Next Article in Journal
Synergistic Enhancement of Stain Resistance in Exterior Wall Coatings Using SiO2-TiO2 Composite Overlay
Previous Article in Journal
A Review of the Developments in Capacity-Uprating Conductors for Overhead Transmission Lines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influences of SiO2 Additions on the Structures and Thermal Properties of AlTaO4 Ceramics as EBC Materials

1
Faculty of Materials Science and Engineering, Kunming University of Science and Technology, Kunming 650093, China
2
National-Local Joint Engineering Laboratory for Technology of Advanced Metallic Solidification Forming and Equipment, Kunming 650093, China
*
Authors to whom correspondence should be addressed.
Coatings 2025, 15(10), 1204; https://doi.org/10.3390/coatings15101204 (registering DOI)
Submission received: 4 September 2025 / Revised: 6 October 2025 / Accepted: 10 October 2025 / Published: 13 October 2025
(This article belongs to the Section Ceramic Coatings and Engineering Technology)

Abstract

Ceramic matrix composites (CMCs) are extensively utilized in aero engines due to their high-temperature stability; however, they are prone to environmental corrosion at high temperatures, and environmental barrier coatings (EBCs) are necessary to resist oxidation and corrosion. Among various EBC materials, AlTaO4 offers high cost-effectiveness and low thermal expansion coefficients (TECs), but its resistance to SiO2 erosion and high-temperature stability remain unclear. We investigated the influences of SiO2 additions on the structures and thermal properties of AlTaO4; and AlTaO4 mixtures containing 10 wt.% SiO2 were kept at 1400 °C for 30–120 h. AlTaO4 exhibited excellent high-temperature phase stability, and SiO2 dissolved into AlTaO4 to generate a solid solution. XRD Rietveld refinement was employed to confirm the position of Si in the lattices, while SEM and EDS characterizations demonstrated the homogeneous distribution of Si, Al, and Ta elements. At 1200 °C, the TECs of SiO2-AlTaO4 (4.65 × 10−6 K−1) were close to those of SiC (4.5–5.5 × 10−6 K−1). Additionally, the addition of SiO2 could reduce TECs of AlTaO4, a feature that helped alleviate the interface thermal stress between AlTaO4 and the Si bond coat in the EBC systems. At 900 °C, the thermal conductivity was reduced by 26.9% compared to that of AlTaO4, and the lowest value was 1.65 W·m−1·K−1. Accordingly, SiO2 will enter the lattices of AlTaO4 after heat treatments at 1400 °C, and SiO2 additions will reduce the thermal conductivity and TECs of AlTaO4, which is beneficial for its EBC applications.

1. Introduction

With the continuous developments in aero engine technology, the turbine inlet temperature of advanced aero engines has risen to 1600 °C, which exceeds the melting points of high-temperature nickel-based alloys [1,2,3]. Ceramic matrix composites (CMCs) exhibit excellent high-temperature stability and can withstand long-term exposure to temperatures of up to 1600 °C, leading to their performance being significantly superior to that of nickel-based alloys [4,5,6,7]. The applications of CMCs can effectively enhance engine thermal efficiency and reduce cooling air demand. Using CMCs to manufacture engine turbine components can reduce the weight by 50%–67%, which significantly lowers engine weight and further improves aircraft fuel efficiency [8]. For example, SiC fiber-reinforced SiC (SiCf/SiC) composites have a density of approximately 2 g·cm−3, which is much less than that of high-temperature alloys [3,9,10,11,12,13,14]. Some CMCs can achieve twice the strength of nickel-based alloys and combine high specific strength and specific modulus. Therefore, CMCs can withstand high stress and maintain good mechanical properties in the harsh engine environment.
In the actual operating conditions of gas turbine engines, CMCs operate in air at temperatures exceeding 1200 °C, where they react with environmental dust (primarily calcium-magnesium-alumino-silicate (CMAS) [15,16,17,18], fuel impurities, and high-temperature water vapor, and a SiO2 passivation layer is formed [19]. The dense SiO2 layer is capable of inhibiting the further oxidation of both the Si bond coat and the CMCs. However, high-temperature steam induces further corrosion of SiO2, resulting in the formation of volatile Si(OH)4 and subsequent degradation of CMCs [20]. To address this issue, environmental barrier coatings (EBCs) are applied to the CMC’s surface to isolate high-temperature corrosive media [21,22,23].
EBCs should exhibit low thermal conductivity for protecting the substrate; match the thermal expansion coefficients (TECs) of the CMCs to avoid coating cracking at high temperatures [24,25]; possess a high melting point to ensure operations in high-temperature environments; and exhibit resistance to erosion by CMAS and high-temperature water vapor [26,27]. Current EBC candidates include aluminates [28,29], phosphates [30,31,32], hafnates [33,34], zirconates [35,36,37,38], tantalates [39,40,41], and niobates [42,43,44,45,46]. However, each of these material systems has inherent drawbacks that restrict their applications [47,48,49,50]. Among various materials, AlTaO4 is regarded as a promising EBC material, attributed to its excellent thermal stability, low thermal expansion coefficients (TECs), and superior mechanical properties [51]. According to the selling prices on the official website of Aladdin Reagents as of 3 October 2025, 100 g of Al2O3 costs RMB 142.90, but commonly used rare earth oxides, such as 100 g of Sm2O3, are priced at RMB 2829.90, and 100 g of Sc2O3 is priced at RMB 2758.90. From an economic perspective, AlTaO4 demonstrates significant advantages compared to RETaO4 during the synthesis process [52]. Since the Si bond coat reacts continuously with oxygen in the environment during long-term service in high-temperature environments, a thermally grown oxide (TGO: SiO2) layer is gradually formed. The stability, densification, and bonding state with the top coating of this TGO layer directly affect the protective lifespan of the entire coating system [53,54,55,56,57]. Therefore, it is particularly critical to evaluate the interfacial interaction between AlTaO4 and the SiO2 derived from the Si bond coat. Specifically, two key aspects need to be considered: first, whether the presence of SiO2 will form a second phase with AlTaO4 (which may lead to performance failure), and second, whether the difference in thermal expansion coefficients between SiO2 and AlTaO4 EBCs will affect the structural integrity of AlTaO4 during service. The chemical compatibility between the two ultimately determines the service performance of AlTaO4 EBCs under high temperatures.
Thus, this study focuses on 1400 °C (a typical high-temperature service range) and systematically investigates the chemical reaction behavior between AlTaO4 and SiO2 at this temperature. It further evaluates the effects of SiO2 introduction and prolonged high-temperature environments on the applicability of AlTaO4 as EBCs in high-temperature service scenarios, providing an experimental basis for the subsequent composition design and process optimization of AlTaO4 EBCs.

2. Materials and Methods

2.1. Synthesis of Materials

The pristine powders of Ta2O5 (purity > 99.9%, particle size ≤ 30 μm, Macklin, Shanghai, China) and Al2O3 (purity > 99.9%, particle size ≤ 30 μm, Macklin, China) were mixed in a stoichiometric ratio and homogenized via wet planetary ball milling at a rotational speed of 300 r/min for a duration of 12 h. The homogenized slurry was dried in an oven at approximately 70 °C for 24 h. After complete drying, the resultant powder was sintered in a tube furnace under ambient air atmosphere at 1500 °C for a holding time of 6 h. Subsequently, the sintered powder was removed, ground, and sieved through a 300-mesh screen to obtain the pristine AlTaO4 powder (purity > 99.9%, particle size ≤ 50 μm).
The AlTaO4 powder and SiO2 powder (purity > 99.9%, particle size ≤ 30 μm, Macklin, China) were weighed at a mass ratio of 9:1 and homogenized via wet planetary ball milling at a rotational speed of 300 r/min for a duration of 12 h. The homogenized slurry was dried in an oven at approximately 70 °C for 24 h. Upon complete drying, the resultant powder was uniaxially pressed into disc-shaped green bodies with a height of 2 mm and a radius of 7.5 mm under a pressure of 320 MPa. These green discs were then sintered in a tube furnace under ambient air atmosphere at 1400 °C for holding times of 30, 60, 90, and 120 h, respectively. In the present study, the samples corresponding to the aforementioned sintering times were designated as the 30 h, 60 h, 90 h, and 120 h samples, respectively. The bulk density of the prepared samples was measured using the Archimedes method, and the porosity (φ) was calculated by comparing the measured bulk density with the theoretical density.

2.2. Structural Identification

Phase characterization of the sintered samples was conducted using an X-ray diffractometer (XRD, MiniFlex 600, Rigaku, Tokyo, Japan), and subsequent Rietveld refinement of the obtained XRD patterns was performed via the General Structure Analysis System (GSAS-II, SVN version 5455) software to quantify phase composition. For microstructural and elemental analysis, a scanning electron microscope (SEM, JSM-7800F, JEOL, Tokyo, Japan) equipped with an energy-dispersive spectrometer (EDS, 6751A-6UUS-SN, Thermo Scientific, Waltham, MA, USA) was employed to observe grain size, surface pores, and elemental distribution. The Raman spectrum was measured using a laser confocal Raman microscope (LABRAM Soleil, HORIBA, Palaiseau, France) with an excitation wavelength of 532 nm, over the range of 50–1100 cm−1.

2.3. Measurements of Thermal Properties

The TECs of the samples were measured using a dilatometer (DIL 402, NETZSCH, Selb, Germany) at a heating rate of 5 °C/min. For thermal diffusivity (α) testing, a laser flash apparatus (LFA457, NETZSCH, Germany) was employed, with the measurement conducted at a heating rate of 10 °C/min. The thermal diffusivity was determined using the Radiation Pulse correction model available in the system, with a deviation of less than 2%. Three independent measurements were conducted at each point. Due to the extremely small test error, no explanation was provided in this paper. The thermal conductivity (k) was further calculated via Equation (1), using the experimentally obtained thermal diffusivity, density (ρ), and specific heat (Cp) [58].
k = α ρ C p

3. Results and Discussion

3.1. Crystal Structure

Figure 1a shows the XRD patterns of AlTaO4 and SiO2 mixtures after heat treatment at 1400 °C for various durations. The peaks of each sample matched well with the standard card of AlTaO4 (PDF#01-076-0363), which had a monoclinic phase and space group C2/m.
The ionic radii of Al3+, Si4+, and Ta5+ were 50, 42, and 69 pm [59], respectively. The difference in ionic radii between Al3+ and Si4+, calculated via Equation (2), was 16%, and that between Ta5+ and Si4+ was 39.13%. XRD results indicated that SiO2 had dissolved into the AlTaO4 lattice to form a SiO2-AlTaO4 solid solution.
Δ r =   r 1 r 2 r 1
In Figure 1b, a shift in characteristic peaks was observed. The main peaks after heat treatment for 30–60 h showed a progressive left shift, while those heat-treated for 90–120 h gradually shifted to the right compared with the sample heated for 60 h. According to Bragg’s equation, an increase in interplanar spacing (d) gave rise to a reduction in diffraction angle (2θ), and this resulted in a left shift in the diffraction peaks. After 120 h of heat treatment, the opposite process might have occurred. The above phenomenon was presumably caused by the position of Si in the AlTaO4 lattice, which was influenced by heat treatment time. The sig and Rwp values in Figure 2 indicated high confidence in the XRD refinement results.
Table 1 presents the changes in lattice constants, unit cell volume, theoretical density, and porosity. The refinement results were consistent with those reported in previous work [51].
Specifically, for samples heat-treated for 30–60 h, the unit cell volume was expanded, and the theoretical density was decreased. For the samples heat-treated for 90–120 h, the unit cell volume began to shrink, and the theoretical density increased compared to the samples heat-treated for 30–60 h. The above results indicate that the heat treatment time indeed caused changes in the position of Si atoms in the AlTaO4 lattice. Si occupied interstitial sites in AlTaO4 after heat treatment for 30–60 h, and an interstitial solid solution was formed, resulting in lattice expansion. After 60–90 h heat treatment, the lattice thermal vibrations of AlTaO4 were further intensified to increase atomic spacing and expand interstitial sizes. Si atoms moved from interstitial sites to lattice sites, and a substitutional solid solution was triggered to reduce the unit cell volume. Figure 3 shows the crystal structure of AlTaO4 when Si occupies both the Ta and Al atomic sites. In the lattices of AlTaO4, Al and Ta atoms were bonded to six O atoms, and Si atoms had a probability of 5% to occupy the sites of Al and Ta atoms. The polyhedron distortion (Δd) was calculated using Equation (3) [60,61]:
Δ d   =   1 n n L L L 2
where L represents the bond length of A–O (A = Al, Ta), L represents the average bond length of [AlO6] and [TaO6], and n represents the number of bonds as illustrated in Figure 3.
The results are shown in Figure 4. The total distortion index showed a trend of increasing and then decreasing, which was consistent with the XRD refinement results. The distortion was dominated by the Al-O bonds, and this was consistent with the changes in the position of Si.

3.2. Microstructures and Elemental Distributions

Figure 5 shows the surface morphology of SiO2-AlTaO4 and the corresponding EDS mapping for samples heat-treated for different durations, and it illustrates the changes in grain size and surface morphology. It can be denoted that the grain size of SiO2-AlTaO4 ceramics after heat treatment for different durations ranged from submicron to micron scale. As the heat treatment time increased, the grains gradually grew and arranged more regularly, and the sample heat-treated for 120 h had the largest grains.
EDS mapping revealed the homogeneous distribution of Si, Al, Ta, and O elements within the grains without obvious segregation, and the microstructure of SiO2-AlTaO4 remained largely unaffected by SiO2 addition and heat treatment duration. Figure 6 showed the atomic ratio of each element based on the EDS spectra, which was consistent with the experimental design.

3.3. Raman Spectra

The Raman spectrum of SiO2-AlTaO4 is shown in Figure 6b. Obviously, the characteristic peak positions in the Raman spectra of all samples maintained good consistency, indicating that the core structure of the samples remained unchanged during the 30–120 h heat treatment process, with only slight changes in peak intensity and peak shape. Raman spectrum correction was performed on the 30 h sample, and the results are shown in Figure 6c. The peaks at wavenumbers of 98.38, 134.85, 169.92, 201.20, 244.28, and 284.71 cm−1 corresponded to the characteristic peaks of the Al-Ta-O framework. The peaks at 349.59, 418.47, 445.36, 515.28, and 581.26 cm−1 mostly corresponded to Ta-O bonds or Al-O bonds. The strong peak at 947.84 cm−1 was a characteristic identification peak of AlTaO4, corresponding to the characteristic stretching vibration of [TaO6] octahedrons [62,63]. The appearance of the weak peak at 465.57 cm−1 confirmed the existence of Si-O bonds. The above results indicated that there was no second-phase SiO2 in SiO2-AlTaO4.

3.4. Thermal Properties

Based on the above results, a solid solution phase was formed after heat treatments. Therefore, it was important to evaluate changes in the thermal properties of SiO2-AlTaO4. Figure 7a showed a comparison of thermal expansion rates between SiO2-AlTaO4, AlTaO4, and AlNbO4 [51]. The results showed that the slopes of the thermal expansion rate curves remained stable as the temperature increased, which indicated good high-temperature stability. The thermal expansion rate of SiO2-AlTaO4 was significantly lower than that of AlTaO4 and AlNbO4. Mechanistically, thermal expansion essentially resulted from the increase in average atomic distance due to intensified atomic thermal vibrations. Figure 7b showed their TECs, which had a positive correlation with temperature.
The TECs of SiO2-AlTaO4 increased rapidly initially and gradually stabilized at temperatures beyond 1100 °C and reached approximately 4.65 × 10−6 K−1 at 1200 °C, which were close to the TECs of SiC (4.5–5.5 × 10−6 K−1) [4,64]. The TECs of SiO2-AlTaO4 were significantly lower than those of AlTaO4 and AlNbO4, which could be attributed to the following reasons. (i) After SiO2 and AlTaO4 formed a solid solution, Si ions were introduced into the AlTaO4 lattices, which caused the lattice distortion and micro-local stress. The stress restrained atomic thermal vibrations and made it difficult for atoms to expand freely as the temperature increased, thereby reducing the TECs [65,66]. (ii) The reduced lattice distortion decreased the vibrational entropy to improve lattice order and enhanced the bonding strength among different atoms, which further reduced TECs [67]. (iii) The SiO2-AlTaO4 ceramics had porosities of 30%–40%, which were much higher than the porosities of AlTaO4 and AlNbO4 reported in Ref. [51], and a high porosity would decrease TECs. Obviously, the SiO2 would react with AlTaO4 to form a solid solution, and the TECs were decreased at the same time, which was good for reducing the interface thermal stress between AlTaO4 and the bond coat in the EBC system.
The low thermal conductivity of EBCs can provide high thermal insulation for CMCs, thereby extending their service life. Figure 8a showed the thermal diffusivity of SiO2-AlTaO4, which ranged from 1.39 to 0.60 mm2·s−1 at 25–900 °C. Compared to AlTaO4, SiO2-AlTaO4 had a lower thermal diffusivity below 400 °C, and they had similar thermal diffusivity at 400–900 °C. Also, SiO2-AlTaO4 had a lower thermal diffusivity than AlNbO4, Yb2Si2O7, and other EBC materials [51,68,69]. Figure 8b compared the thermal conductivity of SiO2-AlTaO4 with other oxides [51,68,69,70,71]; and their thermal conductivity showed a decreasing trend with the increasing temperature, which was consistent with the behavior of insulating materials.
As the temperature increased, the atomic thermal vibrations intensified, and phonon scattering was enhanced, resulting in a decrease in thermal conductivity. The thermal conductivity of SiO2-AlTaO4 exhibited a decreasing trend with increasing temperature, ranging from 2.91 to 1.65 W·m−1·K−1 at temperatures between 25 °C and 900 °C. It could be seen that the thermal conductivity of SiO2-AlTaO4 was significantly lower than that of AlTaO4. At 900 °C, AlTaO4 exhibited a thermal conductivity of 2.26 W·m−1·K−1, whereas the thermal conductivity of SiO2-AlTaO4 was only 1.65 W·m−1·K−1; obviously, the introduction of SiO2 led to a 26.9% reduction in the thermal conductivity of AlTaO4. Compared to other oxides, SiO2-AlTaO4 ceramics exhibited a low thermal conductivity, and SiO2 addition enhanced the phonon scattering in AlTaO4.

4. Conclusions

This study investigated the high-temperature reaction behaviors between SiO2 and AlTaO4, and the influences of SiO2 additions on thermal conductivity and TECs were studied. The conclusions were:
(1)
The SiO2-AlTaO4 samples heat-treated at 1400 °C for 30–120 h maintained the monoclinic phase, and Si occupied the interstitial sites to form an interstitial solid solution at 30–60 h, where the Si atom migrated to lattice sites to form a substitutional solid solution at 90–120 h. AlTaO4 exhibited excellent high-temperature stability, and the addition of SiO2 did not alter its crystal structure.
(2)
SEM and EDS mapping revealed the uniform distribution of Si, Al, Ta, and O elements in each sample, without obvious segregation. The grain sizes were on the micron scale, gradually increasing in size and becoming more regularly arranged with extended heat treatment time. The XRD, SEM, and EDS results synergistically indicated the good phase stability of AlTaO4.
(3)
The TECs of SiO2-AlTaO4 were approximately 4.65 × 10−6 K−1 at 1200 °C, which were lower than those of AlTaO4 and AlNbO4. The SiO2 additions could reduce the TECs of AlTaO4, thereby reducing the thermal stress between AlTaO4 and the Si bond coat. The thermal conductivity at 900 °C was 26.9% lower than that of AlTaO4, and the lowest value reached 1.65 W·m−1·K−1 as the phonon scattering strength was increased. Accordingly, the reaction between SiO2 and AlTaO4 could reduce the TECs and thermal conductivity, which were good for the EBC applications.

Author Contributions

Methodology, L.Z., L.C. and J.W.; Validation, B.W.; Investigation, B.W. and Z.G.; Data curation, B.W.; Writing—original draft, B.W.; Writing—review & editing, B.W., L.C. and J.F.; Supervision, L.Z., L.C., J.W. and J.F.; Project administration, L.C. and J.F. All authors have read and agreed to the published version of the manuscript.

Funding

Thanks for the support from the National Natural Science Foundation (Nos. 52562006 and 52502065), National Key Research and Development Program of China (No. 2022YFB3708600), Yunnan Fundamental Research Projects (No. 202201BE070001-008), General Project in Yunnan Province (Nos. 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

The data presented in this study are available on request from thecorresponding author due to legal.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Shin, S.; Wang, Q.Y.; Luo, J.; Chen, R.K. Advanced materials for high-temperature thermal transport. Adv. Funct. Mater. 2020, 30, 1904815. [Google Scholar] [CrossRef]
  2. George, E.P.; Raabe, D.; Ritchie, R.O. High-entropy alloys. Nat. Rev. Mater. 2019, 4, 515–534. [Google Scholar] [CrossRef]
  3. Chen, L.; Li, B.H.; Feng, J. Rare-earth tantalates for next-generation thermal barrier coatings. Prog. Mater Sci. 2024, 144, 101265. [Google Scholar] [CrossRef]
  4. Lee, K.N. Current status of environmental barrier coatings for Si-Based ceramics. Surf. Coat. Technol. 2000, 133–134, 1–7. [Google Scholar] [CrossRef]
  5. Chen, S.Y.; Chen, Z.K.; Wang, J.M.; Zeng, Y.; Song, W.L.; Xiong, X.; Li, X.C.; Li, T.Q.; Wang, Y.C. Insight into the effect of Ti substitutions on the static oxidation behavior of (Hf,Ti)C at 2500 °C. Adv. Powder Mater. 2024, 3, 100168. [Google Scholar] [CrossRef]
  6. Sciti, D.; Vinci, A.; Zoli, L.; Galizia, P.; Failla, S.; Mungiguerra, S.; Di Martino, G.D.; Cecere, A.; Savino, R. Propulsion tests on ultra-high-temperature ceramic matrix composites for reusable rocket nozzles. J. Adv. Ceram. 2023, 12, 1345–1360. [Google Scholar]
  7. Guo, L.X.; Huang, S.W.; Li, W.; Lv, J.S.; Sun, J. Theoretical design and experimental verification of high-entropy carbide ablative resistant coating. Adv. Powder Mater. 2024, 3, 100213. [Google Scholar] [CrossRef]
  8. He, X.; Yuan, X.H.; Xu, H.; Song, P.; Yu, X.; Li, C.; Huang, T.H.; Li, Q.L.; Lü, K.Y.; Feng, J.; et al. Analysis of structure and microhardness of Al2O3-40 wt.% TiO2/NiCoCrAl gradient coating with in-situ needle-like phase reinforcement after high-temperature treatment. Ceram. Int. 2019, 45, 14546–14554. [Google Scholar] [CrossRef]
  9. Madar, R. Silicon carbide in contention. Nature 2004, 430, 974–975. [Google Scholar] [CrossRef] [PubMed]
  10. Kedir, N.; Garcia, E.; Kirk, C.; Gao, J.L.; Guo, Z.R.; Zhai, X.D.; Sun, T.; Fezzaa, K.; Sampath, S.; Chen, W.W. Impact damage of narrow silicon carbide (SiC) ceramics with and without environmental barrier coatings (EBCs) by various foreign object debris (FOD) simulants. Surf. Coat. Technol. 2021, 407, 126779. [Google Scholar] [CrossRef]
  11. Zou, B.L.; Khan, Z.S.; Fan, X.Z.; Huang, W.Z.; Gu, L.J.; Wang, Y.; Xu, J.Y.; Tao, S.Y.; Yang, K.Y.; Ma, H.M.; et al. A new double layer oxidation resistant coating based on Er2SiO5/LaMgAl11O19 deposited on C/SiC composites by atmospheric plasma spraying. Surf. Coat. Technol. 2013, 219, 101–108. [Google Scholar]
  12. Zhang, S.C.; Wu, L.H.; Sun, X.K.; Sun, H.R.; Ai, B.; Chen, Y.F. Ultra-low thermal conductivity multilayer composites for insulation in low pressure environments. Adv. Ceram. 2023, 44, 442–450. [Google Scholar]
  13. Liu, X.F.; He, J.F.; Li, Y.G.; Li, H.; Lei, W.; Jia, Q.L.; Zhang, S.W.; Zhang, H.J. Foam-gelcasting preparation of porous SiC ceramic for high-temperature thermal insulation and infrared stealth. Rare Met. 2023, 42, 3829–3838. [Google Scholar]
  14. Zhang, J.X.; Zhang, J.; Ye, X.L.; Ma, X.M.; Liu, R.; Sun, Q.L.; Zhou, Y.L. Ultralight and compressive SiC nanowires aerogel for high-temperature thermal insulation. Rare Met. 2023, 42, 3354–3363. [Google Scholar]
  15. Wright, A.J.; Mock, C.; Sharobem, T.; Sotiropoulos, N.; Dambra, C.; Keyes, B.; Ghoshal, A. Integrated microstructural and chemical approach for improving CMAS resistance in thermal and environmental barrier coatings. Coatings 2025, 15, 680. [Google Scholar] [CrossRef]
  16. Wu, Y.; Guo, X.; He, D.Y. Research progress of CMAS corrosion and protection method for thermal barrier coatings in aero-engines. China Surf. Eng. 2023, 36, 1–13. [Google Scholar]
  17. Bryce, K.; Majee, B.P.; Huang, L.P.; Lian, J. A systematic study of thermomechanical properties and calcium-magnesium-aluminosilicate (CMAS) corrosion of multicomponent rare-earth phosphates. J. Adv. Ceram. 2024, 13, 1807–1822. [Google Scholar]
  18. Zhang, L.Y.; Luo, K.R.; Chen, L.; Wang, J.K.; Li, B.H.; Feng, J. A4Ta2O9 (A = Ca, Mg) tantalate as novel thermal barrier coatings materials with superior calcium-magnesium-alumino-silicate corrosion resistance. J. Eur. Ceram. Soc. 2025, 45, 117090. [Google Scholar] [CrossRef]
  19. Naraparaju, R.; Schulz, U.; Mechnich, P.; Döbber, P.; Seidel, F. Degradation study of 7wt.% yttria stabilised zirconia (7YSZ) thermal barrier coatings on aero-engine combustion chamber parts due to infiltration by different CaO-MgO-Al2O3-SiO2 variants. Surf. Coat. Technol. 2014, 260, 73–81. [Google Scholar] [CrossRef]
  20. Niu, Z.B.; Li, D.X.; Jia, D.C.; Yang, Z.H.; Lin, K.P.; Riedel, R.; Colombo, P.; Zhou, Y. Oxidation behavior of amorphous and nanocrystalline SiBCN ceramics—Kinetic consideration and microstructure. Adv. Powder Mater. 2024, 3, 100163. [Google Scholar] [CrossRef]
  21. Richards, B.T.; Wadley, H.N.G. Plasma spray deposition of tri-layer environmental barrier coatings. J. Eur. Ceram. Soc. 2014, 34, 3069–3083. [Google Scholar] [CrossRef]
  22. An, K.; Wang, Y.Q.; Sui, Y.; Qing, Y.Q.; Tong, W.; Wang, X.Z.; Liu, C.S. Recent advances of cerium compounds in functional coatings: Principle, strategies and applications. J. Rare Earths 2025, 43, 227–245. [Google Scholar] [CrossRef]
  23. Xie, Y.S.; Peng, Y.; Deng, Z.Z.; Zhu, Z.; Cheng, Y.; Ma, D.H.; Zhu, L.Y.; Zhang, X.H. Synergistic regulation of temperature resistance and thermal insulation performance of zirconia-based ceramic fibers. Rare Met. 2023, 42, 4189–4200. [Google Scholar] [CrossRef]
  24. Yang, D.L.; Liu, J.L.; Zhang, J.G.; Liang, X.H.; Zhang, X.F. In situ high-Temperature tensile fracture mechanism of PS-PVD EBCs. Coatings 2022, 12, 655. [Google Scholar] [CrossRef]
  25. 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]
  26. Cao, G.; Wang, S.Q.; Ding, Z.Y.; Wang, Y.H.; Liu, Z.G.; Ouyang, J.H.; Wang, Y.M.; Wang, Y.J. High temperature corrosion behavior and the inhibition of apatite formation evolving from corrosion of (Y1−xYbx)2SiO5 in water vapor environment. Appl. Surf. Sci. 2022, 601, 154284. [Google Scholar]
  27. Dong, S.J.; Lü, K.Y.; Wang, Y.H.; Zeng, J.Y.; Yuan, J.Y.; Jiang, J.N.; Deng, L.H.; Cao, X.Q. High-temperature corrosion of HfSiO4 environmental barrier coatings exposed to water vapor/oxygen atmosphere and molten calcium magnesium aluminosilicate. Corros. Sci. 2022, 197, 110081. [Google Scholar]
  28. Xiang, H.M.; Feng, Z.H.; Li, Z.P.; Zhou, Y.C. Crystal structure, mechanical and thermal properties of Yb4Al2O9: A combination of experimental and theoretical investigations. J. Eur. Ceram. Soc. 2017, 37, 2491–2499. [Google Scholar] [CrossRef]
  29. Morán-Ruiz, A.; Vidal, K.; Larrañaga, A.; Arriortua, M.I. Characterization of Ln4Al2O9 (Ln = Y, Sm, Eu, Gd, Tb) rare-earth aluminates as novel high-temperature barrier materials. Ceram. Int. 2018, 44, 8761–8767. [Google Scholar] [CrossRef]
  30. Bryce, K.; Shih, Y.-T.; Huang, L.; Lian, J. Calcium-magnesium-aluminosilicate (CMAS) corrosion resistance of high entropy rare-earth phosphate (Lu0.2Yb0.2Er0.2Y0.2Gd0.2)PO4: A novel environmental barrier coating candidate. J. Eur. Ceram. Soc. 2023, 43, 6461–6472. [Google Scholar] [CrossRef]
  31. Zhang, P.X.; Wang, E.H.; Guo, C.Y.; Yang, T.; Hou, X.M. High-entropy rare earth phosphates (REPO4, RE = Ho, Tm, Yb, Lu, Dy, Er and Y) with excellent comprehensive properties. J. Eur. Ceram. Soc. 2024, 44, 1873–1879. [Google Scholar] [CrossRef]
  32. Karaulov, A.G.; Zoz, E.I. Phase formation in the ZrO2-HfO2-Gd2O3 and ZrO2-HfO2-Yb2O3 systems. Refract. Ind. Ceram 1999, 40, 479–483. [Google Scholar] [CrossRef]
  33. Mikuśkiewicz, M.; Migas, D.; Moskal, G. Synthesis and thermal properties of zirconate, hafnate and cerate of samarium. Surf. Coat. Technol. 2018, 354, 66–75. [Google Scholar] [CrossRef]
  34. Lopato, L.M.; Red’Ko, V.P.; Gerasimyuk, G.I.; Shevchenko, A.V. ChemInform abstract: Synthesis and properties of the compounds M4Zr3O12 and M4Hf3O12 (M: Rare earth element). ChemInform 1991, 22. [Google Scholar] [CrossRef]
  35. Mikuśkiewicz, M.; Moskal, G.; Stopyra, M.; Korol, J. Thermal diffusivity and conductivity of Pr, Eu and Ho zirconates. J. Therm. Anal. Calorim. 2025, 150, 1195–1204. [Google Scholar] [CrossRef]
  36. Chen, Q.; Xu, J.; Wang, H.C.; Wu, J.Z.; Dong, H.; Zhai, B.X.; He, J.; Gao, F. Thermophysical properties of Yb3+-doped nonstoichiometric gadolinium zirconate ceramics. Ceram. Int. 2024, 50, 47813–47823. [Google Scholar] [CrossRef]
  37. Fan, Y.; Bai, Y.L.; Li, Q.; Lu, Z.Y.; Chen, D.; Liu, Y.C.; Li, W.X.; 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]
  38. Herweyer, L.A.; Opila, E.J. High-temperature Na2SO4 interaction with air plasma sprayed Yb2Si2O7 + Si EBC system: Topcoat behavior. J. Am. Ceram. Soc. 2021, 104, 6496–6507. [Google Scholar] [CrossRef]
  39. Chen, L.; Hu, M.Y.; Wu, P.; Feng, J. Thermal expansion performance and intrinsic lattice thermal conductivity of ferroelastic RETaO4 ceramics. J. Am. Ceram. Soc. 2019, 102, 4809–4821. [Google Scholar] [CrossRef]
  40. Ogawa, T.; Matsudaira, T.; Yokoe, D.; Kawai, E.; Kawashima, N.; Fisher, C.A.J.; Habu, Y.; Kato, T.; Kitaoka, S. Spontaneously formed nanostructures in double perovskite rare-earth tantalates for thermal barrier coatings. Acta Mater. 2021, 216, 117152. [Google Scholar] [CrossRef]
  41. Chen, L.; Hu, M.Y.; Zheng, X.D.; Feng, J. Characteristics of ferroelastic domains and thermal transport limits in HfO2 alloying YTaO4 ceramics. Acta Mater. 2023, 251, 118870. [Google Scholar] [CrossRef]
  42. Tian, L.J.; Peng, F.; Song, X.M.; Zheng, W.; Liu, Z.W.; Huang, Y.L.; Zeng, Y. Enhancing the thermodynamic properties of rare-earth niobates through high-entropy and composite engineering. Ceram. Int. 2024, 50, 19488–19501. [Google Scholar] [CrossRef]
  43. Chen, L.; Luo, K.R.; Li, B.H.; Hu, M.Y.; Feng, J. Mechanical property enhancements and amorphous thermal transports of ordered weberite-type RE3Nb/TaO7 high-entropy oxides. J. Adv. Ceram. 2023, 12, 399–413. [Google Scholar] [CrossRef]
  44. Yang, J.; Li, L.; Li, C.R.; Shi, D.D.; Pan, W. Phase transformation inhibition and improved thermo-mechanical properties of rare earth niobates for thermal barrier coatings. Ceram. Int. 2024, 50, 29799–29805. [Google Scholar] [CrossRef]
  45. Gan, M.D.; Lai, L.P.; Wang, J.K.; Wang, J.; Chen, L.; He, J.J.; Feng, J.; Chong, X. Suppressing the oxygen-ionic conductivity and promoting the phase stability of the high-entropy rare earth niobates via Ta substitution. J. Mater. Sci. Technol. 2025, 209, 79–94. [Google Scholar] [CrossRef]
  46. Chen, L.; Wang, Y.T.; Hu, M.Y.; Zhang, L.Y.; Wang, J.K.; Zhang, Z.B.; Liang, X.B.; Guo, J.; Feng, J. Achieved limit thermal conductivity and enhancements of mechanical properties in fluorite RE3NbO7 via entropy engineering. Appl. Phys. Lett. 2021, 118, 071905. [Google Scholar] [CrossRef]
  47. Song, G.D.; Yang, J.P.; Dong, Y.J.; Cui, C. Failure mechanism of plasma sprayed physical vapor deposition (PS-PVD) thermal barrier coating under high temperature and high velocity aviation kerosene gas thermal shock. Corros. Prot. 2022, 43, 20–25. [Google Scholar]
  48. Wei, L.L.; Guo, L.; Li, M.Z.; Guo, H.B. Calcium-magnesium-alumina-silicate (CMAS) resistant Ba2REAlO5 (RE = Yb, Er, Dy) ceramics for thermal barrier coatings. J. Eur. Ceram. Soc. 2017, 37, 4991–5000. [Google Scholar] [CrossRef]
  49. Zhang, Z.J.; Sun, J.; Liu, G.H.; Han, Y.; Liu, W.; Li, Y.; Wang, W.; Liu, X.Y.; Zhang, P.; Pan, W.; et al. High toughness and CMAS resistance of REAlO3/RE2Zr2O7 (RE = La, Nd, Sm, Eu, Gd, and Dy) composites with eutectic composition for thermal barrier coatings. J. Adv. Ceram. 2024, 13, 800–809. [Google Scholar] [CrossRef]
  50. Li, H.Y.; Jia, Z.J.; Zhang, L.; Tian, Y.; Liu, X.G.; Bao, Y.W.; Wan, D.T. A review on the development of prestressed ceramics. Adv. Ceram. 2024, 45, 100–109. [Google Scholar]
  51. Chen, L.; Hu, M.Y.; Feng, J. Mechanical properties, thermal expansion performance and intrinsic lattice thermal conductivity of AlMO4 (M = Ta, Nb) ceramics for high-temperature applications. Ceram. Int. 2019, 45, 6616–6623. [Google Scholar] [CrossRef]
  52. Aladdin Reagents. Available online: https://www.aladdin-e.com/zh_cn/ (accessed on 3 October 2025).
  53. Bai, Y.; Han, Z.H.; Li, H.Q.; Xu, C.; Xu, Y.L.; Ding, C.H.; Yang, J.F. Structure-property differences between supersonic and conventional atmospheric plasma sprayed zirconia thermal barrier coatings. Surf. Coat. Technol. 2011, 205, 3833–3839. [Google Scholar] [CrossRef]
  54. Liu, R.X.; Liang, W.P.; Miao, Q.; Zhao, H.; Zhang, X.F.; Zhang, M.; Yan, R.X.; Ramasubramanian, B.; Ramakrishna, S. Revealing the oxidation growth mechanism and crack evolution law of novel Si-HfO2/Yb2Si2O7/Yb2SiO5 environmental barrier coatings during thermal cycling. J. Adv. Ceram. 2024, 13, 1677–1696. [Google Scholar] [CrossRef]
  55. Chen, L.; Wang, W.J.; Li, J.H.; Yang, G.J. Suppressing the phase-transition-induced cracking of SiO2 TGOs by lattice solid solution. J. Eur. Ceram. Soc. 2023, 43, 3201–3215. [Google Scholar] [CrossRef]
  56. Opila, E.J. Oxidation and volatilization of silica formers in water vapor. J. Am. Ceram. Soc. 2003, 86, 1238–1248. [Google Scholar] [CrossRef]
  57. Parker, C.G.; Opila, E.J. Stability of the Y2O3-SiO2 system in high-temperature, high-velocity water vapor. J. Am. Ceram. Soc. 2020, 103, 2715–2726. [Google Scholar] [CrossRef]
  58. Schlichting, K.W.; Padture, N.P.; Klemens, P.G. Thermal conductivity of dense and porous yttria-stabilized zirconia. J. Mater. Sci. 2001, 36, 3003–3010. [Google Scholar] [CrossRef]
  59. Shannon, R.D.; Prewitt, C.T. Effective ionic radii in oxides and fluorides. Acta Cryst. B 1969, 25, 925–946. [Google Scholar] [CrossRef]
  60. Robinson, K.; Gibbs, G.V.; Ribbe, P.H. Quadratic elongation: A quantitative measure of distortion in coordination polyhedra. Science 1971, 172, 567–570. [Google Scholar] [CrossRef]
  61. Tian, Z.L.; Lin, C.F.; Zheng, L.Y.; Sun, L.C.; Li, J.L.; Wang, J.Y. Defect-mediated multiple-enhancement of phonon scattering and decrement of thermal conductivity in (YxYb1-x)2SiO5 solid solution. Acta Mater. 2018, 144, 292–304. [Google Scholar] [CrossRef]
  62. Du, Z.J.; Shan, Z.T.; Qiao, A.; Tao, H.Z.; Yue, Y.Z. Tantalum-lanthanum mixing effect on structural and mechanical properties of aluminate glasses. J. Non-Cryst. Solids. 2025, 650, 123340. [Google Scholar] [CrossRef]
  63. Xie, M.X.; Xiang, C.Y.; Zeng, X.L.; Zhang, X.S.; Zhang, L.J.; Xiong, J.T. Optimization of the Raman spectroscopic method for crystal orientation and residual stress of aluminum oxide. J. Mater. Res. Technol. 2024, 33, 8265–8276. [Google Scholar] [CrossRef]
  64. Spitsberg, I.; Steibel, J. Thermal and environmental barrier coatings for SiC/SiC CMCs in aircraft engine applications. Int. J. Appl. Ceram. Technol. 2004, 1, 291–301. [Google Scholar] [CrossRef]
  65. Ning, D.; Huang, H.J.; Chao, R.X.; Lin, S.; Bi, Y.T. Lattice distortion, mechanical and thermodynamic properties of (TiZrHf)C and (TiZrHf)N ceramics. Appl. Phys. A 2023, 129, 720. [Google Scholar]
  66. Zhang, H.; Su, J.B.; Duo, S.W.; Zhou, X.; Yuan, J.Y.; Dong, S.J.; Yang, X.; Zeng, J.Y.; Jiang, J.N.; Deng, L.H.; et al. Thermal and mechanical properties of Ta2O5 doped La2Ce2O7 thermal barrier coatings prepared by atmospheric plasma spraying. J. Eur. Ceram. Soc. 2019, 39, 2379–2388. [Google Scholar] [CrossRef]
  67. Zhu, S.S.; Zhu, J.P.; Ye, S.B.; Yang, K.J.; Li, M.L.; Wang, H.L.; He, J.L. High-entropy rare earth titanates with low thermal conductivity designed by lattice distortion. J. Am. Ceram. Soc. 2023, 106, 6279–6291. [Google Scholar] [CrossRef]
  68. Tian, Z.L.; Zhang, J.; Zhang, T.Y.; Ren, X.M.; Hu, W.P.; Zheng, L.Y.; Wang, J.Y. Towards thermal barrier coating application for rare earth silicates RE2SiO5 (RE = La, Nd, Sm, Eu, and Gd). J. Eur. Ceram. Soc. 2019, 39, 1463–1476. [Google Scholar] [CrossRef]
  69. Liu, X.Y.; Zhang, P.; Huang, M.Z.; Han, Y.; Xu, N.; Li, Y.; Zhang, Z.J.; Pan, W.; Wan, C.L. Effect of lattice distortion in high-entropy RE2Si2O7 and RE2SiO5 (RE = Ho, Er, Y, Yb, and Sc) on their thermal conductivity: Experimental and molecular dynamic simulation study. J. Eur. Ceram. Soc. 2023, 43, 6407–6415. [Google Scholar] [CrossRef]
  70. Moskal, G.; Jasik, A.; Mikuśkiewicz, M.; Jucha, S. Thermal resistance determination of Sm2Zr2O7 + 8YSZ composite type of TBC. Appl. Surf. Sci. 2020, 515, 145998. [Google Scholar] [CrossRef]
  71. Vourdas, N.; Marathoniti, E.; Pandis, P.K.; Argirusis, C.; Sourkouni, G.; Legros, C.; Mirza, S.; Stathopoulos, V.N. Evaluation of LaAlO3 as top coat material for thermal barrier coatings. Trans. Nonferrous Met. Soc. China 2018, 28, 1582–1592. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of SiO2-AlTaO4 ceramics. (a) 2-Theta = 10–70°; (b) 2-Theta = 27–31°.
Figure 1. XRD patterns of SiO2-AlTaO4 ceramics. (a) 2-Theta = 10–70°; (b) 2-Theta = 27–31°.
Coatings 15 01204 g001
Figure 2. The XRD Rietveld refinements of SiO2-AlTaO4. (a) 30 h; (b) 60 h; (c) 90 h; and (d) 120 h.
Figure 2. The XRD Rietveld refinements of SiO2-AlTaO4. (a) 30 h; (b) 60 h; (c) 90 h; and (d) 120 h.
Coatings 15 01204 g002
Figure 3. Crystal structures of SiO2-AlTaO4 ceramics. (a) Three-dimensional structure; (b) Viewing along the b axis.
Figure 3. Crystal structures of SiO2-AlTaO4 ceramics. (a) Three-dimensional structure; (b) Viewing along the b axis.
Coatings 15 01204 g003
Figure 4. The distortion degree of polyhedra in lattices of SiO2-AlTaO4 ceramics obtained from the XRD refinement.
Figure 4. The distortion degree of polyhedra in lattices of SiO2-AlTaO4 ceramics obtained from the XRD refinement.
Coatings 15 01204 g004
Figure 5. Typical microstructures and EDS mapping of SiO2-AlTaO4 after different heat treatment times: (a) 30 h; (b) 60 h; (c) 90 h; and (d) 120 h.
Figure 5. Typical microstructures and EDS mapping of SiO2-AlTaO4 after different heat treatment times: (a) 30 h; (b) 60 h; (c) 90 h; and (d) 120 h.
Coatings 15 01204 g005
Figure 6. (a) EDS spectra of SiO2-AlTaO4 after heat treatment at 1400 °C for 30 h; (b) Raman spectra from 50 to 1100 cm−1 for each sample; and (c) The fitted Raman spectrum of the 30 h.
Figure 6. (a) EDS spectra of SiO2-AlTaO4 after heat treatment at 1400 °C for 30 h; (b) Raman spectra from 50 to 1100 cm−1 for each sample; and (c) The fitted Raman spectrum of the 30 h.
Coatings 15 01204 g006
Figure 7. Thermal expansion performance of SiO2-AlTaO4, AlTaO4, and AlNbO4 ceramics [51]: (a) Thermal expansion rate; (b) TECs.
Figure 7. Thermal expansion performance of SiO2-AlTaO4, AlTaO4, and AlNbO4 ceramics [51]: (a) Thermal expansion rate; (b) TECs.
Coatings 15 01204 g007
Figure 8. Thermal properties of SiO2-AlTaO4 and other oxides [51,68,69,70,71]: (a) Thermal diffusivities; (b) Thermal conductivities.
Figure 8. Thermal properties of SiO2-AlTaO4 and other oxides [51,68,69,70,71]: (a) Thermal diffusivities; (b) Thermal conductivities.
Coatings 15 01204 g008
Table 1. Lattice constants and theoretical density (ρ) of SiO2-AlTaO4 ceramics obtained from the XRD Rietveld refinements.
Table 1. Lattice constants and theoretical density (ρ) of SiO2-AlTaO4 ceramics obtained from the XRD Rietveld refinements.
Samplea (Å)b (Å)c (Å)β (°)V3)ρ
(g·cm−3)
Porosity
30 h12.1353.7766.449107.730281.4696.23830.3%
60 h12.1363.7796.449107.710281.7476.23236.1%
90 h12.1353.7776.449107.720281.5596.23639.7%
120 h12.1483.7666.458107.705281.4556.23838.9%
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

Wu, B.; Zhang, L.; Chen, L.; Wang, J.; Gao, Z.; Feng, J. Influences of SiO2 Additions on the Structures and Thermal Properties of AlTaO4 Ceramics as EBC Materials. Coatings 2025, 15, 1204. https://doi.org/10.3390/coatings15101204

AMA Style

Wu B, Zhang L, Chen L, Wang J, Gao Z, Feng J. Influences of SiO2 Additions on the Structures and Thermal Properties of AlTaO4 Ceramics as EBC Materials. Coatings. 2025; 15(10):1204. https://doi.org/10.3390/coatings15101204

Chicago/Turabian Style

Wu, Bingyan, Luyang Zhang, Lin Chen, Jiankun Wang, Zipeng Gao, and Jing Feng. 2025. "Influences of SiO2 Additions on the Structures and Thermal Properties of AlTaO4 Ceramics as EBC Materials" Coatings 15, no. 10: 1204. https://doi.org/10.3390/coatings15101204

APA Style

Wu, B., Zhang, L., Chen, L., Wang, J., Gao, Z., & Feng, J. (2025). Influences of SiO2 Additions on the Structures and Thermal Properties of AlTaO4 Ceramics as EBC Materials. Coatings, 15(10), 1204. https://doi.org/10.3390/coatings15101204

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop