Next Article in Journal
Adsorption Study of Uremic Toxins (Urea, Creatinine, and Uric Acid) Using Modified Clinoptilolite
Previous Article in Journal
Effect of Heat Treatment Temperature on the Microstructure and Properties of Titanium-Clad Steel Plate Prepared by Vacuum Hot Rolling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Regulations of Thermal Expansion Coefficients of Yb1−xAlxTaO4 for Environmental Barrier Coatings Applications

Faculty of Materials Science and Engineering, Kunming University of Science and Technology, Kunming 650093, China
*
Authors to whom correspondence should be addressed.
Coatings 2024, 14(9), 1097; https://doi.org/10.3390/coatings14091097
Submission received: 30 July 2024 / Revised: 22 August 2024 / Accepted: 22 August 2024 / Published: 31 August 2024
(This article belongs to the Special Issue Glass Materials and Coatings: Analysis, Preparation and Application)

Abstract

:
Environmental barrier coatings (EBCs) are widely used to protect ceramic matrix composites (CMCs, SiCf/SiC, and Al2O3f/Al2O3), and they should have low thermal expansion coefficients (TECs) matching the CMCs and excellent mechanical properties to prolong their lifetime. Current EBC materials have disadvantages of phase transitions and insufficient mechanical properties, which affect their working temperatures and lifetime. It is necessary to develop new oxide EBCs. Ytterbium tantalate (YbTaO4) is a stable and novel EBC material, and we have improved the mechanical properties and TECs of Yb1−xAlxTaO4 (x = 0.05, 0.1, 0.2, 0.3, 0.4, 0.5) ceramics by replacing Yb with Al. XRD, SEM, and EDS are used to verify the crystal and microstructures, and nano-indentation is used to measure the modulus and hardness when changes in TECs are measured within a thermal expansion device. The results show that the phase structure of Yb1−xAlxTaO4 (x = 0.05, 0.1, 0.2, 0.3, 0.4, 0.5) is stable at 25–1400 °C within air atmosphere, and their high-temperature TECs (6.4–8.9 × 10−6 K−1, 1400 °C) are effectively regulated by introductions of different contents of Al, which enlarge their engineering applications for SiCf/SiC and Al2O3f/Al2O3 CMCs. The evolutions of TECs are analyzed from structural characteristics and phase compositions, and the increased TECs make Yb1−xAlxTaO4 potential EBCs for Al2O3 matrixes. Due to the high bonding strength of Al–O bonds, hardness, as well as Young’s modulus, are enhanced with the increasing Al content, with Yb1−xAlxTaO4 (x = 0.05, 0.1, 0.2, 0.3, 0.4, 0.5) having a nano-hardness of 3.7–12.8 GPa and a Young’s modulus of 100.9–236.6 GPa. The TECs of YbTaO4 are successfully regulated to expand their applications, and they match those of Al2O3 and SiC matrixes, as well as displaying improved mechanical properties. This work promotes applications of YbTaO4 as potential EBCs and provides a new way to regulate the TECs of tantalates.

1. Introduction

Currently, high-temperature nickel-based alloys are the primary materials used as hot-end components of aviation engines. With the rapid advancement of aircraft components and the increases in the power of engines, the maximum working temperature (1150 °C) of nickel-based alloys cannot meet their requirements. Thermal barrier coatings (TBCs) are applied to the surface of nickel-based components to provide thermal insulation accompanied with cooling technologies, which increase working temperatures [1,2,3]. However, the effects of TBCs are limited, and the traditional zirconate-based TBCs have working temperatures of less than 1200 °C [4]. Additionally, alloys have a high density, which increases the overall weight of engines to reduce fuel efficiency.
Accordingly, lightweight high-temperature ceramic matrix composites (CMCs) with operating temperatures of 1000–1650 °C are noticed for their outstanding performances [5,6,7,8,9]. Using CMCs can reduce the weight of these components and significantly enhance the thrust-to-weight ratio [10]. Common CMCs include Al2O3, C, and SiC fiber-reinforced Al2O3, as well as C fiber-toughened SiC (Cf/SiC) and SiC fiber-reinforced SiC (SiCf/SiC) [11,12,13]. CMCs are susceptible to corrosion by molten salts, oxygen, and steam at high temperatures, which significantly reduce their lifespan [14,15,16,17,18]. Based on the above situations, environmental barrier coatings (EBCs) are mostly applied to shield CMCs from damages caused by high-temperature steam and oxygen, thereby extending their service life [19,20,21]. Oxide EBCs should have low thermal expansion coefficients (TECs, 3~9 × 10−6 K−1) to match those of CMCs, and the interface thermal stress should be reduced as much as possible. The main functions of EBCs are isolating high-temperature steam and oxides to avoid the premature failure of CMCs; thus, EBCs must be stable at high temperatures. It can be seen that excellent high-temperature stability, adjustable TECs, and outstanding mechanical properties are essential for novel EBC materials.
The first-generation EBCs are mullite, which has low TECs and outstanding chemical compatibility with SiC. However, mullite will produce cracks during thermal cycling [22]. By combining mullite with yttria-stabilized zirconia (YSZ), researchers have improved the properties of mullite. YSZ has high TECs, which lead to thermal stress, and the resultant cracks allow water vapor to penetrate through EBCs to corrode CMCs [23,24]. Barium strontium aluminosilicate (BSAS) shows a substantial improvement in performance, but its high-temperature volatility and poor chemical compatibility with SiO2 limit its applications [25]. Rare earth silicates (Rare earth = RE, RE2SiO5, and RE2Si2O7) with low TECs and excellent SiO2 chemical compatibility have replaced previous EBCs [26,27]. RE2SiO5 can withstand high temperatures for extended periods; however, they are unstable at temperatures beyond 1200 °C, and the phase transformations between polymorphs can produce strain and cracks which cause failure. It can be seen that excellent high-temperature stability is essential for EBCs. Additionally, the TECs of CMCs span from about 3 × 10−6 K−1 to 9 × 10−6 K−1 [17,28,29,30,31,32], indicating that low TECs are necessary for EBCs.
Chen et al. [33,34] have reported that YbTaO4 and AlTaO4 are stable without phase changes at 25~1500 °C, and they show excellent chemical compatibility with SiO2. At 1200 °C, the TECs of YbTaO4 and AlTaO4 are 7.6 × 10−6 K−1 and 5.6 × 10−6 K−1, respectively. The above characteristics suggest that YbTaO4 and AlTaO4 are novel oxide EBCs. YbTaO4 has weaknesses of a relatively low modulus and low hardness, which are not beneficial for its EBC applications. It is necessary to optimize the properties of YbTaO4 EBCs, including TECs and mechanical properties. To expand the engineering applications of YbTaO4, we adjust its TECs by Al-substituted Yb to form Yb1−xAlxTaO4 (x = 0.05, 0.1, 0.2, 0.3, 0.4, 0.5) oxides. AlTaO4 has a higher hardness and modulus than YbTaO4, and Al-substituting Yb can enhance the mechanical properties of YbTaO4 [35,36]. The results indicate that Yb1-xAlxTaO4 ceramics have TECs matching SiCf/SiC and Al2O3f/Al2O3 CMCs, and their TECs are successfully regulated by Al substitution. Furthermore, the hardness and modulus of YbTaO4 are simultaneously enhanced to improve their lifetime. In a future study, it is necessary to research the CaO-MgO-AlO1.5-SiO2 corrosion and oxidation resistance properties of YbTaO4. Overall, Yb1−xAlxTaO4 ceramics have significant potential as EBCs according to their adjustable TECs and improved mechanical properties.

2. Experiments

2.1. Sample Synthesis

The synthesis of Yb1−xAlxTaO4 samples (x = 0.05, 0.1, 0.2, 0.3, 0.4, 0.5) used Yb2O3 (purity ≥ 99.99%), Al2O3 (purity ≥ 99.99%), and Ta2O5 (purity ≥ 99.99%) powders from Aladdin, China, and their particle sizes were less than 10 μm. The raw powders were maintained at 1000 °C for 3 h to remove any moisture and then were mixed based on their chemical formula. Anhydrous ethanol was mixed with the weighted powders, and the mixtures were ball-milled for 12 h at 300 rpm. The mixed powders were dried at 100 °C for 12 h. The powders were subsequently pressed under a pressure of 6 MPa for 2 min to obtain ceramic disks. Finally, the disks were sintered at 1500~1700 °C in atmospheric conditions for 3–6 h to produce dense Yb1−xAlxTaO4 ceramics.

2.2. Structural Analysis

An X-ray diffraction (XRD) device (MiniFlex600, Rigaku, Tokyo, Japan) was used to characterize phases of Yb1−xAlxTaO4. The microstructure, as well as elemental distribution and content, were analyzed using energy-dispersive X-ray energy spectroscopy (EDS, Quantax 200, Bruker, Karlsruhe, Germany) and scanning electron microscopy (SEM, SIGMA-300, ZEISS, Oberkochen, Germany). The actual density (ρ) was obtained by the Archimedes principle:
ρ = m 1 m 3     m 2
where m1 was the initial weight, m2 was the weight placed in water, and m3 was the saturated wet weight of the sample.

2.3. Properties Measurements

The TECs (αp) were measured within a thermal expansion device (DIL 402, NETZSCH, Bavaria, Germany) at 30–1400 °C based on the thermal expansion rate (dL/L0):
α p = dL / L 0 Δ T
where dL, L0, and ΔT were elongations in length, original length, and change in temperature, respectively.
The Young’s modulus and hardness of Yb1−xAlxTaO4 were measured using a nano-indentation device (Nanomechanisc, Inc. iMicro, Milpitas, CA, USA) according to the nano-indentation depth (h ≤ 0.2 μm) specified in the national standard GB/T21838.1-2008 [37]. The load was 50 mN, and the holding time was 10 s because the nano Young’s modulus and hardness were stable when the load was higher than 20 mN in our previous study. A holding time of 10 s was used to stabilize the contact area between the sample and the nano-indenter.

3. Results and Discussion

3.1. Crystal and Microstructure

Figure 1a shows experimental XRD patterns of Yb1−xAlxTaO4 ceramics compared with the standard ICDD cards. When x is 0.05–0.1, Yb1−xAlxTaO4 is crystallized into a monoclinic-prime (m′) phase, and its corresponding standard card is PDF#24-1416. When x = 0.2–0.5, XRD patterns correspond to the standard card of PDF#24-1413 and 24-1416, indicating that Yb1−xAlxTaO4 (x = 0.2–0.5) is a mixture of m′ and monoclinic (m) phases. Figure 1b shows that the XRD peaks of Yb1−xAlxTaO4 are obviously shifted to a higher angle with an increment in Al content. The Al3+ ion has a shorter ionic radius than Yb3+, and Al3+ substitutes Yb3+ will contract lattice volumes [37]. The cell parameters and theoretical densities of Yb1−xAlxTaO4 are obtained from XRD Rietveld refinement and are shown in Figure 2a–f and Table 1. For most samples, the weighting factor (Rwp) is less than 15% and the fitting factor (x2) is less than 3, which indicates a high credibility of lattice parameters [38].
Figure 3 shows the crystal structures of Yb1−xAlxTaO4 with different space groups. The unit cell of the m′ phase has twisted [YbO8] dodecahedra and [TaO6] octahedra. Yb and Ta are layered in different planes, and the structure relies on the connection between [YbO8] dodecahedra when two kinds of polyhedral are periodically continued along the a-axis. The primary distinctions between the m and m′ phases are that the oxygen coordination number of Ta atoms in the m phase is 4 and 6 for the m′ phase. According to Hazen and Prewitt’s theory, the thermal expansion of metallic ions in polyhedron is affected by their oxygen coordination number [39,40,41,42]. The higher the coordination number, the lower the TECs. Thus, m-Yb1−xAlxO4 may have higher TECs than those with the m′ phase, and it will be discussed in Section 3.3.
Figure 4a–f shows that the grain sizes of each sample are less than 15 µm. The elemental distribution is analyzed by EDS mapping, and the elemental distribution of Yb1−xAlxTaO4 is uniform when x = 0.05 and 0.1, while Yb1−xAlxTaO4 ceramics are two-phase composites when x is greater than 0.1. Figure 4 shows SEM in the backscatter mode, while the differences in color are not obvious because they consist of the same elements. The points marked as 1~8 in Figure 4c–f are measured using an EDS point scan to estimate their elemental content, as shown in Table 2. The Yb content in point 1 is less than that in point 2, and a similar situation is detected in the other three samples. It is believed that segregations of Yb and Al exist in Yb1−xAlxTaO4 solid solutions; subsequently, this leads to co-formations of m and m′ phases. Previous studies prove that lanthanide contraction leads to the m-m′ phase transition in RETaO4 ceramics [28,34,39], and Al3+ substituting Yb3+ reduces the A-site ionic radius in ABO4-type YbTaO4. The formations of m + m′ Yb1−xAlxTaO4 composites indicate that the solid solubility of Al in YbTaO4 is less than 0.2. The EDS point scan is semi-quantitative, and it is unable to obtain the accurate chemical forms of m- and m′-phase Yb1−xAlxTaO4 ceramics.

3.2. Mechanical Properties

Young’s modulus is directly related to the bonding strength of crystal [43]. In a practical working environment, turbine engines are subject to impact friction from objects entrained in high-speed airflow. Therefore, a high modulus and high hardness are needed to improve the service life of YbTaO4. Young’s modulus (100.9–236.6 GPa) is shown in Figure 5a for Yb1−xAlxTaO4, which is lower than that of YSZ (250 GPa) but similar to those of La2Zr2O7 (237 GPa) and Yb2SiO5 (160 GPa) [44,45,46]. The highest modulus is 236.6 GPa when x = 0.5, and the lowest modulus is 100.9 GPa when x = 0.05. The substitution between Al and Yb increases the bonding strength, which enhances the modulus. Figure 5b shows the hardness (3.7–12.8 GPa) for Yb1−xAlxTaO4. The highest hardness (12.8 GPa) of Yb0.5Al0.5TaO4 is higher than other EBC materials, and this phenomenon can be explained by the enhanced bonding strength. When Al substitutes Yb, the Al-O bond is stronger than the Yb-O bond, which increases the bonding strength and hardness. Both RETaO4 and AlTaO4 are monoclinic phases, and AlTaO4 has a far higher bonding strength and modulus than most RETaO4, except ScTaO4. Thus, using Al to substitute Yb can enhance the hardness and modulus of YbTaO4.

3.3. Thermal Expansion Performance

Figure 6a shows the thermal expansion rates of Yb1−xAlxTaO4, and their slopes are almost constant with the increasing temperature, which can prove their excellent phase stability. Figure 6b shows that the TECs of Yb1−xAlxTaO4 increase steadily with the increasing temperature, and they are approximately 6.0~9.0 × 10−6 K−1 at 1400 °C. The TECs of single-phase Yb1−xAlxTaO4 are lower than that of RE2SiO5, while the TECs of bi-phase Yb1−xAlxTaO4 are similar to or slightly greater than that of RE2SiO5 [46]. The TECs of Yb1−xAlxTaO4 ceramics are similar to those of CMCs, including Al2O3f/Al2O3 (8–9 × 10−6 K−1) and SiCf/SiC (4–6 × 10−6 K−1). Indeed, the TECs of Yb0.5Al0.5TaO4 match that of Al2O3f/Al2O3, and the TECs of Yb0.95Al0.05TaO4 match that of SiCf/SiC. The mathematical expression for TECs is [47]
TECs T = TECs 0 T + TECs 1 ( T )
TECs 0 T = 3 k B 2 φ r 0 U T T D 3 g 0 x D
TECs 1 T = 3 k B 2 φ r 0 U k B T 2 U T T D 3 g 1 x D
g 0 x D = 0 T D T x 4 e x e x 1 2 d x
g 1 x D = 0 T D T x 5 e x ( 1 + e x ) e x   1 3 d x
where φ is the reciprocal of width for the energy barrier, kB is the Boltzmann’s constant, r0 is the equilibrium inter-ion distance, TD is the Debye temperature, and U is the crystal-binding energy. The ionic radius of Al3+ is 0.535 Å, and the ionic radius of Yb3+ is 0.863 Å. The equilibrium inter-ionic distance decreases with the increasing Al content, and the crystal lattice is more tightly packed. The r0 is inversely proportional to the TECs; subsequently, the TECs are increased with the increasing Al content. The crystal-binding energy (U) is [48]
U = N A A z + z e 2 r 0 ( 1     1 B )
where NA is the Avogadro’s constant, A is the Madelung’s constant, z is the ionic charge, and B is the Born’s index. A decrease in the equilibrium distance causes an increase in the crystal-binding energy which reduces the TECs, which is contrary to the experimental results. The TECs are not determined by a single index since Yb1−xAlxTaO4 is a composite when x is higher than 0.2. Yang et al. [49] have found that the TECs of m-phase RETaO4 are higher than that of m′-phase RETaO4 because m′-phase RETaO4 has a higher crystal-binding energy. For m-phase RETaO4, the cations and anions are located at two different positions, whereas the cation and anion are located at three different positions in m′-phase RETaO4. The structural differences affect their thermodynamic stability and force constants. The force constant is directly proportional to the bonding strength indicated by the Debye temperature (TD) [50,51]:
T D = h k B 3 e 4 π V 1 3 V M
where e is the total atomic weight per cell and h is the Planck’s constant. The Debye temperature is positively correlated with the interatomic bonding force. The strength of the Al-O bond is stronger than that of the Yb-O bond, leading to the Debye temperature being increased. Thus, the TECs of single-phase m′-Yb1−xAlxTaO4 are reduced with an increasing Al content when x is 0.05~0.2. For x = 0.3~0.5, the TECs of Yb1−xAlxTaO4 are dramatically increased due to the production of m-Yb1−xAlxTaO4. Furthermore, thermal expansion is the non-harmonic oscillations of lattices reflected by the Grüneisen parameter (γ):
γ = α p · V · K T C V
where V is the molar volume, KT is the isothermal bulk modulus, and CV is the constant volume heat capacity. Without considering CV and KT, the TECs of the m phase are higher than those of the m′ phase. Therefore, the Grüneisen parameter of the m phase is larger than that of the m′ phase, and a similar situation is reported in Sm1−xYbxTaO4 ceramics [39]. There is a theory suggesting that the TECs of metal–oxygen bonds (TECsM) are related to their oxygen coordination number (P) and cation charge (Z) [52,53,54]:
TECs M = 32.9 · ( 0.75   Z P ) × 10 6
Equation (11) expresses that increasing the oxygen coordination number can enhance the TECs when the cation charge is unchanged, indicating that the TECs of Ta-O bonds in [TaO4] of m-Yb1−xAlxTaO4 should be lower than that of Ta-O bonds in [TaO6] of the m′ phase. However, a current study of Yb1−xSmxTaO4 shows that the m phase has higher TECs than the m′ phase attributed to the higher an-harmonic vibrations of lattices. It is believed that crystal structures are the dominators of the TECs for tantalates.

4. Conclusions

In this study, Yb1−xAlxTaO4 (x = 0.05, 0.1, 0.2, 0.3, 0.4, 0.5) ceramics are successfully prepared to regulate their properties. The substitution between Al and Yb improves the hardness and modulus because Al-O bonds are stronger than Yb-O bonds. The nano hardness of Yb1−xAlxTaO4 (x = 0.2, 0.3, 0.4, 0.5) (9.6–12.8 GPa) and the Young’s modulus (146.8–236.6 GPa) are increased effectively compared with YbTaO4 and other EBC materials. The TECs of Yb1−xAlxTaO4 are 6.4–8.9 × 10−6 K−1 at 1400 °C, which are similar to those of the Al2O3 and SiC matrixes (3–9 × 10−6 K−1). For single-phase m′-Yb1−xAlxTaO4 (x = 0.05, 0.1), the addition of Al can slightly reduce the TECs because Al-O bonds are stronger than Yb-O bonds, which leads to a decrement in the an-harmonic vibrations of lattices. For bi-phase Yb1−xAlxTaO4 (x = 0.2–0.5), their TECs increase with the increasing Al content due to the formation of the m phase. Crystal structures are the dominators of the TECs for Yb1−xAlxTaO4 tantalates, which govern their lattice an-harmonic vibrations. Thus, the TECs of YbTaO4 can be regulated by Al substitution. Accompanied with the enhanced mechanical properties, the studied Yb1−xAlxTaO4 ceramics are promising EBCs.

Author Contributions

Formal analysis, J.L., L.C., L.Z., X.C. and J.F.; Investigation, J.L., L.C., L.Z., X.C., C.X. and T.L.; Writing—original draft, J.L.; Writing—review and editing, L.C. and J.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Yunnan Major Scientific and Technological Projects (Grant No. 202302AG050010), General Project in Yunnan Province (Grant No. 202201AT070192 and 202101BE070001-011), and Open Project of Yunnan Precious Metals Laboratory (No. 2023050240).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data underlying the results presented in this paper are not publicly available at this time due to privacy concerns but may be obtained from the authors upon reasonable request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Vaßen, R.; Jarligo, M.O.; Steinke, T.; Mack, D.E.; Stöver, D. Overview on Advanced Thermal Barrier Coatings. Surf. Coat. Technol. 2010, 205, 938–942. [Google Scholar] [CrossRef]
  2. Zhao, P. Mechanical Properties, Thermophysical Properties and Electronic Structure of Yb3+ or Ce4+-Doped La2Zr2O7-Based TBCs. J. Rare Earths 2023, 41, 588–598. [Google Scholar] [CrossRef]
  3. Wu, Y.; Guo, X.; He, D. Research Progress of CMAS Corrosion and Protection Method for Thermal Barrier Coatings in Aero-engines. China Surf. Eng. 2023, 36, 1–13. [Google Scholar]
  4. Wang, Y.; Ouyang, J.; Wei, T.; Cao, G.; Liu, Z.; Ding, Z.; Wang, Y.; Wang, Y. Mechanical Properties, Thermal Conductivity and Defect Formation Energies of Samarium Immobilization in Gd2Zr2O7: First-Principles Study and Irradiation Experiment. J. Rare Earths 2023, 41, 422–433. [Google Scholar] [CrossRef]
  5. Ma, Z.; Liu, L.; Zheng, W. Environmental Barrier Coating for Aeroengines: Materials and Properties. Adv. Ceram. 2019, 40, 331–344. [Google Scholar]
  6. Chen, W.; Duan, H.; Shi, J.-L.; Qian, Y.; Wan, J.; Zhang, X.; Sheng, H.; Guan, B.; Wen, R.; Yin, Y.; et al. Bridging Interparticle Li+ Conduction in a Soft Ceramic Oxide Electrolyte. J. Am. Chem. Soc. 2021, 143, 5717–5726. [Google Scholar] [CrossRef]
  7. Liu, L.; Yang, X.; Xiao, G.; Wei, D.; Du, Y. Thermal Performance Affected by the Mesoscopic Characteristics of the Ceramic Matrix Composite for Hypersonic Vehicle. J. Phys. Conf. Ser. 2021, 1786, 012004. [Google Scholar]
  8. Wang, X.; Gao, X.; Zhang, Z.; Cheng, L.; Ma, H.; Yang, W. Advances in Modifications and High-Temperature Applications of Silicon Carbide Ceramic Matrix Composites in Aerospace: A Focused Review. J. Eur. Ceram. Soc. 2021, 41, 4671–4688. [Google Scholar] [CrossRef]
  9. Yan, Z.; Peng, H.; Ji, X.; Zhang, D. Preparation of Yb2Si2O7 Precursor and Its Agglomerated Powder for Low Pressure Plasma Spraying. Therm. Spray Technol. 2022, 14, 38–45+22. [Google Scholar]
  10. Kimmel, J.; Miriyala, N.; Price, J.; More, K.; Tortorelli, P.; Eaton, H.; Linsey, G.; Sun, E. Evaluation of CFCC Liners with EBC after Field Testing in a Gas Turbine. J. Eur. Ceram. Soc. 2002, 22, 2769–2775. [Google Scholar] [CrossRef]
  11. Liu, Y.; Yu, H.; Meng, H.; Chu, Y. Atomic-Level Insights into the Initial Oxidation Mechanism of High-Entropy Diborides by First-Principles Calculations. J. Mater. 2024, 10, 423–430. [Google Scholar] [CrossRef]
  12. Wen, Z.; Tang, Z.; Liu, Y.; Zhuang, L.; Yu, H.; Chu, Y. Ultrastrong and High Thermal Insulating Porous High-Entropy Ceramics up to 2000 °C. Adv. Mater. 2024, 36, 2311870. [Google Scholar] [CrossRef]
  13. Du, J.; Liu, R.; Wan, F.; Li, D.; Li, J.; Wang, Y. Failure Mechanism of Ytterbium Silicate/Silicon Bi-Layer Environmental Barrier Coatings on SiCf/SiC Composites upon Long-Time Water Vapor and Oxygen Corrosion Test. Surf. Coat. Technol. 2022, 447, 128871. [Google Scholar] [CrossRef]
  14. Yang, F.; Fan, Y.; Zhao, J.; Liu, Y.; Chu, K.; Li, Y.; Li, W.; Liu, B. Mechanical and Thermal Properties for Corrosion Products of Lutetium Silicates against CMAS. J. Am. Ceram. Soc. 2024, 107, 5285–5297. [Google Scholar] [CrossRef]
  15. Fan, Y.; Bai, Y.; Li, Q.; Lu, Z.; Chen, D.; Liu, Y.; Li, W.; Liu, B. Compositional Design and Phase Formation Capability of High-Entropy Rare-Earth Disilicates from Machine Learning and Decision Fusion. npj Comput. Mater. 2024, 10, 95. [Google Scholar] [CrossRef]
  16. Richards, B.T.; Sehr, S.; De Franqueville, F.; Begley, M.R.; Wadley, H.N.G. Fracture Mechanisms of Ytterbium Monosilicate Environmental Barrier Coatings during Cyclic Thermal Exposure. Acta Mater. 2016, 103, 448–460. [Google Scholar] [CrossRef]
  17. Yuan, K.; Hou, W.; Ji, X.; Lu, X. Study on Microstructure Evolution in Ytterbium Silicate Environmental Barrier Coatings under Thermal Shock Tests. Therm. Spray Technol. 2023, 15, 13–18. [Google Scholar]
  18. Tian, J.; Chen, L.; Chen, X.; Luo, K.; Li, B.; Zhang, D.; Wang, M.; Xu, B.; Ren, Z.; Yan, S.; et al. Regulation of Crystal and Microstructures of RETaO4 (RE = Nd, Sm, Gd, Ho, Er) Powders Synthesized via Co-Precipitation. J. Rare Earths 2024, in press. [Google Scholar] [CrossRef]
  19. Deijkers, J.A.; Wadley, H.N.G. A Duplex Bond Coat Approach to Environmental Barrier Coating Systems. Acta Mater. 2021, 217, 117167. [Google Scholar] [CrossRef]
  20. Chen, L.; Li, B.; Guo, J.; Zhu, Y.; Feng, J. High-Entropy Perovskite RETa3O9 Ceramics for High-Temperature Environmental/Thermal Barrier Coatings. J. Adv. Ceram. 2022, 11, 556–569. [Google Scholar] [CrossRef]
  21. Turcer, L.R.; Sengupta, A.; Padture, N.P. Low Thermal Conductivity in High-Entropy Rare-Earth Pyrosilicate Solid-Solutions for Thermal Environmental Barrier Coatings. Scr. Mater. 2021, 191, 40–45. [Google Scholar] [CrossRef]
  22. Federer, J.I. Alumina Base Coatings for Protection of SiC Ceramics. J. Mater. Eng. 1990, 12, 141–149. [Google Scholar] [CrossRef]
  23. Lee, K.N.; Miller, R.A. Development and Environmental Durability of Mullite and Mullite/YSZ Dual Layer Coatings for SiC and Si3N4 Ceramics. Surf. Coat. Technol. 1996, 86–87, 142–148. [Google Scholar] [CrossRef]
  24. Jiang, C.; Feng, M.; Yu, C.; Bao, Z.; Zhu, S.; Wang, F. Thermal Cycling Behavior of Nanostructured and Conventional Yttria-Stabilized Zirconia Thermal Barrier Coatings via Air Plasma Spray. Rare Met. 2023, 42, 3859–3869. [Google Scholar] [CrossRef]
  25. Cojocaru, C.V.; Kruger, S.E.; Moreau, C.; Lima, R.S. Elastic Modulus Evolution and Behavior of Si/Mullite/BSAS-Based Environmental Barrier Coatings Exposed to High Temperature in Water Vapor Environment. J. Therm. Spray Technol. 2011, 20, 92–99. [Google Scholar] [CrossRef]
  26. Chen, P.; Xiao, P.; Li, Z.; Wang, Y.; Tang, X.; Li, Y. Water Vapor Corrosion Behavior and Failure Mechanism of Air Sprayed Bi-Layer Yb2Si2O7/SiC and Tri-Layer Yb2Si2O7/(SiCw-Mullite)/SiC Environmental Barrier Coating. Adv. Powder Mater. 2023, 2, 100064. [Google Scholar] [CrossRef]
  27. Dong, Y.; Ren, K.; Wang, Q.; Shao, G.; Wang, Y. Interaction of Multicomponent Disilicate (Yb0.2Y0.2Lu0.2Sc0.2Gd0.2)2Si2O7 with Molten Calcia-Magnesia-Aluminosilicate. J. Adv. Ceram. 2022, 11, 66–74. [Google Scholar] [CrossRef]
  28. Chen, L.; Wang, J.; Li, B.; Luo, K.; Feng, J. Simultaneous Manipulations of Thermal Expansion and Conductivity in Symbiotic ScTaO4/SmTaO4 Composites via Multiscale Effects. J. Adv. Ceram. 2023, 12, 1625–1640. [Google Scholar] [CrossRef]
  29. Meng, H.; Yu, R.; Tang, Z.; Wen, Z.; Yu, H.; Chu, Y. Formation Ability Descriptors for High-Entropy Diborides Established through High-Throughput Experiments and Machine Learning. Acta Mater. 2023, 256, 119132. [Google Scholar] [CrossRef]
  30. Gatzen, C.; Mack, D.E.; Guillon, O.; Vaßen, R. YAlO3—A Novel Environmental Barrier Coating for Al2O3/Al2O3–Ceramic Matrix Composites. Coatings 2019, 9, 609. [Google Scholar] [CrossRef]
  31. Chen, Z.; Lin, C.; Zheng, W.; Song, X.; Jiang, C.; Niu, Y.; Zeng, Y. Influence of Average Radii of RE3+ Ions on Phase Structures and Thermal Expansion Coefficients of High-Entropy Pyrosilicates. J. Adv. Ceram. 2023, 12, 1090–1104. [Google Scholar] [CrossRef]
  32. Song, C.; Ye, F.; Cheng, L.; Liu, Y.; Zhang, Q. Long-Term Ceramic Matrix Composite for Aeroengine. J. Adv. Ceram. 2022, 11, 1343–1374. [Google Scholar] [CrossRef]
  33. Chen, L.; Li, B.; Feng, J. Rare-Earth Tantalates for Next-Generation Thermal Barrier Coatings. Prog. Mater Sci. 2024, 144, 101265. [Google Scholar] [CrossRef]
  34. Chen, L.; Hu, M.; Guo, J.; Chong, X.; Feng, J. Mechanical and Thermal Properties of RETaO4 (RE = Yb, Lu, Sc) Ceramics with Monoclinic-Prime Phase. J. Mater. Sci. Technol. 2020, 52, 20–28. [Google Scholar] [CrossRef]
  35. Zhang, X.; Li, M.; Zhang, A.; Guo, S.; Mao, J.; Deng, C.; Wang, P.; Deng, C.; Feng, J.; Liu, M.; et al. Al-Modification for PS-PVD 7YSZ TBCs to Improve Particle Erosion and Thermal Cycle Performances. J. Adv. Ceram. 2022, 11, 1093–1103. [Google Scholar] [CrossRef]
  36. Zhang, J.; Tan, X.; Fan, X.; Mao, J.; Deng, C.-M.; Liu, M.; Zhou, K.; Zhang, X. Thermal Insulation Performance of 7YSZ TBCs Adjusted via Al Modification. Rare Met. 2023, 42, 994–1004. [Google Scholar] [CrossRef]
  37. Shi, X.; Liu, W.; Li, M.; Sun, Q.; Xu, S.; Du, D.; Zou, J.; Chen, Z. A Solvothermal Synthetic Environmental Design for High-Performance SnSe-Based Thermoelectric Materials. Adv. Energy Mater. 2022, 12, 2200670. [Google Scholar] [CrossRef]
  38. McCusker, L.B.; Von Dreele, R.B.; Cox, D.E.; Louër, D.; Scardi, P. Rietveld Refinement Guidelines. J. Appl. Crystallogr. 1999, 32, 36–50. [Google Scholar] [CrossRef]
  39. Chen, L.; Hu, M.; Feng, J. Defect-Dominated Phonon Scattering Processes and Thermal Transports of Ferroelastic (Sm1-XYbX)TaO4 Solid Solutions. Mater. Today Phys. 2023, 35, 101118. [Google Scholar] [CrossRef]
  40. Dolan, M.D.; Harlan, B.; White, J.S.; Hall, M.; Misture, S.T.; Bancheri, S.C.; Bewlay, B. Structures and anisotropic thermal expansion of the α, β, γ, and δ polymorphs of Y2Si2O7. Power Diffr. 2008, 23, 20–25. [Google Scholar] [CrossRef]
  41. Chen, L.; Hu, M.; Zheng, X.; Feng, J. Characteristics of Ferroelastic Domains and Thermal Transport Limits in HfO2 Alloying YTaO4 Ceramics. Acta Mater. 2023, 251, 118870. [Google Scholar] [CrossRef]
  42. Kong, F.; Tao, S.; Qian, B.; Gao, L. Facile Synthesis of MTaO4 (M = Al, Cr and Fe) Metal Oxides and Their Application as Anodes for Lithium-Ion Batteries. Ceram. Int. 2018, 44, 8827–8831. [Google Scholar] [CrossRef]
  43. Nie, G.; Bao, Y.; Wan, D.; Tian, Y. Evaluating High Temperature Elastic Modulus of Ceramic Coatings by Relative Method. J. Adv. Ceram. 2017, 6, 288–303. [Google Scholar] [CrossRef]
  44. Ren, X.; Pan, W. Mechanical Properties of High-Temperature-Degraded Yttria-Stabilized Zirconia. Acta Mater. 2014, 69, 397–406. [Google Scholar] [CrossRef]
  45. Liu, B.; Liu, Y.; Zhu, C.; Xiang, H.; Chen, H.; Sun, L.; Gao, Y.; Zhou, Y. Advances on Strategies for Searching for next Generation Thermal Barrier Coating Materials. J. Mater. Sci. Technol. 2019, 35, 833–851. [Google Scholar] [CrossRef]
  46. Ren, X.; Tian, Z.; Zhang, J.; Wang, J. Equiatomic Quaternary (Y1/4Ho1/4Er1/4Yb1/4)2SiO5 Silicate: A Perspective Multifunctional Thermal and Environmental Barrier Coating Material. Scr. Mater. 2019, 168, 47–50. [Google Scholar] [CrossRef]
  47. Ruffa, A.R. Temperature Dependence of the Elastic Shear Moduli of the Cubic Metals. Phys. Rev. B 1977, 16, 2504–2514. [Google Scholar] [CrossRef]
  48. Quane, D. Crystal Lattice Energy and the Madelung Constant. J. Chem. Educ. 1970, 47, 396. [Google Scholar] [CrossRef]
  49. Yang, K.; Chen, L.; Wu, F.; Zheng, Q.; Li, J.; Song, P.; Wang, Y.; Liu, R.; Feng, J. Thermophysical Properties of Yb(TaxNb1−x)O4 Ceramics with Different Crystal Structures. Ceram. Int. 2020, 46, 28451–28458. [Google Scholar] [CrossRef]
  50. Han, Y.; Yu, R.; Liu, H.; Chu, Y. Synthesis of the Superfine High-Entropy Zirconate Nanopowders by Polymerized Complex Method. J. Adv. Ceram. 2022, 11, 136–144. [Google Scholar] [CrossRef]
  51. Anderson, O.L. A Simplified Method for Calculating the Debye Temperature from Elastic Constants. J. Phys. Chem. Solids 1963, 24, 909–917. [Google Scholar] [CrossRef]
  52. Hazen, R.M.; Prewitt, C.T. Effects of Temperature and Pressure on Interatomic Distances in Oxygen-Based Minerals. Am. Mineral. 1977, 62, 309–315. [Google Scholar]
  53. Hazen, R.M. Effects of Temperature and Pressure on the Crystal Structure of Ferromagnesian Olivine. Am. Mineral. 1977, 62, 286–295. [Google Scholar]
  54. Ridley, M.; Gaskins, J.; Hopkins, P.; Opila, E. Tailoring Thermal Properties of Multi-Component Rare Earth Monosilicates. Acta Mater. 2020, 195, 698–707. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of Yb1−xAlxTaO4 (x = 0.05, 0.1, 0.2, 0.3, 0.4, 0.5) ceramics; (a) XRD patterns compared with standard PDF cards; (b) shifts of XRD peaks.
Figure 1. XRD patterns of Yb1−xAlxTaO4 (x = 0.05, 0.1, 0.2, 0.3, 0.4, 0.5) ceramics; (a) XRD patterns compared with standard PDF cards; (b) shifts of XRD peaks.
Coatings 14 01097 g001
Figure 2. XRD Rietveld refinements of Yb1−xAlxTaO4 (x = 0.05, 0.1, 0.2, 0.3, 0.4, 0.5) ceramics: (a) x = 0.05; (b) x = 0.1; (c) x = 0.2; (d) x = 0.3; (e) x = 0.4; (f) x = 0.5.
Figure 2. XRD Rietveld refinements of Yb1−xAlxTaO4 (x = 0.05, 0.1, 0.2, 0.3, 0.4, 0.5) ceramics: (a) x = 0.05; (b) x = 0.1; (c) x = 0.2; (d) x = 0.3; (e) x = 0.4; (f) x = 0.5.
Coatings 14 01097 g002
Figure 3. Crystal structures of Yb1−xAlxTaO4 ceramics with different phases: (a) m′ phase; (b) m phase.
Figure 3. Crystal structures of Yb1−xAlxTaO4 ceramics with different phases: (a) m′ phase; (b) m phase.
Coatings 14 01097 g003
Figure 4. Microstructures and EDS mapping of Yb1−xAlxTaO4 ceramics: (a) x = 0.05; (b) x = 0.1; (c) x = 0.2; (d) x = 0.3; (e) x = 0.4; (f) x = 0.5.
Figure 4. Microstructures and EDS mapping of Yb1−xAlxTaO4 ceramics: (a) x = 0.05; (b) x = 0.1; (c) x = 0.2; (d) x = 0.3; (e) x = 0.4; (f) x = 0.5.
Coatings 14 01097 g004
Figure 5. Mechanical properties of Yb1−xAlxTaO4 (x = 0.05, 0.1, 0.2, 0.3, 0.4, 0.5) and other compounds [44,45,46]: (a) Young’s modulus; (b) hardness.
Figure 5. Mechanical properties of Yb1−xAlxTaO4 (x = 0.05, 0.1, 0.2, 0.3, 0.4, 0.5) and other compounds [44,45,46]: (a) Young’s modulus; (b) hardness.
Coatings 14 01097 g005
Figure 6. Thermal expansion performance of Yb1−xAlxTaO4 (x = 0.05, 0.1, 0.2, 0.3, 0.4, 0.5) ceramics [46]: (a) thermal expansion rate; (b) TECs as a function of temperature.
Figure 6. Thermal expansion performance of Yb1−xAlxTaO4 (x = 0.05, 0.1, 0.2, 0.3, 0.4, 0.5) ceramics [46]: (a) thermal expansion rate; (b) TECs as a function of temperature.
Coatings 14 01097 g006
Table 1. Lattice parameters (a, b, c/Å), β angle (β/°), phase content (wt%), unit cell volume (V3), density (ρ/g·cm−3), and relative density R/%) of Yb1−xAlxTaO4 ceramics.
Table 1. Lattice parameters (a, b, c/Å), β angle (β/°), phase content (wt%), unit cell volume (V3), density (ρ/g·cm−3), and relative density R/%) of Yb1−xAlxTaO4 ceramics.
SamplesPhaseContentabcβVρρR
0.05m1005.075.265.4496.17144.1929.49697
0.1m1005.075.435.2696.18144.0619.35797
0.2m91.55.075.435.2596.16144.0058.97298
m8.56.895.2710.86133.13287.507
0.3m85.55.075.435.2596.16143.9998.65097
m14.56.815.2510.89132.95285.041
0.4m80.95.075.255.4396.18143.6438.30997
m19.16.865.2310.84133.03284.464
0.5m76.05.075.255.4396.16143.5188.01796
m24.06.835.1910.85132.92281.963
Table 2. The compositions of Yb1−xAlxTaO4 (x = 0.2, 0.3, 0.4, 0.5) ceramics, as marked in Figure 4, obtained from an EDS point scan.
Table 2. The compositions of Yb1−xAlxTaO4 (x = 0.2, 0.3, 0.4, 0.5) ceramics, as marked in Figure 4, obtained from an EDS point scan.
SamplesPositionYb (at%)Al (at%)Ta (at%)O (at%)
Yb0.8Al0.2TaO413.819.7315.2261.25
210.9520.8419.5348.68
Yb0.7Al0.3TaO435.6220.6617.9855.47
49.9315.7516.5657.77
Yb0.6Al0.4TaO453.8620.7117.3958.04
611.4417.8618.6652.04
Yb0.5Al0.5TaO473.6117.3015.6763.41
810.3915.8217.4456.35
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

Liao, J.; Chen, L.; Zhang, L.; Chen, X.; Xu, C.; Li, T.; Feng, J. Regulations of Thermal Expansion Coefficients of Yb1−xAlxTaO4 for Environmental Barrier Coatings Applications. Coatings 2024, 14, 1097. https://doi.org/10.3390/coatings14091097

AMA Style

Liao J, Chen L, Zhang L, Chen X, Xu C, Li T, Feng J. Regulations of Thermal Expansion Coefficients of Yb1−xAlxTaO4 for Environmental Barrier Coatings Applications. Coatings. 2024; 14(9):1097. https://doi.org/10.3390/coatings14091097

Chicago/Turabian Style

Liao, Jiaxin, Lin Chen, Luyang Zhang, Xunlei Chen, Cheng Xu, Tianyu Li, and Jing Feng. 2024. "Regulations of Thermal Expansion Coefficients of Yb1−xAlxTaO4 for Environmental Barrier Coatings Applications" Coatings 14, no. 9: 1097. https://doi.org/10.3390/coatings14091097

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