Next Article in Journal
Micro-Computed Tomographic Evaluation of the Shaping Ability of Vortex Blue and TruNatomyTM Ni-Ti Rotary Systems
Previous Article in Journal
Intermolecular Interactions in Molecular Ferroelectric Zinc Complexes of Cinchonine
Previous Article in Special Issue
Toughening Mechanism of CaAl12O19 in Red Mud–Al2O3 Composite Ceramics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Understanding the Interface Characteristics Between TiB2(0001) and L12-Al3Zr(001): A First-Principles Investigation

1
State Key Laboratory of Featured Metal Materials and Life-Cycle Safety for Composite Structures, MOE Key Laboratory of New Processing Technology for Nonferrous Metals and Materials, Center of Ecological Collaborative Innovation for Aluminum Industry in Guangxi, and School of Resources, Environment and Materials, Guangxi University, Nanning 530004, China
2
Ningbo Boway Alloy Material Co., Ltd., Ningbo 315000, China
3
SAIC-GM-Wuling Automobile Co., Ltd., Liuzhou 545007, China
4
Guangxi Youhe Metal Materials Technology Co., Ltd., Pingguo 531400, China
*
Authors to whom correspondence should be addressed.
Crystals 2024, 14(11), 979; https://doi.org/10.3390/cryst14110979
Submission received: 18 October 2024 / Revised: 5 November 2024 / Accepted: 11 November 2024 / Published: 14 November 2024

Abstract

:
This study employs first-principles calculation methods to explore the characteristics of the TiB2(0001)/L12-Al3Zr(001) interface, including the atomic structure, adhesion work, interfacial energy, and electronic structure of various interface models. Considering four different terminations and three different stacking positions, twelve potential interface models were investigated. Surface tests revealed that a stable interface could be formed when a 9-layer TiB2(0001) surface is combined with a 7-layer ZrAl-terminated and a 9-layer Al-terminated Al3Zr(001) surface. Among these interfaces, the bridge-site stacking at the T/Al termination (TAB), hollow-site stacking at the Ti/ZrAl termination (TZH), top-site stacking at the B/Al termination (BAT), and hollow-site stacking at the B/ZrAl termination (BZH) were identified as the optimal structures. Particularly, the TAB interface exhibits the strongest adhesion strength and the lowest surface energy, indicating the highest stability. A Detailed analysis of the electronic structure further reveals that most interfaces predominantly exhibit covalent bonding, with the TAB, TZH, and BZH interfaces primarily featuring covalent bonds, while the BAT interface displays a combination of ionic and covalent bonds. The study ultimately ranks the stability of the interfaces from highest to lowest as TAB, BZH, TZH, and BAT.

1. Introduction

Aluminum–silicon alloys, characterized by low density, high strength, high wear resistance, and good machinability, are extensively used in mechanical manufacturing, automotive, maritime, and aerospace industries [1,2]. The performance of the alloy largely depends on the microstructure of its grains. However, in aluminum–silicon alloys, the coarse eutectic silicon phase can lead to grain coarsening, thereby diminishing the material’s performance [3]. Studies have shown that the addition of Zr to aluminum alloys can significantly enhance their performance. In these Zr-doped aluminum alloys, nanoscale Al3Zr precipitates have been observed and are considered to contribute to grain refinement [4].
Al3Zr in aluminum alloys exists in two structures, D023 and L12 [5,6]. D023-Al3Zr is a stable structure formed by tetragonal lattice distortion and intra-cell atomic migration, and it exhibits higher brittleness [7]. L12-Al3Zr is a metastable cubic structure, primarily precipitated from supersaturated aluminum solid solutions during rapid cooling processes [8]. L12-Al3Zr is completely coherent with the α-Al matrix, and it significantly contributes to inhibiting grain recrystallization and enhancing alloy strength through its shearing suppression effect on dislocation movement [9,10]. Studies have indicated that L12-Al3Zr serves as an effective heterogeneous nucleation substrate for α-Al at the atomic level [11,12], and first-principles calculations have confirmed that liquid Al atoms tend to nucleate on the L12-Al3Zr(001) surface at the ZrAl terminated [13].
Adding Al-Ti-B grain refiners to aluminum melt is also a widely applied method for improving the microstructure of grains, which contain strong nucleating particles like TiB2 [14,15]. TiB2 serves as an effective nucleating particle and plays a crucial role in the refinement of α-Al grains [16]. It has been reported that the refining mechanism of Al-Ti-B grain refiners is the formation of a two-dimensional Al3Ti compound on the (0001) TiB2 surface, which effectively promotes heterogeneous nucleation of α-Al [17]. Reports indicate that the distribution of TiB2 particles follows a log-normal distribution, with most nanoparticles being smaller than 100 nanometers [18]. Nanoscale TiB2 particles can pin dislocations at grain boundaries and inhibit the growth of recrystallized grains. Additionally, TiB2 is a boride with a hexagonal structure, known for its excellent stiffness, hardness, and wear resistance, making it an outstanding reinforcing material [19,20].
When both Al-Ti-B grain refiners and excess Zr and Si elements are present in the aluminum melt, the grain refining capability is significantly reduced, leading to the so-called “poisoning” phenomenon of the grain refiner [21,22]. The prevailing view on the mechanism of the “poisoning” phenomenon is that Zr atoms in the Al melt cause the dissolution of Al3Ti on the TiB2(0001) surface, thereby reducing the nucleation capability of the refiner [15,23,24,25]. Some studies also speculate that TiB2 loses its ability to refine Al grains due to the formation of interfacial nucleation with Al3Zr [26]. Therefore, to further elucidate the mechanism of the “poisoning” phenomenon, it is necessary to study the relationship between the two nucleating particles, TiB2 and Al3Zr. Current research on the interface relationship between TiB2 and D023-Al3Zr is already quite extensive. Researchers have observed the precipitation of D023-Al3Zr at the TiB2 interface using scanning electron microscopy, with the orientation relationships identified as (0001)TiB2//( 1 ( _ ) 0 3 ( _ ) )D023-Al3Zr, (10 1 ( _ ) 0)TiB2//(10 5 ( _ ) )D023-Al3Zr, and [12 1 ( _ ) 0]TiB2//[010]D023-Al3Zr [27,28]. However, research on the interface relationship between TiB2 and L12-Al3Zr remains quite scarce. During the solidification process of the alloy melt, the precipitation of the D023-Al3Zr phase at the TiB2 interface nearly depletes Zr from the matrix, leading to scant precipitation of L12-Al3Zr around the TiB2 clusters. This situation makes it difficult to observe the interface relationship between TiB2 and L12-Al3Zr through experimental means [29]. Therefore, simulations based on Density Functional Theory can provide a convenient means to study the potential interface relationship between these two phases.
This article employs first-principles calculation methods to systematically investigate the structural and electronic properties of the TiB2(0001)/L12-Al3Zr(001) interface at the atomic level. To gain a more comprehensive understanding of the bonding behavior at the TiB2(0001)/L12-Al3Zr(001) interface, we also predicted the ideal adhesive work and interfacial energy. Finally, the electronic structure of TiB2(0001)/L12-Al3Zr(001) was analyzed through the differential charge density and a density of states analysis.

2. Calculation Method

First-principles calculations were performed using the Cambridge Serial Total Energy Package (CASTEP) code [30]. The interactions between ionic cores and valence electrons were described using ultrasoft pseudopotentials [31]. Within the framework of the Generalized Gradient Approximation (GGA), the Perdew–Burke–Ernzerhof (PBE) function was employed as the exchange-correlation potential [32]. Theoretical calculations were performed with a plane-wave cutoff energy of 380 eV. For the calculations of bulk and surfaces, Brillouin zone sampling was carried out using the Monkhorst–Pack method, with K-points set to 10 × 10 × 10 for bulk and 6 × 6 × 1 for surfaces [33]. During the relaxation of all atomic models, the Broyden–Fletcher–Goldfarb–Shanno (BFGS) minimization scheme was used for ensemble optimization steps, with all atomic structures and volumes being fully relaxed [34]. The structural optimizations of these systems were relaxed until the total energy was less than 1 × 10−5 eV/atom. The convergence tolerances for the forces between the atoms, maximum stress, and maximum ionic displacement were set to 0.01 eV/Å, 0.02 GPa, and 5 × 10−4 Å, respectively. Additionally, for the interfaces, a vacuum layer of 14 Å was chosen to isolate the free surfaces, ensuring sufficient space to prevent their interactions.

3. Results and Discussion

3.1. Crystal Properties

The schematic diagram of the crystal structure model used in the calculations is shown in Figure 1. From Figure 1a, it can be seen that the crystal structure model of the bulk TiB2 was of the AlB2 type, with the space group P6/mmm. As shown in Figure 1b, the crystal structure model of the bulk Al3Zr was of the Cu3Au type, with the space group Pm-3m.
To ensure the accuracy of the calculations in this paper, especially the accuracy of the pseudopotentials used, a series of calculations were performed on the lattice constants and formation enthalpies of TiB2 and Al3Zr using the Perdew–Burke–Ernzerhof (PBE) within the Generalized Gradient Approximation. The lattice constants were obtained directly from the calculations, and the formation enthalpies of TiB2 and Al3Zr were derived using Equations (1) and (2) [35]:
Δ γ H T i B 2 = E b u l k T i B 2 N b u l k T i T i B 2 × E b u l k T i N b u l k T i T i N b u l k B ( T i B 2 ) × E b u l k B N b u l k B B N b u l k T i T i B 2 + N b u l k B T i B 2
Δ γ H A l 3 Z r = E b u l k A l 3 Z r N b u l k A l A l 3 Z r × E b u l k A l N b u l k A l A l N b u l k Z r A l 3 Z r × E b u l k Z r N b u l k Z r Z r N b u l k A l A l 3 Z r + N b u l k Z r A l 3 Z r
where Δ γ H T i B 2 and E b u l k T i B 2 represent the formation enthalpy and the total energy of bulk TiB2, respectively; E b u l k T i denotes the energy of a Ti atom in the bulk, and E b u l k B is the energy of boron in its conventional cell; N b u l k T i T i B 2 and N b u l k B ( T i B 2 ) are the numbers of Ti and B atoms in the bulk model of TiB2, respectively; and N b u l k T i T i and N b u l k B ( B ) represent the numbers of atoms in the conventional cells of titanium and boron, respectively. Equation (2) follows the same rationale and will not be reiterated here.
Table 1 presents the lattice constants and formation enthalpies of TiB2 and L12-Al3Zr, comparing them with other computational and experimental results. In this work, the lattice constants of TiB2 differ by less than 0.004 Å from the theoretical values reported in the literature [36,37], and the deviation from the experimental values reported in the literature does not exceed 0.006 Å [38]. The calculated formation enthalpy of TiB2 had a maximum deviation of 0.013 eV/atom from other theoretical values obtained from calculations and a difference of 0.062 eV/atom compared with the experimental formation enthalpies. For the L12-Al3Zr unit cell structure, the maximum discrepancy in lattice constants compared to theoretical values from the other literature was 0.035 Å, with a formation enthalpy error within 0.021 eV/atom [11,39], and the discrepancy from the experimental lattice constant reference values was 0.008 Å [40]. These results indicate that our current computational methods and settings are reliable, and the calculation accuracy was sufficient.

3.2. Surface Convergence and Surface Energy

Due to the significant difference in lattice constants between the primitive cells of TiB2(0001) and the Al3Zr(001) surfaces, direct construction of the interface model was not feasible. To satisfy the periodicity conditions and lattice relationships, it was necessary to find matching lattice constants for the TiB2(0001) and Al3Zr(001) surfaces to build the interface model. After testing, it was found that the lattice constants of the TiB2(0001)-( 3 × 2 )R30° surface unit and the Al3Zr(001)-( 2 × 2 )R45° surface unit are numerically similar, and the mismatch rate at the interface can be calculated as follows:
ξ = 2 A A A l 3 Z r + A T i B 2 1
where in the above equation, A represents the interfacial area after the TiB2(0001) and Al3Zr(001) surfaces form an interface, while A A l 3 Z r and A T i B 2 , respectively, refer to the areas of the Al3Zr(001) and TiB2(0001) surfaces before forming the interface. Using Equation (1), the mismatch rate of this interface was calculated to be less than 1%. Therefore, the TiB2(0001)/Al3Zr(001) interface model was considered reasonable.
Before establishing the interface model, it was necessary to perform convergence tests on the atomic thickness of the phases on both sides of the interface to ensure that the atoms deep within the surface exhibited characteristics similar to those of bulk phase atoms. Typically, the convergence of the surface energy for both the TiB2(0001) and Al3Zr(001) surfaces was used to determine this. Since both the TiB2(0001) and Al3Zr(001) surfaces are non-polar, the convergence could be assessed by measuring the percentage change in relaxation between the atomic layers ( Δ i j , where i , j represent the atomic layers in the surface model).
The change in interlayer distance ( Δ i j ) is defined as follows:
Δ i j = d i j d i j 0 d i j 0 × 100 %
where Δ i j is the initial distance between layers i and j in the ideal crystal structure and is the vertical distance between the adjacent i and j layers after relaxation.
After the geometric optimization of the atoms, the changes in the interlayer distances of the TiB2(0001) and Al3Zr(001) surfaces are shown in Table 2. For TiB2, when the number of atomic layers nn in the Ti-terminated and B-terminated TiB2(0001) surfaces was ≥9, the changes in the interlayer distances in TiB2(0001) were essentially converged. A comparison of the percentage changes in the interlayer distances of the same number of atomic layers shows that the relaxation of the outermost interlayer distance on the B-terminated surface was much greater than that on the Ti-terminated surface. The results indicate that the outermost interlayer distances on the Ti-terminated and B-terminated surfaces with nine atomic layers decreased by 7.7% and 9.6% of their bulk volumes, respectively. From Table 2, it can be seen that for Al3Zr, the interlayer distances on the Al-terminated side essentially converged when nn was ≥9, while on the ZrAl-terminated side, a good convergence effect was achieved when nn was ≥7.
For simulations related to surfaces, the more atomic layers the surface model contains, the more precise the results become. However, as the number of atoms increases, the computational resources required also increase. Therefore, to further determine the minimum thickness of the TiB2(0001) and Al3Zr(001) surfaces that satisfy the properties of the bulk phase, and to reduce the error in determining the number of layers, we further discussed the impact of different layer thicknesses on the convergence of surface energy. This approach makes the selection of convergence layer numbers more reliable [41].
Surface energy is an important parameter for measuring surface stability. For non-stoichiometric TiB2 surface models, the chemical potentials of both Ti and B should be considered. Therefore, the surface energy of the TiB2(0001) surface was calculated using a generalized definition of surface energy [42,43]:
γ T i B 2 0001 = E s u r f T i B 2 0001 N T i B 2 0001 T i × μ b u l k T i B 2 + 2 N T i B 2 0001 T i N T i B 2 0001 B × μ s l a b B 2 × A s u r f T i B 2 0001
In the formula, γ T i B 2 ( 0001 ) represents the surface energy of the TiB2(0001) surface. E s u r f T i B 2 0001 refers to the total energy of the TiB2(0001) surface used in the calculation. N T i B 2 0001 T i is the number of Ti atoms on the TiB2(0001) surface. N T i B 2 0001 B represents the number of B atoms on the TiB2(0001) surface. μ b u l k T i B 2 is the chemical potential of bulk TiB2. μ s l a b B represents the chemical potential of B in the surface. A s u r f T i B 2 ( 0001 ) is the area of the TiB2(0001) surface model.
It is well known that the chemical potentials of various elements in the surface composition are always lower than the corresponding bulk chemical potentials because the surface optimizes energy through a specific ordered arrangement [44]. This implies that μ s l a b B μ b u l k B and μ s l a b T i μ b u l k T i . Therefore, the range of chemical potential for B is as follows:
μ b u l k T i B 2 = μ b u l k T i + 2 μ b u l k B + Δ γ H T i B 2 , 1 2 Δ γ H T i B 2 μ s l a b B μ b u l k B 0
In the formula mentioned above, Δ γ H T i B 2 represents the formation enthalpy of TiB2 previously discussed. μ b u l k T i is the chemical potential of Ti in the bulk phase. μ b u l k B is the chemical potential of B in the α-B12 bulk phase. μ b u l k T i B 2 is the chemical potential of the TiB2 bulk phase.
The Al3Zr(001) surface model used in the calculations was also non-stoichiometric, hence the surface energy of the Al3Zr(001) surface could be determined using Equation (7):
γ Z r A l 3 001 = E s u r f Z r A l 3 001 N Z r A l 3 001 Z r × μ b u l k Z r A l 3 + 3 N Z r A l 3 001 Z r N Z r A l 3 001 A l × μ s l a b A l 2 × A s u r f Z r A l 3 001
Similarly, the range for the chemical potential of aluminum is as follows:
μ b u l k A l 3 Z r = μ b u l k Z r + 3 μ b u l k A l + Δ γ H A l 3 Z r ,   1 3 Δ γ H A l 3 Z r μ s l a b A l μ b u l k A l 0
Table 3 lists the surface energy data for different structures of the TiB2(0001) surface and the Al3Zr(001) surface under conditions rich in B and Al, respectively. The surface energies of the Ti-terminated and B-terminated ends of TiB2(0001) converged as the number of atomic layers increased, reaching convergence when the atomic layers were ≥9. Similarly, the Al-terminated end of Al3Zr(001) converged at the atomic layers ≥7; the ZrAl-terminated end converged when the atomic layers were ≥9. The smaller the surface energy value, the better the stability of the surface, thus the stability of the surface was consistent with the assessment of the interlayer distances. This implies that choosing a 9-layer TiB2(0001) in conjunction with a 7-layer ZrAl-terminated end and a 9-layer Al-terminated end of Al3Zr(001) to construct the TiB2/Al3Zr interface is feasible for subsequent research.
The surface energy of non-stoichiometric surfaces is influenced by the chemical potential difference between the surface atoms and the bulk atoms. Therefore, we calculated the surface energy for a 9-layer TiB2(0001) with different B chemical potential differences ( μ s l a b B μ b u l k B ) , and also for a 7-layer ZrAl-terminated and a 9-layer Al-terminated Al3Zr(001) with different Al chemical potential differences ( μ s l a b A l μ b u l k A l ) . The relationship between the calculated surface energies and the different chemical potential differences is shown in Figure 2. From Figure 2a, it can be seen that the surface energy range for the B-terminated TiB2(0001) was from 6.124 to 2.876 J/m2; for the Ti-terminated TiB2(0001), the range was from 2.106 to 5.354 J/m2. Under Ti-rich conditions, the surface energy of the Ti-terminated side was lower than that of the B-terminated side, indicating that the Ti-terminated TiB2(0001) surface was more stable at this time. However, as the B chemical potential difference increased, the surface energy of the Ti-terminated side gradually increased, while that of the B-terminated side gradually decreased. When the B chemical potential difference was greater than −0.613 eV, the B-terminated TiB2(0001) surface became more stable. This is consistent with the results reported in the literature [45]. From Figure 2b, it is evident that the surface energy of L12-Al3Zr(001) also showed a linear relationship with the chemical potential difference in Al. The surface energy range for the Al-terminated Al3Zr(001) was from 1.545 to 1.053 J/m2; for the ZrAl-terminated Al3Zr(001), the range was from 1.213 to 1.705 J/m2. The results indicate that as the chemical potential difference shifted from Zr-rich conditions to Al-rich conditions, the surface energy of the Al-terminated end showed a decreasing trend, while the surface energy of the ZrAl end showed an increasing trend. Within the range of −0.487 eV to −0.322 eV, the surface energy of the Al end was approximately lower than that of the ZrAl end, and in the range from −0.322 eV to 0 eV, the opposite was true. This suggests that the Al-terminated end of Al3Zr(001) is more stable in an aluminum-rich environment, while the ZrAl-terminated end is more stable in a zirconium-rich environment.

3.3. Interface Characteristics of TiB2(0001)/Al3Zr(001)

3.3.1. Atomic Structure of the Interface

Based on the results of the convergence tests, the TiB2/Al3Zr interface was constructed as follows: nine layers of TiB2(0001) were stacked on top of the Al3Zr(001) surface, with nine layers chosen for the Al-terminated surface of Al3Zr(001) and seven layers for the ZrAl-terminated surface. The mismatch degree of this interface was less than 1%, indicating that it was a typical coherent interface. Considering that the TiB2(0001) and Al3Zr(001) surfaces each have two different terminations (Ti-terminated, B-terminated and Al-terminated, ZrAl-terminated) and three different stacking sequences—top site (TS), bridge site (BS), and hollow site (HS)—a total of 12 interface models were established. These models are illustrated in Figure 3. The “interface name” was designated based on the first letter of the terminations and stacking order of TiB2(0001) and Al3Zr(001).

3.3.2. Interface Adhesion Work

The strength of the bonding at the interface can be described by the adhesion work. The adhesion work between the TiB2(0001) surface and the Al3Zr(001) surface can be determined using the following definition [46]:
W a d = E A l 3 Z r s l a b + E T i B 2 s l a b E A l 3 Z r T i B 2 t o t a l A
where E A l 3 Z r / T i B 2 t o t a l and A represent the total energy and the interface area of the TiB2(0001)/Al3Zr(001) interface, respectively. E T i B 2 s l a b and E A l 3 Z r s l a b are the energies of the relaxed TiB2(0001) and Al3Zr(001) slabs, respectively.
Table 4 displays the interfacial distances and corresponding adhesion work for 12 different interfaces. From Table 4, it is evident that the stacking position significantly impacted both the equilibrium distance and the adhesion work of the interfaces. Within the same stacking order, the adhesion work of the interfaces between the Ti-terminated and Al-terminated sides of the TiB2/Al3Zr was higher than that of other terminated interfaces, indicating better stability for interfaces involving the Ti termination compared with those with the Al termination.
For interfaces between the B-terminated and ZrAl-terminated sides, the adhesion work for the BZT, BZB, and BZH interfaces were very close, at 2.22, 2.22, and 2.29 J/m2, respectively, suggesting that the stacking order had a minimal impact on the bonding strength of these interfaces. However, for interfaces between the Ti-terminated and ZrAl-terminated sides, the effect of the stacking order was relatively more significant. The TZH interface, for example, had an adhesion work of 1.97 J/m2, which was substantially higher than that of the TZT and TZB interfaces, yet their interfacial distances were the opposite. This is because, in the TZH interface, the Ti atoms interacted with the surrounding two Zr atoms and two Al atoms, affecting the interfacial distance. Generally, the higher the W a d , the stronger the bonding force between the interface atoms, hence the TAB interface was the most stable among the 12 interface structures, possessing the greatest bonding strength.

3.3.3. Interfacial Energy

Interfacial energy is a key thermodynamic quantity used to measure the stability of an interface. Generally, the lower the interfacial energy, the more stable the interface structure. The formula for calculating the interfacial energy of the TiB2(0001)/Al3Zr interface is as follows [47,48]:
γ i n t = γ T i B 2 0001 + γ Z r A l 3 001 W a d
In the formula, γ T i B 2 ( 0001 ) represents the interfacial energy; γ T i B 2 ( 0001 ) and γ Z r A l 3 ( 001 ) , respectively, denote the surface energies of the TiB2(0001) and Al3Zr(001) surfaces; and W a d represents the adhesion work of the corresponding interface.
Based on the previous analysis, we have identified interface structures with the lowest interfacial energies for further study: the TAB, TZH, BAT, and BZH interfaces. Figure 4 illustrates the variation in interfacial energy for these four interfaces as a function of the chemical potential difference between boron in the slab and bulk ( μ s l a b B μ b u l k B ) . Throughout the range of ( μ s l a b B μ b u l k B ) , the interfacial energies for TAB, TZH, BAT, and BZH were, respectively, 1.730 to 4.430 J/m2, 1.848 to 5.509 J/m2, 2.455 to 6.118 J/m2, and 2.895 to 5.595 J/m2. Evidently, for interfaces on the titanium side, TAB exhibits the lowest interfacial energy and the highest interface stability, consistent with the results obtained from the adhesion work analysis. For interfaces on the boron side, under titanium-rich conditions, BZH displayed relatively lower interfacial energy, indicating a more stable interface structure. In contrast, under boron-rich conditions, BAT showed better thermodynamic stability.

3.4. Electron Structure

The bonding at interfaces fundamentally involves interactions among the electrons located at the interface. To further analyze the interactions within the TiB2(0001)/Al3Zr(001) interface, we selected representative models from four different terminated interfaces and calculated their differential charge density maps. These interfaces were the TAB, BAT, TZH, and BZH interfaces. Each of these interface models exhibited the highest adhesion work and the lowest interfacial energy within their respective terminated interface models. The differential charge density for the TiB2(0001)/Al3Zr(001) interface can be expressed as follows [47]:
Δ ρ = ρ t o a l ρ T i B 2 0001 ρ A l 3 Z r 001
In this formula, ρ t o a l is the total charge density of the TiB2(0001)/Al3Zr(001) interface; and ρ T i B 2 ( 0001 ) and ρ A l 3 Z r ( 001 ) represent the independent charge densities of the TiB2(0001) surface and the Al3Zr(001) surface that make up the interface structure, respectively.
Figure 5 shows the differential charge density results for the four interface models. Red and blue represent the accumulation and loss of charge, respectively. From Figure 5a, it can be seen that there was a significant charge accumulation between the Ti and Al atoms at the TAB interface, which was due to the loss of charge from the Al and Ti atoms. This indicates the formation of strong Ti−Al covalent bonds at the L12-Al3Ti(111) surface. Additionally, some charge accumulated between the Al and Zr atoms, suggesting the formation of Al−Zr covalent bonds. As shown in Figure 5b, electron transfer occurred between the Al and B atoms, indicating the presence of ionic bonds; furthermore, there was a small accumulation of charge between the Al and B atoms, and the presence of B-Al covalent bonds. Figure 5c describes the differential charge density at the TZH interface, clearly showing Ti-Zr covalent bonding. On the BZH interface, strong covalent bonds were formed between the Zr and B atoms. It is noteworthy that Zr formed covalent bonds with both Al atoms in the same layer and those in the second layer, but the covalent bond between Zr and Al in the same layer was stronger than that between Zr and the second-layer Al, which was related to the distances between the atoms.
To further investigate the electronic structure and analyze the bonding nature of the atoms at the TiB2(0001)/Al3Zr(001) interface, we further analyzed the density of states (DOSs) of the stable interfaces, with results shown in Figure 6. Notable distinctions were observed in the sharpness of the peaks associated with the same elements in the TiB2(0001) slab near the interface junction, and similarly in the Al3Zr(001) slab. This indicates that the bonding process in these four interface models was primarily contributed by the outer electrons of the first layer of atoms at the interface end.
Continuing the analysis with a focus on the TAB interface model, we aimed to understand the bonding characteristics of the interface. From Figure 6a, it is evident that the first layer of Ti at the TiB2(0001) end and the first layer of Al at the Al3Zr(001) end exhibited a peak between 0 and 5 eV, primarily contributed by the Ti 3d and Al 3p orbitals. Moreover, hybridization between Ti 3d and Al 3p also occurred around 10 eV. Notably, there was intense hybridization near 8 eV between the first layer of Ti at the TiB2(0001) end and the second layer of Zr at the Al3Zr(001) end, mainly contributed by the Ti 3d and Zr 4d orbitals. The peak intensities indicated strong hybridization at the TAB interface.
In Figure 6b, the hybridization between the first layer of B on the TiB2(0001) side and the first layer of Al on the Al3Zr(001) side was also evident. A distinct peak appeared near the Fermi level, mainly due to the contributions of the B 2p and Al 3s orbitals. Additionally, the 2p orbital of the first-layer B atoms showed a prominent peak around 5 eV, and there were indications of increased orbital density around 5 eV in the 3p orbital of the first-layer Al atoms. This suggests a possible overlap between the B 2p and Al 3p orbitals, potentially pointing to the presence of hybridization. Such hybridization could lead to the formation of strong B-Al covalent bonds at the interface.
Figure 6c shows significant hybridization between Ti 3d and Zr 4d in the 5 to 10 eV range, forming Ti-Zr covalent bonds. The interface composition in Figure 6d, which consisted of the first layer of B at the TiB2(0001) end and the first layer at the ZrAl end of Al3Zr(001), however, showed no significant common peaks, with only slight hybridization near the Fermi level.
The density of states analysis in Figure 6 revealed variations in the electronic states of Al atoms when adjacent to Zr compared with their presence in layers without Zr. This observation suggests differences in orbital interactions, likely due to hybridization, between Al and neighboring atoms under different compositional contexts. Notably, the analysis showed that the hybridization patterns were more pronounced in the TAB interface, where peak intensities suggested a stronger covalent character in the bonding interactions, particularly involving Al 3p and Zr 4d orbitals. This finding underscores the impact of atomic positioning and local chemical environment on the electronic properties at the interface. Therefore, it can be concluded that the hybridization at the TAB interface exhibited a higher intensity and complexity; based on the analysis of the density of states and differential charge density, the order of interface stability is as follows: TAB interface > BZH interface > TZH interface > BAT interface. This is consistent with the previous analyses of adhesion work and interfacial energy.

4. Conclusions

To understand the interfacial bonding characteristics of the TiB2(0001)/Al3Zr(001) interface, first-principles studies were conducted focusing on the adhesion work, interfacial energy, and electronic structure. This study considered 12 different TiB2/Al3Zr interfaces featuring four different terminations and three stacking sequences, with a detailed investigation on four of the most stable structures. The main findings of the study are summarized as follows:
(1)
Simulation tests have demonstrated that a stable interface structure can be formed when a 9-layer TiB2(0001) surface is combined with a 7-layer ZrAl-terminated and a 9-layer Al-terminated Al3Zr(001) surface.
(2)
For interfaces with the same termination in TiB2/Al3Zr, the study identifies the bridge-site stacking (TAB) at the T/Al termination, hollow-site stacking (TZH) at the Ti/ZrAl termination, top-site stacking (BAT) at the B/A termination, and hollow-site stacking at the B/ZrAl termination as the optimal structures. Among these interfaces, the TAB interface has the strongest adhesion strength due to the highest adhesion work of 2.41 J/m2.
(3)
The interfacial energies of TAB, TZH, BAT, and BZH across the entire range of ( μ s l a b B μ b u l k B ) are, respectively, 1.730 to 4.430 J/m2, 1.848 to 5.509 J/m2, 6.118 to 2.455 J/m2, and 5.595 to 2.895 J/m2. Evidently, TAB exhibits the lowest surface energy and the highest interface stability.
(4)
Electronic structure analysis indicates that the TAB, TZH, and BZH interfaces primarily feature covalent bonding, while the BAT interface shows both ionic and covalent bonds. According to the analysis of the electronic structure, the stability order of the TiB2(0001) and Al3Zr(001) interfaces is as follows: TAB interface > BZH interface > TZH interface > BAT interface.

Author Contributions

Conceptualization, H.N.; Methodology, H.N.; Software, J.L. and Z.X.; Validation, G.W.; Formal analysis, Z.C.; Investigation, W.Z.; Data curation, Z.Y.; Writing—original draft preparation, X.P. and L.Y.; Writing—review and editing, M.P.; Visualization, H.N.; Supervision, X.P.; Project administration, H.T.; Funding acquisition, H.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research work is supported by the National Natural Science Foundation of China-Guangxi Joint Fund (U20A20276), the Guangxi Natural Science Foundation (2021GXNSFAA220088, 2021GXNSFBA220080), the National Natural Science Foundation of China (51761002, 51961008), the Special Funds for Local Scientific and Technological Development Funds Guided by the Central Government in 2021 (GuiKeZY21195030), the Guangxi Science and Technology Base and Talent Project in 2022 (GuiKeAD21238010), the Key research and development project of Guangxi (GuiKeAB22080061), and the Research Project of Guangxi Key Laboratory of Processing for Non-ferrous Metallic and Featured Materials (2021GXMPSF06).

Data Availability Statement

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

Conflicts of Interest

Author Liujiang Yang was employed by the company Jiangsu Runyang Yueda Century Photovoltaic Technology. Author Mingjun Pang was employed by the company SAIC-GM-Wuling Automobile. Author Gaobao Wang was employed by the company Guangxi Youhe Metal Materials Technology Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Cheng, W.; Liu, C.; Ge, Z. Optimizing the mechanical properties of Al–Si alloys through friction stir processing and rolling. Mater. Sci. Eng. A 2021, 804, 140786. [Google Scholar] [CrossRef]
  2. Jin, S.; Luo, Z.; An, X.; Liao, X.; Li, J.; Sha, G. Composition-dependent dynamic precipitation and grain refinement in Al-Si system under high-pressure torsion. J. Mater. Sci. Technol. 2021, 68, 199–208. [Google Scholar] [CrossRef]
  3. Yin, J.; Li, Y.; Niu, Y.; Zhang, Z.; Song, D.; Chen, Y.; Liu, Y.; Zhong, Y. Effect of particle size on the microstructure and tensile property of the Al-28.5 Si alloy prepared by continuous powder extrusion. Mater. Res. Express 2021, 8, 026524. [Google Scholar] [CrossRef]
  4. Wang, F.; Qiu, D.; Liu, Z.-L.; Taylor, J.A.; Easton, M.A.; Zhang, M.-X. The grain refinement mechanism of cast aluminium by zirconium. Acta Mater. 2013, 61, 5636–5645. [Google Scholar] [CrossRef]
  5. Xu, J.-h.; Freeman, A. Bandfilling and structural stability of trialuminides: YAl3, ZrAl3, and NbAl3. J. Mater. Res. 1991, 6, 1188–1199. [Google Scholar] [CrossRef]
  6. Moon, K.I.; Kim, S.C.; Lee, K.S. A study on the microstructure of D023 Al3Zr and L12 (Al+ 12.5 at.% Cu) 3Zr intermetallic compounds synthesized by PBM and SPS. Intermetallics 2002, 10, 185–194. [Google Scholar] [CrossRef]
  7. Xu, J.-H.; Freeman, A. Phase stability and electronic structure of ScAl3 and ZrAl3 and of Sc-stabilized cubic ZrAl3 precipitates. Phys. Rev. B 1990, 41, 12553. [Google Scholar] [CrossRef]
  8. Yang, Q.-B.; Deng, Y.-J.; Mou, Y.; Zhang, Z.-Q.; Li, W.-G.; Qing, L. Effect of Al3Zr particles on hot-compression behavior and processing map for Al-Cu-Li based alloys at elevated temperatures. Trans. Nonferrous Met. Soc. China 2020, 30, 872–882. [Google Scholar] [CrossRef]
  9. Li, Y.; Lu, B.; Yu, W.; Fu, J.; Xu, G.; Wang, Z. Two-stage homogenization of Al-Zn-Mg-Cu-Zr alloy processed by twin-roll casting to improve L12 Al3Zr precipitation, recrystallization resistance, and performance. J. Alloys Compd. 2021, 882, 160789. [Google Scholar] [CrossRef]
  10. Wang, Y.; Wu, X.; Cao, L.; Tong, X.; Zou, Y.; Zhu, Q.; Tang, S.; Song, H.; Guo, M. Effect of Ag on aging precipitation behavior and mechanical properties of aluminum alloy 7075. Mater. Sci. Eng. A 2021, 804, 140515. [Google Scholar] [CrossRef]
  11. Yang, T.; Han, X.; Ding, Z.; Wang, Y.; Li, J. Electronic and structural properties of low-index L12–Al3Zr surfaces by first-principle calculations. Calphad 2019, 66, 101645. [Google Scholar] [CrossRef]
  12. Yang, T.; Wei, M.; Ding, Z.; Han, X.; Li, J. First-principle calculations on the Al/L12-Al3Zr heterogeneous nucleation interface. Calphad 2020, 69, 101768. [Google Scholar] [CrossRef]
  13. Yang, T.; Han, X.; Li, W.; Chen, X.; Liu, P. Angular dependent potential for Al-Zr binary system to study the initial heterogeneous nucleation behavior of liquid Al on L12-Al3Zr. Comput. Mater. Sci. 2023, 230, 112480. [Google Scholar] [CrossRef]
  14. Birol, Y. Effect of silicon content in grain refining hypoeutectic Al–Si foundry alloys with boron and titanium additions. Mater. Sci. Technol. 2012, 28, 385–389. [Google Scholar] [CrossRef]
  15. Wang, Y.; Fang, C.; Zhou, L.; Hashimoto, T.; Zhou, X.; Ramasse, Q.; Fan, Z. Mechanism for Zr poisoning of Al-Ti-B based grain refiners. Acta Mater. 2019, 164, 428–439. [Google Scholar] [CrossRef]
  16. Mao, G.; Tong, G.; Gao, W.; Liu, S.; Zhong, L. The poisoning effect of Sc or Zr in grain refinement of Al-Si-Mg alloy with Al-Ti-B. Mater. Lett. 2021, 302, 130428. [Google Scholar] [CrossRef]
  17. Fan, Z.; Wang, Y.; Zhang, Y.; Qin, T.; Zhou, X.; Thompson, G.; Pennycook, T.; Hashimoto, T. Grain refining mechanism in the Al/Al–Ti–B system. Acta Mater. 2015, 84, 292–304. [Google Scholar] [CrossRef]
  18. Tang, Y.; Chen, Z.; Borbely, A.; Ji, G.; Zhong, S.; Schryvers, D.; Ji, V.; Wang, H. Quantitative study of particle size distribution in an in-situ grown Al–TiB2 composite by synchrotron X-ray diffraction and electron microscopy. Mater. Charact. 2015, 102, 131–136. [Google Scholar] [CrossRef]
  19. Wang, Q.; Li, Y.; Chen, S.; Liu, X.; Chen, Z.; Wang, M.; Zhu, H.; Wang, H. Interface alloying design to improve the dispersion of TiB2 nanoparticles in Al composites: A first-principles study. J. Phys. Chem. C 2021, 125, 5937–5946. [Google Scholar] [CrossRef]
  20. Liu, Z.; Dong, Z.; Cheng, X.; Zheng, Q.; Zhao, J.; Han, Q. On the supplementation of magnesium and usage of ultrasound stirring for fabricating in situ TiB 2/A356 composites with improved mechanical properties. Metall. Mater. Trans. A 2018, 49, 5585–5598. [Google Scholar] [CrossRef]
  21. Li, Y.; Hu, B.; Liu, B.; Nie, A.; Gu, Q.; Wang, J.; Li, Q. Insight into Si poisoning on grain refinement of Al-Si/Al-5Ti-B system. Acta Mater. 2020, 187, 51–65. [Google Scholar] [CrossRef]
  22. Li, Y.; Jiang, Y.; Hu, B.; Li, Q. Novel Al-Ti-Nb-B grain refiners with superior efficiency for Al-Si alloys. Scr. Mater. 2020, 187, 262–267. [Google Scholar] [CrossRef]
  23. Xiao, H.-Y.; Li, Y.-G.; Geng, J.-W.; Li, H.-P.; Wang, M.-L.; Dong, C.; Li, Z.-G.; Wang, H.-W. Effects of nano-sized TiB2 particles and Al3Zr dispersoids on microstructure and mechanical properties of Al–Zn–Mg–Cu based materials. Trans. Nonferrous Met. Soc. China 2021, 31, 2189–2207. [Google Scholar] [CrossRef]
  24. Liu, S.-D.; Zhang, X.-M.; Chen, M.-A.; You, J.-H.; Zhang, X.-Y. Effect of Zr content on quench sensitivity of AIZnMgCu alloys. Trans. Nonferrous Met. Soc. China 2007, 17, 787–792. [Google Scholar] [CrossRef]
  25. Pang, X.; Yang, L.; Yang, J.; Pang, M.; Xu, Z.; Li, A.; Wei, B.; Tang, H. Understanding the poisoning mechanisms of Si and Zr atoms on L12 Al3Ti (111) surface: A first-principles investigation. Vacuum 2023, 210, 111891. [Google Scholar] [CrossRef]
  26. Wu, J.; Ruan, Q.; Chen, S.; Meng, C.; Xu, Z.; Wei, C.; Tang, H.; Wang, J. Insights into poisoning mechanism of Zr by first principle calculation on adhesion work and adsorption energy between TiB2, Al3Ti, and Al3Zr. Metals 2022, 12, 286. [Google Scholar] [CrossRef]
  27. Huang, J.; Xiang, Z.; Tang, Y.; Li, J.; Shen, G.; Shi, G.; Ma, X. Precipitation behavior of D023, L12 type Al3zr and interfacial characteristics in Tib2 reinforced Al-Zn-Mg-Cu-Zr composite. SSRN. Available online: https://papers.ssrn.com/abstract_id=4100289 (accessed on 17 October 2024).
  28. Huang, J.; Xiang, Z.; Tang, Y.; Li, J.; Shen, G.; Shi, G.; Ma, X.; Chen, Y.; Chen, Z. Interfacial precipitation of the in-situ TiB2-reinforced Al–Zn–Mg–Cu–Zr composite. Mater. Chem. Phys. 2023, 301, 127561. [Google Scholar] [CrossRef]
  29. Hu, H.; Zhao, M.; Wu, X.; Jia, Z.; Wang, R.; Li, W.; Liu, Q. The structural stability, mechanical properties and stacking fault energy of Al3Zr precipitates in Al-Cu-Zr alloys: HRTEM observations and first-principles calculations. J. Alloys Compd. 2016, 681, 96–108. [Google Scholar] [CrossRef]
  30. Segall, M.; Lindan, P.J.; Probert, M.a.; Pickard, C.J.; Hasnip, P.J.; Clark, S.; Payne, M. First-principles simulation: Ideas, illustrations and the CASTEP code. J. Phys. Condens. Matter 2002, 14, 2717. [Google Scholar] [CrossRef]
  31. Vanderbilt, D. Soft self-consistent pseudopotentials in a generalized eigenvalue formalism. Phys. Rev. B 1990, 41, 7892. [Google Scholar] [CrossRef]
  32. Wu, Z.; Cohen, R.E. More accurate generalized gradient approximation for solids. Phys. Rev. B 2006, 73, 235116. [Google Scholar] [CrossRef]
  33. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188. [Google Scholar] [CrossRef]
  34. Head, J.D.; Zerner, M.C. A Broyden—Fletcher—Goldfarb—Shanno optimization procedure for molecular geometries. Chem. Phys. Lett. 1985, 122, 264–270. [Google Scholar] [CrossRef]
  35. Zhang, H.; Xiong, H.; Wang, D.; Wang, W. Mechanism of NbC heterogeneous nucleation on TiN in microalloyed steel: A first-principles study. Comput. Mater. Sci. 2018, 146, 126–133. [Google Scholar] [CrossRef]
  36. Wang, Q.; Liu, C.; Yao, R.; Zhu, H.; Liu, X.; Wang, M.; Chen, Z.; Wang, H. First-principles study on the stability and work function of low-index surfaces of TiB2. Comput. Mater. Sci. 2020, 172, 109356. [Google Scholar] [CrossRef]
  37. Liu, R.; Yin, X.; Feng, K.; Xu, R. First-principles calculations on Mg/TiB2 interfaces. Comput. Mater. Sci. 2018, 149, 373–378. [Google Scholar] [CrossRef]
  38. Topor, L.; Kleppa, O. Enthalpies of formation of first-row transition-metal diborides by a new calorimetric method. J. Chem. Thermodyn. 1985, 17, 1003–1016. [Google Scholar] [CrossRef]
  39. Ghosh, G.; Vaynman, S.; Asta, M.; Fine, M. Stability and elastic properties of L12-(Al, Cu) 3 (Ti, Zr) phases: Ab initio calculations and experiments. Intermetallics 2007, 15, 44–54. [Google Scholar] [CrossRef]
  40. Desch, P.; Schwarz, R.; Nash, P. Formation of metastable L12 phases in Al3Zr and Al-12.5% X-25% Zr (X ≡ Li, Cr, Fe, Ni, Cu). J. Less Common Met. 1991, 168, 69–80. [Google Scholar] [CrossRef]
  41. Li, Y.; Gao, Y.; Xiao, B.; Min, T.; Fan, Z.; Ma, S.; Yi, D. Theoretical study on the electronic properties and stabilities of low-index surfaces of WC polymorphs. Comput. Mater. Sci. 2011, 50, 939–948. [Google Scholar] [CrossRef]
  42. Rapcewicz, K.; Chen, B.; Yakobson, B.; Bernholc, J. Consistent methodology for calculating surface and interface energies. Phys. Rev. B 1998, 57, 7281. [Google Scholar] [CrossRef]
  43. Batyrev, I.; Alavi, A.; Finnis, M.W. Ab initio calculations on the Al2O3 (0001) surface. Faraday Discuss. 1999, 114, 33–43. [Google Scholar] [CrossRef]
  44. Song, M.; Huang, B.; Zhang, M.; Li, J. Study of formation behavior of TiC ceramic obtained by self-propagating high-temperature synthesis from Al–Ti–C elemental powders. Int. J. Refract. Met. Hard Mater. 2009, 27, 584–589. [Google Scholar] [CrossRef]
  45. Han, Y.; Dai, Y.; Shu, D.; Wang, J.; Sun, B. First-principles study of TiB2 (0001) surfaces. J. Phys. Condens. Matter 2006, 18, 4197. [Google Scholar] [CrossRef] [PubMed]
  46. Jin, W.; Li, L.; Zhang, S.; Yang, H.; Gao, K.; Pang, X.; Volinsky, A.A. First principles calculations of interfacial properties and electronic structure of the AlN (0 0 0 1)/Ti (0 0 0 1) interface. Chem. Phys. Lett. 2018, 713, 153–159. [Google Scholar] [CrossRef]
  47. Pang, X.; Yang, J.; Pang, M.; He, J.; Yang, W.; Qin, H.; Zhan, Y. Theoretical understanding of atomic and electronic structures of the ZrC (111)/Cu (111) interface. J. Alloys Compd. 2019, 791, 431–437. [Google Scholar] [CrossRef]
  48. Xiong, H.; Liu, Z.; Zhang, H.; Du, Z.; Chen, C. First principles calculation of interfacial stability, energy and electronic properties of SiC/ZrB2 interface. J. Phys. Chem. Solids 2017, 107, 162–169. [Google Scholar] [CrossRef]
Figure 1. Crystal structures of TiB2 and Al3Zr: (a) TiB2; (b) Al3Zr.
Figure 1. Crystal structures of TiB2 and Al3Zr: (a) TiB2; (b) Al3Zr.
Crystals 14 00979 g001
Figure 2. Relationship between the surface energy and chemical potential: (a) the function of surface energy with boron chemical potential difference for the Ti-terminated and B-terminated TiB2 surfaces; (b) the function of surface energy with aluminum chemical potential difference for the Al-terminated and ZrAl-terminated Al3Zr surfaces. Vertical dashed lines indicate the stability range of the compounds.
Figure 2. Relationship between the surface energy and chemical potential: (a) the function of surface energy with boron chemical potential difference for the Ti-terminated and B-terminated TiB2 surfaces; (b) the function of surface energy with aluminum chemical potential difference for the Al-terminated and ZrAl-terminated Al3Zr surfaces. Vertical dashed lines indicate the stability range of the compounds.
Crystals 14 00979 g002
Figure 3. Twelve models for the TiB2/Al3Zr interface: (a) TAT interface; (b) TAB interface; (c) TAH interface; (d) BAT interface; (e) BAB interface; (f) BAH interface; (g) TZT interface; (h) TZB interface; (i) TZH interface; (j) BZT interface; (k) BZB interface; (l) BZH interface.
Figure 3. Twelve models for the TiB2/Al3Zr interface: (a) TAT interface; (b) TAB interface; (c) TAH interface; (d) BAT interface; (e) BAB interface; (f) BAH interface; (g) TZT interface; (h) TZB interface; (i) TZH interface; (j) BZT interface; (k) BZB interface; (l) BZH interface.
Crystals 14 00979 g003aCrystals 14 00979 g003b
Figure 4. The relationship between the twelve types of TiB2(0001)/Al3Zr(001) interfaces and varying chemical potentials of boron.
Figure 4. The relationship between the twelve types of TiB2(0001)/Al3Zr(001) interfaces and varying chemical potentials of boron.
Crystals 14 00979 g004
Figure 5. Differential charge density of the TiB2(0001)/Al3Zr(001) interfaces: (a) TAB interface; (b) BAT interface; (c) TZH interface; (d) BZH interface.
Figure 5. Differential charge density of the TiB2(0001)/Al3Zr(001) interfaces: (a) TAB interface; (b) BAT interface; (c) TZH interface; (d) BZH interface.
Crystals 14 00979 g005
Figure 6. Partial density of states for the TiB2(0001)/Al3Zr(001) interfaces: (a) TAB interface; (b) BAT interface; (c) TZH interface; (d) BZH interface. Vertical dashed lines indicate the Fermi level.
Figure 6. Partial density of states for the TiB2(0001)/Al3Zr(001) interfaces: (a) TAB interface; (b) BAT interface; (c) TZH interface; (d) BZH interface. Vertical dashed lines indicate the Fermi level.
Crystals 14 00979 g006aCrystals 14 00979 g006b
Table 1. Calculated and experimental lattice constants and formation energies of bulk TiB2 and bulk Al3Zr.
Table 1. Calculated and experimental lattice constants and formation energies of bulk TiB2 and bulk Al3Zr.
PhasesMethodLattice Constant (Å) Δ r H ( e V / a t o m )
abc
TiB2Present3.0263.0263.229−1.072
Calculation a3.0343.0343.224−1.079
Calculation b3.0303.0303.221−1.085
Experiment c3.0323.0323.229−1.134
Al3ZrPresent4.0824.0824.082−0.487
Calculation d4.1174.1174.117−0.471
Calculation e4.0964.0964.096−0.466
Experiment f4.0904.0904.090-
a Ref. [36] GGA-PBE. b Ref. [37] GGA-PBE. c Ref. [38]. d Ref. [11] GGA-PBE. e Ref. [39] GGA-PW91. f Ref. [40].
Table 2. Interlayer relaxation percentages for TiB2(0001) and Al3Zr(001) surfaces at different slab thicknesses.
Table 2. Interlayer relaxation percentages for TiB2(0001) and Al3Zr(001) surfaces at different slab thicknesses.
SurfaceTerminationInterlayerSlab Thickness, n
357911
TiB2(0001)TiΔ12−6.96−7.69−7.54−7.73−7.63
Δ23 −1.61−2.09−2.00−1.86
Δ34 −1.98−2.51−2.29
Δ45 −2.72−2.14
Δ56 −2.73
BΔ12−8.15−10.20−9.53−9.61−9.71
Δ23 0.470.11−0.110.02
Δ34 −2.31−2.24−2.40
Δ45 −3.16−2.69
Δ56 −2.71
Al3Zr(001)AlΔ12−1.453.021.330.630.91
Δ23 −1.52−1.201.681.05
Δ34 0.652.892.62
Δ45 −0.81−0.46
Δ56 0.67
Zr-AlΔ12−15.33−18.17−16.38−16.68−16.92
Δ23 −4.95−5.10−4.43−4.73
Δ34 4.282.513.21
Δ45 0.09−0.39
Δ56 0.89
Table 3. The surface energy (J/m2) calculation results of TiB2(0001) and Al3Zr(001) surfaces.
Table 3. The surface energy (J/m2) calculation results of TiB2(0001) and Al3Zr(001) surfaces.
LayerTiB2(0001)Al3Zr(001)
Ti-TerminatedB-TerminatedZrAl-TerminatedAl-Terminated
35.4423.0141.7681.193
55.4472.9871.7211.113
75.3922.9141.7071.073
95.3542.8761.7051.053
115.3052.8361.6371.017
135.2642.7791.6861.062
Table 4. The comparison of interfacial distance (d0) and adhesion work (Wad) for 12 different interfaces.
Table 4. The comparison of interfacial distance (d0) and adhesion work (Wad) for 12 different interfaces.
Terminations of TiB2(0001)Terminations of Al3Zr(001)Stacking SequenceInterface Named0 (Å)Wad (J/m2)
TiAlTSTAT2.322.37
BSTAB2.342.41
HSTAH2.182.40
TiZr-AlTSTZT2.601.26
BSTZB2.871.26
HSTZH2.411.97
BAlTSBAT2.071.92
BSBAB1.981.87
HSBAH1.921.88
BZr-AlTSBZT2.272.22
BSBZB2.162.22
HSBZH2.182.29
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

Pang, X.; Yang, L.; Nong, H.; Pang, M.; Wang, G.; Li, J.; Chen, Z.; Zeng, W.; Xiao, Z.; Yang, Z.; et al. Understanding the Interface Characteristics Between TiB2(0001) and L12-Al3Zr(001): A First-Principles Investigation. Crystals 2024, 14, 979. https://doi.org/10.3390/cryst14110979

AMA Style

Pang X, Yang L, Nong H, Pang M, Wang G, Li J, Chen Z, Zeng W, Xiao Z, Yang Z, et al. Understanding the Interface Characteristics Between TiB2(0001) and L12-Al3Zr(001): A First-Principles Investigation. Crystals. 2024; 14(11):979. https://doi.org/10.3390/cryst14110979

Chicago/Turabian Style

Pang, Xingzhi, Loujiang Yang, Hang Nong, Mingjun Pang, Gaobao Wang, Jian Li, Zhenchao Chen, Wei Zeng, Zhihang Xiao, Zengxiang Yang, and et al. 2024. "Understanding the Interface Characteristics Between TiB2(0001) and L12-Al3Zr(001): A First-Principles Investigation" Crystals 14, no. 11: 979. https://doi.org/10.3390/cryst14110979

APA Style

Pang, X., Yang, L., Nong, H., Pang, M., Wang, G., Li, J., Chen, Z., Zeng, W., Xiao, Z., Yang, Z., & Tang, H. (2024). Understanding the Interface Characteristics Between TiB2(0001) and L12-Al3Zr(001): A First-Principles Investigation. Crystals, 14(11), 979. https://doi.org/10.3390/cryst14110979

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