrophobic [27,28], and organic–inorganic hybrid coating [29,30,31].
The hybrid coatings elaborated through the sol–gel methodology [32,33,34] have offered advantages because, at low temperatures, the process controls the organic and inorganic composition of the coating and reaches a high level of purity [35]. A variety of hybrid organic–inorganic materials based on polymethylmethacrylate (PPMA) with various metal oxides ( SiO 2 ,   TiO 2 ,   and   ZrO 2 ) have been proposed [36,37,38,39,40,41,42,43,44,45]. The polystyrene (PS) polymer has participated as the organic part in combination with SiO 2 , ZrO 2 , Al 2 O 3 , MnO 2 , and TiO 2 in hybrid composites, which have been applied as dielectric materials and in searching for surface hydrophobicity or for better optoelectronic properties [46,47,48,49,50,51,52,53].
The organic and inorganic components may present a certain level of incompatibility between them, and to face this problem, coupling agents have been used [34,35]. After polymerization and polycondensation of the organic and inorganic phases, these components are linked through molecular coupling [54]. For example, 3-(Trimethoxysilyl) propyl methacrylate (TMSPM) is the coupling agent commonly used for the formation of hybrid materials [34,35,45]. The TMSPM allows the coupling through the silane groups, with which the inorganic phase is attached, while the organic phase is attached with acrylate as a coupling agent [48].
In order to improve the corrosion resistance of AM–magnesium alloys against the presence of chlorides, in this research sol–gel method was applied for the synthesis of the polystyrene–zirconium dioxide ( PS–ZrO 2 ) . The hybrid material was deposited by spin-coating on the AM60 alloy and AM60–AlN nanocomposite metallic substrates, which were exposed to a simulated marine environment solution (SME). The hydrophobicity property of the coating and its roughness were evaluated. Immersion tests were performed to monitor the changes in time of SME pH and concentration of Mg ion release. The hybrid coating PS–ZrO 2 surface morphology and composition, as well as their change after the exposure to SME, were performed by scanning electron microscopy and energy-dispersive X-ray spectroscopy (SEM-EDS). X-ray photoelectron spectroscopy (XPS) and X-ray diffraction (XRD) were used to characterize the hybrid coating deposited on the alloy substrates. Electrochemical impedance spectroscopy (EIS) diagrams were acquired to characterize the interface of the hybrid coating–alloy–electrolyte (SME solution) on which the corrosion process occurs.

2. Materials and Methods

2.1. PS–ZrO2 Hybrid Synthesis

ZrO 2 has attractive properties, such as thermal and chemical stability, high mechanical and abrasion resistance, low thermal conductivity, and low toxicity, as well as providing corrosion protection of metal substrates [55,56,57,58]. The methodology for the synthesis of PS–ZrO 2 hybrid material by means of the sol–gel method used in this research has been previously described [48] and has been similar to other hybrid material systems studied as PMMA–SiO 2 , PMMA–TiO 2 , and PMMA–ZrO 2 [34,36,37,38,39,42,59].
In this study, zirconium isopropoxide ( Zr ( OPr ) 4 ) and styrene monomer (ST) were used as the inorganic and organic precursors, and 3-(trimethoxysilyl)propyl methacrylate (TMSPM) was used as the coupling agent. Anhydrous ethanol (EtOH) and nitric acid were employed as solvent and catalyst, respectively, with a molar relation of 1:30:1 ( Zr ( OPr ) 4 : EtOH : HNO 3 ) for the preparation of the inorganic component (Solution 1). NaOH was used to remove the 4-tert-butylcatechol (4-TBC), which acts as a polymerization inhibitor in the styrene monomer using a molar relation of 1:0.11 (ST:OH), and then it was filtered. Benzoyl peroxide (BPO) was added to this solution with a ratio of 1:0.0006 (ST:BPO) as a polymerization initiator for the preparation of the organic component (Solution 2). EtOH and deionized water were added to TMSPM with a molar relation of 1:1:6 (TMSPM:EtOH: H 2 O ). Hydrochloric acid was incorporated, obtaining a homogeneous solution due to the hydrolysis of the coupling agent (Solution 3). The reagents for the synthesis of PS–ZrO 2 are summarized in Table 1. The three resulting solutions were mixed to obtain a homogeneous hybrid solution.

2.2. Deposition of PS–ZrO2 Hybrid on AM60 and AM60–AlN Alloy Surfaces

The hybrid solution was stored for 24 h, leaving it to age before PS–ZrO 2 was deposited on the metal samples. The AM60–AlN nanocomposite and AM60 alloy used as substrates were provided in the form round bar, its nominal composition, according to the producer (Magontec, Bottrop, Germany), in weight percent, is 6.0 Al; 0.2–0.4 Mn and the remainder being Mg. The manufacturing and incorporation of aluminum nitride nanoparticles (AlN, 1.0 wt.%, average diameter of 80 nm) in the AM60 for the formation of the AM60–AlN nanocomposite has been reported previously [11,13,60]. The surface of substrates (10 mm in diameter and thickness of 2 mm) was polished (to 2000 grain size of silicon carbide), sonicated in ethanol, and dried at room temperature.
Hybrid material coatings were performed by spin-coating using 0.5 mL of PS–ZrO 2 solution with a speed of 3000 rpm for 30 s. Afterward, the coated samples were taken to a vacuum drying oven (ADP-200C, Yamato Scientific Company Ltd., Tokyo, Japan), heated to 200   C for a period of 1 h, and stored in a desiccator to prevent the corrosion of the surfaces.

2.3. Roughness and Wettability of PS–ZrO2 Hybrid Coating on Mg–Al Alloy Surfaces

The roughness of the coated and non-coated alloy samples was measured using a 3D optical profilometer (Contour GT-K 3D, Bruker, Madison, WI, USA), and their surface wettability was determined by the contact angle (CA) of deionized water drops (with a volume of 1 μL) in contact with the alloy surfaces, measured after 1 min with the goniometer equipment (VCA-optima, AST Products Inc., Billerica, MA, USA), according to the sessile drop method at room temperature. Recorded images were obtained by means of a camera installed in the goniometer and positioned on the tested surface.

2.4. Immersion Test

The samples of AM60–AlN nanocomposite and the AM60 alloy coated with PS–ZrO 2 hybrid were immersed in 20 mL of simulated marine environment (SME) solution (Table 2), according to the ASTMG31-12a standard [61], for a period of up to 30 days. The change in time of SME pH solution and the concentration of the released Mg 2 + ions into the solution (Hanna Instruments, HI83200, Woonsocket, RI, USA) were measured.

2.5. SEM-EDS, XPS, and XRD Surface Analysis

The morphology and composition of the PS–ZrO 2 -hybrid-coated AM60 and AM60–AlN samples were analyzed before and after being immersed in the SME solution with a scanning electron microscope and energy dispersive spectroscopy (SEM-EDS, XL-30 ESEM-JEOL JSM-7600F, JEOL Ltd., Tokyo, Japan). Additional information was provided by X-ray photoelectron spectroscopy (XPS, K-Alpha Surface Analyzer, Thermo Scientific, Waltham, MA, USA), in which spectra binding energies were normalized to C1s carbon peak at 284.8 eV. X-ray diffraction patterns (Siemens D-500, Siemens D-5000, Munich, Germany; 2θ, 34 kV/25 mA CuKα) were used to determine possible crystal structures in the hybrid material.

2.6. Electrochemical Test

The electrochemical corrosion activity of the PS–ZrO 2 -hybrid-coated AM60 and AM60–AlN samples was through electrochemical impedance spectroscopy (EIS) during the sample’s immersion for 15 days in the solution SME at 21 °C. The conventional three-electrode cell (inside a Faraday cage) of working electrode (tested sample of 0 . 78   cm 2 of area), auxiliary electrode of Pt mesh, and saturated calomel as a reference electrode (SCE, Gamry Instruments, Philadelphia, PA, USA), was connected to the potentiostat/galvanostat/ZRA (Gamry Instruments, Interface-1000E, Philadelphia, PA, USA). The EIS diagrams were potentiostat/galvanostat/ZRA (Gamry Instruments, Interface-1000E, Philadelphia, PA, USA). The EIS diagrams were collected at a perturbation amplitude of ±10 mV vs. OCP (after 1 h of stabilization) at frequencies from 100 kHz to 10 mHz. The EIS spectra were analyzed with the Gamry Echem Analyst software (Gamry Instruments, Philadelphia, PA, USA).

3. Results and Discussion

3.1. Surface Characterization PS–ZrO2 Hybrid Coating

3.1.1. SEM-EDS Analysis

Figure 1 presents the homogeneous morphology of the spin-coated PS–ZrO 2 , well dispersed on the Mg–Al alloy surfaces of AM60–AlN (Figure 1a,c,e) and AM60 (Figure 1b,d,f). Table 3 summarizes the EDS elemental analysis (wt.%) of several randomly selected areas, labeled as “zones” in Figure 1, in which the elemental content is mainly the same.
A magnification (×20,000) of zones (Figure 1c,d) suggested that there is air, probably trapped inside the micro/nanostructures of the coating, which could reduce the contact between aggressive substances and alloy surface in the air sites, providing corrosion resistance to the alloy [62].
Despite the good dispersion of the PS–ZrO 2 coating on the studied Mg–Al alloy surfaces, some isolated areas of micro-defects have been observed (Figure 1e,f) that could generate channels connecting the SME solution with the alloy substrate and facilitate the passage of Cl , for example. EDS analysis revealed a high C content (Table 3), corresponding to the organic polystyrene ( PS ,   ( C 8 H 8 ) n ) , as a part of the hybrid coating, and a high content of O and Zr, confirming the presence of ZrO 2 , the inorganic part of the coating. The presence of silicon (Si) in all zones was ascribed to the coupling agent (3-(trimethoxysilyl) propyl methacrylate), which acted as a link between the organic and inorganic components of the hybrid PS–ZrO 2 coating. The decrease in Mg and Al contents (not corresponding to those of the Mg–Al tested alloys) was due to the physical barrier provided by the hybrid PS–ZrO 2 coating.
The C, Si, O, and Zr mappings (Figure 2) show the distribution of these elements and their contribution along the coated alloy surfaces of PS–ZrO2–AM60–AlN (Figure 2a) and PS–ZrO2–AM60 (Figure 2b). The reported AlMn and β–Mg17Al12 particles [13,14] were detected under the coating.
The cross-sectional SEM images (Figure 3) showed three well-defined zones corresponding to the epoxy resin (upper zone), hybrid coating of PS–ZrO2 (central zone), and Mg matrix (lower zone).
The thickness of the PS–ZrO 2 coatings on the AM60–AlN surface was ≈ 280   ±   25   nm , while on AM 60 , it was ≈ 330   ±   16   nm ; the difference was ascribed to the roughness difference in the studied alloy surfaces. The EDS element mappings, carried out in the yellow marked zones (Figure 3a,b), indicated that the upper zones present a high carbon content, corresponding to the epoxy resin, while in the central region, there is a set of elements, such as C, O, and Zr presenting the organic and inorganic components of the hybrid coating; in the lower zone, the high contents of Mg and Al correspond to the Mg–Al matrix of the AM60 and AM60–AlN alloys.

3.1.2. X-ray Photoelectron Spectroscopy (XPS)

The XPS spectra of the PS–ZrO 2 hybrid coating deposited on the AM60–AlN nanocomposite and AM60 alloy substrates were similar (Figure 4), and they were analyzed based on the binding energies of C1s, Si2p, Zr3d, and O1s. The C1s signal revealed the contribution of three peaks, ascribed to several bonds: C–C and C–H (at 284.8 eV) of hydrocarbon and phenyl groups, characteristic of the polystyrene [63]; C–O–C (at 286.0 eV) of the ether groups; and O–C=O (288.90 eV) of the double ester group, belonging to TMSPM coupling agent [29,64]. The binding energy of Si2p (at 102.3 eV) was attributed to the Si–O bond of the coupling agent (TMSPM) [48,65]. The high-resolution spectrum of Zr3d presents a doublet of two spin–orbital components, Zr 3 d 5 / 2 and Zr 3 d 3 / 2 , whose binding energies are approximately 182.5 and 184.9 eV, respectively, where the shift indicates the presence of Zr 4 + species [66], considering that for Zr 0 + , the Zr 3 d 5 / 2 and Zr 3 d 3 / 2 components have binding energies of 178.7 and 181.1 eV, respectively [67]. The species of Zr 4 + have allowed the interaction of the inorganic ZrO 2 to form Zr–O–Zr and Zr–OH bonds as a consequence of the ionization of the species O 2 and OH (the peaks of O1s at 530.1 and 531.1 eV, respectively) [68,69]. The O1s binding energy at 531.8 eV was ascribed to the Si–O–Zr bonds [70] because of the abundance of Si–OH groups present in the silane coupling agent (TMSPM) after its hydrolysis; the energies located at 532.6 eV and 533.7 eV belong to the characteristic groups of ether and ester, respectively, as a part of the TMSPM [71]. The Mg1s binding energy at 1304.5 eV corresponds to the Mg–O bonds, by which the hybrid coating was attached to the Mg matrix [72].

3.1.3. XRD Analysis

Diffraction patterns shown in Figure 5 of the AM60–AlN and AM60 substrates, before and after being coated with the hybrid PS–ZrO 2 , did not reveal the presence of crystalline structure in the hybrid material; the characteristic peaks of Mg, Al–Mn, and β–Mg17Al12 have been previously detected and reported [13]. It has been suggested that if the ZrO 2 is present as an amorphous phase, it could reduce the sites for the diffusion of the Cl ions through such film and, thereby, improve the corrosion resistance of the metal substrate [73].

3.2. Surface Roughness and Contact Angle

The average roughness ( R a ) of the AM60–AlN nanocomposite and AM60 alloy surfaces, with and without the hybrid coating of PS–ZrO 2 , are compared in Figure 6. It has been reported that the introduction of aluminum nitride (AlN) nanoparticles favored a reduction in grain size, which in fact, allowed a slight decrease in the roughness of the AM60 alloy of approximately 15% (Figure 6a,b) [13,14]. The hybrid coating of PS–ZrO 2 presented a 5% higher roughness value deposited on AM60 than on AM60–AlN because of the initial roughness difference. Such different roughness of the PS–ZrO 2 deposits has led to the presence of more or less trapped air inside the micro/nano-structured PS–ZrO 2 as sites of micro-defects (Figure 1e,f).
Hydrophobicity of any coating material is a property of wide interest when it is deposited on a metal surface as a protective material, thereby reducing contact with a humid and aggressive aqueous environment and, thus, improving the corrosion resistance of the metal substrate. Through the measurement of the contact angle ( θ c ) [74], the material may be classified as hydrophilic ( θ c < 90 ° ) , hydrophobic ( 90 ° < θ c < 150 ° ) , or superhydrophobic ( θ c > 150 ° ) [75,76,77,78,79].
Figure 7 presents the recorded images of the contact angle (CA) of deionized water drops on non-coated AM60–AlN nanocomposite and the AM60 alloy surfaces (Figure 7a,b) compared to those coated with the hybrid deposit of PS–ZrO 2 (Figure 7c,d).
The CA values revealed that the nature of the uncoated surfaces of the tested Mg–Al alloys is hydrophobic ( θ CA   >   90 ° ) [74]: the AM60–AlN nanocomposite presented a contact angle of 111 . 76 ± 2 . 93 ° (Figure 7a), whereas 110 . 70 ± 1 . 70 ° was obtained for the AM60 alloy (Figure 7b). However, after the deposit of the PS–ZrO 2 hybrid coating, the wettability of the surfaces changed, presenting a reduction in the contact angle values: for the PS–ZrO2–AM60–AlN, the CA = 83 . 37   ± 0 . 86 ° (Figure 7c), which was close to the hydrophobic border ( θ CA   <   90 ° ) ; and for PS–ZrO 2 –AM60, the CA value was 93 . 60   ± 1 . 87 ° (Figure 7d), with a wettability still in the hydrophobic range. The change in the contact angle of the coated surfaces may attribute to the increase in the PS–ZrO 2 surface roughness (Figure 7c,d), according to the suggestions of Wenzel [80] and Caxie–Baxter [81,82]. In the presence of air trapped on the surface ( PS–ZrO 2 nonuniform surface), the contact with liquids will be interrupted; however, it was a reduction in the CA, associated with the hydroxyl groups of Zr–OH and Si–OH present in the inorganic components of the hybrid coating, due to their incomplete condensation [48,83], which led to a decrease in the benefits that the microstructure of the hybrid material could provide.

3.3. Solution Monitoring

The change in time of SME solution pH was monitored for a period of 30 days during the immersion of the hybrid-coated Mg–Al alloy samples (Figure 8). The initial value of pH = 7.94 shifted to a more alkaline value of pH ≈ 8.64 after 7 days because the SME solution caused corrosion (Reactions 1–3) of the Mg matrix, which occurred in those sites where the hybrid material of PS–ZrO 2 presented some micro-defects (Figure 1e,f) and the H 2 bubbles (Reaction 3) continued to come out. The observed pH behavior has been reported previously during the exposure of AM60 and AM60–AlN to SME solution [14]. After this period of 7 days, the pH diminished, and this fact was associated with the formation of Mg ( OH ) 2 , an insoluble corrosion product, obstructing those micro-cracks and the initially formed micro-defects, which area acted as a physical barrier, hindering the progress of the corrosion process of the Mg matrix. However, in those sites, the insoluble Mg ( OH ) 2 product may suffer a localized attack from the chloride ions (SME solution) and be partially dissolved, giving the origin of released Mg 2 + ions (Reaction 4) and activating the corrosion process an increase in Mg2+ concentration and pH after 10 days.
Mg ( s ) Mg ( ac ) 2 + + 2 e
Mg 2 + + OH Mg ( OH ) 2
2 H 2 O + 2 e H 2 ( g ) + 2 OH ( ac )
Mg ( OH ) 2 + 2 Cl MgCl 2 + 2 OH Mg 2 + + 2 Cl
Figure 9 compares the concentrations of Mg 2 + ions released into the SME solution because of the progress in Reaction (4) during the exposure of PS–ZrO2–AM60–AlN and PS–ZrO2–AM60 alloys, non-coated and coated with the hybrid PS–ZrO 2 for 30 days to SME solution. From the non-coated alloy surface of AM60, 333.33   ±   11 . 54   mg   L 1 of Mg 2 + ions were released, while from the AM60–AlN nanocomposite, the concentration was 353 . 33   ±   11 . 54   mg   L 1 [14] because of the shift of pH to more alkaline values, which led to an instability of the AlN particles [19]. The progress in the release of the Mg ions has been associated with the presence of Al–Mn intermetallic particles, active cathodic sites in AM60 allow and AM60–AlN nanocomposite, in which vicinity the Mg matrix (active anode) suffers accelerated localized corrosion attack [13,14,15,16,17,18]. The results indicated that the hybrid PS–ZrO 2 coating deposited on AM60–AlN and AM60 surfaces reduced the release of Mg 2 + by approximately 89% and 91%, respectively, because of the obstructed micro-localized sites of defects and impaired diffusion of chloride ions through the formed layer of corrosion product.

3.4. SEM-EDS Analysis of the PS–ZrO2-Coated Alloy Surfaces after Exposure to SME Solution

The SEM images in Figure 10 illustrate the morphology of the AM60–AlN and AM60 magnesium alloys’ surfaces coated with the PS–ZrO 2 after their exposure for 30 days to SME marine–coastal simulate solution. The micro-cracks that appeared on the PS–ZrO 2 hybrid coating may be considered a consequence of the exerted pressure by H 2 bubbles during the localized corrosion of the Mg matrix, which had the possibility to occur because of the initial micro-defects on the coating surface (Figure 1e,f). Once formed, these new micro-cracks may favor the progress of Mg alloy corrosion activity.
The EDS elemental analysis (Table 4) revealed two typical zones labeled as “1” and “2” (Figure 10c,d). Zone “1” corresponds to the Al–Mn intermetallic particles and cathodic active, characteristics of the AM60 Mg alloy matrix, which have been stable and resistive against attacks by chloride ions and changes in the pH of the SME solution. This fact allows us to confirm the cathodic activity of AlMn, previously reported [15,16,17]. Zone “2” presented the composition of the hybrid layer of PS–ZrO 2 deposited on the Mg alloys after immersion for 30 days in the SME solution (Table 4).
The elemental mapping of PS–ZrO 2 -coated AM60–AlN nanocomposite and AM60 alloy (Figure 11) after 30 days of exposure in SME solution confirmed the presence of elements characteristics of intermetallic cathodic particles of Al–Mn and the attached to them those of the ZrO 2 , as well as the presence of β–Mg17Al12 cathodic particles, which resisted to the corrosion attacks. It has been reported [84] that the incorporation of ZrO 2 in an organic coating has managed to reduce the advance of corrosive species ( Cl ions) towards the substrate because ZrO 2 (an inorganic element) allowed the formation of a more complete network, obstructing the diffusion of species through the coating. This behavior may be related to the good chemical stability of ZrO 2 [85,86].
After the chemical removal of the surface layers, according to the ASTM G1-03 [87] formed during the exposure of the coated and non-coated Mg alloys for 30 days to the SME model solution, the SEM images in Figure 12 show the surface appearance and the EDS analysis of the zone of interest is resumed in Table 5. Zone 1 suggested the presence of the cathodic particles of Al–Mn, while Zone 3 indicates the particles of β–Mg17Al12, which were not attacked by the corrosion process. Zone 2 presents the Mg matrix. The low contents of C and Si have been parts of the organic material of the PS and the coupling agent of TMSPM. The localized attacks were more intensive in the non-coated AM60–AlN nanocomposite and AM60 alloy.
Additional mapping (Figure 13) confirmed the presence of those characteristic particles, as suggested above (Table 5).
Figure 14 groups several SEM images of the cross-sections of the surfaces of the AM60–AlN (Figure 14a) and AM 60 (Figure 14b) magnesium alloys coated with PS–ZrO 2 deposits. The visualized depths of the localized attack towards the matrix have average values of depth of penetration of ≈28.08 µm on the AM60–AlN surface and ≈17.90 µm on the AM 60 . In the absence of the studied hybrid deposit, the average penetration depths were ≈175.70 µm and ≈121.40 µm, respectively, reaching a maximum in the nanocomposite AM60–AlN of ≈246.40 µm and ≈178.00 µm on the AM60 surfaces after the exposure for 30 days to SME model solution (Figure 14c,d) [14]. The comparison of these results allowed us to consider a reduction in the localized attack by ≈85% due to the protective effect of the PS–ZrO 2 deposit against the chloride attack of the marine–coastal environment (SME model solution).
However, because of the available defects on the hybrid coating surface (Figure 1e,f), the chloride ions (SME solution) and oxygen diffusion processes were facilitated, and they were able to penetrate through the hybrid material and attack the Mg matrix. On the other hand, the ZrO 2 , a good semiconductor [88], could serve as local cathodes, which were not attacked during the corrosion process and were maintained on the Mg alloys’ surfaces after the removal of the corrosion layers.

3.5. Electrochemical Impedance Spectroscopy

The electrochemical impedance, visualized using Bode and Nyquist diagrams (EIS, a non-destructive technique), was elaborated to characterize the interface of the hybrid-coated Mg alloys after their exposure for 1 day and 15 days to SEM marine environment model solution (Figure 15). The Nyquist diagrams (Figure 15a,b) revealed two capacitive semi-circles associated with two time constants at higher and medium frequencies (HF and MF), respectively. The diameters of the HF-capacitive loops were associated with the particularities of the formed corrosion layer on the Mg matrix in the presence of the hybrid PS–ZrO 2 coating, while the MF-capacitive loops may relate to the charge transfer processes of the hydrogen evolution ( H 2 ) and the Mg 2 + released through the double layer [89,90,91]. The Bode plots (Figure 15) of the phase angle were found to be in good agreement with the observed changes in the Nyquist diagrams. The phase angle between 70 ° and 80 ° confirmed that the interfaces of the studied alloys are capable of accumulating electrical charges, which in fact, will complicate the mass transfer process through the electrode interface and, consequently, the progress in the corrosion process because of the hybrid PS–ZrO2 coating.
The quantification of the EIS data, which characterized the activity of the coated Mg alloys, was carried out according to the equivalent circuit (EC) present in Figure 16 and the values are summarized in Table 6, and they were compared to those of non-coated surfaces (Table 7). The EC includes the following components: Rs is the solution resistance; R 1 denotes the resistance of the layer on the metal substrate, and the constant phase element CPE 1 denotes the “capacitance”, representing the hybrid coating and the formed corrosion layer later with the progress in the corrosion process; and R 2 and CPE 2 as “capacitance” are characteristic of the charge transfer process at the coated substrate/electrolyte interface [90,92,93]. The values of Rp (polarization resistance) were calculated as Rp = R 1 + R 2 .
The comparison of the R p values (Table 6) allowed us to consider that the hybrid coating of PS–ZrO 2 deposited on the Mg alloy AM60 surface can be attributed to greater resistance against the corrosion process, presenting an increase in its Rp by 37% (at 15 days in SME model solution), compared to that of 22% for the coated composite AM60–AlN. In the case of non-coated AM surfaces, the Rp of AM60 was 9% higher than that of the AM60–AlN (Table 7). These facts were related to the AlN hydroxide phase’s solubility, which raises the pH in the range of 5.5–12; during the exposure of the coated AM alloys, the pH of the SME model solution was ≈8.5 (Figure 8). In the presence of chloride ions, it has been reported that the AlN may transform into Al ( OH ) 3 [19].

4. Conclusions

  • A hybrid coating of polystyrene (PS)– ZrO 2 material was developed by the sol–gel technique and deposited by spin-coating method on AM60 and nanocomposite AM60–AlN magnesium alloy surfaces to enhance the corrosion resistance in marine environments.
  • The PS–ZrO 2 coating was dispersed homogeneously on the alloy substrates, presenting isolated micro–nano-structure defects with air trapped inside, which led to an increase in roughness of ≈4 times. The average thickness of the hybrid coating was ≈ 305   ±   20   nm . The XRD patterns revealed no crystalline structure of the hybrid organic–inorganic coating. The deposit of PS–ZrO 2 reduced the contact angle of the Mg substrates, and their wettability was close to the hydrophobic border (θCA 90°–94°), associated with the hydroxyl groups of Zr–OH and Si–OH incomplete condensation.
  • During the exposure of the hybrid-coated substrates for 30 days to SME solution, simulating marine–coastal environment, the initial value of pH = 7.94 shifted to a more alkaline pH ≈ 8.54 because the SME solution caused corrosion of the Mg matrix, which occurred in those sites where the hybrid material of PS–ZrO 2 presented some micro-defects and the H 2 bubbles continued to come out.
  • The results indicated that the hybrid PS–ZrO 2 coating reduced the release of Mg 2 + by approximately 90% and 91% compared to that of non-coated AM magnesium alloy substrates, because of the obstructed micro-localized defects by corrosion products, which impaired the diffusion of chloride ions through the Mg matrix.
  • After the chemical removal of the surface layers formed during the exposure to SME solution, the SEM images showed that the localized pitting attack occurred in the vicinity of the Al–Mn and β–Mg17Al12 intermetallic cathodic particles, suggested by EDS analysis.
  • Cross-section images revealed that the average value of depth of penetration (≈23 µm) was reduced by ≈85% compared to that of non-coated substrates due to the protective effect of the PS–ZrO 2 hybrid coating on AM magnesium alloy substrates exposed to marine–coastal simulated ambient (SME).
  • The polarization values of R p calculated from EIS indicated that the R p of the PS–ZrO 2 -coated AM60 alloy increased by 37% and that of the composite AM60–AlN increased by 22%; these values were considered as a protection gain against the corrosion in the presence of chloride ions.
  • The corrosion protection efficiency of the hybrid PS–ZrO 2 against the presence of chlorides should be improved by modifying the concentration of the precursors and/or applying a drying process that uses a temperature program ramp.

Author Contributions

Conceptualization and methodology, L.V. and L.C. performed the preparation of samples and the corrosion tests; L.C., D.S.-A. and R.R.-B. contributed to the synthesis and deposit of the hybrid material; L.C. and L.V. performed the formal analysis of the results and the writing of the original draft and editing. L.V. supervised the project. All correspondence should be addressed to L.V. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data are available upon request from the corresponding author.

Acknowledgments

Luis Chávez acknowledges the Mexican National Council for Science and Technology (CONACYT) for the scholarship for his Ph.D. study. The authors gratefully thank the National Laboratory of Nano and Biomaterials (LANNBIO-CINVESTAV) for allowing the use of SEM-EDS and XPS facilities; thanks also go to Victor Rejón, Daniel Aguilar, and Willian Cauich for their support in data acquisition. The authors thank Carlos Ávila and Agustin Galindo for their technical assistance and for allowing access to the Research Laboratory and Technological Development in Advanced Coating (LIDTRA-CINVESTAV Queretaro).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Predko, P.; Rajnovic, D.; Grilli, M.L.; Postolnyi, B.O.; Zemcenkovs, V.; Rijkuris, G.; Pole, E.; Lisnanskis, M. Promising methods for corrosion protection of magnesium alloys in the case of Mg-Al, Mg-Mn-Ce and Mg-Zn-Zr: A recent progress review. Metals 2021, 11, 1133. [Google Scholar] [CrossRef]
  2. Ehrenberger, S.; Dieringa, H.; Friedrich, H.E. Life Cycle Assessment of Magnesium Components in Vehicle Construction; German Aerospace Center: Bremen, Germany, 2013. [Google Scholar]
  3. Mohrbacher, H. High-Performance Steels for Sustainable Manufacturing of Vehicles. In Green and Sustainable Manufacturing of Advanced Material; Singh, M., Ohji, T., Asthana, R., Eds.; Elsevier: Amsterdam, The Netherlands, 2015; pp. 135–163. [Google Scholar]
  4. Helms, H. Fuel Saving by Light-Weighting for European Articulated Trucks; Institute for Energy and Environmental Research: Heidelberg, Germany, 2005; Available online: http://www.alcoa.com/global/en/environment/pdf/fuel_saving_by_lightweighting.pdf (accessed on 12 May 2023).
  5. Schumann, S.; Friedrich, H.E. Engineering Requirements, Strategies and Examples. In Magnesium Technology: Metallurgy, Design Data, Applications; Friedrich, H.E., Mordike, B.L., Eds.; Springer: Berlin/Heidelberg, Germany, 2006; pp. 499–632. [Google Scholar]
  6. Zeng, R.-C.; Zhang, J.; Huang, W.-J.; Dietzel, W.; Kainer, K.U.; Blawert, C.; Ke, W. Review of Studies on Corrosion of Magnesium Alloys. Trans. Nonferrous Met. Soc. China 2006, 16, s763–s771. [Google Scholar] [CrossRef]
  7. Somekawa, H.; Mukai, T. Effect of grain refinement on fracture toughness in extruded pure magnesium. Scr. Mater. 2005, 53, 1059–1064. [Google Scholar] [CrossRef]
  8. Kang, S.-H.; Lee, Y.S.; Lee, J.H. Effect of grain refinement of magnesium alloy AZ31 by severe plastic deformation on material characteristics. J. Mater. Process. Technol. 2008, 201, 436–440. [Google Scholar] [CrossRef]
  9. Mukai, T.; Yamanoi, M.; Watanabe, H.; Ishikawa, K.; Higashi, K. Effect of grain refinement on tensile ductility in ZK60 magnesium alloy under dynamic loading. Mater. Trans. 2001, 42, 1177–1181. [Google Scholar] [CrossRef]
  10. Saboori, A.; Padovano, E.; Pavese, M.; Badini, C. Novel magnesium Elektron21-AlN nanocomposites produced by ultrasound-assisted casting; microstructure, thermal and electrical conductivity. Materials 2017, 11, 27. [Google Scholar] [CrossRef]
  11. Lerner, M.I.; Glazkova, E.A.; Lozhkomoev, A.S.; Svarovskaya, N.V.; Bakina, O.V.; Pervikov, A.V.; Psakhie, S.G. Synthesis of Al nanoparticles and Al/AlN composite nanoparticles by electrical explosion of aluminum wires in argon and nitrogen. Powder Technol. 2016, 295, 307–314. [Google Scholar] [CrossRef]
  12. Lerner, M.; Vorozhtov, A.; Guseinov, S.; Storozhenko, P. Metal Nanopowders Production. In Metal Nanopowders: Production, Characterization, and Energetic Applications, 1st ed.; Gromov, A.A., Teipel, U., Eds.; Wiley-VCH: Hoboken, NJ, USA, 2014; pp. 79–106. [Google Scholar]
  13. Chávez, L.; Veleva, L.; Feliu, S., Jr.; Giannopoulou, D.; Dieringa, H. Corrosion Behavior of Extruded AM60-AlN Metal Matrix Nanocomposite and AM60 Alloy Exposed to Simulated Acid Rain Environment. Metals 2021, 11, 990. [Google Scholar] [CrossRef]
  14. Chávez, L.; Veleva, L.; Sánchez, G.; Dieringa, H. AM60-AlN Nanocomposite and AM60 Alloy Corrosion Activity in Simulated Marine-Coastal Ambience. Metals 2022, 12, 1997. [Google Scholar] [CrossRef]
  15. Lunder, O.; Nordien, J.H.; Nisancioglu, K. Corrosion resistance of cast Mg-Al alloys. Corros. Rev. 1997, 15, 439–470. [Google Scholar] [CrossRef]
  16. Davoodi, A.; Pan, J.; Leygraf, C.; Norgren, S. The role of intermetallic particles in localized corrosion of an aluminum alloy studied by SKPFM and integrated AFM/SECM. J. Electrochem. Soc. 2008, 155, C211–C218. [Google Scholar] [CrossRef]
  17. Pawar, S.; Zhou, X.; Thompson, G.E.; Scamans, G.; Fan, Z. The role of intermetallics on the corrosion initiation of twin roll cast AZ31 Mg alloy. J. Electrochem. Soc. 2015, 162, C442–C448. [Google Scholar] [CrossRef]
  18. Asmussen, R.M.; Binns, W.J.; Parfov-Nia, R.; Jakupi, P.; Shoesmith, D.W. The stability of aluminum-manganese intermetallic phases under the microgalvanic coupling conditions. Mater. Corros. 2016, 67, 39–50. [Google Scholar] [CrossRef]
  19. Svedberg, L.M.; Arndt, K.C.; Cima, M.J. Corrosion of aluminum nitride (AlN) in aqueous cleaning solutions. J. Am. Ceram. Soc. 2000, 83, 41–46. [Google Scholar] [CrossRef]
  20. Schwalm, R. Introduction to Coatings Technology. In UV Coatings: Basics, Recent Developments and New Applications; Schwalm, R., Ed.; Elsevier Science: Amsterdam, The Netherlands, 2006; pp. 1–18. [Google Scholar]
  21. Wang, J.; Pang, X.; Jahed, H. Surface protection of Mg alloys in automotive applications: A review. AIMS Mater. Sci. 2019, 6, 567–600. [Google Scholar] [CrossRef]
  22. Zhang, D.; Peng, F.; Liu, X. Protection of magnesium alloys: From physical barrier coating to smart self-healing coating. J. Alloys Compd. 2021, 853, 157010–157031. [Google Scholar] [CrossRef]
  23. Umehara, H.; Takaya, M.; Terauchi, S. Chrome-free surface treatments for magnesium alloy. Surf. Coat. Technol. 2003, 169, 666–669. [Google Scholar] [CrossRef]
  24. Gonzalez-Nunez, M.A.; Nunez-Lopez, C.A.; Skeldon, P.; Thompson, G.E.; Karimzadeh, H.; Lyon, P.; Wilks, T.E. A non-chromate conversion coating for magnesium alloys and magnesium-based metal matrix composites. Corros. Sci. 1995, 37, 1763–1772. [Google Scholar] [CrossRef]
  25. Gray, J.; Luan, B. Protective coatings on magnesium and its alloys—A critical review. J. Alloys Compd. 2002, 336, 88–113. [Google Scholar] [CrossRef]
  26. Hu, R.-G.; Zhang, S.; Bu, J.-F.; Lin, C.-J. Recent progress in corrosion protection of magnesium alloys by organic coatings. Prog. Org. Coat. 2012, 73, 129–141. [Google Scholar] [CrossRef]
  27. Yao, W.; Wu, L.; Huang, G.; Jiang, B.; Atrens, A.; Pan, F. Superhydrophobic coatings for corrosion protection of magnesium alloys. J. Mater. Sci. Technol. 2020, 52, 100–118. [Google Scholar] [CrossRef]
  28. Shao, W.; Kan, Q.; Bai, X.; Wang, C. Robust Superhydrophobic Coatings for Enhanced Corrosion Resistance and Dielectric Properties. Coatings 2022, 12, 1655. [Google Scholar] [CrossRef]
  29. Dos Santos, F.C.; Harb, S.V.; Menu, M.-J.; Turq, V.; Pulcinelli, S.H.; Santilli, C.V.; Hammer, P. On the structure of high performance anticorrosive PMMA–siloxane–silica hybrid coatings. RSC Adv. 2015, 5, 106754–106763. [Google Scholar] [CrossRef]
  30. Taghavikish, M.; Surya, S.; Dutta, N.K. Novel thiol-ene hybrid coating for metal protection. Coatings 2016, 6, 17. [Google Scholar] [CrossRef]
  31. Malucelli, G. Hybrid organic/inorganic coatings through dual-cure processes: State of the art and perspectives. Coatings 2016, 6, 10. [Google Scholar] [CrossRef]
  32. Al-Kandary, S.; Ali, A.A.M.; Ahmad, Z. New polyimide-silica nano-composites from the sol-gel process using organically-modified silica network structure. J. Mater. Sci. 2006, 41, 2907–2914. [Google Scholar] [CrossRef]
  33. Rubio, E.; Almaral, J.; Ramírez-Bon, R.; Castaño, V.; Rodríguez, V. Organic–inorganic hybrid coating (poly (methyl methacrylate)/monodisperse silica). Opt. Mater. 2005, 27, 1266–1269. [Google Scholar] [CrossRef]
  34. Morales-Acosta, M.D.; Alvarado-Beltrán, C.G.; Quevedo-López, M.A.; Gnade, B.E.; Mendonza-Galván, A.; Ramírez-Bon, R. Adjustable structural, optical and dielectric characteristics in sol–gel PMMA–SiO2 hybrid films. J. Non-Cryst. Solids 2013, 362, 124–135. [Google Scholar] [CrossRef]
  35. Alvarado-Rivera, J.; Muñoz-Saldaña, J.; Ramírez-Bon, R. Nanoindentation testing of SiO2-PMMA hybrid films on acrylic substrates with variable coupling agent content. J. Sol-Gel Sci. Technol. 2010, 54, 312–318. [Google Scholar] [CrossRef]
  36. Martínez-Landeros, V.H.; Gnade, B.E.; Quevedo-López, M.A.; Ramírez-Bon, R. Permeation studies on transparent multiple hybrid SiO2-PMMA coatings-Al2O3 barriers on PEN substrates. J. Sol-Gel Sci. Technol. 2011, 59, 345–351. [Google Scholar] [CrossRef]
  37. Morales-Acosta, M.D.; Quevedo-López, M.A.; Ggade, B.E.; Ramírez-Bon, R. PMMA−SiO2 organic–inorganic hybrid films: Determination of dielectric characteristics. J. Sol-Gel Sci. Technol. 2011, 58, 218–224. [Google Scholar] [CrossRef]
  38. Almaral-Sánchez, J.L.; Rubio, E.; Mendoza-Galván, A.; Ramírez-Bon, R. Red colored transparent PMMA-SiO2 hybrid films. J. Phys. Chem. Solids 2005, 66, 1660–1667. [Google Scholar] [CrossRef]
  39. Alvarado-Beltrán, C.G.; Almaral-Sánchez, J.L.; Quevedo-López, M.A.; Ramírez-Bon, R. Dielectric Gate Applications of PMMA-TiO2 Hybrid Films in ZnO-Based Thin Film Transistors. Int. J. Electrochem. Sci. 2015, 10, 4068–4082. [Google Scholar]
  40. Ohlmaier-Delgadillo, F.; Castillo-Ortega, M.M.; Ramírez-Bon, R.; Armenta-Villegas, L.; Rodríguez-Félix, D.E.; Santacruz-Ortega, H.; Castillo-Castro, T.C.; Santos-Sauceda, I. Photocatalytic properties of PMMA-TiO2 class I and class II hybrid nanofibers obtained by electrospinning. J. Appl. Polym. Sci. 2016, 133, 44334–44342. [Google Scholar] [CrossRef]
  41. Aziz, N.A.A.; Achoi, M.F.; Abdullah, S.; Rusop, M. Structural and optical properties of nanohybrid PMMA/TiO2. Adv. Mater. Res. 2013, 667, 63–67. [Google Scholar] [CrossRef]
  42. Alvarado-Beltrán, C.G.; Almaral-Sánchez, J.L.; Mejia, I.; Quevedo-López, M.A.; Ramírez-Bon, R. Sol-gel PMMA-ZrO2 hybrid layers as gate dielectric for low-temperature ZnO-based thin-film transistors. ACS Omega 2017, 2, 6968–6974. [Google Scholar] [CrossRef]
  43. Hu, Y.; Gu, G.; Zhou, S.; Wu, L. Preparation and properties of transparent PMMA/ZrO2 nanocomposites using 2-hydroxyethyl methacrylate as a coupling agent. Polymer 2011, 52, 122–129. [Google Scholar] [CrossRef]
  44. Reyes-Acosta, M.A.; Torres-Huerta, A.M.; Domínguez-Crespo, M.A.; Flores-Vela, A.I.; Dorantes-Rosales, H.J.; Ramírez-Meneses, E. Influence of ZrO2 nanoparticles and thermal treatment on the properties of PMMA/ZrO2 hybrid coatings. J. Alloys Compd. 2015, 643, S150–S158. [Google Scholar] [CrossRef]
  45. Alvarado-Beltrán, C.G.; Almaral-Sánchez, J.L.; Ramírez-Bon, R. Synthesis and properties of PMMA-ZrO2 organic-inorganic hybrid films. J. Appl. Polym. Sci. 2015, 132, 42738–42744. [Google Scholar] [CrossRef]
  46. Mezan, S.O.; Jabbar, A.H.; Hamzah, M.Q.; Tuama, A.N.; Hasan, N.N.; Roslan, M.S.; Agam, M.A. Synthesis, characterization, and properties of polystyrene/SiO2 nanocomposite via sol-gel process. AIP Conf. Proc. 2019, 2151, 020034. [Google Scholar]
  47. Zhu, S.-Y.; Zhang, X.-M.; Chen, W.-X.; Feng, L.-F. Synthesis, characterization, and properties of polystyrene/SiO2 hybrid materials via sol–gel process. Polym. Compos. 2015, 36, 482–488. [Google Scholar] [CrossRef]
  48. Sánchez-Ahumada, D.; Verastica-Ward, L.J.; Orozco, M.; Vargas-Hernández, D.; Castro-Beltran, A.; Ramírez-Bon, R.; Alvarado-Beltrán, C.G. In-situ low-temperature synthesis of PS-ZrO2 hybrid films and their characterization for high-k gate dielectric application. Prog. Org. Coat. 2021, 154, 106188. [Google Scholar] [CrossRef]
  49. Bahgat Radwan, A.; Abdullah, A.M.; Mohamed, A.M.A. New electrospun polystyrene/Al2O3 nanocomposite superhydrophobic coatings; synthesis, characterization, and application. Coatings 2018, 8, 65. [Google Scholar] [CrossRef]
  50. Maul, J.; Frushour, B.G.; Kontoff, J.R.; Eichenauer, H.; Ott, K.-H.; Schade, C. Polystyrene and Styrene Copolymers. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH: Weinheim, Germany, 2007; pp. 477–484. [Google Scholar]
  51. Hou, W.; Wang, Q. UV-driven reversible switching of a polystyrene/titania nanocomposite coating between superhydrophobicity and superhydrophilicity. Langmuir 2009, 25, 6875–6879. [Google Scholar] [CrossRef] [PubMed]
  52. Xu, X.; Zhang, Z.; Guo, F.; Yang, J.; Zhu, X.; Zhou, X.; Xue, Q. Fabrication of bionic superhydrophobic manganese oxide/polystyrene nanocomposite coating. J. Bionic Eng. 2012, 9, 11–17. [Google Scholar] [CrossRef]
  53. Hou, W.; Wang, Q. Wetting behavior of a SiO2–polystyrene nanocomposite surface. J. Colloid Interface Sci. 2007, 316, 206–209. [Google Scholar] [CrossRef]
  54. Sánchez-Ahumada, D.; Verastica-Ward, L.J.; Gálvez-López, M.F.; Castro-Beltrán, A.; Ramírez-Bon, R.; Alvarado-Beltrán, C.G. Low-temperature synthesis and physical characteristics of PS-TiO2 hybrid films for transparent dielectric gate applications. Polymer 2019, 172, 170–177. [Google Scholar] [CrossRef]
  55. Tahmasebpour, M.; Babaluo, A.A.; Aghjeh, M.R. Synthesis of zirconia nanopowders from various zirconium salts via polyacrylamide gel method. J. Eur. Ceram. Soc. 2008, 28, 773–778. [Google Scholar] [CrossRef]
  56. Guo, G.-Y.; Chen, Y.-L. A nearly pure monoclinic nanocrystalline zirconia. J. Solid State Chem. 2005, 178, 1675–1682. [Google Scholar] [CrossRef]
  57. Pang, X.; Zhitomirsky, I.; Niewczas, M. Cathodic electrolytic deposition of zirconia films. Surf. Coat. Technol. 2005, 195, 138–146. [Google Scholar] [CrossRef]
  58. Wang, J.; Liu, X.; Ren, S.; Guan, F.; Yang, S. Mechanical properties and tribological behavior of ZrO2 thin films deposited on sulfonated self-assembled monolayer of 3-mercaptopropyl trimethoxysilane. Tribol. Lett. 2005, 18, 429–436. [Google Scholar] [CrossRef]
  59. Kumar, A.; Mondal, S.; Rao, K.K. Low temperature solution processed high-κ ZrO2 gate dielectrics for nanoelectonics. Appl. Surf. Sci. 2016, 370, 373–379. [Google Scholar] [CrossRef]
  60. Dieringa, H.; Katsarou, L.; Buzolin, R.; Szakácks, G.; Horstmann, M.; Wolff, M.; Mendis, C.; Vorozhtsov, S.; StJhon, D. Ultrasound assisted casting of an AM60 based metal matrix nanocomposite, its properties, and recyclability. Metals 2017, 7, 388. [Google Scholar] [CrossRef]
  61. ASTM-NACE/ASTM G31-12a; Standard Guide for Laboratory Immersion Corrosion Testing of Metals. ASTM International: West Conshohocken, PA, USA, 2021; pp. 1–9.
  62. Zhang, Y.; Cao, H.; Huang, H.; Wang, Z. Hydrophobic modification of magnesium hydroxide coating deposited cathodically on magnesium alloy and its corrosion protection. Coatings 2019, 9, 477. [Google Scholar] [CrossRef]
  63. Goodson, M.L.; Lagle, R.; Guggilla, P. X-Ray Photoelectron Spectroscopy of Polystyrene Composite Films. J. Adv. Mater. Sci. Eng. 2022, 2, 1–5. [Google Scholar] [CrossRef]
  64. Foti, L.; Sionek, A.; Stori, E.M.; Soares, P.P.; Pereira, M.M.; Krieger, M.A.; Petzhold, C.L.; Schreiner, W.H.; Soares, M.J.; Goldenberg, S.; et al. Electrospray induced surface activation of polystyrene microbeads for diagnostic applications. J. Mater. Chem B 2015, 3, 2725–2731. [Google Scholar] [CrossRef] [PubMed]
  65. Wang, Y.; Zhang, D.; Shi, L.; Li, L.; Zhang, J. Novel transparent ternary nanocomposite films of trialkoxysilane-capped poly (methyl methacrylate)/zirconia/titania with incorporating networks. Mater. Chem. Phys. 2008, 110, 463–470. [Google Scholar] [CrossRef]
  66. Tsunekawa, S.; Asami, K.; Ito, S.; Yashima, M.; Sugimoto, T. XPS study of the phase transition in pure zirconium oxide nanocrystallites. Appl. Surf. Sci. 2005, 252, 1651–1656. [Google Scholar] [CrossRef]
  67. Bespalov, I.; Datler, M.; Buhr, S.; Drachsel, W.; Rupprechter, G.; Suchorski, Y. Initial stages of oxide formation on the Zr surface at low oxygen pressure: An in situ FIM and XPS study. Ultramicroscopy 2015, 159, 147–151. [Google Scholar] [CrossRef]
  68. Dupin, J.-C.; Gonbeau, D.; Vinatier, P.; Levasseur, A. Systematic XPS studies of metal oxides, hydroxides and peroxides. Phys. Chem. Chem. Phys. 2000, 2, 1319–1324. [Google Scholar] [CrossRef]
  69. Li, K.; Guo, L.; Wang, Y.; Huang, J. Synthesis and thermal performance of polymer precursor for ZrC ceramic. Ceram. Int. 2021, 47, 28806–28810. [Google Scholar] [CrossRef]
  70. Lee, S.H.; Jeong, S.; Moon, J. Nanoparticle-dispersed high-k organic–inorganic hybrid dielectrics for organic thin-film transistors. Org. Electron. 2009, 10, 982–989. [Google Scholar] [CrossRef]
  71. López, G.P.; Castner, D.G.; Ratner, B.D. XPS O 1s binding energies for polymers containing hydroxyl, ether, ketone and ester groups. Surf. Interface Anal. 1991, 17, 267–272. [Google Scholar] [CrossRef]
  72. Moulder, J.F.; Stickle, W.F.; Sobol, P.E.; Bomben, K.D. Handbook of X-Ray Photoelectron Spectroscopy; Physical Electronics Inc.: Chanhassen, MN, USA, 1995. [Google Scholar]
  73. Stambolova, I.; Stoyanova, D.; Shipochka, M.; Boshkova, N.; Eliyas, A.; Simeonova, S.; Grozev, N.; Boshkov, N. Surface morphological and chemical features of anticorrosionZrO2-TiO2 coatings: Impact of zirconium precursor. Coatings 2021, 11, 703. [Google Scholar] [CrossRef]
  74. Ahmad, D.; van den Boogaert, I.; Miller, J.; Presswell, R.; Jouhara, H. Hydrophilic and hydrophobic materials and their applications. Energy Sources Part A Recovery Util. Environ. Eff. 2018, 40, 2686–2725. [Google Scholar] [CrossRef]
  75. Davaasuren, G.; Ngo, C.-V.; Oh, H.-S.; Chun, D.-M. Geometric study of transparent superhydrophobic surfaces of molded and grid patterned polydimethylsiloxane (PDMS). A. Surf. Sci. 2014, 314, 530–536. [Google Scholar] [CrossRef]
  76. Elkhyat, A.; Courderot-Masuyer, C.; Gharbi, T.; Humbert, P. Influence of the hydrophobic and hydrophilic characteristics of sliding and slider surfaces on friction coefficient: In vivo human skin friction comparison. Skin Res. Tech. 2004, 10, 215–221. [Google Scholar] [CrossRef]
  77. Huang, S.; Xu, J.; Liang, C.; Zhang, X. Size distribution measurement of packed tower drift based on hydrophobic materials. Appl. Therm. Eng. 2016, 99, 873–879. [Google Scholar] [CrossRef]
  78. Stanton, M.M.; Ducker, R.E.; MacDonald, J.C.; Lambert, C.R.; McGimpsey, W.G. Super-hydrophobic, highly adhesive, polydimethylsiloxane (PDMS) surfaces. J. Colloid Interface Sci. 2012, 367, 502–508. [Google Scholar] [CrossRef]
  79. Zhang, Y.; Sundararajan, S. Superhydrophobic engineering surfaces with tunable air-trapping ability. J. Micromech. Microeng. 2008, 18, 035024. [Google Scholar] [CrossRef]
  80. Wenzel, R.N. Resistance of solid surfaces to wetting by water. Ind. Eng. Chem. 1936, 28, 988–994. [Google Scholar] [CrossRef]
  81. Cengiz, U.; Cansoy, C.E. Applicability of Cassie–Baxter equation for superhydrophobic fluoropolymer–silica composite films. App. Surf. Sci. 2015, 335, 99–106. [Google Scholar] [CrossRef]
  82. Cassie, A.; Baxter, S. Wettability of porous surfaces. Trans. Faraday Soc. 1944, 40, 546–551. [Google Scholar]
  83. Li, B.; Yin, X.; Xue, S.; Mu, P.; Li, J. Facile fabrication of graphene oxide and MOF-based superhydrophobic dual-layer coatings for enhanced corrosion protection on magnesium alloy. Appl. Surf. Sci. 2022, 580, 152305. [Google Scholar] [CrossRef]
  84. Xu, W.; Wang, Z.; Han, E.-H.; Wang, S.; Liu, Q. Corrosion performance of nano-ZrO2 modified coatings in hot mixed acid solutions. Materials 2018, 11, 934. [Google Scholar] [CrossRef]
  85. Hu, C.; Zheng, Y.; Qing, Y.; Wang, F.; Mo, C.; Mo, Q. Preparation of Poly (o-toluidine)/Nano Zirconium Dioxide (ZrO2)/Epoxy Composite Coating and Its Corrosion Resistance. J. Inorg. Organomet. Polym. Mater. 2015, 25, 583–592. [Google Scholar] [CrossRef]
  86. Zhang, W.; Ji, G.; Bu, A.; Zhang, B. Corrosion and tribological behavior of ZrO2 films prepared on stainless steel surface by the sol–gel method. ACS Appl. Mater. Interfaces 2015, 7, 28264–28272. [Google Scholar] [CrossRef]
  87. ASTM G1-03; Standard Practice for Preparing, Cleaning and Evaluating Corrosion Test Specimens. ASTM International: West Conshohocken, PA, USA, 2003.
  88. Von Hauff, P.; Foroughi-Abari, A.; Bothe, K.; Cadien, K.; Barlage, D. ZrO2 on GaN metal oxide semiconductor capacitors via plasma assisted atomic layer deposition. Appl. Phys. Lett. 2013, 102, 251601. [Google Scholar] [CrossRef]
  89. Ashassi-Sorkhabi, H.; Moradi-Alavian, S.; Esrafili, M.D.; Kazempour, A. Hybrid sol-gel coatings based on silanes-amino acids for corrosion protection of AZ91 magnesium alloy: Electrochemical and DFT insights. Prog. Org. Coat. 2019, 131, 191–202. [Google Scholar] [CrossRef]
  90. Guo, X.; An, M.; Yang, P.; Su, C.; Zhou, Y. Property characterization and formation mechanism of anticorrosion film coated on AZ31B Mg alloy by SNAP technology. J. Sol-Gel Sci. Technol. 2009, 52, 335–347. [Google Scholar] [CrossRef]
  91. Ashraf, M.A.; Liu, Z.; Peng, W.-X.; Yoysefi, N. Amino acid and TiO2 nanoparticles mixture inserted into sol-gel coatings: An efficient corrosion protection system for AZ91 magnesium alloy. Prog. Org. Coat. 2019, 136, 105296. [Google Scholar] [CrossRef]
  92. Ashassi-Sorkhabi, H.; Moradi-Alavian, S.; Kazempour, A. Salt-nanoparticle systems incorporated into sol-gel coatings for corrosion protection of AZ91 magnesium alloy. Prog. Org. Coat. 2019, 135, 475–482. [Google Scholar] [CrossRef]
  93. Dong, Q.; Dai, J.; Qjan, K.; Liu, H.; Zhou, X.; Yao, Q.; Lu, M.; Chu, C.; Xue, F.; Bai, J. Dual self-healing inorganic-organic hybrid coating on biomedical Mg. Corros. Sci. 2022, 200, 110230. [Google Scholar] [CrossRef]
Figure 1. SEM images of the hybrid coating deposited on two different Mg–Al alloy surfaces: (a) PS–ZrO 2 –AM60–AlN (×1000-SEI mode); (b) PS–ZrO 2 –AM60 (×1000-SEI mode); (c) Image 1a (×20,000-SEI mode); (d) Image 1b (×20,000-SEI mode); (e) Image 1a (×15,000-LABE mode); (f) Image 1b (×15,000-LABE mode).
Figure 1. SEM images of the hybrid coating deposited on two different Mg–Al alloy surfaces: (a) PS–ZrO 2 –AM60–AlN (×1000-SEI mode); (b) PS–ZrO 2 –AM60 (×1000-SEI mode); (c) Image 1a (×20,000-SEI mode); (d) Image 1b (×20,000-SEI mode); (e) Image 1a (×15,000-LABE mode); (f) Image 1b (×15,000-LABE mode).
Coatings 13 01059 g001
Figure 2. SEM images (×1000) and maps of elements on (a) PS–ZrO2–AM60–AlN- and (b) PS–ZrO2–AM60-coated surfaces.
Figure 2. SEM images (×1000) and maps of elements on (a) PS–ZrO2–AM60–AlN- and (b) PS–ZrO2–AM60-coated surfaces.
Coatings 13 01059 g002
Figure 3. Cross-sectional SEM images of Mg–Al coated with hybrid PS–ZrO2: (a) AM60–AlN (×20,000) and (b) AM60 (×20,000).
Figure 3. Cross-sectional SEM images of Mg–Al coated with hybrid PS–ZrO2: (a) AM60–AlN (×20,000) and (b) AM60 (×20,000).
Coatings 13 01059 g003
Figure 4. XPS spectra of PS–ZrO2 deposited of the AM60–AlN and AM60 magnesium substrates: (a) C1s, (b) Si2p, (c) Zr3d, (d) O1s, and (e) Mg1s.
Figure 4. XPS spectra of PS–ZrO2 deposited of the AM60–AlN and AM60 magnesium substrates: (a) C1s, (b) Si2p, (c) Zr3d, (d) O1s, and (e) Mg1s.
Coatings 13 01059 g004
Figure 5. XRD patterns of non-coated and PS–ZrO2-hybrid-coated substrates.
Figure 5. XRD patterns of non-coated and PS–ZrO2-hybrid-coated substrates.
Coatings 13 01059 g005
Figure 6. Roughness surface values (Ra) of (a) AM60–AlN, (b) AM60, (c) PS–ZrO2–AM60–AlN, and (d) PS–ZrO2–AM60.
Figure 6. Roughness surface values (Ra) of (a) AM60–AlN, (b) AM60, (c) PS–ZrO2–AM60–AlN, and (d) PS–ZrO2–AM60.
Coatings 13 01059 g006
Figure 7. Contact angle (CA) of deionized water drops on the surfaces of (a) AM60–AlN, (b) AM60, (c) PS–ZrO2–AM60–AlN, and (d) PS–ZrO2–AM60.
Figure 7. Contact angle (CA) of deionized water drops on the surfaces of (a) AM60–AlN, (b) AM60, (c) PS–ZrO2–AM60–AlN, and (d) PS–ZrO2–AM60.
Coatings 13 01059 g007
Figure 8. Change in time of SME solution pH during the immersion of PS–ZrO2–AM60–AlN- and PS–ZrO2–AM60-hybrid-coated surfaces for up to 30 days.
Figure 8. Change in time of SME solution pH during the immersion of PS–ZrO2–AM60–AlN- and PS–ZrO2–AM60-hybrid-coated surfaces for up to 30 days.
Coatings 13 01059 g008
Figure 9. Change over time of the concentration of Mg 2 + ions released from non-coated and coated AM60–AlN and AM60 surfaces with the hybrid deposit of PS–ZrO2 exposed for 30 days to SME solution.
Figure 9. Change over time of the concentration of Mg 2 + ions released from non-coated and coated AM60–AlN and AM60 surfaces with the hybrid deposit of PS–ZrO2 exposed for 30 days to SME solution.
Coatings 13 01059 g009
Figure 10. SEM images of (a) PS–ZrO2–AM60–AlN (×500-SEI mode), (b) PS–ZrO2–AM60 (×500-SEI mode), (c) PS–ZrO2–AM60–AlN (×1000-LABE mode), and (d) PS–ZrO2–AM60 (×1000-LABE mode) surfaces after exposure for 30 days to SME solution.
Figure 10. SEM images of (a) PS–ZrO2–AM60–AlN (×500-SEI mode), (b) PS–ZrO2–AM60 (×500-SEI mode), (c) PS–ZrO2–AM60–AlN (×1000-LABE mode), and (d) PS–ZrO2–AM60 (×1000-LABE mode) surfaces after exposure for 30 days to SME solution.
Coatings 13 01059 g010
Figure 11. SEM images (×1000) and maps of elements on (a) PS–ZrO2–AM60–AlN and (b) PS–ZrO2–AM60 after exposure for 30 days to SME solution.
Figure 11. SEM images (×1000) and maps of elements on (a) PS–ZrO2–AM60–AlN and (b) PS–ZrO2–AM60 after exposure for 30 days to SME solution.
Coatings 13 01059 g011
Figure 12. SEM images (×1000) of surfaces after removal of the formed layers during the exposure for 30 days to SME solution: (a) PS–ZrO2–AM60–AlN, (b) PS–ZrO2–AM60, (c) AM60–AlN, and (d) AM60.
Figure 12. SEM images (×1000) of surfaces after removal of the formed layers during the exposure for 30 days to SME solution: (a) PS–ZrO2–AM60–AlN, (b) PS–ZrO2–AM60, (c) AM60–AlN, and (d) AM60.
Coatings 13 01059 g012
Figure 13. SEM images (×3000) and maps of elements on (a) PS–ZrO2–AM60–AlN and (b) PS–ZrO2–AM60 surfaces after removal of the layers formed during the exposure to SME solution.
Figure 13. SEM images (×3000) and maps of elements on (a) PS–ZrO2–AM60–AlN and (b) PS–ZrO2–AM60 surfaces after removal of the layers formed during the exposure to SME solution.
Coatings 13 01059 g013
Figure 14. SEM images (×500) of cross-sections on the surfaces of (a) PS–ZrO2–AM60–AlN, (b) PS–ZrO2–AM60, (c) AM60–AlN, and (d) AM60 (×250) after their exposure for 30 days to chloride SME solution.
Figure 14. SEM images (×500) of cross-sections on the surfaces of (a) PS–ZrO2–AM60–AlN, (b) PS–ZrO2–AM60, (c) AM60–AlN, and (d) AM60 (×250) after their exposure for 30 days to chloride SME solution.
Coatings 13 01059 g014
Figure 15. Nyquist diagrams and Bode plots of phase angle of AM60 and AM60–AlN coated with the hybrid PS–ZrO2 (in black color), compared to those of non-coated surfaces after immersion in SME model solution for (a) 1 day and (b) 15 days.
Figure 15. Nyquist diagrams and Bode plots of phase angle of AM60 and AM60–AlN coated with the hybrid PS–ZrO2 (in black color), compared to those of non-coated surfaces after immersion in SME model solution for (a) 1 day and (b) 15 days.
Coatings 13 01059 g015
Figure 16. Equivalent circuit of PS–ZrO2-coated AM60 and AM60–AlN surfaces, and non-coated, during the exposure to SME solution.
Figure 16. Equivalent circuit of PS–ZrO2-coated AM60 and AM60–AlN surfaces, and non-coated, during the exposure to SME solution.
Coatings 13 01059 g016
Table 1. Reagents used for the synthesis of the hybrid PS–ZrO2.
Table 1. Reagents used for the synthesis of the hybrid PS–ZrO2.
PrecursorsSolventsCatalysisAnti-InhibitorInitiator
Zirconium isopropoxide
( Zr ( OPr ) 4 )
Anhydrous ethanol (EtOH)Nitric acid
( HNO 3 )
Sodium hydroxide (NaOH) Benzyl peroxide
(BPO)
Styrene monomer
(ST)
Deionized water
( H 2 O )
Hydrochloric acid
HCl
3-(trimethoxysilyl)propyl methacrylate (TMSPM)
Table 2. Composition of simulated marine–coastal environment (SME, pH = 7.94) [14].
Table 2. Composition of simulated marine–coastal environment (SME, pH = 7.94) [14].
Reagents NaCl Na 2 SO 4 NaHCO 3
Concentration 5 . 84   g   L 1 4 . 09   g   L 1 0 . 20   g   L 1
Table 3. EDS elemental analysis (wt.%) of several randomly selected zone areas (Figure 1) of the PS–ZrO2–AM60–AlN (Zones 1–2) and PS–ZrO2–AM60 (Zones 3–4) tested samples.
Table 3. EDS elemental analysis (wt.%) of several randomly selected zone areas (Figure 1) of the PS–ZrO2–AM60–AlN (Zones 1–2) and PS–ZrO2–AM60 (Zones 3–4) tested samples.
ElementCOMgAlSiZr
Zone 120.2419.6046.062.202.729.17
Zone 219.8719.3945.583.032.859.28
Zone 318.1720.2547.161.782.789.85
Zone 418.1920.1646.662.102.969.93
Table 4. Elemental analysis (wt.%) of two zones (Figure 10) of PS–ZrO2–AM60–AlN and PS–ZrO2–AM60 surface layers after exposure for 30 days to SME solution.
Table 4. Elemental analysis (wt.%) of two zones (Figure 10) of PS–ZrO2–AM60–AlN and PS–ZrO2–AM60 surface layers after exposure for 30 days to SME solution.
ElementCONaMgAlSiSClMnZr
Zone 1-19.365.032.1918.791.49--41.6111.53
Zone 214.5640.221.1318.624.041.431.192.05-16.76
Table 5. Elemental analysis (wt.%) of zones of interest (Figure 12) on the coated and non-coated surfaces after the removal of the layers formed during the exposure for 30 days to SME solution.
Table 5. Elemental analysis (wt.%) of zones of interest (Figure 12) on the coated and non-coated surfaces after the removal of the layers formed during the exposure for 30 days to SME solution.
ElementCOMgAlSiMn
Zone 12.511.142.1736.921.1856.09
Zone 24.461.8289.064.66--
Zone 35.672.5857.6434.11--
Table 6. Fitting parameters from EIS data of PS–ZrO2-hybrid-coated AM60–AlN and AM60 surfaces after their immersion for 1 and 15 days in SME chloride solution.
Table 6. Fitting parameters from EIS data of PS–ZrO2-hybrid-coated AM60–AlN and AM60 surfaces after their immersion for 1 and 15 days in SME chloride solution.
PS–ZrO2–AM60–AlN
Time (Days) R s
( Ω   cm 2 )
CPE 1
( μ S   s n   cm 2 )
n 1 R 1
( k Ω   cm 2 )
CPE 2
( μ S   s n   cm 2 )
n 2 R 2
( k Ω   cm 2 )
R p
( k Ω   cm 2 )
169.16 ± 0.547.85 ± 0.210.91 ± 0.017.22 ± 0.140.66 ± 0.130.84 ± 0.112.10 ± 0.279.32 ± 0.30
15 71.35 ± 0.4937.18 ± 0.640.88 ± 0.0111.36 ± 0.243.97 ± 0.530.99 ± 0.132.61 ± 0.2413.97± 0.37
PS ZrO 2 AM 60
189.93 ± 0.733.76 ± 0.080.88 ± 0.0110.63 ± 0.270.26 ± 0.020.84 ± 0.054.75 ± 0.2615.35 ± 0.37
1580.20 ± 0.54 39.48 ± 0.660.88 ± 0.0113.23 ± 0.253.361 ± 0.550.99 ± 0.173.93 ± 0.2217.16± 0.33
Table 7. Fitting parameters from EIS data of non-coated AM60–AlN and AM60 after their immersion for 1 and 15 days in SME solution.
Table 7. Fitting parameters from EIS data of non-coated AM60–AlN and AM60 after their immersion for 1 and 15 days in SME solution.
AM60–AlN
Time (Days) R s
( Ω   cm 2 )
CPE 1
( μ S   s n   cm 2 )
n 1 R 1
( k Ω   cm 2 )
CPE 2
( mS   s n   cm 2 )
n 2 R 2
( k Ω   cm 2 )
R p
( k Ω   cm 2 )
159.89 ± 0.4510.49 ± 0.260.93 ± 0.017.00 ± 0.150.54 ± 0.080.87 ± 0.082.72 ± 0.289.77 ± 0.32
15 66.56 ± 0.4343.23 ± 0.720.92 ± 0.049.23 ± 0.166.51 ± 0.380.97 ± 0.222.17 ± 0.2211.40 ± 0.27
AM 60
168.62 ± 0.5011.45 ± 0.310.94 ± 0.016.12 ± 0.140.48 ± 0.080.86 ± 0.0842.37 ± 0.248.49 ± 0.28
1571.21 ± 0.4639.86 ± 0.670.93 ± 0.0110.06 ± 0.166.13 ± 0.470.97 ± 0.2252.45 ± 0.2512.51 ± 0.30
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

Chávez, L.; Veleva, L.; Sánchez-Ahumada, D.; Ramírez-Bon, R. Hybrid Coating of Polystyrene–ZrO2 for Corrosion Protection of AM Magnesium Alloys. Coatings 2023, 13, 1059. https://doi.org/10.3390/coatings13061059

AMA Style

Chávez L, Veleva L, Sánchez-Ahumada D, Ramírez-Bon R. Hybrid Coating of Polystyrene–ZrO2 for Corrosion Protection of AM Magnesium Alloys. Coatings. 2023; 13(6):1059. https://doi.org/10.3390/coatings13061059

Chicago/Turabian Style

Chávez, Luis, Lucien Veleva, Diana Sánchez-Ahumada, and Rafael Ramírez-Bon. 2023. "Hybrid Coating of Polystyrene–ZrO2 for Corrosion Protection of AM Magnesium Alloys" Coatings 13, no. 6: 1059. https://doi.org/10.3390/coatings13061059

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