Next Article in Journal
Effects of Synchronous Bias Mode and Duty Cycle on Microstructure and Mechanical Properties of AlTiN Coatings Deposited via HiPIMS
Previous Article in Journal
Preparation Method and Application of Nanobubbles: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

In Situ Growth of High-Performance ZnAl-Layered Double Hydroxides/Al2O3 Composite Coatings on Aluminum Alloy

College of Mechanical and Electrical Engineering, Harbin Engineering University, Harbin 150001, China
*
Author to whom correspondence should be addressed.
Coatings 2023, 13(9), 1511; https://doi.org/10.3390/coatings13091511
Submission received: 4 August 2023 / Revised: 19 August 2023 / Accepted: 21 August 2023 / Published: 26 August 2023

Abstract

:
In this research, high-performance ZnAl-layered double hydroxides (ZnAl-LDHs) with different Zn2+ concentrations were prepared on the surface of an anodized 1060 aluminum alloy using the in situ growth method, and the influence of Zn2+ concentration on corrosion resistance and tribological behavior was further explored. The surface morphology, element distribution, phase composition, and mechanical performance of ZnAl-LDHs/anodic aluminum oxide (AAO) composite coatings were tested and analyzed. Subsequently, the investigation of the anti-corrosion properties and tribological behavior of as-prepared composite coatings and AAO were conducted experimentally. The results show that with the increase in Zn2+ concentration, the thickness and bonding strength of the coatings gradually increase, and the hardness progressively decreases. The ZnAl-LDHs/AAO composite coating with a Zn2+ concentration of 0.3 mol/L exhibits the best anti-corrosion ability, which has the minimum corrosion current density of 2.435 × 10−9 A/cm2. This can be attributed to the uniform and compact ZnAl-LDH films where the porous structure of the AAO is well sealed. Moreover, the excellent friction-reduction and anti-wear properties of ZnAl-LDHs/AAO composite coatings with larger Zn2+ concentrations are verified using a ball-on-disc rotating wear tester under dry wear. A reasonable mechanism for improving tribological properties of resulting ZnAl-LDHs/AAO composite coatings is proposed.

1. Introduction

Aluminum (Al) and its alloys are widely used in the aerospace, transportation, and marine industries due to their high strength ratio, good electrical and thermal conductivity, and excellent manufacturing characteristics [1,2,3]. The 1060 Al alloy is cost-effective and exhibits the advantages of light weight, high strength-to-weight ratio, excellent electrical conductivity, and good recyclability. However, some technical disadvantages remain unsolved. Al alloys possess a low hardness, a high coefficient of friction (COF) (ca. 0.4–0.8), and high wear rates (ca. 10−4–10−5 mm3/m) [4,5]. Moreover, the uppermost thin protective oxide coating (ca. 1.5–10 nm) with high passivity on the surface of the Al alloy could be readily destroyed under the extreme chemical, physical, or mechanical conditions, and thus leads to increased cracks and corrosion [6,7]. To further expand the application field of Al alloys, it is of great significant to protect Al alloys from corrosion and wear damage, and further develop the coatings with both excellent corrosion resistance and tribological properties.
Anodic oxidation is a low-cost and simple operation that is beneficial to large-scale manufacturing [8,9,10], which has become a prevailing technique to obtain superior corrosion resistance and tribological properties upon the surface of diverse metal alloys such as magnesium (Mg), titanium, and aluminum [11,12,13]. The factors affecting the anodization mainly include current density, oxidation duration, electrolyte parameters (i.e., concentration, temperature, and pH value), and so on [14,15,16]. Research has found a growth in the thickness of the anodized membrane with increasing oxidation duration, and the anodic aluminum oxide (AAO) has an amorphous structure, where the chemical composition mainly consisted of Al2O3 [17]. In addition, various aqueous electrolytes (e.g., weak acid, strong acid, and alkaline) were used to prepare the AAO membranes, in which the anodic coatings obtained from weak acidity solutions such as malic acid have fewer pores [18,19,20]. Although the anodization of Al alloys produces a protective film, physical defects (e.g., bulges, cracks, and pores) upon the surface of AAO are unavoidable and exhibit a high specific surface area [21]. This leads to a limited protection of the Al alloy substrate against attacks from aggressive media due to the existence of channels, which provide deep cracks and interconnected pores. Therefore, the porous AAO membrane structure and larger COF limit its application in some extreme working conditions [22]. It is essential to remove the inherent defects and develop efficient post-sealing treatments to AAO coatings.
Layered double hydroxides (LDHs), a kind of anionic clay with a special layer structure, can load various functional anions. Their general formula can be expressed as [M2+1−xM3+x(OH)2]x+(An−)x/n∙yH2O (M2+ and M3+ are metal cations in the laminates, and An− is the charge-balancing anion between the layers). Owing to the diversity of chemical composition, unique structures, exchange ability, high thermal stability, and promising actively protective properties [23,24,25,26,27], LDHs have been widely investigated in metal corrosion protection [28], lubricant additives [11,29,30], and other uses such as fuel cell, electrode material, electrocatalysts, and supercapacitor [31,32,33,34]. Many different methods employed for the fabrication of LDH coatings have been reported in the literature, among which the in situ growth approach seems to be one of the most flexible, simplest, and inexpensive techniques for preparing LDH films on the surface of metal alloys [11,20,22,35].
Recently, applications of LDHs as a post-sealing treatment to protect metal alloys (e.g., Al, Mg) has drawn increasing attention. Chen F et al. utilized the inhibitor-loaded LDHs formed on the MAO ceramic layer to heal the structural defects and found that the vanadate-loaded LDH nanoplatelets significantly improved the corrosion resistance, yielding a long-term active protection for the MAO-covered Al alloy substrate [22]. Wu L et al. synthesized the MgAl-LDH films on an anodized Mg alloy and found that the films with bilayer structure significantly increased the impedance of anodic films, demonstrating an effective protection for the Mg alloy [24]. Xue L et al. fabricated three-dimensional ZnAl-LDHs/AAO and NiAl-LDHs/AAO membranes using porous AAO templates as an Al3+ source and substrate through a simple precipitant-free in situ growth technique, and found that the crystallinity and morphology of samples can be regulated artificially by the reaction time, acidity of reaction solutions, and concentration of metal ions, among which the weak acidic conditions are convenient to prepare the LDHs/AAO membranes [36]. In addition, Ding P et al. found that the kind of metal cations has a great effect on the growth and crystallization process of the LDH crystals, which results in the different structure and morphologies of the LDH coatings [35]. Ni2+ shows excellent action for synthesizing LDH films with homogenous morphology, and the obtained Ni-based LDH lamellas had small size, large surface area, and high bulk density, exhibiting an improved protective effect for the Al alloy [35,37]. However, the investigation for tribological properties of in situ growth LDHs/AAO composite coatings on the Al alloys has not been reported yet, especially regarding the influence of Zn2+ concentration on the corrosion resistance and tribological behavior of ZnAl-LDHs/AAO composite coatings.
LDH coatings prepared on the anodized Al alloy surface using in situ growth have a series of performance advantages: the coating is compact and uniform, the structure is controllable, and the bonding is good to the matrix. In this paper, we explored the influence of different Zn2+ concentrations (0.05 mol/L, 0.10 mol/L, 0.20 mol/L, and 0.30 mol/L) in terms of surface morphology, element distribution, and phase composition of ZnAl-LDHs/AAO composite coatings by using scanning electron microscopy (SEM), X-ray diffraction (XRD), and Fourier Transform Infrared Spectroscopy (FT-IR) first, and then concentrated on the hardness and interfacial bonding strength. The electrochemical tests were carried out using an electrochemical workstation composed of a three-electrode system, and the tribological experiments were conducted on a ball-on-disc rotating wear tester under dry wear. The anti-corrosion and tribological properties of the as-prepared composite coatings were discussed, and the effect of Zn2+ concentration on the anti-corrosion property and tribological performance of ZnAl-LDHs/AAO composite coatings was analyzed. Based on the results of COF, wear loss, and worn surface studies, the lubrication mechanism for ZnAl-LDHs/AAO composite coatings was proposed, which could further optimize the lubricating behavior of two-dimensional layered materials.

2. Materials and Methods

2.1. Preparation of ZnAl-LDHs/AAO Composite Coatings

Commercial 1060 Al alloy (with nominal composition, wt%: Fe 0.243, Si 0.057, Cu 0.004, Ti 0.021, Zn ≤ 0.05, Mn ≤ 0.03, Mg ≤ 0.03%, and balance Al) was used as the substrate. Samples with dimensions of 40 mm × 30 mm × 0.5 mm were polished with water-resistant sandpaper of grade 400, 800, 1200, 1500, and 2000 grit successively, followed by cleaning in a mixed solution of alcohol and acetone (3:1 vol.) ultrasonically, rinsing in deionized water, and drying in ambient atmosphere naturally. All of the analytical grade chemicals were utilized without further purification. Deionized water was employed as a solvent.
The overall manufacturing process of preparing ZnAl-LDHs/AAO composite coatings on the surface of the Al alloy substrate is briefly illustrated in Figure 1. The obtained samples were put into alkali washing solution (NaOH (5 g/L), Na2CO3 (20 g/L) and Na3PO4 (20 g/L)) for 3 min and moved into acid washing solution (HNO3 (40 g/L)) for polishing of 15 s. Then, all samples were washed with deionized water and dried. After the pre-treatment, the 1060 Al plates were anodized in phosphoric acid electrolyte (0.3 mol/L) for 1 h under current density of 0.01 A/cm2 by using the direct current (DC) electrolytic treatment and a circulating water-cooling device (T = 20 °C). A certain amount of ZnSO4·7H2O and (NH4)2SO4 were dissolved in 60 mL of deionized water to obtain solutions with different Zn2+ concentrations (0.05 mol/L, 0.1 mol/L, 0.20 mol/L, and 0.30 mol/L) and their pH values were adjusted to 8.5 by adding ammonia water (1.2 mol/L). Subsequently, the AAO substrate was placed in the mixed solution using an autoclave at 90 °C for 24 h. There are two procedures of ionization equilibrium and coordination equilibrium in the formation of LDH. The aluminum source on the anodized aluminum surface continuously consumed the OH- in solution, causing the ammonia to undergo ionization equilibrium in a positive direction. The occurrence of ionization equilibrium caused the concentration of NH3 to become smaller, which promoted the dissociation of [Zn(NH3)6]2+, which in turn caused Zn2+ to be released and the concentration of NH3 to increase. At this time, the presence of Al3+, Zn2+, and OH provided conditions for the growth of ZnAl-LDH films, which can be generated in situ on the AAO surface. After the in situ growth of LDH films on the surface of AAO, the substrates were taken out, rinsed with deionized water 6 times, and dried at 70 °C. Finally, the ZnAl-LDH/AAO composite coatings were successfully synthesized.

2.2. Characterization and Tests

To observe the cross-sectional profiles of ZnAl-LDHs/AAO composite coatings, specimens were completely wrapped using epoxy resin, left to set for 12 h, the cross-section of the specimens were then sandpapered until smooth. The surface morphologies and cross-sectional profiles of ZnAl-LDHs/AAO composite coatings were examined using SEM (FEI, Hillsboro, OR, USA) and the chemical composition was characterized using the energy dispersive spectrum (EDS). The crystal structure of the samples was confirmed using XRD (DX-2700B, Pingdingshan, China) over a 2θ range of 5–70° and a scanning speed of 5° min−1. During the testing process, ZnAl-LDHs films were scraped from the surface of the AAO with a razor blade, and the powder was added to an appropriate amount of potassium bromide for grinding, and finally, the samples were prepared by KBr presses for infrared spectroscopy. The functional groups of the samples were acquired using spectroscopic analysis using the FT-IR (Spectrum 100, PerkinElmer, Waltham, MA, USA) in the range of 500 cm−1 to 4000 cm−1 at room temperature. The average hardness value (a mean of six measurements) was tested using a micro-hardness tester (HV-1000IS, Shanghai, China) with a diamond tetrapyramid indenter. The interfacial bonding force was examined by a pull-off tester (Ogdensburg, NY, USA), special glue was employed to bond the spindle and sample, and then stayed 12 h at room temperature.
The electrochemical characteristics of the samples were evaluated using an electrochemical workstation (CHI660B, Chenhua, Shanghai, China) after being soaked in 3.5 wt% NaCl solution for 72 h at room temperature with a typical three-electrode system. The samples were used as the working electrode, while a platinum sheet and a saturated calomel electrode (SCE) were the counter and the reference electrodes, respectively. The test frequency of electrochemical impedance spectroscopy (EIS) ranged from 1000 Hz down to 10 Hz, and the scanning voltage was 10 mV. The experiments were scanned from −0.5 V to +0.5 V with respect to the open circuit potential at a scan rate of 1 mV/s. The corrosion potential (Ecorr) and corrosion current density (icorr) were obtained from the potentiodynamic polarization curves. The tribological experiment was conducted on a friction-abrasion testing machine (SFT-2M, Lanzhou, China), with a GCr15 steel ball (a diameter of 5 mm) as friction counterpart, a load of 1 N, a rotation speed of 200 RPM (revolutions per min), and a rotation radius of 2 mm. Subsequently, the worn surfaces of the samples were observed using SEM. The mass losses of samples were measured using an electronic analytical balance (EX225D, Ohaus, Parsippany, NJ, USA, with an accuracy of 0.1 mg and 0.01 mg readability); each group was measured five times to ensure the repeatability of the data and arithmetic means are used as the final experimental results.

3. Results and Discussion

3.1. Chemical and Phase Composition

The diffraction patterns of the ZnAl-LDHs/AAO composite coatings with different Zn2+ concentrations are shown in Figure 2a. The (003), (006), and (012) reflection peaks appear at around 10°, 19.9°, and 34°, indicating the successful synthesis of ZnAl-LDHs for all Zn2+ concentrations, which is consistent with the results of the research by Shen et al. [38]. The appearance of the characteristic peak of Al2O3 in all groups suggests that the composite coatings consist of ZnAl-LDHs and AAO. With the increase in Zn2+ concentration in the solution, there is an obvious augment in the intensity of the LDH phases and a peak (003) narrowing, indicating more LDH platelets with high crystallinity contained in the deposited coatings. The phenomena may be due to the formation of compact LDH lamellas. Meanwhile, there is an obvious decrease in the intensity of the Al2O3 phases, which can be attributed to the cover by the denser deposited LDH films. The characteristic diffraction peak of LDH films (003) undergoes a blue shift at the concentration of 0.2 mol/L and 0.3 mol/L. According to the Bragg reflection, the calculated value of basal interlayer spacing d (003) is 0.890 nm at the concentration of 0.05 mol/L and 0.10 mol/L. Therein, with the increase in Zn2+ concentration, the basal interlayer distance increases to d (003) = 1.132 nm and d (003) = 1.168 nm at the concentrations of 0.20 mol/L and 0.30 mol/L, respectively. Here, the increased Zn2+ concentration can both change the LDH’s structural characteristics (morphologies and size) and the basal interlayer spacing between LDH layers.
Figure 2b shows the FT-IR spectra of ZnAl-LDHs/AAO composite coatings with different Zn2+ concentration, where the peaks at 1138 cm−1 and 1624 cm−1 can be attributed to the intercalated sulfate ions and the bending vibration of hydroxyl in interlayer water molecules, respectively. The absorption peak at 3450 cm−1 can be assigned to the γ(O-H) stretching vibration and the interaction of interlayer water molecules between LDH plates. The peaks that occur at the range of 617–837 cm−1 belong to M-OH characteristic absorption peaks of metal–oxygen bonds between LDH layers. These values corresponded exactly to the characteristic peaks of ZnAl-LDH [39]. Compared with the FT-IR spectra under different Zn2+ concentrations, the absorption peak of SO42− and the M-O absorption peak of the interlayer metal–oxygen bonds are more obvious; this corresponds to the XRD results. Hence, ZnAl-LDH films with complete functional groups and lamellae structure can be formed under different concentrations of Zn2+.

3.2. Surface and Cross-Sectional Morphology

The SEM images of ZnAl-LDH films at different Zn2+ concentrations are shown in Figure 3. All the ZnAl-LDH nanoparticles possessed typical hexagonal symmetry structure with uniform size [40]. Apparently, the homogeneous ZnAl-LDH platelets are all perpendicular to the AAO substrate, which is consistent with the results of Wu [41]. With the increase in Zn2+ concentration, it can be observed that the LDH nanosheets become more compact and the size of voids between them is reduced. Therein, the ZnAl-LDH nanosheets at the concentration of 0.3 mol/L possess a minimum size of about 3.75 μm and the resulting films are denser than those of other groups. In addition, the content ratio of the Al and Zn elements is about 1:12 (see Figure 3d). This is because that the AAO substrate is covered by the LDH films to a greater extent in this Zn2+ concentration, which can also be verified by the covered “flower-like” cluster structure from Figure 3c.
The cross-sectional photographs of the ZnAl-LDHs/AAO composite coatings with different Zn2+ concentrations are shown in Figure 4. Based on our earlier published works, the structure of AAO includes two parts (inner barrier layer and outer porous oxide layer), which also can be observed in the LDHs/AAO composite coatings [42]. It is clearly visible that the ZnAl-LDH films become thicker with the increased Zn2+ concentration, among which the film thickness in the concentration of 0.3 mol/L is the largest (99.26 μm). Wang et al. prepared MgAl-LDHs/Al2O3 composite coatings on the surface of Mg alloy AZ31 and found that the thickness of composite coatings is enhanced with the increased concentration of Al2O3 nanoparticles [11]. When the Zn2+ concentration ranges from 0.05 mol/L to 0.30 mol/L, the growth rate for the thickness of LDH films gradually becomes lower, indicating a limited improved effect of larger Zn2+ concentration on the thickness growth of LDH films.

3.3. Hardness and Interfacial Bonding Strength

Figure 5a shows the hardness of AAO and ZnAl-LDHs/AAO composite coatings with different Zn2+ concentrations, where the curve of ZnAl-LDH film thickness is added as a reference. It can be seen that the hardness of the composite coatings gradually decreases with the rise in Zn2+ concentration. There is a minimum value in the concentration of 0.30 mol/L, and the corresponding value is 112 HV, which is about 50% of that of the AAO (230 HV). The results indicate that the structure of ZnAl-LDHs/AAO composite coatings is composed of the hard AAO and soft LDH films. However, the thickness of the ZnAl-LDH films at different Zn2+ concentrations is inversely correlated with those of the hardness.
The interfacial bonding strength of samples was examined by a coating adhesion automatic scratch tester, and the average value (a mean of three measurements) was taken as the interfacial bonding strength. The interfacial bonding force between the ZnAl-LDH films and AAO matrix is shown in Figure 5b, it can be seen that the interfacial bonding forces of samples at different Zn2+ concentrations are 7.20 MPa, 8.50 MPa, 8.75 MPa, and 8.85 MPa, respectively. The bonding force gradually enhances with the increase in Zn2+ concentration, and reaches a maximum of 8.85 MPa at 0.30 mol/L. This may be due to the fact that the higher Zn2+ concentration offers more divalent Zn2+ for growth of LDH nanosheets and the further compact LDH films, which can be confirmed from the SEM images of ZnAl-LDH films in Figure 3. In addition, Zhang K et al. found that the bonding strength between micro-arc oxidation (MAO) coatings and 7N01Al alloy increases with the augment of NaAlO2 concentration to a certain extent [43]. Here, the greater the power used for growth of the coating, the more beneficial it is for the metallurgical combination of the substrate and coatings, and the bonding strength will be enhanced accordingly. As a result, the less energy consumed by the electrolyte, the better the interfacial bonding strength. The energy employed in the electrolyte decreases with the reduction in its resistance, which is mainly connected to the concentration of conductive particles. Therefore, the higher the Zn2+ concentration, the lower the corresponding resistance of the electrolyte will be, which will decrease the energy consumed by the electrolyte and the interfacial bonding strength will gradually enhance.

3.4. Corrosion Resistance of ZnAl-LDHs/AAO Composite Coatings

The potentiodynamic polarization curves are used to investigate the anti-corrosion behavior of different coated samples, as shown in Figure 6. The corrosion potential (Ecorr) and corrosion current density (icorr) derived from the curves are summarized in Table 1. Evidently, a higher corrosion potential or a lower corrosion current density represent a lower corrosion rate and a better corrosion resistance in a typical polarization curve. The AAO has the most negative Ecorr (−0.775 V) and the highest icorr (3.020 × 10−7 A/cm2) among all groups, which indicates that the AAO is easily corroded in chloride-containing solutions. All groups with ZnAl-LDHs/AAO composite coatings have the larger Ecorr values, meanwhile, their icorr values are two orders of magnitude lower than that of the bare AAO substrate. The reason may be that the ZnAl-LDH nanosheets passivate most of the active Al sites on the substrate surface to increase the potential [37]. Importantly, all coated groups appear to be very effective in protecting the AAO matrix from corrosion, the ZnAl-LDHs/AAO composite coating at the concentration of 0.30 mol/L has the highest Ecorr (−0.462) and the lowest icorr (2.435 × 10−9 A/cm2), indicating a superior corrosion resistance.
Figure 7 shows the experimental data of the electrochemical impedance spectroscopy (EIS). The Nyquist curve (see Figure 7a) of ZnAl-LDHs/AAO composite coatings is composed of a capacitive arc, which evaluates the anti-corrosion behavior by the size of the capacitive arc radius. It can be observed that each coating has a larger capacitive arc radius and plays a role in blocking the corrosive medium as a resistive barrier layer during immersion. It is well-known that the low-frequency impedance modulus of Z (10−2–0 Hz) in the Bode plot (see Figure 7b) can reflect the corrosion resistance of the coatings [44]. Compared with other concentration conditions, the |Z| value of the low-frequency region of the LDHs/AAO composite coating in the concentration of 0.3 mol/L is significantly higher, which indicates that the composite coating prepared under this concentration has the best corrosion resistance. This result is consistent with the potentiodynamic polarization curve. To further confirm the corrosion resistance of ZnAl-LDHs/AAO composite coatings with different Zn2+ concentrations, the physical model of ZnAl-LDHs/AAO composite coatings and the corresponding equivalent circuit are established, as shown in Figure 8. The fitting results are shown in Table 2. Rsol represents the solution resistance, Rct is used to describe the interface transfer resistance, Rox represents the resistance between the LDH film and oxide film. It can be observed that Rct of ZnAl-LDHs/AAO composite coating at the concentration of 0.30 mol/L is the largest, which has the best corrosion resistance. ZnAl-LDHs/AAO composite coatings with good crystallinity were formed at different concentrations of Zn2+. With the continuous increase in Zn2+ concentration, the distance between LDH nanosheets gradually increased and the distribution of LDH nanosheets became more compact. Hence, LDH nanosheets with higher concentrations of Zn2+ played a better role in sealing AAO film. Additionally, the thickness of the ZnAl-LDHs/AAO composite coatings became thicker, and the binding force became stronger with increased Zn2+ concentration. And the effect of inhibiting the penetration behavior of corrosive ions was enhanced, which significantly improved the corrosion resistance of ZnAl-LDHs/AAO composite coatings.

3.5. Tribological Behavior of ZnAl-LDHs/AAO Composite Coatings

Figure 9a shows the relationship between the COF and sliding time under dry wear for all groups. It can be observed that the COF of AAO exhibit significant fluctuation at the initial stage of the test, and then gradually increase up to a relatively stable value (0.7) throughout the test. The larger COF can be attributed to the decreased effective contact area; the earlier published studies have proved that the actual contact area of surface affects the tribological properties significantly [45,46]. Likewise, bulges, cracks, and pores upon the surface of AAO reduced the effective contact area, which will generate severe abrasive wear, “edge crushing phenomena”, along the edges of these defects and finally lead to more wear debris left on the surfaces [45,47]. The COF of samples with lower Zn2+ concentrations (i.e., 0.05 mol/L and 0.10 mol/L) gradually become stable after the initial running-in stage, giving a relatively constant value about 0.2 and 0.35 until the end of the tests, respectively. At low Zn2+ concentrations, it is easy for ZnAl-LDH nanosheets to peel away from the substrate surface caused by the lower interfacial bonding force (see Figure 5b), which resulted in the larger COF values. However, the sample with the Zn2+ concentration of 0.20 mol/L has a long-term large COF of approximately 0.4 in the initial friction stage due to the reduced actual effective contact area resulted from relatively rough surface caused by the formation of a large number of “flower-like” LDH nanosheets (see Figure 3c), which results in the largest wear loss (1.1 mg) among all groups (see Figure 9b). In the case of the sample with a Zn2+ concentration of 0.30 mol/L, a flat and low COF curve can be observed, which is reduced by 85.71% of that of AAO and also exhibits little fluctuation. The reason for this phenomenon is that the sample with the concentration of 0.30 mol/L, possessing a uniform and compact surface (see Figure 3d), has fewer collisions among micro-peaks during sliding process. The COF curves of samples with different Zn2+ concentrations show larger distinction in terms of values, stability, and trend, indicating the significant effect for Zn2+ concentration of ZnAl-LDHs on the COF. It was obvious that the COF curves for each group with ZnAl-LDHs/AAO composite coatings gradually fluctuated with the increase in degree of wear, but there was no evident transition point, implying that the LDH composite films still existed [48,49].
The wear losses of samples with different Zn2+ concentrations under dry wear are shown in Figure 9b. It can be observed that the wear losses have a reduced trend with the increased Zn2+ concentration, but there is a rise in the concentration of 0.20 mol/L, which should be attributed to the long-term large COF caused by the rough peaks of surface in the initial friction stage. High contact pressure due to the rough surface would crush particles and break the layered structure of ZnAl-LDHs, which also restricts its lubrication properties (see Figure 3 and Figure 10). Therein, the wear loss of the sample with the Zn2+ concentration of 0.30 mol/L is the lowest (0.20 mg), which is caused by the lower average COF and larger interfacial bonding strength. Additionally, the published literature also found that the thickness and pore size of the AAO coatings has an effect on the tribological behavior to a certain degree [50,51]. As a result, the AAO parameters and LDH nanosheets may have a coupling impact on the friction and wear performance of ZnAl-LDHs/AAO composite coatings.
To further explain the effect of different Zn2+ concentration on the tribological behavior of ZnAl-LDHs/AAO composite coatings under dry wear, the SEM images and EDS results of the worn surfaces are acquired, as shown in Figure 10. The wear track widths of samples with the Zn2+ concentration of 0.05 mol/L, 0.10 mol/L, 0.20 mol/L, and 0.30 mol/L are 343.21 μm, 312.51 μm, 340.42 μm, and 296.77 μm, respectively. Therein, the wear scar width of the sample with the concentration of 0.30 mol/L is the minimum and the reduced width is nearly 23% of that of the sample with the concentration of 0.20 mol/L, showing excellent wear resistance. In addition, there were no evident streaks and debris generated on wear traces compared with other groups. This result is consistent with the wear losses of samples. The EDS results show that a stable friction-reducing LDH films on the surface of the LDHs/AAO composite coatings after friction and wear tests, which further confirms the formation of the ZnAl-LDH film and its friction-reduction and wear resistance. In summary, the lubricating mechanism of ZnAl-LDH nanosheets as lubricant additives is that nanoparticles may fill pores, and the friction force is reduced by the relative sliding motion of uniform and smooth LDH layers, which further forms a protective film on the surfaces of friction pairs [39].
The friction and wear process can be divided into two stages: The first stage is the friction and wear of ZnAl-LDH films, which corresponds to the initial running-in period of the sliding, and the second stage is the friction and wear of ZnAl-LDHs/AAO composite coatings. The plastic deformation of matrix surface can be contributed to by transitory heat owing to dissipated energy caused by rough surface or asperity collisions, resulting in relatively higher resistance to sliding [52]. The ZnAl-LDH films play a role of lubrication in the process of friction due to its relatively weak van der Waals forces for easy shear between their laminates, which significantly reduces the COF for the AAO matrix. Moreover, wear loss and worn scars obtained an the effective reduction under the action of the LDHs with largest Zn2+ concentration (0.30 mol/L) compared with the other groups. As shown in Figure 11, the tribological model of the ZnAl-LDHs/AAO composite coated samples is established. Smaller LDH nanosheets were easily embedded in the pores and defects of sliding surfaces, forming a thick tribofilm, which also smoothed the collision of asperities during actual friction process. The friction-reduction and anti-wear mechanism of ZnAl-LDHs/AAO composite coatings can be summarized as follows: The soft ZnAl-LDH films and hard AAO substrate act the bearing effect together, and the ZnAl-LDH films with composite flake structure are destroyed and deformed due to the action of load extrusion and frictional tangential force. Finally, the ZnAl-LDH lubricating films are formed between the friction pairs. Meanwhile, the LDH nanosheets stored in the cracks and holes of the AAO matrix will constantly move to the interface to replenish the lubricating films due to the action of load and tangential force friction. In addition, the LDH transfer films, namely, lubricating membrane, formation during the sliding friction process, are transferred from AAO to the surface of friction counterpart (ball), which changes the contact model and a lubricating interface is formed here. Therefore, the synergistic effect of the hard AAO matrix, ZnAl-LDH films, and LDH transfer films endows the ZnAl-LDHs/AAO composite coatings with excellent tribological properties.

4. Conclusions

In this paper, ZnAl-LDHs/AAO composite coatings were synthesized by anodization and the in situ growth method on the surface of an Al alloy substrate at different Zn2+ concentrations (0.05 mol/L, 0.10 mol/L, 0.20 mol/L, and 0.30 mol/L). The corrosion resistance and tribological properties of ZnAl-LDHs/AAO composite coatings were tested and analyzed. The effect of Zn2+ concentration on the anti-corrosion property and tribological behavior of ZnAl-LDHs/AAO composite coatings was discussed. The key conclusions are summarized as follows:
The prepared ZnAl-LDH nanosheets are all perpendicular to the AAO substrate, and the resulting ZnAl-LDHs/AAO composite coatings become uniform and compact with the increased Zn2+ concentration, but some “flower-like” ZnAl-LDH platelets can be found on the surface with an Zn2+ concentration of 0.2 mol/L. Meanwhile, the thickness and interfacial bonding strength of ZnAl-LDH films are both enhanced with the increase in Zn2+ concentration.
All ZnAl-LDHs/AAO composite-coated samples have improved corrosion resistance, friction-reduction, and anti-wear properties. The sample with the highest Zn2+ concentration possesses better corrosion resistance and acceptable tribological properties, especially 0.30 mol/L, which exhibits the best protected action for the Al alloy substrate.
The sliding pairs change from substrate/steel ball to ZnAl-LDH films/transferred films, and a lubricating interface formed here enhances the lubricity for the contact model. Moreover, the increased bonding force between the ZnAl-LDHs and AAO substrate suppresses the excessive removal of ZnAl-LDH films, which provides good durability of lubrication.

Author Contributions

Conceptualization, L.Y.; Methodology, L.Y. and C.N.; Software, C.Z., Z.S., C.N., R.L., Z.W. and Y.W.; Validation, Z.S.; Writing—original draft, L.Y. and C.Z.; Writing—review & editing, C.Z. All authors have read and agreed to the published version of the manuscript.

Funding

The research was supported by the National Natural Science Foundation of China (No. 52075112), the National Key R & D Program of China for Young Scientists (No. 2021YFB2011200), and the Major Scientific and Technological Project of Heilongjiang Province for Ten Million Project (No. 2021ZX05A03).

Institutional Review Board Statement

This study does not involve experiments onhuman tissue.

Data Availability Statement

The original codes and data used to support the findings of this study are available from the corresponding author upon request.

Conflicts of Interest

The authors declare no potential conflict of interest with respect to the research, authorship, and/or publication of this article.

References

  1. Ma, Z.; Feng, A.; Chen, D.; Shen, J. Recent advances in friction stir welding/processing of aluminum alloys: Microstructural avolution and mechanical properties. Crit. Rev. Solid State 2018, 43, 269–333. [Google Scholar] [CrossRef]
  2. Karthik, G.M.; Kim, H.S. Heterogeneous aspects of additive manufactured metallic parts: A review. Met. Mater. Int. 2021, 27, 1–39. [Google Scholar] [CrossRef]
  3. Guo, F.; Wu, S.; Liu, J.; Wu, Z.; Fu, S.; Ding, S. A time-domain stepwise fatigue assessment to bridge small-scale fracture mechanics with large-scale system dynamics for high-speed maglev lightweight bogies. Eng. Fract. Mech. 2021, 248, 107711. [Google Scholar] [CrossRef]
  4. Zhang, F.; Zhao, L.; Chen, H.; Xu, S.; Evans, D.; Duan, X. Corrosion resistance of superhydrophobic layered double hydroxide films on aluminum. Angew. Chem. Int. Ed. 2008, 47, 2466–2469. [Google Scholar] [CrossRef]
  5. Bai, B.N.P.; Ramasesh, B.S.; Surappa, M.K. Dry sliding wear of A356-Al-SiCp somposites. Wear 1992, 157, 295–304. [Google Scholar] [CrossRef]
  6. Li, S.; Yao, W.; Liu, J.; Yu, M.; Ma, K. Effect of SiC nanoparticle concentration on the properties of oxide films formed on Ti-10V-2Fe-3Al alloy. Vacuum 2016, 123, 1–7. [Google Scholar] [CrossRef]
  7. Li, S.; Yao, W.; Liu, J.; Yu, M.; Wu, L.; Ma, K. Study on anodic oxidation process and property of composite film formed on Ti-10V-2Fe-3Al alloy in SiC nanoparticle suspension. Surf. Coat. Technol. 2015, 277, 234–241. [Google Scholar] [CrossRef]
  8. Cui, L.; Gao, S.; Li, P.; Zeng, R.; Zhang, F.; Li, S.; Han, E. Corrosion resistance of a self-healing micro-arc oxidation/polymethyltrimethoxysilane composite coating on magnesium alloy AZ31. Corros. Sci. 2017, 118, 84–95. [Google Scholar] [CrossRef]
  9. Lu, X.; Blawert, C.; Tolnai, D.; Subroto, T.; Kainer, K.; Zhang, T.; Wang, F.; Zheludkevich, M. 3D reconstruction of plasma electrolytic oxidation coatings on Mg alloy via synchrotron radiation tomography. Corros. Sci. 2018, 139, 395–402. [Google Scholar] [CrossRef]
  10. Totaro, P.; Khusid, B. Multistep anodization of 7075-T6 aluminum alloy. Surf. Coat. Technol. 2021, 421, 127407. [Google Scholar] [CrossRef]
  11. Wu, L.; Ding, X.; Zheng, Z.; Ma, Y.; Atrens, A.; Chen, X.; Xie, Z.; Sun, D.; Pan, F. Fabrication and characterization of an actively protective Mg-Al LDHs/Al2O3 composite coating on magnesium alloy AZ31. Appl. Surf. Sci. 2019, 487, 558–568. [Google Scholar] [CrossRef]
  12. Wu, L.; Liu, L.; Zhan, Q.; Wen, C.; Hao, X.; Liu, J.; Chen, Y.; Atrens, A. Morphology, microstructure and tribological properties of anodic films formed on Ti10V2Fe3Al alloy in different electrolytes. Rare Met. 2021, 40, 2905–2916. [Google Scholar] [CrossRef]
  13. Zaraska, L.; Sulka, G.; Szeremeta, J.; Jaskula, M. Porous anodic alumina formed by anodization of aluminum alloy (AA1050) and high purity aluminum. Electrochim. Acta 2010, 55, 4377–4386. [Google Scholar] [CrossRef]
  14. Kikuchi, T.; Taniguchi, T.; Suzuki, R.; Natsui, S. Fabrication of a plasma electrolytic oxidation/anodic aluminum oxide multi-layer film via one-step anodizing aluminum in ammonium carbonate. Thin Solid Film. 2020, 697, 137799. [Google Scholar] [CrossRef]
  15. Bensalah, W.; Feki, M.; Wery, M.; Ayedi, H. Chemical dissolution resistance of anodic oxide layers formed on aluminum. Trans. Nonferrous Met. Soc. China 2011, 21, 1673–1679. [Google Scholar] [CrossRef]
  16. Montero-Moreno, J.M.; Sarret, M.; Muller, C. Influence of the aluminum surface on the final results of a two-step anodizing. Surf. Coat. Technol. 2007, 201, 6352–6357. [Google Scholar] [CrossRef]
  17. Ko, C.; Kuo, Y.; Chen, S.; Chen, S.; Guo, J.; Wang, Y. Formation of aluminum composite passive film on magnesium alloy by integrating sputtering and anodic aluminum oxidation processes. Thin Solid Films 2020, 709, 138151. [Google Scholar] [CrossRef]
  18. Sul, Y.; Johansson, C.; Jeong, Y.; Albrektsson, T. The electrochemical oxide growth behaviour on titanium in acid and alkaline electrolytes. Med. Eng. Phys. 2001, 23, 329–346. [Google Scholar] [CrossRef]
  19. Song, H.; Park, S.; Jeong, S.; Park, Y. Surface characteristics and bioactivity of oxide films formed by anodic spark oxidation on titanium in different electrolytes. J. Mater. Process. Technol. 2009, 209, 864–870. [Google Scholar] [CrossRef]
  20. Liu, L.; Wu, L.; Chen, X.; Sun, D.; Chen, Y.; Zhang, G.; Ding, X.; Pan, F. Enhanced protective coatings on Ti-10V-2Fe-3Al alloy through anodizing and post-sealing with layered double hydroxides. J. Mater. Sci. Technol. 2020, 37, 104–113. [Google Scholar] [CrossRef]
  21. Rahimi, S.; Khiabani, A.B.; Yarmand, B.; Kolahi, A. Comparison of corrosion and antibacterial properties of Al alloy treated by plasma electrolytic oxidation and anodizing methods. Mater. Today-Proc. 2018, 5, 15667–15676. [Google Scholar] [CrossRef]
  22. Zhang, J.; Song, B.; Wei, Q.; Bourell, D.; Shi, Y. A review of selective laser melting of aluminum alloys: Processing, microstructure, property and developing trends. J. Mater. Sci. Technol. 2019, 35, 270–284. [Google Scholar] [CrossRef]
  23. Feng, Y.; Williams, G.; Leroux, F.; Taviot-Gueho, C.; O’Hare, D. Selective anion-exchange properties of second-stage layered double hydroxide heterostructures. Chem. Mater. 2006, 18, 4312–4318. [Google Scholar] [CrossRef]
  24. Wu, L.; Zhang, G.; Tang, A.; Liu, Y.; Atrens, A.; Pan, F. Communication-fabrication of protective layered double hydroxide films by conversion of anodic films onmagnesium alloy. J. Electrochem. Soc. 2017, 164, 339–341. [Google Scholar] [CrossRef]
  25. Zhang, G.; Wu, L.; Tang, A.; Ma, Y.; Song, G.; Zheng, D.; Jiang, B.; Atrens, A.; Pan, F. Active corrosion protection by a smart coating based on a MgAl-layered double hydroxide on a cerium-modified plasma electrolytic oxidation coating on Mg alloy AZ31. Corros. Sci. 2018, 139, 370–382. [Google Scholar] [CrossRef]
  26. Wu, L.; Yang, D.; Zhang, G.; Zhang, Z.; Zhang, S.; Tang, A.; Pan, F. Fabricationand characterization of Mg-M layered double hydroxide films on anodized magnesium alloy AZ31. Appl. Surf. Sci. 2018, 431, 77–186. [Google Scholar] [CrossRef]
  27. Zhang, G.; Wu, L.; Tang, A.; Weng, B.; Atrens, A.; Ma, S.; Liu, L.; Pan, F. Sealing of anodized magnesium alloy AZ31 with MgAl layered double hydroxides layers. RSC Adv. 2018, 8, 2248–2259. [Google Scholar] [CrossRef]
  28. Wang, C.; Huang, Y.; Li, J.; Wang, M.; Du, X.; Chen, D. Preparation of superhydrophobic Li-Al-Ala LDH/SA film with enhanced corrosion resistance and mechanical stability on AZ91D Mg alloy. J. Mater. Sci. 2022, 57, 14780–14798. [Google Scholar] [CrossRef]
  29. Wang, H.; Liu, Y.; Liu, W.; Wang, R.; Wen, J.; Sheng, H.; Peng, J.; Erdemir, A.; Luo, J. Tribological behavior of NiAI-Layered double hydroxide nanoplatelets as oil-based lubricant additives. ACS Appl. Mater. Interfaces 2017, 9, 30891–30899. [Google Scholar] [CrossRef]
  30. Wang, H.; Liu, Y.; Liu, W.; Liu, Y.; Wang, K.; Li, J.; Ma, T.; Eryilmaz, O.; Shi, Y.; Erdemir, A.; et al. Superlubricity of polyalkylene glycol aqueous solutions enabled by ultrathin layered double hydroxide nanosheets. ACS Appl. Mater. Interfaces 2019, 11, 20249–20256. [Google Scholar] [CrossRef]
  31. Zhao, S.; Tsen, W.; Hu, F.; Zhong, F.; Liu, H.; Wen, S.; Zheng, G.; Qin, C.; Gong, C. Layered double hydroxide-coated silica nanospheres with 3D architecture-modified composite anion exchange membranes for fuel cell applications. J. Mater. Sci. 2020, 55, 2967–2983. [Google Scholar] [CrossRef]
  32. Zhang, Y.; Wuu, D.; Huang, J. Preparation of AgNWs@NiO-Co3O4 dopant material for an activated carbon thin-film electrode of pseudocapacitors. J. Mater. Sci. 2021, 56, 15229–15240. [Google Scholar]
  33. Zhang, Z.; Zhou, J.; Wei, H.; Dai, Y.; Li, S.; Shi, H.; Xu, G. Construction of hierarchical NiFe-LDH/FeCoS2/CFC composites as efficient bifunctional electrocatalysts for hydrogen and oxygen evolution reaction. J. Mater. Sci. 2022, 55, 16625–16640. [Google Scholar] [CrossRef]
  34. Qi, J.; Ruan, C.; Hu, R.; Sui, Y.; He, Y.; Meng, Q.; Wei, F.; Ren, Y.; Wei, W. Flexible wire-shaped symmetric supercapacitors with Zn-Co layered double hydroxide nanosheets grown on Ag-Coated cotton wire. J. Mater. Sci. 2022, 55, 16683–16696. [Google Scholar]
  35. Ding, P.; Li, Z.; Wang, Q.; Zhang, X.; Tang, S.; Song, N.; Shi, L. In situ growth of layered double hydroxide films under dynamic processes: Influence of metal cations. Mater. Lett. 2012, 77, 1–3. [Google Scholar] [CrossRef]
  36. Xue, L.; Lu, Z.; Cheng, Y.; Sun, X.; Lin, H.; Xiao, X.; Liu, X.; Zhuo, S. Three-dimensional layered double hydroxide membranes: Fabrication technique, growth mechanism, and enhanced photocatalytic activity. Chem. Commun. 2018, 54, 8494–8497. [Google Scholar] [CrossRef]
  37. Liu, J.; Shi, H.; Yu, M.; Du, R.; Rong, G.; Li, S. Effect of divalent metal ions on durability and anticorrosion performance of layered double hydroxides on anodized 2A12 aluminum alloy. Surf. Coat. Technol. 2019, 373, 56–64. [Google Scholar] [CrossRef]
  38. Shen, G.; Zhang, L.; Gu, Z.; Zheng, Z.; Liu, Y.; Tan, G.; Jie, X. Zinc aluminum-layered double hydroxide(LDH)-graphene oxide(GO) lubricating and corrosion-resistant composite coating on the surface of magnesium alloy. Surf. Coat. Technol. 2022, 437, 128354. [Google Scholar]
  39. Zhang, Y.; Wang, Y.; Li, C.; Zhou, B.; Cheng, K.; Wei, Y. Preparation and corrosion resistance of the ZnAl-LDHs film on 6061 Al alloy surface. Acta. Metall. Sin. 2018, 54, 1417–1427. [Google Scholar]
  40. Li, S.; Bhushan, B. Lubrication performance and mechanisms of Mg/Al-, Zn/Al-, and Zn/Mg/Al-layered double hydroxide nanoparticles as lubricant additives. Appl. Surf. Sci. 2016, 378, 308–319. [Google Scholar] [CrossRef]
  41. Wu, H.; Zhang, Y.; Long, S.; Zhang, L.; Jie, X. Tribological behavior of graphene anchored Mg-Al layered double hydroxide film on Mg alloy pre-sprayed Al coating. Appl. Surf. Sci. 2020, 530, 146536. [Google Scholar]
  42. Ying, L.; Xie, Y.; Fu, Z.; Shi, H.; Wang, G. Study on direct current electro-deposition of copper nanowires in anodic alumina membrane pores. Coatings 2019, 9, 378. [Google Scholar]
  43. Zhang, K.; Yu, S. Preparation of wear and corrosion resistant micro-arc oxidation coating on 7N01 aluminum alloy. Surf. Coat. Technol. 2020, 388, 125453. [Google Scholar] [CrossRef]
  44. Snihirova, D.; Lamaka, S.; Montemor, M. “SMART” protective ability of water based epoxy coatings loaded with CaCO3 microbeads impregnated with corrosion inhibitors applied on AA2024 substrates. Electrochim. Acta 2012, 83, 439–447. [Google Scholar] [CrossRef]
  45. Long, R.; Zhao, C.; Jin, Z.; Zhang, Y.; Pan, Z.; Sun, S.; Gao, W. Influence of groove dimensions on the tribological behavior of textured cylindrical roller thrust bearings under starved lubrication. Ind. Lubr. Tribol. 2021, 73, 971–979. [Google Scholar] [CrossRef]
  46. Zhao, C.; Long, R.; Zhang, Y.; Wang, Y.; Wang, Y. Influence of characteristic parameters on the tribological properties of vein-bionic textured cylindrical roller thrust bearings. Tribol. Int. 2022, 175, 107861. [Google Scholar] [CrossRef]
  47. Rosenkranz, A.; Stratmann, A.; Gachot, C.; Burghardt, G.; Jacobs, G.; Mucklich, F. Improved wear behavior of cylindrical roller thrust bearings by three-beam laser interference. Adv. Eng. Mater. 2016, 18, 854–862. [Google Scholar]
  48. Xie, H.; Cheng, Y.; Li, S.; Cao, J.; Cao, L. Wear and corrosion resistant coatings on surface of cast A356 aluminum alloy by plasma electrolytic oxidation in moderately concentrated aluminate electrolytes. Trans. Nonferrous Met. Soc. China 2017, 27, 336–351. [Google Scholar] [CrossRef]
  49. Martini, C.; Ceschini, L.; Tarterini, F.; Paillard, J.; Curran, J. PEO layers obtained from mixed aluminate-phosphate baths on Ti-6Al-4V: Dry sliding behaviour and influence of a PTFE topcoat. Wear 2010, 269, 747–756. [Google Scholar] [CrossRef]
  50. Soffritti, C.; Fortini, A.; Nastruzzi, A.; Sola, R.; Merlin, M.; Garagnani, G. Dry sliding behavior of an aluminum alloy after innovative hard anodizing treatments. Materials 2021, 14, 3281. [Google Scholar] [CrossRef]
  51. Kim, H.; Kim, D.; Ahn, H. Tribological properties of nanoporous anodic aluminum oxide film. Surf. Coat. Technol. 2010, 205, 1431–1437. [Google Scholar] [CrossRef]
  52. Meyer, E. Controlling friction atom by atom. Science 2015, 348, 1089. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic diagram for the manufacturing process of ZnAl-LDHs/AAO composite coatings on the surface of 1060 Al alloys.
Figure 1. Schematic diagram for the manufacturing process of ZnAl-LDHs/AAO composite coatings on the surface of 1060 Al alloys.
Coatings 13 01511 g001
Figure 2. XRD patterns (a) and FT-IR spectra (b) of ZnAl-LDHs/AAO composite coatings with different Zn2+ concentrations.
Figure 2. XRD patterns (a) and FT-IR spectra (b) of ZnAl-LDHs/AAO composite coatings with different Zn2+ concentrations.
Coatings 13 01511 g002
Figure 3. SEM surface micrographs and EDS analysis of ZnAl-LDHs/AAO composite coatings with different Zn2+ concentrations.
Figure 3. SEM surface micrographs and EDS analysis of ZnAl-LDHs/AAO composite coatings with different Zn2+ concentrations.
Coatings 13 01511 g003
Figure 4. Cross-sectional SEM micrographs of ZnAl-LDHs/AAO composite coatings with different Zn2+ concentrations.
Figure 4. Cross-sectional SEM micrographs of ZnAl-LDHs/AAO composite coatings with different Zn2+ concentrations.
Coatings 13 01511 g004
Figure 5. The hardness (a), and interfacial bonding force (b) of ZnAl-LDHs/AAO composite coatings with different Zn2+ concentrations.
Figure 5. The hardness (a), and interfacial bonding force (b) of ZnAl-LDHs/AAO composite coatings with different Zn2+ concentrations.
Coatings 13 01511 g005
Figure 6. Potentiodynamic polarization curves of the AAO and ZnAl-LDHs/AAO composite coatings with different Zn2+ concentrations.
Figure 6. Potentiodynamic polarization curves of the AAO and ZnAl-LDHs/AAO composite coatings with different Zn2+ concentrations.
Coatings 13 01511 g006
Figure 7. EIS curves of the AAO and ZnAl-LDHs/AAO composite coatings with different Zn2+ concentrations: Nyquist curve (a) and Bode plot (b).
Figure 7. EIS curves of the AAO and ZnAl-LDHs/AAO composite coatings with different Zn2+ concentrations: Nyquist curve (a) and Bode plot (b).
Coatings 13 01511 g007
Figure 8. Physical model of ZnAl-LDHs/AAO composite coatings (a) and corresponding equivalent circuit (b).
Figure 8. Physical model of ZnAl-LDHs/AAO composite coatings (a) and corresponding equivalent circuit (b).
Coatings 13 01511 g008
Figure 9. Coefficient of friction curves (a) and wear losses (b) of the AAO and ZnAl-LDHs/AAO composite coatings with different Zn2+ concentrations under dry wear.
Figure 9. Coefficient of friction curves (a) and wear losses (b) of the AAO and ZnAl-LDHs/AAO composite coatings with different Zn2+ concentrations under dry wear.
Coatings 13 01511 g009
Figure 10. SEM photographs and EDS data of worn surfaces for the ZnAl-LDHs/AAO composite coatings with different Zn2+ concentrations at low and high magnification: 0.05 mol/L (a,b); 0.10 mol/L (c,d); 0.20 mol/L (e,f); 0.30 mol/L (g,h).
Figure 10. SEM photographs and EDS data of worn surfaces for the ZnAl-LDHs/AAO composite coatings with different Zn2+ concentrations at low and high magnification: 0.05 mol/L (a,b); 0.10 mol/L (c,d); 0.20 mol/L (e,f); 0.30 mol/L (g,h).
Coatings 13 01511 g010
Figure 11. Model of friction-reduction and wear resistance mechanism of ZnAl-LDHs/AAO composite coatings.
Figure 11. Model of friction-reduction and wear resistance mechanism of ZnAl-LDHs/AAO composite coatings.
Coatings 13 01511 g011
Table 1. Potentiodynamic polarization parameters of different groups.
Table 1. Potentiodynamic polarization parameters of different groups.
GroupsEcorr (V vs. SCE)icorr (A/cm2)
AAO−0.7753.020 × 10−7
0.05 mol/L−0.6405.416 × 10−9
0.10 mol/L−0.5413.479 × 10−9
0.20 mol/L−0.4993.415 × 10−9
0.30 mol/L−0.4622.435 × 10−9
Table 2. Fitting results for EIS data of ZnAl-LDHs/AAO composite coatings.
Table 2. Fitting results for EIS data of ZnAl-LDHs/AAO composite coatings.
GroupsRsol (ohm)CLDHs (F)RLDHs (ohm)Cox (F)Rox (ohm)Cdl (F)Rct (ohm)χ2
AAO10.61--2.927 × 10−76.778 × 1035.953 × 10−81.345 × 1068.2 × 10−4
0.05 mol/L9.523.231 × 10−82.543 × 1035.010 × 10−82.451 × 1043.549 × 10−84.598 × 1063.6 × 10−3
0.10 mol/L8.041.576 × 10−84.396 × 1035.375 × 10−82.029 × 1042.721 × 10−85.527 × 1062.4 × 10−3
0.20 mol/L10.231.431 × 10−84.523 × 1036.678 × 10−81.297 × 1041.943 × 10−87.321 × 1062.8 × 10−3
0.30 mol/L9.641.303 × 10−86.041 × 1038.846 × 10−81.186 × 1041.209 × 10−81.289 × 1074.2 × 10−3
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

Ying, L.; Zhao, C.; Sun, Z.; Nie, C.; Liu, R.; Wang, Z.; Wu, Y. In Situ Growth of High-Performance ZnAl-Layered Double Hydroxides/Al2O3 Composite Coatings on Aluminum Alloy. Coatings 2023, 13, 1511. https://doi.org/10.3390/coatings13091511

AMA Style

Ying L, Zhao C, Sun Z, Nie C, Liu R, Wang Z, Wu Y. In Situ Growth of High-Performance ZnAl-Layered Double Hydroxides/Al2O3 Composite Coatings on Aluminum Alloy. Coatings. 2023; 13(9):1511. https://doi.org/10.3390/coatings13091511

Chicago/Turabian Style

Ying, Lixia, Chao Zhao, Zexian Sun, Chongyang Nie, Ruxin Liu, Zhiyong Wang, and Yunlong Wu. 2023. "In Situ Growth of High-Performance ZnAl-Layered Double Hydroxides/Al2O3 Composite Coatings on Aluminum Alloy" Coatings 13, no. 9: 1511. https://doi.org/10.3390/coatings13091511

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