Next Article in Journal
Non-Waste Technology for Utilization of Tree Branches
Previous Article in Journal
A Fault Diagnosis Model for Tennessee Eastman Processes Based on Feature Selection and Probabilistic Neural Network
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

First-Principles Investigation of the Shear Properties and Sliding Characteristics of c-ZrO2(001)/α-Al2O3(11¯02) Interfaces

State Key Laboratory for Mechanical Structure Strength and Vibration, Department of Engineering Mechanics, Xi’an Jiaotong University, Xi’an 710049, China
*
Author to whom correspondence should be addressed.
Appl. Sci. 2022, 12(17), 8869; https://doi.org/10.3390/app12178869
Submission received: 22 July 2022 / Revised: 29 August 2022 / Accepted: 2 September 2022 / Published: 4 September 2022
(This article belongs to the Section Surface Sciences and Technology)

Abstract

:
The ideal mechanical shear properties and sliding characteristics of c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interfaces are examined through simulated shear deformations using first-principles calculations. We investigate three types of interface models, abbreviated as O-, 2O-, and Zr- models, when shear displacements are applied along the < 1 1 ¯ 01 > and < 11 2 ¯ 0 > directions of their Al2O3 lattice. The theoretical shear strength and unstable stacking energy of the ZrO2/Al2O3 interfaces are discussed. In the process of the ZrO2/Al2O3 interfacial shear deformation, we find that the sliding of the ZrO2 atomic layers, accompanied by the shifting of Zr atoms and Al atoms near the interface, plays a dominant role; in addition, the ZrO2/Al2O3 interfaces fail within the ZrO2 atomic layer. Among the three models, the O- model exhibits the strongest shear resistance; whereas the Zr- model is the most prone to slip. Furthermore, their tensile and shear strengths are compared; moreover, their potential applications are provided.

1. Introduction

Thermal barrier coatings (TBCs) are widely used in the protection of hot components in gas turbine engines; they can reduce the corrosive degradation of turbine blades and improve the efficiency of the engine [1,2,3,4,5,6,7]. Typically, TBCs consist of four material layers: (i) superalloy substrate; (ii) metallic bond coat (BC); (iii) thermally grown oxide (TGO); and (iv) ceramic top coat (TC) [8].
Yttria-stabilized zirconia (YSZ) is a frequently used TC material owing to its low thermal conductivity and high thermal expansion coefficient (TEC). When TBCs operate at high temperatures, there are many factors affecting their lifetime; such as the growth of TGO [9,10], phase transformation [11,12], the influence of calcium-magnesium-alumina-silicate (CMAS) [13,14], sintering at elevated temperatures [15,16], and residual stress [17,18]. The lifetime of TBCs at high temperatures is reduced for the above factors, so that the TBCs cannot efficiently protect turbine engines. Guan et al. proposed a master–slave model with a fluid-thermo-structure.
(FTS) interaction for the thermal fatigue life prediction of TBCs [19]. As the thickness of the TGO increased, the stresses at the TC/TGO and TGO/BC interfaces increased; resulting in the fracture of the TC/TGO or TGO/BC interfaces [20]. It is widely recognized that the lifetime of TBCs at high temperatures is mainly affected by the mechanical properties of the TC/TGO and TGO/BC interfaces. Investigating the mechanical properties of interfaces directly from experiments remains a challenge because of the lack of experimental conditions and difficulty of sample preparation. Atomistic modeling and simulation, which provide the intrinsic atomic properties of interface structures, are often used to investigate the mechanical properties of interfaces and identify their failure mechanisms [21]. The shear strength and sliding behavior of the Ni(111)/α-Al2O3(0001) interface were investigated by Guo et al. [22] using first-principles calculations. The ideal mechanical strength of the ZrO2(111)/Ni(111) interface was calculated through simulated tensile and shear deformations using first-principles calculations [23]. However, despite these previous efforts, the mechanical strength and shear behavior of ZrO2/Al2O3 interfaces under first-principles shear simulation are still not well examined. The difficulty of this type of research is mainly due to the complexity of the ceramic/ceramic interface structures and the large number of computations involved in the shear simulation process.
In this study, shear simulations of c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interfaces were carried out using first-principles calculations. The shear directions of the three models were applied along the < 1 1 ¯ 01 > and < 11 2 ¯ 0 > directions of the Al2O3 lattice. For all three models, we examined the ideal shear strength, unstable stacking energy, and evolution of the atomic structures to clarify the sliding mechanisms of the c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interfaces. With these results, we further compared their shear and tensile properties to clarify the overall features of the c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) system.

2. Methodology

In this study, we examine c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interfaces using the SIESTA [24] code based on density functional theory (DFT). We adopted the all-electron projector augmented wave (PAW) method within the generalized gradient approximation (GGA) for the electron exchange and correlation. For our calculations, a 5 × 5 × 1 Monkhost-Pack scheme was used for the k-point sampling of the Brillouin zone. The cutoff energy for the plane-wave basis was 500 eV, and the self-consistent energy relaxation criterion for electron was 10−5 eV. Relaxation is performed using the conjugate gradient (CG) algorithm, and the convergence criterion is a Hellman–Feynman force of less than 0.01 eV/Å.
Ideal coherent c-ZrO2/α-Al2O3 interface models are discussed in ref. [25]. The c-ZrO2 surface was cut from its crystal on the (001) plane, the α-Al2O3 surface was cut from its crystal on the ( 1 1 ¯ 02 ) plane, and the lattice misfit of the c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interface was acceptable [26]. The layering sequence in the α-Al2O3( 1 1 ¯ 02 ) surface termination is |O—Al—O—Al—O|O—Al—O—Al—O|... This α-Al2O3( 1 1 ¯ 02 ) surface termination is a compact surface with the lowest energy. Compared to the α-Al2O3( 1 1 ¯ 02 ) surface termination, the c-ZrO2(001) surface termination is more complex. The layering sequence in the c-ZrO2(001) surface termination was Zr-O-Zr-..., O-Zr-O-..., and 2O-Zr-... [27]. In this study, three types of c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interface models were considered. In the following, these three models are named Zr-, O-, and 2O-, respectively. Figure 1 shows a schematic of the c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interfacial shear simulation. The supercell consists of a slab of α-Al2O3( 1 1 ¯ 02 ) sandwiched between two slabs of c-ZrO2(001); furthermore, this unit cell is orthogonal, 5.11 Å × 4.86 Å in dimension. According to preliminary calculations, the atomic relaxation layers of the c-ZrO2 slab and the α-Al2O3 slab were determined to be three and five layers, respectively. The atom layers outside the relaxation layers are considered to be rigid regions, represented by the two shaded rectangles in Figure 1; namely, the c-ZrO2 and α-Al2O3 blocks. The c-ZrO2 block moves a distance of δ relative to the α-Al2O3 block along the < 1 1 ¯ 01 > or < 11 2 ¯ 0 > direction in the α-Al2O3 lattice. The atoms outside the rectangles were then relaxed until the convergence criterion was satisfied.

3. Results and Discussion

3.1. Shear Strength Parameters

Figure 2 shows the shear stress–shear displacement curves for the c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interface. The peak shear stresses are 5.04 and 5.69 GPa at the critical shear displacements; and δc, 1.43 Å, and 1.85 Å along the < 1 1 ¯ 01 > and < 11 2 ¯ 0 > slip directions of the O- model (see Figure 2a,b), respectively. In addition, for the 2O- model, the peak shear stresses along the < 1 1 ¯ 01 > and < 11 2 ¯ 0 > directions are 3.66 and 4.52 GPa at 1.43 and 2.24 Å (see Figure 2c,d), respectively. Furthermore, for the Zr- model shown in Figure 2e,f, the peak shear stresses for the Zr- model are 2.94 and 3.95 GPa at 1.23 and 2.24 Å in the < 1 1 ¯ 01 > and < 11 2 ¯ 0 > directions, respectively. Compared with the 2O- and Zr- shear deformation models, the peak shear stress in the O- model is greater in the same slip direction. This implies that the termination of c-ZrO2(001) has a significant influence on the adhesive strength of the interface. In addition, the peak shear stress in the < 11 2 ¯ 0 > direction is higher than that in the < 1 1 ¯ 01 > direction for all three models. This indicates that the < 11 2 ¯ 0 > direction has a stronger shear resistance than the < 1 1 ¯ 01 > direction in the interface slip process. In Figure 2a–f, the maximum shear stress values in region II are 3.13, 5.11, 2.65, 2.98, 2.51, and 2.59 GPa, respectively. The maximum shear stress value in region I is larger than that in region II in the same model.
The shear moduli of the α-Al2O3 single crystals obtained from the calculations and experiments are 168 GPa and 163 GPa, respectively [28,29]. The shear moduli of the c-ZrO2 single crystals obtained by calculation and experiment are 106 and 81 GPa, respectively [30,31]. The maximum theoretical shear modulus of the supercell is estimated to be 38.7 GPa. Compared with the shear moduli of the α-Al2O3 and c-ZrO2 single crystals, this value is much smaller. These results indicate that the c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interfaces are weaker than those of bulk α-Al2O3 and c-ZrO2. Furthermore, this implies that the c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interfaces are not simply complements of bulk α-Al2O3 and c-ZrO2.
The energy per unit area of the slip plane Φ is defined as Φ = τ d δ ; τ and δ are the shear stress and displacement, respectively. The maximum value of Φ is termed the unstable stacking energy γus [32], which is the area between δ = 0 and the first δ = δt at which τ = 0 again. It is an important parameter for characterizing the interfacial shear strength. The mechanical properties of ZrO2/Al2O3 interfaces can be more comprehensively understood using the ideal shear strength combined with the unstable stacking energy. As shown in Figure 3, the unstable stacking energy γus of the O- model is larger than that of the other two models in the same slip direction; whereas the unstable stacking energy γus in the < 11 2 ¯ 0 > direction is higher than that in the < 1 1 ¯ 01 > direction for all three models. For the O- model, the calculated unstable stacking energies γus along the < 1 1 ¯ 01 > and < 11 2 ¯ 0 > slip directions are 399 and 612 mJ/m2, respectively. For the 2O- model, the calculated unstable stacking energies γus along the < 1 1 ¯ 01 > and < 11 2 ¯ 0 > slip directions are 349 and 494 mJ/m2, respectively. Furthermore, the calculated unstable stacking energies γus for the Zr- model along the < 1 1 ¯ 01 > and < 11 2 ¯ 0 > slip directions are 245 and 437 mJ/m2, respectively.
The mechanical properties of the three interface models can be analyzed by combining the calculated ideal shear strength with the unstable stacking energy. The present results indicate that the O- model has the strongest shear resistance, and the Zr- model is the most prone to slip among the three models. It can also be concluded from the results that the shear resistance in the < 11 2 ¯ 0 > slip direction is higher than that in the < 1 1 ¯ 01 > direction in the same model. The O-< 11 2 ¯ 0 > model c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interface may be useful for optimizing TBC structures to meet harsh requirements.

3.2. Failure Mechanism

The evolution of the atomic structures at the c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interface during the shearing process is shown in Figure 4. The atom indices Zr1 and Zr2 denote the first Zr layer atom and second Zr layer in the c-ZrO2 block from the interface, respectively. The same procedure is applied to Al1 and Al2. The capital letters of the markers in Figure 4 correspond to those in Figure 2. The failure process of the interfaces is investigated by combining Figure 2 and Figure 4.
From Figure 2, it can be seen that the shear stresses exhibit an obvious rebound phenomenon in region I. The shear stress increases gradually until it reaches the first peak value (point B), drops slightly to point C, and then continues to increase to the maximum value (point D). This rebound is mainly due to the shifting of Zr atoms, which represents the process of breaking and rebonding of the specific Zr-O and Zr-Al bonds near the interface. For the O-< 1 1 ¯ 01 > model, Figure 2a shows the breaking of the Zr1-O2 and Zr2-Al2 bonds; it also shows the rebonding of the Zr1-O1 and Zr1-Al2 bonds. For the O-< 11 2 ¯ 0 > model shown in Figure 2b, the Zr1-O1 and Zr1-O3 bonds break; in addition, the Zr1-Al2, Zr1-O2, and Zr2-O4 bonds rebond. For the 2O-model, the Zr1-O3 and Al1-O2 bonds break; the Zr1-O1, Zr1-O4, Zr1-Al1, and Al2-O2 bonds rebond along the < 1 1 ¯ 01 > direction, as shown in Figure 2c; the Zr1-O3, Zr1-O4, and Zr1-O5 bonds break; and the Zr-O1 and Zr1-Al1 bonds rebond along the < 11 2 ¯ 0 > direction (Figure 2d). For the Zr-model, the Zr2-O3 bond breaks, the Zr2-O2 bond rebonds, and the Zr2-Al2 bond rebonds along the < 1 1 ¯ 01 > direction (Figure 2e); the Zr2-O1 bond breaks and the Zr2-Al2 bond rebonds along the < 11 2 ¯ 0 > direction, as shown in Figure 2f.
The ZrO2 atomic layers slide along the shear direction, accompanied by the shifting of Zr atoms, until the shear stresses reach the maximum value; this is shown in Figure 4 from A to D. The ZrO2 atom layers undergo reconstruction from D to E, as shown in Figure 4. It is apparent that the models failed within the ZrO2 atomic layer. The shear stresses decrease rapidly after point D in Figure 2, mainly because the ZrO2 atom layers reconstruct intensively in the direction opposite to that of the shear accompanying the shifting of the Al atoms. Furthermore, the interfacial atomic configuration is transformed near both stress peaks; this indicates that the reconstruction of the interface structure has a significant influence on the mechanical strength of the interface region to resist sliding deformation. This phenomenon is consistent with the conclusions of Farkas [33]. Thus, it is concluded that in the process of the ZrO2/Al2O3 interfacial shear deformation, the sliding of the ZrO2 atomic layers, accompanied by the shifting of Zr and Al atoms near the interface, will play a dominant role. According to the comparison results of the maximum shear stress in regions I and II, it can also be concluded that the mechanical strength of the reconstructed interfaces is lower than that of the initial interfaces. Therefore, the reconstructed interfaces are more unstable and prone to slip than the initial interface.

3.3. Discussion

In this study, shear simulations of the c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interface models along the < 1 1 ¯ 01 > and < 11 2 ¯ 0 > directions of the α-Al2O3 lattice were performed using first-principles calculations. The O-< 11 2 ¯ 0 > model has the strongest shear resistance and the Zr-< 1 1 ¯ 01 > model is the most prone to slip. It indicates that the mechanical properties of c-ZrO2(001)/α-Al2O3 ( 1 1 ¯ 02 ) interfaces are affected by the interfacial atomic configuration and shear direction. In the process of the ZrO2/Al2O3 interfacial shear deformation, the ZrO2 atomic layers slide along the shear direction, accompanied by the shifting of Zr and Al atoms near the interface; moreover, shear failure generally occurs in the ZrO2 atomic layers. Furthermore, the reconstruction of the interface structure has a significant influence on the mechanical strength of the interface region that resists sliding deformation. In the Ni(111)/α-Al2O3 (0001) interface, Guo et al. [22] also observed the similar phenomenon. However, in the Al-terminated O-site interface, they observed that the maximum shear strength was along the Ni <110> direction; while the maximum unstable stacking energy was along the Ni <112> direction. This an intriguing observation; the reason for this phenomenon is the shifting of Al atoms near the interface during the sliding process. For our models, the shifting of Zr and Al atoms near the interface makes the sliding process more stable. In addition, the shear strength of the Ni/Al2O3 interfaces and the ZrO2/Al2O3 interfaces are weak; indicating that shear failure occurs at any interface in TBC.
We examined the ZrO2/Al2O3 interfaces using first-principle calculations to understand the tensile strength and failure properties [25]. In that work, the ideal tensile strengths and work of separation of the c-ZrO2(001)/α-Al2O3 ( 1 1 ¯ 02 ) interfaces were obtained. Table 1 summarizes the ideal tensile strength, the work of separation, the ideal shear strength, and the unstable stacking energy. For all the models in Table 1, the O-< 11 2 ¯ 0 > model is the most stable model; having the maximum tensile strength and shear strength. The Zr-< 1 1 ¯ 01 > model is the most unstable model; it has the lowest tensile strength and shear strength. These properties are closely related to the strength of the bonds at the interface. Compared with the tensile strengths, the shear strengths of the same model are much smaller. It indicates that ZrO2/Al2O3 interfaces are more prone to shear failure than tensile failure. This is because of the different failure mechanisms between tensile failure and shear failure. Owing to the tensile failure of the interface, all the atomic bonds across the interface need to be broken; whereas shear failure only needs to be partially broken. Therefore, shear failure requires less stress and energy than tensile failure does.
The cohesive zone model (CZM) is a widely accepted model in the elastic–plastic fracture mechanics community. The core task of developing a material-dependent CZM is to extract material parameters, such as the cohesive energy and maximum cohesive strength in the cohesive laws. In the continuum mechanics framework, there are many ways to develop cohesive zone models; in addition, more accurate cohesive zone models can be constructed based on first-principles results [34,35,36]. Figure 5 shows the shear stress-displacement curves fitted by the Fourier series; the fitting curve is basically consistent with the exponential cohesive zone model [37] in the inset. We chose the O- and Zr- models for the highest and lowest mechanical strengths, respectively. The cohesive shear strength τ’max and work of shear separation ϕt in the exponential CZM can be identified by the ideal shear strength τmax and unstable stacking energy γus, respectively. Hence, it is necessary to determine the mechanical properties of the interface, such as the theoretical tensile strength, work of separation, theoretical shear strength, and unstable stacking energy.
In this study, shear simulations of c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interface models along the < 1 1 ¯ 01 > and < 11 2 ¯ 0 > directions of the α-Al2O3 lattice were performed using first-principles calculations. In practice, c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interfaces may slip along other directions and undergo different structural transformations. Therefore, further studies on the shear behavior of c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interfaces along other directions can provide more information for optimizing the TBC structure to meet the harsh requirements. In addition, the shear simulation method implemented in this study is the constrained shear method; which produces normal stress in the simulation process. Umeno [38] observed that the mechanical properties of silicon are affected by the internal displacement of atoms and normal stresses inside the cell. Based on this, we can investigate the effect of normal stresses on the ideal shear strength of the c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interface using the unconstrained shear method.

4. Conclusions

Shear simulations of the c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interfaces were carried out using first-principles calculations. The O- model sliding along the Al2O3 < 11 2 ¯ 0 > direction is the most stable, with a maximum shear strength of 5.69 GPa and an unstable stacking energy of 612 mJ/m2. The Zr- model sliding along the Al2O3 < 1 1 ¯ 01 > direction is the most unstable; it has the minimum shear strength (2.94 GPa) and minimum unstable stacking energy (245 mJ/m2). In the process of c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interfacial shear deformation, the sliding of the Zr atomic layers, accompanied by the shifting of Zr atoms and Al atoms near the interface, plays a dominant role; in addition, shear failure generally occurs in the ZrO2 atomic layers. The shear strength values of the c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interface are much smaller than the theoretical tensile strength in the same model; this indicates that the ZrO2(001)/Al2O3 interfaces are more prone to shear failure than tensile failure.

Author Contributions

Conceptualization, Z.B. and F.S.; methodology, Z.B. and F.S.; software, Z.B.; validation, Z.B. and F.S.; formal analysis, Z.B.; investigation, Z.B.; resources, F.S.; data curation, Z.B.; writing—original draft, Z.B.; writing—review and editing, F.S.; visualization, Z.B.; supervision, F.S.; project administration, F.S.; funding acquisition, F.S. All authors have read and agreed to the published version of the manuscript.

Funding

This study was partially supported by the Natural Science Foundation of China through Grant no. 12072248.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Evans, A.; Hutchinson, J. The mechanics of coating delamination in thermal gradients. Surf. Coat. Technol. 2007, 201, 7905–7916. [Google Scholar] [CrossRef]
  2. Cheng, B.; Yang, N.; Zhang, Q.; Zhang, M.; Zhang, Y.-M.; Chen, L.; Yang, G.-J.; Li, C.-X.; Li, C.-J. Sintering induced the failure behavior of dense vertically crack and lamellar structured TBCs with equivalent thermal insulation performance. Ceram. Int. 2017, 43, 15459–15465. [Google Scholar] [CrossRef]
  3. Venkadesan, G.; Muthusamy, J. Experimental investigation of Al2O3/8YSZ and CeO2/8YSZ plasma sprayed thermal barrier coating on diesel engine. Ceram. Int. 2019, 45, 3166–3176. [Google Scholar] [CrossRef]
  4. Chen, L.; Gao, L.-L.; Yang, G.-J. Imaging Slit Pores Under Delaminated Splats by White Light Interference. J. Therm. Spray Technol. 2018, 27, 319–335. [Google Scholar] [CrossRef]
  5. Bolelli, G.; Righi, M.G.; Mughal, M.Z.; Moscatelli, R.; Ligabue, O.; Antolotti, N.; Sebastiani, M.; Lusvarghi, L.; Bemporad, E. Damage progression in thermal barrier coating systems during thermal cycling: A nano-mechanical assessment. Mater. Des. 2019, 166, 107615. [Google Scholar] [CrossRef]
  6. Fattahi, M.; Nezafat, N.B.; Ţălu, Ş; Solaymani, S.; Ghoranneviss, M.; Elahi, S.M.; Shafiekhani, A.; Rezaee, S. Topographic characterization of zirconia-based ceramics by atomic force microscopy: A case study on different laser irradiations. J. Alloy Compd. 2020, 831, 154763. [Google Scholar] [CrossRef]
  7. Boissonnet, G.; Chalk, C.; Nicholls, J.; Bonnet, G.; Pedraza, F. Phase stability and thermal insulation of YSZ and erbia-yttria co-doped zirconia EB-PVD thermal barrier coating systems. Surf. Coat. Technol. 2020, 389, 125566. [Google Scholar] [CrossRef]
  8. Padture, N.P.; Gell, M.; Jordan, E.H. Thermal Barrier Coatings for Gas-Turbine Engine Applications. Science 2002, 296, 280–284. [Google Scholar] [CrossRef]
  9. Jonnalagadda, K.P.; Eriksson, R.; Yuan, K.; Li, X.-H.; Ji, X.; Yu, Y.; Peng, R.L. A study of damage evolution in high purity nano TBCs during thermal cycling: A fracture mechanics based modelling approach. J. Eur. Ceram. Soc. 2017, 37, 2889–2899. [Google Scholar] [CrossRef]
  10. Yang, J.; Wang, L.; Li, D.; Zhong, X.; Zhao, H.; Tao, S. Stress Analysis and Failure Mechanisms of Plasma-Sprayed Thermal Barrier Coatings. J. Therm. Spray Technol. 2017, 26, 890–901. [Google Scholar] [CrossRef]
  11. Jamali, H.; Loghman-Estarki, M.R.; Razavi, R.S.; Mozafarinia, R.; Edris, H.; Bakhshi, S. Comparison of thermal shock behavior of nano-7YSZ, 15YSZ and 5.5 SYSZ thermal barrier coatings produced by APS method. Ceramics 2016, 60, 210–219. [Google Scholar] [CrossRef]
  12. Loghman-Estarki, M.; Razavi, R.S.; Jamali, H. Thermal stability and sintering behavior of plasma sprayed nanostructured 7YSZ, 15YSZ and 5.5SYSZ coatings at elevated temperatures. Ceram. Int. 2016, 42, 14374–14383. [Google Scholar] [CrossRef]
  13. Zeng, J.; Sun, J.; Zhang, H.; Yang, X.; Qiu, F.; Zhou, P.; Niu, W.; Dong, S.; Zhou, X.; Cao, X. Lanthanum magnesium hexaluminate thermal barrier coatings with pre-implanted vertical microcracks: Thermal cycling lifetime and CMAS corrosion behaviour. Ceram. Int. 2018, 44, 11472–11485. [Google Scholar] [CrossRef]
  14. Bal, E.; Karabaş, M.; Taptik, I.Y. The effect of CMAS interaction on thermal cycle lifetime of YSZ based thermal barrier coatings. Mater. Res. Express 2018, 5, 065201. [Google Scholar] [CrossRef]
  15. Cocks, A.; Fleck, N. Constrained sintering of an air-plasma-sprayed thermal barrier coating. Acta Mater. 2010, 58, 4233–4244. [Google Scholar] [CrossRef]
  16. Cocks, A.; Fleck, N.; Lampenscherf, S. A brick model for asperity sintering and creep of APS TBCs. J. Mech. Phys. Solids 2014, 63, 412–431. [Google Scholar] [CrossRef]
  17. Zhang, X.; Li, C.; Withers, P.J.; Markocsan, N.; Xiao, P. Determination of local residual stress in an air plasma spray thermal barrier coating (APS-TBC) by microscale ring coring using a picosecond laser. Scr. Mater. 2019, 167, 126–130. [Google Scholar] [CrossRef]
  18. Li, C.; Zhang, X.; Chen, Y.; Carr, J.; Jacques, S.; Behnsen, J.; di Michiel, M.; Xiao, P.; Cernik, R. Understanding the residual stress distribution through the thickness of atmosphere plasma sprayed (APS) thermal barrier coatings (TBCs) by high energy synchrotron XRD; digital image correlation (DIC) and image based modelling. Acta Mater. 2017, 132, 1–12. [Google Scholar] [CrossRef]
  19. Guan, P.; Ai, Y.; Fei, C.; Yao, Y. Thermal Fatigue Life Prediction of Thermal Barrier Coat on Nozzle Guide Vane via Master–Slave Model. Appl. Sci. 2019, 9, 4357. [Google Scholar] [CrossRef]
  20. Dong, H.; Yang, G.; Li, C.-X.; Luo, X.-T.; Li, C.-J. Effect of TGO Thickness on Thermal Cyclic Lifetime and Failure Mode of Plasma-Sprayed TBCs. J. Am. Ceram. Soc. 2014, 97, 1226–1232. [Google Scholar] [CrossRef]
  21. Finnis, M. The theory of metal - ceramic interfaces. J. Phys. Condens. Matter 1996, 8, 5811–5836. [Google Scholar] [CrossRef]
  22. Guo, X.; Shang, F. Shear strength and sliding behavior of Ni/Al2O3 interfaces: A first-principle study. J. Mater. Res. 2012, 27, 1237–1244. [Google Scholar] [CrossRef]
  23. Guo, X.; Zhang, Y.; Jung, Y.-G.; Li, L.; Knapp, J.; Zhang, J. Ideal tensile strength and shear strength of ZrO2(111)/Ni(111) ceramic-metal Interface: A first principle study. Mater. Des. 2016, 112, 254–262. [Google Scholar] [CrossRef]
  24. Soler, J.M.; Artacho, E.; Gale, J.D.; García, A.; Junquera, J.; Ordejón, P.; Sánchez-Portal, D. The SIESTA method for ab initio order-N materials simulation. J. Phys. Condens. Matter 2002, 14, 2745–2779. [Google Scholar] [CrossRef]
  25. Bao, Z.Y.; Shang, F.L. First-principles investigation of the tensile strength and fracture property of c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interfaces. Mater. Sci. Eng. B. 2022, submitted.
  26. Christensen, A.; Carter, E.A. First-principles characterization of a heteroceramic interface: ZrO2(001) deposited on an α − Al2O3( 1 1 ¯ 02 ) substrate. Phys. Rev. B 2000, 62, 16968–16983. [Google Scholar] [CrossRef]
  27. Eichler, A.; Kresse, G. First-principles calculations for the surface termination of pure and yttria-doped zirconia surfaces. Phys. Rev. B 2004, 69, 045402. [Google Scholar] [CrossRef]
  28. Duan, W.; Karki, B.B.; Wentzcovitch, R.M. High-pressure elasticity of alumina studied by first principles. Am. Miner. 1999, 84, 1961–1966. [Google Scholar] [CrossRef]
  29. Xu, Q.; Salles, N.; Chevalier, J.; Amodeo, J. Atomistic simulation and interatomic potential comparison in α-Al2O3: Lattice, surface and extended-defects properties. Model. Simul. Mater. Sci. Eng. 2022, 30, 035008. [Google Scholar] [CrossRef]
  30. Cousland, G.; Cui, X.; Smith, A.; Stampfl, A.; Stampfl, C. Mechanical properties of zirconia, doped and undoped yttria-stabilized cubic zirconia from first-principles. J. Phys. Chem. Solids 2018, 122, 51–71. [Google Scholar] [CrossRef]
  31. Liu, H.; Inokoshi, M.; Nozaki, K.; Shimizubata, M.; Nakai, H.; Too, T.D.C.; Minakuchi, S. Influence of high-speed sintering protocols on translucency, mechanical properties, microstructure, crystallography, and low-temperature degradation of highly translucent zirconia. Dent. Mater. 2022, 38, 451–468. [Google Scholar] [CrossRef] [PubMed]
  32. Rice, J.R. Dislocation nucleation from a crack tip: An analysis based on the Peierls concept. J. Mech. Phys. Solids 1992, 40, 239–271. [Google Scholar] [CrossRef]
  33. Farkas, D. Atomistic theory and computer simulation of grain boundary structure and diffusion. J. Phys. Condens. Matter 2000, 12, R497–R516. [Google Scholar] [CrossRef]
  34. Urata, S.; Hirobe, S.; Oguni, K.; Li, S. Atomistic to continuum simulations of fracture and damage evolutions in oxide glass and glass-ceramic materials: A critical review. J. Non Cryst. Solids X 2022, 15, 100102. [Google Scholar] [CrossRef]
  35. Wang, R.; Liu, Y.; Mao, J.; Liu, Z.; Hu, D. Cyclic cohesive zone model damage parameter acquisition for fatigue crack growth considering crack closure effect. Int. J. Fatigue 2022, 163, 107021. [Google Scholar] [CrossRef]
  36. Needleman, A. A Continuum Model for Void Nucleation by Inclusion Debonding. J. Appl. Mech. 1987, 54, 525–531. [Google Scholar] [CrossRef]
  37. Xu, X.-P.; Needleman, A. Void nucleation by inclusion debonding in a crystal matrix. Model. Simul. Mater. Sci. Eng. 1993, 1, 111–132. [Google Scholar] [CrossRef]
  38. Umeno, Y.; Kitamura, T. Ab initio simulation on ideal shear strength of silicon. Mater. Sci. Eng. B 2002, 88, 79–84. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interfacial shear simulation. The red, pink, and blue spheres represent O, Al, and Zr atoms, respectively. To better demonstrate the system, the model has been duplicated in the horizontal direction.
Figure 1. Schematic diagram of the c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interfacial shear simulation. The red, pink, and blue spheres represent O, Al, and Zr atoms, respectively. To better demonstrate the system, the model has been duplicated in the horizontal direction.
Applsci 12 08869 g001
Figure 2. Shear stress–displacement curves of the c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interface models: (a) O-< 1 1 ¯ 01 >, (b) O-< 11 2 ¯ 0 >, (c) 2O-< 1 1 ¯ 01 >, (d) 2O-< 11 2 ¯ 0 >, (e) Zr-< 1 1 ¯ 01 >, and (f) Zr-< 11 2 ¯ 0 >. The horizontal lines are the boundaries between region I and II. The shear stresses in region I are positive values, instead those in region II are negative. Specific Zr-O and Zr-Al bonds are also depicted at certain stages.
Figure 2. Shear stress–displacement curves of the c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interface models: (a) O-< 1 1 ¯ 01 >, (b) O-< 11 2 ¯ 0 >, (c) 2O-< 1 1 ¯ 01 >, (d) 2O-< 11 2 ¯ 0 >, (e) Zr-< 1 1 ¯ 01 >, and (f) Zr-< 11 2 ¯ 0 >. The horizontal lines are the boundaries between region I and II. The shear stresses in region I are positive values, instead those in region II are negative. Specific Zr-O and Zr-Al bonds are also depicted at certain stages.
Applsci 12 08869 g002
Figure 3. The energy Φ as a function of the shear displacement δ of the c-ZrO2(001)/α-Al2O3 ( 1 1 ¯ 02 ) interface models.
Figure 3. The energy Φ as a function of the shear displacement δ of the c-ZrO2(001)/α-Al2O3 ( 1 1 ¯ 02 ) interface models.
Applsci 12 08869 g003
Figure 4. The evolution of the atomic structures at the c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interfaces during the shearing process: (a) O-< 1 1 ¯ 01 >, (b) O-< 11 2 ¯ 0 >, (c) 2O-< 1 1 ¯ 01 >, (d) 2O-< 11 2 ¯ 0 >, (e) Zr-< 1 1 ¯ 01 >, and (f) Zr-< 11 2 ¯ 0 >. The red, grey, and green spheres represent O, Al, and Zr atoms, respectively.
Figure 4. The evolution of the atomic structures at the c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interfaces during the shearing process: (a) O-< 1 1 ¯ 01 >, (b) O-< 11 2 ¯ 0 >, (c) 2O-< 1 1 ¯ 01 >, (d) 2O-< 11 2 ¯ 0 >, (e) Zr-< 1 1 ¯ 01 >, and (f) Zr-< 11 2 ¯ 0 >. The red, grey, and green spheres represent O, Al, and Zr atoms, respectively.
Applsci 12 08869 g004
Figure 5. The shear stress–displacement curves of the c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interface models fitted by the Fourier series.
Figure 5. The shear stress–displacement curves of the c-ZrO2(001)/α-Al2O3( 1 1 ¯ 02 ) interface models fitted by the Fourier series.
Applsci 12 08869 g005
Table 1. The basic mechanical properties of the c-ZrO2(001)/α-Al2O3 ( 1 1 ¯ 02 ) interfaces, including the tensile strength (σmax), the work of separation (Wsep), the shear strength (τmax), and the unstable stacking energy (γus).
Table 1. The basic mechanical properties of the c-ZrO2(001)/α-Al2O3 ( 1 1 ¯ 02 ) interfaces, including the tensile strength (σmax), the work of separation (Wsep), the shear strength (τmax), and the unstable stacking energy (γus).
Modelsσmax
(GPa)
Wsep
(mJ/m2)
Shear Directionτmax
(GPa)
γus
(mJ/m2)
O-9.0121208< 1 1 ¯ 01 >5.04399
< 11 2 ¯ 0 >5.69612
2O-7.452829< 1 1 ¯ 01 >3.66349
< 11 2 ¯ 0 >4.52494
Zr-6.893715< 1 1 ¯ 01 >2.94245
< 11 2 ¯ 0 >3.95437
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bao, Z.; Shang, F. First-Principles Investigation of the Shear Properties and Sliding Characteristics of c-ZrO2(001)/α-Al2O3(11¯02) Interfaces. Appl. Sci. 2022, 12, 8869. https://doi.org/10.3390/app12178869

AMA Style

Bao Z, Shang F. First-Principles Investigation of the Shear Properties and Sliding Characteristics of c-ZrO2(001)/α-Al2O3(11¯02) Interfaces. Applied Sciences. 2022; 12(17):8869. https://doi.org/10.3390/app12178869

Chicago/Turabian Style

Bao, Zeying, and Fulin Shang. 2022. "First-Principles Investigation of the Shear Properties and Sliding Characteristics of c-ZrO2(001)/α-Al2O3(11¯02) Interfaces" Applied Sciences 12, no. 17: 8869. https://doi.org/10.3390/app12178869

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