You are on page 1of 10

Ceramics International 43 (2017) 3346–3355

Contents lists available at ScienceDirect

Ceramics International
journal homepage: www.elsevier.com/locate/ceramint

First-principles study of structural, mechanical, and thermodynamic MARK


properties of cubic Y2O3 under high pressure

Xian Zhanga, , Wenhua Guia, Qingfeng Zengb
a
School of Advanced Materials and Nanotechnology, Xidian University, Xi’an 710071, PR China
b
International Center for Materials Descovery, School of Materials Science and Engineering, Northwestern Polytechnical University, Xi’an 710072, Shaanxi,
PR China

A R T I C L E I N F O A BS T RAC T

Keywords: The structural, mechanical, and thermodynamic properties of cubic Y2O3 crystals at different hydrostatic
Y2O3 pressures and temperatures are systematically investigated based on density functional theory within the
First principles generalized gradient approximation. The calculated ground state properties, such as equilibrium lattice
Mechanical properties parameter a0, the bulk modulus B0, and its pressure derivative B0′ are in favorable agreement with the
Thermodynamic properties
experimental and available theoretical values. The pressure dependence of a/a0 and V/V0 are also investigated.
Quasi-harmonic Debye model
Furthermore, the elastic constants Cij, bulk modulus B, shear modulus G, Young's modulus E, the ductile or
brittle (B/G), Vickers hardness Hv, isotropic wave velocities and sound velocities are calculated in detail in a
pressure range from 0 to 14 GPa. It was found that the Debye temperature decreases monotonically with an
increase in pressure, the calculated elastic anisotropic factors indicate that Y2O3 has low anisotropy at zero
pressure, and that its elastic anisotropy increases as the pressure increases. Finally, the thermodynamic
properties of Y2O3, such as the dependence of the heat capacities CV and CP, the thermal expansion coefficient α,
the isothermal bulk modulus, and the Grüneisen parameter γ on temperature and pressure, are discussed from
0 to 2000 K and from 0 to 14 GPa, respectively, applying the non-empirical Debye model in the quasi-harmonic
approximation.

1. Introduction six oxygen atoms in the form of a perfect (distorted) octahedron


(namly, Y1O6 and Y2O6 polyhedral units), and O atom is surrounded
The binary sesquioxides R2O3 are promising materials in the by four Y atoms in the form of a distorted tetrahedron. This compound
science and technology of many applications such as high-temperature exhibit three structural polymorphisms: cubic, hexagonal corundum-
materials, electronics, catalysts, photonics, nuclear materials, chemi- type (space group R3c, No. 167) and monoclinic Rh2O3-type (space
cals, biomaterials, etc. Therefore, a better understanding of binary group Pbca, No. 61) structures, commonly known as C-Y2O3, H-Y2O3,
systems can help in the design of new oxide materials with desirable and M-Y2O3, respectively. The cubic form of Y2O3 being stable at room
properties and would stimulate further advances in the field. In recent temperature and ambient pressure. The structural stability of yttria
years, the most widely studied materials employed for optoelectronic with pressure and temperature makes it useful in many industrial
devices that accomplish these high-pressure and high-temperature (HP applications and technological areas. Due to its excellent chemical
and HT) properties are several sesquioxide like: Y2O3, In2O3, Sb2O3, stability [11,12], low emission and small absorption coefficient in the
Sc2O3, Bi2O3, and Ti2O3 [1–10]. Several cubic-bixbyite structure under IR region at high temperature, outstanding refractoriness (T≈2410 ℃),
HP-HT conditions showed a transformation to a very dense and optical transparency over a broad spectral range (0.2–0.8 µm), large
unusual oxide structure of the α-Gd2S3-type with Pnma space group. band gap (5.5–6.0 eV), high hardness, low phonon energy (430 cm–1),
Yttria (Y2O3), with a cubic bixbyite structure, is an important rare- high refraction index (1.92–1.99), Y2O3 transparent ceramics are used
earth sesquioxide. The cubic structure of Y2O3 with space group of Ia3- as excellent IR-windows materials and used in phosphor matrixes,
(Th7 ) (No. 206) contains a total of 16 formula units (80 atoms) in unit scintillators, missile domes, waveguides and anti-wear anti-reflection
cell. The unit cell contains two nonequivalent yttrium Wyckoff sites, Y1 coatings for mid-infrared lenses, solid oxide fuel cells, high-tempera-
atom located at the 8b sitesand Y2 at the 24d sites, and one type of O ture protective coatings [13–16]. Recently, Y2O3 has also received
atom located at 4e Wyckoff sites, respectively. Y1 (Y2) is surrounded by attention as a promising candidate for replacing silicon dioxide (SiO2)


Corresponding author.
E-mail address: zhxian@mail.xidian.edu.cn (X. Zhang).

http://dx.doi.org/10.1016/j.ceramint.2016.11.176
Received 4 October 2016; Received in revised form 15 November 2016; Accepted 24 November 2016
Available online 27 November 2016
0272-8842/ © 2016 Elsevier Ltd and Techna Group S.r.l. All rights reserved.
X. Zhang et al. Ceramics International 43 (2017) 3346–3355

as a gate dielectric material in metal-oxide-semiconductor (MOS) ionic Hellmann–Feynman force of within 0.01 eV/Å, a maximum
transistors and put it down to its high dielectric contant (ε=14–19) stress of within 0.02 GPa, and a maximum ionic displacement of
[17–20]. Moreover, Y2O3 is also an important oxide-based phosphor within 5.0×10–4 Å.
material and is used as a host material in rare-earth-doped lasers [21]. The investigations on the thermodynamic properties of Y2O3 under
In recent years, the significance of yttria has attracted extensive high pressure and temperature are of great practical significance in
experimental and theoretical studies on the electronic, mechanical and engineering applications. In this paper, to investigate the thermody-
opticalproperties [22–25]. The measurement of the elastic moduli of namic properties of Y2O3, we have used the quasi-harmonic Debye
yttria from 300 to 1473 K have been reported by James et al. [26], and model, as implemented in the Gibbs program [49]. The non-equili-
found it to be stable up to about 2705 K. Ab-initio molecular dynamics brium Gibbs function G*(V; P, T) can be written as follows:
[27] study of the structure of Y2O3 up to 5000 K indicated melting at
G*(V ; P, T ) = E (V ) + pV + AVib [Θ(V ); T ] (1)
3150 K. Raman spectra of C-type bixbyite structured R2O3 (R=Yb, Sc,
Er, Y, Ho, Gd, and Sm) are studied by Abrashev et al. [28]. The stability where E(V) represents the total energy per unit cell, and can be
of the crystalline phases of Y2O3 at high pressure has also been studied obtained by electronic structure calculations, P is the constant hydro-
by calculating [29] enthalpy in various phases using the ab-initio static pressure condition, Θ (V) is the Debye temperature, and AVib is
method. Phase transitions and thermodynamic properties of yttria, the vibrational Helmholtz free energy, which can be written using the
Y2O3 has been studied by Bose et al. [30]. Ramzan et al. [31] have been quasi-harmonic approximation [50] and the Debye model of phonon
extensively researched the electronic, mechanical and optical proper- density of states as [51],
ties of Y2O3 with the GGA-PBE approximation the HSE06 hybrid-
⎡9 Θ ⎛ Θ ⎞⎤
density functional. In 1990, Ching et al. [32] repoted a self-consistent AVib (Θ; T ) = nkT ⎢ + 3 ln(1 − e−Θ / T ) − D⎜ ⎟⎥
band structure and optical properties calculation of Y2O3, using the ⎣8 T ⎝ T ⎠⎦ (2)
orthogonalized linear combination of atomic orbitals (OLCAO) method
where n is the number of atoms per formula unit, D (Θ/T) is the Debye
in the local-density approximation (LDA). Ahuja et al. [33] carried out
integral. For an isotropic solid, Θ can be taken as [51],
a study on the electronic and optical properties of ceramic Sc2O3 and
Y2O3 using compton spectroscopy and first principles calculations. 3 y x3
Badehian et al. [34] studied the elastic, structural, electronic, thermo-
D (y ) =
y3
∫0 x
e −1
dx
(3)
dynamic, and optical properties of cubic yttria ceramic within the
density functional theory, implemented in the WIEN2k (2011) code ℏ 1/3 BS
Θ= (6π 2V 1/2n ) f (σ )
[35]. Zheng et al. [36] investigated the native point defects in bulk kB M (4)
yttria through the first- principles calculation using pseudopotential
where the Poisson ratio σ is taken to be 0.25, f (σ) is given by Francisco
plane waves approach. Swamy et al. [37] analyzed the thermodynamic
et al. [52], and BS denotes the adiabatic bulk modulus, which is
properties and phase diagram of Y2O3. Ou et al. [38] studied the
approximated by the static compressibility and expressed by [52],
vacancy formation energies and electronic structures of Y2O3 by first
principles. Mueller et al. [39] used the linear muffin-tin orbital (LMTO) ⎛ d2E (V ) ⎞
method within the atomic sphere approximation for the calculation of BS ≅ B(V ) = V ⎜ ⎟
⎝ dV 2 ⎠ (5)
the electronic structure of yttria. Mudavakkat et al. [40] evaluated the
structure, morphology, and optical properties of nanocrystalline yt- ⎧ ⎡ −1⎫1/3
⎪ ⎛ 21 + σ ⎞3/2 ⎛ 11 + σ ⎞3/2⎤ ⎪
trium oxide. Manning et al. [41] measured the elastic properties of ⎢
f (σ ) = ⎨3 2⎜ ⎟ +⎜ ⎟ ⎥ ⎬
polycrystalline yttrium oxide, holmium oxide, and erbium oxide at high ⎪ ⎢⎣ ⎝ 31 − σ ⎠ ⎝ 31 − σ ⎠ ⎥⎦ ⎪
⎩ ⎭ (6)
temperature. Also, some studies have reported [42] that bulk modulus
values deviate too much from experimental values (around 150 GPa Therefore, the non-equilibrium Gibbs function G*(V; P, T) as a
[26,43]). While the bulk modulus and elastic moduli are valuable for function of V, P, and T can be minimized with respect to volume V by
designing such optical components. taking,
In the present study, we have investigated the effect of tstructural, ⎛ ∂G*(V ; P, T ) ⎞
mechanical, and thermodynamic properties of cubic Y2O3 at high ⎜ ⎟ =0
pressures (0–14 GPa) and temperatures (0–2000 K), by using first- ⎝ ∂V ⎠ P, T (7)
principles calculations combined with the quasi-harmonic Debye
Other thermodynamic properties can be obtained by solving Eq. (7)
model.
and the equation of state (EOS), and the isothermal bulk modulus BT,
the heat capacity at constant volume CV, and the heat capacity at
2. Computational details and theoretical method
constant pressure CP, and the thermal expansion coefficient α are given
by [53],
In the present work, the Cambridge Serial Total Energy Package
(CASTEP) [44] based on density functional theory (DFT) was used, ⎛ ∂p ⎞
BT (p , T ) = − V ⎜ ⎟
under ambient pressure conditions. Furthermore, the norm-conserving ⎝ ∂V ⎠ (8)
pseudo-potential [45] was adopted to describe the interactions of
valence electrons (O 2s22p4 and Y 4s24p65s24d1) with ion cores. The ⎡ ⎛Θ⎞ 3Θ / T ⎤
CV = 3nkB⎢4D⎜ ⎟ − −Θ / T ⎥
exchange-correlation energy was evaluated using GGA-PBE [46]. A ⎣ ⎝T ⎠ e − 1⎦ (9)
quasi-Newton minimization algorithm using the Broyden–Fletcher–
Goldfarb–Shanno (BFGS) [47] scheme was used, which provides a very Cp = Cv(1 + αγT ) (10)
efficient and robust way to explore the optimum crystal structure with
γCv
minimum energy. The k-point meshes for Brillouin zone sampling are α=
BT V (11)
constructed using the Monkhorst–Pack scheme [48]. A plane wave cut-
off energy of 600 eV and a 4×4×4 k-point mesh are found to be where γ is the Grüneisen parameter, which is defined as,
sufficient to converge the total energy to better than 1 meV/atom. This
d ln Θ(V )
assured a high level of convergence with respect to all parameters: a γ=−
d ln V (12)
self-consistent computational convergence precision of within 5.0×10–
7
eV/atom, a total energy variation of 5.0×10–6 eV/atom, a maximum

3347
X. Zhang et al. Ceramics International 43 (2017) 3346–3355

GGA-PBE calculations are underestimated to experimental data


(10.604 Å) [55] by about 0.41% and 0.09%, and GGA+U calculations
are overestimated by about 0.745% for lattice parameters a, respec-
tively. Bulk modulus is a fundamental physical property of solids and it
can also be used as a measure of the average bond strengths of atoms of
the given crystals. Therefore, its value is an important indicator for the
several purposes. It also can be seen that our result of bulk modulus B0
and its pressure derivative B0′, compressibility K are well consistent
with the experimental data [26,56–60] and theoretical data
Fig. 1. The structure of cubic Y2O3. (a). A conventional cell containing 80 atoms, (b) and [30,31,40,60]. As can be seen from Table 1, there are very a few
(c). two different sites for Y atoms, donoted as Y1 and Y2, respectively. (d) the site of an O differences between structural parameters with coulomb potential (U)
atom. The red spheres represent oxygen atoms and the gray spheres represent yttrium and without U. Thus we only report calculation results that did not
atoms. (For interpretation of the references to color in this figure legend, the reader is consider coulomb potential Agoston et al. [4] and Gomis et al. [9]
referred to the web version of this article.)
measured the bulk modulus B0, its pressure derivative B0′ within
density functional theory calculations and obtained value
3. Results and discussion B0=178.87 GPa, B0′=5.15 for C-In2O3 and B0=147 GPa, B0′=5 for C-
Ti2O3, respectively. Form the results obtained for bulk modulus, it
3.1. Electronic and structural properties observed that our results (see Table 1) are in good agreement with
theoretical date [4,9]. This is because the C-Ti2O3, C-In2O3 and C-Ti2O3
As shown in Fig. 1, Y2O3 has a cubic bixbyite structure with the are very similar in structure and properties. It is also indicated that the
space group Ia3 (No. 206) at ambient temperature and pressure calculated method performed in our work is reasonable.
conditions, which is similar to In2O3. The cubic unit cell contains 80 In order to show how the structural parameters under pressure in
atoms (two primitive cells of Y2O3) with two nonequivalent yttrium this compound behave, the equilibrium ratios of normalized lattice
cation sites, Y1 and Y2, occupying the 8b (1/4, 1/4, 1/4) and 24d (– parameters a/a0 and normalized volume V/V0 as functions of pressure
0.029, 0, 0.25) crystallographic positions, respectively, and one type of are plotted in Fig. 3, where a0 and V0 are the equilibrium structure
O atom at the 48e (0.3910, 0.1510, 0.3804) site. Y1 (Y2) is surrounded parameters at T=0 K and P=0 GPa, respectively.
by six oxygen atoms in the form of a perfect (distorted) octahedron By fitting the calculated data to a third-order polynomial, we
(Y1O6 and Y2O6 polyhedral units), and the O atom is surrounded by obtained their relationships at the temperature T=0 K,
four Y atoms in the form of a distorted tetrahedron. We have used the
experimental data for initial input with a lattice constant of 10.604 Å, a / a 0 = 1.00073–1.85 × 10–3P + 4.659 × 10–5P 2 –1.81187 × 10–7P 3 (14)
and optimized the ion positions. –3 –4 2 –7 3
To determine the ground state structure of Y2O3, several different V / V0 = 1.00262–5.57 × 10 P + 1.1875 × 10 P –2.88194 × 10 P
lattice parameter a are set to calculate the total energy E corresponding (15)
primitive cell volume V. The plot of calculated total energies versus We note that with the increase in pressure, the ratios of lattice
volume for Y2O3 is given in Fig. 2. By fitting the calculated energy- parameter a/a0 decrease and the volume V/V0 shrinks constantly,
volume (E-V) to the third-order Birch-Murnaghan equation of state indicating that the cell is being compressed. The reason of this change
(EOS) [54], in which the energy-volume relationship is expanded as, is that when pressure increases, atoms in the interlayers become closer,
⎡ B ′ −1 ⎤ and their interactions become stronger. The effects of temperature and
B0V0 ⎢ 1 ⎛ V0 ⎞ 0 V B0′ ⎥ pressure on the cell are opposite.
E (V ) = E 0 + ′ ⎢
⎜ ⎟ + 0 −
B0 ⎣ (B0−1) ⎝ V ⎠

V (B0−1) ⎥⎦

Generally, electronic properties of solid materials are studied by
(13)
calculating the energy band structure and density of states. In our
we can determine the equilibrium structure parameters, the bulk work, the energy band gap value has been calculated to be 4.296 eV,
modulus B0 and its pressure derivative B0′. All the equilibrium 5.748 eV, and 4.133 eV using GGA-PBE, GGA+U, and LDA, respec-
parameters using LDA, GGA-PBE, and GGA+U approximations are tively (listed in Table 2, together with other theoretical results and
given in Table 1, together with other theoretical data and available available experimental data). The calculated GGA+U band gap is larger
experimental data for comparison. It may be noted that the LDA and than that of obtained from GGA-PBE and LDA, and is in good
agreement with the value 5.5–6 eV obtained through experiment
-409.675 GGA [20,39,61,62]. The results show that the overall band profiles improve
-409.700 BM EOS Fit the band gap energy with respect to the other theatrical results
-409.725 [27,30,41] and they are in good agreement with the experimental
results. Our calculated band gaps with LDA and GGA-PBE are under-
-409.750
estimated in comparison with the experimental data. The reason of this
Energy(Hartree)

-407.625
-407.640 GGA+U underestimation is that the Kohn–Sham one particle equation does not
-407.655 BM EOS Fit provide the quasiparticle excitation energies. To our knowledge, the
-407.670 modified GGA+U formalism better band gap than the GGA because it
-407.685 has orbital independent exchange-correlation potential which depends
-407.700 only on semilocal quantities.
-409.250
LDA
-409.275
BM EOS Fit 3.2. Mechanical properties and anisotropy
-409.300
-409.325
-409.350 The elastic constants determine the response of a crystal to external
-409.375 force, and they have a significant role in our understanding of the
560 570 580 590 600 610 620 630 640 650 660
3 mechanical behaviors of materials. These properties play an important
Volume(bohr ) part in providing valuable information about the binding characteristic
Fig. 2. Energy as a function of the primitive unit cell volume of cubic Y2O3 using the between adjacent atomic planes, anisotropic character of binding, and
GGA-PBE, GGA+U, and LDA approximation. structural stability. Hence, to study the stability of C-Y2O3, we have

3348
X. Zhang et al. Ceramics International 43 (2017) 3346–3355

Table 1
The equilibrium lattice constant (a0), equilibrium volume (V0), bulk modulus (B0), the pressure derivative of bulk modulus (B0′ ), compressibility coefficient K (10−3 GPa) for Y2O3
obtained in our work compared with other studies.

Method a0 (Å) V0 (Å) B0 (GPa) B0′ K (GPa)

This work GGA-PBE 10.594 594.43 165.2 5.003 6.05


GGA+U 10.683 609.65 155.6 4.849 6.43
LDA 10.560 588.73 145.4 5.010 6.88
Other studies Potential model 10.61a 597.2a 138a
Ab-initio 10.63a 600.0a 170a
GGA-PBE 10.71b 612.0b 153.86b
HSE06 10.60b 601.2b 150.64b
GGA-PBE 10.630c 136.85c 4.4230c 7.12c
GGA-WC 10.658c 140.81c 4.7526c 7.1c
LDA 10.527c 140.36c 4.5554c 7.3c
Brillouin spectroscopy 10.392d 180–183d 4.01d, 7.65d 6.92d 5.494d
Experiment 10.6018e 14.904e 149.5 ± 1.0e 6.6e
10.604f 144h, 146.2i 6.9h, 6.8i
10.603g 148.9j 6.7j

a
Ref. [30].
b
Ref. [31].
c
Ref. [34].
d
Ref. [42].
e
Ref. [26].
f
Ref. [55].
g
Ref. [56].
h
Ref. [57].
i
Ref. [58].
j
Ref. [59].

1.00 of the lattice primitive cell volume V. The energy of a strained system is
a/a0 expressed as [63]
V/V0 ⎛ ⎞
1
E (V , δ ) = E (V0, 0) + V0⎜⎜∑ τiξiδi + ∑ Cijδiξiδj ⎟⎟
Structural parameter ratios

0.98 ⎝ i 2 ij ⎠ (16)

where E (V0, 0) is the energy of the unstrained system with the


equilibrium volume V0, and δ is the small-strain corresponding
0.96 volume, τi is the stress tensor, ξi is a Voigt index factor, and Cijis the
elastic tensor. For cubic crystals, there are only three independent
elastic constants, namely, C11, C12, and C44. And two shear moduli, c
and c′, where c=C44, c′=(C11−C12)/2, corresponding to shear along the
0.94 (100) and (110) planes, respectively, which can be used to estimate
distortion in a certain plane of Y2O3 when it is expanded or com-
0 2 4 6 8 10 12 14 pressed. The conditions for the mechanical stability of crystal struc-
Pressure (GPa)
tures are given by [64]
Fig. 3. The normalized parameters a/a0 and V/V0 as functions of pressure.
k
Cαα > 0, j
C11 > | k
C12|, j
C11 + 2 k
C12 > 0 (17)
calculated the elastic constants at normal a pressure using CASTEP WherekCαα = Cαα − p (α = 1, 4) and kC12 = C12 + p .
package. Once three elastic constants C11, C12, and C44 have been calculated,
The elastic constants are defined by means of a Taylor expansion of bulk modulus B, and isotropic shear modulus G can be calculated
the total energy E(V, δ), for the system with respect to a small strain δ simply using the subsequent clear expressions:

Table 2
Calculated energy band gap for cubic Y2O3 using various approximations.

Present work Other calculations Experiment

GGA-PBE GGA+U LDA GGA-PBE mBJ-GGA LDA LMTO HSE06

a b b d
Band gap/eV 4.296 5.748 4.13 4.3 5.7 4.5 4.5 6.0e 5.5f, 6.0g
4.8b 4.5c 5.8h

a
Ref. [27].
b
Ref. [30].
c
Ref. [40].
d
Ref. [60].
e
Ref. [31].
f
Ref. [20].
g
Ref. [61].
h
Ref. [62].

3349
X. Zhang et al. Ceramics International 43 (2017) 3346–3355

Table 3
Values of the elastic constant Cij (GPa), bulk modulus B (GPa), shear modulus G (GPa), Young’s modulus E (GPa), and Possion’s ratio ν for various approximations for cubic Y2O3.

C11 C44 C12 B GV GR GH E ν

This work GGA-PBE 243.3 85.9 126.2 165.2 74.9 72.4 73.7 192.4 0.306
GGA+U 233.1 79.5 116.8 155.6 70.9 69.3 70.2 182.8 0.304
LDA 212.3 88.0 115.9 148.0 72.1 66.2 69.2 179.4 0.298
Theoretical [30] 206a 233b 59a 103a
72b 139b
Theoretical [31] 219.3 72.79 120.68 62.43 165 0.32
Experiment [26] 223.7 ± 0.6 74.6 ± 0.7 112.4 ± 1.1 149.5 66.3 ± 0.8 173 ± 2 0.307 ± 0.003

GV, GR, and GH, represent the Voigt, Reuss, and average polycrystalline shear moduli, respectively.
a
Potential model.
b
ab initio.

GV + GR it also indicates that Y2O3 becomes increasingly difficult to compress as


G=
2 (18) the pressure increases. The variation in the modulus (B, G, E) in
Fig. 4(b), Young's modulus and the shear modulus exhibit a slight
BV + BR
B= decrease as the pressure increases. Table 4 also shows that the
2 (19)
hydrostatic pressure has a remarkable effect on the microhardness,
As is known to all, the arithmetic average of the Voigt (BV, GV) and which decreases with increasing hydrostatic pressure. The similar
the Reuss (BR, GR) can be commonly used to estimate the bulk modulus elastic properties of Y2O3 using PBE-GGA approximation at 0–7 GPa
B and shear modulus G, which can be evaluated by the Voigt–Reuss– have also been calculated by Badehian et al. [34], and obtained the
Hill scheme [65]
C11 + 2C12 Table 4
BV = BR = Values of the elastic constant Cij (GPa), bulk modulus B (GPa), shear modulus G (GPa),
3 (20)
B/G, Young’s modulus E (GPa), universal elastic anisotropy index AU, Possion’s ratio ν,
Cauchy pressure CP (GPa), compressibility coefficient K (10–3 GPa–1), and Vickers
C11 − C12 + 3C44
GV = hardness Hv (GPa), for Y2O3 with GGA-PBE approroximation.
5 (21)
P (GPa)
5C44(C11 − C12 )
GR =
4C44 + 3(C11 − C12 ) (22) 0 2 4 6 8 10 12 14

Here, GV is Voigt's shear modulus corresponding to the upper C11 (GPa) 243.3 249.1 258.7 269.2 278.3 285.7 292.2 298.2
bound of G values, and GR is Reuss's shear modulus for cubic crystals C44 (GPa) 85.9 84.6 85.5 85.6 86.4 85.0 84.4 83.5
C12 (GPa) 126.2 137.0 147.0 156.2 169.1 179.2 189.4 198.0
corresponding to the lower bound values.
B (GPa) 165.2 174.4 184.2 193.8 205.5 214.7 223.6 231.4
Young's modulus E, Poisson ratio v, and compressibility coefficient G (GPa) 73.7 71.7 72.0 72.5 71.9 70.5 69.2 68.0
K are important fundamental parameter closely connected to many GV (GPa) 74.9 73.2 73.6 74.0 73.7 72.3 71.2 70.1
physical properties such as internal strain, thermoelastic stress, bond- GR (GPa) 72.4 70.3 70.5 71.0 70.1 68.6 67.2 65.9
ing forces, hardness, sound velocity, fracture toughness. The expres- B/G 2.24 2.43 2.56 2.67 2.86 3.05 3.23 3.40
E (GPa) 192.4 189.2 191.2 193.4 193.1 190.6 188.2 185.8
sions for the Young's modulus, Poisson's ratio and compressibility AU 0.179 0.206 0.221 0.210 0.258 0.268 0.300 0.320
coefficient are given by ν 0.306 0.319 0.327 0.334 0.343 0.352 0.36 0.366
CP (GPa) 40.3 52.5 61.6 70.5 82.7 94.2 105.0 114.5
9BG
E= Hv (GPa) 7.705 6.898 6.536 6.24 5.748 5.272 4.866 4.536
3B + G (23) K (10–3 GPa–1) 6.052 5.734 5.428 5.159 4.866 4.657 4.471 4.321

3B − 2G
ν=
2(3B + G ) (24)
320 320
K=
1
=
3 C11 C12 B G
B C11 + 2C12 (25) C44 280
E
280 Cauchy pressure
Our theoretical elastic constants C11, C12, C44, bulk modulus B
(GPa), Young's modulus E (GPa), shear modulus G (GPa), and 240
Elastic constants (GPa)

Poisson's ratios ν for Y2O3 at P=0 GPa and T=0 K are summarized in 240
Modulus (GPa)

Table 3, together with the available theoretical [30,31] or experiment 200


[26] data. The obtained elastic constant, bulk modulus, shear modulus, 200
Young's modulus, Poisson's ratio, and other important physical quan- 160
tities at 0–14 GPa using GGA-PBE approroximation are listed in
Table 4. When the pressure is less than 14 GPa, the calculated elastic 160
120
constants listed in Table 4, satisfied the above stability certeria [see Eq.
(17)], indicating that the crystal structure of Y2O3 remains elastically 120 80
stable, i.e., it is still cubic.
The variation in the elastic constants C11, C12, C44, modulus (B, G,
80 40
E) and cauchy pressure CP of Y2O3 with hydrostatic pressure are
0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14
plotted in Fig. 4. It is noted that C11 and C12 increase linearly with an
increase in pressure, while the varying amplitude of C44 is moderate Pressure (GPa) Pressure (GPa)
compared with C11 and C12 (Fig. 4a). It is clear that the C11 and C12 are Fig. 4. (a) The elastic constant as a function of pressure of Y2O3 at T=0 K. (b) Moduli
more sensitive to the change of pressure than the C44, at the same time, versus pressure.

3350
X. Zhang et al. Ceramics International 43 (2017) 3346–3355

elastic constants satisfy the cubic stability condition and showed that
Y2O3 is elastically stable. It can be seen that the gareement between our
results and previous theoretical date [34] is good. Pettifor [66] reported
that the angular character of atomic bonding in metals and compounds
is depicted through the Cauchy pressure, which is defined as
CP=C12−C44. The negative (positive) values show the brittle (ductile)
nature of the compound. Table 4 shows that the Cauchy pressure CP is
40.3 GPa at 0 GPa. According to above empirical laws, the Y2O3
behaves in a ductile manner under pressure. Furthermore, in this
work, the effect of the hydrostatic pressure on the Cauchy pressure was
also investigated, and the results are plotted in Fig. 4(b). Evidently, an
increase in the hydrostatic pressure causes an increase in the Cauchy
pressure, indicating an increase in the ductility of crystalline Y2O3.
Elastic anisotropy or isotropy is an important physical property of
materials, and has a vital role in technological and industrial applica-
tions. Ranganathan et al. [67] introduced the concept of the universal
anisotropy index: for cubic Voigt and Reuss approximations, the
anisotropy index can be simplified to AU=5 (GV/GR−1), and for an
isotropic material, AU is equal to zero, and deviation of AU from zero
indicates the presence of elastic anisotropy. The dependence of the
calculated anisotropy factor AU on pressure for Y2O3 is shown Table 4.
It is obvious that Y2O3 has low anisotropy at zero pressure, and the
anisotropy of Y2O3 becomes stronger as the pressure increases.
A three-dimensional (3D) surface representation of the elastic
anisotropy of cubic Y2O3, expressed by the variation of Young's
modulus with crystal direction, is further considered. The directional
dependence of Young's modulus can be described as [68],

1 ⎛ 1 ⎞
= S11 − 2⎜S11 − S12 − S44⎟(l12l22 + l12l32 + l22l32 )
E ⎝ 2 ⎠ (26)

where l1=sinθ cosφ, l2=sinθ sinφ, and l3=cosθ and represent the
direction cosines with respect to the x, y and z directions of the lattice,
and S11, S12, and S44 are the elastic compliance coefficients.
The 3D directional dependence of an isotropic system exhibits a
spherical shape, while the degree of deviation from spherical reflects
the extent of anisotropy. The projected configurations of the directional
Young's moduli on the xy, yz, and xz planes revealed significant
similarity, denoting that the strength of Y2O3 would not change with
the crystallographic plane. As an example, the 3D representation and
corresponding 2D projections of Young's modulus for Y2O3 compounds
are presented in Fig. 5a–d at four different hydrostatic pressures (i.e.,
0, 4, 8, and 14 GPa). It can be seen that the projected configuration of
the directional Young's modulus would vary from a square to quasi-
square shape as the hydrostatic pressure rises from 0 to 14 GPa. This
suggests that a Y2O3 single crystal tends to become slightly anisotropic
at higher hydrostatic pressures. This result is consistent with that of the
pressure- dependent anisotropy index AU values listed in Table 4.
The angular character of atomic bonding is related to material
characteristics such as brittleness or ductility. Ductility and brittleness
are both crucial features in the manufacture of materials, and have a
close relationship with B and G. According to the Pugh criterion [69],
the critical threshold value for distinguishing the physical properties of
materials is around 1.75. If B/G > 1.75, a polycrystal possesses a Fig. 5. Directional dependence of Young’s modulus for Y2O3 (GPa) and its projection
onto the x–y, y–z, and x–z planes associated with four different hydrostatic pressures:
ductile character, otherwise the material exhibits brittle behavior.
(a) 0 GPa, (b) 4 GPa, (c) 8 GPa, and (ed) 14 GPa.
The computed B/G ratio for Y2O3 was about 2.24, as listed in
Table 4, indicating that crystalline Y2O3 is a ductile material.
increases slightly with pressure, ranging from 0.306 to 0.366 under
Poisson's ratio is generally used to describe the stability of a crystal
pressure between 0 and 14 GPa, The results indicates that the ionic
against shear and provides more information about the characteristics
contribution to interatomic bonding and the interatomic forces are
of the bonding forces than elastic constants [51]. The values of the
mainly central forces for Y2O3.
Poisson ratio ν for covalent materials is small (ν=0.1), whereas for ionic
As a measure to resist penetration, deformation, abrasion, and
materials a typical value of ν is 0.25. In this study, the calculated
wear, hardness plays an important role in industrial applications.
poisson ratio ν is also given in Table 4, the value of ν for Y2O3 is about
Given to the above investigation, Y2O3 quite possibly possesses
0.306 with GGA-PBE method at P=0 GPa and T=0 K, so a considerable
excellent hardness. Recently, many studies have shown that the elastic
ionic contribution should be assumed for this compound. As shown in
modulus G and B both influence Hv for many materials. In this work,
the Table 4, the result shows the Poisson ratio for cubic phase Y2O3

3351
X. Zhang et al. Ceramics International 43 (2017) 3346–3355

Table 5
Calculated density ρ, shear-wave velocity VS, longitudinal velocity VL, average sound
velocity VM, Debye temperature ΘD, and melting temperature Tm of Y2O3.

P (GPa) ρ (g cm–3) VL (m/s) VS (m/s) VM (m/s) ΘD (K) Tm (K)

0 5.033 7.235 3.825 4.276 556.113 1530 ± 300


2 5.087 7.285 3.755 4.204 549.122 1584 ± 300
4 5.139 7.385 3.744 4.197 549.700 1642 ± 300
6 5.188 7.483 3.738 4.193 550.943 1699 ± 300
8 5.235 7.587 3.705 4.162 548.420 1768 ± 300
10 5.281 7.645 3.653 4.108 542.811 1822 ± 300
12 5.325 7.702 3.605 4.058 537.632 1875 ± 300
14 5.369 7.746 3.559 4.011 535.212 1921 ± 300

we adopt the Tian's formula to predict the hardness [35]: Hv=0.92(G/


B)1.137G0.708. The results of Vickers hardness Hv calculated above
equation of Y2O3 under pressure are listed in Table 4. The calculated
results are 7.705 GPa at P=0 GPa. Obviously, the Hv show a monotonic
dencrease with pressure in the range of 0–14 GPa, indicating that the
Hv of Y2O3 is significantly influenced by pressure.
Fig. 6. (a) Longitudinal wave velocity VL, shear-wave velocity VS, and average velocity
The Debye temperature, as an important fundamental parameter, is
VM versus pressure. (b) The calculated pressure dependences of the sound velocity in
closely related to many physical properties of solids, such as elastic Y2O3.
constants, melting temperature, and specific heat. Based on the Debye
theory, the vibration of a solid can be considered as an elastic wave. The melting temperature under pressure for Y2O3 is listed in Table 5.
The Debye temperature ΘD can be calculated from the elastic constant Quantitative analysis of Table 5 indicates that the melting temperature
data, since ΘD is proportional to the averaged sound velocity VM, via increase with increasing pressure.
[70], The single-crystal elastic wave velocities can be computed from the
elastic constants by resolution of the Christoffel equation [73]:
h ⎡ 3n ⎛ NAρ ⎞⎤
1/3
ΘD = ⎢ ⎜ ⎟⎥ VM (cijklkjkk−ρɷ2)ul=0, where cijkl is the elastic constant tensor element,
k ⎣ 4π ⎝ M ⎠⎦ (27) ρ is the mass density, ω is the vibrational angular frequency, ul is the
where h is Planck's constant, k is Boltzmann's constant, NA is displacement amplitude, and kj is the wave vector component of the
Avogadro's number, M is the molecular weight, ρ is the density, and vibrational wave. The solutions of this equation are of two types: a
n is the number of atoms per formula unit. In the present paper, for longitudinal wave VLA with polarization parallel to the direction of
Y2O3, n=5 and M=225.8099. The averaged sound velocity can be propagation and VTA1 and VTA2 donoted the shear-wave velocities
computed from [71] polarized in the (110) and (001) planes along the propagation
direction, respectively. The calculated elastic wave velocities along
1⎛ 1 2 ⎞
−1/3
[100], [110], and [111] directions are depicted in Fig. 6(b). It is found
VM = ⎜⎜ 3 + 3 ⎟⎟ that Y2O3 has low anisotropy in both longitudinal and shear-wave
3 ⎝ VL VS ⎠ (28)
velocities. Both the longitudinal velocities along the directions [100],
where VL and VS are the longitudinal and shear-wave velocities, [110], and [111] slowly increase with an increase in pressure, while the
respectively, shear-wave velocities along the directions [00,1], [110], and [111]
move slightly downward instead of upward.
VL = [(B + 4/3G )/ ρ]1/2 (29)

VS = (G / ρ)1/2 3.3. Thermodynamics properties


(30)

The velocity of sound for longitudinal and shear waves (VL and VS), Study of thermodynamic properties of materials is of great im-
Debye average velocity (VM), Debye temperature (ΘD) have been portance in order to extend our knowledge about their specific
investigated using Eqs. (27)–(30), respectively, shown in Table 5. behaviors when they are put under severe constraints such as high
The sound velocities depend on elastic moduli through the bulk pressure and high temperature environment.
modulus (B) and shear modulus (G). Thus, the greater the elastic In our work, the thermodynamic properties of cubic Y2O3 are
modulus, the greater is the sound velocity. Moreover, we calculated the determined in the temperature range 0–2000 K in GGA-PBE approx-
pressure dependence of mass density, melting temperature Tm listed in imation, where the quasi-harmonic model remains fully valid. The
Table 5 and also the sound velocities versus pressure up to 14 GPa are effect of pressure is studied in a pressure range from 0 to 14 GPa. We
depicted in Fig. 6(a). It is observed that the sound velocities VS and VM, began with a study of the isothermal bulk modulus. Fig. 7 presents in
and Debye temperature decrease monotonously as pressure increases, detail the variation of the isothermal bulk modulus with different
while longitudinal shear waves VL and mass density increases gradually temperature and different pressure. The variation of the bulk modulus
with increasing pressure. It should be noted that our calculated Debye with pressure at various temperatures is shown in Fig. 7(a). It is seen
temperature at T=0 K and P=0 GPa is 556.113 K, which is less than the that the bulk modulus B rapidly increases almost linearly with pressure
result of Ref. [34] (ΘD=533.42 K). The values of the Debye temperature at different temperatures. The variation of bulk modulus with tem-
decrease with increasing pressure from 0 to 14 GPa in Table 5. peratures at P=0, 2, 4, 6, 8, 10, 12, and 14 GPa is shown in Fig. 7(b). It
Specifically, there is about 1.25%, 1.15%, 0.93%, 1.38%, 2.39%, is clear that when the temperatures is less than 100 K, the isothermal
3.32%, and 3.76% decreases in the Debye temperature as hydrostatic bulk modulus nearly remains constant (≈157.9 GPa at zero tempera-
pressure increase, indicating that hydrostatic pressure has less influ- ture and zero pressure), but it drops remarkably at temperatures higher
ence on the Debye temperature of Y2O3. than 100 K up to 2000 K. Correspondingly, when T < 100 K, the
The melting temperature Tm of cubic cystalline solids has been primitive cell volume of both compounds has a small change; when
estimated using an empirical relation [72]: Tm=553+5.91C11 ± 300 K. T > 100 K, the cell volume changes rapidly as T increases, and rapid

3352
X. Zhang et al. Ceramics International 43 (2017) 3346–3355

(a) (b)
240
0 K
230 100 K
300 K
220 500 K 14GPa
1000 K 12GPa
210
Bulk modulus (GPa)

1500 K
2000 K 10GPa
200
8GPa
190
6GPa
180 4GPa
2GPa
170
0GPa
160
150
140
0 2 4 6 8 10 12 14 0 500 1000 1500 2000
Pressure (GPa) Temperature (K)
Fig. 7. Bulk modulus B of Y2O3 as a function of pressure and temperature. Fig. 9. (a) Pressure and (b) temperature dependence of the Grüneisen parameter γ for
cubic Y2O3.
volume variation makes the B rapidly decreases. The calculated zero-
pressure bulk modulus of Y2O3 is 154.5 GPa and 142.8 GPa at T=300 data obtained from various sources [75] (≈120–140 J mol−1 K−1) and
and 2100 K, respectively. An inspection of Fig. 7(a) and (b) reveals that the data given in the JANAF table [76] (≈116.74 J mol−1 K−1). From
the effect of increasing pressure on the isothermal bulk modulus is the Fig. 8, one can also see that the influences of the temperature on the
same as that of decreasing temperature. It is obvious that the effect of heat capacity are much more significant than that of the pressure on it.
pressure on the bulk modulus is more important than that of The Grüneisen parameter γ is a very important physical parameter
temperature. in relation to condensed matter. It can describe alterations in the
The heat capacity can be used to analyze the vibrational properties vibration of a crystal lattice resulting from an increase or decrease in
of solids, serving as a bridge between thermodynamics and microscopic volume with temperature change. The study of γ is very important for
structure. It provides an essential insight into vibrational properties, thermodynamic properties (such as the temperature dependence of
and has many applications. Fig. 8 illustrates the temperature depen- phonon frequencies), elasticity, and the non-syntony of matter. This
dence of the isochoric heat capacity CV and the isobaric heat capacity parameter indicates the anharmonicity of atomic vibrations, and
CP of Y2O3 at 0, 2, 4, 6, 8, 10, 12, and 14 GPa. A comparison of Fig. 8(a) reveals the relationship between the expansion coefficient and other
with (b) reveals that CV and CP exhibit very similar behavior, with both properties. In Fig. 9, we have plotted the Grüneisen parameter γ of
increasing sharply with rising temperature and being proportional to Y2O3 at various pressure and temperature. It can be observed that the γ
T3 at low temperatures (T < 500 K). However, at high temperatures, the decreases dramatically with an increase in pressure at a given
CV approaches a constant value, CP increases monotonously with temperature. Meanwhile, at higher temperatures, the γ decreases more
increments of the temperature. The values follow the Debye model at rapidly with the increasing pressure. As shown in Fig. 7(b), at higher
low temperature (CV(T) ~T3) and the classical behavior (CV(T) ~3 R for temperatures (T > 500 K), γ increases monotonously with temperature
mono-atomic solids) is found at sufficient high temperatures, obeying at a given pressure. These results are due to the fact that the effect of
Dulong and Petit's Rule [74]. Our calculated values temperature on the Grüneisen parameter γ is not as significant as that
(≈123.74 J mol−1 K−1) are close to the experimental and theoretical of pressure, and there will be a large thermal expansion at a low
pressure. The indisputable fact is that the effect of increased pressure
on the material is the same as decreased temperature of the material.
The thermal expansion coefficient α is a significant thermodynamic
parameter in theoretical and practical applications, it can be obtained
from the temperature derivative of lattice constant. The variations of
the thermal expansion a with temperature at different pressures (P=0,
2, 4, 6, 8, 10, 12, and 14 GPa) is presented in Fig. 10 for Y2O3
compound. In α–T graphs, it is noted that α increases exponentially
(α∝T3) at lower temperatures (T < 600 K) and increases slowly (ap-
proaches a linear increases) at higher temperatures (T > 600 K), we
have finally found that the thermal expansion coefficient α converges to
a constant value at high temperatures. This is due to the inadequacy of
the quasi-harmonic approximation at high temperatures and low
pressures. The effects of pressure on the thermal expansion.
coefficient α are very small at low temperatures, the effects are
increasingly obvious as the temperature increases. In this work, the
thermal expansion coefficient has been calculated as 0.153×10−5 (K−1)
at T=300 K at zero presssure and as 0.511×10−5 (K−1) at 2000 K,
respectively. Our calculated value is very compatible with
0.203×10−5 (K−1) at T=300 K which has been calculated by Chase [76].
Fig. 8. Temperature and pressure dependence of the heat capacity for Y2O3.

3353
X. Zhang et al. Ceramics International 43 (2017) 3346–3355

0.6 resources by the High-performance Super Computing Center in Xidian


University.
0.5
References
Thermal expension (10-5K-1)

0.4
[1] J. Qi, J.F. Liu, Y. He, W. Chen, C. Wang, Compression behavior and phase
transition of cubic In2O3 nanocrystals, J. Appl. Phys. 109 (2011) (063520-1-6).
0 GPa [2] T. de Boer, M.F. Bekheet, A. Gurlo, R. Riedel, A. Moewes, Band gap and electronic
0.3
2 GPa structure of cubic, rhombohedral, and orthorhombic In2O3 polymorphs: experi-
4 GPa ment and theory, Phys. Rev. B 93 (2016) (155205-1-7).
0.2 6 GPa [3] A. Gurlo, D. Dzivenko, P. Kroll, R. Riedel, High-pressure high-temperature
8 GPa synthesis of Rh2O3-II-type In2O3 polymorph, Phys. Stat. Sol. 2 (2008) 269–271.
[4] P. Agoston, K. Albe, Thermodynamic stability, stoichiometry, and electronic
10 GPa structure of bcc-In2O3 surfaces, Phys. Rev. B 84 (2011) (045311-1-20).
0.1
12 GPa [5] B.G. Domene, J.A. Sans, F.J. Manjón, Sergey V. Ovsyannikov, L.S. Dubrovinsky,
14 GPa D.M. Garcia, O. Gomis, D. Errandonea, H. Moutaabbid, Y.L. Godec, H.M. Ortiz,
0.0 A. Muñoz, P.R. Hernández, C. Popescu, Synthesis and high-pressure study of
0 500 1000 1500 2000 corundum type In2O3, J. Phys. Chem. C 119 (2015) 29076–29087.
Temperature (K) [6] A.L.J. Pereira, L. Gracia, D.S. Perez, R. Vilaplana, F.J. Manjon, D. Errandonea,
M. Nalin, A. Beltran, Structural and vibrational study of cubic Sb2O3 under high
Fig. 10. Temperature-dependent behavior of the expansion coefficient α for Y2O3 at pressure, Phys. Rev. B 85 (2012) (174108-1-11).
[7] S.V. Ovsyannikov, E. Bykova, M. Bykov, M.D. Wenz, A.S. Pakhomova, K. Glazyrin,
P=0, 2, 4, 6, 8, 10, 12, and 14 GPa.
H.P. Liermann, L. Dubrovinsky, Structural and vibrational properties of single
crystals of Scandia, Sc2O3 under high pressure, J. Appl. Phys. 118 (2015) (165901-
4. Conclusions 1-8).
[8] A.J. Pereira, D. Errandonea, A.B. Gracia, O. Gomis, J.A. Sans, B.G. Domene,
A. Miquel-Veyrat, F.J. Manjón1, A. Muñoz, C. Popescu, Structural study of α-Bi2O3
In this work, the structural, elastic, and thermodynamic properties under pressure, J. Phys.: Cond. Mater. 25 (2013) (475402-475402).
of cubic Y2O3 or α-Y2O3 under pressure have been predicted by first- [9] O. Gomis, D. Santamaría-Pérez, J. Ruiz-Fuertes, J.A. Sans, R. Vilaplana,
principles calculations in combination with the quasi-harmonic Debye H.M. Ortiz, B. García-Domene, F.J. Manjón, D. Errandonea, P. Rodríguez-
Hernández, A. Muñoz, M. Mollar, High-pressure structural and elastic properties of
model. The value of the equilibrium lattice parameter of the ground Ti2O3, J. Appl. Phys. 116 (2014) (133521-1-9).
state structure are found to be 10.560 Å, 10.594 Å, and 10.683 Å, using [10] S.V. Ovsyannikov, X. Wu, G. Garbarino, M.N. nez-Regueiro, V.V. Shchennikov,
LDA, GGA-PBE, and GGA+U approximations, respectively, which is in J.A. Khmeleva, A.E. Karkin, N. Dubrovinskaia, L. Dubrovinsky, High-pressure
behavior of structural, optical, and electronic transport properties of the golden
good agreement with the experimental data (10.604 Å). The pressure Th2S3-type Ti2O3, Phys. Rev. B 88 (2013) (184106-1-15).
dependence of a/a0 and V/V0 are also investigated, and both are found [11] X.-R. Hou, S.-M. Zhou, W.-J. Li, Y.-K. Li, Study on the effect and mechanism of
to decrease with increasing pressure. By fitting the third-order Birche– zirconia on the sinterability of yttria transparent ceramic, J. Eur. Ceram. Soc. 30
(2010) 3125–3129.
Munaghan EOS, the bulk modulus B0 and its pressure derivative B0′
[12] J. Zhang, L.-Q. An, M. Liu, S. Shimai, S.-W. Wang, Sintering of Yb3+:Y2O3
are determined as 145.4 GPa, 165.2 GPa, 155.6 GPa and 5.010, 5.003, transparent ceramic in hydrogen atmosphere, J. Eur. Ceram. Soc. 29 (2009)
4.849, respectively. The band gap of Y2O3 are calculated, and show that 305–309.
[13] C. Greskovich, S. Duclos, Ceramic scintillators, Annu. Rev. Mater. Sci. 27 (1997)
the GGA+U approximations is more accurate, with a value of 5.743 eV
69–88.
(experimental data, 5.5–6 eV). The predicted elastic constants and [14] D.C. Harris, Durable 3–5μm transmitting infrared window materials, Infrared
elastic modulus of Y2O3 do satisfy the mechanical stability criteria, Phys. Technol. 39 (1998) 185–201.
implying that this cubic single crystal is a mechanically stable system at [15] L. Lou, W. Zhang, A. Brioude, C.L. Luyer, J. Mugnier, Preparation and character-
ization of sol–gel Y2O3 planar waveguides, Opt. Mater. 18 (2001) 331–336.
pressure less than 14 GPa. The compressibility coefficient K [16] R. Korenstein, P. Cremin, T.E. Varitimos, R. Tustison, Optical properties of durable
(6.333×10−3 GPa at 0 GPa), the ratio B/G (2.33 at 0 GPa), and oxide coatings for infrared applications, Proc. SPIE 5078 (2003) 169–178.
Poisson's ratios ν for Y2O3 are obtained under pressure, which [17] R. McKee, F. Walker, M. Chisholm, Crystalline oxides on silicon-alternative
dielectrics for advanced transistor technologies, Mat. Res. Soc. Symp. Proc. 567
indicates that Y2O3 is a potentially compressible and behaves in a (1999) 415–425.
ductile manner at pressures up to 14 GPa. Moreover, the Vickers [18] G.D. Wilk, R.M. Wallace, J.M. Anthony, High-k gate dielectrics: current status
hardness, the Debye temperature, the melting temperature, the iso- andmaterials properties considerations, J. Appl. Phys. 89 (2001) 5243–5275.
[19] L.-A. Ragnarsson, S. Guha, M. Copel, E. Cartier, N.A. Bojarczuk, J. Karasinski,
tropic wave velocities, and the sound velocities as well as the elastic Molecular-beam-deposited yttrium-oxide dielectrics in aluminum-gated metal–
anisotropy of Y2O3 are also investigated at various pressures. The oxide–semiconductor field-effect transistors: effective electron mobility, Appl.
Debye temperature decreases monotonically with the increase of Phys. Lett. 78 (2001) 4169–4171.
[20] M.K. Bera, Y. Liu, L.M. Kyaw, Y.J. Ngoo, S.P. Singh, E.F. Chor, Positive threshold-
pressure, and Y2O3 has low anisotropy for both longitudinal and
voltage shift of Y2O3 gate dielectric InAlN/GaN-on-Si(111) MOSHE-MTs with
shear-wave velocities. respect to HEMTs, ECS J. Solid State Sci. Technol. 3 (2014) 120–126.
In addition, we carried out a study of the thermodynamic properties [21] I.M. Odeh, Fabrication and optical constants of amorphous copper nitride thin
films prepared by ion beam assisted dc magnetron reactive sputtering, J. Alloy.
using the quasi-harmonic Debye model at pressures of 0–14 GPa and
Compd. 454 (2008) 102–105.
temperatures of 0–2000 K. The heat capacity (CV and CP), the thermal [22] J.X. Zhao, Y.J. Zhang, H.Y. Gong, Y.B. Zhang, X.L. Wang, X. Guo, Y.J. Zhao,
expansion coefficient α, and the Grüneisen parameter γ are system- Fabrication of high-performance Y2O3 stabilized hafnium dioxid refractories,
atically calculated. We predict that the thermal expansion coefficient α Ceram. Int. 41 (2015) 5232–5238.
[23] L. Zhang, J. Feng, W. Pan, Vacuum sintering of transparent Cr:Y2O3 ceramics,
of Y2O3 is 0.153×10−5 K−1 at 0 GPa and ambient temperature. It can be Ceram. Int. 41 (2015) 8755–8760.
found that the thermal expansion coefficient converges to a nearly [24] D.Y. Yan, X.D. Xu, H. Lu, Y.W. Wang, P. Liu, J. Zhang, Fabrication and properties
constant value at high pressures and temperatures. The heat capacity of Y2O3 transparent ceramic by sintering aid combinations, Ceram. Int. 42 (2016)
16640–16643.
CV is proportional to T3 at low temperatures, and it converges to the [25] B. Ahmadi, S.R. Reza, M. Ahsanzadeh-Vadeqani, M. Barekat, Mechanical and
Dulong–Petit limit (≈123.74 J mol−1 K−1) at high temperatures. We optical properties of spark plasma sintered transparent Y2O3 ceramics, Ceram. Int.
hope our theoretical study do some help to the experiment. 42 (2016) 17081–17088.
[26] J.W. Palko, W.M. Kriven, S.V. Sinogeikin, J.D. Bass, A. Sayir, Elastic constants of
yttria (Y{sub2}O{sub3}) monocrystals to high temperatures, J. Appl. Phys. 89
(2001) 7791–7796.
Acknowledgments [27] A.B. Belonoshko, G. Gutierrez, R. Ahuja, B. Johansson, Molecular dynamics
simulation of the structure of yttria Y2O3 phases using pairwise interactions, Phys.
This study is financially supported by the National Natural Science Rev. B 64 (2001) 607–611.
[28] M.V. Abrashev, N.D. Todorov, J. Geshev, Raman spectra of R2O3 (R-rare earth)
Foundation of China (Grants nos. 51372203, 51332004, 51571166 and
sesquioxides with C-type bixbyite crystal structure: a comparative study, J. Appl.
11304238),We also thanks for the support for the computational

3354
X. Zhang et al. Ceramics International 43 (2017) 3346–3355

Phys. 116 (2014) (103508–1–7). niobium and tantalium nitrides M4N5 and M5N6 stoichiometry, Solid State Sci. 14
[29] H. Yusa, T. Tsuchiya, N. Sata, Y. Ohishiz, Dense yttria phase eclipsing the A-type (2012) 80–83.
sesquioxide structure: high-pressure experiments and ab initio calculation, Inorg. [51] M.A. Blanco, A.M. Pendás, E. Francisco, J.M. Recio, R. Franco, Thermodynamical
Chem. 49 (2010) 4478–4485. properties of solids from microscopic theory: applications to MgF2 and Al2O3, J.
[30] P.P. Bose, M.K. Gupta, R. Mittal, S. Rols, S.N. Achary, A.K. Tyagi, S.L. Chaplot, Mol. Struct. Theochem. 368 (1996) 245–251.
Phase transitions and thermodynamic properties of yttria, Y2O3: inelastic neutron [52] E. Francisc, M.A. Blanco, G. Sanjurjo, Atomistic simulation of SrF2 polymorphs,
scattering shell model and first-principles calculations, Phys. Rev. B 84 (2011) Phys. Rev. B 63 (2001) 094107–094116.
(094301–094301–11). [53] R.M. Wentzcovitch, K.J. Chang, M.L. Cohen, Electronic and structural properties of
[31] M. Ramzan, Y. Li, R. Chimata, R. Ahuja, Electronic, mechanical and optical BN and BP, Phys. Rev. B 34 (1986) 1071–1103.
properties of Y2O3 with hybrid density functional (HSE06), Comput. Mater. Sci. 71 [54] F. Birch, Finite elastic strain of cubic crystals, Phys. Rev. 71 (1947) 809–824.
(2013) 19–24. [55] M.G. Paton, E.N. Maslen, A refinement of the crystal structure Yttria, Acta Cryst. 19
[32] W.Y. Ching, Y.-N. Xu, Electronic and optical properties of Yttria, Phys. Rev. Lett. 65 (1965) 307–310.
(1990) 895–898. [56] V. Swamya, N.A. Dubrovinskaya, L.S. Dubrovinsky, High-temperature powder x-
[33] B.L. Ahuja, S. Sharma, N.L. Heda, S. Tiwari, K. Kumar, B.S. Meena, S. Bhatt, ray diffraction of yttria to melting point, J. Mater. Res. 14 (1999) 456–459.
Electronic and optical properties of ceramic Sc2O3 and Y2O3: compton spectroscopy [57] W.J. Tropf, D.C. Harris, Mechanical, thermal, and optical properties of Yttria and
and first principles calculations, J. Phys. Chem. Solids 92 (2016) 53–63. Lanthana-doped Yttria, Proc. SPIE 1112 (1989) 9–19.
[34] H.A. Badehian, H. Salehi, M. Ghoohestani, First-principles study of elastic, [58] W.R. Manning, O. Hunter, B.R. Powell, Elastic properties of polycrystalline Yttrium
structural, electronic, thermodynamical, and optical properties of Yttria (Y2O3) Oxide, Dysprosium oxide, Holmium oxide, and Erbium oxide: room temperature
ceramic in cubic phase, J. Am. Ceram. Soc. 96 (2013) 1832–1840. measurements, J. Am. Ceram. Soc. 52 (1969) 436–442.
[35] P. Blaha, K. Schwarz, G.K.H. Madsen, D. Kvasnicka, J. Luitz, K. WIEN2k, An [59] O. Tevet, O. Yeheskel, Elastic moduli of transparent Yttria, J. Am. Ceram. Soc. 82
Augmented Plane Wave, Local Orbitals Program for Calculating Crystal Properties, (1999) 136–144.
Techn. University Wien, Austria, Schwarz, 2001. [60] Y. Nigara, Measurement of the optical constants of Yttrium oxide, Jpn. J. Appl.
[36] J.X. Zheng, G. Ceder, T. Maxisch, W.K. Chim, W.K. Choi, Native point dfects in Phys. 7 (1968) 404–408.
Yttria as a high-dielectric-constant gate oxide material: a first-principles study, [61] P.W. Robertson, J. Peacock, Band offsets and Schottky barrier heights of high
Phys. Rev. B 73 (2006) 104101–104108. dielectric constant oxides, J. Appl. Phys. 92 (2002) 4712–4721.
[37] V. Swamy, H.J. Seifert, F. Aldinger, Thermodynamic properties of Y2O3 phases and [62] J.H. Jeong, J.S. Bae, S.S. Yi, J.C. Park, Y.S. Kim, Photoluminescence characteristics
the yttrium–oxygen phase diagram, J. Alloy. Compd. 269 (1998) 201–207. of Li-doped Y2O3:Eu3+ thin-film phosphors on sapphire substrates, J. Phys.:
[38] Y.D. Ou, W.S. Lai, Vacancy formation and clustering behavior in Y2O3 by first Condens. Matter 15 (2003) 567–572.
principles, Nucl. Instrum. Methods Phys. Res. Sect. B 269 (2011) 1720–1723. [63] P. Ravindran, F. Lars, P.A. Korzhavyi, B. Johansson, J. Wills, O. Eriksson, Density
[39] D.R. Mueller, D.L. Ederer, J. van Ek, W.L. O’Brien, Q.Y. Dong, J. Jia, T.A. Callcott, functional theory for calculation of elastic properties of orthorhombic crystals:
Soft-x-ray emission and the local p-type partial density of electronic states in Y2O3: application to TiSi2, J. Appl. Phys. 84 (1998) 4891–4904.
experiment and theory, Phys. Rev. B 54 (1996) 15034–15039. [64] R. Hill, The elastic properties of crystalline aggregate, Proc. Soc. A, vol. 65, pp.
[40] V.H. Mudavakkat, V.V. Atuchin, V.N. Kruchinin, A. Kayani, C.V. Ramana, 349–354, 1952.
Structure, morphology and optical properties of nanocrystalline yttrium oxide [65] G.V. Sin’ko, N.A. Smirnow, Ab initio calculations of elastic constants and thermo-
(Y2O3) thin films, Opt. Mater. 34 (2012) 893–900. dynamic properties of bcc, fcc, and hcp Al crystals under pressure, J. Phys.
[41] W.R. Manning, J.R. Hunter, Elastic properties of polycrystalline Yttrium oxide, Condens. Matter 14 (2002) 6989–7005.
Holmium oxide, and Erbium oxide: high temperature measurements, J. Am. [66] D.G. Pettifor, Theoretical predictions of structure and related properties of
Ceram. Soc. 52 (2006) 492–496. intermetallics, Mater. Sci. Technol. 8 (1992) 345–349.
[42] Y.-N. Xu, Z.-Q. Gu, W.Y. Ching, Electronic, structural, and optical properties of [67] S.I. Ranganathan, M.O. Starzewski, Universal elastic anisotropy index, Phys. Rev.
crystalline yttria, Phys. Rev. B 56 (1997) 14993–15000. Lett. 101 (2008) (055504–1–4).
[43] M.O. Marlowe, D.R. Wilder, Elasticity and internal friction of polycrystalline [68] J.F. Nye, Physical Properties of Crystals: Their Representation by Tensors and
Yttrium oxide, J. Am. Ceram. Soc. 48 (1965) 227–233. Matrices, Oxford University Press, NewYork, 1985, p. 145.
[44] V. Milman, B. Winkler, J.A. White, C.J. Pickard, M.C. Payne, E.V. Akhmatskay, [69] S.F. Pugh, Relations between the elastic moduli and the plastic properties of
R.H. Nobes, Electronic structure, properties, and phase stability of inorganic polycrystalline pure metal, Philos. Mag. 45 (1954) 823–826.
crystals: a pseudopotential plane-wave study, Int. J. Quantum Chem. 77 (2000) [70] O.L. Anderson, A simplified method for calculating the debye temperature from
895–900. elastic constants, J. Phys. Chem. Solids 24 (1963) 909–917.
[45] D.R. Hamann, M. Schluter, C. Chiang, Norm-conserving pseudopotentials, Phys. [71] L.Y. Lu, Y. Cheng, X.R. Chen, Thermodynamic properties of MgO under high
Rev. Lett. 43 (1979) 1494–1497. pressure from first-principles calculations, J. Zhu. Physica B 370 (2005) 236–242.
[46] J.P. Perdew, Y. Wang, Accurate and simple analytic representation of the electron- [72] M. Alouani, R.C. Albers, M. Methfessel, Calculated elastic constants and structural
gas correlation energy, Phys. Rev. B 45 (1992) 13244–13249. properties of Mo and MoSi2, Phys. Rev. B 43 (1991) 6500–6509.
[47] D. Goldfarb, A family of variable metric methods derived by variational means, [73] L.D. Landau, E.M. Lifschitz, Theory of elasticity, Course Theor. Phys. (1980) 7.
Math. Comput. 24 (1970) 23–26. [74] A.T. Petit, P.L. Dulong, Recherches sur quelques points importants de la Théorie de
[48] H.J. Monkhorst, J.D. Pack, Special points for Brillouin-zone integrations, Phys. la Chaleur, Ann. Chim. Phys. 10 (1819) 395–413.
Rev. B 13 (1976) 5188–5192. [75] T.S. Yashvili, D.S. Tsagareishvili, G.G. Gvelesiani, Enthalpy and specific heat of
[49] M.A. Blanco, E. Francisco, V. Luaña, GIBBS: isothermal-isobaric thermodynamics yttrium sesquioxide and cerium dioxide at high temperatures, Izv. Akad. Nauk
of solids from energy curves using a quasi-harmonic Debye model, Comput. Phys. SSSR 46 (1967) 409–414.
Commun 158 (2004) 57–72. [76] M.W. Chase, NIST-JANAF thermochemical tables (4rd ed.)J. Phys. Chem. Ref.
[50] T. Chihi, M. Fatmi, A. Bouhemadou, First-principles prediction of metastable Data 9 (1998) 1–25.

3355

You might also like