Next Article in Journal
Oxidative and Glycation Stress Biomarkers: Advances in Detection Technologies and Point-of-Care Clinical Applications
Previous Article in Journal
Ytterbium(III) Tricyanomethanides with Sodium and Potassium: Similarities and Differences Between NaYb[C(CN)3]4 and KYb[C(CN)3]4
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Unveiling Vacancy-Driven Stability: Atomic and Electronic Insights into Ni/Al2O3 Interfaces

1
College of Urban Rail Transit, Jilin Railway Technical University, Jilin 132299, China
2
College of Mechanical and Electrical Engineering, Jilin Institute of Chemical Technology, Jilin 132022, China
3
State Key Laboratory of Advanced Welding and Joining, Harbin Institute of Technology, Harbin 150001, China
4
College of New Energy and Materials, Northeast Petroleum University, Daqing 163711, China
*
Author to whom correspondence should be addressed.
Molecules 2025, 30(21), 4285; https://doi.org/10.3390/molecules30214285
Submission received: 29 September 2025 / Revised: 29 October 2025 / Accepted: 31 October 2025 / Published: 4 November 2025

Abstract

The Ni/Al2O3 interface bears the load transfer and energy dissipation, which determines the service performance of the composite materials. In this study, three distinct vacancy-defect-modified interface models (D1, D2, and D3, corresponding to vacancies in the first, second, and third layers of the Ni substrate surface, respectively) were constructed to systematically investigate the regulatory mechanism of vacancies on interfacial stability. The underlying mechanism of vacancy-enhanced interfacial stability was elucidated from both atomic-scale structural and electronic property perspectives. The results demonstrate that the D1, D2, and D3 structures increase the adhesion work of the interface by 2.0%, 6.7%, and 0.3%, respectively. This enhancement effect mainly stems from vacancy-induced atomic relaxation at the interface, which optimizes the equilibrium interfacial spacing and effectively releases residual strain energy. Further electronic structure analysis reveals a notable increase in charge density at the vacancy-modified interface (particularly in the D2 structure), indicating that vacancy defects promote charge transfer and redistribution by altering local electron distribution. More importantly, the bonding strength of the interface exhibits a positive correlation with electron orbital hybridization intensity, where stronger s-, p-, and d-orbit hybridization directly leads to a more stable interface. These findings provide atomic- and electronic-scale insights into the mechanistic role of vacancy defects in governing bonding at the Ni/Al2O3 interface.

1. Introduction

Over the past few years, metal matrix ceramic composites (MMCCs) have gradually replaced traditional metallic materials in marine equipment applications, demonstrating broad prospects. Among them, Ni-Al2O3 composites have emerged as crucial functional coating materials due to their excellent corrosion resistance, high-temperature strength, and thermal stability [1,2,3]. Notably, the Ni/Al2O3 interface, as a critical region of the composite, not only performs essential functions in energy transfer and load distribution but also directly determines the overall performance of the composite material. However, harsh marine conditions (e.g., erosion and abrasion) impose increasingly demanding requirements on interfacial adhesion [4]. Therefore, investigating the microscopic mechanisms and optimization strategies of the Ni/Al2O3 interface holds significant scientific and practical importance.
To this day, Ni-Al2O3 composites highlight various fabrication methods to enhance their mechanical properties and corrosion resistance. For instance, Upadhyaya et al. [5] demonstrated that electroless Ni plating significantly improves the corrosion resistance of cold-sprayed Ni/Al2O3 interfaces by effectively mitigating chloride ion penetration. Building on this, Chang et al. [6] fabricated Ni-Al2O3 composite coatings via electrodeposition, revealing a direct proportional relationship between coating hardness and Al2O3 content. Further improving performance, Chen et al. [7] applied laser remelting to Ni-Al2O3 nanocomposite coatings, which simultaneously enhanced both tensile strength and elongation. In contrast, Dhand et al. [8] experimentally confirmed that thermally sprayed Ni-Al2O3 coatings exhibit superior wear resistance compared to their cold-sprayed counterparts. Beyond conventional reinforcements, Gong et al. [9] discovered that incorporating rare earth elements (Y, Ce) into ceramic/metal systems not only enhances corrosion resistance but also improves thermal shock resistance, offering a promising avenue for extreme-environment applications. However, these experimental approaches primarily provide empirical correlations without fully revealing the underlying atomic and electronic-scale mechanisms of the Ni/Al2O3 interface.
First-principles calculation methods, owing to their foundation in fundamental quantum mechanical principles, have become one of the most important theoretical approaches for investigating solid–solid interfaces [10,11,12]. Extensive studies have been conducted on metal–ceramic interfaces using this approach, revealing key insights into interface stability and bonding mechanisms [13,14,15,16,17,18]. For instance, Xue et al. [13] analyzed three interface configurations between Ti and different terminations (O, single-Al, and double-Al) of Al2O3 (0001), identifying the O-terminated interface as the most stable due to its ability to form stronger ionic bonds with Ti. Shi et al. [14] demonstrated that the O-terminated Al2O3 (0001)/Ni (111) interface exhibits the highest binding strength among the studied configurations. Liu et al. [15] further demonstrated that the O-terminated Ni/Cr2O3 interface represents the most stable configuration. Lin et al. [16] computationally showed that Al/TiC interfaces exhibit greater stability than Al/TiN interfaces. Beyond termination effects, researchers have explored broader material combinations. This electronic structure perspective was further quantified by Li et al. [17], who identified strong Fe-C ionic interactions as the primary contributor to enhanced adhesion strength in Fe/SiC interfaces through detailed charge density difference analysis. Bao et al. [18] found that dopant Yttrium substantially improves both the tensile strength and work of separation at Ni/Al2O3 interfaces. In fact, the interface region of a composite may contain various types of defects, such as vacancies, dislocations, and segregations. Based on this, the current work focuses specifically on investigating the influence of vacancy defects on properties of the Ni/Al2O3 interface.
Most recently, our team has systematically investigated the failure mechanisms and optimization strategies of ideal Ni/Al2O3 interfaces [19,20]. Based on these considerations, the present work selects the O-terminated Ni/Al2O3 interface, which exhibits the highest binding strength, as the primary research subject. The specific research content is as follows. Firstly, the γ-Ni (111)/α-Al2O3 (0001) interface structures with vacancies at different locations were constructed and optimized. Subsequently, the atomic relaxation behavior at the interface was analyzed to establish the relationship between vacancy positions and the bonding strength of the interface. Finally, the electronic structures of various interface configurations were examined to elucidate the influence of vacancy defects on interfacial bonding mechanisms. This study is expected to provide fundamental guidance for the interface theory of the Ni-Al2O3 composite.

2. Results and Discussion

2.1. Bonding Strength of the Interface

The work of adhesion (Wad) is defined as the energy required to separate a unit area of an interface between two dissimilar materials into two free surfaces [21]. This thermodynamic parameter provides a quantitative description of interfacial adhesion strength. Generally, a higher value of Wad at the interface corresponds to greater bonding strength. To investigate the effect of vacancy defects on the Ni/Al2O3 interface, it is necessary to calculate the Wad for both ideal and defective Ni/Al2O3 interface structures using Equation (1) [22]:
W a d = E N i + E A l 2 O 3 E N i / A l 2 O 3 A
where ENi and EAl2O3 represent the energy of the Ni-only block and the Al2O3-only block, respectively, when the opposing surfaces of each model are replaced by vacuum layers. ENi/Al2O3 is the total energy of the Ni/Al2O3 interface structure, and A denotes the interfacial area with the unit of m2.
The results of Wad for ideal and defect-containing Ni/Al2O3 interfaces are shown in Figure 1, with the following ranking: D2 (7.54607 J/m2) > D1 (7.21353 J/m2) > D3 (7.09151 J/m2) > Ideal (7.06946 J/m2). Herein, the structures corresponding to ideal, D1, D2, and D3 are provided in Figure 11 of Section 3.2. Compared to the ideal interface, the Wad for the D1 and D2 structures shows a significant increase, rising by 2.0% and 6.7%, respectively, while the increase for the D3 structure is negligible. This demonstrates that first- and second-layer vacancy defects enhance interfacial bonding, whereas third-layer vacancies, located farther from the interface, have negligible effects on bonding strength. Related studies have confirmed that vacancy defects can positively influence interfacial properties. For instance, Chen et al. [23] revealed that graphene/Al interfaces with single-vacancy defects exhibit significantly higher Wad than ideal interfaces. Similarly, Liu et al. [24] demonstrated that vacancy defects enhance the stability of Cu/graphene interfaces.
Figure 2 displays the atomic structure of the ideal and defective Ni/Al2O3 interface after optimization. Relative to the ideal Ni/Al2O3 interface, the vacancies can induce varying degrees of atomic migration at the interface, which microscopically reflects the balanced interatomic interactions. In particular, it can be observed that the bond length and bond angle between atoms in the interface region change, which indicates that the formation of vacancies changes the microstructure of the interface. The change in these structural parameters confirms that vacancy defects fundamentally reconstruct the atomic configuration of the interface by inducing local lattice distortion. As shown in Figure 2b, the D1 vacancy configuration exhibits modified interfacial relaxation due to missing first-layer Ni atoms. Both O and Ni atoms at the interface migrate toward the interfacial center, indicating enhanced interatomic interactions within the defect-containing region. For the D2 structure (Figure 2c), it demonstrates that the vacancy causes the release of constraints on the first layer of Ni atoms, which move toward the center of the interface by 0.78 Å. This phenomenon reduces the interfacial spacing, thereby enhancing the bonding strength between Ni and Al2O3. Notably, a collective rightward displacement occurs in the top three atomic layers of the Ni substrate. In Figure 2d, the D3 structure shows an overall parallel movement of the interface layer atoms, while the interface spacing does not change (see red dotted line). In fact, this microstructure reconstruction will directly affect the electron density distribution of the interface system and then change the composition and distribution characteristics of the interface binding energy.
This phenomenon demonstrates that vacancies in the first and second atomic layers of the Ni significantly influence interfacial bonding, while their effects diminish with increasing depth. In summary, near-surface vacancy defects in the Ni matrix positively influence the interface properties, which aligns with the results of Wad above. These findings collectively explain the variations in the interfacial bonding strength across different vacancy configurations, establishing a clear structure–property relationship between vacancy geometry and interface performance. To reveal the underlying mechanisms, further electronic-scale analysis will be conducted on the interfaces with vacancy defects, focusing on charge density redistribution and bonding orbital interactions.

2.2. Electronic Properties of Interface

Generally, the distribution and transfer of electrons at the interface can regulate the breaking and reconstruction of chemical bonds, thereby governing the interfacial bonding mechanism. Accordingly, this research discusses the electronic dynamics at the ideal and defective Ni/Al2O3 interface, including charge density, charge density difference, overlap population of the bond, and density of states.
Figure 3 displays the charge density of ideal and defective Ni/Al2O3 interfaces, where the color gradient from blue to red indicates progressively increasing charge concentrations. At non-interface regions, the charge distribution concentration in the atomic gap of the Ni matrix is significantly stronger than the Al2O3 side, which is determined by their respective crystal structure characteristics. The non-uniform charge-sharing zones between the Ni and O atoms are observed at the interface, indicating that vacancy defects induce different degrees of dynamic movement of electrons. Compared to the ideal Ni/Al2O3 interface, the D1 and D2 structures (Figure 3b,c) exhibit a marked increase in electron density between Ni and O atoms at the interface, as evidenced by the color gradient. Particularly, the D2 interface displays the strongest charge density, whereas D3 approaches ideal interface characteristics. Indeed, the charge concentration of the interface can represent the strength of the covalent bond. Therefore, the higher the charge density at the interface, the stronger the bonding strength of the interface, which is also consistent with the above Wad.
To further analyze electron transfer behavior at the Ni/Al2O3 interface, differential charge density was calculated for both ideal and defective structures, as shown in Figure 4. Here, the red-colored areas correspond to electron accumulation, whereas the blue-colored areas represent electron dissipation. The comparative analysis in Figure 4a–c demonstrates that the Ni atoms at the D1 and D2 interfaces form a deeper electron depletion region while the O atoms exhibit a more pronounced charge aggregation region relative to the ideal Ni/Al2O3 interface. This indicates that electrons are transferred from Ni to O atoms, forming a Ni-O ionic bond. Taking into account the observed charge density distribution, it can be inferred that the Ni/Al2O3 interface is predominantly characterized by ionic Ni–O bonds, with a minor contribution from covalent interactions. The presence of vacancies alters the local atomic structure, leading to changes in interlayer spacing and subsequent modifications in chemical bond strength, thereby affecting interfacial properties. Previous studies [14] have also confirmed a direct correlation between the charge distribution of Ni–O bonds and interlayer spacing at the Ni/Al2O3 interface.
In other words, the bond strength of Ni-O at the D1 and D2 interfaces is significantly greater than that at the ideal interface. In particular, the Ni-O ionic bond characteristic of the D2 interface is the most prominent. As evidenced in Figure 4d, the D3 interface exhibits an electron transfer situation similar to that of the ideal interface, indicating that vacancy defects away from the surface have a negligible effect on the electron transfer at the interface. In summary, the charge transfer characteristics observed across vacancy-modified interfaces exhibit strong consistency with the corresponding result of Wad, collectively validating the causal relationship between defect-induced electronic redistribution and interfacial bonding enhancement.
To gain deeper insights into the influence of vacancy defects on the bonding mechanism at the Ni/Al2O3 interface, it is essential to investigate the hybridization behavior of electronic orbitals through projected density of states (PDOS). Figure 5 presents the PDOS plots for ideal and defective Ni/Al2O3 interfaces, and the black dashed line indicates the Fermi level. In all interfacial configurations, the PDOS of interfacial O atoms exhibits a pronounced peak above zero at the Fermi level (marked by black dashed lines), in contrast to bulk O-center atoms, demonstrating that the O atoms at the interface have been metallized by Ni, forming partial metallic bonds. The PDOS spectrum of the D1 structure is shown in Figure 5b. In the energy range of −20 to −17.5 eV, the PDOS peaks of Ni and O atoms at the interface exhibit significant shape similarity (as indicated by the colored arrows). This corresponds to a typical bonding peak, indicating that electron orbitals are hybridized between atoms. It is worth noting that compared to the ideal Ni/Al2O3 interface, the D1 structure resulted in a greater number of bonding peaks, which means enhanced electron orbital hybridization at the interface, resulting in stronger covalent bonds. Previous studies have demonstrated that a larger overlap area of hybrid orbital bonding peaks at the interface correlates with higher bonding strength of the interface [25]. Further analysis of Figure 5c reveals that the D2 structure exhibits a greater overlap area of the bonding peak compared to the ideal interface, suggesting a corresponding enhancement in the bonding strength of the interface. In contrast, the PDOS spectrum of the D3 structure (Figure 5d) shows no significant differences. In fact, the electron orbital hybridization between atoms at the interface serves as the microscopic manifestation of interfacial reactions, which aligns with Chun et al.’s [26] conclusion that the Ni/Al2O3 interface produces NiO and NixAl1-x. However, the Ni/Al2O3 interface with vacancy defects can modulate the strength of electron orbital hybridization, leading to variations in the types and quantities of interfacial reaction products, thereby affecting the bonding properties of the interface. These results demonstrate the first- and second-layer vacancy-driven bonding strength of the interface by increasing the degree of electron orbital hybridization at the interface, thereby exhibiting superior bonding characteristics.
In order to directly quantify the binding strength of the Ni/Al2O3 interface, the electron overlap population (EOP) of the bonds between the Ni and O atoms at the interface is calculated, as shown in Figure 6. Positive values of EOP represent a covalent character, with larger values indicating stronger chemical bonds [27]. As shown in Figure 6b, the D1 structure shows a 39.3% increase in the EOP value of O1-Ni3 bonds compared to that of the ideal interface. For the D2 structure (Figure 6c), the EOP value of O2-Ni1 decreased by 26.3%, whereas the O1-Ni4 and O3-Ni3 pairs exhibited increases by 22.2% and 30.3%, respectively. It can therefore be inferred that the formation of vacancies at the interface induces electron transfer between atoms. An increase in the EOP value signifies enhanced interatomic interaction, corresponding to a strengthening of interfacial bonding. Another study [28] demonstrated that a decrease in the EOP value of the metal–ceramic interface region leads to the failure of the structure. Interestingly, both D1 and D2 structures exhibit not only increased total EOP but also the breaking of old bonds and the formation of new bonds, likely caused by lattice distortion induced by vacancy defects. In contrast, the D3 structure shows negligible variation. Thus, electron distribution and transfer constitute the fundamental determinants of the bonding performance of the Ni/Al2O3 interface.
In summary, the effects of vacancy defects on Ni/Al2O3 interfaces operate through three synergistic mechanisms. First, the atomic relaxation induced by vacancy defects significantly alters the local geometric configuration. Second, charge redistribution leads to a remarkable enhancement of interfacial electron density. Third, the strengthened orbital hybridization effect at the interface further intensifies covalent interactions. These collective behaviors jointly induce variations in the Wad at the Ni/Al2O3 interface, thereby enhancing its interfacial bonding strength. This discovery not only deepens our understanding of interface behavior in metal–ceramic composites but also provides new theoretical foundations for optimizing a composite interface design. Based on these findings, we plan to systematically investigate the effects of different defect types (e.g., interstitial atoms and dislocations) on interfacial properties, aiming to establish comprehensive theoretical guidance for developing high-performance composite materials.

3. Calculation Method and Details

3.1. Calculation Settings

All first-principles calculations in this work, encompassing bulk, surface, and interface behaviors, were conducted using the density functional theory (DFT)-based Cambridge Sequential Total Energy Package (CASTEP) [29,30]. Ultrasoft pseudopotentials (USPPs) [31] were employed to enhance computational efficiency in treating electron-ion interactions. To obtain the ground state configuration, the Broyden–Fletcher–Goldfarb–Shannon (BFGS) [32,33] algorithm was used for atom structures. The calculation parameters are set as follows. The Perdew–Burke–Ernzerhof (PBE) functional within the generalized gradient approximation (GGA) [34] framework was employed as the exchange–correlation functional to describe the Ni/Al2O3 interface system.
Here, plane-wave cutoff energy and k-point sampling density were carried out by convergence tests, as shown in Figure 7 and Figure 8. Thus, the cutoff energy was set to 380 eV with a k-point mesh of 8 × 8 × 1 for Ni/Al2O3 interface systems.
Additionally, the total energy was calculated as a function of interface distance. As shown in Figure 9, the structure is most stable when the interface distance is 1.7 Å. Furthermore, the convergence criteria were set as follows: energy change tolerance of 1.0 × 10−5 eV/atom, maximum force below 0.03 eV/Å, and maximum displacement of less than 0.001 Å. These parameters ensured rigorous geometric convergence of all systems.

3.2. Model Building

Previous experimental studies [35] have confirmed that stable interfaces in heterogeneous materials tend to form between surfaces with the lowest surface energy. Accordingly, the Ni (111) [36] and Al2O3 (0001) [37] surfaces were selected for this study. It should be noted that the Al2O3 (0001) surface exhibits three distinct terminations: 1Al-, 2Al-, and O-terminated configurations. Our earlier work [19] demonstrated that the characteristic properties of Al2O3 can be accurately captured with atomic layer thicknesses of 12, 14, and 13 layers for the 1Al-, 2Al-, and O-terminated Al2O3 (0001) surfaces, respectively. Similarly, Ni (111) requires at least five atomic layers to adequately represent its bulk characteristics [36]. However, the most stable interface configuration is achieved between a 5-layer Ni (111) surface and a 13-layer O-terminated Al2O3 (0001) surface. Based on these considerations, we constructed three distinct Ni-Al2O3 interface models as illustrated in Figure 10.
According to the methodology established by Zhang et al. [38], a systematic comparative analysis was conducted by maintaining a constant vacancy concentration of 25% while strategically varying the spatial distribution of vacancy sites. This involves the introduction of vacancy defects in the first, second, and third atomic layers of Ni (hereafter referred to as D1, D2, and D3, respectively), as shown in Figure 11.

4. Conclusions

This study systematically investigates the atomic and electronic properties of Ni (111)/Al2O3 (0001) interface structures modified by vacancy defects at different sites, focusing on the role of vacancies in interfacial bonding mechanisms.
(1) The results of Wad indicate that vacancy defects have a positive effect on the bonding performance of Ni/Al2O3 interfaces. Specifically, vacancy defects in the first (D1) and second (D2) layers of Ni increase the Wad by 2% and 6.7%, respectively, whereas the influence diminishes with increasing vacancy-defect depth.
(2) Various types of modified vacancy defects can induce atomic relaxation behaviors at the Ni/Al2O3 interface. For the D1 and D2 structures, the atoms at the surface move toward the center of the interface and shorten the interface spacing, which in turn increases the binding strength of the interface. Note that the atoms at the D3 interface move in parallel and do not change the interface spacing. Therefore, the occurrence of atomic relaxation behavior is a microscopic manifestation of the release of stress, making the Ni/Al2O3 interface structure more stable.
(3) The formation of vacancies alters the charge concentration in the shared region of the Ni/Al2O3 interface. The D1 and D2 structures exhibit significantly increased charge concentration, indicating electron transfer from both sides to the interface region. Specifically, the D2 structure exhibited the highest charge density, resulting in the strongest bonding strength of the interface.
(4) Vacancy defects can influence the hybridization behavior of electron orbitals at the interface. The D1 and D2 structures enhance the bonding strength of the interface by promoting electron orbital hybridization between Ni and O atoms at the Ni/Al2O3 interface, specifically manifested as an increase in the number of bonding peaks and an enlargement of the overlapping area of bonding peaks in the PDOS spectra. The greater the electron orbital hybridization, the stronger the bonding strength of the interface.

Author Contributions

Software, H.Y.; formal analysis, R.L.; investigation, R.L.; data curation, L.D. and D.K.; writing—original draft, L.D. and R.L.; writing—review and editing, L.D.; supervision, R.L. and H.Y.; project administration, H.Y. and D.K.; resources, H.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Acknowledgments

The authors would like to acknowledge the technical support from Harbin Institute of Technology.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Pan, X.Q.; Wang, J.J. Application of Surface Engineering Techniques to Corrosion Control over Offshore Engineering Equipment. Key Eng. Mater. 2008, 373, 551–555. [Google Scholar] [CrossRef]
  2. Abbas, M.K.G.; Ramesh, S.; Tasfy, S.; Lee, K.S. A state-of-the-art review on alumina toughened zirconia ceramic composites. Mater. Today Commun. 2023, 37, 106964. [Google Scholar] [CrossRef]
  3. Wang, Y.; Fan, L.; Chen, H. Effect of strength and toughness of ceramic particles on impact abrasive wear resistance of iron matrix composites. Mater. Today Commun. 2023, 35, 105689. [Google Scholar] [CrossRef]
  4. Kota, N.; Charan, M.S.; Laha, T.; Roy, S. Review on development of metal/ceramic interpenetrating phase composites and critical analysis of their properties. Ceram. Int. 2022, 48, 1451–1483. [Google Scholar] [CrossRef]
  5. Upadhyaya, R.; Tailor, S.; Shrivastava, S.; Gladkova, A.A.; Modi, S.C. Effect of electroless Ni plating on the properties of cold-sprayed Ni-Al2O3 coatings. Surf. Innov. 2017, 5, 97–105. [Google Scholar] [CrossRef]
  6. Chang, L.M.; Liu, J.H.; Zhang, R.J. Hardness and wear resistance of Ni/Al2O3 composite coatings by electrodeposition. Adv. Mater. Res. 2011, 154, 654–657. [Google Scholar]
  7. Dhand, D.; Grewal, J.S.; Kumar, P. Study and comparison of wear behaviour of Ni-Al2O3 coatings deposited by hot and cold spray. Surf. Topogr. Metrol. Prop. 2021, 9, 045056. [Google Scholar] [CrossRef]
  8. Chen, J.S.; Yang, J.M.; Qiao, B. Influence of Laser Remelting on Tensile Properties of Nanocomposite Ni-Al2O3 Coatings. Mater. Trans. 2012, 53, 2205–2207. [Google Scholar]
  9. Gong, W.B.; Li, R.W.; Li, Y.P.; Sun, D.; Wang, W. Stabilization and corrosion resistance under high-temperature of nanostructured CeO2/ZrO2-Y2O3 thermal barrier coating. Acta Metall. Sin. 2013, 49, 593–598. [Google Scholar] [CrossRef]
  10. Finnis, W.M.; Kruse, C.; SchÖnberger, U. First principles calculations for metal/ceramic interfaces. MRS Online Proc. Libr. 1994, 357, 427–434. [Google Scholar] [CrossRef]
  11. Yu, M.; Zhang, X.; Li, W.; Ai, T.; Dong, H.; Yang, Z. Study of Mo3NiB3 and Mo2NiB2 cermets by first-principles calculation and experiment. Next Mater. 2025, 7, 100324. [Google Scholar] [CrossRef]
  12. Zhang, H.; Li, J.; Bian, Y.; Cen, W.; Yin, D. The energetic, electronic and mechanical properties of metal (Fe/Co/Ni/Cu) || ceramics (TiN) interfaces: A first-principles study. Mater. Today Commun. 2025, 42, 111281. [Google Scholar] [CrossRef]
  13. Xue, H.; Wei, X.; Guo, W.; Zhang, X. Bonding mechanism study of active Ti element and α-Al2O3 by using first-principle calculation. J. Alloys Compd. 2020, 820, 153070. [Google Scholar] [CrossRef]
  14. Shi, S.; Tanaka, S.; Kohyama, M. First-principles study of the tensile strength and failure of α-Al2O3 (0001)/Ni (111) interfaces. Phys. Rev. B-Condens. Matter Mater. Phys. 2007, 76, 075431. [Google Scholar] [CrossRef]
  15. Liu, H.; Li, Y.; Zhang, C.; Dong, N.; Lan, A.; Li, H.; Dong, H.; Han, P. Effects of aluminum diffusion on the adhesive behavior of the Ni (111)/Cr2O3 (0001) interface: First principle study. Comput. Mater. Sci. 2013, 78, 116–122. [Google Scholar]
  16. Lin, Z.; Peng, X.; Fu, T.; Zhao, Y.; Feng, C.; Huang, C.; Wang, Z. Atomic structures and electronic properties of interfaces between aluminum and carbides/nitrides: A first-principles study. Phys. E Low-Dimens. Syst. Nanostructures 2017, 89, 15–20. [Google Scholar] [CrossRef]
  17. Li, Y.; Shi, M.; Gao, T.; Chen, C. Adhesion and mechanical properties of Fe/SiC interfaces analyzed at the atomic level: Insight from DFT calculations. Surf. Coat. Technol. 2025, 502, 131983. [Google Scholar] [CrossRef]
  18. Bao, Z.; Guo, X.; Shang, F. An atomistic investigation into the nature of fracture of Ni/Al2O3 interface with yttrium dopant under tension. Eng. Fract. Mech. 2015, 150, 239–247. [Google Scholar] [CrossRef]
  19. Li, R.W.; Liu, S.B.; Zhang, X.F.; Chen, Q.; Yang, H. Unveiling the secrets of the nickel-metal/alumina-ceramic interface during stretching: Atomic and electronic insights. Mater. Today Commun. 2024, 39, 109365. [Google Scholar] [CrossRef]
  20. Sun, F.Q.; Zhang, X.F.; Li, L.; Chen, Q.; Kong, D.; Yang, H.; Li, R. Exploring How Dopants Strengthen Metal-Ni/Ceramic-Al2O3 Interface Structures at the Atomic and Electronic Levels. Molecules 2025, 30, 1990. [Google Scholar] [CrossRef]
  21. Finnis, M.W. The theory of metal-ceramic interfaces. J. Phys. Condens. Matter 1996, 8, 5811. [Google Scholar] [CrossRef]
  22. Zhang, W.; Smith, J.R.; Evans, A.G. The connection between ab initio calculations and interface adhesion measurements on metal/oxide systems: Ni/Al2O3 and Cu/Al2O3. Acta Mater. 2002, 50, 3803–3816. [Google Scholar] [CrossRef]
  23. Chen, Y.; Liu, Y.; Zhou, F.; Chen, M.; Qu, N.; Liao, M.; Zhu, J. The interface properties of defective graphene on aluminium: A first-principles calculation. Comput. Mater. Sci. 2021, 188, 110157. [Google Scholar] [CrossRef]
  24. Liu, Y.; An, L.; Gong, L. First-principles study of Cu adsorption on vacancy-defected/Au-doped graphene. Mod. Phys. Lett. B 2018, 32, 1850139. [Google Scholar] [CrossRef]
  25. Li, R.W.; Chen, Q.C.; Ouyang, L.; Zhang, Y.; Nie, B.; Ding, Y. Insight into the strengthening mechanism of α-Al2O3/γ-Fe ceramic-metal interface doped with Cr, Ni, Mg, and Ti. Ceram. Int. 2021, 47, 22810–22820. [Google Scholar] [CrossRef]
  26. Chun, H.; Choi, D.; Kang, J.; Park, J.S.; Han, B. First-Principles computational study of Ni/α-Al2O3 hybrid interface reactions under extreme thermodynamic conditions. Appl. Surf. Sci. 2019, 509, 144861. [Google Scholar] [CrossRef]
  27. Segall, M.D.; Shah, R.; Pickard, C.J.; Payne, M.C. Population analysis of plane-wave electronic structure calculations of bulk materials. Phys. Rev. B 1996, 54, 16317. [Google Scholar] [CrossRef]
  28. Zhang, H.Z.; Liu, L.M.; Wang, S.Q. First-principles study of the tensile and fracture of the Al/TiN interface. Comput. Mater. Sci. 2007, 38, 800–806. [Google Scholar] [CrossRef]
  29. Mattsson, A.E.; Schultz, P.A.; Desjarlais, M.P.; Mattsson, T.R.; Leung, K. Designing meaningful density functional theory calculations in materials science—A primer. Model. Simul. Mater. Sci. Eng. 2004, 13, R1. [Google Scholar] [CrossRef]
  30. Segall, M.D.; Lindan, P.J.D.; Probert, M.J.; Pickard, C.J.; Hasnip, P.J.; Clark, S.J.; Payne, M.C. First-principles simulation: Ideas, illustrations and the CASTEP code. J. Phys. Condens. Matter 2002, 14, 2717. [Google Scholar] [CrossRef]
  31. Laasonen, K.; Pasquarello, A.; Car, R.; Lee, C.; Vanderbilt, D. Car-Parrinello molecular dynamics with Vanderbilt ultrasoft pseudopotentials. Phys. Rev. B 1993, 47, 10142. [Google Scholar] [CrossRef] [PubMed]
  32. Pfrommer, B.G.; Côté, M.; Louie, S.G.; Cohen, M.L. Relaxation of crystals with the quasi-Newton method. J. Comput. Phys. 1997, 131, 233–240. [Google Scholar] [CrossRef]
  33. Fischer, T.H.; Almlof, J. General methods for geometry and wave function optimization. J. Phys. Chem. 1992, 96, 9768–9774. [Google Scholar] [CrossRef]
  34. Perdew, J.P.; Burke, K.; Wang, Y. Generalized gradient approximation for the exchange-correlation hole of a many-electron system. Phys. Rev. B 1996, 54, 16533. [Google Scholar] [CrossRef]
  35. Fujimura, N.; Nishihara, T.; Goto, S.; Xu, J.; Ito, T. Control of preferred orientation for ZnOx films: Control of self-texture. J. Cryst. Growth 1993, 130, 269–279. [Google Scholar] [CrossRef]
  36. Yang, J.; Wang, Y.; Huang, J.; Ye, Z.; Sun, X.; Chen, S.; Zhao, Y. First-principles calculations on Ni/W interfaces in Steel/Ni/W hot isostatic pressure diffusion bonding layer. Appl. Surf. Sci. 2019, 475, 906–916. [Google Scholar] [CrossRef]
  37. Wang, X.G.; Chaka, A.; Scheffler, M. Effect of the Environment on α-Al2O3 (0001) Surface Structures. Phys. Rev. Lett. 2000, 84, 3650. [Google Scholar] [CrossRef]
  38. Zhang, X.F.; Li, R.W.; Chen, Q.C.; Kong, D.; Yang, H. Influence of Vacancy on the Bonding Properties of γ-Fe (111)/α-Al2O3 (0001) Interface: A First-Principles Study. Materials 2025, 18, 4666. [Google Scholar] [CrossRef]
Figure 1. The work of adhesion of ideal and defective Ni/Al2O3 interface structures.
Figure 1. The work of adhesion of ideal and defective Ni/Al2O3 interface structures.
Molecules 30 04285 g001
Figure 2. Local structure of ideal and defective Ni/Al2O3 interface after relaxation (unit of length/Å). Herein, (a) Ideal structure, (b) First-layer vacancy structure, (c) Second-layer vacancy structure, and (d) Third-layer vacancy structure.
Figure 2. Local structure of ideal and defective Ni/Al2O3 interface after relaxation (unit of length/Å). Herein, (a) Ideal structure, (b) First-layer vacancy structure, (c) Second-layer vacancy structure, and (d) Third-layer vacancy structure.
Molecules 30 04285 g002aMolecules 30 04285 g002b
Figure 3. Charge density of ideal and defective Ni/Al2O3 interface structures. Herein, (a) Ideal structure, (b) First-layer vacancy structure, (c) Second-layer vacancy structure, and (d) Third-layer vacancy structure.
Figure 3. Charge density of ideal and defective Ni/Al2O3 interface structures. Herein, (a) Ideal structure, (b) First-layer vacancy structure, (c) Second-layer vacancy structure, and (d) Third-layer vacancy structure.
Molecules 30 04285 g003
Figure 4. Charge density differences in ideal and defective Ni/Al2O3 interface structures. Herein, (a) Ideal structure, (b) First-layer vacancy structure, (c) Second-layer vacancy structure, and (d) Third-layer vacancy structure.
Figure 4. Charge density differences in ideal and defective Ni/Al2O3 interface structures. Herein, (a) Ideal structure, (b) First-layer vacancy structure, (c) Second-layer vacancy structure, and (d) Third-layer vacancy structure.
Molecules 30 04285 g004
Figure 5. The PDOS of ideal and defective Ni/Al2O3 interface structures (arrows of the same color represent the bonding peaks between different atoms). Herein, (a) Ideal structure, (b) First-layer vacancy structure, (c) Second-layer vacancy structure, and (d) Third-layer vacancy structure.
Figure 5. The PDOS of ideal and defective Ni/Al2O3 interface structures (arrows of the same color represent the bonding peaks between different atoms). Herein, (a) Ideal structure, (b) First-layer vacancy structure, (c) Second-layer vacancy structure, and (d) Third-layer vacancy structure.
Molecules 30 04285 g005
Figure 6. The overlap population of bonds for the ideal and defective Ni/Al2O3 interface. Herein, (a) Ideal structure, (b) First-layer vacancy structure, (c) Second-layer vacancy structure, and (d) Third-layer vacancy structure.
Figure 6. The overlap population of bonds for the ideal and defective Ni/Al2O3 interface. Herein, (a) Ideal structure, (b) First-layer vacancy structure, (c) Second-layer vacancy structure, and (d) Third-layer vacancy structure.
Molecules 30 04285 g006
Figure 7. Convergence verification of total energy for different Ecut. (a) Ni and (b) Al2O3.
Figure 7. Convergence verification of total energy for different Ecut. (a) Ni and (b) Al2O3.
Molecules 30 04285 g007
Figure 8. Convergence verification of total energy for different k-points.
Figure 8. Convergence verification of total energy for different k-points.
Molecules 30 04285 g008
Figure 9. The relationship between total energy and interfacial distance for the Ni/Al2O3 structure.
Figure 9. The relationship between total energy and interfacial distance for the Ni/Al2O3 structure.
Molecules 30 04285 g009
Figure 10. The establishment process of Ni(111)/Al2O3(0001) interface configuration.
Figure 10. The establishment process of Ni(111)/Al2O3(0001) interface configuration.
Molecules 30 04285 g010
Figure 11. The ideal and defective Ni/Al2O3 interface structures.
Figure 11. The ideal and defective Ni/Al2O3 interface structures.
Molecules 30 04285 g011
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

Duan, L.; Li, R.; Yang, H.; Kong, D. Unveiling Vacancy-Driven Stability: Atomic and Electronic Insights into Ni/Al2O3 Interfaces. Molecules 2025, 30, 4285. https://doi.org/10.3390/molecules30214285

AMA Style

Duan L, Li R, Yang H, Kong D. Unveiling Vacancy-Driven Stability: Atomic and Electronic Insights into Ni/Al2O3 Interfaces. Molecules. 2025; 30(21):4285. https://doi.org/10.3390/molecules30214285

Chicago/Turabian Style

Duan, Lili, Renwei Li, Haifeng Yang, and Dehao Kong. 2025. "Unveiling Vacancy-Driven Stability: Atomic and Electronic Insights into Ni/Al2O3 Interfaces" Molecules 30, no. 21: 4285. https://doi.org/10.3390/molecules30214285

APA Style

Duan, L., Li, R., Yang, H., & Kong, D. (2025). Unveiling Vacancy-Driven Stability: Atomic and Electronic Insights into Ni/Al2O3 Interfaces. Molecules, 30(21), 4285. https://doi.org/10.3390/molecules30214285

Article Metrics

Back to TopTop