Next Article in Journal
Research Progress on Tunable External Cavity Semiconductor Lasers in Visible and Near-Infrared Wavebands
Previous Article in Journal
Advances and Prospects of Nanomaterial Coatings in Optical Fiber Sensors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Y2O3 Particle Size on the Microstructure and Properties of Ni-Co-Y2O3 Composite Coatings

1
School of Mechanical Engineering, Shenyang University, Shenyang 110044, China
2
Key Laboratory of Research and Application of Multiple Hard Films of Liaoning Province, Shenyang University, Shenyang 110044, China
*
Author to whom correspondence should be addressed.
Coatings 2025, 15(9), 1009; https://doi.org/10.3390/coatings15091009
Submission received: 30 July 2025 / Revised: 22 August 2025 / Accepted: 25 August 2025 / Published: 1 September 2025

Abstract

In this study, Ni-Co-Y2O3 composite coating was prepared by electrodeposition, and the effect of Y2O3 particle size on the microstructure and properties of the coating was investigated. The samples were analyzed by XRD, SEM, AFM, EDS, cyclic voltammetry, XPS, hardness, and corrosion resistance test. The results indicate that the diffraction peak of the coating prepared with 50 nm particles exhibits reduced intensity and broadening, whereas the coating prepared with 100 nm particles displays a sharper and more pronounced peak. The onset reduction potential and the performance of the reduction reaction are influenced by particle size. When the particle size is 50 nm, the reduction process is less favorable, with an onset reduction potential of −0.9 V; in contrast, when the particle size is 100 nm, the reduction occurs more readily, with an onset reduction potential of −0.8 V. XPS analysis reveals that the chemical environment of elements varies with particle size. Regarding hardness, the coating prepared by combining different Y2O3 particle sizes exhibits higher hardness compared to that prepared using a single particle size, which can be attributed to the synergistic effect. In terms of corrosion resistance, the coating prepared with 100 nm Y2O3 particles demonstrates superior corrosion resistance, whereas the coating prepared with mixed particle sizes shows reduced stability and is more susceptible to corrosion. The coating prepared by mixing Y2O3 with particle size of 50 nm and 100 nm has a small friction coefficient. In summary, the particle size of Y2O3 has a significant influence on the microstructure, hardness, and corrosion resistance of Ni-Co-Y2O3 composite coatings.

1. Introduction

In the field of modern materials science, surface engineering technologies are rapidly advancing. As a key approach to surface modification, composite coatings have attracted extensive attention and have been the subject of in-depth research. Composite coatings prepared by electrodeposition exhibit smooth surface morphology, strong adhesion to various substrates, high hardness, excellent oxidation resistance, and superior corrosion resistance. These properties enable them to demonstrate broad application potential across multiple industrial sectors, including aerospace, automotive manufacturing, and electronic equipment [1,2,3,4,5,6,7].
Among them, Ni-Co coatings offer advantages such as high hardness, excellent wear resistance, and superior corrosion resistance, and have been widely applied across various industries. Cobalt atoms in the Ni-Co alloy coating substitute for nickel atoms in the lattice, causing lattice distortion and increasing the number of grain boundaries. Co addition refines the grain size and enhances the compactness of the composite coating. Consequently, the corrosion resistance and high-temperature stability of the composite coating are significantly improved compared to pure Ni coatings. Moreover, the Ni-Co structure exhibits hydrogen evolution catalytic activity [8,9,10,11]. However, the demand for enhanced performance to meet increasingly challenging applications has driven researchers to explore novel strategies for improving its properties. To enhance the performance of nickel-based coatings under harsh conditions, various types of ceramic reinforcement particles have been incorporated, including SiC [12], Y2O3 [13], SiO2 [14], ZrO2 [15], Al2O3 [16], WC [17], CNTs [18], CeO2 [19], and MoS2 [20].
Among these nanoparticles, Y2O3 serves as an additive in ceramic materials and can improve the grain boundary structure of ceramics, promoting uniform grain growth and thereby enhancing mechanical properties such as hardness, toughness, and wear resistance. Y2O3 offers multiple advantages in composite coatings. As a rare earth element compound, Y2O3 enhances the adhesion of oxide films and inhibits elemental interdiffusion through grain boundary segregation. Y2O3 promotes the formation of a protective oxide film, improves compactness, and exhibits excellent hot corrosion resistance, particularly in environments containing sulfur or salt fog [10]. Additionally, Y2O3 can act as a heterogeneous nucleation site during electrodeposition, refining grains and reducing electrochemical resistance more effectively than Al2O3 [16]. By refining the microstructure and reducing the number of active sites for corrosion microcells, Y2O3 significantly enhances corrosion resistance across various environments. In contrast, Al2O3 can lead to localized porosity, while SiO2 may dissolve in acidic conditions [14]. Moreover, Y2O3 possesses a high melting point, excellent high-temperature phase stability, and strong chemical inertness. In comparison, Al2O3 nanoparticles tend to agglomerate, and SiO2 may undergo phase transformations or form volatile oxides at elevated temperatures. As hard particles, Y2O3 has advantages over MoS2 in enhancing hardness [20].
In addition, Y2O3 is also widely used in the fabrication of high-temperature transparent ceramics, electronic ceramics, and structural ceramics. When incorporated into Ni-Co alloy coatings, Y2O3 can inhibit grain growth during the deposition process, promoting grain refinement and an increase in grain boundary area. This facilitates the formation of a new phase structure in the Ni-Co alloy, which hinders dislocation movement, thereby improving coating hardness and enhancing resistance to deformation [21,22,23,24].
Y. J. Xue et al. [25] electrodeposited a Ni-Y2O3 nanocomposite coating from a nickel sulfamate solution containing Y2O3 particles. The results showed that, compared with the pure Ni coating, the Ni-Y2O3 nanocomposite coating exhibited finer grains and enhanced corrosion resistance. Furthermore, the Ni-Y2O3 nanocomposite coating prepared under ultrasonic conditions displayed a denser microstructure, finer grains, and superior corrosion resistance. Xiaofang Yi et al. [26] prepared both Ni-Co composite coatings and Ni-Co-Y2O3 composite coatings via electrodeposition. The results indicate that the Ni-Co-Y2O3 composite coating exhibits higher electrocatalytic activity for the hydrogen evolution reaction compared to the Ni-Co composite coating. Y. B. Zhou [27] prepared a Ni-Y2O3 composite coating through the co-deposition of Ni and Y2O3 particles. Compared with the electrodeposited pure Ni coating, the Ni-Y2O3 composite coating exhibits superior corrosion resistance and alters the oxidation growth mechanism of nickel. Tian et al. [28] co-deposited nano-Y2O3 particles and Ni from a plating solution via electroplating to fabricate a Ni-Y2O3 composite coating. As the Y2O3 concentration increased from 20 g/L to 80 g/L, the Y2O3 content in the coating rose from 1.56% to 4.4%. The co-deposition of Y2O3 reduced the friction coefficient and wear weight loss. Under the experimental conditions, the coating exhibited optimal microhardness and tribological properties when the Y2O3 content reached 4.4%.
The performance of the coating is largely determined by its microstructure, and the particle size of the reinforcing phase is also a key influencing factor. So far, numerous studies have been conducted on Ni-Co coatings and Ni-Y2O3 composite coatings. However, limited research has focused on the influence of Y2O3 particle size on the microstructure and properties of the coating. Y2O3 particles of different sizes exhibit variations in specific surface area, surface energy, and reactivity, which directly affect their dispersion behavior in the plating solution. Meanwhile, the incorporation of particles with different sizes can alter the internal stress distribution and surface roughness of the coating, thereby influencing its overall performance. Smaller particles are more readily dispersed in the electrolyte through mechanical stirring or ultrasonic treatment, which helps reduce coating defects. However, if the particle size is excessively small, surface energy increases significantly, leading to particle agglomeration and the formation of micron-sized aggregates, which compromises coating uniformity. Conversely, excessively large particles may cause stress concentration at the coating–substrate interface, potentially resulting in delamination or spalling. Therefore, in this paper, Ni-Co-Y2O3 composite coatings were prepared by incorporating Y2O3 particles with individual particle sizes of 50 nm and 100 nm, as well as a 1:1 mixture of both sizes. The microstructure, surface properties, and electrochemical properties of the three types of Ni-Co-Y2O3 composite coatings were systematically analyzed using XPS, AFM, and electrochemical techniques. The influence of Y2O3 particle size on the microstructure and performance of the Ni-Co-Y2O3 composite coatings was investigated.

2. Preparation and Characterization

Ni-Co-Y2O3 composite coatings were prepared using an electrodeposition technique, with a copper sheet serving as the cathode substrate. The preparation steps and deposition process are illustrated in Figure 1. A 99% pure nickel sheet was utilized as the anode, maintaining a distance of 2 cm between the cathode and anode. The electrolyte composition and operating conditions for the preparation of Ni-Co-Y2O3 coatings are detailed in Table 1. Y2O3 nanoparticles were homogeneously dispersed in the solution through rotor agitation using a magnetic stirrer (DF-101S Suzhou Weier Laboratory Supplies Co., Ltd., Suzhou, China). The copper substrate was polished sequentially with 600#, 1200#, and 1500# sandpaper, followed by cleaning with deionized water after alkaline and acid washing treatments. Subsequently, the substrate was dried and set aside. This pretreatment process effectively enhances the adhesion strength between the coating and the substrate. After pretreatment, the samples were immersed in the solution for electrodeposition to prepare the Ni-Co-Y2O3 coating. The manufacturer of Y2O3 is Epure, with a purity of ≥ 99.9%. Mixed particle size group: 50 nm and 100 nm Y2O3 particles were pre-mixed in a 1:1 mass ratio, then ultrasonically dispersed to form a uniform mixed suspension. The suspension was added to the electrolyte at the same total concentration as the single particle size group, and the particles were kept in suspension in the electrolyte by magnetic stirring. The electrodeposition duration was 75 min, after which the prepared samples were rinsed with distilled water and dried using an air blower.
The surface morphology of the Ni-Co-Y2O3 composite coatings was investigated using a Hitachi S4800 scanning electron microscope (SEM; Hitachi, Tokyo, Japan). An energy dispersive X-ray detector (EDS; Oxford, UK) coupled with the SEM was employed to determine the chemical composition and Y2O3 content of the coating. The impact of varying particle sizes on the surface roughness of the deposited nanocoatings was assessed using atomic force microscopy (AFM; Dimension Icon). Additionally, the influence of different particle sizes on the phase structure of the deposited nanocoatings was analyzed using X-ray diffraction (XRD; D8 Advance; the ceramic closed tube is the core, and the Cu target is standard. Japanese scientific instrument manufacturers, Japan). The effect of particle size on the chemical composition of the deposited nanocoatings was further evaluated using X-ray photoelectron spectroscopy (XPS; Thermo Escalab 250 Xi). Vickers microhardness measurements were conducted using a 402 MVD hardometer (Shanghai Takezu Analytical Instruments Co., Ltd., Shanghai, China). The hardness test conditions included an objective lens magnification of 40×, a brightness value of 100, a loading pressure of 100 g held for 15 s, and the final hardness value of the composite coating was calculated as the average of five measurements. Electrochemical experiments were conducted using a three-electrode cell at the IVIUM electrochemical workstation (Ivium Technologies, Eindhoven, The Netherlands). The platinum wire (1 mm in diameter) was connected as the working electrode (WE), with 1 cm × 1 cm platinum plate as the counter electrode (CE) and saturated calomel electrode as the reference electrode (RE). Cyclic voltammograms were recorded at a potential scanning rate of 5 mV/s, with test voltages ranging from −2.5 V to 1.5 V and a step size of 10 mV. The corrosion resistance of the composite coatings was evaluated using both the polarization curve method and the AC impedance method. The AC impedance tests were performed in a 3.5% NaCl corrosive medium over a frequency range of 0.01 Hz to 100,000 Hz, maintaining a scanning rate of 5 mV/s. The polarization curves were scanned from potentials of −0.8 V to 0.2 V, while the remaining parameters were consistent with those described above.

3. Results and Discussion

3.1. Cyclic Voltammetry Tests

Figure 2 shows the cyclic voltammetry curves of three Ni-Co-Y2O3 composite coatings. The reduction potential of Ni-Co-Y2O3 (50 nm) coating is −0.9 V. The Y2O3 particles with a 50 nm particle size are smaller, reducing the effective reaction area, thereby increasing the resistance of electron transfer, resulting in the reduction reaction requiring a more negative potential to drive. In addition, particle agglomeration restricts the diffusion of reactant ions (Ni2+, Co2+), further increasing reaction resistance. The negative reduction potential (−0.9 V) indicates that the reduction reaction is more difficult and that the deposition process demands higher energy input. In contrast, the reduction potential of the Ni-Co-Y2O3 (100 nm) coating is −0.8 V. The larger Y2O3 particles at this size enhance the effective reaction area and facilitate electron conduction, thereby reducing electron transfer resistance. The increased particle spacing also improves the diffusion of reactants and mitigates diffusion limitations [29].

3.2. Phase Analysis

Figure 3 shows the XRD patterns of Ni-Co-Y2O3 composite coatings (50 nm, 100 nm, and 50 nm and 100 nm mixed) prepared by different particle sizes of Y2O3, and analyzes the crystal structure characteristics of the coating. The XRD pattern of the Ni-Co-Y2O3 composite coating was presented by comparing the diffraction peaks of the coating with the standard diffraction peaks compiled by the Joint Committee on Powder Diffraction Standards (JCPDS). The diffraction peaks corresponding to Ni, Co, and Y2O3 could be identified in the composite coating. The XRD pattern reflects the phase composition and crystal structure of the material by measuring the X-ray diffraction intensity at different 2θ angles. There is a weak peak between 40° and 45°, which corresponds to the diffraction peak of Y2O3. Moreover, there are phases formed by Ni and Co in the coating, and Y2O3 particles are successfully compounded. Compared with Ni-Co-Y2O3 (50 nm) coating, the change of Co/Ni peak position and strength may be caused by the change of stress state and crystal growth direction caused by the increase in coating thickness, which also indicates that Y2O3 is successfully compounded.

3.3. XPS Analysis

The spectral peak characteristics of Ni, Co, and Y elements in the coatings were systematically investigated and compared. Figure 4 presents the XPS survey spectrum of the composite coating. Figure 5 shows the XPS detection diagrams of the coatings prepared with 50 nm and 100 nm. Table 2 presents the content of each element. In the Ni 2p spectrum, the Ni 2p3/2 peaks for the Ni-Co-Y2O3 (50 nm) coating were observed at 852.83 eV and 873.35 eV, whereas those for the Ni-Co-Y2O3 (100 nm) coating were located at 851.93 eV and 873.06 eV, indicating a shift towards lower binding energy by approximately 0.9 eV. The peak intensity of the Ni-Co-Y2O3 (100 nm) coating is significantly higher, suggesting modifications in the chemical environment of Ni, variations in its relative content, or alterations in surface enrichment. The Co 2p3/2 peaks are observed at 779.94 eV and 795.68 eV in (a), and at 780.57 eV and 796.09 eV in (b). These peaks exhibit a shift towards lower binding energy, accompanied by an increase in peak intensity, which indicates changes in the valence state, content, or surface distribution of Co. The Y 3d5/2 peak is located at 158.03 eV in (a) and shifts to 158.21 eV in (b), corresponding to a shift towards higher binding energy by approximately 0.18 eV. Additionally, the peak intensity is enhanced, implying modifications in the chemical environment of Y, variations in its relative content, or an increase in surface distribution density [30,31].
As illustrated in Figure 6, in the Ni 2p spectrum, the Ni 2p3/2 peak is observed at approximately 852.32 eV and 872.44 eV, exhibiting a relatively weak intensity. A peak potentially associated with metallic nickel (Ni0) may be present at lower binding energies, indicating a low content of metallic Ni in the coating. The peak at 869.50 eV is potentially associated with Ni+, indicating that nickel in the coating exhibits multiple oxidation states. In the Co 2p spectrum, the Co 2p3/2 peak is observed at approximately 779 eV and 795 eV, with the peak at 778–780 eV being correlated with Co2+. The broadened peak suggests that cobalt exists in multiple chemical states or distinct crystal structures. During the growth of the coating, Co forms alloys or compounds with Ni and Y, leading to changes in the electronic structure, which are manifested as broad peaks in the XPS spectrum. In the Y 3d spectrum, the Y 3d5/2 peak is located near 158.40 eV and 160.48 eV, corresponding to Y3+ in Y2O3. The mixture of Y2O3 particles with sizes of 50 nm and 100 nm results in the broadening or shifting of the Y binding energy, and the variations in the local chemical environment of Y influence its distribution and chemical state.

3.4. Microstructure of the Plated Layer

Figure 7 shows the cross-sectional morphology of the coatings prepared with different particle sizes and mixed particle sizes. The thickness of the coating was 38 μm, 32 μm, and 33 μm, respectively. Figure 8 shows the surface morphology of the coatings prepared with different particle sizes and mixed particle sizes. At a magnification of 2000× and 5000×, the Ni-Co-Y2O3 (100 nm) coating exhibits uniformly distributed surface particles with consistent spacing, indicating that the nanoparticles were well dispersed during the deposition process. The EDS spectrum reveals that the coating primarily consists of Ni, Co, and Y elements, with Ni and Co present in higher concentrations and Y in a relatively lower amount. Since EDS cannot distinguish between nanoparticles of different sizes, only the total Y content can be determined. The Y content in the three composite coatings is 1.3 wt%, 1.4 wt%, and 1.7 wt%, respectively. These results demonstrate that during the electroplating process, metal ions and nanoparticles were uniformly deposited onto the substrate surface, and Y2O3 nanoparticles participated stably in the deposition reaction without significant agglomeration or loss.
Figure 9 is the AFM image of the composite coating. From the 3D image, the surface of Ni-Co-Y2O3 (50 nm) coating shows more peak–valley structures, and the height difference is relatively large. This may be due to the formation of relatively dense and irregular accumulation of 50 nm particles during the coating process. The surface of Ni-Co-Y2O3 (100 nm) coating is relatively flat, the peak–valley structure is relatively small and the height difference is small. The peak–valley structure on the surface of Ni-Co-Y2O3 (50 nm and 100 nm mixed) coating is very significant, and the height difference is also very large. This may be due to the complex microstructure formed by the mixing of 50 nm and 100 nm particles, and the interaction of particles with different sizes.

3.5. Hardness Analysis

Figure 10 is the hardness analysis histogram. Table 3 presents the hardness measurement data of the composite coating. It can be observed from the chart that the hardness of the Ni-Co-Y2O3 (50 nm and 100 nm mixed) coating is significantly higher than that of the other two coatings. In the Ni-Co-Y2O3 (50 nm) composite coating, the relatively small nanoparticle size leads to a higher proportion of grain boundaries. An excessive number of grain boundaries can impede dislocation movement, thereby increasing the hardness to some extent. When the nanoparticle size increases to 100 nm, a more favorable microstructure may form. This results in reduced local stress concentration and improved overall load-bearing capacity and hardness. The coating prepared with a mixture of particle sizes contains a larger number of 50 nm particles per unit volume, which can more effectively pin dislocation movement, hinder lattice slip, and provide high-frequency short-range obstacles [32]. During the electrodeposition process, 50 nm particles inhibit the local migration of grain boundaries, while 100 nm particles hinder the movement of large-scale grain boundaries. Under this synergistic effect, the grain structure is significantly refined [33]. From the perspective of nanoparticle synergy, the 50 nm particles effectively impede dislocation movement due to their small size, thereby enhancing grain boundary strengthening. Meanwhile, the 100 nm particles may improve stress distribution over a larger scale, enhance load transfer capability, and thus synergistically improve the overall hardness of the coating [34,35,36]. Table 4 and Table 5 present comparative data graphs for hardness. The hardness of Ni-Co-Y2O3 (50 nm and 100 nm mixed) coating is significantly higher than that of the coating prepared by Lin Shengnan [36].

3.6. Corrosion Resistance Analysis

Figure 11 shows the polarization curves of the three coatings. Table 6 presents the parameters of the polarization curve. Table 7 is the statistical table of corrosion results. Among them, the corrosion current density of Ni-Co-Y2O3 (100 nm) coating is relatively low and lower than the other two, the corrosion rate is slow and the corrosion resistance is good. The corrosion potential of the coating is relatively positive, which means that the coating is less prone to oxidation reaction in the corrosion system. The corrosion current density of Ni-Co-Y2O3 (50 nm and100 nm mixed) coating is relatively high, and the corrosion potential is relatively negative, which means that the coating is more prone to oxidation in the corrosion system, in a more unstable state, and more susceptible to corrosion [37,38].
Figure 12 presents the Nyquist plots of the three coatings. As shown in the diagram, the Ni-Co-Y2O3 (100 nm) coating exhibits a high charge transfer resistance, indicating that the electron transfer between the coating and the corrosive medium is significantly hindered, thereby effectively suppressing the corrosion reaction. This suggests that the Ni-Co-Y2O3 (100 nm) coating can efficiently prevent the electrochemical interaction between ions in the corrosive medium and the metallic components of the coating, thus enhancing its corrosion resistance. In contrast, the Ni-Co-Y2O3 (50 nm and 100 nm mixed) coating shows a lower charge transfer resistance in the Nyquist plot. During the corrosion process, electrons are more readily transferred between the coating and the corrosive medium, facilitating the corrosion reaction and resulting in relatively weaker corrosion resistance. When 50 nm and 100 nm Y2O3 particles are combined to prepare the coating, the significant difference in particle size makes it difficult to form a uniform microstructure during deposition. In a corrosive environment, the aggressive medium is more likely to penetrate through these non-uniform interfaces and structural defects, reaching the substrate surface and initiating a corrosion reaction.
Figure 13 presents the Bode plots of the three coatings. The Ni-Co-Y2O3 (100 nm) coating exhibits a large phase angle in the low-frequency region. A larger phase angle indicates stronger capacitive behavior, suggesting that the coating possesses significant capacitance characteristics. During the corrosion process, these characteristics can function similarly to a capacitor by storing and releasing charges, thereby suppressing the initiation of corrosion reactions. According to Table 8, the charge transfer resistance (Rct) of the Ni-Co-Y2O3 (100 nm) coating is 16,050 Ω, which is considerably higher than that of the other two coatings. This further supports the conclusion drawn from the Nyquist diagram—that a higher charge transfer resistance effectively inhibits the corrosion reaction. In contrast, the Ni-Co-Y2O3 (50 nm and 100 nm mixed) coating shows a relatively small phase angle in the low-frequency region, indicating weaker capacitance characteristics and a higher susceptibility to corrosion.

3.7. Frictional Wear

Figure 14 shows the friction and wear diagram of the three composite coatings. As shown in Figure 14, when the 50 nm and 100 nm Y2O3 particles are mixed, the prepared coating has a smaller friction coefficient. On the one hand, the surface microstructure is optimized. Small particles fill the gap of large particles to make the surface smoother and reduce the occlusion of surface protrusions during friction. As the strengthening phase, Y2O3 particles with different particle sizes work together to form a more effective strengthening network inside the coating. The larger particles provide a larger support skeleton, while the smaller particles hinder the movement of dislocations, making it more difficult for the coating material to deform plastically. At the same time, the particles with different particle sizes are evenly distributed in the Ni-Co matrix and support each other to form a denser structure. As a strengthening phase, the microstructure is refined and the dislocation movement is hindered by the synergistic effect, which makes the coating more difficult to deform and have higher hardness.

4. Conclusions

Ni-Co-Y2O3 composite coatings were successfully prepared by electrodeposition. The effects of Y2O3 particle size on the microstructure and properties of Ni-Co-Y2O3 coating were studied.
  • The particle size of Y2O3 has a significant influence on the reduction potential of the coating. XPS analysis reveals that the chemical environment of the elements changes with increasing particle size. Hardness testing indicates that the Ni-Co-Y2O3 (50 nm and 100 nm mixed) coating benefits from the synergistic effect between large and small particles, which effectively hinders dislocation movement and optimizes load transfer. As a result, its hardness is higher than that of coatings with a single particle size.
  • The Ni-Co-Y2O3 (50 nm and 100 nm mixed) coating showed poor stability and was prone to corrosion through the corrosion resistance test, and the particle adhesion was weak. The coating prepared by mixing Y2O3 with particles of 50 nm and 100 nm in size had a smaller friction coefficient.

Author Contributions

Data curation, investigation, writing—original draft, L.Q.; supervision, H.K.; investigation, T.Y.; writing—original draft preparation, J.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Applied Basic Research Program of Liaoning Provincial Department of Science and Technology of China, grant Number 2025JH2/101300035.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data that support the findings of this study are included within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bakonyi, I.; Hansal, W.E.G. Development of pulse-plating technology for the preparation of coatings with varying composition along their thickness: A historical overview. Trans. IMF 2018, 96, 237–243. [Google Scholar] [CrossRef]
  2. Aliofkhazraei, M.; Walsh, F.C.; Zangari, G.; Köçkar, H.; Alper, M.; Rizal, C.; Magagnin, L.; Protsenko, V.; Arunachalam, R.; Rezvanian, A.; et al. Development of electrodeposited multilayer coatings: A review of fabrication, microstructure, properties and applications. Appl. Surf. Sci. Adv. 2021, 6, 100141. [Google Scholar] [CrossRef]
  3. Li, B.; Li, D.; Chen, W.; Liu, Y.; Zhang, J.; Wei, Y.; Zhang, W.; Jia, W. Effect of current density and deposition time on microstructure and corrosion resistance of Ni-W/TiN nanocomposite coating. Ceram. Int. 2019, 45, 4870–4879. [Google Scholar] [CrossRef]
  4. Xia, F.; Li, Q.; Ma, C.; Guo, X. Preparation and characterization of Ni–AlN nanocoatings deposited by magnetic field assisted electrodeposition technique. Ceram. Int. 2020, 46, 2500–2509. [Google Scholar] [CrossRef]
  5. Xia, F.; Yan, P.; Ma, C.; Wang, B.; Liu, Y. Effect of different heat-treated temperatures upon structural and abrasive performance of Ni-TiN composite nanocoatings. J. Mater. Res. Technol. 2023, 27, 2874–2881. [Google Scholar] [CrossRef]
  6. Ma, C.; He, H.; Xia, F.; Xiao, Z.; Liu, Y. Performance of Ni–SiC composites deposited using magnetic-field-assisted electrodeposition under different magnetic-field directions. Ceram. Int. 2023, 49, 35907–35916. [Google Scholar] [CrossRef]
  7. Gao, M.; Liu, L.; Huang, Z.; Jiang, C.; Wang, X.; Huang, F.; Li, X.; Zhang, X. Electrodeposition and properties of Ni–W–Ti2AlC composite coatings. Ceram. Int. 2024, 50, 18832–18842. [Google Scholar] [CrossRef]
  8. Bakhit, B.; Akbari, A. Synthesis and characterization of Ni–Co/SiC nanocomposite coatings using sediment co-deposition technique. J. Alloys Compd. 2013, 560, 92–104. [Google Scholar] [CrossRef]
  9. Karimzadeh, A.; Rouhaghdam, A.S.; Aliofkhazraei, M.; Miresmaeili, R. Sliding wear behavior of Ni–Co–P multilayer coatings electrodeposited by pulse reverse method. Tribol. Int. 2020, 141, 105914. [Google Scholar] [CrossRef]
  10. Allahyarzadeh, M.H.; Aliofkhazraei, M.; Rouhaghdam, A.S.; Torabinejad, V. Electrochemical tailoring of ternary Ni-W-Co (Al2O3) nanocomposite using pulse reverse technique. J. Alloys Compd. 2017, 705, 788–800. [Google Scholar] [CrossRef]
  11. Ranjith, B.; Kalaignan, G.P. Ni–Co–TiO2 nanocomposite coating prepared by pulse and pulse reversal methods using acetate bath. Appl. Surf. Sci. 2010, 257, 42–47. [Google Scholar] [CrossRef]
  12. Wasekar, N.P.; Latha, S.M.; Ramakrishna, M.; Rao, D.S.; Sundararajan, G. Pulsed electrodeposition and mechanical properties of Ni-W/SiC nano-composite coatings. Mater. Des. 2016, 112, 140–150. [Google Scholar] [CrossRef]
  13. Zhang, R.; Li, Z.; Yu, X.; Cui, G. Characterisation and properties of Ni–W–Y2O3–ZrO2 nanocomposite coating. Surf. Eng. 2019, 35, 578–587. [Google Scholar] [CrossRef]
  14. Sadreddini, S.; Afshar, A. Corrosion resistance enhancement of Ni-P-nano SiO2 composite coatings on aluminum. Appl. Surf. Sci. 2014, 303, 125–130. [Google Scholar] [CrossRef]
  15. Beltowska-Lehman, E.; Bigos, A.; Indyka, P.; Chojnacka, A.; Drewienkiewicz, A.; Zimowski, S.; Kot, M.; Szczerba, M.J. Optimisation of the electrodeposition process of Ni-W/ZrO2 nanocomposites. J. Electroanal. Chem. 2018, 813, 39–51. [Google Scholar] [CrossRef]
  16. Li, B.; Zhang, W.; Huan, Y.; Dong, J. Synthesis and characterization of Ni-B/Al2O3 nanocomposite coating by electrodeposition using trimethylamine borane as boron precursor. Surf. Coat. Technol. 2018, 337, 186–197. [Google Scholar] [CrossRef]
  17. Elkhoshkhany, N.; Hafnway, A.; Khaled, A. Electrodeposition and corrosion behavior of nano-structured Ni-WC and Ni-Co-WC composite coating. J. Alloys Compd. 2017, 695, 1505–1514. [Google Scholar] [CrossRef]
  18. Firdouz, Z.; Tripathi, P.; Mondal, K.; Balani, K. Effect of carbonaceous reinforcements on anticorrosive and magnetic properties of Ni-Cu based composite coatings prepared by pulsed electrodeposition. Surf. Coat. Technol. 2022, 441, 128560. [Google Scholar] [CrossRef]
  19. Wang, C.; Shen, L.; Qiu, M.; Tian, Z.; Jiang, W. Characterizations of Ni-CeO2 nanocomposite coating by interlaced jet electrodeposition. J. Alloys Compd. 2017, 727, 269–277. [Google Scholar] [CrossRef]
  20. He, Y.; Wang, S.C.; Walsh, F.C.; Chiu, Y.L.; Reed, P.A.S. Self-lubricating Ni-P-MoS2 composite coatings. Surf. Coat. Technol. 2016, 307, 926–934. [Google Scholar] [CrossRef]
  21. Li, B.; Zhang, W.; Mei, T.; Miao, Y. Fabrication of Ni–B/TiC–Y2O3 nanocomposites by one-step electrodeposition at different duty cycle and evaluation of structural, surface and performance as protective coating. J. Alloys Compd. 2020, 823, 153888. [Google Scholar] [CrossRef]
  22. Yao, Q.L.; He, M.; Kong, Y.R.; Gui, T.; Lu, Z.H. Y2O3-functionalized graphene-immobilized Ni–Pt nanoparticles for enhanced hydrous hydrazine and hydrazine borane dehydrogenation. Rare Met. 2023, 42, 3410–3419. [Google Scholar] [CrossRef]
  23. Zhou, X.; Wang, Y.; Liang, Z.; Jin, H. Electrochemical deposition and nucleation/growth mechanism of Ni–Co–Y2O3 multiple coatings. Materials 2018, 11, 1124. [Google Scholar] [CrossRef] [PubMed]
  24. Zhang, Y.; Qian, S.Q.P.; Fu, X.F. Experimental study on friction-assisted electroforming of Ni-Co alloys. Mater. Res. Express 2023, 10, 026508. [Google Scholar] [CrossRef]
  25. Xue, Y.J.; Shen, C.; Li, J.S.; Li, H.; Si, D.H. Corrosion resistance of Ni-Y2O3 composite coating prepared by electrodeposition under ultrasonic condition. Adv. Mater. Res. 2010, 97, 1235–1238. [Google Scholar] [CrossRef]
  26. Yi, X.; He, P.; Chen, Y.; Wang, W.; Yang, D.; Dong, F. Y2O3-Modified Ni-Co Composite Coating as Cathode Materials for Hydrogen Evolution Reaction on Titanium Substrate. ECS Trans. 2010, 28, 13. [Google Scholar] [CrossRef]
  27. Zhou, Y.B.; Sun, J.F.; Wang, S.C.; Zhang, H.J. Oxidation of an electrodeposited Ni–Y2O3 composite film. Corros. Sci. 2012, 63, 351–357. [Google Scholar] [CrossRef]
  28. Tian, L.; Xu, J. Electrodeposition and characterization of Ni–Y2O3 composite. Appl. Surf. Sci. 2011, 257, 7615–7620. [Google Scholar] [CrossRef]
  29. Kan, H.M.; Meng, Y.Y.; Reddy, R.G. Influence of particle size and surfactants on uniformity and quantity of silicon carbide particles in electrodeposited nickel-silicon carbide coatings. J. Cent. South Univ. 2021, 28, 1627–1636. [Google Scholar] [CrossRef]
  30. Biesinger, M.C.; Lau, L.W.; Gerson, A.R.; Smart, R.S.C. Resolving surface chemical states in XPS analysis of first row transition metals, oxides and hydroxides: Sc, Ti, V, Cu and Zn. Appl. Surf. Sci. 2010, 257, 887–898. [Google Scholar] [CrossRef]
  31. Grey, L.H.; Nie, H.Y.; Biesinger, M.C. Defining the nature of adventitious carbon and improving its merit as a charge correction reference for XPS. Appl. Surf. Sci. 2024, 653, 159319. [Google Scholar] [CrossRef]
  32. Wu, G.; Li, N.; Zhou, D.; Cang, Z. The microstructure and high temperature properties of electrodeposited Co-Ni-Al2O3 composite coatings. J. Compos. Mater. 2004, 02, 8–13. [Google Scholar]
  33. Zhao, M.; Wang, J.; Li, H.; Yang, H.; Zhou, Y.; Li, J. Grain size effects on mechanical behavior of Al0.25CoCrFeNi high-entropy alloy at room and cryogenic temperatures. Intermetallics 2025, 183, 108792. [Google Scholar] [CrossRef]
  34. Akbarpour, M.R.; Asl, F.G.; Rashedi, H. Pulse-reverse electrodeposition of Ni-Co/graphene composite films with high hardness and electrochemical behaviour. Diam. Relat. Mater. 2023, 133, 109720. [Google Scholar] [CrossRef]
  35. Karmakar, R.; Maji, P.; Ghosh, S.K. A review on the nickel based metal matrix composite coating. Met. Mater. Int. 2021, 27, 2134–2145. [Google Scholar] [CrossRef]
  36. Lin, S.N.; Niu, L.P.; Wang, Y.Y.; Zhou, X.Y.; Liang, Z.P.; Liu, X.; Zhang, T.A. Orthogonal optimization of composite electrodeposition nickel-cobalt-yttrium oxide process in sulfamate system. Plat. Finish. 2020, 39, 532–535. [Google Scholar]
  37. Raghavendra, C.R.; Basavarajappa, S.; Sogalad, I.; Kumar, S. A review on Ni based nano composite coatings. Mater. Today Proc. 2021, 39, 6–16. [Google Scholar] [CrossRef]
  38. Shozib, I.A.; Ahmad, A.; Abdul-Rani, A.M.; Beheshti, M.; Aliyu, A.A.A. A review on the corrosion resistance of electroless Ni-P based composite coatings and electrochemical corrosion testing methods. Corros. Rev. 2022, 40, 1–37. [Google Scholar] [CrossRef]
Figure 1. Electrodeposition coating process diagram.
Figure 1. Electrodeposition coating process diagram.
Coatings 15 01009 g001
Figure 2. Cyclic voltammograms of electrolytes with different particle sizes.
Figure 2. Cyclic voltammograms of electrolytes with different particle sizes.
Coatings 15 01009 g002
Figure 3. XRD patterns of coatings prepared by Y2O3 with different particle sizes.
Figure 3. XRD patterns of coatings prepared by Y2O3 with different particle sizes.
Coatings 15 01009 g003
Figure 4. XPS survey spectrum of the composite coating.
Figure 4. XPS survey spectrum of the composite coating.
Coatings 15 01009 g004
Figure 5. XPS test of composite coatings: (a) 50 nm Y2O3, (b) 100 nm Y2O3.
Figure 5. XPS test of composite coatings: (a) 50 nm Y2O3, (b) 100 nm Y2O3.
Coatings 15 01009 g005
Figure 6. XPS test of Ni-Co-Y2O3 (50 nm and 100 nm mixed) coatings.
Figure 6. XPS test of Ni-Co-Y2O3 (50 nm and 100 nm mixed) coatings.
Coatings 15 01009 g006
Figure 7. Sections of the three coatings: (a) 50 nm Y2O3, (b) 100 nm Y2O3, (c) 50 nm and 100 nm Y2O3 mixture.
Figure 7. Sections of the three coatings: (a) 50 nm Y2O3, (b) 100 nm Y2O3, (c) 50 nm and 100 nm Y2O3 mixture.
Coatings 15 01009 g007
Figure 8. SEM and EDS analysis of the three coatings: (a,d) 50 nm Y2O3, (b,e) 100 nm Y2O3, (c,f) 50 nm and 100 nm Y2O3 mixture.
Figure 8. SEM and EDS analysis of the three coatings: (a,d) 50 nm Y2O3, (b,e) 100 nm Y2O3, (c,f) 50 nm and 100 nm Y2O3 mixture.
Coatings 15 01009 g008
Figure 9. AFM images of composite coatings: (a) 50 nm Y2O3, (b) 100 nm Y2O3, (c) 50 nm and 100 nm Y2O3 mixture.
Figure 9. AFM images of composite coatings: (a) 50 nm Y2O3, (b) 100 nm Y2O3, (c) 50 nm and 100 nm Y2O3 mixture.
Coatings 15 01009 g009
Figure 10. Histogram of hardness analysis of three composite coatings.
Figure 10. Histogram of hardness analysis of three composite coatings.
Coatings 15 01009 g010
Figure 11. Polarization curves of three Ni-Co-Y2O3 coatings.
Figure 11. Polarization curves of three Ni-Co-Y2O3 coatings.
Coatings 15 01009 g011
Figure 12. Impedance curves of three Ni-Co-Y2O3 coatings.
Figure 12. Impedance curves of three Ni-Co-Y2O3 coatings.
Coatings 15 01009 g012
Figure 13. Bode plots of three Ni-Co-Y2O3 coatings.
Figure 13. Bode plots of three Ni-Co-Y2O3 coatings.
Coatings 15 01009 g013
Figure 14. Friction and wear diagram of composite coating.
Figure 14. Friction and wear diagram of composite coating.
Coatings 15 01009 g014
Table 1. Process conditions of electrodeposition coating preparation.
Table 1. Process conditions of electrodeposition coating preparation.
Electrolyte ComponentsConcentration (g/L)
NiSO4·7H2O300
NiCl2·6H2O80
CoSO4·7H2O15
H3BO340
SDS4
Y2O3 (50 nm)6
Y2O3 (100 nm)6
Y2O3 (50 nm and 100 nm mixed)6
Deposition parametersAmount
Solution pH4.3
Temperature (°C)55
Current density (A/dm2)4
Magnetic agitation rate (rpm)350
Electrodeposition cell voltage (V)1.1
Deposition time (minute)75
Bath volume (mL)100
Y2O3 particle size50 nm; 100 nm
Table 2. The content of each element.
Table 2. The content of each element.
SampleY 3dC 1sO 1sCo 2pNi 2p
50 nm0.7569.9270.731.63
100 nm0.9864.4231.071.072.47
mixed1.8156.1434.251.875.92
Table 3. Hardness measurement results data of composite coatings.
Table 3. Hardness measurement results data of composite coatings.
Sample1
Hardness (HV)
2
Hardness (HV)
3
Hardness (HV)
4
Hardness (HV)
5
Hardness (HV)
50 nm400.7402.5404.1406.3400.4
100 nm509.2510.4511.3511.3510.0
mixed675.8679.2683.5687.1675.9
Table 4. Comparison table of hardness data.
Table 4. Comparison table of hardness data.
Y2O3 (size)50 nm50 nm100 nmmixed
Hardness (HV)538.85 [36]402.8510.6680.3
Table 5. Hardness data statistics chart.
Table 5. Hardness data statistics chart.
StatisticsError BarsStandard DeviationStatistical Significance
50 nm1.102.46p < 0.0001
100 nm0.5051.13p < 0.001
mixed2.214.93p < 0.0001
Table 6. Polarization curve parameter table.
Table 6. Polarization curve parameter table.
SamplesEcorr (V)Icorr (A/cm2)βa (V/dec)−βc (V/dec)Corrosion Rate
(mm/y)
50 nm−0.482.036 × 10−582.496.40.0137
100 nm−0.468.719 × 10−663.582.60.0091
mixed−0.512.381 × 10−575.990.70.0106
Table 7. Statistical table of corrosion results.
Table 7. Statistical table of corrosion results.
StatisticsError BarsStandard Deviation
data0.01460.0252
Table 8. Electrochemical analysis data.
Table 8. Electrochemical analysis data.
SamplesRS (Ω)CPEf (μF.cm−2)Rf (Ω)CPEdl (μF.cm−2)Rct (Ω)χ2
50 nm5.587128.24163292.34286≈0.002
100 nm5.06167.47862038716050≈0.004
mixed5.297106.23213130.89458≈0.005
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

Qi, L.; Kan, H.; Yue, T.; Wu, J. Effect of Y2O3 Particle Size on the Microstructure and Properties of Ni-Co-Y2O3 Composite Coatings. Coatings 2025, 15, 1009. https://doi.org/10.3390/coatings15091009

AMA Style

Qi L, Kan H, Yue T, Wu J. Effect of Y2O3 Particle Size on the Microstructure and Properties of Ni-Co-Y2O3 Composite Coatings. Coatings. 2025; 15(9):1009. https://doi.org/10.3390/coatings15091009

Chicago/Turabian Style

Qi, Linxin, Hongmin Kan, Tingting Yue, and Jiang Wu. 2025. "Effect of Y2O3 Particle Size on the Microstructure and Properties of Ni-Co-Y2O3 Composite Coatings" Coatings 15, no. 9: 1009. https://doi.org/10.3390/coatings15091009

APA Style

Qi, L., Kan, H., Yue, T., & Wu, J. (2025). Effect of Y2O3 Particle Size on the Microstructure and Properties of Ni-Co-Y2O3 Composite Coatings. Coatings, 15(9), 1009. https://doi.org/10.3390/coatings15091009

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