1. Introduction
Electrodeposition has evolved into a sophisticated methodology for engineering composite coatings through strategic material integration, offering cost-effective solutions for combating corrosion and wear in industrial applications [
1]. Conventional metallic coatings like pure cobalt or nickel provide foundational protection via dense metal matrices [
2,
3], while modern composite electrodeposition extends functionality through synergistic combinations of metallic alloys and ceramic reinforcement phases [
4,
5,
6]. This technological progression addresses escalating demands from sectors including offshore energy systems and aerospace components, where surfaces endure simultaneous mechanical stress and corrosive attacks.
The architecture of advanced composite coatings follows two primary design paradigms—(1) alloying strategically selected metals (e.g., Ni-Co systems for enhanced hardness through solid solution strengthening [
7]), and (2) incorporating secondary-phase particles to establish multifunctional material systems [
8]. Among potential additives, ceramic nanoparticles (NPs) such as SiC, Al
2O
3, and SiO
2 demonstrate particular promise due to their exceptional hardness (>25 GPa for SiC), thermal stability (>1600 °C decomposition temperature), and electrochemical inertness [
3,
9]. Notable examples include Ni/Al
2O
3 composites, which achieve twice the wear resistance to that of pure nickel coatings [
4], as well as Ni/SiC composites that show a 60% lower corrosion current density compared to pure nickel in electrochemical tests [
10]. By adding βSiC to the Ni-Co composite coating, fine-grained strengthening and dispersion strengthening were achieved, increasing the hardness by 40% and reducing the friction coefficient by 57 [
11]. The Co-Ni-Ce/TiC coating achieves a wear resistance five times that of the copper substrate through grain refinement (50.6 nm) and the lubrication effect of TiC, providing a new solution for the protection of copper molds [
12].
Despite these advantages, the practical implementation of nanoparticle-reinforced coatings confronts a fundamental challenge—the inherent tendency of nanoscale additives to agglomerate due to high surface energy [
13]. Such agglomeration of nanoparticles can lead to microstructural heterogeneity, potentially forming localized micron-scale clusters and weakening matrix–particle interfacial bonding [
14], which collectively contributes to the degradation of coating mechanical integrity. Conventional dispersion strategies exhibit inherent drawbacks. Mechanical agitation merely achieves macroscopic suspension but fails to suppress the Brownian agglomeration of nanoclusters [
15], while surfactant additives often induce residual tension stresses within composite coatings, impairing mechanical integrity [
16]. Although pulsed magnetic fields show particle alignment potential [
17], the requisite equipment complexity undermines industrial scalability.
Recent advances in surface chemistry provide alternative approaches. Binary surfactant systems (Span80–Tween60) reduce SiC agglomerate size by approximately 66% through steric stabilization [
10]. Residual organic layers from surfactants weaken interfacial bonding strength by forming thermodynamically unstable interfaces, which accelerate environmental degradation (e.g., water diffusion) and reduce mechanical integrity [
18,
19]. Laser surface remelting enhances particle distribution homogeneity but promotes grain coarsening (e.g., columnar grains > 1–2 μm) [
20,
21], with elevated energy inputs risking thermal degradation in temperature-sensitive substrates [
22]. These limitations underscore the need for surface modification methods that simultaneously enhance dispersion stability and preserve interfacial compatibility.
The co-deposition of nanoparticles into metal layers during electrolysis emerges from the synergistic interplay of physical adsorption, mass transport, and electrochemical reactions. Nanoparticles transported via diffusion, migration, and convection mechanisms adsorb onto the cathode surface [
23], where they co-deposit with electrochemically reduced metal ions to form uniform nanocomposite coatings. Electrolyte flow dynamics—especially through stirring or electrode rotation—critically govern particle delivery kinetics and distribution uniformity [
24]. Following initial physical adsorption, nanoparticles become incorporated into the evolving metal matrix through synchronized reduction–deposition processes. Crucially, while elevated current densities accelerate metal ion reduction, they may concurrently reduce nanoparticle embedding efficiency; however, extended deposition times substantially increase particle loading within the metal matrix [
25].
As shown in
Figure 1, this investigation presents an innovative chemical pretreatment protocol to optimize SiC nanoparticle dispersion in Co/SiC nanocomposite coatings. Through alkaline etching that removes oxide layers and tailors surface chemistry, we successfully engineered SiC surfaces with enhanced hydrophilicity and electrostatic stabilization. The resulting coatings exhibit synergistic enhancements including improved wear resistance, validating the method’s effectiveness for marine and aerospace applications.
Our approach avoids common pitfalls in nanoparticle-enhanced coatings, e.g., eliminating organic surfactants that degrade interfacial bonding through residual impurities, preserving matrix microstructure by employing calcining processing instead of high-energy methods, and enabling substrate-independent application through aqueous-phase chemistry. The combined surface engineering and electrochemical optimization strategy was systematically validated through multiscale characterization and mechanical testing, while industrial scalability was confirmed via competitive cost-performance metrics.
3. Results
Figure 2 presents the macroscopic morphology of SiC particles under three treatment conditions. Calcined samples exhibited compact agglomerates with dark grayish–blue coloration, distinct from the pale-gray loose aggregates of non-treated particles. The chromatic differentiation became markedly more pronounced upon DI water immersion, whereby treated variants developed a greenish–gray slurry through rapid hydration, while non-treated particles maintained buoyancy with a characteristic floating behavior. This chromatic variation likely indicates enhanced phase purity through oxide layer removal, as corroborated by subsequent EDS analysis. Notably, the treated particles demonstrated superior aqueous wettability, revealing a pronounced improvement in hydrophilicity following surface activation via alkaline treatments.
Three types of nanoparticles were dispersed in deionized water at a ratio of 5 g/30 mL.
Figure 3 illustrates the morphological evolution of the nanoparticle suspensions under three treatment conditions—immediately after aqueous introduction (
Figure 3a), post-mechanical homogenization (
Figure 3b), and following 24 h of quiescent settling (
Figure 3c). As shown in
Figure 3a, untreated nanoparticles exhibited buoyancy at the water–air interface upon direct introduction, whereas both TAE-SiC and HE-SiC achieved homogeneous aqueous dispersion. Mechanical agitation (300 rpm for 15 min) was employed to disperse nanoparticles within the plating bath. Untreated particles demonstrated significant agglomeration along centrifuge tube walls (
Figure 3b), indicating poor colloidal stability. Post-sedimentation analysis (
Figure 3c) revealed that the untreated and TAE-SiC systems maintained a metastable dispersion after 24 h, whereas HE-SiC underwent incipient settling due to a reduced ζ-potential (−15 mV vs. −25 mV for TAE-SiC). Optical microscopy of the dried suspensions (
Figure 3d) confirmed macro-scale agglomeration in the untreated and HE-SiC systems, with clusters exceeding 1 μm in diameter. In contrast, TAE-SiC retained nanoscale dispersion (average cluster size: 300 nm), appearing optically transparent under bright-field imaging.
The zeta potential and particle size distribution of the suspensions were systematically characterized using a zeta potential analyzer (
Table 1 and
Figure 4; the colors in the bar graphs represent distinct measurement batches for visual differentiation, without additional significance). Size distribution analysis (
Figure 4a) revealed that untreated particles (Origin) formed polydisperse clusters (>1 μm), while HE-SiC particles showed monodisperse aggregates averaging 1.3 μm. TAE-SiC yielded nanoparticles with a mean hydrodynamic diameter of 300 nm (sub-optical resolution), producing visually homogeneous suspensions, as corroborated by macroscopic observations in
Figure 3. Zeta potential measurements (
Figure 4b) demonstrated a superior electrochemical stability for TAE-SiC (−25 mV), contrasting with the lower potentials of Origin-SiC (−15 mV) and HE-SiC (−15 mV).
The higher ζ potential magnitude of TAE-SiC indicates a stronger electrostatic repulsion at the cathode surface. To quantify the effect of the electric field on particle movement, the electrophoretic mobilities (μ_e) of both TAE-treated and untreated SiC were calculated from their respective ζ potentials using the Helmholtz–Smoluchowski equation, as follows: μ_e = (ε0εᵣ ζ)/η, where ε0 is the vacuum permittivity (8.854 × 10−12 F/m), εᵣ is the relative permittivity of water (78.5 at 25 °C), and η is the dynamic viscosity of water (8.91 × 10−4 Pa·s at 25 °C). The calculated μ_e values were −17.37 μm·cm/V·s and −10.42 μm·cm/V·s for the TAE-treated and untreated SiC particles, respectively. This significant difference confirms that the TAE treatment substantially enhanced the electrophoretic migration of the particles away from the cathode.
The enhanced dispersion stability of TAE-SiC stems from the following synergistic mechanisms: (1) a reduced particle mass (<300 nm) minimizes gravitational settling, and (2) an elevated surface charge (−25 mV) maximizes electrostatic repulsion through extended Derjaguin–Landau–Verwey–Overbeek (DLVO) interactions. Conversely, Origin particles’ paradoxical stability arises from hydrophobic surface-induced air encapsulation during agitation, generating buoyant hollow microclusters stabilized by Laplace pressure (
Figure 3d). This micromechanical stabilization mechanism compensates for their unfavorable electrokinetic profile.
Figure 5a presents the SEM micrographs of Origin-SiC, HE-SiC, and TAE-SiC, all exhibiting polyhedral morphologies with comparable size distributions (50–200 nm). EDS quantification (
Figure 5b) was detected ~15 at% oxygen in Origin-SiC and HE-SiC, confirming residual surface oxides. TAE-SiC exhibited near-complete oxide removal, verifying successful deoxidation via thermal activation. Despite the theoretical Si:C stoichiometric atomic ratio (1:1), all particles showed carbon enrichment (58–68 at% C vs. 32–38 at% Si), indicative of graphitic residues from SiC synthesis. TAE-SiC processing did not reduce the carbon content, confirming its inability to eliminate carbonaceous contaminants. The observed hydrophilicity enhancement primarily originated from oxide removal rather than carbon reduction, as surface hydroxyl groups from residual oxides impede water wettability.
The X-ray diffraction analysis of the three particle types (
Figure 6) revealed well-defined peaks matching β-SiC reference patterns. All samples exhibited characteristic β-SiC diffraction peaks at 35.7° (111), 60.0° (220), 71.8° (311), and 75.5° (222) 2θ. TAE-SiC showed sharper (111) and (311) diffraction peaks compared to the Origin-SiC and HE-SiC variants (
Figure 6), indicating enhanced crystallinity through thermal activation. Broad scattering features between 20 and 30° 2θ were observed in Origin-SiC and HE-SiC (
Figure 6 inset), suggesting residual amorphous Si-O surface species. The absence of amorphous scattering signals in TAE-SiC’s XRD profiles confirms effective oxide removal via alkaline calcination, directly correlating with their improved hydrophilicity (as demonstrated in
Figure 3).
The phase purity and surface chemistry of the SiC nanoparticles were further probed through X-ray photoelectron spectroscopy (XPS).
Figure 7 contrasts the wide-scan survey spectra of Origin-SiC and TAE-SiC, revealing distinct compositional shifts aligned with our EDS and XRD observations (
Figure 5 and
Figure 6). Critically, TAE-SiC exhibits a 43.6% reduction in oxygen atomic percent (12.93 at% vs. 22.92 at% for Origin-SiC;
Table 2), quantitatively confirming oxide removal via thermal–alkaline etching. Concomitantly, silicon content increased by 16.3% (37.41 at% vs. 32.17 at%), while carbon rose marginally (49.66 at% vs. 44.91 at%)—consistent with the preserved graphitic residues noted earlier.
High-resolution spectra (
Figure 8) elucidate the chemical state’s evolution. In Origin-SiC (
Figure 8a–c), the O 1s signal at 533.95 eV signifies stoichiometric SiO
2, while peaks at 532.65 eV (Si–O–C) and 533.28 eV (Si–O–Si) denote oxygen-bridged contaminants [
26]. The C 1s spectrum (
Figure 8b) confirms carbon heterogeneity, whereby the 283.68 eV peak corresponds to SiC-bonded carbon, the 284.90/284.86 eV peaks indicate sp
2/sp
3 carbon impurities, and the 289.52 eV peak reflects Si–O–C moieties [
27]. These dual oxygen/carbon contamination layers impede hydrophilicity and promote agglomeration (
Figure 3).
TAE-SiC (
Figure 8d–f) demonstrates markedly altered bonding. The O 1s signal (
Figure 8d) collapses at >533 eV, with residual peaks at 532.39 eV (Si–O–C) and 533.24 eV (substoichiometric Si–O
x, x < 2). The Si 2p spectrum (
Figure 8f) shows a dominant Si–C peak (101.76 eV) alongside a diminished feature at 102.96 eV, indicating ultrathin surface oxidation (<2 nm) rather than bulk SiO
2. Concurrently, C 1s (
Figure 8e) retains the SiC signature (283.88 eV) and graphitic carbon (284.98 eV), while the 286.01 eV peak (C–O) aligns with residual Si–O–C. Critically, the Si–O–Si/SiO
2 suppression correlates with the reduced FWHM for Si 2p (1.63 eV vs. 2.02 eV;
Table 2), evidencing improved crystallinity and surface homogeneity post-TAE.
TAE-SiC and Origin-SiC composites were prepared at 50 g/L loading in cobalt electrolyte (CoSO4·7H2O 240 g/L, H3BO3 30 g/L, pH 5.0). Notably, Origin-SiC required extended dispersion (10 h mechanical stirring at 300 rpm + 1 h ultrasonication) to achieve comparable colloidal stability with TAE-SiC (2 h stirring + 1 h ultrasonication), reflecting the improved hydrophilicity from surface oxide removal (as per XRD results).
Under varying current densities, the coatings exhibited a clustered agglomerate structure, characterized by dendritic flakes and near-spherical clusters. As shown in
Figure 9, Origin-SiC-enhanced coatings displayed loosely packed agglomerates with secondary dendritic flakes (1–3 μm) and nanosized surface particulates. At 2 A/dm
2, tightly bonded agglomerates (>10 μm diameter) exhibited negligible porosity. However, increasing the current density to 8 A/dm
2 reduced agglomerate size to <5 μm and increased surface porosity, accompanied by thickened dendritic flakes and spheroidized clusters. In contrast, TAE-SiC-enhanced coatings displayed a denser matrix with indistinct agglomerate boundaries at low current densities (<4 A/dm
2), where dendritic features appeared as prismatic protrusions. Elevated current densities (>6 A/dm
2) induced boundary definition, micron-particle attachment (~1 μm diameter), and visible porosity. During co-deposition, micro- and nanoparticles migrating to the cathode surface acted as heterogeneous nucleation sites for cobalt ion reduction, forming agglomerates via particle encapsulation. High current densities or low nucleation densities favored rapid cobalt growth along specific crystallographic orientations, generating dendritic flakes.
As shown in
Figure 10, energy-dispersive X-ray spectroscopy (EDS) revealed oxygen contamination (~1 at%) in coatings containing Origin-SiC, whereas TAE-SiC-enhanced coatings exhibited no detectable oxygen (<0.01 at%). However, TAE-SiC-enhanced coatings displayed a 34% reduction in silicon content (3.83 at% vs. 5.82 at% for Origin particles at 2 A/dm
2), indicating lower SiC incorporation efficiency. This discrepancy correlates with zeta potential measurements—TAE-SiC exhibited enhanced surface charge (−25 mV) compared to Origin-SiC (−15 mV), generating a stronger electrostatic repulsion within the electric field. Consequently, TAE-SiC migrated away from the cathode during electrodeposition, reducing near-electrode particle concentration by ~30% and diminishing co-deposition efficiency. At current densities exceeding 4 A/dm
2, rapid cobalt deposition (0.56 μm/min at 4A/dm
2; 1.24 μm/min at 8 A/dm
2) outpriced particle migration, enabling the hydrodynamic entrapment of TAE-SiC near the cathode. This resulted in a 120% increase in SiC content (5.9 at% at 8 A/dm
2; 2.68 at% at 4 A/dm
2), as metallic encapsulation overrode electrostatic effects.
Figure 11 displays cross-sectional SEM micrographs of TAE-SiC-enhanced cobalt-based coatings, demonstrating that increasing current density from 4 to 6 A/dm
2 promotes both deposition rate and particle incorporation efficiency. Electrophoretic deposition dominated the particle transport process, as supported by zeta potential measurements (−25 mV for TAE-SiC).
In contrast,
Figure 12 shows untreated (Origin-SiC) particles’ distribution in coatings deposited under various parameters (2 A/dm
2 for 1 h; 4–8 A/dm
2 for 0.5 h each). The untreated particles demonstrate current density-independent embedment behavior, with uniform distribution across tested conditions. This fundamental difference highlights the critical role of surface activation in achieving current-responsive particle incorporation behavior.
Figure 13 presents comparative surface topographies of TAE-SiC- and Origin-SiC-enhanced coatings after standardized tribological testing. Nanoscale reinforcements exhibit distinct wear characteristics compared to micron-sized analogs, with grain refinement correlating to more continuous abrasion grooves that are consistent with micro-plowing mechanisms. TAE-SiC-enhanced coatings show shorter average scratch lengths relative to untreated specimens, as measured from three SEM fields per sample. Surface morphological analysis reveals enhanced hook-shaped furrow formation in TAE-SiC-enhanced coatings, potentially indicative of interacting adhesive–abrasive mechanisms. While the recorded mass losses under ASTM G65 Method E conditions show minimal difference (1.6923 ± 0.0035 g vs. 1.7107 ± 0.0062 g), with corresponding wear rates of (2.037 ± 0.004) × 10
3 mm
3/(N·m) vs. (2.059 ± 0.007) × 10
3 mm
3/(N·m), repeated trials with larger sample sizes are recommended to confirm this trend. Notably, oxygen content variations between treated and untreated particles suggest that surface chemistry may mediate these wear responses, though direct causalities remain to be established.
The Tafel polarization analysis reveals that composite coatings with TAE-SiC exhibit a significantly enhanced corrosion resistance compared to their untreated counterparts (
Figure 14 and
Table 3). At current densities of 2–4 A/dm
2, the corrosion potential (Ecorr) of TAE-SiC-enhanced coatings positively shifted by 61–120 mV (e.g., −0.52 ± 0.003 V at 2 A/dm
2 vs. untreated −0.58 ± 0.003 V), indicating reduced thermodynamic corrosion susceptibility. Combined with SEM morphology (
Figure 9) and zeta potential analysis (
Figure 4), this improvement originates from TAE-induced oxide layer removal (EDS confirmed oxygen content elimination), which mitigates interfacial galvanic corrosion. Although the corrosion current density (Icorr) of TAE-SiC-enhanced coatings (1.81 × 10
−5~2.83 × 10
−5 A/cm
2) partially overlaps with that of untreated systems (0.96 × 10
−5~4.06 × 10
−5 A/cm
2), distinct mechanisms are identified through polarization resistance (Rp) and microstructures. At 2 A/dm
2, TAE-SiC-enhanced coatings achieve an Rp of 3.22 × 10
4 Ω·cm
2 (
Figure 9—dense, non-porous surface), attributed to uniform passivation derived from well-dispersed nanoparticles (size ~300 nm; ζ-potential −25 mV). In contrast, untreated coatings experience Rp collapse to 2.47 × 10
4 Ω·cm
2 at 8 A/dm
2 (
Figure 9—loose clusters), resulting from micron-sized agglomerate-induced crack propagation (>1 μm clusters in
Figure 3c). Furthermore, TAE-SiC-enhanced coatings exhibit minor Icorr fluctuations (±15%) under varying current densities (4 → 8 A/dm
2), whereas their untreated counterparts demonstrate over threefold Icorr variation. This distinct behavior confirms stabilized co-deposition via electrostatic repulsion (ζ-potential −25 mV) for TAE-SiC, whereas hydrophobic agglomerates randomly entrained at high current densities accelerate chaotic corrosion path formation.