Next Article in Journal
An Innovative Wood Fire-Retardant Coating Based on Biocompatible Nanocellulose Surfactant and Expandable Graphite
Previous Article in Journal
Effect of Organic–Inorganic Mixed Intumescent Flame Retardants on Fire-Retardant Coatings
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigation on the Thermal–Mechanical Properties of YbRESiO5 (RE = Yb, Eu, Gd, Ho, Tm, Lu, Y, Sc): First-Principles Calculations and Thermal Performance Experiments

1
Key Laboratory of Pressure System and Safety, Ministry of Education, East China University of Science and Technology, Shanghai 200237, China
2
Key Laboratory for Advanced Corrosion and Protection of Aviation Materials, AECC Beijing Institute of Aeronautical Materials, Beijing 100195, China
3
Shanghai Institute of Aircraft Mechanics and Control, Shanghai 200092, China
*
Authors to whom correspondence should be addressed.
These authors have contributed equally.
Coatings 2024, 14(8), 1035; https://doi.org/10.3390/coatings14081035
Submission received: 26 July 2024 / Revised: 13 August 2024 / Accepted: 13 August 2024 / Published: 14 August 2024

Abstract

:
Environmental barrier coatings are critically needed in the future to safeguard SiC-based ceramic matrix composites (SiC CMCs) used in gas turbines. Element doping of rare earth monosilicates could further improve the properties of the coating. The crystal structure, elastic properties, and resistance to water vapor corrosion of Yb2SiO5 and YbRESiO5 (where RE = Sc, Y, Eu, Gd, Ho, Tm, Lu) were examined in this study using first-principles calculations. When RE is Yb, the material is Yb2SiO5. Based on the outcomes of the calculation, we prepared YbRESiO5 (RE = Sc, Yb, Eu) and studied the thermodynamic properties. The findings show that YbReSiO5’s resistance to water vapor corrosion is as follows: YbLuSiO5 < YbEuSiO5 < YbGdSiO5 < YbYSiO5 < Yb2SiO5 < YbTmSiO5 < YbHoSiO5 < YbScSiO5. YbScSiO5 has a lower unit cell volume, average Re-O bond length, and thermal expansion coefficient than Yb2SiO5, while YbEuSiO5 has the reverse pattern. Moreover, of the eight materials, YbScSiO5 has the greatest elastic modulus and lattice distortion. After doping with Eu, YbEuSiO5 exhibits a decrease in thermal conductivity by nearly thirty percent compared to Yb2SiO5, due to the formation of oxygen vacancies. The development of environmental barrier coating materials may benefit from these discoveries.

1. Introduction

As the aviation sector develops, thrust–weight ratios increase in new generations, and internal gas temperatures exceed the bearing limits of nickel-based superalloys [1,2]. According to the needs of the current development, the surface temperature of the hot end components of the new generation of aero-engines can reach 1400 °C, while the maximum working temperature of single-crystal nickel-based superalloys is 1100 °C and Ni3Al single-crystals have a high-temperature limit of 1200 °C, so there is an urgent need to develop a new type of material matrix [3]. The most promising high-temperature structural materials to replace nickel-based superalloys in the hot sections of aircraft engines are SiC-based ceramic matrix composites (SiC CMCs) [1,4,5]. SiC CMCs have a high temperature resistance, low density, high strength, high elastic modulus, oxidation resistance, erosion resistance, and crack insensitivity [6]. But in real high-temperature service conditions, SiC CMCs face substantial hurdles, especially with regard to considerable water vapor corrosion issues [4]. Poor environmental durability still prevents them from being used in combustion environments due to the presence of two chemical degradation mechanisms, recession and thermal corrosion. (i) Recession is the loss of silicon-based ceramics due to repeated oxidation and volatilization when exposed to high-velocity water vapor in a turbine. (ii) Thermal corrosion is caused by alkaline salts in the combustion environment, which form liquid silicates and cause craters in silicon-based ceramics [7,8,9]. Alkaline salts are salts formed by the reaction of an alkaline metal (e.g., sodium, potassium) or an alkaline earth metal (e.g., calcium, magnesium) with an acid. Experimental results from NASA et al. show that the loss rate of the ceramic surface is about 270 μm/kh in a combustion environment (1200 °C; pressure 1013.25 KPa; airflow velocity 90 m/s). For the protection of the substrate, environmental barrier coatings (EBCs) must be used. Due to their resistance to water vapor corrosion, high temperature phase stability, and matching thermal expansion coefficient with SiC CMCs substrates, rare earth silicate materials in EBC systems are currently receiving a lot of attention [1,10,11].
EBCs have evolved through three generations since the 1990s [12]. The initial generation of EBC systems, YSZ/Mullite, could not meet the practical application requirements under continuous thermal cycling because of crack propagation, which limited its ability to resist a few hundred hours in a water vapor atmosphere at 1300 °C [13,14]. Later, BSAS (1-xBaO-xSrO-Al2O3-2SiO2, 0 ≤ x ≤ 1) materials with low thermal expansion coefficients (4 × 10−6·K−1–5.15 × 10−6·K−1, and thus a close match with SiC, 4.02 × 10−6·K−1) and a low elastic modulus (60–70 Gpa) were used to develop the second generation of EBCs [15]. These materials replaced YSZ as the surface coating, or they enhanced Mullite with BSAS + Mullite to form the BSAS/Mullite/Si system. Unfortunately, at 1584 K, the BSAS material system experiences a glass phase shift that shortens the coating’s lifespan by causing cracking [13,14,16]. BASA’s volatilization loss for 1000 h at 1400 °C, 6 atm total pressure and 24 m/s gas flow rate is as high as 70 μm. The third-generation rare earth silicate system has a high melting point (1950 °C), a low thermal conductivity (2.3 W·m−1·K−1–1.5 W·m−1·K−1), and a coefficient of thermal expansion (6–8 × 10−6·K−1) similar to that of silicon-based ceramic materials (5 × 10−6·K−1). And according to research, it can serve thousands of hours in a water vapor environment at 1482 °C, with excellent water vapor corrosion resistance. Therefore, rare earth silicate materials are drawing more and more attention [13,14,17,18,19].
Two distinct crystal phases can be distinguished in RE2SiO5 based on variations in rare earth elements (RE). All RE2SiO5 are monoclinic and stable as either space group P21/c or C2/c, which are generally termed the X1 and X2 structure. Re2SiO5 forms the X1 phase when the radius of the rare earth elements (RE = La, Ce, Pr, Nd, Sm, Eu, Tb, and Gd) is bigger, i.e., in the atomic numbers from 57 to 64. RE2SiO5 forms the X2 phase when the radius of the rare earth elements (RE = Dy, Ho, Er, Tm, Yb, and Lu) is lower [20], i.e., in the atomic numbers from 66 to 71. According to Mackenzie Ridley’s research, RE2SiO5 is in the X2 phase when the average radius of rare earth elements is less than 0.95 Å [21]. Two crystal structures are present in rare earth monosilicates; the X1 phase is stable at low temperatures (800–1200 °C) and the X2 phase emerges and is stable at high temperatures (1450–1950 °C) [22]. Volume differences brought forth by changes in the crystal shape affect stress fields and raise the possibility of cracking. Consequently, X2-phase monosilicates at high temperatures (1450–1550 °C) are frequently utilized in Environmental Barrier Coatings (EBCs). Re2SiO5, in X2 phase compounds, includes one silicon atom (Si), five oxygen atoms (O1–O5), and two distinct locations for rare earth elements (RE1 and RE2). While the other oxygen atom forms a loose link with the rare earth elements, the four separate oxygen atoms (O1–O4) combine to create a [SiO4] tetrahedron with Si. The two distinct rare earth atoms have coordination numbers of 7 and 6, respectively, resulting in the compounds [REO7] and [REO6]. Whereas [REO7] and [REO6] are comparatively softer, the [SiO4] tetrahedral structure is strong and has stiff characteristics. The mechanical and thermal properties of the material are greatly impacted by these features. The rigidity of the [SiO4] tetrahedra endows the material with a higher modulus of elasticity and lower coefficient of thermal expansion, and thus the material typically exhibits better structural stability. In contrast, the weaker bonding of the [REO7] and [REO6] units allows for increased flexibility of the material, but may result in higher coefficients of thermal expansion and lower moduli of elasticity [20,21,23].
For multi-component rare earth monosilicates, the introduction of rare earth elements needs to be carefully selected, and different rare earth elements will have an impact on the Si-O bond strength and mechanical properties. It has been shown that the strength of the Si-O bond is proportional to the water vapor corrosion resistance of multi-component rare earth monosilicates, and the higher the strength of the Si-O bond, the stronger the water vapor corrosion resistance. For example, K. N. Lee [17] experimentally showed that the water vapor corrosion resistance of Yb2SiO5 is greater than that of Lu2SiO5. And Han et al. [24] calculated the Si-O bond length of Yb2SiO5 to be 1.6137 Å by the first nature principle, which is smaller than the Si-O bond length of Lu2SiO5 of 1.6246 Å [24]. The flexural strength, elastic modulus, and thermal shock resistance of monoclinic Re2SiO5 for orthosilicates are dependent on the mass and radius of the rare earth cations, as demonstrated by both experimental and Density Functional Theory (DFT) calculations. For example, according to the experiments, RE2SiO5 increases the elastic modulus from Tb2SiO5 (143.72 Gpa) to Lu2SiO5 (171.3 Gpa) and the shear modulus from Tb2SiO5 (61 Gpa) to Lu2SiO5 (71 Gpa) as the radius of the rare earth elements decreases [7].
In order to study the lattice distortions as well as the mechanical properties of Yb2SiO5 and YbRESiO5 (where RE = Sc, Y, Eu, Gd, Ho, Tm, Lu), we calculated the crystal structures and the elastic properties of the materials using first-principles calculation. Furthermore, the Si-O bond lengths of eight different rare earth monosilicates were simulated to assess the materials’ resistance to water vapor corrosion. Additionally, Yb2SiO5, YbScSiO5, and YbEuSiO5 bulk materials were produced utilizing solid-phase reaction techniques based on the outcomes of first-principles calculations in order to precisely investigate the impact of various rare earth cations on the mechanical and thermodynamic properties of the materials. Microstructural analysis and thermodynamic property evaluations of the composite materials were conducted. Last but not least, by investigating the effects of the crystal structure, coefficient of thermal expansion, thermal conductivity, and elastic constants of different rare earth monosilicates, this study may provide a theoretical framework for the creation of novel EBC materials.

2. Computational Models and Experimental Procedures

2.1. Computational Models

One silicon atom (Si), five oxygen atom sites (O1–O5), and two distinct rare earth atom sites (Yb1 and Yb2) make up the Yb2SiO5 unit cell. While the remaining oxygen atom loosely connects with the rare earth atoms, four of the oxygen atoms (O1–O4) form a [SiO4] tetrahedron with the silicon atom. The two distinct rare earth atoms have coordination numbers of 7 and 6, respectively, resulting in the formation of [YbO7] and [YbO6] polyhedra [25]. Therefore, in the actual process, because larger atomic radius elements tend to occupy sites with larger coordination numbers, in the model of Yb2SiO5, to better approximate real conditions, rare earth elements (RE = Eu, Gd, Ho, Tm) with atomic radii larger than Yb were added into the Re1 site, while rare earth elements (RE = Sc, Y, Lu) with atomic radii smaller than Yb were added into the RE2 site [26]. The conventional and primitive cells of Yb2SiO5 are shown in Figure 1 a and b, respectively.
CASTEP was used to do computations from first principles. The most stable crystal structure at zero temperature was obtained by geometric optimization of the crystal structure of each rare earth silicate, which was then analyzed before the material’s resistance to water vapor corrosion and its thermodynamic properties were assessed. The crystal structure optimization method was carried out by expanding Bloch waves in reciprocal space with a cutoff energy of 450 eV using a plane wave basis set under periodic boundary conditions [27]. The first irreducible Brillouin zone was used as the discrete 2 × 3 × 4k sample grid for the Monkhorst–Pack technique of energy integration. The first irreducible Brillouin zone is the smallest, uniquely defined region in reciprocal space that contains all the distinct momentum states of a crystal, crucial for simplifying electronic band structure calculations. Regarding exchange-correlation energies, the local density approximation (LDA) was employed [28]. To maximize the crystal structure, separate modifications were performed to the lattice constants and the relative locations of interior atoms. We used the Broyden–Fletcher–Goldfarb–Shanno (BFGS) minimization approach to minimize both the interatomic forces and the overall energy [29]. We computed the electrostatic interactions between valence electrons and ionic cores using ultra-soft pseudopotentials. Atomic forces within 0.03 eV/Å, energy differences within 1 × 10−5 eV/atom, maximum stress within 0.05 GPa, and maximum ion displacement within 1 × 10−3 Å were the convergence conditions for geometric optimization. A distance of 3 Å was chosen as the Mulliken population cutoff distance (which is the threshold distance used to determine which atomic contributions to electronic charge density are considered significant in the Mulliken population analysis) [24].
By linearly fitting stress against applied homogeneous elastic strains, the elastic constants were found. The strain–stress relationship approach developed by Milman and Warren was used to derive second-order elastic constants [30]. Four sets of uniform strain configurations (which refer to four different strain states applied in various directions and magnitudes to study changes in a material’s structure and properties) were used to calculate the 13 independent elastic constants (c11, c22, c33, c44, c55, c66, c12, c13, c23, c15, c25, c35, and c46) of Yb2SiO5. Internal atomic degrees of freedom were optimized to yield the stresses. The linear fitting of computed stresses as a function of strain allowed for the determination of elastic stiffness. Maximum ion displacement differences were defined at 2 × 10−4 Å, total energy differences within 2 × 10−6 eV/atom, and Hellmann–Feynman force (which is the force on an atom or particle derived directly from the gradient of the potential energy with respect to atomic positions) differences within 0.006 eV/Å as the convergence criterion for optimizing internal atomic degrees of freedom. The average sound velocity ( v m ), Debye temperature ( θ D ), bulk modulus ( B ), shear modulus ( G ), Young’s modulus ( E ), and Poisson’s ratio (υ) were then calculated using the elastic stiffness constants ( c i j ) [31].

2.2. Material Preparation and Microstructure Characterization

The Yb2SiO5, YbScSiO5, and YbEuSiO5 bulk materials were prepared using the solid-state reaction method. The raw materials included Sc2O3, Ho2O3, Yb2O3 (≥99.9 wt.% from Shanghai Maclin Biochemical Technology Co., Ltd., Shanghai, China), and SiO2 (≥99.99 wt.% from Shanghai Maclin Biochemical Technology Co., Ltd.). The powdered raw materials were first placed in crucibles and dried in an oven for two hours at 393 K. After that, they were weighed using a molar ratio of 2:1 for silicon to rare earth elements. After that, a clean stainless-steel tank was filled with the materials, milling balls, and anhydrous ethanol in a mass ratio of 1:2:1. The tank was then subjected to a 10 h ball milling process at 270 revolutions per minute. The slurry that had been ground was dried in ovens for 12 h at 353 K and 393 K. Subsequently, the coarse oxides were crushed into powders using an agate grinding disc and sieved through an 80-mesh screen to achieve fine powders. After that, these fine powders were put into molds measuring φ15 mm, and they were uniaxially pressed for five minutes at 15 MPa before oil pressure was applied. For thirty minutes, an oil pressure of 300 MPa was applied. To obtain the final products, the samples were sintered for 10 h at 1823 K.
Phase analysis of the sintered samples was carried out with an X-ray diffractometer (XRD, D/max, 2550VB/PC, Rigaku, Tokyo, Japan) filtered with Cu Kα radiation at 18 KV accelerating voltage and 450 mA current. Using a step size of 10°/min, the diffraction angles were scanned from 10° to 80°. A scanning electron microscope (SEM, HITACHI S-3400 N, Japan) fitted with an energy-dispersive spectrometer (EDS, AMETEK APOLLO X, Berwyn, PA, USA) was used for microstructural examination.

2.3. Thermal Properties

Using standard metallographic techniques, all samples used in the thermal treatment experiments were cut from sintered bulk silicate.
Using a laser flash analyzer (NETZSCH, LFA 427, Germany), the bulk silicate’s thermal diffusivity was determined throughout a temperature range of 300 K to 1473 K. Sample dimensions were 12.7 mm × 2 mm. To reduce heat radiation, a thin layer of graphite was applied to the samples’ two sides prior to testing.
The thermal conductivity (k) of the material was calculated using the following formula, derived from thermal diffusivity ( D t h ), density ( ρ ), and specific heat capacity ( c p ) [32]:
k = D t h · ρ · c p
Additionally, the density ( ρ ) was determined using Archimedes’ method (which is using the principle of buoyancy to determine the volume of an irregularly shaped object) [33], while the specific heat capacity ( c p ) was calculated using the Neumann–Kopp law (which states that the heat capacity of a substance is proportional to the number of atoms or molecules in a compound) [34]. Furthermore, the thermal conductivity of fully dense samples ( k 0 ) was computed [35].
k 0 = k 1 4 ψ 3
k is the theoretical thermal conductivity, ψ is the porosity. Additionally, thermal expansion coefficient-related parameters of bulk silicates were measured using a vertical high-temperature dilatometer (NETZSCH, DIL 402C, Germany) in the range of 303 K to 1473 K. The formula for calculating the thermal expansion coefficient ( a ) is as follows [31]:
a = L L · L
Here, L represents the original length of the sample, and L denotes the length change of the sample under the temperature difference T . The samples used were rectangular rods measuring 2 mm × 6 mm × 11 mm in the three dimensions.

3. Result and Discussion

3.1. First-Principles Calculations

After geometric optimization, the computed and theoretical lattice parameters of Yb2SiO5 are shown in Table 1. Deviation in this section is defined as the gap between the geometrically optimized lattice parameters of the material and the theoretical lattice parameters. Less than 3% separates the lattice constants from their theoretical values, indicating that the computations are reliable.
The crystal structures of YbRESiO5 (RE = Sc, Y, Ho, Eu, Tm, Lu, Gd) following geometric optimization are displayed in Figure 2. Due of variations in rare earth element types, the locations of Si and O atoms vary during geometric optimization, leading to changes in lattice constants and subsequently affecting the unit cell volumes. The crystal forms with various rare earth element dopants following geometric optimization are shown in Figure 2.
After geometric optimization with addition of different rare earth elements, Table 2 shows the lattice parameters and unit cell volumes of YbRESiO5. According to computational analysis, the largest unit cell volume is produced by doping with the largest atomic radius element, Eu, and the smallest unit cell volume is produced by adding the element Sc. The overall trend is YbScSiO5 (750.4240 Å 3 ) < YbHoSiO5 (794.8210 Å 3 ) < Yb2SiO5 (794.7620 Å 3 ) < YbTmSiO5 (800.1510 Å 3 ) < YbGdSiO5 (806.6770 Å 3 ) < YbYSiO5 (809.2980 Å 3 ) < YbLuSiO5 (812.4510 Å 3 ) < YbEuSiO5 (813.6240 Å 3 ).

3.1.1. Crystal Lattice Distortion Calculation

To quantitatively describe the distortion of polyhedra, the following equation is used to calculate the lattice distortion within the material [25]:
d = 1 n n d M O n [ d M O R e ] [ d M O R e ] 2
Here, d M O R e represents the average bond length of Re-O and Si-O in YbReSiO5 (Re = Sc, Y, Lu, Eu, Gd, Tm, Ho), while d M O n represents the bond length between the rare earth element (Re) or silicon (Si) and the nth oxygen atom.
Table 3 gives the bond lengths, average bond lengths, and lattice distortions of [SiO4], [REO6], and [REO7] in YbRESiO5 (RE = Sc, Y, Lu, Eu, Gd, Tm, Ho). We discovered that adding Sc, the element with the smallest atomic radius in our investigation, drastically alters the bond lengths and lattice distortions of the material when compared to Yb2SiO5. In comparison to Yb2SiO5, YbScSiO5 has the highest lattice distortions in [REO6] and [REO7], rising by approximately 50% and 60%, respectively. It also has the shortest average RE-O bond lengths in [REO6] and [REO7]. Compared to Yb2SiO5, its average RE-O bond lengths decrease by 0.5% and 4.4%. Conversely, among elements that increase the RE-O bond lengths, adding Eu resulted in the longest RE-O bond lengths in [REO6] and [REO7] for YbEuSiO5, rising by roughly 1.3% and 1.5%, respectively, compared to Yb2SiO5.
In this study, the average RE-O bond lengths in [REO6] polyhedra for the eight different rare earth monosilicates are ordered as follows: YbScSiO5 (2.10674 Å) < YbHoSiO5 (2.21699 Å) < YbTmSiO5 (2.21344 Å) < Yb2SiO5 (2.21750 Å) < YbLuSiO5 (2.24643 Å) < YbGdSiO5 (2.23314 Å) < YbEuSiO5 (2.23771 Å). In [REO7] polyhedra, the average RE-O bond lengths are ordered as: YbScSiO5 (2.29198 Å) < Yb2SiO5 (2.30098 Å) < YbTmSiO5 (2.30253 Å) < YbYSiO5 (2.31480 Å) < YbLuSiO5 (2.31456 Å) < YbHoSiO5 (2.32545 Å) < YbGdSiO5 (2.32157 Å) < YbEuSiO5 (2.33001 Å). The lattice distortion in [REO7] is ranked as follows: YbEuSiO5 (1.26024‰) < YbGdSiO5 (1.27335‰) < Yb2SiO5 (1.63300‰) < YbTmSiO5 (1.68259‰) < YbLuSiO5 (1.82196‰) < YbYSiO5 (2.07000‰) < YbHoSiO5 (2.37837‰) < YbScSiO5 (3.31675‰). The lattice distortion in [REO6] polyhedra is ranked as: YbYSiO5 (0.06170‰) < YbLuSiO5 (0.07019‰) < YbGdSiO5 (0.11984‰) < Yb2SiO5 (0.12334‰) < YbTmSiO5 (0.20894‰) < YbHoSiO5 (0.31473‰) < YbEuSiO5 (0.37126‰) < YbScSiO5 (0.47114‰). Among the eight rare earth monosilicates, YbScSiO5 has the biggest lattice distortion and the shortest RE-O bond length. We relate this to the considerable difference in atomic radii between Yb and Sc elements. Additionally, the introduction of various elements also affects the lattice distortion of [SiO4] tetrahedra, but because of their rigid nature, the amplitude of variation is not significant in this study and thus is not discussed in detail here. YbScSiO5 exhibits the largest lattice distortion and shortest RE-O bond length as a result of this heterogeneity in rare earth cation radii.

3.1.2. Water Vapor Corrosion Resistance

Testing the water vapor corrosion resistance in environments similar to combustion but without alumina is necessary because the majority of current experiments on environmental barrier coatings for water vapor corrosion are carried out in alumina tubes, and there are reports that alumina contamination may alter the water vapor corrosion resistance of RE2SiO5 [9]. It is difficult, though, to carry out useful research in such hot conditions without adding alumina materials. Thus, in order to acquire theoretical information on the water vapor corrosion resistance of eight materials, this work uses first-principles calculations. It has been demonstrated that using first-principles calculations to predict the properties of compounds with various components but the same crystal structure works well [24]. For example, the strength of the Si-O linkages can reflect the water vapor corrosion resistance of RE2Si2O7 with the same crystal structure.
The Si-O bonds in rare earth silicates are broken down by steam during high-temperature water vapor corrosion, which corrodes environmental barrier coatings by Si-O degradation. The pace of corrosion is slower if the Si-O bonds are generally stable; on the other hand, the rate of corrosion is higher if the Si-O bonds are weak. As a result, by determining the strength of the Si-O linkages, the water vapor corrosion resistance of rare earth monosilicates may be compared [36]. Si-O bond density, which is defined as the ratio of Si-O bond energy to Si-O bond length, usually directly correlates with the strength of Si-O bonds. Determining the Si-O bond information through first-principles calculations on materials enables evaluation of the water vapor corrosion performance of various rare earth monosilicates.
Eight distinct rare earth monosilicates are shown in Figure 3 and Table 4 along with their Si-O bond densities, Milliken populations, and Si-O bond lengths. Our calculations show that, out of the eight materials, YbScSiO5 has the highest resistance to water vapor corrosion because it has the shortest Si-O link length (1.60243 Å), the highest Milliken population (0.63500), and the greatest bond density (0.3963 Å−1). Following Eu addition, YbEuSiO5 has the longest Si-O bond length (0.61000 Å); nevertheless, it also has a lower Milliken population (0.61000) and bond density (0.3794 Å−1) than Yb2SiO5, suggesting that it is less resistant to water vapor corrosion than Yb2SiO5. The water vapor corrosion resistance of Yb2SiO5 is generally enhanced by adding Sc, Ho, and Tm, and decreased by doping with Gd, Eu, Y, and Lu. Overall, the ranking of water vapor corrosion resistance among the eight different rare earth monosilicates is as follows: YbLuSiO5 (0.3764 Å1) < YbEuSiO5 (0.3794 Å−1) < YbGdSiO5 (0.3810 Å−1) < YbYSiO5 (0.3811 Å−1) < Yb2SiO5 (0.3844 Å−1) < YbTmSiO5 (0.3851 Å−1) < YbHoSiO5 (0.3860 Å−1) < YbScSiO5 (0.3963 Å−1).

3.1.3. Elastic Constants

The elastic constants of YbRESiO5 (RE = Yb, Eu, Gd, Ho, Tm, Lu, Y, Sc) that were determined using first-principles techniques are shown in Table 5. The average sound velocity, Debye temperature, bulk modulus, shear modulus, and Young’s modulus of Yb2SiO5 are all markedly increased by the equimolar addition of Sc, while the addition of Gd has the reverse effect. Poisson’s ratio is the lowest when adding Ho. The order of magnitude for average sound velocity is as follows: YbGdSiO5 (2875.6370 m/s) < YbEuSiO5 (3027.7198 m/s) < YbHoSiO5 (3047.4524 m/s) < YbTmSiO5 (3115.4777 m/s) < Yb2SiO5 (3116.0476 m/s) < YbLuSiO5 (3136.1570 m/s) < YbYSiO5 (3282.3078) < YbScSiO5 (3669.7119 m/s). The ranking of the Debye temperature is YbGdSiO5 (367.7621 K) < YbEuSiO5 (392.4537 K) < YbHoSiO5 (394.2189 K) < YbTmSiO5 (395.7621 K) < Yb2SiO5 (400.2373 K) < YbLuSiO5 (403.6193 K) < YbYSiO5 (419.4569 K) < YbScSiO5 (481.4030 K). The ranking of Young’s modulus is YbGdSiO5 (125.8217 Gpa) < YbEuSiO5 (134.4732 Gpa) < YbHoSiO5 (139.7532 Gpa) < YbYSiO5 (145.4985 Gpa) < YbTmSiO5 (145.6084 Gpa) < YbLuSiO5 (151.7248 Gpa) < Yb2SiO5 (153.4019 Gpa) < YbScSiO5 (162.0600 Gpa).
Studies have indicated that the atomic number of the rare earth element has a substantial impact on the elastic modulus and flexural strength of X2-RE2SiO5 samples. This discovery could be explained by the rare earth element’s decreased ionic radius increasing rare earth–oxygen interactions. In X2-RE2SiO5 samples, the bond strength can be represented by the cation field strength (CFS) [7]. The calculation formula for CFS is as follows [37]:
C F S = Z c / r c 2
Z c represents the cation charge, and r c 2 represents the cation radius squared.
It is clear from the data in Table 6 that when the ionic radius of rare earth elements falls, so does the bonding strength between oxygen and rare earth in X2-RE2SiO5. Consequently, the cation field strength (CFS) of cations having larger radii, including Gd3+, Eu3+, Ho3+, Tm3+, and Y3+, is lower than that of Yb2SiO5. As a result, when adding cations such as Gd3+, Eu3+, Ho3+, Tm3+, and Y3+, the Young’s modulus lowers. Of the eight rare earth monosilicates, the element Sc3+ with the shortest cation radius has the highest CFS and, consequently, the highest Young’s modulus. The findings of the computed elastic modulus, however, do not quite match the CFS ranking. This disparity can be explained by the fact that materials containing Yb frequently show aberrant characteristics in comparison to other lanthanide elements, such as a different melting and boiling point and density. This observation might be explained by the fact that Yb ([Xe]4f146s2) has a closed-shell atomic configuration, meaning that there are only two 6s valence electrons accessible for chemical bonding [7].

3.2. Phases and Microstructure of Sintered Bulk Materials

Three materials’ XRD patterns are displayed in Figure 4: Yb2SiO5, YbScSiO5, and YbEuSiO5. There is no second phase in the prepared silicate bulk materials, according to the measured diffraction peaks between the prepared monosilicates and the standard PDF of the Yb2SiO5 card (97-000-4446). Shifts in the diffraction peaks of YbScSiO5 and YbEuSiO5 relative to the standard Yb2SiO5 PDF card are detected upon doping Yb2SiO5 with Sc and Eu. The diffraction peaks move to larger angles when the smaller cation radius element (Sc3+) is incorporated into Yb2SiO5. This is due to a change in lattice parameters. On the other hand, adding a higher cation radius element (Eu3+) increases the lattice parameters and causes the peaks of diffraction to move to smaller angles. These outcomes agree with the XRD patterns in Figure 4 and the simulated computations in Table 2. The unit cell volume of Yb2SiO5 drops by 5.1% when equal moles of Sc are added, but increases by 2.3% when doped with the same moles of Eu. Consequently, YbScSiO5 (750.4240 Å 3 ) < Yb2SiO5 (794.7620 Å 3 ) < YbEuSiO5 (813.6240 Å 3 ) is the order in which the unit cell volumes of the three materials are arranged.
As seen in the XRD patterns in Figure 4, which show no further phases, the produced rare earth monosilicates often have few pores and no noticeable fissures. Furthermore, porosity levels within 5% were demonstrated by the density and porosity calculations of Yb2SiO5, YbScSiO5, and YbEuSiO5, as indicated in Table 3. The microstructures of cross-sections of the materials Yb2SiO5, YbScSiO5, and YbEuSiO5 are displayed in Figure 5. A homogeneous distribution of elements is revealed by EDS scans of the elemental distribution in the three materials (Figure 6). When paired with the XRD results shown in Figure 4, these results show that the Yb2SiO5, YbScSiO5, and YbEuSiO5 sample preparation procedures were accurate and produced materials that were reasonably pure. The results of various densities and porosities of the three materials are shown in Table 7.

3.3. Thermal Performance Test

3.3.1. Thermal Conductivity

Figure 7 shows the thermal expansion coefficients and specific heat capacities of Yb2SiO5, YbScSiO5, and YbEuSiO5 within the temperature range of 300 K to 1473 K. For these three rare earth monosilicates, the thermal expansion coefficients often decrease with increasing temperature, which is consistent with the typical trend of phonon thermal conductivity [25,38]. Both the thermal conductivity and the thermal expansion coefficient, however, exhibit a rising trend at 800 °C, most likely as a result of radiation heat transfer effects, which have also been noted in other tests [39]. Consequently, statistics on laser thermal conductivity above 800 °C should be regarded as approximations. The thermal diffusivity of Yb2SiO5 ranges from 0.470 mm2·s−1 to 0.788 mm2·s−1 between 300 K and 1473 K. YbScSiO5 exhibits a thermal diffusivity ranging from 0.506 mm2·s−1 to 0.751 mm2·s−1. The lowest thermal expansion coefficient is shown by YbEuSiO5, which ranges from 0.352 mm2·s−1 to 0.610 mm2·s−1.
Three distinct rare earth monosilicates’ actual and theoretical thermal conductivities are displayed in Figure 8. In this investigation, the porosity of all three samples stayed below 5%. The estimated theoretical value of 2.17 W·m−1·K−1 for Yb2SiO5 at ambient temperature closely matches the reported value of 2.2 W·m−1·K−1 for Yb2SiO5 in the literature [40]. The theoretical thermal conductivity of Yb2SiO5 at normal temperatures is 2.07 W·m−1·K−1. Microstructural variations, such as variations in grain size and the impact of infrared radiation at high temperatures, may be the cause of the poorer thermal conductivity and thermal diffusivity of Yb2SiO5 in this investigation. Additionally, YbScSiO5 has a range of 1.85 W·m−1·K−1 to 1.67 W·m−1·K−1, YbEuSiO5 exhibits the lowest thermal conductivity among the three materials, ranging from 1.44 W·m−1·K−1 to 1.07 W·m−1·K−1, and Yb2SiO5 varies from 2.07 W·m−1·K−1 to 1.52 W·m−1·K−1. Overall, the addition of Sc elements results in very little change in thermal conductivity and thermal diffusivity; at 650 K, YbScSiO5 exhibits slightly higher values than Yb2SiO5, but before this temperature, the tendency is the opposite. On the other hand, YbEuSiO5′s diffusivity and thermal conductivity dramatically drop with the addition of Eu. The experiment’s observation of radiation heat transfer may be the cause of the rise in thermal conductivity at 1073 K.
This study states that the thermal conductivity of Sc2SiO5 material is 1.67 times greater than that of Yb2SiO5 and varies from 2.50 W·m−1·K−1 to 3.50 W·m−1·K−1 between 400 K and 1673 K. Eu2SiO5 has a lower thermal conductivity than Yb2SiO5, ranging from 1.50 W·m−1·K−1 to 1.25 W·m−1·K−1. As a result, it drops to 1.25 W·m−1·K−1, which is less than Yb2SiO5 [26,35]. Thus, it can be deduced that there may be a positive correlation between the thermal conductivity of rare earth monosilicate materials doped with binary equimolar elements and the thermal conductivity of their parent materials. Additionally, YbScSiO5 and YbEuSiO5 exhibit greater specific heat capacity than Yb2SiO5, with the inclusion of Sc elements having a particularly notable effect. Yb2SiO5 with Sc may somewhat raise the material’s thermal conductivity, while doping the material with Eu may result in a minor drop in thermal conductivity. Nonetheless, the square root of thermal conductivity and phonon scattering coefficient are inversely proportional in materials with added point defects because of the large lattice distortion in YbScSiO5 and the adherence to phonon heat transfer principles in those materials. Higher phonon scattering coefficients can result from larger lattice distortions and mismatched rare earth cation weights and radii. This could account for YbScSiO5’s weaker thermal diffusivity and conductivity than Yb2SiO5 at temperatures lower than 600 K.
The outer electron structure of Eu, which is 4f76s2, is another factor contributing to YbEuSiO5’s lowest thermal conductivity. Its outer electron structure becomes 4f7 when it loses two electrons. Based on Hund’s rule, 4f7 is in a stable half-filled state. Eu2+ ions, which encourage the creation of oxygen vacancies and drastically lower thermal conductivity, may therefore be present in YbEuSiO5 [26].
Overall, compared to Yb2SiO5, the insertion of Sc components exhibits little overall change in thermal conductivity and thermal diffusivity. On the other hand, compared to Yb2SiO5, the addition of Eu elements leads to much decreased thermal diffusivity and thermal conductivity because of the development of oxygen vacancies. In particular, YbEuSiO5 has a minimum thermal conductivity of 1.07 W·m−1·K−1 in the temperature range of 303 K to 1473 K, which is 29% less than Yb2SiO5’s (1.52 W·m−1·K−1). Moreover, YbEuSiO5’s (1.44 W·m−1·K−1 to 1.07 W·m−1·K−1) thermal conductivity fluctuates less in this temperature range than Yb2SiO5’s (2.07 W·m−1·K−1 to 1.52 W·m−1·K−1). Furthermore, the stability of the temperature and stress fields is aided by the low range of thermal conductivity parameters. For materials that are monosilicate, these elements are essential.

3.3.2. CTE

Figure 9 shows the coefficient of thermal expansion (CTE) and ΔL/L curves of Yb2SiO5, YbScSiO5, and YbEuSiO5 from 303 K to 1473 K. Compared to Yb2SiO5, YbEuSiO5 exhibits an increasing trend in ΔL/L, while YbScSiO5 shows a decreasing trend. Here, ΔL is the only representation of the change because L, the original length, is a constant. In contrast to Yb2SiO5, the effects of Sc and Eu doping on ΔL are opposite. In order to get precise coefficients of thermal expansion for monosilicate materials, as shown in Figure 8, it is noted that, for Yb2SiO5, the CTE varies between 5.86 × 10−6·K−1 and 8.05 × 10−6·K−1 over the 600 K to 1473 K temperature range. The results of this investigation closely match those of Tian et al. and mistakes lower than 0.5 × 10−6·K−1 can often be ignored, as they are most likely the result of experimental and sampling mistakes. The addition of Eu to YbEuSiO5 raises its CTE in comparison to Yb2SiO5, while Sc doping lowers the CTE of YbScSiO5, following the ΔL/L trends. YbScSiO5 and YbEuSiO5 show CTE values of 5.64 × 10−6·K−1 to 7.56 × 10−6·K−1 and 6.16 × 10−6·K−1 to 8.6 × 10−6·K−1, respectively, throughout the 600 K to 1473 K range. Compared to Yb2SiO5, the addition of Sc lowers the material’s overall CTE as well as the amplitude of its temperature-dependent CTE variation. On the other hand, Eu’s introduction has the opposite impact.
The coefficient of thermal expansion (CTE) of YbScSiO5 doped with Sc is often lower than that of Yb2SiO5, but the CTE of YbEuSiO5 doped with Eu is higher. YbScSiO5 has the lowest CTE at 1473 K, 7.56 × 10−6·K−1, 7% less than Yb2SiO5, whereas YbEuSiO5 has the greatest CTE at 8.6 × 10−6·K−1, 7.5% more than Yb2SiO5.
The coefficient of thermal expansion (CTE) of rare earth monosilicates should be correlated with the parent silicate, based on the experimental data presented in this work [31]. Based on Mackenzie et al.’s research, we discovered that Sc2SiO5 has a thermal expansion coefficient between 4 × 10−6·K−1 and 6 × 10−6·K−1 in the 200 °C to 1200 °C temperature range. This is much lower than that of Yb2SiO5. As reported by Chen et al., we discovered that the thermal expansion coefficient of Eu2SiO5 is significantly larger than that of Yb2SiO5 in the temperature range of 200 °C to 1200 °C, ranging from 7 × 10−6·K−1 to 8.5 × 10−6·K−1 [26,40]. This could be the reason for the difference in material CTE seen when doping Yb2SiO5 with Sc versus Eu. Because this phenomenon is not examined in further detail in the current work, more theoretical and experimental research is required to validate it.

4. Conclusions

The crystal structure, elastic constants, Si-O bond lengths, and bond energies of YbRESiO5 (RE = Yb, Eu, Gd, Ho, Tm, Lu, Y, Sc) were investigated in this work using first-principles calculations. Using a solid-state reaction, dense polycrystalline ceramics of Yb2SiO5, YbScSiO5, and YbEuSiO5 were created. The created ceramic samples were subjected to thermal tests. The following is a summary of the primary conclusions:
(1)
According to first-principles calculations, the crystal structure and elastic properties of Yb2SiO5 are affected differently when equimolar amounts of rare earth elements RE (RE = Eu, Gd, Ho, Tm, Lu, Y, Sc) are added. In particular, doping the material with Eu, which has the largest atomic radius, and Sc, which has the smallest, seems to have opposite effects. While YbEuSiO5 has the opposite behavior, YbScSiO5 has the biggest lattice distortion, average sound velocity, Debye temperature, and Young’s modulus. It also has the least average RE-O bond length and unit cell volume. The material’s Young’s modulus is increased when Yb2SiO5 is doped with Sc, mostly as a result of an increase in cation field strength and a decrease in average cation radius.
(2)
The Si-O bond lengths and bond energies of YbRESiO5 (RE = Yb, Eu, Gd, Ho, Tm, Lu, Y, Sc) were calculated using first-principles methods. The seven aforementioned elements can be divided into two groups based on the outcomes of the simulation: In comparison to Yb2SiO5, one group (Gd, Eu, Y, and Lu) decreases the bond energy and lengthens the Si-O bond when replacing Yb, which lowers the material’s resistance to oxygen and water corrosion. When doped, the other group (Sc, Ho, and Tm) shortens the Si-O bond and lengthens the bond energy, improving the material’s resistance to oxygen and water corrosion.
(3)
According to thermal conductivity measurements, below 600 K, YbScSiO5 has somewhat less thermal diffusivity and thermal conductivity than Yb2SiO5. They are marginally greater than Yb2SiO5 beyond 600 K, but the overall difference is negligible. When Eu is added, YbEuSiO5 exhibits much reduced thermal conductivity and diffusivity in comparison to Yb2SiO5. In particular, for YbEuSiO5, the minimum thermal conductivity is 1.072 W·m−1·K−1 between 400 °C and 1200 °C, while for Yb2SiO5, it is 1.52 W·m−1·K−1. This indicates that when YbEuSiO5 is compared to Yb2SiO5, its minimum thermal conductivity decreases by 29.4%. The primary cause of YbEuSiO5’s notable reduction in thermal conductivity is the addition of Eu, which results in the introduction of oxygen vacancies.
(4)
Coefficient of thermal expansion (CTE) testing results show that YbScSiO5 has a lower CTE than Yb2SiO5, but adding Eu raises the CTE of YbEuSiO5. In particular, at 1473 K, YbScSiO5 exhibits the lowest CTE (7.56 × 10−6·K−1) compared to Yb2SiO5 (7% lower) at the same temperature. YbEuSiO5, on the other hand, has the highest CTE at 8.6 × 10−6·K−1, which is 7.5% more than Yb2SiO5. According to an analysis of the experimental data, there may be a positive correlation between the base material and the CTE of multi-component silicates.

Author Contributions

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

Funding

This research was sponsored by the National High Technology Research and Development Program of China (2023YFB3711200), the National Natural Science Foundation of China (52175136, 52130511), the Science Center for Gas Turbine Project (P2021-A-IV-002), the Shanghai Joint Innovation Program in the Field of Commercial Aviation Engines, the Shanghai Gaofeng Project for University Academic Program Development, and the Key Research and Development Projects in Anhui Province (2022a05020004).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Liao, W.; Tan, Y.; Zhu, C.; Teng, Z.; Jia, P.; Zhang, H. Synthesis, microstructures, and corrosion behaviors of multi-components rare-earth silicates. Ceram. Int. 2021, 47, 32641–32647. [Google Scholar] [CrossRef]
  2. Chen, H.; Xiang, H.; Dai, F.-Z.; Liu, J.; Zhou, Y. High entropy (Yb0.25Y0.25Lu0.25Er0.25)2SiO5 with strong anisotropy in thermal expansion. J. Mater. Sci. Technol. 2020, 36, 134–139. [Google Scholar] [CrossRef]
  3. Yang, Z.; Yang, K.; Wang, W.; Yang, T.; Fang, H.; Qiang, L.; Zhang, X.; Zhang, C. Investigation of Thermal Shock Behavior of Multilayer Thermal Barrier Coatings with Superior Erosion Resistance Prepared by Atmospheric Plasma Spraying. Coatings 2022, 12, 804. [Google Scholar] [CrossRef]
  4. More, K.L.; Tortorelli, P.F.; Ferber, M.K.; Keiser, J.R. Observations of Accelerated Silicon Carbide Recession by Oxidation at High Water-Vapor Pressures. J. Am. Ceram. Soc. 2008, 83, 211–213. [Google Scholar] [CrossRef]
  5. Opila, E.J. Oxidation and Volatilization of Silica Formers in Water Vapor. J. Am. Ceram. Soc. 2004, 86, 1238–1248. [Google Scholar] [CrossRef]
  6. Opila, E.J.; Smialek, J.L.; Robinson, R.C.; Fox, D.S.; Jacobson, N.S. SiC Recession Caused by SiO2 Scale Volatility under Combustion Conditions: II, Thermodynamics and Gaseous-Diffusion Model. J. Am. Ceram. Soc. 2004, 82, 1826–1834. [Google Scholar] [CrossRef]
  7. Tian, Z.; Zheng, L.; Wang, J.; Wan, P.; Li, J.; Wang, J. Theoretical and experimental determination of the major thermo-mechanical properties of RE2SiO5 (RE = Tb, Dy, Ho, Er, Tm, Yb, Lu, and Y) for environmental and thermal barrier coating applications. J. Eur. Ceram. Soc. 2016, 36, 189–202. [Google Scholar] [CrossRef]
  8. Fan, W.; Zou, B.; Fan, C.; Wang, Y.; Cai, X.; Tao, S.; Xu, J.; Wang, C.; Cao, X.; Zhu, L. Fabrication and oxidation resistant behavior of plasma sprayed Si/Yb2Si2O7/LaMgAl11O19 coating for Cf/SiC composites at 1673K. Ceram. Int. 2017, 43, 392–398. [Google Scholar] [CrossRef]
  9. Rohbeck, N.; Morrell, P.; Xiao, P. Degradation of ytterbium disilicate environmental barrier coatings in high temperature steam atmosphere. J. Eur. Ceram. Soc. 2019, 39, 3153–3163. [Google Scholar] [CrossRef]
  10. Raj, R. Fundamental research in structural ceramics for service near 2000 C. J. Am. Ceram. Soc. 1993, 76, 2147–2174. [Google Scholar] [CrossRef]
  11. 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]
  12. Price, J.R.; Van Roode, M.; Stala, C. Ceramic Oxide-Coated Silicon Carbide for High Temperature Corrosive Environments. Key Eng. Mater. 1992, 72–74, 71–84. [Google Scholar] [CrossRef]
  13. Cao, G.; Wang, Y.-H.; Ding, Z.-Y.; Yang, H.-L.; Liu, Z.-G.; Ouyang, J.-H.; Wang, Y.-M.; Wang, Y.-J. Tunable corrosion resistance of rare-earth monosilicate to molten calcia-magnesia-aluminosilicate glass by RE-doping strategy. Corros. Sci. 2022, 202, 110319. [Google Scholar] [CrossRef]
  14. Lee, K.N.; Miller, R.A. Oxidation Behavior of MuIlite-Coated SiC and SiC/SiC Composites under Thermal Cycling between Room Temperature and 1200–1400 °C. J. Am. Ceram. Soc. 2005, 79, 620–626. [Google Scholar] [CrossRef]
  15. 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]
  16. Cojocaru, C.V.; Lévesque, D.; Moreau, C.; Lima, R.S. Performance of thermally sprayed Si/mullite/BSAS environmental barrier coatings exposed to thermal cycling in water vapor environment. Surf. Coat. Technol. 2013, 216, 215–223. [Google Scholar] [CrossRef]
  17. Lee, K.N.; Fox, D.S.; Bansal, N.P. Rare earth silicate environmental barrier coatings for SiC/SiC composites and Si3N4 ceramics. J. Eur. Ceram. Soc. 2005, 25, 1705–1715. [Google Scholar] [CrossRef]
  18. Nasiri, N.A.; Patra, N.; Horlait, D.; Jayaseelan, D.D.; Lee, W.E. Thermal Properties of Rare-Earth Monosilicates for EBC on Si-Based Ceramic Composites. J. Am. Ceram. Soc. 2016, 99, 589–596. [Google Scholar] [CrossRef]
  19. Xu, Y.; Hu, X.; Xu, F.; Li, K. Rare earth silicate environmental barrier coatings: Present status and prospective. Ceram. Int. 2017, 43, 5847–5855. [Google Scholar] [CrossRef]
  20. Felsche, J. The Crystal Chemistry of the Rare-Earth Silicates; Springer: Berlin/Heidelberg, Germany, 1973; pp. 99–197. [Google Scholar]
  21. Ridley, M.J.; Tomko, K.Q.; Tomko, J.A.; Hoglund, E.R.; Howe, J.M.; Hopkins, P.E.; Opila, E.J. Tailoring thermal and chemical properties of a multi-component environmental barrier coating candidate (Sc0.2Nd0.2Er0.2Yb0.2Lu0.2)2Si2O7. Materialia 2022, 26, 101557. [Google Scholar] [CrossRef]
  22. Wang, Y.; Niu, Y.; Zhong, X.; Shi, M.; Mao, F.; Zhang, L.; Li, Q.; Zheng, X. Water vapor corrosion behaviors of plasma sprayed RE2SiO5 (RE = Gd, Y, Er) coatings. Corros. Sci. 2020, 167, 108529. [Google Scholar] [CrossRef]
  23. Qian, B.; Wang, Y.; Zu, J.; Xu, K.; Shang, Q.; Bai, Y. A review on multicomponent rare earth silicate environmental barrier coatings. J. Mater. Res. Technol. 2024, 29, 1231–1243. [Google Scholar] [CrossRef]
  24. Han, J.; Wang, Y.; Liu, R.; Jiang, D. Study on water vapor corrosion resistance of rare earth monosilicates RE2SiO5 (RE = Lu, Yb, Tm, Er, Ho, Dy, Y, and Sc) from first-principles calculations. Heliyon 2018, 4, e00857. [Google Scholar] [CrossRef] [PubMed]
  25. Xiang, H.; Feng, Z.; Zhou, Y. Mechanical and thermal properties of Yb2SiO5: First-principles calculations and chemical bond theory investigations. J. Mater. Res. 2014, 29, 1609–1619. [Google Scholar] [CrossRef]
  26. Chen, Z.; Tian, Z.; Zheng, L.; Ming, K.; Ren, X.; Wang, J.; Li, B. (Ho0.25Lu0.25Yb0.25Eu0.25)2SiO5 high-entropy ceramic with low thermal conductivity, tunable thermal expansion coefficient, and excellent resistance to CMAS corrosion. J. Adv. Ceram. 2022, 11, 1279–1293. [Google Scholar] [CrossRef]
  27. Segall, M.D.; Lindan, P.J.D.; Probert, M.J.; Pickard, C.J.; Hasnip, P.J.; Clark, S.J.; Payne, M.C. First-principles simulation: Ideas, illustrations and the CASTEP code. J. Phys. Condens. Matter 2002, 14, 2717–2744. [Google Scholar] [CrossRef]
  28. Ceperley, D.M.; Alder, B.J. Ground state of the electron gas by a stochastic method. Phys. Rev. Lett. 1980, 45, 566–569. [Google Scholar] [CrossRef]
  29. Pfrommer, B.G.; Côté, M.; Louie, S.G.; Cohen, M.L. Relaxation of Crystals with the Quasi-Newton Method. J. Comput. Phys. 1997, 131, 233–240. [Google Scholar] [CrossRef]
  30. Milman, V.; Warren, M.C. Elasticity of hexagonal BeO. J. Phys. Condens. Matter 2000, 13, 241–251. [Google Scholar] [CrossRef]
  31. Li, K.; Wang, W.; Yang, S.; Liu, Y.; Yang, T. First-principles calculations and thermal-mechanical experimental studies on middle-entropy rare-earth disilicates. Ceram. Int. 2024, 50, 22290–22305. [Google Scholar] [CrossRef]
  32. Leitner, J.; Chuchvalec, P.; Sedmidubský, D.; Strejc, A.; Abrman, P. Estimation of heat capacities of solid mixed oxides. Thermochim. Acta 2002, 395, 27–46. [Google Scholar] [CrossRef]
  33. Bruce, D.; Paradise, P.; Saxena, A.; Temes, S.; Clark, R.; Noe, C.; Benedict, M.; Broderick, T.; Bhate, D. A critical assessment of the Archimedes density method for thin-wall specimens in laser powder bed fusion: Measurement capability, process sensitivity and property correlation. J. Manuf. Process. 2022, 79, 185–192. [Google Scholar] [CrossRef]
  34. Leitner, J.; Voňka, P.; Sedmidubský, D.; Svoboda, P. Application of Neumann–Kopp rule for the estimation of heat capacity of mixed oxides. Thermochim. Acta 2010, 497, 7–13. [Google Scholar] [CrossRef]
  35. Charvat, F.R.; Kingery, W.D. Thermal Conductivity: XIII, Effect of Microstructure on Conductivity of Single-Phase Ceramics. J. Am. Ceram. Soc. 2006, 40, 306–315. [Google Scholar] [CrossRef]
  36. Hao, S.; Oleksak, R.P.; Doğan, Ö.N.; Gao, M.C. Computational Design of High-Entropy Rare Earth Disilicates as Next-Generation Thermal/Environmental Barrier Coatings. Acta Mater. 2023, 258, 119225. [Google Scholar] [CrossRef]
  37. Menke, Y.; Peltier-Baron, V.; Hampshire, S. Effect of rare-eath cations on properties of sialon glasses. J. Non Cryst. Solids 2000, 276, 145–150. [Google Scholar] [CrossRef]
  38. Giri, A.; Braun, J.L.; Rost, C.M.; Hopkins, P.E. On the minimum limit to thermal conductivity of multi-atom component crystalline solid solutions based on impurity mass scattering. Scr. Mater. 2017, 138, 134–138. [Google Scholar] [CrossRef]
  39. Wright, A.J.; Wang, Q.; Huang, C.; Nieto, A.; Chen, R.; Luo, J. From high-entropy ceramics to compositionally-complex ceramics: A case study of fluorite oxides. J. Eur. Ceram. Soc. 2020, 40, 2120–2129. [Google Scholar] [CrossRef]
  40. 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. (a) Conventional cell and (b) primitive cell of Yb2SiO5.
Figure 1. (a) Conventional cell and (b) primitive cell of Yb2SiO5.
Coatings 14 01035 g001
Figure 2. Geometrically optimized crystal structure of YbRESiO5 (RE = Sc, Y, Lu, Tm, Ho, Gd, Eu). (a) YbScSiO5, (b) YbYSiO5, (c) YbLuSiO5, (d) YbTmSiO5, (e) YbHoSiO5, (f) YbGdSiO5, and (g) YbEuSiO5.
Figure 2. Geometrically optimized crystal structure of YbRESiO5 (RE = Sc, Y, Lu, Tm, Ho, Gd, Eu). (a) YbScSiO5, (b) YbYSiO5, (c) YbLuSiO5, (d) YbTmSiO5, (e) YbHoSiO5, (f) YbGdSiO5, and (g) YbEuSiO5.
Coatings 14 01035 g002aCoatings 14 01035 g002b
Figure 3. (a) Si-O bond lengths, (b) Milliken population, and (c) density of Si-O bonds for YbRESiO5 (RE = Yb, Sc, Y, Lu, Tm, Ho, Gd, Eu).
Figure 3. (a) Si-O bond lengths, (b) Milliken population, and (c) density of Si-O bonds for YbRESiO5 (RE = Yb, Sc, Y, Lu, Tm, Ho, Gd, Eu).
Coatings 14 01035 g003
Figure 4. XRD patterns of YbReSiO5 (Re = Yb, Sc, Eu).
Figure 4. XRD patterns of YbReSiO5 (Re = Yb, Sc, Eu).
Coatings 14 01035 g004
Figure 5. (a) Microstructure of Yb2SiO5, (b) microstructure of YbScSiO5, and (c) microstructure of YbEuSiO5.
Figure 5. (a) Microstructure of Yb2SiO5, (b) microstructure of YbScSiO5, and (c) microstructure of YbEuSiO5.
Coatings 14 01035 g005
Figure 6. (a) Element distribution of Yb2SiO5, (b) element distribution of YbScSiO5, and (c) element distribution of YbEuSiO5.
Figure 6. (a) Element distribution of Yb2SiO5, (b) element distribution of YbScSiO5, and (c) element distribution of YbEuSiO5.
Coatings 14 01035 g006
Figure 7. (a) Thermal diffusivity of Yb2SiO5, YbScSiO5, and YbEuSiO5, (b) specific heat capacity of Yb2SiO5, YbScSiO5, and YbEuSiO5.
Figure 7. (a) Thermal diffusivity of Yb2SiO5, YbScSiO5, and YbEuSiO5, (b) specific heat capacity of Yb2SiO5, YbScSiO5, and YbEuSiO5.
Coatings 14 01035 g007
Figure 8. (a) Experimental thermal conductivity of Yb2SiO5, YbScSiO5, and YbEuSiO5, (b) theoretical thermal conductivity of Yb2SiO5, YbScSiO5, and YbEuSiO5.
Figure 8. (a) Experimental thermal conductivity of Yb2SiO5, YbScSiO5, and YbEuSiO5, (b) theoretical thermal conductivity of Yb2SiO5, YbScSiO5, and YbEuSiO5.
Coatings 14 01035 g008
Figure 9. (a) Thermal expansion coefficient (CTE) of Yb2SiO5, YbScSiO5, and YbEuSiO5, (b) ΔL/L curves of Yb2SiO5, YbScSiO5, and YbEuSiO5.
Figure 9. (a) Thermal expansion coefficient (CTE) of Yb2SiO5, YbScSiO5, and YbEuSiO5, (b) ΔL/L curves of Yb2SiO5, YbScSiO5, and YbEuSiO5.
Coatings 14 01035 g009
Table 1. The lattice parameters of Yb2SiO5 after lattice optimization and the theoretical lattice parameters.
Table 1. The lattice parameters of Yb2SiO5 after lattice optimization and the theoretical lattice parameters.
ParameterYb2SiO5
Theory [20]
Yb2SiO5
Calculation
Deviation
(%)
a ( Å ) 14.280014.19610.5021
b ( Å ) 10.280010.08521.9052
c ( Å ) 6.65306.56141.3013
γ(°)122.200122.22500.0220
Table 2. Geometrically optimized lattice parameters of YbRESiO5 (RE = Sc, Y, Lu, Tm, Ho, Gd, Eu).
Table 2. Geometrically optimized lattice parameters of YbRESiO5 (RE = Sc, Y, Lu, Tm, Ho, Gd, Eu).
Parameter a ( Å ) b ( Å ) c ( Å ) γ(°) V ( Å 3 )
YbYSiO514.274910.18176.5710122.0710809.2980
YbLuSiO514.320310.18516.5834122.2090812.4510
YbHoSiO514.174110.10466.5834122.3620794.8210
YbGdSiO514.236010.16556.5936122.2250806.6770
YbScSiO513.91889.946326.4012122.1350750.4240
YbTmSiO514.149310.11006.6152122.2680800.1510
Yb2SiO514.191610.08586.5616122.2250794.7620
YbEuSiO514.276010.16326.6205122.2340813.6240
Table 3. Lattice distortion values of different rare earth monociliate.
Table 3. Lattice distortion values of different rare earth monociliate.
CompoundBondPolyhedraBond Length (Å)Average Bond
Length (Å)
Degree of
Distortion (‰)
Yb2SiO5RE1-O[REO7]2.17365, 2.26578, 2.27687, 2.28556,
2.29827, 2.3003, 2.50653
2.300981.63300
RE2-O[REO6]2.17305, 2.20835, 2.21083, 2.22665, 2.22665, 2.247912.217500.12334
Si-O[SiO4]1.5915, 1.60382, 1.60696, 1.62281.606270.04824
YbTmSiO5RE1-O[REO7]2.1845, 2.23035, 2.24948, 2.28916,
2.31377, 2.35425, 2.4962
2.302531.68259
RE2-O[REO6]2.15346, 2.20741, 2.21105, 2.22142, 2.22615, 2.261162.213440.20894
Si-O[SiO4]1.59125, 1.60026, 1.60967, 1.622921.606020.05333
YbEuSiO5RE1-O[REO7]2.22305, 2.26695, 2.31138, 2.31634, 2.32237, 2.36489, 2.505132.330011.26024
RE2-O[REO6]2.15427, 2.22478, 2.23663, 2.25526, 2.26148, 2.293842.237710.37126
Si-O[SiO4]1.59478, 1.60092, 1.61483, 1.619831.607590.03971
YbGdSiO5RE1-O[REO7]2.2152, 2.26505, 2.29988, 2.31368,
2.31368, 2.35175, 2.49526
2.321571.27335
RE2-O[REO6]2.19257, 2.21548, 2.23843, 2.23147, 2.25323, 2.267692.233140.11984
Si-O[SiO4]1.59321, 1.60148, 1.6151, 1.620291.607520.04468
YbHoSiO5RE1-O[REO7]2.20092, 2.25368, 2.28601, 2.30648, 2.30662, 2.34039, 2.584062.325452.37837
RE2-O[REO6]2.20766, 2.22172, 2.22778, 2.23061, 2.27294, 2.141232.216990.31473
Si-O[SiO4]1.59119, 1.5994, 1.61279, 1.621941.606330.05451
YbLuSiO5RE1-O[REO7]2.17016, 2.28852, 2.28869, 2.30331, 2.30836, 2.31393, 2.528942.314561.82196
RE2-O[REO6]2.21116, 2.23925, 2.24528, 2.25512, 2.25873, 2.269282.246430.07019
Si-O[SiO4]1.59393, 1.60716, 1.60992, 1.618731.607440.03059
YbScSiO5RE1-O[REO7]2.15248, 2.21908, 2.23995, 2.2501,
2.28767, 2.29817, 2.59642
2.291983.31675
RE2-O[REO6]2.11166, 2.02535, 2.08059, 2.11853, 2.13039, 2.173942.106740.47114
Si-O[SiO4]1.59075, 1.5956, 1.60682, 1.616541.602430.05863
YbYSiO5RE1-O[REO7]2.17206, 2.28342, 2.29012, 2.29846, 2.30223, 2.30753, 2.549752.314802.07000
RE2-O[REO6]2.19854, 2.23202, 2.23482, 2.23886, 2.24511, 2.255532.234150.06170
Si-O[SiO4]1.5952, 1.60773, 1.60961, 1.619041.606200.02906
Table 4. Si-O bond information for YbRESiO5 (RE = Yb, Sc, Y, Lu, Tm, Ho, Gd, Eu).
Table 4. Si-O bond information for YbRESiO5 (RE = Yb, Sc, Y, Lu, Tm, Ho, Gd, Eu).
CompoundBond Average (Å)Population of BondDensity (Å−1)
Yb2SiO51.606270.617500.3844
YbTmSiO51.606030.618500.3851
YbEuSiO51.607590.610000.3794
YbGdSiO51.607520.612500.3810
YbHoSiO51.606200.62000.3860
YbLuSiO51.607440.605000.3764
YbScSiO51.602430.635000.3963
YbYSiO51.606300.617200.3811
Table 5. Elastic constants of YbRESiO5 (RE = Yb, Tm, Eu, Gd, Lu, Ho, Sc, Y).
Table 5. Elastic constants of YbRESiO5 (RE = Yb, Tm, Eu, Gd, Lu, Ho, Sc, Y).
Compound v m
(m/s)
ɵ D
(K)
B
(Gpa)
G
(Gpa)
E
(Gpa)
v
Yb2SiO53116.0476400.2373116.961859.8568153.40190.2696
YbTmSiO53115.4777395.7621105.478657.3296145.60840.2699
YbEuSiO53027.7198392.4537102.446657.1057134.47320.2677
YbGdSiO52875.6370367.762192.029449.4669125.82170.2721
YbLuSiO53136.1570403.6193106.271960.1105151.72480.2621
YbHoSiO53047.4524394.218997.266355.4342139.75320.2605
YbScSiO53669.7119481.4030126.260563.0056162.06000.2861
YbYSiO53282.3078419.456995.067953.6648145.49850.2625
Table 6. CFS of rare earth cations in YbRESiO5 (RE = Yb, Sc, Y, Lu, Tm, Ho, Gd, Eu).
Table 6. CFS of rare earth cations in YbRESiO5 (RE = Yb, Sc, Y, Lu, Tm, Ho, Gd, Eu).
CompoundAverage Ionic RadiusCFS
Yb2SiO50.864.056
YbTmSiO50.8654.009
YbEuSiO50.9053.662
YbGdSiO50.93.370
YbHoSiO50.883.387
YbLuSiO50.8554.104
YbScSiO50.8354.303
YbYSiO50.8753.918
Table 7. Theoretical density ρ (RE2SiO5), mass fraction ω (RE2SiO5), ω (SiO2), experimental density ρ (Exp.), bulk density (d), and porosity ( ψ ) of Yb2SiO5, YbScSiO5, and YbEuSiO5.
Table 7. Theoretical density ρ (RE2SiO5), mass fraction ω (RE2SiO5), ω (SiO2), experimental density ρ (Exp.), bulk density (d), and porosity ( ψ ) of Yb2SiO5, YbScSiO5, and YbEuSiO5.
Compound ρ (Re2SiO5)
(g·cm−3)
ω (Re2SiO5)
(wt.%)
ω (SiO2)
(wt.%)
ρ ( E x p . )
(g·cm−3)
d
(%)
ψ
(%)
Yb2SiO57.2886.7713.237.1297.802.20
YbEuSiO55.7681.5718.435.4995.314.69
YbScSiO56.6986.1313.876.4095.664.34
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

Yang, S.; Wang, T.; Li, K.; Wang, W.; Liu, Y.; Yang, T. Investigation on the Thermal–Mechanical Properties of YbRESiO5 (RE = Yb, Eu, Gd, Ho, Tm, Lu, Y, Sc): First-Principles Calculations and Thermal Performance Experiments. Coatings 2024, 14, 1035. https://doi.org/10.3390/coatings14081035

AMA Style

Yang S, Wang T, Li K, Wang W, Liu Y, Yang T. Investigation on the Thermal–Mechanical Properties of YbRESiO5 (RE = Yb, Eu, Gd, Ho, Tm, Lu, Y, Sc): First-Principles Calculations and Thermal Performance Experiments. Coatings. 2024; 14(8):1035. https://doi.org/10.3390/coatings14081035

Chicago/Turabian Style

Yang, Shilong, Tianying Wang, Kaibin Li, Weize Wang, Yangguang Liu, and Ting Yang. 2024. "Investigation on the Thermal–Mechanical Properties of YbRESiO5 (RE = Yb, Eu, Gd, Ho, Tm, Lu, Y, Sc): First-Principles Calculations and Thermal Performance Experiments" Coatings 14, no. 8: 1035. https://doi.org/10.3390/coatings14081035

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