Next Article in Journal
Research on Thermal Effect and Laser-Induced Damage Threshold of 10.6 µm Antireflection Coatings Deposited on Diamond and ZnSe Substrates
Previous Article in Journal
A Study on the Spectral Characteristics of 83.4 nm Extreme Ultraviolet Filters
Previous Article in Special Issue
The Damage Evolution of a Cr2O3-TiO2 Coating Subjected to Cyclic Impact and Corrosive Environments and the Influence of a Nickel Intermediate Layer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tuning Nanocrystalline Heterostructures for Enhanced Corrosion Resistance: A Study on Electrodeposited Ni Coatings

1
College of Mechanical and Electrical Engineering, Nanjing Forestry University, Nanjing 210037, China
2
NOMATEN Centre of Excellence, National Centre for Nuclear Research, 05-400 Otwock, Poland
3
Jiangsu Key Laboratory of Advanced Metallic Materials, Southeast University, Nanjing 211189, China
4
Department of Chemistry, University of Helsinki, P.O. Box 55, 00014 Helsinki, Finland
*
Author to whom correspondence should be addressed.
Coatings 2025, 15(5), 534; https://doi.org/10.3390/coatings15050534
Submission received: 28 March 2025 / Revised: 25 April 2025 / Accepted: 27 April 2025 / Published: 30 April 2025

Abstract

:
Tailoring the microstructural heterogeneity of metallic coatings is a promising strategy for enhancing their corrosion resistance; however, its systematic optimization remains underexplored. Here in, we present a one-step, scalable electrodeposition strategy to fabricate Ni coatings with tunable nanocrystalline heterostructures on Cu substrates by varying the current density from 1 mA/cm2 to 50 mA/cm2. The coating with a current density of 10 mA/cm2, featuring a heterogeneous nanograin structure of coexisting small and large grains, exhibited optimal corrosion resistance in 3.5 wt.% NaCl solution, with a low self-corrosion current density of 4.48 µA/cm2. Electrochemical impedance spectroscopy (EIS) and molecular dynamics (MD) simulations revealed that the heterostructure dispersed Cl adsorption sites and promoted passivation. High-resolution transmission electron microscopy (HRTEM) revealed that as the current density increased from 10 mA/cm2 to 50 mA/cm2, the corrosion product transitioned from a crystalline NiOOH structure to an amorphous structure, which correlated with a reduced corrosion resistance. The heterogeneous microstructure enhances durability, offering a cost-effective and alloy-free alternative for offshore applications. These findings provide a theoretical and experimental basis for designing advanced corrosion-resistant coatings.

1. Introduction

Ni-based coatings have been widely utilized in engineering applications, particularly in offshore environments, owing to their corrosion resistance and mechanical reliability [1,2]. However, the increasing demands of modern engineering, such as extended service life, resistance to corrosion, and cost-effectiveness, present challenges that conventional Ni-based coatings may struggle to address fully. To overcome these shortcomings, an effective strategy is compositional plainification [3], a design philosophy that improves material performance by microstructural optimization alone [4], avoiding the complexity and cost of alloying elements [5]. Unlike alloyed coatings, such as high-entropy alloys, which depend on compositional diversity to enhance corrosion resistance at the cost of processing complexity and resource demands, compositional plainification uses controlled grain size refinement and structural heterogeneity to achieve superior properties [6,7,8]. Recent advances have highlighted nanostructured coatings with heterogeneous microstructures [9], which are promising alternatives. These heterostructured materials exhibit synergistic properties, including enhanced mechanical strength [10], oxidation resistance [11], and corrosion resistance [12]. These properties make them suitable candidates for next-generation coatings in harsh environments.
Electrodeposition is an advantageous technique for synthesizing nanostructured coatings. It provides a scalable, single-step process for customizing microstructural features, such as grain size and phase distribution [13,14]. Electrodeposition has been successfully applied to fabricate Ni-based coatings, including pure Ni and its binary to quinary alloys [15,16,17,18]. However, its potential for producing nanocrystalline heterostructures remains unexplored. Previous studies on electrodeposition provide a foundation for this work. Studies on electrodeposited Ni-based alloys have demonstrated that deposition parameters, such as current density, electrolyte composition, and temperature, influence grain refinement and coating properties [19,20,21,22]. Among these, our earlier investigation showed that tuning the current density is a straightforward and customizable approach to introducing heterostructures into electrodeposited coatings [16]. However, a systematic understanding of how these parameters control the formation of heterostructured coatings and their corrosion mechanisms remains limited. This gap implies the need for a comprehensive approach that integrates experimental synthesis with thorough characterization and corresponding simulations.
The approach in the current work yields Ni coatings with tunable microstructures designed to meet the rigorous demands of offshore engineering. To clarify the relationship between nanoscale architecture and corrosion resistance, the samples were characterized using a suite of experimental techniques, including X-ray diffraction (XRD), scanning electron microscopy (SEM), and electrochemical tests. The adsorption mechanisms governing the corrosion behavior were simulated using molecular dynamics (MD). The corrosion products were characterized using high-resolution transmission electron microscopy (HRTEM) to reveal the transitions in the corrosion mechanism. By linking the interplay between structural heterogeneity and corrosion resistance, this work provides actionable insights into the design of durable and cost-effective coatings. These findings offer a pathway for the development of durable coatings tailored for offshore energy and chemical engineering.

2. Materials and Methods

2.1. Electrodeposition Preparation

Pure Cu (≥99.9%) was selected as the substrate material. The substrates were machined into small pieces (15 mm × 15 mm × 1 mm), polished with SiC papers and diamond slurry, and etched in 10 vol. % sulfuric acid for 30 s, rinsed with ultrapure water, and dried with nitrogen gas. During this process, the surface was degreased by ultrasonic cleaning. Nanocrystalline Ni coatings were electrodeposited on pure copper using the direct current method at room temperature [16]. Each liter of the electrolytic bath solution comprised 76.1 mM NiSO4, 466.4 mM C6H8O7, 163.2 mM C6H5Na3O7, 1368.9 mM NaCl, 404.3 mM H3BO3, 7.3 mM C7H4NNaO3S, and 0.3 mM C12H25SO4Na (pH = 2.5). The electrolytic bath was prepared using ultrapure deionized water. A square platinum sheet electrode (1 cm × 1 cm) was used as the anode. The electrodeposition time was 30 min for each sample. During electrodeposition, different current densities were applied: 1 mA/cm2, 5 mA/cm2, 10 mA/cm2, 20 mA/cm2 and 50 mA/cm2. Based on the differences between the pre-and post-deposition weights, the estimated thicknesses for current densities of 1 mA/cm2, 5 mA/cm2, 10 mA/cm2, 20 mA/cm2, and 50 mA/cm2 were 0.3 μm, 1.9 μm, 3.7 μm, 6.2 μm, and 13.4 μm, respectively.

2.2. Material Characterization

The phase structure of the electrodeposited Ni coatings was identified using XRD (Ultima IV, Rigaku Corporation, Tokyo, Japan). Cu Kα radiation was used for the measurements. The operating voltage and current were 40 kV and 30 mA, respectively. The surface morphology was observed using field emission SEM (FE-SEM, Nova NanoSEM 450, FEI, Hillsboro, OR, USA) at an operating voltage of 15 kV. The electrochemical properties were evaluated using an electrochemical workstation (CHI-660E, Chenhua, Shanghai, China) with a standard three-electrode configuration. During the measurements, the test medium was a 3.5 wt.% NaCl solution at room temperature. The three-electrode system consisted of the prepared electrodeposited Ni coating as the working electrode, a saturated calomel electrode (SCE) as the reference electrode, and a platinum mesh electrode as the counter electrode. The exposed area was 1.0 cm2. Before the electrochemical measurements, the open-circuit potential (OCP) of each sample was monitored for 10 min. After the OCP measurements, the potential stabilized for all samples, confirming a steady state for subsequent electrochemical impedance spectroscopy (EIS) and polarization tests. During the EIS test, the frequency was set within a range from 0.1 Hz to 10 MHz, the sinusoidal perturbation was 5 mV, and the spectroscopy data were obtained at 0 V. After the EIS tests, the data were fitted using the ZSimpWin 3.30d software. Potentiodynamic polarization tests were performed at a scanning rate of 5.0 mV/s to evaluate the uniform corrosion of the coating. The results were further analyzed using the Tafel extrapolation method. To understand the corrosion products, freestanding TEM samples were mechanically peeled off from the samples after polarization testing. HRTEM tests were carried out using scanning TEM (STEM, Talos F200X, Thermo Fisher Scientific, Waltham, MA, USA) at an operating voltage of 200 kV. Energy dispersive spectroscopy (EDS, X-Max, Oxford Instruments, High Wycombe, UK) was used to analyze element distribution.

2.3. Adsorption Simulation

MD simulations were performed using LAMMPS (version: 19 Nov. 2024, Development) [22] to study Cl adsorption on Ni surfaces by comparing the grain boundary and interior regions. A symmetric tilt {210}[001] grain boundary was constructed using Atomsk 0.13.1 [23] by rotating two [001] grains ±26.57°. Cl ions were placed above the Ni surface, with two at the grain boundary and five in the grain interior to maintain charge neutrality. For the qualitative simulation, the interactions were modeled via a hybrid potential combining the embedded atom method (EAM) for Ni–Ni interactions [24] and customized Lennard–Jones (LJ) potentials for Ni–Cl (ε = 0.0025382 eV, σ = 2.300 Å) and Cl–Cl (ε = 0.00986272 eV, σ = 3.516 Å) interactions, with long-range Coulombic forces truncated at 10.0 Å and calculated using the particle–particle–mesh (PPPM) method. The Ni substrate was fixed, and simulations were performed at 300 K for 50,000 steps, that is, 25 ps, using a Nosé–Hoover thermostat. The adsorption energies were calculated from the stabilized potential energy of Cl ions in each region, and the results were visualized using OVITO 3.11.3 [25].

3. Results

Figure 1 shows the XRD patterns of the as-prepared Ni coatings at various current densities from 1 mA/cm2 to 50 mA/cm2. All the samples have two face-centered cubic (FCC) structures, i.e., electrodeposited Ni and the Cu substrate. The characteristic Ni peaks correspond to the (111)FCC, (200)FCC, and (220)FCC planes. Interestingly, these Ni peaks vary with the applied current density. This reflects the structural differences among the electrodeposited Ni coatings. At the lowest current density, the Ni peaks are broad and weak. This suggests the presence of fine grains in the samples. With increasing current density, the Ni peaks become relatively sharp and intense. This indicates a larger grain size in the sample. At current densities ranging from 1 mA/cm2 to 20 mA/cm2, the samples exhibit a preferred orientation along the (111) plane, corresponding to the plane with the lowest surface energy in FCC metals. At the largest current density of 50 mA/cm2, the samples exhibit a strong (220) orientation in FCC crystal growth. This aligns with our previous work [16], which demonstrates that high current densities produce strongly oriented coatings. The shift from the close-packed (111) plane to the (220) plane increases the surface energy, contributing to reduced corrosion resistance. Cu diffraction peaks remain stable as the current density increases from 1 mA/cm2 to 20 mA/cm2, and the intensity decreases at a current density of 50 mA/cm2, indicating a thicker Ni coating.
Figure 2 shows the surface morphologies of the electrodeposited Ni coatings obtained at different current densities from 1 mA/cm2 to 50 mA/cm2. The grain size increases with the current density, which is consistent with the XRD results. Notably, the grain size heterogeneity varies with the tuned current density. At the lowest current density, the grains are uniformly small. At a current density of 5 mA/cm2, the grains become heterogeneous, combining small and large grains, suggesting a transition in the nucleation–growth balance. The heterogeneity peaks at a current density of 10 mA/cm2, with small and large grains coexisting. At a current density of 20 mA/cm2, heterogeneity remains, but larger grains dominate. At the largest current density of 50 mA/cm2, the grains are overgrown, and the heterogeneity is almost lost. This corresponds to the texturing observed in the XRD results at the highest current density. This heterogeneity reflects the interplay between nucleation and growth during electrodeposition [26]. The peak heterogeneity in the sample with a current density of 10 mA/cm2 arises from the balance between nucleation and growth. A sufficiently high deposition rate [27] promotes grain growth, preferentially enlarging some grains, while numerous active nucleation sites sustain smaller grains. At lower current densities, nucleation dominates growth. With further increases in the current density, the loss of heterogeneity indicates that growth gradually dominates.
Polarization curves were measured to evaluate the corrosion resistance of the Ni coatings in a 3.5 wt.% NaCl solution (Figure 3a). The self-corrosion current density, Icorr, and self-corrosion potential, Ecorr, were fitted based on the polarization curves and are listed in Figure 3b. The samples exhibit different corrosion behaviors in 3.5 wt.% NaCl solution. The self-corrosion current density is the lowest in the sample with a current density of 10 mA/cm2 (4.48 µA/cm2) and the highest in the sample with a current density of 1 mA/cm2 (24.61 µA/cm2). The self-corrosion potential is most positive for the samples with current densities of 5 mA/cm2, 10 mA/cm2, and 20 mA/cm2 (−0.38 V, −0.37 V, and −0.39 V, respectively). For the sample with the largest current density, the self-corrosion current density increases to 16.47 µA/cm2, whereas the self-corrosion potential becomes most negative, i.e., −0.68 V. This indicates a reduced corrosion resistance and a change in the oxide kinetics. Beyond grain heterogeneity, the observed (220) texture contributes to this degradation, as its less close-packed structure increases the surface energy and enhances Cl adsorption compared to the (111)-oriented surfaces at lower current densities.
Figure 4 shows the EIS Nyquist plots of the electrodeposited Ni coatings with the fitting curves. The semicircles in the curves show the charge transfer and mass transport processes at the interface between the Ni coating and the 3.5 wt% NaCl solution. The Nyquist plots with current densities from 5 mA/cm2 to 50 mA/cm2 show much larger semicircles than those with a current density of 1 mA/cm2. An equivalent circuit, R1(Q1R2)(Q2R3) [28], was selected to fit the EIS data, where R1 is the solution resistance, Q1 and Q2 are constant phase elements, and R2 and R3 represent the charge transfer (Faraday) processes. This circuit was chosen to capture the two-time constants observed in the Nyquist plots, consistent with the electrochemical behavior of the Ni coatings in 3.5 wt.% NaCl solution.
The calculated values of the parameters are listed in Table 1. R2 is the charge-transfer resistance at the interface between the Ni coatings and the 3.5 wt% NaCl solution, while R3 represents a deeper charge-transfer process. The maximum R2 is 4586 Ω·cm2 in the sample with a current density of 10 mA/cm2. This indicates a strong charge-transfer resistance. The minimum value is 23.6 Ω·cm2 in the sample with a current density of 1 mA/cm2. This suggests an enhanced ion exchange across the reactive surface. Conversely, the maximum R3 is 7739 Ω·cm2 in the sample with a current density of 50 mA/cm2. This implies an additional mechanism, possibly the formation of a thicker corrosion product layer during the secondary charge transfer. However, the lower R2 and negative self-corrosion potential of the sample indicate that this layer offers limited corrosion protection.
The corrosion products of the electrodeposited Ni coatings prepared at a current density of 10 mA/cm2 were characterized using TEM after Tafel polarization. In Figure 5a, the bright-field image shows the petal-like nanostructures. In Figure 5b,c, HAADF-EDS mapping reveals the presence of Ni. As shown in Figure 5d, the HRTEM results confirm the corrosion products. The fast Fourier transform (FFT) pattern inset is obtained from the area in the red box with Z = [001]Ni. The inset of the inverse FFT (IFFT) results shows that the d-spacing of (011)Ni is 0.26 nm. Interestingly, Figure 5e shows NiOOH at the periphery of Ni. Compared with Ni, NiOOH exhibits a larger d-spacing of 0.59 nm for the (0001) plane [29]. NiOOH facilitates the formation of a passivation layer [30,31]. In 3.5 wt.% NaCl medium, it provides initial passivation [32], reducing coating dissolution.
Figure 6 shows the TEM results of the corrosion products of the Ni coatings at a current density of 50 mA/cm2. In Figure 6a, the bright-field image shows irregular, loosely packed nanostructures, which differ from the compact clusters of the sample with a current density of 10 mA/cm2. The inset selected-area electron diffraction (SAED) patterns in Figure 6a display diffuse rings. This confirms the amorphous nature [33] of the corrosion products. Porous amorphous corrosion products reduce protective efficacy by facilitating Cl penetration [34], likely due to their higher permeability and reduced barrier properties. A crystalline oxide inner layer is typically covered by an outer NiOOH/Ni(OH)2 layer at high potentials [35]. Under high current density, elevated overpotential, or prolonged exposure, the NiO layer grows rapidly and disorderedly due to irreversible oxidation reactions [35]. This disordered growth produces loosely packed structures, increasing susceptibility to Cl adsorption and subsequent corrosion. This property, coupled with hydration, remarkably increases the surface area exposed to corrosive species, promoting localized corrosion, such as pitting, and increasing corrosion rates, making it less protective than NiOOH. Thus, these amorphous corrosion products indicate a significant corrosion severity. This transition likely results from the elevated anodic overpotential for the sample with the largest current density, which accelerates Ni oxidation and promotes the rapid growth of a disordered structure [35], while the lower overpotential for the samples with lower current densities supports the formation of ordered NiOOH. Correspondingly, the sample exhibited reduced corrosion resistance, as indicated by a high Icorr of 16.47 µA/cm2 and Ecorr of −0.68 V. The highest R3, 7739 Ω·cm2, was also observed in the EIS results. This can be interpreted as resistance associated with a deeper, diffusion-limited process within the amorphous corrosion product, such as NiO·nH2O. Figure 6b,c confirms the presence of Ni through HAADF-EDS mapping. In Figure 6d, no lattice fringes are observed in the HRTEM images. Corresponding to the red boxed region in Figure 6d, the FFT inset shows diffuse and broad rings. These results confirm the non-crystalline character, i.e., the lack of long-range order.

4. Discussion

The corrosion resistance of the Ni-coated electrode, which was electrodeposited in 3.5 wt.% NaCl solution is fundamentally related to their microstructure. In this study, the microstructure was customized by tuning the applied current density during electrodeposition (from 1 to 50 mA/cm2). The prepared heterostructures remarkably enhanced corrosion resistance by leveraging extended grain diversity to mitigate corrosion. MD simulations were carried out to understand the adsorption mechanism of Cl on the Ni surface in the earliest stage of corrosion (Figure 7). MD simulations revealed stronger Cl adsorption at the grain boundaries (−4.36 eV) than at the grain interiors (−1.74 eV), offering additional insights that complement the density functional theory (DFT) calculations [36]. The deeper energy well at the boundary (2.62 eV difference) shows enhanced electrostatic and van der Waals interactions [37], indicating that the grain boundaries are preferential sites for attack before passivation layers such as NiOOH or NiO·nH2O are formed. Thus, grain boundaries serve as initial adsorption sites before passivation layers such as NiOOH or NiO·nH2O. Furthermore, the texture evolution from (111) to (220) at 50 mA/cm2 amplifies this degradation, highlighting the critical role of the grain structure and crystal orientation in governing the corrosion behavior.
In the sample with the smallest current density, the high density of nanoscale boundaries amplifies this weakness, as reflected in its electrochemical properties. In such nanocrystalline materials, triple junctions, i.e., points where three-grain boundaries intersect, exacerbate this effect below the critical grain size. Naturally, these junctions increase the defect density and provide pathways for corrosive species, such as Cl, due to the segregation of impurities or alloying elements at the triple junction [38]. Additionally, the triple-junction diffusion coefficient for Ni (at d = 70 nm) exceeds that of the grain boundaries (3.5 × 10−9 m2/s vs. 2.8 × 10−26 m2/s at 298 K) [39]. This enhances atomic mobility and potentially affects passive film formation during corrosion. The triple-junction volume fraction Vtj is estimated as follows [40],
V t j = V i c V g b
where Vic and Vgb are the intercrystalline and grain boundary volume fractions, respectively [40],
V i c = 1 ( d δ d ) 3
V g b = [ 3 δ ( d δ ) 2 ] d 3
where d is the grain size and where δ is the grain boundary thickness. This indicates a significant triple-junction network in the nanocrystalline material that can enhance atomic mobility compared to coarser structures [39]. In contrast, the prepared heterogeneous Ni coatings render this susceptibility a favorable mechanism. Compared to uniform nanograins, the mixture of small and large nanograins disperses Cl adsorption and mitigates localized corrosion, which may promote the formation of a NiOOH protective layer.
Synergy spans multiple scales in different heterogeneous materials, such as Fe–Cr–Ni alloys [41], AA 7075 Al alloys [42], and CoCrFeMnNi high-entropy alloys [43]. As corrosion progresses beyond initiation, nanoscale or ultrafine regions accelerate passivation kinetics, while microscale or coarse regions enhance durability and reduce oxidation and pitting susceptibility across diverse material systems and conditions [41,42,43]. The role of structural heterogeneity in corrosion resistance can be understood using a mechanical analogy. Similar to the way smaller nanograins impede dislocation motion to enhance strength, they also promote passivation. Larger grains provide structural stability during corrosion, similar to their role in maintaining ductility in heterogeneous materials [9]. This strategy exemplifies the potential of compositional plainification to achieve superior performance, such as corrosion resistance, through structural tuning rather than compositional complexity [3]. This approach optimizes material performance via tailored interface engineering with simpler compositions and pure elements. Traditional alloying, as seen in steel, superalloys, and high-entropy alloys strengthened with elements like Cr [44], often compromises manufacturing sustainability and complicates recycling [45]. As demonstrated with pure Ni coatings, compositional plainification enables stable nanoscale interfaces to enhance corrosion resistance without alloying. The heterogeneous structure can also enhance the mechanical properties, such as hardness, wear resistance, and fatigue resistance [9]. This work highlights the role of structural heterogeneity in corrosion properties, supporting a sustainable approach to high-performance coating design.

5. Conclusions

This study introduces a one-step, scalable electrodeposition approach to fabricate Ni coatings with tunable nanocrystalline heterostructures, and their corrosion resistance can be optimized by varying the current density from 1 to 50 mA/cm2. The experimental results demonstrate that the heterostructure formed at 10 mA/cm2, characterized by coexisting small and large grains, exhibits superior corrosion resistance in a 3.5 wt.% NaCl solution, with a self-corrosion current density as low as 4.48 µA/cm2, outperforming other conditions. Molecular dynamics simulations revealed that stronger Cl adsorption at grain boundaries than at grain interiors initiates corrosion, whereas the heterostructure mitigates this effect by dispersing the adsorption sites. High-resolution transmission electron microscopy further revealed that as the current density increased from 10 mA/cm2 to 50 mA/cm2, the corrosion product shifted from a crystalline NiOOH structure to an amorphous structure, reducing the passivation efficacy. This structural optimization strategy avoids the complexity of alloying, leverages the heterogeneous microstructure of pure Ni to enhance performance, and offers a cost-effective and manufacturable coating option for offshore energy and chemical engineering. By integrating electrochemical testing, microstructural characterization, and simulations, this work elucidates the multiscale synergistic mechanisms underpinning enhanced corrosion resistance, providing clear guidance for designing high-performance metallic coatings.

Author Contributions

W.H.: Conceptualization, Methodology, Investigation, Visualization, Writing—original draft, Writing—review and editing, and Supervision. Z.Z.: Investigation, Formal analysis, Visualization, Writing—original draft. X.H.: Investigation. Y.W.: Investigation. S.W.: Investigation. X.L.: Investigation. S.L.: Investigation. S.Z.: Investigation. F.F.: Writing—review and editing. J.J.: Writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Natural Science Foundation of Jiangsu Province, China (BK20220428) and Jiangsu Students’ Platform for Innovation and Entrepreneurship Training Program (202210298091Y). W.H. acknowledges support from the European Union Horizon 2020 research and innovation program under grant agreement no. 857470, the European Regional Development Fund via the Foundation for Polish Science International Research Agenda PLUS program grant no. MAB PLUS/2018/8. This publication was partly created within the framework of the project of the Minister of Science and Higher Education, Support for the activities of Centres of Excellence established in Poland under Horizon 2020, under contract no. MEiN/2023/DIR/3795.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data will be made available upon request.

Acknowledgments

W.H. acknowledges the computational resources provided by the High-Performance Cluster at the National Centre for Nuclear Research in Poland.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Wang, Y.; Zhou, X.; Lu, Y.; Li, X. Enhanced mechanical properties in bulk nanograined Ni with high-density fivefold twins. Small 2025, 21, 2410202. [Google Scholar] [CrossRef]
  2. Huang, C.; Zhang, Z.; Yan, J.; Sun, L.; Wang, J. Enhancing wear and corrosion resistance of electroless Ni-P coatings in CO2-saturated NaCl solution through polytetrafluoroethylene incorporation. Corros. Sci. 2024, 226, 111620. [Google Scholar] [CrossRef]
  3. Li, X.; Lu, K. Improving sustainability with simpler alloys. Science 2019, 364, 733–734. [Google Scholar] [CrossRef]
  4. Liu, D.; Wang, D.; Hong, T.; Wang, Z.; Wang, Y.; Qin, Y.; Su, L.; Yang, T.; Gao, X.; Ge, Z.; et al. Lattice plainification advances highly effective SnSe crystalline thermoelectrics. Science 2023, 380, 841–846. [Google Scholar] [CrossRef]
  5. Naghdi, A.; Domínguez-Gutiérrez, F.J.; Huo, W.Y.; Karimi, K.; Papanikolaou, S. Dynamic nanoindentation and short-range order in equiatomic NiCoCr medium-entropy alloy lead to novel density wave ordering. Phys. Rev. Lett. 2024, 132, 116101. [Google Scholar] [CrossRef]
  6. Zhao, S.; Zhang, R.; Yu, Q.; Ell, J.; Ritchie, R.O.; Minor, A.M. Cryoforged nanotwinned titanium with ultrahigh strength and ductility. Science 2021, 373, 1363–1368. [Google Scholar] [CrossRef]
  7. Li, Y.; Yuan, G.; Li, L.; Kang, J.; Yan, F.; Du, P.; Raabe, D.; Wang, G. Ductile 2-GPa steels with hierarchical substructure. Science 2023, 379, 168–173. [Google Scholar] [CrossRef]
  8. Li, X.Y.; Zhou, X.; Lu, K. Rapid heating induced ultrahigh stability of nanograined copper. Sci. Adv. 2020, 6, eaaz8003. [Google Scholar] [CrossRef]
  9. Zhu, Y.; Wu, X. Heterostructured materials. Prog. Mater. Sci. 2023, 131, 101019. [Google Scholar] [CrossRef]
  10. Wang, Y.; Zhu, Y.; Yu, Z.; Zhao, J.; Wei, Y. Hetero-zone boundary affected region: A primary microstructural factor controlling extra work hardening in heterostructure. Acta Mater. 2022, 241, 118395. [Google Scholar] [CrossRef]
  11. Kong, Y.; Peng, K.; Huang, H. Highly controllable additive manufacturing of heterostructured nickel-based composites. Int. J. Mach. Tool. Manuf. 2024, 195, 104112. [Google Scholar] [CrossRef]
  12. Peng, B.; Jin, J.; Liu, Y.; Lu, C.; Li, L.; Yan, M. Towards peculiar corrosion behavior of multi-main-phase Nd-Ce-Y-Fe-B permanent material with heterogeneous microstructure. Corros. Sci. 2020, 177, 108972. [Google Scholar] [CrossRef]
  13. Zhao, R.; Wang, H.; Du, H.; Yang, Y.; Gao, Z.; Qie, L.; Huang, Y. Lanthanum nitrate as aqueous electrolyte additive for favourable zinc metal electrodeposition. Nat. Commun. 2022, 13, 3252. [Google Scholar] [CrossRef]
  14. Ye, C.; Jin, H.; Shan, J.; Jiao, Y.; Li, H.; Gu, Q.; Davey, K.; Wang, H.; Qiao, S.-Z. A Mo5N6 electrocatalyst for efficient Na2S electrodeposition in room-temperature sodium-sulfur batteries. Nat. Commun. 2021, 12, 7195. [Google Scholar] [CrossRef]
  15. Li, Y.; Cai, X.; Zhang, G.; Xu, C.; Gao, W.; An, M. Optimization of electrodeposition nanocrytalline Ni-Fe alloy coatings for the replacement of Ni coatings. J. Alloy. Comp. 2022, 903, 163761. [Google Scholar] [CrossRef]
  16. Huo, W.; Wang, S.; Fang, F.; Tan, S.; Kurpaska, Ł.; Xie, Z.; Kim, H.S.; Jiang, J. Microstructure and corrosion resistance of highly <111> oriented electrodeposited CoNiFe medium-entropy alloy films. J. Mater. Res. Technol. 2022, 20, 1677–1684. [Google Scholar] [CrossRef]
  17. Li, S.; Li, H.; Zhai, Z.; Cao, X.; Liu, D.; Jiang, J. Corrosion resistance and tribological behavior of FeCoCrNi@GO/Ni high entropy alloy-based composite coatings prepared by electrodeposition. Surf. Coating. Technol. 2024, 477, 130379. [Google Scholar] [CrossRef]
  18. Dong, M.; Liu, P.; Wang, C.; Wang, Y.; Tang, X.; He, M.; Liu, J. Microstructure and properties of FeCoNiCr and FeCoNiCrW high entropy alloy coatings by electrodeposition. Intermetallics 2024, 175, 108492. [Google Scholar] [CrossRef]
  19. Wang, Y.; Guan, L.; He, Z.; Zhang, S.; Singh, H.; Hayat, M.D.; Yao, C. Influence of pretreatments on physicochemical properties of Ni-P coatings electrodeposited on aluminum alloy. Mater. Des. 2021, 197, 109233. [Google Scholar] [CrossRef]
  20. Daneshnia, A.; Raeissi, K.; Salehikahrizsangi, P. Rapid one-step electrodeposition of robust superhydrophobic and oleophobic Ni coating with anti-corrosion and self-cleaning properties. Surf. Coating. Technol. 2022, 450, 129007. [Google Scholar] [CrossRef]
  21. Wang, Y.; Zhang, G.; He, Z.; Chen, J.; Gao, W.; Cao, P. Superhydrophobic Ni nanocone surface prepared by electrodeposition and its overall performance. Surf. Coating. Technol. 2023, 464, 129548. [Google Scholar] [CrossRef]
  22. Thompson, A.P.; Aktulga, H.M.; Berger, R.; Bolintineanu, D.S.; Brown, W.M.; Crozier, P.S.; in’t Veld, P.J.; Kohlmeyer, A.; Moore, S.G.; Nguyen, T.D.; et al. LAMMPS—A flexible simulation tool for particle-based materials modeling at the atomic, meso, and continuum scales. Comp. Phys. Comm. 2022, 271, 108171. [Google Scholar] [CrossRef]
  23. Hirel, P. Atomsk: A tool for manipulating and converting atomic data files. Comput. Phys. Comm. 2015, 197, 212–219. [Google Scholar] [CrossRef]
  24. Foiles, S.M.; Baskes, M.I.; Daw, M.S. Embedded-atom-method functions for the fcc metals Cu, Ag, Au, Ni, Pd, Pt, and their alloys. Phys. Rev. B 1986, 33, 7983–7991. [Google Scholar] [CrossRef]
  25. Stukowski, A. Visualization and analysis of atomistic simulation data with OVITO—The Open Visualization Tool. Modelling Simul. Mater. Sci. Eng. 2010, 18, 015012. [Google Scholar] [CrossRef]
  26. Lin, J.; Kilani, M.; Baharfar, M.; Wang, R.; Mao, G. Understanding the nanoscale phenomena of nucleation and crystal growth in electrodeposition. Nanoscale 2024, 16, 19564–19588. [Google Scholar] [CrossRef]
  27. Chen, L.; Zhang, G.; Zhou, G.; Xiang, C.; Miao, X.; Liu, L.; An, X.; Lan, H.; Liu, H. In situ visual observation of surface energy-controlled heterogeneous nucleation of metal nanocrystals. Small 2024, 20, 2401674. [Google Scholar] [CrossRef]
  28. Zhou, J.; Cheng, Y.; Wan, Y.; Chen, H.; Wang, Y.; Ma, K.; Yang, J. Enhancement mechanisms of self-lubricating Ti3SiC2 ceramic doping in CoCrFeNi high-entropy alloy via high-speed laser cladding: Tribology and electrochemical corrosion. Surf. Coating. Technol. 2024, 480, 130554. [Google Scholar] [CrossRef]
  29. Dai, L.; Fang, C.; Yao, F.; Zhang, X.; Xu, X.; Han, S.; Deng, J.; Zhu, J.; Sun, J. Thickness-dependent β/γ-NiOOH transformation of Ni-MOFs in oxygen evolution reaction. Appl. Surf. Sci. 2023, 623, 156991. [Google Scholar] [CrossRef]
  30. Qin, L.; Lian, J.; Jiang, Q. Effect of grain size on corrosion behavior of electrodeposited bulk nanocrystalline Ni. Trans. Nonferr. Metal. Soc. China 2010, 20, 82–89. [Google Scholar] [CrossRef]
  31. Hall, D.S.; Lockwood, D.J.; Bock, C.; MacDougall, B.R. Nickel hydroxides and related materials: A review of their structures, synthesis and properties. Proc. R. Soc. A Math. Phys. Eng. Sci. 2015, 471, 20140792. [Google Scholar] [CrossRef] [PubMed]
  32. El-Tantawy, Y.A.; Al-Kharafi, F.M. Role of Cl− in breakdown of Ni passivity in aqueous NaOH solutions. Electrochim. Acta 1982, 27, 691–699. [Google Scholar] [CrossRef]
  33. Wan, C.; Zhang, Z.; Dong, J.; Xu, M.; Pu, H.; Baumann, D.; Lin, Z.; Wang, S.; Huang, J.; Shah, A.H.; et al. Amorphous nickel hydroxide shell tailors local chemical environment on platinum surface for alkaline hydrogen evolution reaction. Nat. Mater. 2023, 22, 1022–1029. [Google Scholar] [CrossRef] [PubMed]
  34. Marcus, P.; Maurice, V. Atomic level characterization in corrosion studies. Philos. T. R. Soc. A 2017, 375, 20160414. [Google Scholar] [CrossRef]
  35. Davies, D.E.; Barker, W. Influence of pH on corrosion and passivation of nickel. Corrosion 1964, 20, 47t–53t. [Google Scholar] [CrossRef]
  36. Yang, S.; Liang, G.; Huang, Y.; Hao, X.; Zhao, J.; Lv, M. Adsorption structure and properties of Ni/Fe electrodeposition interface: A DFT study. Model. Simul. Mater. Sci. Eng. 2024, 32, 055024. [Google Scholar] [CrossRef]
  37. Zhang, P.; Yang, Y.; Duan, X.; Liu, Y.; Wang, S. Density functional theory calculations for insight into the heterocatalyst reactivity and mechanism in persulfate-based advanced oxidation reactions. ACS Catal. 2021, 11, 11129–11159. [Google Scholar] [CrossRef]
  38. Wasekar, N.P. The influence of grain size and triple junctions on corrosion behavior of nanocrystalline Ni and Ni-W alloy. Scr. Mater. 2022, 213, 114604. [Google Scholar] [CrossRef]
  39. Chen, Y.; Schuh, C.A. Contribution of triple junctions to the diffusion anomaly in nanocrystalline materials. Scr. Mater. 2007, 57, 253–256. [Google Scholar] [CrossRef]
  40. Palumbo, G.; Thrope, S.J.; Aust, K.T. On the contribution of triple junctions to the structure and properties of nanocrystalline materials. Scr. Metall. Mater. 1990, 24, 1347–1350. [Google Scholar] [CrossRef]
  41. Mahesh, B.V.; Raman, R.K.S.; Koch, C.C. Bimodal grain size distribution: An effective approach for improving the mechanical and corrosion properties of Fe–Cr–Ni alloys. J. Mater. Sci. 2012, 47, 7735–7743. [Google Scholar] [CrossRef]
  42. Tian, W.; Li, S.; Wang, B.; Liu, J.; Yu, M. Pitting corrosion of naturally aged AA 7075 aluminum alloys with bimodal grain size. Corros. Sci. 2016, 113, 1–16. [Google Scholar] [CrossRef]
  43. Wang, J.; Zhang, Z.; Dai, H.; Fujiwara, H.; Chen, X.; Ameyama, K. Enhanced corrosion resistance of CoCrFeMnNi high entropy alloy using heterogeneous structure design. Corros. Sci. 2022, 209, 110761. [Google Scholar] [CrossRef]
  44. Li, Y.; Olejarz, A.; Kurpaska, Ł.; Lu, E.; Alava, M.J.; Kim, H.S.; Huo, W. Designing cobalt-free face-centered cubic high-entropy alloys: A strategy using d-orbital energy level. Int. J. Refract. Metal. Hard Mater. 2024, 124, 106834. [Google Scholar] [CrossRef]
  45. Huo, W.; Wang, S.; Zhang, X.; Ren, K.; Tan, S.; Fang, F.; Xie, Z.; Jiang, J. A strategy to improve the performance of TiO2 nanotube array film photocatalysts by magnetron-sputtered amorphous BiFeO3. Vacuum 2022, 202, 111135. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of Ni coatings with various current densities from 1 mA/cm2 to 50 mA/cm2.
Figure 1. XRD patterns of Ni coatings with various current densities from 1 mA/cm2 to 50 mA/cm2.
Coatings 15 00534 g001
Figure 2. SEM images of Ni coatings at various current densities. (a) 1 mA/cm2. (b) 5 mA/cm2. (c) 10 mA/cm2. (d) 20 mA/cm2. (e) 50 mA/cm2.
Figure 2. SEM images of Ni coatings at various current densities. (a) 1 mA/cm2. (b) 5 mA/cm2. (c) 10 mA/cm2. (d) 20 mA/cm2. (e) 50 mA/cm2.
Coatings 15 00534 g002
Figure 3. (a) Polarization curves of Ni coatings with various current densities from 1 mA/cm2 to 50 mA/cm2. (b) Fitted self-corrosion current density and potential.
Figure 3. (a) Polarization curves of Ni coatings with various current densities from 1 mA/cm2 to 50 mA/cm2. (b) Fitted self-corrosion current density and potential.
Coatings 15 00534 g003
Figure 4. (a) EIS Nyquist plots of Ni coatings with various current densities from 1 mA/cm2 to 50 mA/cm2 corresponding to the fitting curves. (b) Enlarged part of the Figure shows the low-magnitude data.
Figure 4. (a) EIS Nyquist plots of Ni coatings with various current densities from 1 mA/cm2 to 50 mA/cm2 corresponding to the fitting curves. (b) Enlarged part of the Figure shows the low-magnitude data.
Coatings 15 00534 g004
Figure 5. TEM results of the sample at a current density of 10 mA/cm2 after polarization testing. (a) Bright-field image. (b) HAADF image. (c) EDS mapping of Ni. (d,e) HRTEM images of the corrosion products.
Figure 5. TEM results of the sample at a current density of 10 mA/cm2 after polarization testing. (a) Bright-field image. (b) HAADF image. (c) EDS mapping of Ni. (d,e) HRTEM images of the corrosion products.
Coatings 15 00534 g005
Figure 6. TEM results of the sample at a current density of 50 mA/cm2 after polarization testing. (a) Bright-field image. (b) HAADF image. (c) EDS mapping of Ni. (d) HRTEM results of the corrosion products.
Figure 6. TEM results of the sample at a current density of 50 mA/cm2 after polarization testing. (a) Bright-field image. (b) HAADF image. (c) EDS mapping of Ni. (d) HRTEM results of the corrosion products.
Coatings 15 00534 g006
Figure 7. MD simulation results for Cl adsorption on the Ni surfaces, including the grain boundary and grain interior regions.
Figure 7. MD simulation results for Cl adsorption on the Ni surfaces, including the grain boundary and grain interior regions.
Coatings 15 00534 g007
Table 1. Calculated parameters of the equivalent circuits of the EIS results.
Table 1. Calculated parameters of the equivalent circuits of the EIS results.
Current Density
(mA/cm2)
R1
(Ω·cm2)
Q1-Y0
−1·cm2·Sn)
Q1-nR2
(Ω·cm2)
Q2-Y0
−1·cm2·Sn)
Q2-nR3
(Ω·cm2)
115.620.00193123.60.008550.422836.96
515.858.716 × 10−40.789311072.995 × 10−50.91581410
1034.542.641 × 10−50.90545862.983 × 10−40.87992473
2028.542.866 × 10−50.931235544.238 × 10−40.84311773
5025.822.689 × 10−40.979232952.713 × 10−50.95027739
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

Huo, W.; Zhang, Z.; Huang, X.; Wang, Y.; Wang, S.; Lu, X.; Li, S.; Zhu, S.; Fang, F.; Jiang, J. Tuning Nanocrystalline Heterostructures for Enhanced Corrosion Resistance: A Study on Electrodeposited Ni Coatings. Coatings 2025, 15, 534. https://doi.org/10.3390/coatings15050534

AMA Style

Huo W, Zhang Z, Huang X, Wang Y, Wang S, Lu X, Li S, Zhu S, Fang F, Jiang J. Tuning Nanocrystalline Heterostructures for Enhanced Corrosion Resistance: A Study on Electrodeposited Ni Coatings. Coatings. 2025; 15(5):534. https://doi.org/10.3390/coatings15050534

Chicago/Turabian Style

Huo, Wenyi, Zeling Zhang, Xuhong Huang, Yueheng Wang, Shiqi Wang, Xiaoheng Lu, Shuangxiao Li, Senlei Zhu, Feng Fang, and Jianqing Jiang. 2025. "Tuning Nanocrystalline Heterostructures for Enhanced Corrosion Resistance: A Study on Electrodeposited Ni Coatings" Coatings 15, no. 5: 534. https://doi.org/10.3390/coatings15050534

APA Style

Huo, W., Zhang, Z., Huang, X., Wang, Y., Wang, S., Lu, X., Li, S., Zhu, S., Fang, F., & Jiang, J. (2025). Tuning Nanocrystalline Heterostructures for Enhanced Corrosion Resistance: A Study on Electrodeposited Ni Coatings. Coatings, 15(5), 534. https://doi.org/10.3390/coatings15050534

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