Abstract

Titanium dioxide (TiO2) is one of the important functional materials owing to its diverse applications in many fields of chemistry, physics, nanoscience, and technology. Hundreds of studies on its physicochemical properties, including its various phases, have been reported experimentally and theoretically, but the controversial nature of relative dielectric permittivity of TiO2 is yet to be understood. Toward this end, this study was undertaken to rationalize the effects of three commonly used projector augmented wave (PAW) potentials on the lattice geometries, phonon vibrations, and dielectric constants of rutile (R-)TiO2 and four of its other phases (anatase, brookite, pyrite, and fluorite). Density functional theory calculations within the PBE and PBEsol levels, as well as their reinforced versions PBE+U and PBEsol+U (U = 3.0 eV), were performed. It was found that PBEsol in combination with the standard PAW potential centered on Ti is adequate to reproduce the experimental lattice parameters, optical phonon modes, and the ionic and electronic contributions of the relative dielectric permittivity of R-TiO2 and four other phases. The origin of failure of the two soft potentials, namely, Ti_pv and Ti_sv, in predicting the correct nature of low-frequency optical phonon modes and ion-clamped dielectric constant of R-TiO2 is discussed. It is shown that the hybrid functionals (HSEsol and HSE06) slightly improve the accuracy of the above characteristics at the cost of a significant increase in computation time. Finally, we have highlighted the influence of external hydrostatic pressure on the R-TiO2 lattice, leading to the manifestation of ferroelectric modes that play a role in the determination of large and strongly pressure-dependent dielectric constant.
1. Introduction
Semiconducting crystalline TiO2 has been experimentally observed in three different temperature phases.1−4 These include the rutile phase, the anatase phase, and the brookite phase. The rutile phase, with a band gap of 3.0 eV,5 exists at temperatures around 1097 and 1172 K and is stable at ambient conditions.1,2 Allen et al.6 have found that a mixture of the brookite and anatase phases of TiO2 can be observable up to 772 K, whereas that of the anatase, brookite, and rutile phases can be observable around 872 K, and the rutile-only phase is observable at and above 1097 K. The anatase and brookite phases of TiO2 are semiconducting systems with band gaps of 3.4 eV7 and 3.3 eV,8 respectively.
One of the fundamental physical properties of rutile TiO2 (hereafter, R-TiO2) is the relative dielectric permittivity (expressed as εr)9−16 that provides insight into the applicability of Ti-based materials in high energy density storage devices,17,18 supercapacitors,19 among other technological areas.20 A number of theoretical21−23 and experimental11−16,24 investigations have been centered on the determination of εr. Specifically, Schöche et al.25 have discussed the low- and high-frequency contributions to εr reported by various authors. From these studies, it is apparent that different research groups have reported different values of εr (εr values between 100 and 10,000). The variability in the observed or proposed values of εr has caused many research attentions owing to the underlying controversy. Bonkerud and co-workers16 have recently uncovered that the giant εr reported in early studies11,12 was due to an incorrectly designed experiment or an incorrect interpretation. Their argument was based on the early capacitance measurements that have considered the full thickness of the crystal, where metallic contacts were deposited on opposite sides. In any case, R-TiO2 ceramics when co-doped with Nb + Tr (Tr = Al, In, Ga, Tl, Dy, Gd, La, Y, Sm, V, Li) feature colossal dielectric permittivity.26−28 The experimental εr values for these systems vary between 10,000 and 100,000. They are yet to be confirmed by first principles and/or density functional theory (DFT) calculations.
Lee et al.29 have studied the dielectric properties of R-TiO2 using a variety of exchange correlation energy DFT functionals, including the local density approximation (LDA), the LDA+U, the generalized gradient approximation (GGA) method, and the PBE0 and HSE06 hybrid functionals. The essence of the study was that the dielectric properties of R-TiO2 vary upon changing the theoretical methods applied, and the frequencies of phonon vibrations may or may not be softened at high-symmetry points in k-space. The εr values calculated using LDA were in close agreement with experiment, and those calculated with the HSE06 functional and LDA+U method were overestimated and severely underestimated, respectively. For instance, the ionic contribution to εr was reported to be 63.7, 40.0, 176.5, 282.1, and 376.4 with LDA+U (U = 3.0 eV), LDA+U (U = 4.36 eV), LDA, PBE0, and HSE06, respectively (see the actual data in Table 1 of ref (29)). Our inspection of their data shows that εr is not meaningful with LDA+U and is partially (or significantly) overestimated with the LDA (or PBE0 and HSE06) methods compared to the corresponding experimental mean value of 159.7.14 Furthermore, the LDA result outlined above was abnormally different from that of 133.5 reported by Shojaee et al.,23 even though the HSE functional gave a value of 129.3 (exact values 159 and 70 along the x–y plane and z-direction, respectively).30 Interestingly, however, the electronic contribution to εr did not show any strong dependence on the quality of theoretical methods applied (PBE, PBE0, etc.).21,31−33 The incipient ferroelectric behavior and the low-frequency dielectric constant are directly linked to lattice dynamics, with a strong dependence on the low-frequency transverse optical phonon mode A2u.34,35
In this study, we have theoretically investigated the lattice properties, zone-center phonon vibrations, and the low- and high-frequency-dependent (ionic and electronic contributions, respectively) dielectric permittivity of rutile (R-)TiO2 using DFT at PBEsol36 and PBE37,38 levels, with and without incorporating the Hubbard term U.39,40 Density functional perturbation theory (DFPT) was employed for the evaluation of phonon vibrations and dielectric properties.41−43 The same properties were also investigated for the anatase and brookite phases, as well as for the two high-pressure structures of TiO2 (called pyrite44−46 and fluorite45−48) for comparison.
The main objective of our study is to demonstrate the dependency of the above properties on the two DFT functionals employed, as well as that on the three commonly used variants of the projector augmented wave (PAW) potential49,49,50 available in the Vienna ab initio simulation package (VASP).51−53 The study is necessary because it is imperative to clarify how and to what extent the promotion of the number of core electrons to valence electrons in the PAW potential(s) affects the accuracy of the physical properties of TiO2 (e.g., lattice geometry, relative permittivity, and phonon frequencies). The two GGA functionals were chosen because PAW potentials are specifically designed for use in conjunction with PBE- and PBE-like functionals (e.g., PBEsol); they are computationally inexpensive and are at the forefront of large-scale, high-throughput calculations for in-silico design of new functional materials. To verify the conclusions arrived at using the two PBE-based functionals, similar calculations for R-TiO2 were performed at a relatively high level using hybrid functionals HSE0654 and HSEsol.55 Finally, we have compared our computed results with the low-temperature experimental results.
2. Computational Details
The electronic structure properties of all the five phases of TiO2 were calculated using PBE and PBEsol; the VASP code was used. Hereafter, we refer the anatase-, brookite-, pyrite-, and fluorite phases of TiO2 to as A-, B-, P-, and F-TiO2, respectively.
The PAW potentials were used, in which the core electrons were frozen, and were replaced by pseudopotentials. There are four versions of PAW potential for Ti, distributed by VASP. Depending on the size of the cutoff radius used to define the potential around the nucleus, the PAW potentials have been considered hard or soft, meaning they differ from each other in the degree to which they treat inner-shell electrons as valence electrons. We used three commonly used versions of the PAW potential for Ti, keeping the same for the O site in TiO2 unchanged. We label the standard PAW potential of Ti, which has four valence electrons 3d34s1, as Ti_std. The remaining two variants of the PAW potential for Ti are augmented versions, labeled as Ti_pv and Ti_sv, which include 2p6 and 2s22p6 semi-core states as part of the basis set for valence bands, respectively. The outermost cutoff radii were 2.8, 2.5, and 2.3 Å for Ti_std, Ti_pv, and Ti_sv, respectively. The Wigner–Seitz radii of the corresponding PAW potentials were 1.323, 1.323, and 1.217 Å, respectively.
The crystal lattices of the five phases of TiO2 were fully relaxed without any constraints, starting from their respective experimental structures wherever feasible (see discussion below). The cutoff energy for the plane-wave basis set was set at 520 eV, and the break condition for the electronic relaxation loop was set at 10–8 eV. The force on each ion was less than 0.002 eV Å–1. The k-meshes of 9 × 9 × 15, 8 × 8 × 10, 6 × 4 × 6, 10 × 10 × 10, and 8 × 8 × 8 were utilized for R-, A-, B-, F-, and P-TiO2, respectively. High precision, together with a blocked Davidson iteration scheme, was invoked.
The ionic41,56−58 and electronic (optic)57−59 contributions (εij(0) and εij(∞), respectively) to dimensionless total relative dielectric permittivity εr (εr = εij(0) + εij(∞)) were calculated at the same levels of theory (PBE and PBEsol) for all the three PAW potentials mentioned above. DFPT was employed since it provides better estimation of the relative dielectric permittivity at the expense of computational time, especially when compared with experiment. The finite difference method (FDM)60−62 is computationally expensive, which was also considered in conjunction with the hybrid functionals and GGA to examine whether εr obtained with this method are comparable with those emerged from DFPT/GGA. The same k-meshes used for the convergence of crystal lattices and evaluation of other properties were used for the calculation of εij(0). The cutoff energy was set at 520, 600, and 700 eV that enabled us to test the dependence of the cutoff energy for plane-wave on εij(0) and εij(∞) and εr.
For reasons discussed in the following section, the hybrid functionals, HSE06 and HSEsol, were employed to obtain the relaxed lattice geometry, phonon modes, and dielectric permittivity of R-TiO2. A 4 × 4 × 6 k-mesh was chosen both for geometry relaxation and phonon mode analysis. The cutoff energy for the plane-wave basis set, the total energy convergence for electronic relaxation, and the residual forces used for ionic relaxation were 520 eV, 10–6 eV, and 0.002 eV Å–1, respectively. The FDM was applied for the calculation of the phonon vibrational modes and relative dielectric permittivity. The cutoff energy for the plane-wave basis set and the ion relaxation loop were 600 and 10–6 eV, respectively.
Phonon dispersion and Raman spectra were calculated using FDM.63,64 The PBEsol relaxation geometries of R-TiO2, A-TiO2, and B-TiO2 were used to construct supercells consisting of 72, 108, and 96 atoms, respectively. A 3 × 3 × 3 k-mesh was used to calculate phonon dispersion for all three systems above. For the calculation of the Raman spectra, the size of the k-mesh was chosen to be 9 × 9 × 15, 8 × 8 × 4, and 6 × 6 × 4 for the corresponding systems, respectively. The cutoff energy was set to 520 eV, and the energy for electronic relaxation loop and the residual force were converged to within 10–7 eV and 0.02 eV Å–1, respectively. For cubic P- and F-TiO2, supercells with cell sizes close to 10 Å were used, consisting of 96 and 192 atoms, respectively. A k-mesh of 2 × 2 × 2 was used, together with the electronic relaxation loop and the residual force that were set converged to within 10–8 eV and 0.02 eV Å–1, respectively. The DFPT formalism was utilized, together with the Phonopy code.65
3. Results and Discussion
3.1. Lattice Properties
The fully relaxed lattice geometries of all the five phases of TiO2 are shown in Figure 1. Tables 123 list the calculated lattice properties of R-, A-, and B-phases of TiO2 and compare them with experiment. The properties include the lattice constants (a, b, c, α, β, and γ), cell volume V, and mass density ρ.
Figure 1.

(a)–(e) Polyhedral models of the PBEsol relaxed unit-cell geometries of the five phases of TiO2, obtained with Ti_std. The orientation of the crystallographic axes a, b, and c is shown for each case.
Table 1. Comparison of PBE, PBEsol, PBE+U, and PBEsol+U Computed Lattice Constants (a = b ≠ c and α = β = γ), Cell Volume (V), and Mass Density (ρ) of R-TiO2 with Experimenta,b.
| PAW | a = b/Å | c/Å | V/Å3 | ρ/gcm–3 | a = b/Å | c/Å | V/Å3 | ρ/gcm–3 |
|---|---|---|---|---|---|---|---|---|
| PBE | PBE+U | |||||||
| Ti_std | 4.661 | 2.970 | 64.52 | 4.11 | 4.687 | 3.023 | 66.41 | 3.99 |
| Ti_pv | 4.652 | 2.968 | 64.23 | 4.13 | 4.673 | 3.014 | 65.82 | 4.03 |
| Ti_sv | 4.645 | 2.969 | 64.05 | 4.14 | 4.665 | 3.007 | 65.42 | 4.05 |
| PBEsol | PBEsol+U | |||||||
| Ti_std | 4.615 | 2.949 | 62.79 | 4.22 | 4.640 | 3.000 | 64.59 | 4.11 |
| Ti_pv | 4.600 | 2.943 | 62.28 | 4.26 | 4.623 | 2.989 | 63.89 | 4.15 |
| Ti_sv | 4.596 | 2.941 | 62.12 | 4.27 | 4.614 | 2.982 | 63.49 | 4.18 |
| HSEsol | ||||||||
| Ti_std | 4.572 | 2.934 | 61.33 | 4.32 | ||||
| Ti_pv | 4.556 | 2.934 | 60.91 | 4.36 | ||||
| Ti_sv | 4.551 | 2.931 | 60.70 | 4.37 | ||||
| HSE06 | ||||||||
| Ti_std | 4.603 | 2.947 | 62.43 | 4.25 | ||||
| Ti_pv | 4.590 | 2.949 | 62.13 | 4.27 | ||||
| Ti_sv | 4.584 | 2.947 | 61.92 | 4.28 | ||||
| expt.66 | 4.601 | 2.964 | 62.745 | 4.23 | 4.601 | 2.964 | 62.745 | 4.23 |
α = β = γ = 90° for the theoretical methods applied.
Three different PAW potentials used.
Table 2. Comparison of PBEsol and PBEsol+U Level Lattice Constants (a = b ≠ c and α = β = γ), Cell Volume (V), and Mass Density (ρ) of A-TiO2 with Experimenta,b.
| PAW | PBEsol |
PBEsol+U |
||||||||
|---|---|---|---|---|---|---|---|---|---|---|
| a/Å | b/Å | c/Å | V/Å3 | ρ/gcm–3 | a/Å | b/Å | c/Å | V/Å3 | ρ/gcm–3 | |
| Ti_std | 3.800 | 3.800 | 9.547 | 137.85 | 3.85 | 3.858 | 3.858 | 9.553 | 142.19 | 3.73 |
| Ti_pv | 3.777 | 3.777 | 9.575 | 136.59 | 3.88 | 3.827 | 3.827 | 9.591 | 140.47 | 3.78 |
| Ti_sv | 3.773 | 3.773 | 9.564 | 136.14 | 3.90 | 3.817 | 3.817 | 9.575 | 139.49 | 3.80 |
| expt.68 | 3.784 | 3.784 | 9.518 | 136.28 | 3.89 | 3.784 | 3.784 | 9.518 | 136.28 | 3.89 |
α = β = γ = 90° for the theoretical methods applied.
Three different PAW potentials used.
Table 3. Comparison of PBEsol and PBEsol+U Level Lattice Constants (a = b ≠ c and α = β = γ), Cell Volume (V), and Mass Density (ρ) of B-TiO2 with Experimenta,b.
| PAW | PBEsol |
PBEsol+U |
||||||||
|---|---|---|---|---|---|---|---|---|---|---|
| a/Å | b/Å | c/Å | V/Å3 | ρ/gcm–3 | a/Å | b/Å | c/Å | V/Å3 | ρ/gcm–3 | |
| Ti_std | 5.149 | 9.203 | 5.463 | 258.88 | 4.10 | 5.228 | 9.262 | 5.519 | 267.22 | 3.97 |
| Ti_pv | 5.127 | 9.186 | 5.449 | 256.64 | 4.13 | 5.203 | 9.243 | 5.489 | 263.94 | 4.02 |
| Ti_sv | 5.124 | 9.177 | 5.444 | 255.95 | 4.15 | 5.190 | 9.227 | 5.476 | 262.23 | 4.05 |
| expt.68 | 5.144 | 9.293 | 5.412 | 258.71 | 4.10 | 5.144 | 9.293 | 5.412 | 258.71 | 4.10 |
| expt.69 | 5.137 | 9.714 | 5.452 | 256.94 | 4.13 | 5.137 | 9.714 | 5.452 | 256.94 | 4.13 |
| expt.70 | 4.981 | 9.082 | 5.53 | 250.14 | 4.24 | 4.981 | 9.082 | 5.53 | 250.14 | 4.24 |
α = β = γ = 90° for the theoretical methods applied.
Three different PAW potentials used.
Experimental lattice constants even reported for the three major phases of TiO2 vary widely from one research group to another.56,66−72 The variation may be due to the differences in the experimental conditions and the fitting methods adopted in measuring and refining the structural data. A similar attribute is notable of the data obtained using two DFT functionals and the three variants of the PAW potential, as well as when the Hubbard U (U = 3.0 eV) term was combined with GGA.
From Table 1, it may be seen that the lattice constants of R-TiO2 follow the trend: a = b ≠ c and α = β = γ = 90°. PBEsol reproduces lattice constants and cell volumes, which are in fair agreement with the experimental range of reported values (not shown). In contrast, the PBE functional systematically overestimates the same properties (Table 1). The same is also observed when both of them were compared with to the high-precision experimental data reported so far with an R-factor (discrepancy index) as low as 1.47%.66 For a particular GGA method, PBE or PBEsol, when passed from Ti_std via Ti_pv to Ti_sv, the lattice constants (a = b and c) and cell volume were systematically decreased; this is opposite to the trend in the mass density that was increased. The result suggests that the lattice constant or cell volume tends to decrease and mass density tends to increase as the PAW potential softens. Conversely, the HSE06 method gave improved lattice properties over HSEsol. The main difference between PBEsol and HSE lies in the lattice volume, in which the predictability of the former is relatively accurate compared to that feasible in the crystal.
The DFT-U method did not show any dramatic improvement in the lattice properties. Rather, it significantly overestimates all the lattice constants, with a consequent increase in the cell volume. The reducing behavior of the lattice constants observed with DFT was observed with DFT+U when passing from Ti_std via Ti_pv to Ti_sv.
A-TiO2 is metastable. It is stable only up to a temperature of 600 °C. Beyond this temperature, a transformation to rutile phase occurs, even though the reported thermal analysis suggests that the transformation from anatase to rutile may start at a temperature of around 400 °C.73 The geometry of A-TiO2 is tetragonal, with space group I41/amd (141), and the unit and primitive cells contain four and two formula units of TiO2, respectively. Table 2 summarizes the conventional unit-cell lattice properties of A-TiO2, obtained from this work and reported experimentally. Since PBEsol-calculated lattice properties agree well with experiment for R-TiO2 (cf. Table 1), we did not consider the PBE and DFT+U methods to compute the lattice properties for A-TiO2. Table 2 summarizes the PBEsol results that demonstrate that Ti_std slightly overestimates the lattice constants and cell volume and underestimates the mass density compared to experiment.68 There is no appreciable improvement in these properties upon increasing the size of the PAW potential from Ti_std through Ti_pv to Ti_sv. PBEsol+U has overestimated the lattice constants and cell volume and underestimated the mass density, all compared to experiment.
The crystal structure of B-TiO2 is orthorhombic. It has eight formula units of TiO2 (space group Pbca (61)). Several experimental studies have reported the crystal structure of B-TiO2, showing a dispersion in the values of the lattice properties. For instance, Murugesan et al.68 (Silva Junior et al.74) [Nishio-Hamane et al.75] {Rezaee et al.70} have reported a = 5.144 Å (5.138; 5.259) [4.596] {4.492}, b = 9.293 (9.194; 9.202) [8.962] {9.101} Å, and c = 5.412 (5.449; 5.47) [4.823] {5.455} Å. Because of the variation in the lattice constants, the nature of both V and ρ is changed appreciably. The V values extracted from the unit cells of the corresponding crystals were 258.711 (256.844; 264.711) [198.656] {245.341} Å3. Our PBEsol results are very close to those reported by Murugesan et al.68 (Table 3). There is a monotonic decrease in the lattice constants with respect to the increase in the number of core electrons as valence electrons in the PAW potentials, an observation which is very similar to that found for R- and A-TiO2 (vide supra).
Table 4 summarizes the calculated lattice properties of P- and F-phases of TiO2. These are compared with the properties of corresponding crystals deposited to the Materials Project database76 (IDs: mp-110259177 and mp-1008677,78 respectively), as a result of the lack of experimentally known data. In any case, the PBEsol lattice constants are all slightly underestimated relative to the PBE, with a concomitant decrease in the cell volume and an increase of mass density. The trend in the cell properties for a given correlated method with respect to the three PAW potentials is consistent with that observed for R-, A-, and B-TiO2 (see above). We have observed that F- and P-TiO2 lattices are relatively denser than that of the brookite, anatase, and rutile lattices of TiO2, in agreement with the computational data of Dharmale et al.44
Table 4. Comparison of Lattice Constants (a = b = c and α = β = γ), Cell Volume (V), and Mass Density (ρ) of P- and F-TiO2, Obtained Using PBEsol and PBE, with Those Catalogued in the Materials Project Database (Calculated with [PBE/Ti_pv]76)a.
| PAW | PBEsol |
PBE |
||||
|---|---|---|---|---|---|---|
| a = b = c | V/Å3 | ρ/gcm–3 | a = b = c | V/Å3 | ρ/gcm–3 | |
| Pyrite (Cubic, P-a3̅) | ||||||
| Ti_std | 4.861 | 114.84 | 4.62 | 4.905 | 118.03 | 4.49 |
| Ti_pv | 4.852 | 114.22 | 4.64 | 4.902 | 117.78 | 4.50 |
| Ti_sv | 4.848 | 113.97 | 4.65 | 4.899 | 117.55 | 4.51 |
| mp-110259177 | 4.902 | 117.79 | 4.50 | 4.902 | 117.79 | 4.50 |
| Fluorite (Cubic, Fm-3̅m) | ||||||
| Ti_std | 4.789 | 109.87 | 4.83 | 4.836 | 113.12 | 4.69 |
| Ti_pv | 4.786 | 109.63 | 4.84 | 4.838 | 113.25 | 4.68 |
| Ti_sv | 4.783 | 109.40 | 4.85 | 4.835 | 113.05 | 4.69 |
| mp-100867778 | 4.84 | 113.38 | 4.68 | 4.840 | 113.38 | 4.68 |
Three different PAW potentials used.
3.2. Phonon Vibrations and Dynamical Stability
Born’s stability criterion79,80 suggests that a crystal lattice is said to be dynamically stable if the normal mode vibration frequencies of all the optical phonons are positive, that is, if the eigenvalues of the dynamical matrix are all positive. To this end, we investigated the nature of phonon vibrational frequencies for all the five phases of TiO2, in combination with three PAW potentials. The frequencies of IR- and Raman-active modes of R- and A-TiO2 are summarized in Tables 5 and 6, respectively, whereas those of B-TiO2 are summarized in Table S1; the PBE and experimental phonon frequencies are included in these tables wherever applicable.
Table 5. Comparison of Zone-Center Fundamental IR and Raman Phonon Frequencies ω (cm–1), as Well as IR Intensities I (km/mol) of Rutile-TiO2, Obtained Using PBEsol and PBE Functionals, in Conjunction with the PAW Potentials Ti_std, Ti_pv, and Ti_sv Centered on Tiabcdef.
| symmetryf | exptg | expth,i | PBEsol |
PBE |
||||||||||
|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
| Ti_std | Ti_pv | Ti_sv | Ti_std | Ti_pv | Ti_sv | |||||||||
| ω | I | Ω | I | ω | I | ω | I | ω | I | ω | I | |||
| B2g(R) | 825.5 | 826 (826.6) | 810.2 | 0.0 | 795.3 | 0.0 | 793.1 | 0.0 | 783.6 | 0.0 | 770.3 | 0.0 | 769.3 | 0.0 |
| A1g(R) | 609.8 | 612 (610.4) | 605.8 | 0.0 | 591.0 | 0.0 | 589.8 | 0.0 | 580.9 | 0.0 | 568.4 | 0.0 | 568.0 | 0.0 |
| Eu(IR) | 500 (494.0) | 495.3 | 20.7 | 476.9 | 13.4 | 478.4 | 13.2 | 481.8 | 15.3 | 467.8 | 10.5 | 468.9 | 10.3 | |
| Eu(IR) | 500 (494.0) | 495.3 | 20.7 | 476.9 | 13.4 | 478.3 | 13.2 | 481.8 | 15.3 | 467.8 | 10.5 | 468.9 | 10.3 | |
| Eg(R) | 445.8 | 447 (444.9) | 459.6 | 0.0 | 448.0 | 0.0 | 447.1 | 0.0 | 440.6 | 0.0 | 430.0 | 0.0 | 429.6 | 0.0 |
| Eg(R) | 445.8 | 447 (444.9) | 459.6 | 0.0 | 448.0 | 0.0 | 447.1 | 0.0 | 440.6 | 0.0 | 430.0 | 0.0 | 429.6 | 0.0 |
| A2g (S) | 405.5 | 0.0 | 413.6 | 0.0 | 415.1 | 0.0 | 407.7 | 0.0 | 414.8 | 0.0 | 416.4 | 0.0 | ||
| B1u(S) | (406.3) | 395.8 | 0.0 | 372.5 | 0.0 | 373.9 | 0.0 | 369.7 | 0.0 | 354.3 | 0.0 | 357.1 | 0.0 | |
| Eu(IR) | 388 (374.4) | 375.0 | 9.7 | 366.4 | 4.0 | 366.9 | 3.7 | 356.9 | 3.2 | 353.9 | 1.0 | 355.1 | 0.9 | |
| Eu(IR) | 388 (374.7) | 375.0 | 9.7 | 366.4 | 4.0 | 366.9 | 3.7 | 356.9 | 3.2 | 353.9 | 1.0 | 355.0 | 0.9 | |
| A2u(IR) | 167 (188.8) | 180.5 | 179.2 | 61.2 | 168.0 | 71.2 | 165.3 | 99.6 | 184.2 | 106.1i | 171.5 | 98.8i | 168.3 | |
| Eu(IR) | 235.5 | 183 (172.7) | 158.7 | 105.2 | 93.0 | 110.1 | 98.4 | 108.5 | 95.4 | 115.5 | 62.8 i | 114.1 | 54.3i | 111.5 |
| Eu(IR) | 235.5 | 183 (172.7) | 158.7 | 105.2 | 93.0 | 110.1 | 98.2 | 108.4 | 95.4 | 115.5 | 62.8 i | 114.1 | 54.5i | 111.7 |
| B1g(R) | 140.2 | 143 (141.6) | 134.2 | 0.0 | 139.7 | 0.0 | 140.5 | 0.0 | 143.9 | 0.0 | 148.5 | 0.0 | 149.0 | 0.0 |
| B1u(S) | (113.0) | 118.3 | 0.0 | 74.9 | 0.0 | 78.3 | 0.0 | 84.8 | 0.0 | 30.1i | 0.0 | 16.7i | 0.0 | |
| A2u(IR) | 0.0 | 13.7 | 0.1 | 8.5 | 0.4 | 8.2 | 0.0 | 8.9 | 0.1 | 8.3 | 0.6 | 9.0 | ||
| Eu(IR) | 0.0 | 9.1 | 0.1 | 8.5 | 0.3 | 8.3 | 0.0 | 8.9 | 0.1 | 8.3 | 0.4 | 11.6 | ||
| Eu(IR) | 0.0 | 9.1 | 0.1 | 13.0 | 0.2 | 12.6 | 0.1 | 14.2 | 0.1 | 13.2 | 0.4 | 10.1 | ||
Experimental frequencies are included wherever available, and an energy cutoff of 700 eV was used.
Negative value in columns 12 and 14 represent imaginary frequency.
Schöche et al.25 reported the four room-temperature IR-active phonon modes (three degenerate Eu and one non-degenerate A2u) to be 188.6 ± 1.2, 379.3 ± 0.2, 500.5 ± 0.3, and 172.3 ± 1.9 cm–1, respectively.
Gervais and Piriou85 reported the four IR-active phonon modes (three degenerate Eu and one non-degenerate A2u) at 189, 381.5, 508, and 172 cm–1, respectively.
Eagles reported the four IR-active phonon modes (three degenerate Eu and one non-degenerate A2u) at 183, 388, 500, and 167 cm–1, respectively.86
R, IR, and S correspond to Raman- and IR-active and silent modes, respectively.
Ref (83).
Ref (82).
Values in parentheses represent coherent inelastic neutron determination of phonon frequencies at high-symmetry points of Brillouin zone.81
Table 6. Comparison of Fundamental IR and Raman Frequencies ω (cm–1), as Well as IR Intensities I (km/mol), Obtained Using PBEsol and PBE Functionals, in Conjunction with PAW Potentials Ti_std, Ti_pv, and Ti_sv Centered on Ti of A-TiO2a,b.
| exptc,d | PBEsol |
PBE |
|||||||||||
|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
| Ti_std | Ti_pv | Ti_sv | Ti_std | Ti_pv | Ti_sv | ||||||||
| ω | I | ω | I | ω | I | ω | I | ω | I | ω | I | ||
| Eg(R) | 640 (632.7) | 640.4 | 0.0 | 619.8 | 0.0 | 620.6 | 0.0 | 616.5 | 0.0 | 599.7 | 0.0 | 599.7 | 0.0 |
| Eg(R) | 640 (632.7) | 640.4 | 0.0 | 619.8 | 0.0 | 620.6 | 0.0 | 616.5 | 0.0 | 599.7 | 0.0 | 599.7 | 0.0 |
| B2u (S) | 548.6 | 0.0 | 535.9 | 0.0 | 534.4 | 0.0 | 531.0 | 0.0 | 519.2 | 0.0 | 518.6 | 0.0 | |
| A1g(R) | 519 (515.4) | 516.7 | 0.0 | 510.8 | 0.0 | 510.7 | 0.0 | 499.6 | 0.0 | 493.9 | 0.0 | 494.4 | 0.0 |
| B1g(R) | 515 (515.4) | 497.5 | 0.0 | 487.3 | 0.0 | 489.0 | 0.0 | 478.5 | 0.0 | 475.1 | 0.0 | 477.6 | 0.0 |
| Eu(IR) | 435 (TO) | 446.0 | 76.7 | 419.0 | 51.1 | 420.7 | 50.1 | 412.0 | 57.9 | 395.3 | 34.6 | 396.5 | 32.8 |
| Eu(IR) | 435 (TO) | 446.0 | 76.7 | 419.0 | 51.1 | 420.7 | 50.1 | 412.0 | 57.9 | 395.3 | 34.6 | 396.5 | 32.8 |
| B1g(R) | 400 (396.0) | 385.4 | 0.0 | 374.2 | 0.0 | 374.1 | 0.0 | 374.9 | 0.0 | 361.7 | 0.0 | 361.8 | 0.0 |
| A2u(IR) | 366 (LO) | 355.9 | 100.3 | 326.6 | 93.3 | 326.9 | 92.0 | 323.3 | 101.5 | 300.0 | 94.0 | 301.2 | 92.8 |
| Eu(IR) | 262 (TO) | 240.8 | 75.4 | 227.6 | 90.5 | 231.5 | 89.2 | 226.3 | 95.8 | 209.0 | 107.9 | 211.0 | 106.9 |
| Eu(IR) | 262 (TO) | 240.8 | 75.4 | 227.6 | 90.5 | 231.3 | 89.2 | 226.3 | 95.8 | 209.0 | 107.9 | 211.0 | 107.0 |
| Eg(R) | 197 (196.3) | 164.7 | 0.0 | 167.1 | 0.0 | 169.2 | 0.0 | 168.4 | 0.0 | 173.4 | 0.0 | 175.9 | 0.0 |
| Eg(R) | 197 (196.3) | 164.7 | 0.0 | 167.1 | 0.0 | 169.2 | 0.0 | 168.3 | 0.0 | 173.4 | 0.0 | 175.9 | 0.0 |
| Eg(R) | 144 (140.9) | 151.4 | 0.0 | 135.2 | 0.0 | 137.9 | 0.0 | 130.7 | 0.0 | 113.9 | 0.0 | 115.2 | 0.0 |
| Eg(R) | 144 (140.9) | 151.3 | 0.0 | 135.2 | 0.0 | 137.6 | 0.0 | 130.7 | 0.0 | 113.9 | 0.0 | 115.1 | 0.0 |
| A2u(IR) | 0.0 | 8.2 | 0.0 | 9.4 | 0.0 | 7.1 | 0.0 | 10.1 | 0.0 | 7.2 | 0.0 | 9.1 | |
| Eu(IR) | 0.0 | 9.9 | 0.0 | 9.3 | 0.1 | 9.2 | 0.0 | 9.8 | 0.0 | 9.4 | 0.0 | 7.2 | |
| Eu(IR) | 0.0 | 9.6 | 0.0 | 7.2 | 0.2 | 9.2 | 0.0 | 8.1 | 0.0 | 9.4 | 0.1 | 9.3 | |
Figures 234 show the phonon dispersion curves of R-, A-, and B-TiO2, respectively. Figure 5a,b illustrates the same curves for F- and P-TiO2, respectively. For the first three cases, the phonon modes are all stable at the center of the zone, Γ(k (0,0,0)), and at other points of high symmetry along the direction of the wave vector of the first Brillouin zone. That is, the optical phonon modes do not involve any imaginary frequencies, confirming that the R-, A-, and B-TiO2 lattices are dynamically stable irrespective of the three PAW potentials utilized.
Figure 2.
Phonon dispersion of R-TiO2, obtained using DFPT in combination with PBEsol and the three PAW potentials centered at the Ti site of the crystal lattice: (a) Ti_std; (b) Ti_pv; and (c) Ti_sv. The phonon frequencies in the vertical axis are in THz (1 THz = 33.3564 cm–1).
Figure 3.
Phonon band structure of A-TiO2, obtained using DFPT in combination with PBEsol and the three variants of the PAW potential centered at the Ti site of the crystal lattice: (a) Ti_std; (b) Ti_pv; and (c) Ti_sv. The phonon frequencies in the vertical axis are in THz (1 THz = 33.3564 cm–1).
Figure 4.
Phonon band structure of B-TiO2, obtained using DFPT in combination with PBEsol and the three variants of the PAW potential centered at the Ti site of the crystal lattice: (a) Ti_std; (b) Ti_pv; and (c) Ti_sv. The phonon frequencies in the vertical axis are in THz (1 THz = 33.3564 cm–1).
Figure 5.
(a,b) Illustration of phonon dispersion of F- and P-TiO2, respectively, obtained using DFPT at the [PBEsol/Ti_std] level. The 4 × 4 × 4 (192 atoms) and 2 × 2 × 2 (96 atoms) supercells were utilized for the phonon calculations for F- and P-TiO2, respectively. Phonon frequencies in the vertical axis are in THz (1 THz = 33.3564 cm–1).
For F-TiO2, the optical phonons are unstable along the path Γ–L–W–X, even though they are stable at the Γ-point (Figure 5a). For P-TiO2, several low-frequency phonon modes are unstable throughout the Brillouin zone including zone center. The instability of the phonon modes is understood owing to their (negative) imaginary frequencies (Figure 5b). The result is not perplexed for pressure-driven crystal lattices, as are commonly observed to be dynamically unstable.
The rutile phase of TiO2 has 15 optical phonon modes. Of these, three are acoustic phonon modes. At Γ(k (0,0,0)), the optical phonons are representation given by 1A1g + 1A2g + 1A2u + 1B1g + 2B1u + 1B2g + 1Eg + 3 Eu, in which Eg (non-polar), B2g, B1g, and A1g are Raman-active and the A2g (non-polar) and B1u modes are infrared (IR) and Raman inactive silent modes.81 Each phonon mode E is twofold degenerate, and the phonon modes A2u and Eu are polar.
The four experimentally determined Raman-active modes of R-TiO2 appear at frequencies ω of 143 (B1g), 447 (Eg), 612 (A1g), and 826 cm–1 (B2g).81,82 The corresponding frequencies measured by Ma et al.83 center at 140.2, 445.8, 609.8, and 825.5 cm–1, respectively. They have also measured a broad and moderately strong band that peaks at 235 cm–1. This mode is composed of high-level anharmonicity and arises from two-phonon scattering. The same feature was observed by Tompsett et al.84
The frequencies of the four vibrational bands observed by Ma et al.83 and others81,82 (see above) are in close match with our results (all within 15 cm–1). For instance, they are centered at frequencies of 134.2 (B1g), 459.6 (Eg), 605.8 (A1g), and 810.2 cm–1 (B2g) with [PBEsol/Ti_std], respectively. When Ti_pv and Ti_sv were used, the two high-frequency modes, A1g and B2g, could be appreciably underestimated, except for the low-frequency phonon modes, B1g and Eg. This comparison leads to a conclusion that the overall nature of the phonon modes reported using coherent inelastic neutron scattering measurements81 matches well with those computed with Ti_std. The PBEsol frequency of the lowest (silent) phonon mode calculated using Ti_std (ω = 118.3 cm–1) is largely overestimated relative to that obtained using Ti_pv and Ti_sv (ω = 74.9 and 78.3 cm–1, respectively). The latter two are unquestionably underestimated by 38.1 and 34.7 cm–1 compared to experiment (ω = 113.0 cm–1, see Table 5).
R-TiO2 has four IR-active optical phonon modes. Schöche et al.25 have found that the IR-active modes (three Eu and one A2u) possess phonon frequencies of 188.6 ± 1.2, 379.3 ± 0.2, 500.5 ± 0.3, and 172.3 ± 1.9 cm–1, respectively. The first three are doubly degenerate modes and the latter one is a non-degenerate mode. Gervais and Piriou85 [Eagles86] have reported that the corresponding phonon mode frequencies should show at frequencies of 189 [183], 381.5 [388], 508 [500], and 172 [167] cm–1, respectively. The IR-active modes reported in different studies are not in disagreement with each other. However, the inelastic neutron scattering study did not observe the degenerate vibrational mode at 381.5 cm–1, assigned by Gervais and Piriou.85 Schöche et al.25 have assigned the two Eu modes to have frequencies of 188.61 and 365.74 cm–1, which are the first transverse optical (TO) and longitudinal optical (LO) modes, respectively.
Indeed, the experimental IR phonon modes are consistent with our PBEsol-based frequencies summarized in Table 5. This functional, together with Ti_std, has underestimated the lowest IR-active mode at 158.7 cm–1, compared to the experimental value of 183 (172.7) cm–1. Ti_pv and Ti_sv underestimated them further to appear at frequencies of 93.0 and 98.3 cm–1, respectively. Interestingly, Ti_sv has predicted the frequencies of the degenerate phonon mode at 98.4 and 98.2 cm–1, thus assigning them as nearly degenerate (which should not be!).
A discrepancy between experiment and calculation can be seen with the non-degenerate IR-active mode A2u. ω for this phonon mode was 180.5, 61.2, and 71.2 cm–1 with Ti_std, Ti_pv, and Ti_sv, respectively. The large difference in the wavenumbers with respect to the PAW potentials can be realized from the phonon dispersion curves shown in Figure 2a–c. From the shape of these curves, it is clear that the phonon modes below 2.5 THz are shallow-like (flattish) around high-symmetry M-point along the path Γ–M–X when the potentials Ti_pv and Ti_sv were used (Figure 2b,c, respectively). This may also mean that a slight distortion of the crystal lattice could cause softening of the phonon mode around the Γ–M–X path. This is not the case with the phonon modes calculated with Ti_std, in which the curves associated with the low-frequency optical modes are relatively steeper (Figure 2a), which match reasonably well with experiment.81
When PBE was used, the discrepancies between the phonon modes at low frequencies were very severe. From the data summarized in Table 5, one can see that the frequencies of all the phonon modes are underestimated for all the three PAW potentials utilized. Furthermore, Ti_pv and Ti_sv have predicted soft modes around the Γ-point that are linked with the three IR-active phonon modes and a silent mode; that is, they link with a doubly degenerate mode Eu, a non-degenerate mode A2u, and a non-degenerate B1u mode. The Ti_pv potential has predicted these phonons at frequencies of 62.8i, 106.1i, and 30.1i, respectively. Similarly, Ti_sv has predicted the corresponding frequencies at 54.4i, 98.8i, and 16.7i, respectively. These soft modes explain why the PBE-relaxed crystal lattice of R-TiO2 is dynamically unstable at the center of the zone, Γ.
Montanari and Harrison87,88 have demonstrated that the A2u (TO) mode becomes soft at the Γ-point if the crystal of R-TiO2 is expanded via LDA or GGA. This may be consistent with our PBE-level observation. The data suggest that the expansion of the rutile lattice by the two soft potentials softens the phonon modes associated with both the IR-active modes Eu and A2u (see columns 12 and 14 of Table 5). The lattice expansion is evidenced in Table 1, especially when comparing the lattice parameters evaluated with Ti_std to those with Ti_pv and Ti_sv.
Mitev et al. have also reported an imaginary frequency of 86i corresponding to the A2u (TO) phonon mode with the PBE functional using the CASTEP code.89 We could not reproduce their result because of differing cutoff criteria used in our calculation. The authors of the study have interpreted that the softening of the A2u (TO) mode at the Γ-point could be responsible for the incipient ferroelectric behavior of R-TiO2 and hence should play some role in the determination of large and strongly temperature-dependent dielectric constant.
We now clarify the origin of the spurious (soft) nature of the phonon modes predicted by the two soft potentials Ti_pv and Ti_sv. For this, we investigated the nature of radial dependence of the charge density in one dimension for each of the three PAW potentials. We have used the PBEsol-, PBEsol+U-, and HSEsol-relaxed lattices of R-TiO2. As shown in Figure 6a–d, the nature of the radial dependence of charge density is very similar regardless of the degree of exchange correlation incorporated in each of the DFT methods chosen. That is, the charge density in the region of space between bonded Ti and O atomic basins in R-TiO2, in the close vicinity of Ti, is steep (attractive) when Ti_std was applied and shallow (somehow repulsive!) when Ti_pv and Ti_sv were employed. These results suggest that the softness of the PAW potential (but certainly not the correlation method) is the sole cause of softening of the low-frequency optical phonon modes as a result of increased repulsion between core electrons in the Ti–O bonding region (cf. Figure 2a,b).
Figure 6.
1D plot of the radial dependence of charge density for R-TiO2, evaluated using several methods in combination with the three PAW potentials Ti_std, Ti_sv, and Ti_pv. Charge density calculations were performed on the (a) PBEsol lattice with PBEsol; (b) PBEsol+U lattice with PBEsol+U; (c) HSE06 lattice with PBEsol, and (d) HSEsol lattice with HSEsol. Values are in logarithmic scale.
A total of 18 phonon modes exist in case of the A-TiO2 lattice.84,90,91 Of these, 3 are acoustic modes and 15 are optical modes. The irreducible representation of the latter 15 modes is given by 1A1g + 1A2u + 2B1g + 1B2u + 3Eg + 2Eu. The modes with a subscript “u” are IR-active phonon modes except B2u (a silent mode), and those with a subscript “g” are Raman-active modes. The PBEsol- and PBE-calculated vibrational frequencies, Table 6, show that the high- and low-frequency phonon modes that are both IR and Raman active are either slightly overestimated or slightly underestimated but were well reproduced by PBEsol in combination with the standard potential Ti_std. In particular, the eight high-frequency Raman and IR modes are all within 15 cm–1 of the experimental values. The three doubly degenerate IR and Raman modes (Eu and Eg) predicted at frequencies of 240.8 (Eu), 164.7 (Eg), and 151.4 (Eg) cm–1 are underestimated or overestimated but all within 35 cm–1 of the experimental wavenumbers (Table 6).84,90 The degree of underestimation of high-frequency phonon vibrations was severe especially with the Ti-centered potentials Ti_pv and Ti_sv; the same behavior was also observed for R-TiO2 (Table 5). As mentioned above, the underestimation should not be due to the short Ti–O bond distance in the crystal lattice calculated using the two soft potentials. We confirmed this by calculating the frequencies of phonon modes with Ti_std, Ti_pv, and Ti_sv using bulk geometries of R- and A-TiO2 obtained with Ti_pv or Ti_sv. As shown in Tables S1 and S2, the phonon frequencies with Ti_std are always higher than those calculated with Ti_pv and Ti_sv, a feature that was emanated irrespective of the type of the crystal lattice used. From this, it is clear that the softer the potential, the shallower the charge density profile in the close vicinity of Ti along the Ti–O bond of TO2, thus softening (lowering) the frequency of phonon vibrations around the zone center. However, the phonon mode softening is less pronounced at the Γ-point for A-TiO2 compared to that found for R-TiO2. The performance of the PBE functional is homologous with that of PBEsol for A-TiO2, but the phonon frequencies are largely underestimated compared to experiment (Table 6).
For B-TiO2, the number of IR- and Raman-active modes is greater than that found in R- and A-TiO2; this is attributed to the greater number of atoms (24) responsible for the unit cell of the brookite lattice. Because of this, and because B-TiO2 has a lower orthorhombic Pbca symmetry, there are 36, 24, 9, and 3 Raman-active (A1g, B1g, B2g, and B3g), IR-active (B1u, B2u, B3u), silent (A1u), and acoustic vibrational phonon modes, respectively. They are represented by the irreducible representations at Γ: 9A1g + 9B1g + 9B2g + 9B3g, 8B1u + 8B2u + 8B3u, 9A1u, and B1u + B2u + B3u. The frequencies of the PBEsol- and PBE-calculated phonon modes compared in Table S3 indicate that the phonon modes evaluated with PBEsol with potential Ti_stv are in better agreement with experiment.84,94
3.3. Dependence of the Magnitude of Relative Dielectric Permittivity on PAW Potentials and DFT Methods
Since rutile and anatase TiO2 are a uniaxial crystal and brookite TiO2 is a biaxial crystal,95 there are two and three independent non-zero components both for the ionic and electronic parts of the dielectric permittivity, respectively. The average of the former property, εr(0), was added to that of the latter property, εr(∞), so as to calculate the relative static dielectric permittivity, εr, for all the three phases of TiO2. They are summarized in Table 7, together with the experimentally reported values wherever available.
Table 7. Comparison of Ionic (Low-Frequency) and Electronic (High-Frequency) Components (εij(0) and εij(∞), Respectively) of the Relative Dielectric Permittivity Tensor εr of R-TiO2, Obtained Using Three Cutoff Energies in Combination with the PBEsol and Three PAW Potentials Ti_std, Ti_pv, and Ti_sva.
| PAW potential | cutoff energy/eV | εij(0) |
εij(∞) |
εr | ||||||
|---|---|---|---|---|---|---|---|---|---|---|
| εxx(0) | εyy(0) | εzz(0) | εr(0) | εxx(∞) | εyy(∞) | εzz(∞) | εr(∞) | εr(0) + εr(∞) | ||
| R-TiO2 | ||||||||||
| Ti_std | 520.0 | 124.0 | 124.0 | 163.0 | 137.0 | 8.0 | 8.0 | 9.6 | 8.5 | 145.5 |
| 600.0 | 123.4 | 123.4 | 162.3 | 136.4 | 8.0 | 8.0 | 9.6 | 8.5 | 144.9 | |
| 700.0 | 122.9 | 122.9 | 161.3 | 135.7 | 8.0 | 8.0 | 9.6 | 8.5 | 144.2 | |
| Ti_pv | 520.0 | 354.1 | 354.1 | 1135.2 | 614.5 | 7.6 | 7.6 | 9.1 | 8.1 | 622.6 |
| 600.0 | 366.0 | 366.0 | 1257.8 | 663.3 | 7.6 | 7.6 | 9.1 | 8.1 | 671.3 | |
| 700.0 | 370.3 | 370.3 | 1327.0 | 689.2 | 7.6 | 7.6 | 9.1 | 8.1 | 697.3 | |
| Ti_sv | 520.0 | 296.2 | 296.2 | 732.9 | 441.8 | 7.5 | 7.5 | 8.9 | 8.0 | 449.9 |
| 600.0 | 309.1 | 309.1 | 799.3 | 472.5 | 7.5 | 7.5 | 8.9 | 8.0 | 480.5 | |
| 700.0 | 327.1 | 327.1 | 966.8 | 540.3 | 7.5 | 7.5 | 8.9 | 7.9 | 548.2 | |
| expt.14c | 111 | 111 | 257 | 159.7 | ||||||
| expt.25 | 84.69 | 84.69 | 152.97 | 107.45 | 5.96 | 5.96 | 7.16 | 6.36 | 113.81 | |
| expt.85 | 84.02 | 84.02 | 153.83 | 107.29 | 5.91 | 5.91 | 7.19 | 6.34 | 113.63 | |
| expt.89 | 81.8 | 81.8 | 167.1 | 110.23 | 6.00 | 6.00 | 7.80 | 6.60 | 116.83 | |
| expt.25 | 87.0 | 87.0 | 163.0 | 112.33 | 6.12 | 6.12 | 7.63 | 6.62 | 118.96 | |
| expt.25 | 89.8 | 89.8 | 166.7 | 115.43 | 6.32 | 6.32 | 7.80 | 6.81 | 122.25 | |
| expt.25 | 86 | 86 | 170 | 114.00 | 6.84 | 6.84 | 8.43 | 7.37 | 121.37 | |
| A-TiO2 | ||||||||||
| Ti_std | 700 | 44.1 | 44.1 | 21.2 | 36.4 | 7.4 | 7.4 | 6.7 | 7.1 | 43.6 |
| Ti_pv | 700 | 54.6 | 54.6 | 23.6 | 44.3 | 7.0 | 7.0 | 6.4 | 6.8 | 51.1 |
| Ti_sv | 700 | 52.4 | 52.4 | 23.3 | 42.7 | 6.9 | 6.9 | 6.3 | 6.7 | 49.4 |
| expt.92 | 45.1 | 45.1 | 22.7 | 37.6 | 5.8 | 5.8 | 5.4 | 5.7 | 43.3 | |
| B-TiO2 | ||||||||||
| Ti_std | 700 | 48.2 | 50.7 | 45.0 | 49.4 | 7.6 | 8.0 | 7.4 | 7.7 | 57.2 |
| Ti_pv | 700 | 52.8 | 70.1 | 47.3 | 56.7 | 7.0 | 7.7 | 7.0 | 7.2 | 64.0 |
| Ti_sv | 700 | 52.4 | 68.6 | 47.9 | 56.3 | 6.9 | 7.6 | 6.9 | 7.1 | 63.4 |
| Ref (23)b | 64.1 | 86.7 | 55.2 | 68.7 | 6.8 | 7.5 | 6.8 | 7.0 | 75.7 | |
| Ref (23)b | 50.2 | 58.2 | 50.0 | 52.8 | 7.0 | 7.5 | 6.9 | 7.1 | 59.9 | |
| expt.97,98 | 93 | 78 | ||||||||
Both low- and room-temperature experimental values of εr reported by various authors are included.
Reported at the LDA and GGA levels, using Quantum Expresso software.
Low-temperature data refer T = 1.6 K.
As outlined in the Introduction section, the major controversy of TiO2 lies in the experimental determination and proposed relative dielectric permittivity values of R-TiO2. Some have reported the colossal nature of the ionic and total dielectric permittivities, and others reported appreciably large numbers for the same property for the same system. For instance, Nicolini, in 1952, has proposed a colossal dielectric permittivity of R-TiO2, with εr approximately 10,000 for ceramic R-TiO2. This value is close to the range 100–10,000 reported by Parker and Wasilik,12 as well as by Chu.13 Parker and Wasilik12 have reported a colossal dielectric permittivity in hydrogenated and reduced oxygen-rich rutile single crystalline TiO2 [e.g., Nicolini (εr = 10,000) and Perker et al. (εr = (10,000–30,000)]. Later, in 1961, Parker reported values of 111 and 257 for ε⊥(0) = εxx(0) = εyy(0) and ε||(0) = εzz(0) along the crystallographic a–/b- and c-axes of R-TiO2 at low temperature, 1.6 K, respectively; these were 86 (58) and 170 (97) at 300 (1000) K, respectively.14 Indeed, these results were in agreement with Samara and Peercy [εr values along the a- and c-axes were 89.8 (114.9) and 166.7 (251) at 296 K (4 K), respectively].24 Bonkerud et al.16 have recently argued that they fail to confirm the results of Nicolini11 and Parker and Wasilik,12 even though they could reconfirm the results individually published by Parker14 and Samara and Peercy.24 Therefore, the anomalous dielectric permittivity reported by various authors has continuously been debated since some of the reported results above are the repercussion of any incorrectly designed experiment or an incorrect interpretation.
The trend in our PBEsol-based values of ε⊥(0) (ε⊥(0) = [εxx(0) + εyy(0)]/2) and ε||(0) (ε||(0) = εzz(0)) (Table 7) calculated with Ti_std is in agreement with those of Samara and Peercy,24 and Parker,14 as well as of Bonkerud et al.16 Because Ti_pv and Ti_sv underestimate wavenumbers of the low-frequency phonons (Table 5) compared to Ti_std, they are the cause of spuriously large εr(0) for R-TiO2. The feature is persistent regardless of the three energy cutoff values invoked during our calculation (Table 7). The discrepancy is understandable since the calculation of the dielectric permittivity due to ions strongly relies on the nature of the PAW potential responsible for the determination of the positive eigenvalues of the dynamical matrix, and the low-frequency phonon mode is the main factor in determining the net magnitude of εr(0). The latter is not surprising given that the εr(0) is inversely proportional to the square of the phonon frequency (εr(0) α 1/ω2),96 so that the smaller the phonon frequency, the larger the value of εr(0). The notion is consistent with the rule of thumb that the combination of a higher frequency and a lower polarization factor should result in a lower dielectric constant in any specific direction. By contrast, εr(∞) is nearly invariant with respect to the cutoff energies used.
Table S4 summarizes εr(0) and εr(∞) of R-TiO2, calculated using PBE. The data reveal that neither of the three PAW potentials, in combination with the three energy cutoffs utilized, are suitable in reproducing experimental εr(0). The εr(0) was predicted to be spuriously large (εr(0 in the range 415–421) with Ti_std and unphysical with Ti_sv and Ti_pv. The εr(∞) values (e.g., 7.9, 8.1, and 8.6 with Ti_sv, Ti_pv, and Ti, respectively) are slightly overestimated relative to the experimentally suggested range (εr(∞) between 6.3 and 7.4).
We have calculated εr(0) and εr(∞) using PBE+U and PBEsol+U (U = 3.0 eV). This was instigated with the perception that the inclusion of the Hubbard on-site effective potential U centered on the d-orbital of the Ti atom provides the expected chemistry of localized electrons and hole polarons and their energetics for rutile, anatase, and brookite TiO2 and that the GGA alone fails to do so.99−101 The εr(0) and εr(∞) values in Table S5, show, first, that εij(0) and εij(∞) computed with a specific functional are largely unaffected changing the cutoff energy for a particular PAW potential. Second, when moving from Ti_std through Ti_pv to Ti_sv, there is a systematic increase in the magnitude of both εij(0) and εij(∞) for a specific functional and cutoff energy. Third, the GGA-based mean value of εij(0) is roughly two or three times smaller when computed with DFT+U (εij(0) in the range 44–69 (35–54) with PBE+U (PBEsol+U)), so there is a sharp disagreement between the calculated and experimental values of εr(0). Similarly, the DFT+U-calculated difference between the parallel and perpendicular components of εij(0) is also unphysically small compared to the notable difference deduced from the corresponding experimental values. Fourth, the εij(∞) are almost oblivious with respect to the two DFT+U methods applied for any particular potential and cutoff energy and are close to the experimental range of reported values (6.3–7.4, see Table 7). The only feature that manifests with the combination between the Hubbard term and DFT is the disappearance of all soft (low-frequency IR-active) phonon modes, meaning that the combination does not predict the dynamic instability of the R-TiO2 lattice predicted by Ti_pv and Ti_sv in combination with PBE (Table 6).
Tables 6 and S4 list the PBEsol- and PBE-calculated εr(0), εr(∞), and εr values for A- and B-TiO2, respectively. For A-TiO2, PBEsol/Ti_std shows the best prediction of εr (εr = 43.6 with PBEsol/Ti_std and 43.3 with experiment). This is compared to those predicted byPBEsol/Ti_pv and PBEsol/Ti_sv that have overestimated εr by 7.8 and 6.1, respectively. The corresponding overestimations by PBE were 10.5, 20.0, and 18.5 for Ti_std, Ti_pv, and Ti_sv, respectively. These values are indeed large, appearing due to numerically small frequencies associated with the low-frequency phonon modes. The lack of detailed experimental data for B-TiO2 did not allow us to estimate the performance of PAW potentials and correlation methods in predicting the accuracy of εr values.
Our result suggests that the εr value for the R-TiO2 lattice is larger than that of the A- and B-TiO2 lattices (εr(R) < εr(B) < εr(A)). For mass density, ρ(A) < ρ(B) < ρ(R).
The ionic and optical dielectric constants for the two cubic phases of TiO2 are listed in Table S6. The former could not be determined correctly and are unphysical, even though the large values of εr(0) indicate the presence of any ferroelectric transition. The obvious reason for this is that the frequencies of the three low-frequency phonon modes were too small for F-TiO2, and many of them were negative for P-TiO2. The optical dielectric constant was close to 10.0–11.0 for both the systems. They are larger than those calculated for R-, A-, and B-TiO2.
We found that ε⊥(0) < ε||(0) for R-TiO2 with PBEsol; it is reversed (ε⊥(0) > ε||(0)) for A-TiO2. The anisotropic nature of εr(0) is similar to that found for the electronic part of εr (i.e., ε⊥(∞) < ε||(∞) for R-TiO2; ε⊥(∞) < ε||(∞) for A-TiO2), showing that the polarization along the c-direction is strongest for the former than for the latter system. This may indicate that the excitonic wavefunctions are extended in one direction than the remaining two, which are expected of low-dimensional excitons.21 As mentioned by Mikami et al.,90 Wemple102 has observed that the difference between the dielectric permittivity of the two TiO2 phases was due to the smaller anion density of anatase TiO2.
We have calculated the dielectric permittivity and phonon frequencies for R-TiO2 using the relaxed lattices of HSEsol and HSE06, in which the frozen phonon approximation was invoked. The results are summarized in Tables S7 and S8. As observed with GGA (see above), the HSE results show that the magnitude of ε(0) depends significantly on the type of the PAW potential employed. Overall, the ε(0) value of [HSE06/Ti_std] agrees well with low-temperature experimental results discussed above.14 The ε(∞) values are also in better agreement with experiment due to a systematic improvement of the band gap with the HSE methods.
3.4. Dependence of Low-Frequency Phonon Modes and Static Dielectric Constant on Hydrostatic Pressure
We have examined the isotropic pressure dependence of the geometry, zone-center phonon modes, and relative dielectric constant of R-TiO2. Figure 7a–c shows the dependence of the harmonic frequencies of the lowest five Γ-center phonon modes on the applied pressure of R-TiO2. Except for the degenerate IR-active mode Eu (ω = 375.0 cm–1 with PBEsol/Ti_std) and the non-degenerate Raman-active mode B1g (ω = 134.2 cm–1 with PBEsol/Ti_std), the phonon frequencies of the other three IR-active bands with mode symmetry A2u, B1u, and Eu are affected by applied pressure. Figure 7 That is, ferroelectricity can be induced to the rutile lattice of TiO2 via negative hydrostatic pressure, and the most sensitive ferroelectric mode is the TO mode A2u. The other two modes (B1u and Eu) soften as the pressure increases, which causes further expansion of the rutile lattice. The softness of the Eu mode, which is the analogue of the A2u mode in the a–b plane, was previously observed.87
Figure 7.
(Left) PBEsol level external (hydrostatic) pressure dependence of the lowest five Γ-centered phonon modes of R-TiO2, obtained with PAW potentials: (a) Ti_std; (b) Ti_pv; (c) Ti_sv. (Right) The dependence of dielectric permittivity (ionic) on applied pressure, obtained PAW potentials: (d) Ti_std; (e) Ti_pv; and (f) Ti_sv. The labeling of the phonon modes is consistent with the data shown in Table 5.
The three PAW potentials give qualitatively similar insight into the development of the three ferroelectric modes, thus leading to the creativity of lattice instability. However, the application of the two soft potentials Ti_pv and Ti_sv confirms that the phonon frequency of the A2u mode vanishes faster than that of the B1u and Eu modes and becomes more negative when the lattice is expanded. That is, for the PAW potentials Ti_pv and Ti_sv, the frequency of the A2u mode vanishes at hydrostatic pressure at about −1.0 GPa, while it is about −9.3 GPa with Ti_std. These results explain why the static dielectric constant increases dramatically and more rapidly with Ti_pv and Ti_sv compared to Ti_std, which may be captured from Figure 7d–f. Also, the anisotropy in the in- and out-of-plane dielectric constants predicted with the two soft potentials are quite significant than that predicted using Ti_std. The component of dielectric constant parallel to the c-axis is increased prominently compared to the in-plane dielectric constants for all the three occasions exploited, meaning that an enhancement of the dielectric constant can be made possible via the expansion of the rutile lattice of TiO2.
4. Conclusions
This study has utilized three most commonly used variants of the PAW potential for Ti to examine the physical properties of TiO2. In particular, the combined application of these pseudopotentials with PBE, PBEsol, and PBEsol+U has enabled us to assess the accuracy of the ground-state lattice properties, phonon vibrations, and relative static dielectric permittivity. We have shown that the change in the size of the potential can have a non-negligible effect on the temperature-dependent lattice geometry of all the five phases of TiO2. Their effect is particularly pronounced when the low frequencies of optical phonon modes, as well as the ionic contribution to the relative dielectric permittivity, were compared. The origin of the underlying difference was uncovered when radial dependence of the charge density was analyzed around the close vicinity of Ti. We have argued that the radial dependence of charge density is relatively steep in the close vicinity of Ti when the standard potential Ti_std was invoked; it was shallow when the two soft potentials Ti_pv and Ti_sv were invoked. Because of this nature of the charge density profile, Ti_std has produced higher phonon frequencies (and relatively accurate εr(0)) than that predicted with Ti_pv and Ti_sv. Higher-level calculations using HSEsol and HSE06 have shown to have some improvement on the accuracy of the dielectric properties, despite being computationally intensive. In addition, we have shown that ferroelectric instability can be induced in the R-TiO2 lattice by applying negative hydrostatic pressure. At least three low-lying phonon modes (A2u, B1u, and Eu) of R-TiO2 are observed to be sensitive not only to the onset of ferroelectric phase transitions but also to the hard and soft nature of the three PAW potentials. Finally, we recommend using the standard PAW potential, Ti_std, for modeling the ground-state lattice properties and phonon-related features of TiO2-based materials, rather than choosing a softer PAW potential.
Acknowledgments
This work was entirely conducted using the computation and laboratory facilities provided by the University of Nagoya and the Research Center for Computational Science,, Okazaki, Japan (Project: 22-IMS-C099). P.R.V. has been appointed as a Reader of the University of Witwatersrand (SA). R.A. thanks Drs. Hiroki Taniguchi and Koichi Hayashi for their fruitful discussions.
Data Availability Statement
This research did not report any data.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.3c02038.
Comparison of the IR and Raman frequencies ω (cm–1) and IR intensities I (km/mol) obtained using PBEsol in combination with PAW potentials Ti_std, Ti_pv, and Ti_sv centered on Ti of R-TiO2; comparison of the IR and Raman frequencies ω (cm–1) and IR intensities I (km/mol) obtained using PBEsol in combination with PAW potentials Ti_std, Ti_pv, and Ti_sv centered on Ti of R-TiO2; comparison of the IR and Raman frequencies and IR intensities obtained using PBEsol and PBE functionals, in conjunction with PAW potentials Ti_std, Ti_pv, and Ti_sv centered on Ti of B-TiO2; comparison of the PBE level electronic and ionic components (εij(0) and εij(∞), respectively) of the relative dielectric permittivity tensor εij for R- and A-TiO2, computed in conjunction with three variants of the PAW potential, viz., Ti_std, Ti_pv, and Ti_sv; comparison of the PBE+U and PBEsol+U (U = 3.0 eV) level electronic and ionic components (εij(0) and εij(∞), respectively) of the relative dielectric permittivity tensor εr for R-TiO2, computed in conjunction with three variants of the PAW potential, viz., Ti_std, Ti_pv, and Ti_sv; comparison of the PBEsol level electronic and ionic components (εij(0) and εij(∞), respectively) of the relative dielectric permittivity tensor εij for F- and P-TiO2, computed in conjunction with three variants of the PAW potential, viz., Ti_std, Ti_pv, and Ti_sv; ionic (low-frequency) and electronic (high-frequency) components (εij(0) and εij(∞), respectively) of the dielectric permittivity tensor εr of R-TiO2, obtained using HSEsol and HSE06 in combination with the three PAW potentials Ti_std; and comparison of the IR and Raman frequencies ω (cm–1) obtained using HSEsol and HSE06 in combination with PAW potentials Ti_std centered on Ti of R-TiO2 (PDF)
Author Contributions
Conceptualization, P.R.V.; computation, data acquisition, formal analysis, and investigation, P.R.V.; computation of phonon modes of R-, A-, and B-TiO2: R.A.; computational verifications and critical discussions, V.A.D. and Y.M.; supervision, R.A.; drawing of figures, preparation of tables, and writing—original draft, P.R.V.; writing—review and editing, P.R.V. and R.A.; All authors have read and agreed to the published version of the manuscript.
This research was funded by JSPS Grant-in-Aid for Transformative Research Areas (A); Grant numbers: 21H05560 and 20H05883.
The authors declare no competing financial interest.
Supplementary Material
References
- Muscat J.; Swamy V.; Harrison N. M. First-Principles Calculations of the Phase Stability of TiO2. Phys. Rev. B 2002, 65, 224112 10.1103/PhysRevB.65.224112. [DOI] [Google Scholar]
- Daβler A.; Feltz A.; Jung J.; Ludwig W.; Kaisersberger E. Characterization of Rutile and Anatase Powders by Thermal Analysis. J. Therm. Anal. 1988, 33, 803–809. 10.1007/BF02138591. [DOI] [Google Scholar]
- Dambournet D.; Belharouak I.; Amine K. Tailored Preparation Methods of TiO2 Anatase, Rutile, Brookite: Mechanism of Formation and Electrochemical Properties. Chem. Mater. 2010, 22, 1173–1179. 10.1021/cm902613h. [DOI] [Google Scholar]
- Li J.-G.; Ishigaki T. Brookite⃗rutile Phase Transformation of TiO2 Studied with Monodispersed Particles. Acta Mater. 2004, 52, 5143–5150. 10.1016/j.actamat.2004.07.020. [DOI] [Google Scholar]
- Pascual J.; Camassel J.; Mathieu H. Fine Structure in the Intrinsic Absorption Edge of TiO2. Phys. Rev. B 1978, 18, 5606–5614. 10.1103/PhysRevB.18.5606. [DOI] [Google Scholar]
- Allen N. S.; Mahdjoub N.; Vishnyakov V.; Kelly P. J.; Kriek R. J. The Effect of Crystalline Phase (Anatase, Brookite and Rutile) and Size on the Photocatalytic Activity of Calcined Polymorphic Titanium Dioxide (TiO2). Polym. Degrad. Stab. 2018, 150, 31–36. 10.1016/j.polymdegradstab.2018.02.008. [DOI] [Google Scholar]
- Tang H.; Lévy F.; Berger H.; Schmid P. E. Urbach Tail of Anatase TiO2. Phys. Rev. B 1995, 52, 7771–7774. 10.1103/PhysRevB.52.7771. [DOI] [PubMed] [Google Scholar]
- Mattsson A.; Österlund L. Adsorption and Photoinduced Decomposition of Acetone and Acetic Acid on Anatase, Brookite, and Rutile TiO2 Nanoparticles. J. Phys. Chem. C 2010, 114, 14121–14132. 10.1021/jp103263n. [DOI] [Google Scholar]
- Carp O.; Huisman C. L.; Reller A. Photoinduced Reactivity of Titanium Dioxide. Prog. Solid State Chem. 2004, 32, 33–177. 10.1016/j.progsolidstchem.2004.08.001. [DOI] [Google Scholar]
- Gyanan; Mondal S.; Kumar A. Tunable Dielectric Properties of TiO2 Thin Film Based MOS Systems for Application in Microelectronics. Superlattices Microstruct. 2016, 100, 876–885. 10.1016/j.spmi.2016.10.054. [DOI] [Google Scholar]
- Nicolini L. A New Dielectric Material. Nature 1952, 170, 938–938. 10.1038/170938a0.13013274 [DOI] [Google Scholar]
- Parker R. A.; Wasilik J. H. Dielectric Constant and Dielectric Loss of TiO2 (Rutile) at Low Frequencies. Phys. Rev. 1960, 120, 1631–1637. 10.1103/PhysRev.120.1631. [DOI] [Google Scholar]
- Chu C. W. New Ordered Phases of Slightly Reduced Rutile and Their Sharp Dielectric Absorptions at Low Temperature. Phys. Rev. B 1970, 1, 4700–4708. 10.1103/PhysRevB.1.4700. [DOI] [Google Scholar]
- Parker R. A. Static Dielectric Constant of Rutile (TiO2), 1.6-1060 °K. Phys. Rev. 1961, 124, 1719–1722. 10.1103/PhysRev.124.1719. [DOI] [Google Scholar]
- Peercy P. S.; Morosin B. Pressure and Temperature Dependences of the Raman-Active Phonons in SnO2. Phys. Rev. B 1973, 7, 2779–2786. 10.1103/PhysRevB.7.2779. [DOI] [Google Scholar]
- Bonkerud J.; Zimmermann C.; Weiser P. M.; Vines L.; Monakhov E. V. On the Permittivity of Titanium Dioxide. Sci. Rep. 2021, 11, 12443. 10.1038/s41598-021-92021-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tse M.-Y.; Wei X.; Hao J. High-Performance Colossal Permittivity Materials of (Nb + Er) Co-Doped TiO2 for Large Capacitors and High-Energy-Density Storage Devices. Phys. Chem. Chem. Phys. 2016, 18, 24270–24277. 10.1039/C6CP02236G. [DOI] [PubMed] [Google Scholar]
- Kawarasaki M.; Tanabe K.; Terasaki I.; Fujii Y.; Taniguchi H. Intrinsic Enhancement of Dielectric Permittivity in (Nb + In) Co-Doped TiO2 Single Crystals. Sci. Rep. 2017, 7, 5351. 10.1038/s41598-017-05651-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Elshahawy A. M.; Elkatlawy S. M.; Shalaby M. S.; Guan C.; Wang J. Surface-Engineered TiO2 for High-Performance Flexible Supercapacitor Applications. J. Electron. Mater. 2023, 52, 1347–1356. 10.1007/s11664-022-10084-0. [DOI] [Google Scholar]
- Osada M.; Kobayashi M.; Kakihana M. Enhanced Dielectric Response Induced by Controlled Morphology in Rutile TiO2 Nanocrystals. J. Ceram. Soc. Jpn. 2013, 121, 593–597. 10.2109/jcersj2.121.593. [DOI] [Google Scholar]
- Thilagam A.; Simpson D. J.; Gerson A. R. A First-Principles Study of the Dielectric Properties of TiO2 polymorphs. J. Phys.: Condens. Matter 2011, 23, 025901 10.1088/0953-8984/23/2/025901. [DOI] [PubMed] [Google Scholar]
- Mahmood T.; Malik H.; Batool R.; Perveen Z.; Saleemi F.; Rasheed H.; Saeed M. A.; Cao C.; Rizwan M. Elastic, Electronic and Optical Properties of Anatase TiO2 under Pressure: A DFT Approach. Chin. J. Phys. 2017, 55, 1252–1263. 10.1016/j.cjph.2017.05.029. [DOI] [Google Scholar]
- Shojaee E.; Abbasnejad M.; Saeedian M.; Mohammadizadeh M. R. First-Principles Study of Lattice Dynamics of TiO2 in Brookite and Cotunnite Structures. Phys. Rev. B 2011, 83, 174302 10.1103/PhysRevB.83.174302. [DOI] [Google Scholar]
- Samara G. A.; Peercy P. S. Pressure and Temperature Dependence of the Static Dielectric Constants and Raman Spectra of TiO2 (Rutile). Phys. Rev. B 1973, 7, 1131–1148. 10.1103/PhysRevB.7.1131. [DOI] [Google Scholar]
- Schöche S.; Hofmann T.; Korlacki R.; Tiwald T. E.; Schubert M. Infrared Dielectric Anisotropy and Phonon Modes of Rutile TiO2. J. Appl. Phys. 2013, 113, 164102. 10.1063/1.4802715. [DOI] [Google Scholar]
- Ke S.; Li T.; Ye M.; Lin P.; Yuan W.; Zeng X.; Chen L.; Huang H. Origin of Colossal Dielectric Response in (In + Nb) Co-Doped TiO2 Rutile Ceramics: A Potential Electrothermal Material. Sci. Rep. 2017, 7, 10144. 10.1038/s41598-017-10562-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhao C.; Wu J. Effects of Secondary Phases on the High-Performance Colossal Permittivity in Titanium Dioxide Ceramics. ACS Appl. Mater. Interfaces 2018, 10, 3680–3688. 10.1021/acsami.7b18356. [DOI] [PubMed] [Google Scholar]
- Hu W.; Lau K.; Liu Y.; Withers R. L.; Chen H.; Fu L.; Gong B.; Hutchison W. Colossal Dielectric Permittivity in (Nb+Al) Codoped Rutile TiO2 Ceramics: Compositional Gradient and Local Structure. Chem. Mater. 2015, 27, 4934–4942. 10.1021/acs.chemmater.5b01351. [DOI] [Google Scholar]
- Lee B.; Lee C.; Hwang C. S.; Han S. Influence of Exchange-Correlation Functionals on Dielectric Properties of Rutile TiO2. Curr. Appl. Phys. 2011, 11, S293–S296. 10.1016/j.cap.2010.11.104. [DOI] [Google Scholar]
- Dou M.; Persson C. Comparative Study of Rutile and Anatase SnO2 and TiO2: Band-Edge Structures, Dielectric Functions, and Polaron Effects. J. Appl. Phys. 2013, 113, 083703 10.1063/1.4793273. [DOI] [Google Scholar]
- Skone J. H.; Govoni M.; Galli G. Self-Consistent Hybrid Functional for Condensed Systems. Phys. Rev. B 2014, 89, 195112 10.1103/PhysRevB.89.195112. [DOI] [Google Scholar]
- Gerosa M.; Bottani C. E.; Di Valentin C.; Onida G.; Pacchioni G. Accuracy of Dielectric-Dependent Hybrid Functionals in the Prediction of Optoelectronic Properties of Metal Oxide Semiconductors: A Comprehensive Comparison with Many-Body GW and Experiments. J. Phys.: Condens. Matter 2018, 30, 044003 10.1088/1361-648X/aa9725. [DOI] [PubMed] [Google Scholar]
- Gerosa M.; Bottani C. E.; Caramella L.; Onida G.; Di Valentin C.; Pacchioni G. Electronic Structure and Phase Stability of Oxide Semiconductors: Performance of Dielectric-Dependent Hybrid Functional DFT, Benchmarked against GW Band Structure Calculations and Experiments. Phys. Rev. B 2015, 91, 155201 10.1103/PhysRevB.91.155201. [DOI] [Google Scholar]
- Lee C.; Ghosez P.; Gonze X. Lattice Dynamics and Dielectric Properties of Incipient Ferroelectric TiO2 Rutile. Phys. Rev. B 1994, 50, 13379–13387. 10.1103/PhysRevB.50.13379. [DOI] [PubMed] [Google Scholar]
- Wehinger B.; Bosak A.; Jochym P. T. Soft Phonon Modes in Rutile TiO2. Phys. Rev. B 2016, 93, 014303 10.1103/PhysRevB.93.014303. [DOI] [Google Scholar]
- Perdew J. P.; Ruzsinszky A.; Csonka G. I.; Vydrov O. A.; Scuseria G. E.; Constantin L. A.; Zhou X.; Burke K. Restoring the Density-Gradient Expansion for Exchange in Solids and Surfaces. Phys. Rev. Lett. 2008, 100, 136406 10.1103/PhysRevLett.100.136406. [DOI] [PubMed] [Google Scholar]
- Perdew J. P.; Burke K.; Ernzerhof M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. 10.1103/PhysRevLett.77.3865. [DOI] [PubMed] [Google Scholar]
- Perdew J. P.; Burke K.; Ernzerhof M. Generalized Gradient Approximation Made Simple [Phys. Rev. Lett. 77, 3865 (1996)]. Phys. Rev. Lett. 1997, 78, 1396–1396. 10.1103/PhysRevLett.78.1396. [DOI] [PubMed] [Google Scholar]
- Dudarev S. L.; Botton G. A.; Savrasov S. Y.; Humphreys C. J.; Sutton A. P. Electron-Energy-Loss Spectra and the Structural Stability of Nickel Oxide: An LSDA+U Study. Phys. Rev. B 1998, 57, 1505–1509. 10.1103/PhysRevB.57.1505. [DOI] [Google Scholar]
- Liechtenstein A. I.; Anisimov V. I.; Zaanen J. Density-Functional Theory and Strong Interactions: Orbital Ordering in Mott-Hubbard Insulators. Phys. Rev. B 1995, 52, R5467–R5470. 10.1103/PhysRevB.52.R5467. [DOI] [PubMed] [Google Scholar]
- Gajdoš M.; Hummer K.; Kresse G.; Furthmüller J.; Bechstedt F. Linear Optical Properties in the Projector-Augmented Wave Methodology. Phys. Rev. B 2006, 73, 045112 10.1103/PhysRevB.73.045112. [DOI] [Google Scholar]
- Wu X.; Vanderbilt D.; Hamann D. R. Systematic Treatment of Displacements, Strains, and Electric Fields in Density-Functional Perturbation Theory. Phys. Rev. B 2005, 72, 035105 10.1103/PhysRevB.72.035105. [DOI] [Google Scholar]
- Pérez-Osorio M. A.; Milot R. L.; Filip M. R.; Patel J. B.; Herz L. M.; Johnston M. B.; Giustino F. Vibrational Properties of the Organic–Inorganic Halide Perovskite CH3NH3PbI3 from Theory and Experiment: Factor Group Analysis, First-Principles Calculations, and Low-Temperature Infrared Spectra. J. Phys. Chem. C 2015, 119, 25703–25718. 10.1021/acs.jpcc.5b07432. [DOI] [Google Scholar]
- Dharmale N.; Chaudhury S.; Mahamune R.; Dash D. Comparative Study on Structural, Electronic, Optical and Mechanical Properties of Normal and High Pressure Phases Titanium Dioxide Using DFT. Mater. Res. Express 2020, 7, 054004 10.1088/2053-1591/ab8d5c. [DOI] [Google Scholar]
- Swamy V.; Muddle B. C. Ultrastiff Cubic TiO2 Identified via First-Principles Calculations. Phys. Rev. Lett. 2007, 98, 035502 10.1103/PhysRevLett.98.035502. [DOI] [PubMed] [Google Scholar]
- Liang Y.; Zhang B.; Zhao J. Mechanical Properties and Structural Identifications of Cubic TiO2. Phys. Rev. B 2008, 77, 094126 10.1103/PhysRevB.77.094126. [DOI] [Google Scholar]
- Niu M.; Cheng D.; Cao D. Fluorite TiO2(111) Surface Phase for Enhanced Visible-Light Solar Energy Conversion. J. Phys. Chem. C 2014, 118, 20107–20111. 10.1021/jp504818j. [DOI] [Google Scholar]
- Abbasnejad M.; Shojaee E.; Mohammadizadeh M. R.; Alaei M.; Maezono R. Quantum Monte Carlo Study of High-Pressure Cubic TiO2. Appl. Phys. Lett. 2012, 100, 261902. 10.1063/1.4730608. [DOI] [Google Scholar]
- Kresse G.; Hafner J. Norm-Conserving and Ultrasoft Pseudopotentials for First-Row and Transition Elements. J. Phys.: Condens. Matter 1994, 6, 8245–8257. 10.1088/0953-8984/6/40/015. [DOI] [Google Scholar]
- Kresse G.; Joubert D. From Ultrasoft Pseudopotentials to the Projector Augmented-Wave Method. Phys. Rev. B 1999, 59, 1758–1775. 10.1103/PhysRevB.59.1758. [DOI] [Google Scholar]
- Kresse G.; Hafner J. Ab Initio Molecular Dynamics for Liquid Metals. Phys. Rev. B 1993, 47, 558–561. 10.1103/PhysRevB.47.558. [DOI] [PubMed] [Google Scholar]
- Kresse G.; Furthmüller J. Efficiency of Ab-Initio Total Energy Calculations for Metals and Semiconductors Using a Plane-Wave Basis Set. Comput. Mater. Sci. 1996, 6, 15–50. 10.1016/0927-0256(96)00008-0. [DOI] [PubMed] [Google Scholar]
- Kresse G.; Furthmüller J. Efficient Iterative Schemes for Ab Initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Rev. B 1996, 54, 11169–11186. 10.1103/PhysRevB.54.11169. [DOI] [PubMed] [Google Scholar]
- Krukau A. V.; Vydrov O. A.; Izmaylov A. F.; Scuseria G. E. Influence of the Exchange Screening Parameter on the Performance of Screened Hybrid Functionals. J. Chem. Phys. 2006, 125, 224106. 10.1063/1.2404663. [DOI] [PubMed] [Google Scholar]
- Schimka L.; Harl J.; Kresse G. Improved Hybrid Functional for Solids: The HSEsol Functional. J. Chem. Phys. 2011, 134, 024116 10.1063/1.3524336. [DOI] [PubMed] [Google Scholar]
- Bokdam M.; Sander T.; Stroppa A.; Picozzi S.; Sarma D. D.; Franchini C.; Kresse G. Role of Polar Phonons in the Photo Excited State of Metal Halide Perovskites. Sci. Rep. 2016, 6, 28618. 10.1038/srep28618. [DOI] [PMC free article] [PubMed] [Google Scholar]
- https://pubs.aip.org/aip/apm/article/7/1/010901/1062013/Dielectric-and-ferroic-properties-of-metal-halide
- Brivio F.; Walker A. B.; Walsh A. Structural and Electronic Properties of Hybrid Perovskites for High-Efficiency Thin-Film Photovoltaics from First-Principles. APL Mater. 2013, 1, 042111 10.1063/1.4824147. [DOI] [Google Scholar]
- Varadwaj P. R.; Marques H. M. The Cs2AgRhCl6 Halide Double Perovskite: A Dynamically Stable Lead-Free Transition-Metal Driven Semiconducting Material for Optoelectronics. Front. Chem. 2020, 8, 796. 10.3389/fchem.2020.00796. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cockayne E.; Burton B. P. Phonons and Static Dielectric Constant in CaTiO3 from First Principles. Phys. Rev. B 2000, 62, 3735–3743. 10.1103/PhysRevB.62.3735. [DOI] [Google Scholar]
- Cockayne E. Comparative Dielectric Response in CaTiO3 and CaAl1/2Nb1/2O3 from First Principles. J. Appl. Phys. 2001, 90, 1459–1468. 10.1063/1.1380991. [DOI] [Google Scholar]
- Umari P.; Pasquarello A. Ab Initio Molecular Dynamics in a Finite Homogeneous Electric Field. Phys. Rev. Lett. 2002, 89, 157602 10.1103/PhysRevLett.89.157602. [DOI] [PubMed] [Google Scholar]
- Parlinski K.; Li Z. Q.; Kawazoe Y. First-Principles Determination of the Soft Mode in Cubic ZrO2. Phys. Rev. Lett. 1997, 78, 4063–4066. 10.1103/PhysRevLett.78.4063. [DOI] [Google Scholar]
- Parlinski K.; Piekarz P. Ab Initio Determination of Raman Spectra of Mg2SiO4 and Ca2MgSi2O7 Showing Mixed Modes Related to LO/TO Splitting. J. Raman Spectrosc. 2021, 52, 1346–1359. 10.1002/jrs.6129. [DOI] [Google Scholar]
- Togo A.; Tanaka I. First Principles Phonon Calculations in Materials Science. Scr. Mater. 2015, 108, 1–5. 10.1016/j.scriptamat.2015.07.021. [DOI] [Google Scholar]
- Mashimo T.; Bagum R.; Ogata Y.; Tokuda M.; Okube M.; Sugiyama K.; Kinemuchi Y.; Isobe H.; Yoshiasa A. Structure of Single-Crystal Rutile (TiO2) Prepared by High-Temperature Ultracentrifugation. Cryst. Growth Des. 2017, 17, 1460–1464. 10.1021/acs.cgd.6b01818. [DOI] [Google Scholar]
- Seitz G.; Penin N.; Decoux L.; Wattiaux A.; Duttine M.; Gaudon M. Near the Ferric Pseudobrookite Composition (Fe2TiO5). Inorg. Chem. 2016, 55, 2499–2507. 10.1021/acs.inorgchem.5b02847. [DOI] [PubMed] [Google Scholar]
- Murugesan S.; Thirumurugesan R.; Mohandas E.; Parameswaran P. X-Ray Diffraction Rietveld Analysis and Bond Valence Analysis of Nano Titania Containing Oxygen Vacancies Synthesized via Sol-Gel Route. Mater. Chem. Phys. 2019, 225, 320–330. 10.1016/j.matchemphys.2018.12.061. [DOI] [Google Scholar]
- Bokhimi X.; Pedraza F. Characterization of Brookite and a New Corundum-like Titania Phase Synthesized under Hydrothermal Conditions. J. Solid State Chem. 2004, 177, 2456–2463. 10.1016/j.jssc.2004.04.021. [DOI] [Google Scholar]
- Rezaee M.; Khoie S. M. M.; Liu K. H. The Role of Brookite in Mechanical Activation of Anatase-to-Rutile Transformation of Nanocrystalline TiO2: An XRD and Raman Spectroscopy Investigation. CrystEngComm 2011, 13, 5055–5061. 10.1039/C1CE05185G. [DOI] [Google Scholar]
- Glover E. N. K.; Ellington S. G.; Sankar G.; Palgrave R. G. The Nature and Effects of Rhodium and Antimony Dopants on the Electronic Structure of TiO2: Towards Design of Z-Scheme Photocatalysts. J. Mater. Chem. A 2016, 4, 6946–6954. 10.1039/C6TA00293E. [DOI] [Google Scholar]
- Quiroz H. P.; Quintero F.; Arias P. J.; Dussan A.; Zea H. R. Effect of Fluoride and Water Content on the Growth of TiO2 Nanotubes Synthesized via Ethylene Glycol with Voltage Changes during Anodizing Process. J. Phys.: Conf. Ser. 2015, 614, 012001 10.1088/1742-6596/614/1/012001. [DOI] [Google Scholar]
- Yang G.-J.; Li C.-J.; Han F.; Huang X.-C. Effects of Annealing Treatment on Microstructure and Photocatalytic Performance of Nanostructured TiO2 Coatings through Flame Spraying with Liquid Feedstocks. J. Vac. Sci. Technol., B: Microelectron. Nanometer Struct.-Process., Meas., Phenom. 2004, 22, 2364–2368. 10.1116/1.1788679. [DOI] [Google Scholar]
- Silva Junior E.; Porta F. A. L.; Liu M. S.; Andrés J.; Varela J. A.; Longo E. A Relationship between Structural and Electronic Order–Disorder Effects and Optical Properties in Crystalline TiO2 Nanomaterials. Dalton Trans. 2015, 44, 3159–3175. 10.1039/C4DT03254C. [DOI] [PubMed] [Google Scholar]
- Nishio-Hamane D.; Zhang M.; Yagi T.; Ma Y. High-Pressure and High-Temperature Phase Transitions in FeTiO3 and a New Dense FeTi3O7 Structure. Am. Mineral. 2012, 97, 568–572. 10.2138/am.2012.3973. [DOI] [Google Scholar]
- Jain A.; Ong S. P.; Hautier G.; Chen W.; Richards W. D.; Dacek S.; Cholia S.; Gunter D.; Skinner D.; Ceder G.; Persson K. A. Commentary: The Materials Project: A Materials Genome Approach to Accelerating Materials Innovation. APL Mater. 2013, 1, 011002 10.1063/1.4812323. [DOI] [Google Scholar]
- mp-1102591: TiO2 (Cubic, Pa-3, 205). Materials Project. https://materialsproject.org/materials/mp-1102591/ (accessed Oct 21, 2022).
- mp-1008677: TiO2 (Cubic, Fm-3m, 225). Materials Project. https://materialsproject.org/materials/mp-1008677/ (accessed Oct 21, 2022).
- Born M. On the Stability of Crystal Lattices I. Math. Proc. Cambridge Philos. Soc. 1940, 36, 160–172. 10.1017/S0305004100017138. [DOI] [Google Scholar]
- Born M.; Huang K.; Born M.; Huang K.. Dynamical Theory of Crystal Lattices; Oxford Classic Texts in the Physical Sciences; Oxford University Press: Oxford, New York, 1998. [Google Scholar]
- Traylor J. G.; Smith H. G.; Nicklow R. M.; Wilkinson M. K. Lattice Dynamics of Rutile. Phys. Rev. B 1971, 3, 3457–3472. 10.1103/PhysRevB.3.3457. [DOI] [Google Scholar]
- Arsov L.; Kormann C.; Plieth W. Electrochemical Synthesis and in Situ Raman Spectroscopy of Thin Films of Titanium Dioxide. J. Raman Spectrosc. 1991, 22, 573–575. 10.1002/jrs.1250221006. [DOI] [Google Scholar]
- Ma H. L.; Yang J. Y.; Dai Y.; Zhang Y. B.; Lu B.; Ma G. H. Raman Study of Phase Transformation of TiO2 Rutile Single Crystal Irradiated by Infrared Femtosecond Laser. Appl. Surf. Sci. 2007, 253, 7497–7500. 10.1016/j.apsusc.2007.03.047. [DOI] [Google Scholar]
- Tompsett G. A.; Bowmaker G. A.; Cooney R. P.; Metson J. B.; Rodgers K. A.; Seakins J. M. The Raman Spectrum of Brookite, TiO2 (Pbca, Z = 8). J. Raman Spectrosc. 1995, 26, 57–62. 10.1002/jrs.1250260110. [DOI] [Google Scholar]
- Gervais F.; Piriou B. Anharmonicity in Several-Polar-Mode Crystals: Adjusting Phonon Self-Energy of LO and TO Modes in Al2O3 and TiO2 to Fit Infrared Reflectivity. J. Phys. C: Solid State Phys. 1974, 7, 2374. 10.1088/0022-3719/7/13/017. [DOI] [Google Scholar]
- Eagles D. M. Polar Modes of Lattice Vibration and Polaron Coupling Constants in Rutile (TiO2). J. Phys. Chem. Solids 1964, 25, 1243–1251. 10.1016/0022-3697(64)90022-8. [DOI] [Google Scholar]
- Montanari B.; Harrison N. M. Pressure-Induced Instabilities in Bulk TiO2 Rutile. J. Phys.: Condens. Matter 2004, 16, 273. 10.1088/0953-8984/16/3/008. [DOI] [Google Scholar]
- Montanari B.; Harrison N. M. Lattice Dynamics of TiO2 Rutile: Influence of Gradient Corrections in Density Functional Calculations. Chem. Phys. Lett. 2002, 364, 528–534. 10.1016/S0009-2614(02)01401-X. [DOI] [Google Scholar]
- Mitev P. D.; Hermansson K.; Montanari B.; Refson K. Soft Modes in Strained and Unstrained Rutile TiO2. Phys. Rev. B 2010, 81, 134303 10.1103/PhysRevB.81.134303. [DOI] [Google Scholar]
- Mikami M.; Nakamura S.; Kitao O.; Arakawa H. Lattice Dynamics and Dielectric Properties of TiO2 Anatase: A First-Principles Study. Phys. Rev. B 2002, 66, 155213 10.1103/PhysRevB.66.155213. [DOI] [Google Scholar]
- Balachandran U.; Eror N. G. Raman Spectra of Titanium Dioxide. J. Solid State Chem. 1982, 42, 276–282. 10.1016/0022-4596(82)90006-8. [DOI] [Google Scholar]
- Gonzalez R. J.; Zallen R.; Berger H. Infrared Reflectivity and Lattice Fundamentals in Anatase TiO2s. Phys. Rev. B 1997, 55, 7014–7017. 10.1103/PhysRevB.55.7014. [DOI] [Google Scholar]
- Beattie I. R.; Gilson T. R.; Anderson J. S. Single Crystal Laser Raman Spectroscopy. Proc. R. Soc. A 1968, 307, 407–429. 10.1098/rspa.1968.0199. [DOI] [Google Scholar]
- Iliev M. N.; Hadjiev V. G.; Litvinchuk A. P. Raman and Infrared Spectra of Brookite (TiO2): Experiment and Theory. Vib. Spectrosc. 2013, 64, 148–152. 10.1016/j.vibspec.2012.08.003. [DOI] [Google Scholar]
- Landmann M.; Rauls E.; Schmidt W. G. The Electronic Structure and Optical Response of Rutile, Anatase and Brookite TiO2. J. Phys.: Condens. Matter 2012, 24, 195503 10.1088/0953-8984/24/19/195503. [DOI] [PubMed] [Google Scholar]
- Zhao X.; Vanderbilt D. First-Principles Study of Structural, Vibrational, and Lattice Dielectric Properties of Hafnium Oxide. Phys. Rev. B 2002, 65, 233106 10.1103/PhysRevB.65.233106. [DOI] [Google Scholar]
- Berberich L. J.; Bell M. E. The Dielectric Properties of the Rutile Form of TiO 2. J. Appl. Phys. 1940, 11, 681–692. 10.1063/1.1712721. [DOI] [Google Scholar]
- Hu W.; Li L.; Li G.; Tang C.; Sun L. High-Quality Brookite TiO2 Flowers: Synthesis, Characterization, and Dielectric Performance. Cryst. Growth Des. 2009, 9, 3676–3682. 10.1021/cg9004032. [DOI] [Google Scholar]
- De Lile J. R.; Kang S. G.; Son Y.-A.; Lee S. G. Investigating Polaron Formation in Anatase and Brookite TiO2 by Density Functional Theory with Hybrid-Functional and DFT+U Methods. ACS Omega 2019, 4, 8056–8064. 10.1021/acsomega.9b00443. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shibuya T.; Yasuoka K.; Mirbt S.; Sanyal B. A Systematic Study of Polarons Due to Oxygen Vacancy Formation at the Rutile TiO2(110) Surface by GGA+U and HSE06 Methods. J. Phys.: Condens. Matter 2012, 24, 435504 10.1088/0953-8984/24/43/435504. [DOI] [PubMed] [Google Scholar]
- Reticcioli M.; Setvin M.; Schmid M.; Diebold U.; Franchini C. Formation and Dynamics of Small Polarons on the Rutile TiO2 (110) Surface. Phys. Rev. B 2018, 98, 045306 10.1103/PhysRevB.98.045306. [DOI] [Google Scholar]
- Wemple S. H. Optical Oscillator Strengths and Excitation Energies in Solids, Liquids, and Molecules. J. Chem. Phys. 1977, 67, 2151–2168. 10.1063/1.435102. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
This research did not report any data.






