Next Article in Journal
Decrypting the Unusual Structure and σ-Hole Interactions of the XC(NO2)3 (X=F, Cl, Br, and I) Compounds Using Quasi-Atomic Orbitals
Previous Article in Journal
High-Resolution Mass Spectrometry for Identification, Quantification, and Risk Assessment of 40 PFAS Migrating from Microwave Popcorn Bags
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Exploring How Dopants Strengthen Metal-Ni/Ceramic-Al2O3 Interface Structures at the Atomic and Electronic Levels

1
Jilin Institute of Chemical Technology, College of Aeronautical Engineering, Jilin 132022, China
2
Public Education Department, Gongqing Institute of Science and Technology, Gongqing 332020, China
3
School of Energy and Power Engineering, Northeast Electric Power University, Jilin 132012, China
4
College of New Energy and Materials, Northeast Petroleum University, Daqing 163711, China
5
State Key Laboratory of Advanced Welding and Joining, Harbin Institute of Technology, Weihai 264209, China
6
Jilin Institute of Chemical Technology, College of Mechanical and Electrical Engineering, Jilin 132022, China
*
Author to whom correspondence should be addressed.
Molecules 2025, 30(9), 1990; https://doi.org/10.3390/molecules30091990
Submission received: 31 March 2025 / Revised: 25 April 2025 / Accepted: 25 April 2025 / Published: 29 April 2025

Abstract

:
The metal-based/ceramic interface structure is a key research focus in science, and addressing the stability of the interface has significant scientific importance as well as economic value. In this project, the work of adhesion, heat of segregation, electronic structure, charge density, and density of states for doped-M (M = Ti, Mg, Cu, Zn, Si, Mn, or Al) Ni (111)/Al2O3 (0001) interface structures are studied using first-principle calculation methods. The calculation results demonstrate that doping Ti and Mg can increase the bonding strength of the Ni–Al2O3 interface by factors of 3.4 and 1.5, respectively. However, other dopants, such as Si, Mn, and Al, have a negative effect on the bonding of the Ni–Al2O3 interface. As a result, the alloying elements may be beneficial to the bonding of the Ni–Al2O3 interface, but they may also play an opposite role. Moreover, the Ti and Mg dopants segregate from the matrix and move to the middle position of the Ni–Al2O3 interface during relaxation, while other dopants exhibit a slight segregation and solid solution in the matrix. Most remarkably, the segregation behavior of Ti and Mg induced electron transfer to the middle of the interface, thereby increasing the charge density of the Ni–Al2O3 interface. For the optimal doped-Ti Ni–Al2O3 interface, bonds of Ti–O and Ti–Ni are found, which indicates that the dopant Ti generates stable compounds in the interface region, acting as a stabilizer for the interface. Consequently, selecting Ti as an additive in the fabrication of metal-based ceramic Ni–Al2O3 composites will contribute to prolonging the service lifetime of the composite.

Graphical Abstract

1. Introduction

Metal-based ceramic composites are advanced, interesting materials whose performances are based on a combination of metal and ceramics. They are widely used in industrial fields, such as aviation, energy, sensors, and automotive [1,2,3,4]. Metal-based ceramic composites mainly include a metal–ceramic coating and metal-based-reinforced ceramic particles. It is worth noting that as a novel Ni–Al2O3 composite, the interface between Ni and Al2O3 determines the stability and lifespan of the composite. Therefore, it is of great scientific significance to explore the properties of the Ni–Al2O3 interface.
Regarding experiments, most studies primarily deal with the production process of metal-based ceramic composites and assess their chemical stability and mechanical characteristics. For instance, research conducted by Martínez-Franco et al. [5] found that a Ni–Al2O3 composite prepared through the high-kinetic processing of ball milling and spark plasma sintering possessed excellent mechanical properties, making it a potential candidate for various high-temperature applications. Just recently, Gao et al. [6] showed that ceramic nanoparticle-enhanced multilayer Ni–Ni–Al2O3 coating structures exhibited superior toughness, wear resistance, and corrosion resistance. Moreover, it was discovered that this multilayer coating structure was less prone to cracking. Shen et al. [7] suggested that an Ni–Al2O3 coating pretreatment is expected to develop as a low-cost, novel, and effective method for promoting nitride. However, Fu et al. [8] suggested that the Ni–Al2O3 interface can crack and expand under hot conditions so that the material could fail early. Our early experiments [9] have also confirmed this. Currently, the biggest challenges faced by Ni–Al2O3 composites during manufacturing and use are the following: (1) poor wettability between the matrix and ceramic; and (2) weak bonding strength at the matrix–ceramic interface. Indeed, effective dopants would be an effective measure to strengthen the Ni–Al2O3 interface binding, as shown in Figure 1. Therefore, it is necessary to conduct a more in-depth exploration of the modification of the Ni–Al2O3 interface, especially at the atomic and electronic scales.
First-principle calculations based on quantum mechanics can study the properties of metal–ceramic interfaces, shedding light on the microscopic features at minimal material dimensions. Hocker et al. [10] applied first-principles calculations to predict the tensile properties of metal–ceramic interfaces at the electronic dimension. Our team [11,12] found that active alloying elements and rare earth elements can act as binders, reducing interfacial spacing and enhancing the bonding strength of the Fe–Al2O3 interface. Research by Chen et al. [13] has shown that doping Mg and Cu can make the charge distribution on the Al–Al2O3 interface uniform, thereby enhancing its mechanical properties. Chen et al. [14] successfully predicted the fracture process of the NiTi–Al2O3 interface, thereby obtaining the failure mechanism of the structures. Guo et al. [15] found that dopant-Ti can form strong Ti–O ionic bonds at the Ag–Al2O3 interface, enhancing the bonding strength of the interface. Consequently, it can be seen that employing the first-principles calculation method can obtain the properties of the interface and reveal the bonding mechanism of heterogeneous materials from the atomic and electronic scales, a task that is unachievable through experimental methods.
Only very recently, our group has conducted research on the properties of Ni–Al2O3 interfaces and discovered that there are three types of Ni–Al2O3 interfaces (single Al-terminated, double Al-terminated, and O-terminated), which have significant differences in their bonding strengths [16]. Among them, the single Al-terminated Ni–Al2O3 interface is the easiest to form in the actual environment, serving as the basic model for this research. This work primarily involves doping alloying elements into the single Al-terminated Ni–Al2O3 interface, aiming to optimize the interfacial environment and strengthen the interfacial bonding. The research details are presented below. Firstly, we utilized the work of adhesion to screen for dopants-M (M = Ti, Mg, Cu, Zn, Si, Mn, or Al), which are advantageous for the bonding of the Ni–Al2O3 interface. Following this, we undertook a detailed atomic and electronic structural analysis of the Ni–Al2O3 interfaces showing positive effects, thereby uncovering their micro-level characteristics. In the final stage, we conducted comprehensive testing on the best-performing doped-Ti Ni–Al2O3 interface conditions with the aim of elucidating the specific mechanisms behind the strengthening effects at this interface.

2. Results and Discussion

2.1. The Binding Strength of the Interface

Generally, the work of adhesion (abbreviated here as the Wad) can be used to evaluate the binding strength of the interface semi-quantitatively, as one of the criteria for judging the bonding properties of heterogeneous materials. In this work, the Wad of the doped-M (M = Ti, Mg, Cu, Zn, Si, Mn, or Al) and non-doped Ni–Al2O3 interface structures are calculated to reveal the effect of doping on the bonding of the interface, as shown in the following Equation (1) [17,18]:
W a d = E N i + M + E A l 2 O 3 E N i + M / A l 2 O 3 A
Here, ENi+M and E A l 2 O 3 denote the total energies of the Ni-side with dopant-M and Al2O3-side, respectively. E N i + M / A l 2 O 3 is the total energy of the doped-M Ni–Al2O3 interface structure, and A stands for the area of the Ni–Al2O3 interface.
Figure 2 shows the Wad of the doped-M and non-doped Ni–Al2O3 interface structures. In contrast to the non-doped Ni–Al2O3 interface, the Wad of the doped-M Ni–Al2O3 interfaces, in descending order, is the doped-Ti (4.23 J/m2) > doped-Mg (1.9 J/m2) > doped-Cu (1.29 J/m2) > non-doped (1.25 J/m2) > doped-Zn (0.85 J/m2), showing that Ti and Mg has a positive effect on strengthening the bonding of the Ni–Al2O3 interface. Yet, the Zn dopant weakens the bonding strength between the Ni and Al2O3. More noteworthy, the Wad of the Ni–Al2O3 interfaces doped with Si, Mn, and Al are negative, suggesting that these dopants would even prevent the formation of the interface. The main reasons for this phenomenon include the compatibility between elements in materials and the activity of the atoms (energy and size, etc.). In a previous study by our group [19], it was also found that titanium can improve the tensile properties of the Fe–Al2O3 interface structure. In this way, Ti seems to be effective in improving the interface properties of metals and Al2O3, which requires further verification with similar metals–Al2O3 structures. Consequently, doping alloy elements in the Ni–Al2O3 interface can have an uncertain effect on the bonding of that interface, which may be enhanced or weakened.
The process of doping will cause changes in the diameter and energy of the atomic position, which will induce a relaxation behavior in the atoms to form a new interface structure. The heat of segregation (ΔGseg) is the driving force that forms the stable compound, which can be used to evaluate the stability of dopants in the interface structure. The calculation of ΔGseg is performed utilizing the following Equation (2) [20]:
Δ G s e g = 1 n E N i / A l 2 O 3 E N i / A l 2 O 3 . n x + n μ x n μ N i
where E N i / A l 2 O 3 and E N i / A l 2 O 3 . n x represent the total energies of the non-doped and doped-M Ni–Al2O3 interface structures, respectively. Here, n denotes the total number of atoms doped with x (where x can be Ti, Mg, Cu, Zn, Si, Mn, or Al), and μ signifies the chemical potential of the dopant element. If ΔGseg is positive, it means that dopant atoms are easy to segregate from the matrix; if ΔGseg is negative, it indicates that the doped atoms are a solid solution in the matrix and not prone to segregation.
Figure 3 shows the ΔGseg of the doped-M and non-doped Ni–Al2O3 interface structures. It is obvious that the ΔGseg of the doped-Ti, Mg, and Cu Ni–Al2O3 interface structures are larger than 0, which are 2.34, 1.42, and 0.86 eV, respectively, indicating that the dopants are in an unstable state in the matrix. However, the ΔGseg of the doped-Zn, Si, Mn, and Al Ni–Al2O3 interfaces are negative and very close to 0, which means that the dopants are more prone to solid solutions in the matrix. From the perspective of the atomic structure, the Ti and Mg atoms have a large relaxation behavior, and the remaining dopants change very little, which matches the result of the ΔGseg. Indeed, a higher heat of segregation corresponds to a lower total energy in the atomic structure’s system, resulting in a more stable structure. Taken together, Ti, Mg, and Cu dopants are beneficial to the stability of the Ni–Al2O3 interface, and hence they are selected for further studies.

2.2. The Electronic Properties of the Interface

The formation of heterogeneous materials is complex, especially with chemical and physical changes at the interface region, which will induce a transfer and redistribution of electrons. To better understand the strengthening mechanism of the Ni–Al2O3 interface by doping elements, it is necessary to study the electronic properties of the interfacial region. Next, we will discuss the spatial distribution of the electrons, the charge density of the plane, the electron overlap population, and the density of states for the Ni–Al2O3 interface structures.
Figure 4 shows the spatial electron distribution of the non-doped and doped-M (M = Ti, Mg, and Cu) Ni–Al2O3 interface structures. In general, the electron distribution positions mean the interactions between the atoms. Compared to the non-doped Ni–Al2O3 interface, Mg and Ti dopants relax towards the center of the interface, transferring their surrounding electrons onto the interface. Interestingly, it is the segregation of the dopant that allows the electrons to be continuously distributed to connect the metal and the ceramic (see Figure 4c,d), which is a microscopic manifestation of the strengthening effect. It is worth noting that the number of electrons at the doped-Ti interface is more than that of other interfaces, indicating that Ti has the most significant effect on the Ni–Al2O3 interface. This phenomenon matches the results of the Wad above. Moreover, we found that Cu had no obvious effects on the electronic environment of the Ni–Al2O3 interface. Consequently, effective dopants can modify the interfacial electronic environment, thereby achieving the purpose of the strengthening effect. To put it differently, the properties of the interface are dictated by the electronic environment.
In order to better understand the characteristics of the Ni–Al2O3 interface, the charge density of the plane was analyzed, as shown in Figure 5. The plane with the greatest number of atoms is presented in this work, where red represents the charge aggregation region and blue represents the charge depletion region.
The distribution of charge density on the Ni-side and Al2O3-side are regular, indicating that the doping behavior of the interfacial region did not affect the non-interfacial region. The introduction of dopants (Ti, Mg, or Cu) at the Ni–Al2O3 interface widens the charge distribution region relative to the non-doped interface (illustrated by the yellow arrow in the Figure 5), marking a stronger bonding interaction from the interface. In particular, the Mg and Ti atoms had relaxed to the middle of the interface, effectively increasing the charge density of the interface. Herein, the segregation behavior of the dopants induces a lattice distortion at the interface, which may be beneficial for the bonding of the Ni–Al2O3 interface. Among all the doped Ni–Al2O3 interfaces, the reinforcement effect of Ti on interfacial bonding was the most significant, making it the best choice for practical applications. Therefore, further research on the strengthening mechanism of doped-Ti Ni–Al2O3 interface is necessary.
The overlap population of bonds provides a direct means to assess whether a valid chemical bond forms between atoms and unveils insights into the chemical changes occurring in the interfacial region. According to the traditional Mulliken formula (Equation (3)) [21], the overlap population of electrons for chemical bonds (Zbond) can be derived by using Equation (4). Both of these equations are expressible as the following:
X M u l l i k e n = A + I 2
Z b o n d = a e b ( Δ X M u l l i k e n ) 2 + c
Here, A represents the affinity energy of the atom, I represents the ionization energy of the atom, and a, b, and c are constants. The Zbond represents the electron overlap population for the bonds, with a greater value indicating a greater strength of bond. A Zbond value greater than 0 indicates that the two atoms are in a bonded state, while a Zbond value less than 0 indicates that the two atoms are in a non-bonded state.
Figure 6 is the overlap population of bonds at the doped-Ti Ni–Al2O3 interface. The values of overlap populations between Ti and O, between Ti and Ni-1, and between Ti and Ni-2 at the interface were 0.46, 0.2, and 0.17 e, respectively. This demonstrates that Ti segregates to the interface and undergoes chemical reactions, forming new chemical bonds with the atoms at both sides of the interface. In fact, the formation of new chemical bonds progressively integrates the two interfacial sides into a unified entity. Therefore, it can be inferred that the Ti undergoes complex chemical reactions in the Ni–Al2O3 interface region and produces new substances that contribute to the interface’s stability, which is the strengthening mechanism of the Ni–Al2O3 interface. As a result, the dopant acts as an adhesive at the interface, which can be understand as a “glue effect”.
In order to reveal the bonding mechanism at the doped-Ti Ni–Al2O3 interface, we calculated the partial density of states (PDOSs) of the atoms. Figure 7 shows the PDOS of the doped-Ti Ni–Al2O3 interface (the positions of the atoms are provided in Figure 6). Relative to the Ni-interior atom, the electron p-orbits of the Ni-1 and Ni-2 at the interface exhibit new peaks ranging from −34.57 to −33.58 eV, matching the shape of the p-orbit of the Ti atom. It can be suggested that the orbital hybridization of electrons has occurred between the Ti and Ni-1, as well as between the Ti and Ni-2, which are a characteristic feature of covalent bonds. Furthermore, the peak of the s-orbit of O-1 atom at the interface becomes stronger within the range of −21.34 to −18.25 eV. The s, p, and d-orbits of the Ti atoms exhibited peaks that match the shape of the s-orbit of the O-1 atom. It can be concluded that electrons are transferred from Ti to O-1, and orbital hybridization of electrons occurs between the Ti and O-1, resulting in the formation of a Ti–O ionic compound. Consequently, the Ti–O1, Ti–Ni1, and Ti–Ni2 bonds exhibit typical bonding peaks (see the green and yellow boxes in Figure 7). This bonding behavior promotes and accelerates the formation of a stable Ni–Al2O3 interface.
Here, our conclusions will provide theoretical guidance for practical applications, advancing the innovation of Ni–Al2O3 composites. In future work, our group will consider the effect of the defects of the materials on the Ni–Al2O3 interface and the strengthening effect of the dopants on the interface under the defects.

3. Calculation Method and Details

3.1. Calculation Parameter

For this research, every calculation, from bulks to surfaces and interfaces, are executed with the Cambridge Sequential Total Energy Package (CASTEP) code [22], which is based on density functional theory (DFT). Initially, the geometry optimization of unit cells, surfaces, and interfaces are carried out using the Broyden–Fletcher–Goldfarb–Shanno (BFGS) minimization algorithm [23]. Convergence tests confirmed that a plane-wave cutoff energy of 380 eV is sufficient. The Brillouin zone was sampled using an 8 × 8 × 1 k-point grid for both interface and surface calculations. To ensure the accuracy of the calculations, the optimization convergence criteria for the original crystal cell structure was set as the following: an energy deviation per atom of 0.01 meV, a stress deviation of 0.05 GPa, and a maximum displacement of 0.001 Å. Subsequently, the ultrasoft pseudopotential (USPP) [24] was used to describe the interaction between the ionic nucleus and valence electrons. The valence electrons of the three atoms are set as O-2s22p4, Al-3s23p1, and Ni-3d84s2, respectively. Ultimately, the generalized gradient approximation (GGA) with the Perdew–Burke–Enzerh (PBE) [25] method served as the electron exchange and correlation function in this work.

3.2. Model Building

According to the actual situation of material, the γ-Ni and α-Al2O3 cells serve as the basic units of the structure. The lattice constants for the Ni unit cell are a = b = c = 3.528 Å, and Al2O3 has a = b = 4.815 Å and c = 13.135 Å. Typically, stable heterogeneous interface systems are composed of crystal planes with the least energy of surface. It has been shown that γ-Ni (111) [17] and α-Al2O3 (0001) [26] belong to the surface structures with the lowest surface energy in the low-index crystal surfaces. Therefore, the α-Al2O3 (0001)/γ-Ni (111) interface system is established, which consists of 5 layers of Ni and 12 layers of Al2O3, as depicted in Figure 8.
Furthermore, the outermost atomic layer on the surface of the material has the greatest influence on the bonding properties of the interface, so the central position of the interface layer was chosen as the doping site to achieve effective contrast, as shown in Figure 9.

4. Conclusions

In this work, we conducted a systematic investigation into the bonding strength and strengthening mechanism of Ni–Al2O3 interfaces with doped-M (M = Ti, Mg, Cu, Zn, Si, Mn, or Al) at the atomic and electronic scales, yielding the following conclusions:
(1) The doping of Ti and Mg can increase the work of adhesion of the Ni–Al2O3 interface by up to factors of 3.38 and 1.52, respectively, while the dopant Cu almost does not change the work of adhesion of the interface. Conversely, the dopants of Zn, Si, Mn, or Al can not only fail to improve the work of adhesion, but also weaken the bonding strength of the interface. Accordingly, different dopants may exert either positive or negative effects on interfacial bonding.
(2) The dopant Ti and Mg atoms will precipitate from the matrix Ni and diffuse into the middle of the Ni–Al2O3 interface in the relaxation process. More importantly, the diffusion behavior of the Ti and Mg can induce the transfer and redistribution of electrons in the Ni–Al2O3 interface region. Compared to the non-doped Ni–Al2O3 interface, the doping of Ti and Mg increase the number of electrons and broaden the area of charge distribution in the interface region. As a result, effective dopants can optimize the electronic structure and enhance interatomic bonding strength at the interface.
(3) Of all the doped Ni–Al2O3 interfaces, Ti is the best candidate. It is found that the stronger bonds of Ti–Ni and Ti–O are formed at the interface, which significantly improved the stability of the Ni–Al2O3 interface. In other words, the dopant Ti appears to act as an adhesive at the Ni–Al2O3 interface.

Author Contributions

Methodology, X.Z.; software, Q.C.; validation, X.Z.; formal analysis, R.L.; investigation, R.L. and D.K.; resources, X.Z.; data curation, L.L. and H.Y.; writing—original draft F.S and R.L. writing—review and editing, R.L.; supervision, R.L. and F.S.; project administration, F.S.; funding acquisition, R.L. All authors have read and agreed to the published version of the manuscript.

Funding

This study were supported by the Department of Science and Technology of Jilin Province (No. 20240101109JC) and the Jilin Institute of Chemical Technology Doctor Scientific Research Starting Foundation (No. 2023015).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in article.

Acknowledgments

The authors would like to acknowledge the technical support from Northeast Electric Power University. The authors acknowledge the assistance of the JLICT Center of Characterization and Analysis.

Conflicts of Interest

There are no conflicts to declare.

References

  1. Yazdani, A.; Isfahani, T. Hardness, wear resistance and bonding strength of nano structured functionally graded Ni-Al2O3 composite coatings fabricated by ball milling method. Adv. Powder Technol. 2018, 29, 1306–1316. [Google Scholar] [CrossRef]
  2. Liu, Y.; Jiang, H.C.; Zhao, X.H.; Deng, X.; Zhang, W. High temperature electrical insulation and adhesion of nanocrystalline YSZ/Al2O3 composite film for thin-film thermocouples on Ni-based superalloy substrates. Appl. Surf. Sci. 2022, 579, 152169. [Google Scholar] [CrossRef]
  3. Opálek, A.; Švec, P.; Žemlička, M.; Štěpánek, M.; Štefánik, P.; Kúdela, S., Jr.; Beronská, N.; Iždinský, K. Ni Porous Preforms Compacted with Al2O3 Particles and Al Binding Agent. Materials 2023, 16, 988. [Google Scholar] [CrossRef] [PubMed]
  4. Ramanavicius, S.; Ramanavicius, A. Insights in the application of stoichiometric and non-stoichiometric titanium oxides for the design of sensors for the determination of gases and VOCs (TiO2-x and TinO2n-1 vs. TiO2). Sensors 2020, 20, 6833. [Google Scholar] [CrossRef]
  5. Martínez-Franco, E.; Trejo-Camacho, J.; Ma, C.; de la Torre, S.D.; García-Moreno, A.; Benítez-Castro, A.; Trapaga-Martinez, G.; Alvarado-Orozco, J.; Muñoz-Saldaña, J. Mechanical characterization by multiscale indentation of particle reinforced Nickel-Alumina metal matrix nanocomposites obtained by high-kinetic processing of ball milling and spark plasma sintering. J. Alloys Compd. 2022, 927, 166880. [Google Scholar] [CrossRef]
  6. Gao, M.Y.; Pei, Z.L.; Song, G.H.; Liu, Z.; Li, H.; Gong, J. Wear resistance of Ni/nano-Al2O3 composite coatings by brush electroplating. J. Mater. Sci. 2024, 59, 7009–7027. [Google Scholar] [CrossRef]
  7. Shen, Q.Q.; Zhang, Y.; Li, X.S.; Wang, L.; Nie, C. Enhanced Wear Resistance and Corrosion Resistance of 304 Austenitic Stainless Steel by Duplex Surface Treatment. Steel Res. Int. 2022, 93, 2100689. [Google Scholar] [CrossRef]
  8. Fu, X.Q. Nanostructure, plastic deformation, and influence of strain rate concerning Ni/Al2O3 interface system using a molecular dynamic study (LAMMPS). Nanomaterials 2023, 13, 641. [Google Scholar] [CrossRef]
  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. Hocker, S.; Schmauder, S.; Bakulin, A.; Kulkova, S. Ab initio investigation of tensile strengths of metal(111)/α-Al2O3(0001) interfaces. Philos. Mag. 2014, 94, 265–284. [Google Scholar] [CrossRef]
  11. Li, R.W.; Chen, Q.C.; Ji, M.X.; Ding, Y. Exploring failure mode and enhancement mechanism of doped rare-earth elements iron-based/alumina-ceramic interface. Ceram. Int. 2022, 48, 7827–7835. [Google Scholar] [CrossRef]
  12. 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]
  13. Chen, Y.T.; Liu, X.H.; Zhang, T.B.; Xie, H.; Zhao, N.; Shi, C.; He, C.; Li, J.; Liu, E. Interface intrinsic strengthening mechanism on the tensile properties of Al2O3/Al composites. Comput. Mater. Sci. 2019, 169, 109131. [Google Scholar] [CrossRef]
  14. Chen, L.; Li, Y.F.; Xiao, B.; Zheng, Q.; Gao, Y.; Zhao, S.; Wang, Z. First-principles calculation on the adhesion strength, fracture mechanism, interfacial bonding of the NiTi(111)//α-Al2O3(0001) interfaces. Mater. Des. 2019, 183, 108119. [Google Scholar] [CrossRef]
  15. Guo, W.B.; She, Z.Y.; Xue, H.T.; Zhang, X. Effect of active Ti element on the bonding characteristic of the Ag (111)/α-Al2O3 (0001) interface by using first principle calculation. Ceram. Int. 2020, 46, 5430–5435. [Google Scholar] [CrossRef]
  16. 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]
  17. Yang, J.; Wang, Y.; Huang, J.H.; 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]
  18. Guénard, P.; Renaud, G.; Barbier, A.; Gautier-Soyer, M. Determination of the α-Ai2O3(0001) Surface Relaxation and Termination by Measurements of Crystal Truncation Rods. Surf. Rev. Lett. 1998, 5, 321–324. [Google Scholar] [CrossRef]
  19. Li, R.W.; Zhang, X.F.; Liu, S.B.; Chen, Q.; Yang, H. Tensile performance and toughening mechanism of the alumina/iron interface doped with Ti: Atomic and electronic insights. Comput. Mater. Sci. 2024, 244, 113152. [Google Scholar] [CrossRef]
  20. Zhang, W.; Smith, J.R.; Wang, X.G.; Evans, A.G. Influence of sulfur on the adhesion of the nickel/alumina interface. Phys. Rev. B 2003, 67, 841–845. [Google Scholar] [CrossRef]
  21. Segall, M.; Shah, R.; Pickard, C.; Payne, M.C. Population analysis of plane-wave electronic structure calculations of bulk materials. Phys. Rev. B Condens. Matter 1996, 54, 16317–16320. [Google Scholar] [CrossRef] [PubMed]
  22. Segall, M.D.; Lindan, P.J.; Probert, M.A.; 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, 2717–2744. [Google Scholar] [CrossRef]
  23. 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]
  24. Vanderbilt, D. Soft self-consistent pseudopotentials in a generalized eigenvalue formalism. Phys. Rev. B Condens. Matter 1990, 41, 7892. [Google Scholar] [CrossRef] [PubMed]
  25. Perdew, J.P.; Burke, K. Ernzerhof, Generalized gradient approximation made simple. Phys. Rev. Lett. 1998, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
  26. Sun, J.; Stirner, T.; Matthews, A. Structure and surface energy of low-index surfaces of stoichiometric α-Al2O3 and α-Cr2O3. Surf. Coat. Technol. 2006, 201, 4205–4208. [Google Scholar] [CrossRef]
Figure 1. An overview of the Ni–Al2O3 interface region.
Figure 1. An overview of the Ni–Al2O3 interface region.
Molecules 30 01990 g001
Figure 2. The work of adhesion (Wad) of non-doped and doped-elements in Ni–Al2O3 interface.
Figure 2. The work of adhesion (Wad) of non-doped and doped-elements in Ni–Al2O3 interface.
Molecules 30 01990 g002
Figure 3. The heat of segregation (ΔGseg) for non-doped and doped Ni–Al2O3 interfaces.
Figure 3. The heat of segregation (ΔGseg) for non-doped and doped Ni–Al2O3 interfaces.
Molecules 30 01990 g003
Figure 4. The electronic distribution state of non-doped and doped Ni–Al2O3 interfaces. (a) non-doped. (b) doped-Cu. (c) doped-Mg. (d) doped-Ti.
Figure 4. The electronic distribution state of non-doped and doped Ni–Al2O3 interfaces. (a) non-doped. (b) doped-Cu. (c) doped-Mg. (d) doped-Ti.
Molecules 30 01990 g004
Figure 5. The charge density of non-doped and doped Ni–Al2O3 interfaces.
Figure 5. The charge density of non-doped and doped Ni–Al2O3 interfaces.
Molecules 30 01990 g005
Figure 6. The overlap population of bonds for the doped-Ti Ni–Al2O3 interface.
Figure 6. The overlap population of bonds for the doped-Ti Ni–Al2O3 interface.
Molecules 30 01990 g006
Figure 7. The PDOS of doped-Ti Ni–Al2O3 interface. (The position of the atoms is shown in Figure 6).
Figure 7. The PDOS of doped-Ti Ni–Al2O3 interface. (The position of the atoms is shown in Figure 6).
Molecules 30 01990 g007
Figure 8. The schematic diagram of the Ni–Al2O3 interface formation.
Figure 8. The schematic diagram of the Ni–Al2O3 interface formation.
Molecules 30 01990 g008
Figure 9. The supercell model of doped-M (M = Ti, Mg, Cu, Zn, Si, Mn, or Al) Ni–Al2O3 interface.
Figure 9. The supercell model of doped-M (M = Ti, Mg, Cu, Zn, Si, Mn, or Al) Ni–Al2O3 interface.
Molecules 30 01990 g009
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

Sun, F.; Zhang, X.; 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. https://doi.org/10.3390/molecules30091990

AMA Style

Sun F, Zhang X, 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(9):1990. https://doi.org/10.3390/molecules30091990

Chicago/Turabian Style

Sun, Fengqiao, Xiaofeng Zhang, Long Li, Qicheng Chen, Dehao Kong, Haifeng Yang, and Renwei Li. 2025. "Exploring How Dopants Strengthen Metal-Ni/Ceramic-Al2O3 Interface Structures at the Atomic and Electronic Levels" Molecules 30, no. 9: 1990. https://doi.org/10.3390/molecules30091990

APA Style

Sun, F., Zhang, X., Li, L., Chen, Q., Kong, D., Yang, H., & Li, R. (2025). Exploring How Dopants Strengthen Metal-Ni/Ceramic-Al2O3 Interface Structures at the Atomic and Electronic Levels. Molecules, 30(9), 1990. https://doi.org/10.3390/molecules30091990

Article Metrics

Back to TopTop