Next Article in Journal
Hydrothermal Synthesis, Characterization and Exploration of Photocatalytic Activities of Polyoxometalate: Ni-CoWO4 Nanoparticles
Next Article in Special Issue
Interactions of Ruddlesden-Popper Phases and Migration-Induced Field-Stabilized Polar Phase in Strontium Titanate
Previous Article in Journal
Effect of the External Velocity on the Exfoliation Properties of Graphene from Amorphous SiO2 Surface
Previous Article in Special Issue
Correlation between Structural and Transport Properties of Ca-Doped La Nickelates and Their Electrochemical Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Comparative Hybrid Hartree-Fock-DFT Calculations of WO2-Terminated Cubic WO3 as Well as SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) Surfaces

1
Institute of Solid State Physics, University of Latvia, 8 Kengaraga Street, LV1063 Riga, Latvia
2
Laboratory of Theoretical and Computational Chemistry, Institute of Theoretical Chemistry, Jilin University, Changchun 130023, China
*
Author to whom correspondence should be addressed.
Crystals 2021, 11(4), 455; https://doi.org/10.3390/cryst11040455
Submission received: 29 March 2021 / Revised: 17 April 2021 / Accepted: 19 April 2021 / Published: 20 April 2021
(This article belongs to the Special Issue Diffusion and Degradation Phenomena in Solid Oxide Materials)

Abstract

:
We performed, to the best of our knowledge, the world’s first first-principles calculations for the WO2-terminated cubic WO3 (001) surface and analyzed the systematic trends in the WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) surface ab initio calculations. According to our first principles calculations, all WO2 or TiO2-terminated WO3, SrTiO3, BaTiO3, PbTiO3 and CaZrO3 (001) surface upper-layer atoms relax inwards towards the crystal bulk, while all second-layer atoms relax upwards. The only two exceptions are outward relaxations of first layer WO2 and TiO2-terminated WO3 and PbTiO3 (001) surface O atoms. The WO2 or TiO2-terminated WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) surface-band gaps at the Γ–Γ point are smaller than their respective bulk-band gaps. The Ti–O chemical bond populations in the SrTiO3, BaTiO3, PbTiO3 and CaTiO3 bulk are smaller than those near the TiO2-terminated (001) surfaces. Conversely, the W–O chemical bond population in the WO3 bulk is larger than near the WO2-terminated WO3 (001) surface.

1. Introduction

Throughout the last 20 years the SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) surfaces have been broadly explored theoretically and experimentally [1,2,3,4,5,6,7,8,9,10]. At the same time, to the best of our knowledge, there are no reports of ab initio calculations dealing with the atomic relaxation and electronic structure of the pristine WO2-terminated WO3 (001) surface in the cubic perovskite-like structure. Nevertheless, a large amount of experimental studies exist dealing with WO3 (001) surfaces [11,12,13,14]. Recent theoretical studies have been devoted, for example, to hydrogen adsorption on the WO3 (001) surface [15], understanding the water splitting process on the WO3 (001) surface [16], and H2O adsorption on the WO3 (001) surface [17].
BaTiO3, PbTiO3 and CaTiO3 perovskites have attracted huge fundamental interest in these materials mostly for their phase transitions. Historically the ABO3 perovskites were highly promising low-cost energy materials. They have been used for numerous optoelectronic and photonic device applications [18]. SrTiO3 perovskite thin films are important for a large amount of technologically important applications [19,20]. For example, they are used for catalysis, optical wave guides, high-capacity memory cells as well as substrates for high-temperature cuprate superconductor growth [19,20]. Barium titanate (BaTiO3) is an excellent photorefractive material [18]. Ferroelectric PbTiO3 thin films have been applied to large numbers of electronic devices, such as non-volatile memory FET [19] and Si monolithic ultrasonic sensors [18]. CaTiO3 is used worldwide in technologically important electronic ceramic materials [18]. Tungsten trioxide (WO3) and its thin films exhibit a large number of novel properties useful for high-technology applications [21]. In particular, WO3 undergoes phase transitions, which are explored for their potential in industrial applications, display systems and microelectronics [21]. It is worth noting that the predictive power of ab initio calculations makes possible the design of new materials for high-technology applications on paper. Nowadays, consumer electronics mostly use lithium-ion batteries containing LiCoO2 cathode, which was discovered in 1980 by J. Goodenough, one of the 2019 Nobel Prize winners for Chemistry [22]. The experimentally detected LiCoO2 average intercalation voltage is 4.0–4.1 V [23]. Based on ab initio calculations by Eglitis and Borstel [24,25,26], it was demonstrated that a novel Li2CoMn3O8 battery cathode material can lead to a high-energy lithium-ion battery working at the 5 V regime.
The SrTiO3, BaTiO3, PbTiO3 and CaTiO3 perovskite cubic unit cells contain five atoms. The A type atom (A = Sr, Ba, Pb or Ca) has the coordinates (0, 0, 0), and it is located in the cube corner position. The Ti atom has the coordinates (½, ½, ½), and it is located in the cube body center position. The 3 O atoms have the coordinates (½, ½, 0), (½, 0, ½), (0, ½, ½), and they are located in the cube face centered positions. All SrTiO3, BaTiO3, PbTiO3 and CaTiO3 cubic perovskites have the same space group Pm3m with the space group number equal to 221. WO3 in its cubic perovskite-like structure has exactly the same space group as ATiO3 perovskites Pm3m, and also the same space group number 221. The only striking difference between the SrTiO3, BaTiO3, PbTiO3 and CaTiO3 cubic perovskites as well as WO3 in its cubic perovskite-like structure is that WO3 has an empty A cation position. Thereby, the cubic perovskite-like unit cell of WO3 contain only four atoms.
The objective of the reported here work was to carry out first-principles calculations for WO2-terminated polar WO3 (001) surfaces in the cubic perovskite-like structure. We compared our WO2-terminated WO3 (001) surface-atomic and electronic-structure ab initio calculations with our results for the related structure TiO2-terminated SrTiO3, BaTiO3, PbTiO3 and CaTiO3 cubic perovskite (001) surfaces. We carefully compared our calculation results for all five of our calculated materials and detected systematic common trends. The results for WO2-terminated WO3 and TiO2-terminated SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) surfaces were summarized and analysed in a way easily readable for a broad audience of scientists.

2. Computational Methods and Surface Models

In order to carry out ab initio DFT-B3LYP or DFT-B3PW calculations, we employed the CRYSTAL computer program package [27]. Unlike the plane-wave codes widely employed in many previous studies [28,29], the CRYSTAL code [27] uses localized Gaussian-type basis sets. In our calculations, we adopted the basis sets (BS) developed for SrTiO3, BaTiO3 and PbTiO3 in [30]. The Hay–Wadt small-core, effective-core pseudopotentials (ECP) were adopted for Ca and Ti atoms [31,32,33]. The small-core ECPs replaced only the inner-core orbitals, while orbitals for subvalence electrons as well as for valence electrons were calculated self-consistently. Oxygen atoms were treated with the all-electron BS. Finally, for the W atom we used BS developed by Cora et al. [34]. Our calculations were performed by means of the B3LYP [35] or B3PW [36,37,38] hybrid exchange–correlation functionals. For all WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 materials we performed the reciprocal space integration with an 8 × 8 × 8 and 8 × 8 × 1 extension of Pack–Monkhorst mesh for the bulk and (001) surfaces of these materials. The CRYSTAL computer program package [27] makes possible the calculation of isolated 2D slabs perpendicular to the Oz direction. In order to compare the performance of different exchange–correlation functionals and choose the best method for our calculations, we calculated the SrTiO3, SrZrO3, BaZrO3, MgF2 and CaF2 bulk Γ–Γ band gaps [30,39,40,41,42] (Table 1 and Figure 1). The experimentally detected SrTiO3, SrZrO3, BaZrO3, MgF2 and CaF2 bulk band gaps at the Γ-point are mentioned in Table 1 for comparison purposes as well as depicted in Figure 1 [43,44,45,46,47].
As can be seen in Table 1, the ab initio Hartree-Fock (HF) calculations, for all five our calculated materials, very strongly overestimate the experimental band gap at Γ-point. Namely, the HF method most strongly (3.29 times) overestimate the experimental SrTiO3 bulk Γ–Γ band gap. Even HF calculated MgF2 bulk Γ–Γ band gap overestimates the experimental value 1.51 times (Table 1 and Figure 1).
From another side, as we can see from Table 1 and Figure 1, the generalized gradient approximations (GGA) to the density functional theory (DFT) systematically and considerably underestimate the experimental Γ–Γ bulk band gap in our calculated ABO3 perovskites as well as MgF2 and CaF2. For example, the PWGGA (6.94 eV) and PBE (6.91 eV) calculated MgF2 bulk band gap at Γ-point is 1.87 and 1.88 times, respectively, smaller than the experimental MgF2 bulk Γ–Γ band gap value of 13.0 eV [37].
To obtain the best possible results, we performed our WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 bulk and (001) surface calculations by means of the B3PW [36,37,38] or B3LYP [35] hybrid exchange–correlation functionals. The hybrid functional incorporates a portion of exact exchange energy density from HF theory (20%) while the rest of the exchange–correlation part is a mixture of different approaches (both exchange and correlation). It is obvious, that the B3PW and B3LYP hybrid exchange–correlation functionals, since they are a superposition of HF and DFT methods as implemented in the CRYSTAL computer code [27], allowed us to achieve as good an agreement as possible between the first principles calculated and the experimentally detected Γ–Γ band gaps for WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 bulk and their (001) surfaces.
In our ab initio calculations we used WO2-terminated WO3 as well as TiO2-terminated SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) slabs containing 9 alternating layers. First our calculated WO3 (001) slab was terminated by WO2 planes from both sides (WO2–O–WO2–O–WO2–O–WO2–O–WO2) from a 19-atom supercell (Figure 2). Another of our calculated SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) slabs was terminated by TiO2 planes from both sides (TiO2–AO–TiO2–AO–TiO2–AO–TiO2–AOvTiO2) and consisted of a 23-atom supercell (Figure 3). Both our calculated slabs were non-stoichiometric and had unit-cell equations W5O14 as well as A4Ti5O14, respectively. To analyse the chemical bonds, effective atomic charges and covalency effects for WO3 and ATiO3 perovskite bulk and (001) surfaces, we used the well-known Mulliken population analysis [48,49,50,51,52].

3. Ab initio Calculation Results for WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 Bulk

As a starting point, by means of the hybrid B3LYP exchange–correlation functional, we calculated the cubic WO3 bulk lattice constant (3.775 Å). Our calculated cubic WO3 constant (3.775 Å) was only slightly larger than the experimental value of a0 = 3.71–3.75 Å [51] (Table 2); nevertheless, it was in almost perfect agreement with the earlier calculation result for the WO3 cubic structure bulk lattice constant calculated by the full-potential linear muffin-tin (FP-LMTO) code equal to 3.78 Å [52]. Our B3PW-calculated SrTiO3 bulk lattice constant (3.904 Å) was only slightly overestimated with respect to the experimental SrTiO3 bulk lattice constant (3.89 Å) extrapolated to 0 K [53] (Table 2). Our ab initio calculation of the BaTiO3 bulk lattice constant (4.008 Å) was in an outstanding agreement with the experimental value of 4.00 Å [53,54,55]. Our B3PW-calculated PbTiO3 bulk lattice constant (3.936 Å) [54,55,56] was only 0.86% under the experimental value of 3.97 Å [57]. Finally, our calculated CaTiO3 bulk lattice constant (3.851 Å) was 1.17% smaller than the experimentally detected (3.8967 Å) [58,59,60] (Table 2).
Our ab initio B3LYP-calculated effective atomic charges for the WO3 bulk were (+3.095e) for the W atom, and (−1.032e) for each of the three O atoms (Table 3). Our B3LYP-calculated effective W atomic charge (+3.095e) was almost two times smaller than the generally accepted classical ionic charge for the W(+6e) atom. In addition, our calculated effective atomic charge for the O (−1.032e) atom was almost two times smaller than the generally accepted O atom classical ionic charge (−2e). In addition, for the SrTiO3, BaTiO3, PbTiO3 and CaTiO3 perovskites, our calculated A atomic charges (+1.871e, +1.797e, +1.354e and +1.782e, respectively) were considerably smaller than those of the classical Sr, Ba, Pb, Ca atom ionic charges (+2e) (Table 3) [61,62,63,64,65,66]. Our B3PW-calculated O atom Mulliken charges in SrTiO3, BaTiO3, PbTiO3 and CaTiO3 perovskites (−1.407e, −1.388e, −1.232e and −1.371e, respectively) are also at least 29.65% smaller than the classical ionic O atomic charge (−2e) [67,68,69]. Finally, our ab initio-calculated SrTiO3, BaTiO3, PbTiO3 and CaTiO3 Ti atomic charges (+2.351e, +2.367e, +2.341e and 2.330e) are more than one-and-a-half times smaller than the formal Ti atom ionic charge (+4e). Our calculated chemical bond population between W and O atoms in WO3 bulk (0.142e) is approximately one-and-a-half times larger than the Ti–O atom chemical bond population in SrTiO3, BaTiO3, PbTiO3 and CaTiO3 perovskites (+0.088e, +0.098e, +0.098e and +0.084e, respectively).
Our B3LYP-calculated WO3 bulk Γ–Γ band gap (4.95 eV) overestimated by 1.21 eV the experimental direct WO3 bulk band gap value at Γ-point of 3.74 eV [70] (Table 4). Moreover, our B3PW-calculated bulk Γ–Γ band gaps for SrTiO3, BaTiO3, PbTiO3 and CaTiO3 perovskites (3.96 eV, 3.55 eV, 4.32 eV and 4.18 eV, respectively) were always slightly overestimated with respect to the experimentally measured direct band gap values at Γ-point for SrTiO3, BaTiO3, PbTiO3 and CaTiO3 perovskites (3.75 eV [43], 3.2 eV [71], 3.4 eV [72] and 3.5 eV [73], respectively) (Table 4).

4. Ab Initio Calculation Results for the WO2-Terminated WO3 as Well as TiO2-Terminated SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) Surfaces

Our B3LYP- or B3PW-calculated atomic displacements for the WO2-terminated WO3 and the TiO2-terminated SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) surface upper-three or two layers are presented in Table 5. According to our B3LYP or B3PW calculations, all atoms of the first (upper) surface layer relaxed inwards, while all second-layer atoms relaxed outwards (Table 5). The only two exceptions to this systematic trend were the outward relaxation of the first layer O atom of the WO2-terminated WO3 (001) surface (+0.42% of a0) and the outward relaxation of the TiO2-terminated PbTiO3 (001) surface first-layer O atom by (0.31% of a0) (Table 5). The first layer metal atom relaxation magnitudes range from −1.71% of a0 for the TiO2-terminated CaTiO3 (001) surface to −3.08% of a0 for the TiO2-terminated BaTiO3 (001) surface (Table 5). The first- and second-layer metal atom displacement magnitudes for WO2-terminated WO3 and TiO2-terminated SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) surfaces were always considerably larger than the respective first- and second-layer O atom displacement magnitudes (Table 5).
Our B3LYP-calculated surface rumpling amplitude s (the relative displacement of an oxygen atom relative to the metal atom in the upper surface layer) for WO2-terminated WO3 (001) surface (+2.49) is in qualitative agreement with our B3PW-calculated surface rumpling amplitudes s for TiO2-terminated BaTiO3, PbTiO3, CaTiO3 and SrTiO3 (001) surfaces (+2.73, +3.12, +1.61 and +2.12, respectively) (Table 6). Our B3PW-calculated surface rumpling amplitude s for TiO2-terminated SrTiO3 (001) surface (+2.12) is in fair agreement with available RHEED (+2.6 [74]) and LEED (+2.1 ± 2 [75]) experimental data (Table 6). Unfortunately, our B3PW-calculated interlayer distance Δd12 for the TiO2-terminated SrTiO3 (001) surface (−5.80) had the opposite sign to the experimentally measured RHEED (+1.8 [74]) and LEED (+1 ± 1 [67]) interlayer distances (Table 6). Finally, our B3PW-calculated interlayer distance Δd23 for the TiO2-terminated SrTiO3 (001) surface (+3.55) is in qualitative agreement with the RHEED experiment result (+1.3), but had the opposite sign to that of the LEED experimental result (−1 ± 1). Nevertheless, it is worth noting that the RHEED (+1.3) and the LEED experiments (−1 ± 1) had opposite signs for the interlayer distance Δd23 (Table 6).
We started the discussion of the electronic structure of WO2-terminated WO3 and TiO2-terminated SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) surfaces with an analysis of charge redistribution in the top-three surface planes (Table 7). The ab initio-calculated atomic displacements, bond populations between the nearest metal and oxygen atoms and the effective atomic charges are collected in Table 7. For example, the effective static atomic charges on WO2-terminated WO3 as well as TiO2-terminated SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) surface upper-layer W and Ti atoms are always reduced in comparison to the bulk WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 crystal charges (−0.312e, −0.06e, −0.06e, −0.062e and −0.052e, respectively). We recently observed a similar effect: the reduction of surface upper-layer metal atomic charges near the ReO2-terminated ReO3 and the ZrO2-terminated SrZrO3, BaZrO3, PbZrO3 and CaZrO3 (001) surfaces [76]. According to our ab initio calculations, the largest upper-layer metal atom displacement was observed for the TiO2-terminated BaTiO3 (001) surface Ba atom (−0.123 Å). Nevertheless, the TiO2-terminated PbTiO3 (001) surface second-layer Pb atom outward displacement (+0.209 Å) was even larger.
Our B3PW-calculated SrTiO3, BaTiO3, PbTiO3 and CaTiO3 bulk Ti-O chemical bond covalency (+0.088e, +0.098e, +0.098e and +0.084e, respectively) were always smaller than near the TiO2-terminated SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) surfaces (0.118e, 0.126e, 0.114e, 0.114e, respectively) (Table 8 and Figure 4). Just opposite situation was obtained for the WO3 crystal: the B3LYP-calculated W-O chemical bond population in the WO3 bulk (0.142e) was larger than near the WO2-terminated WO3 (001) surface (0.108e) (Table 8 and Figure 4). Nevertheless, it is worth noting that the W–O chemical bond population between the W atom on the top layer of WO2-terminated WO3 (001) surface and the O atom on the second layer (0.278e) is the largest one (Table 8), which was in agreement with our previous B3LYP calculations dealing with ReO2-terminated ReO3 (001) surfaces [76].
As can be seen in Table 9 and Figure 5, our B3LYP- or B3PW-calculated WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 bulk band gaps at the Γ–Γ point were always reduced near the WO2- or TiO2-terminated WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) surfaces. The B3PW-calculated SrTiO3 bulk band gap at Γ–Γ point near the TiO2-terminated SrTiO3 (001) surface at Γ–Γ point was reduced only by 0.01 eV. At the same time, our B3LYP-calculated WO3 bulk band gap (4.95 eV) at the Γ–Γ point near the WO2-terminated WO3 (001) surface was reduced by 3.79 eV to 1.16 eV (Table 9 and Figure 5).

5. Conclusions

For the first principles-calculated WO2- or TiO2-terminated WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) surfaces, as a rule, all first-layer surface atoms relax inwards, whereas all second-layer surface atoms relax upwards. The only two exceptions from this systematic trend are the upward relaxation of WO2- or TiO2-terminated WO3 and PbTiO3 (001) surface first-layer O atoms. As a result of our ab initio-calculated atomic relaxation, TiO2-terminated SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) surfaces exhibited a reduction of the interlayer distance Δd12 (−5.80, −5.59, −8.13, −4.46% of a0, respectively) as well as an expansion of Δd23 (+3.55, +2.51, +5.32, +2.75% of a0, respectively). It is worth noting that after geometry optimization, it is very useful to perform ab initio molecular dynamics computations to ensure the stability of the structures over time [77].
The changes in the interlayer distances between the first and second layer (Δd12) were always larger than between the second and third layer (Δd23) for all our calculated perovskites SrTiO3, BaTiO3, PbTiO3 and CaTiO3.
The Ti–O chemical bond population in SrTiO3, BaTiO3, PbTiO3 and CaTiO3 bulk was always smaller than near their TiO2-terminated (001) surface (see Figure 4). In contrast, the W–O chemical bond population in the WO3 bulk (0.142e) was larger than near the WO2-terminated WO3 (001) surface (0.108e). Nevertheless, the largest W–O chemical bond population, according to our ab initio calculations, is between the W atom located on the WO2-terminated WO3 (001) surface and the second-layer O atom (0.278e). It was worth noting, that also for the related material ReO3, according to our calculations [76], the situation was similar. Namely, the Re–O chemical bond population in the ReO3 bulk (0.212e) was larger than near the ReO2-terminated ReO3 (001) surface (0.170e). Nevertheless, the Re–O chemical bond population between the Re atom located on the ReO2-terminated ReO3 (001) surface upper-layer and O atom located on the ReO2-terminated ReO3 (001) surface second layer was the largest (0.262e).
According to our B3LYP or B3PW calculations, the WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 bulk Γ–Γ band gap values (4.95, 3.96, 3.55, 4.32, 4.18 eV, respectively) were always reduced with respect to the bulk near the WO2- or TiO2-terminated WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) surfaces (1.16, 3.95, 2.96, 3.18, 3.30 eV, respectively) (see Figure 5).

Author Contributions

All authors equally contributed to the performed ab initio calculations as well as to the preparation of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the ERAF Project No. 1.1.1.1/18/A/073.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We greatly acknowledge the financial support via the ERAF Project No. 1.1.1.1/18/A/073. Calculations were performed using Latvian Super Cluster (LASC), located in the Center of Excellence at Institute of Solid State Physics, the University of Latvia, which is supported by European Union Horizon 2020 Framework Programme H2020-WIDESPREAD-01-2016-2017-Teaming Phase 2 under Grant Agreement No. 739508, project CAMART.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Solokha, V.; Garai, D.; Wilson, A.; Duncan, D.A.; Thakur, P.K.; Hingerl, K.; Zegenhagen, J. Water splitting on Ti-oxide-terminated SrTiO3 (001). J. Phys. Chem. C 2019, 123, 17232–17238. [Google Scholar] [CrossRef] [Green Version]
  2. Saghayezhian, M.; Sani, S.M.R.; Zhang, J.D.; Plummer, E.W. Rumpling and enhanced covalency at the SrTiO3 (001) surface. J. Phys. Chem. C 2019, 123, 8086–8091. [Google Scholar] [CrossRef]
  3. Foo, G.S.; Hood, Z.D.; Wu, Z.L. Shape effect undermined by surface reconstruction: Ethanol dehydrogenation over shape-controlled SrTiO3 nanocrystals. ACS Catal. 2018, 8, 555–565. [Google Scholar] [CrossRef]
  4. Ryu, G.H.; Lewis, N.P.; Kotsonis, G.N.; Maria, J.P.; Dickey, E.C. Crystallization behavior of amorphous BaTiO3 thin films. J. Mater. Sci. 2020, 55, 8793–8801. [Google Scholar] [CrossRef]
  5. Heifets, E.; Eglitis, R.I.; Kotomin, E.A.; Maier, J.; Borstel, G. Ab initio modeling of surface structure for SrTiO3 perovskite. Phys. Rev. B 2001, 64, 235417. [Google Scholar] [CrossRef] [Green Version]
  6. Ananyev, M.V.; Porotnikova, N.M.; Eremin, V.A.; Kurumchin, E.K. Interaction of O2 with LSM-YSZ Composite Materials and Oxygen Spillover Effect. ACS Catal. 2021, 11, 4247–4262. [Google Scholar] [CrossRef]
  7. Piskunov, S.; Eglitis, R.I. First principles hybrid DFT calculations of BaTiO3/SrTiO3 (001) interface. Solid State Ionics 2015, 274, 29–33. [Google Scholar] [CrossRef]
  8. Sternlicht, H.; Rheinheimer, W.; Dunin-Borkowski, R.E.; Hoffmann, M.J.; Kaplan, W.D. Characterization of grain boundary disconnections in SrTiO3 part I: The dislocation component of grain boundary disconnections. J. Mater. Sci. 2019, 54, 3694–3709. [Google Scholar] [CrossRef]
  9. Borstel, G.; Eglitis, R.I.; Kotomin, E.A.; Heifets, E. Modelling of defects and surfaces in perovskite ferroelectrics. Phys. Stat. Sol. B 2003, 236, 253–264. [Google Scholar] [CrossRef]
  10. Eglitis, R.; Kruchinin, S.P. Ab initio calculations of ABO3 perovskite (001), (011) and (111) nano-surfaces, interfaces and defects. Modern Phys. Lett. B 2020, 34, 2040057. [Google Scholar] [CrossRef]
  11. Dixon, R.A.; Williams, J.J.; Morris, D.; Rebane, J.; Jones, F.H.; Egdell, R.G.; Downes, S.W. Electronic states at oxygen deficient WO3 (001) surfaces: A study by resonant photoemission. Surf. Sci. 1998, 399, 199–211. [Google Scholar] [CrossRef]
  12. Jones, F.H.; Rawlings, K.; Foord, J.S.; Egdell, R.G.; Pethica, J.B.; Wanklyn, B.M.R.; Parker, S.C.; Oliver, P.M. An STM study of surface structures on WO3 (001). Surf. Sci. 1996, 359, 107–121. [Google Scholar] [CrossRef]
  13. Bringaus, R.D.; Höchst, H.; Shanks, H.R. Hydrogen on WO3 (001). Surf. Sci. 1981, 111, 80–86. [Google Scholar] [CrossRef]
  14. Jones, F.H.; Dixon, R.A.; Brown, A. Observation of reduced (1 × 1) terraces on WO3 (001) surfaces using scanning tunneling microscopy. Surf. Sci. 1996, 369, 343–350. [Google Scholar] [CrossRef]
  15. Wang, F.; Valentin, C.D.; Pacchioni, G. DFT study of hydrogen adsorption on the monoclinic WO3 (001) surface. J. Phys. Chem. C 2012, 116, 10672–10679. [Google Scholar] [CrossRef]
  16. Teuch, T.; Klüner, T. Understanding the water splitting mechanism on WO3 (001)—A theoretical approach. J. Phys. Chem. C 2019, 123, 28233–28240. [Google Scholar] [CrossRef]
  17. Albanese, E.; Valentin, C.D.; Pacchioni, G. H2O adsorption on WO3 and WO3-x (001) surfaces. ACS Appl. Mater. Interfaces 2017, 9, 23212–23221. [Google Scholar] [CrossRef]
  18. Dawber, M.; Rabe, K.M.; Scott, J.F. Physics of thin-film ferroelectric oxides. Rev. Mod. Phys. 2005, 77, 1083–1130. [Google Scholar] [CrossRef] [Green Version]
  19. Auciello, O.; Scott, J.F.; Ramesh, R. The physics of ferroelectric memories. Phys. Today 1998, 51, 22–27. [Google Scholar] [CrossRef]
  20. Goniakowski, J.; Finnochi, F.; Noguera, C. Polarity of oxide surfaces and nanostructures. Rep. Progr. Phys. 2008, 71, 016501. [Google Scholar] [CrossRef]
  21. Granqvist, C.G. Electrochromic tungsten oxide films: Review of progress 1993–1998. Solar Energy Mater. Solar Cells 2000, 60, 201–262. [Google Scholar] [CrossRef]
  22. Mizushima, K.; Jones, P.C.; Wiseman, P.J.; Goodenough, J.B. LixCoO2 (0 < x < −1): A new cathode material for batteries of high energy density. Mater. Res. Bull. 1980, 15, 783–789. [Google Scholar] [CrossRef]
  23. Ohzuku, T.; Ueda, A. Solid-State Redox Reactions of LiCoO2 (R3m) for 4 Volt Secondary Lithium Cells. J. Electrochem. Soc. 1994, 141, 2972–2977. [Google Scholar] [CrossRef]
  24. Eglitis, R.I.; Borstel, G. Towards a practical rechargeable 5 V Li ion battery. Phys. Stat. Sol. A 2005, 202, R13–R15. [Google Scholar] [CrossRef]
  25. Eglitis, R.I. Theoretical prediction of the 5 V rechargeable Li ion battery using Li2CoMn3O8 as a cathode. Phys. Scr. 2015, 90, 094012. [Google Scholar] [CrossRef]
  26. Eglitis, R. Ab initio calculations of Li2(Co, Mn)O8 solid solutions for rechargeable batteries. Int. J. Mod. Phys. B 2019, 33, 1950151. [Google Scholar] [CrossRef]
  27. Saunders, V.R.; Dovesi, R.; Roetti, C.; Causa, N.; Harrison, N.M.; Orlando, R.; Zicovich-Wilson, C.M. CRYSTAL-2009 User Manual; University of Torino: Turin, Italy, 2009. [Google Scholar]
  28. Cohen, R.E. Periodic slab LAPW computations for ferroelectric BaTiO3. J. Phys. Chem. Solids 1996, 57, 1393–1396. [Google Scholar] [CrossRef] [Green Version]
  29. Cohen, R.E. Surface effects in ferroelectrics: Periodic slab computations for BaTiO3. Ferroelectrics 1997, 194, 323–342. [Google Scholar] [CrossRef]
  30. Piskunov, S.; Heifets, E.; Eglitis, R.I.; Borstel, G. Bulk properties and electronic structure of SrTiO3, BaTiO3, PbTiO3 perovskites: An ab initio HF/DFT study. Comput. Mater. Sci. 2004, 29, 165–178. [Google Scholar] [CrossRef]
  31. Hay, P.J.; Wadt, W.R. Ab initio effective core potentials for molecular calculations. Potentials for the transition metal atoms Sc to Hg. J. Chem. Phys. 1985, 82, 270–283. [Google Scholar] [CrossRef]
  32. Hay, P.J.; Wadt, W.R. Ab initio effective core potentials for molecular calculations. Potentials for main group elements Na to Bi. J. Chem. Phys. 1985, 82, 284–298. [Google Scholar]
  33. Hay, P.J.; Wadt, W.R. Ab initio effective core potentials for molecular calculations. Potentials for K to Au including the outermost core orbitals. J. Chem. Phys. 1985, 82, 299–310. [Google Scholar] [CrossRef]
  34. Cora, F.; Patel, A.; Harrison, N.M.; Dovesi, R.; Catlow, C.R.A. An ab initio Hartree-Fock study of the cubic and tetragonal phases of bulk tungsten trioxide. J. Am. Chem. Soc. 1996, 118, 12174–12182. [Google Scholar] [CrossRef]
  35. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Perdew, J.P.; Wang, Y. Accurate and simple density functional for the electronic exchange energy: Generalized gradient approximation. Phys. Rev. B 1986, 33, 8800–8802. [Google Scholar] [CrossRef]
  37. Perdew, J.P.; Wang, Y. Erratum: Accurate and simple density functional for the electronic exchange energy: Generalized gradient approximation. Phys. Rev. B 1989, 40, 3399. [Google Scholar] [CrossRef]
  38. Perdew, J.P.; Wang, Y. Accurate and simple analytic representation of the electron-gas correlation energy. Phys. Rev. B 1992, 45, 13244–13249. [Google Scholar] [CrossRef] [PubMed]
  39. Eglitis, R.I.; Rohlfing, M. First-principles calculations of the atomic and electronic structure of SrZrO3 and PbZrO3 (001) and (011) surfaces. J. Phys. Condens. Matter 2010, 22, 415901. [Google Scholar] [CrossRef] [Green Version]
  40. Eglitis, R.I.; Popov, A.I. Systematic trends in (001) surface ab initio calculations of ABO3 perovskites. J. Saudi Chem. Soc. 2018, 22, 459–468. [Google Scholar] [CrossRef]
  41. Vassilyeva, A.F.; Eglitis, R.I.; Kotomin, E.A.; Dauletbekova, A.K. Ab initio calculations of MgF2 (001) and (011) surface structure. Phys. B 2010, 405, 2125–2127. [Google Scholar] [CrossRef]
  42. Shi, H.; Eglitis, R.I.; Borstel, G. Ab initio calculations of the CaF2 electronic structure and F centers. Phys. Rev. B 2005, 72, 045109. [Google Scholar] [CrossRef]
  43. Van Benthem, K.; Elsasser, C.; French, R.H. Bulk electronic structure of SrTiO3: Experiment and theory. J. Appl. Phys. 2001, 90, 6156–6164. [Google Scholar] [CrossRef] [Green Version]
  44. Lee, Y.S.; Lee, J.S.; Noh, T.W.; Byun, D.Y.; Yoo, K.S.; Yamaura, K.; Takayama-Muromachi, E. Systematic trends in the electronic structure parameters of the 4d transition-metal oxides SrMO3 (M = Zr, Mo, Ru and Rh). Phys. Rev. B 2003, 67, 113101. [Google Scholar] [CrossRef] [Green Version]
  45. Robertson, J. Band offsets of wide-band-gap oxides and implications for future electronic devices. J. Vacuum. Sci. Technol. B 2000, 18, 1785–1791. [Google Scholar] [CrossRef]
  46. Lisitsyn, V.M.; Lisitsyna, L.A.; Popov, A.I.; Kotomin, E.A.; Abuova, F.U.; Akilbekov, A.; Maier, J. Stabilization of primary mobile radiation defects in MgF2 crystals. Nucl. Instrum. Methods B 2016, 374, 24–28. [Google Scholar] [CrossRef]
  47. Rubloff, G.W. Far-Ultraviolet Reflectance Spectra and the electronic structure of ionic crystals. Phys. Rev. B 1972, 5, 662–684. [Google Scholar] [CrossRef]
  48. Bochicchio, R.C.; Reale, H.F. On the nature of crystalline bonding: Extension of statistical population analysis to two- and three-dimensional crystalline systems. J. Phys. B At. Mol. Opt. Phys. 1993, 26, 4871–4883. [Google Scholar] [CrossRef]
  49. Eglitis, R.I.; Piskunov, S. First principles calculations of SrZrO3 bulk and ZrO2-terminated (001) surface F centers. Comput. Condens. Matter 2016, 7, 1–6. [Google Scholar] [CrossRef] [Green Version]
  50. Jia, W.; Vikhnin, V.S.; Liu, H.; Kapphan, S.; Eglitis, R.; Usvyat, D. Critical effects in optical response due to charge transfer vibronic excitions and their structure in perovskite-like systems. J. Lumin. 1999, 83, 109–113. [Google Scholar] [CrossRef]
  51. Balászi, C.; Farkas-Jahnke, M.; Kotsis, I.; Petrás, L.; Pfeifer, J. The observation of cubic tungsten trioxide at high-temperature dehydration of tungsten acid hydrate. Solid State Ion. 2001, 141–142, 411–416. [Google Scholar] [CrossRef]
  52. Cora, F.; Stachiotti, M.G.; Catlow, C.R.A.; Rodriguez, C.O. Transition Metal Oxide Chemistry: Electronic Structure Study of WO3, ReO3 and NaWO3. J. Phys. Chem. B 1997, 101, 3945–3952. [Google Scholar] [CrossRef]
  53. Hellwege, K.H.; Hellwege, A.M. Ferroelectrics and Related Substances; Landolt-Bornstein, New Series, Group III; Springer: Berlin/Heidelberg, Germany, 1969; Volume 3. [Google Scholar]
  54. Eglitis, R.I.; Vanderbilt, D. Ab initio calculations of BaTiO3 and PbTiO3 (001) and (011) surface structures. Phys. Rev. B 2007, 76, 155439. [Google Scholar] [CrossRef] [Green Version]
  55. Eglitis, R.I. Ab initio calculations of SrTiO3, BaTiO3, PbTiO3, CaTiO3, SrZrO3, PbZrO3 and BaZrO3 (001), (011) and (111) surfaces as well as F centers, polarons, KTN solid solutions and Nb impurities therein. Int. J. Mod. Phys. B 2014, 28, 1430009. [Google Scholar] [CrossRef] [Green Version]
  56. Eglitis, R.I. Comparative first-principles calculations of SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001), (011) and (111) surfaces. Ferroelectrics 2015, 483, 53–67. [Google Scholar] [CrossRef]
  57. Mabud, S.A.; Glazer, A.M. Lattice parameters and birefringence in PbTiO3 single crystals. J. Appl. Cryst. 1979, 12, 49–53. [Google Scholar] [CrossRef]
  58. Eglitis, R.I.; Vanderbilt, D. Ab initio calculations of the atomic and electronic structure of CaTiO3 (001) and (011) surfaces. Phys. Rev. B 2008, 78, 155420. [Google Scholar] [CrossRef] [Green Version]
  59. Eglitis, R.I. Comparative ab initio calculations of SrTiO3 and CaTiO3 polar (111) surfaces. Phys. Stat. Sol. B 2015, 252, 635–642. [Google Scholar] [CrossRef]
  60. Ali, R.; Yashima, M. Space group and crystal structure of the perovskite CaTiO3 from 296 to 1720 K. J. Solid State Chem. 2005, 178, 2867–2872. [Google Scholar] [CrossRef]
  61. Eglitis, R.I.; Kleperis, J.; Purans, J.; Popov, A.I.; Jia, R. Ab initio calculations of CaZrO3 (011) surfaces: Systematic trends in polar (011) surface calculations of ABO3 perovskites. J. Mater. Sci. 2020, 55, 203–217. [Google Scholar] [CrossRef]
  62. Piskunov, S.; Eglitis, R.I. Comparative ab initio calculations of SrTiO3/BaTiO3 and SrZrO3/PbZrO3 (001) heterostructures. Nucl. Instr. Methods Phys. Res. B 2016, 374, 20–23. [Google Scholar] [CrossRef]
  63. Eglitis, R.I. First-principles calculations of BaZrO3 (001) and (011) surfaces. J. Phys. Condens. Matter 2007, 19, 356004. [Google Scholar] [CrossRef] [Green Version]
  64. Eglitis, R.I.; Purans, J.; Popov, A.I.; Jia, R. Systematic trends in YAlO3, SrTiO3, BaTiO3, BaZrO3 (001) and (111) surface ab initio calculations. Int. J. Mod. Phys. B 2019, 33, 1950390. [Google Scholar] [CrossRef]
  65. Eglitis, R.I.; Vanderbilt, D. First-principles calculations of atomic and electronic structure of SrTiO3 (001) and (011) surfaces. Phys. Rev. B 2008, 77, 195408. [Google Scholar] [CrossRef]
  66. Eglitis, R.I.; Popov, A.I. Comparative ab initio calculations for ABO3 perovskite (001), (011) and (111) surfaces as well as YAlO3 (001) surfaces and F centers. J. Nano Electron. Phys. 2019, 11, 01001. [Google Scholar] [CrossRef] [Green Version]
  67. Eglitis, R.I. Theoretical modelling of the energy surface (001) and topology of CaZrO3 perovskite. Ferroelectrics 2008, 483, 75–85. [Google Scholar] [CrossRef]
  68. Eglitis, R.I.; Piskunov, S.; Zhukovskii, Y.F. Ab initio calculations of PbTiO3/SrTiO3 (001) heterostructures. Phys. Stat. Sol. C 2016, 13, 913–920. [Google Scholar] [CrossRef]
  69. Eglitis, R.I.; Popov, A.I. Ab initio calculations for the polar (001) surfaces of YAlO3. Nucl. Instr. Methods Phys. Res. B 2018, 434, 1–5. [Google Scholar] [CrossRef]
  70. Koffyberg, F.P.; Dwight, K.; Wold, A. Interband transitions of semiconducting oxides determined from photoelectrolysis spectra. Solid State Commun. 1979, 30, 433–437. [Google Scholar] [CrossRef]
  71. Wemple, S.H. Polarization Fluctuations and the Optical-Absorption Edge in BaTiO3. Phys. Rev. B 1970, 2, 2679–2689. [Google Scholar] [CrossRef]
  72. Peng, C.H.; Chang, J.F.; Desu, S. Optical Properties of PZT, PLZT and PNZT Thin Films. Mater. Res. Soc. Symp. Proc. 1991, 243, 21–26. [Google Scholar] [CrossRef]
  73. Ueda, K.; Yanagi, H.; Noshiro, R.; Hosono, H.; Kawazoe, H. Vacuum ultraviolet reflectance and electron energy loss spectra of CaTiO3. J. Phys. Condens. Matter 1998, 10, 3669–3677. [Google Scholar] [CrossRef]
  74. Hikita, T.; Hanada, T.; Kudo, M.; Kawai, M. Structure and electronic state of the TiO2 and SrO-terminated SrTiO3 (100) surface. Surf. Sci. 1993, 287, 377–381. [Google Scholar] [CrossRef]
  75. Bickel, N.; Schmidt, G.; Heinz, K.; Muller, K. Ferroelectric relaxation of the SrTiO3 (100) surface. Phys. Rev. Lett. 1989, 62, 2009–2011. [Google Scholar] [CrossRef] [PubMed]
  76. Eglitis, R.I.; Purans, J.; Gabrusenoks, J.; Popov, A.I.; Jia, R. Comparative ab initio calculations of ReO3, SrZrO3, BaZrO3, PbZrO3 and CaZrO3 (001) surfaces. Crystals 2020, 10, 745. [Google Scholar] [CrossRef]
  77. Car, R.; Parrinello, M. Unified approach for molecular dynamics and density-functional theory. Phys. Rev. Lett. 1985, 55, 2471–2474. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Ab initio calculated and experimentally measured bulk Γ–Γ band gaps for SrTiO3, SrZrO3, BaZrO3, MgF2 and CaF2 obtained by means of different exchange–correlation functionals: (1) PWGGA; (2) B3LYP; (3) B3PW; (4) Experiment; (5) HF.
Figure 1. Ab initio calculated and experimentally measured bulk Γ–Γ band gaps for SrTiO3, SrZrO3, BaZrO3, MgF2 and CaF2 obtained by means of different exchange–correlation functionals: (1) PWGGA; (2) B3LYP; (3) B3PW; (4) Experiment; (5) HF.
Crystals 11 00455 g001
Figure 2. Side view of the nine-layer TiO2-terminated ATiO3 perovskite (001) surface.
Figure 2. Side view of the nine-layer TiO2-terminated ATiO3 perovskite (001) surface.
Crystals 11 00455 g002
Figure 3. Side view of the nine-layer WO2-terminated WO3 polar (001) surface.
Figure 3. Side view of the nine-layer WO2-terminated WO3 polar (001) surface.
Crystals 11 00455 g003
Figure 4. Our ab initio-calculated bulk (1) as well as BO2-terminated (001) surface (2). B–O chemical bond populations (in e) for WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3.
Figure 4. Our ab initio-calculated bulk (1) as well as BO2-terminated (001) surface (2). B–O chemical bond populations (in e) for WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3.
Crystals 11 00455 g004
Figure 5. Our ab initio-calculated WO2-terminated WO3 and TiO2-terminated SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) surface Γ–Γ band gaps (line 1). Experimentally measured bulk Γ–Γ band gaps (line 2). Our ab initio-calculated WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 bulk Γ–Γ band gaps (line 3).
Figure 5. Our ab initio-calculated WO2-terminated WO3 and TiO2-terminated SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) surface Γ–Γ band gaps (line 1). Experimentally measured bulk Γ–Γ band gaps (line 2). Our ab initio-calculated WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 bulk Γ–Γ band gaps (line 3).
Crystals 11 00455 g005
Table 1. By means of different exchange–correlation functionals calculated SrTiO3, SrZrO3, BaZrO3, MgF2 and CaF2 bulk Γ–Γ band gaps (eV). Experimental bulk band gaps at the Γ-point are listed for comparison.
Table 1. By means of different exchange–correlation functionals calculated SrTiO3, SrZrO3, BaZrO3, MgF2 and CaF2 bulk Γ–Γ band gaps (eV). Experimental bulk band gaps at the Γ-point are listed for comparison.
MethodSrTiO3 [30]SrZrO3 [39]BaZrO3 [40]MgF2 [41]CaF2 [42]
Experiment3.75 [43]5.6 [44] 5.3 [45] 13.0 [46]12.1 [47]
B3PW3.965.304.939.4810.96
B3LYP3.895.314.799.4210.85
HF12.3313.5412.9619.6520.77
PWGGA2.313.533.246.948.51
PBE2.353.52-6.918.45
Table 2. Our ab initio-calculated and experimentally measured WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 bulk lattice constants [51,52,53,54,55,56,57,58,59,60].
Table 2. Our ab initio-calculated and experimentally measured WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 bulk lattice constants [51,52,53,54,55,56,57,58,59,60].
MaterialMethodTheoryExperiment
WO3B3LYP3.7753.71–3.75 [51]
FP-LMTO3.78 [52]
SrTiO3B3PW3.9043.89 [53]
BaTiO3B3PW4.008 4.00 [53,54]
PbTiO3B3PW3.9363.97 [54,57]
CaTiO3B3PW3.8513.8967 [58,59,60]
Table 3. Our calculated atomic charges Q(e) as well as bond populations P(e) in WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 bulk materials.
Table 3. Our calculated atomic charges Q(e) as well as bond populations P(e) in WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 bulk materials.
Bulk MaterialsWO3SrTiO3BaTiO3PbTiO3CaTiO3
Ion PropertyB3LYPB3PWB3PWB3PWB3PW
AQ-+1.871+1.797+1.354+1.782
P-−0.010−0.034+0.016+0.006
OQ−1.032−1.407−1.388−1.232−1.371
P+0.142+0.088+0.098+0.098+0.084
BQ+3.095+2.351+2.367+2.341+2.330
Table 4. Ab initio-calculated WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 bulk band gaps for the cubic phase at the Γ–Γ point. Our B3LYP and B3PW calculations were compared with the relevant experimental data.
Table 4. Ab initio-calculated WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 bulk band gaps for the cubic phase at the Γ–Γ point. Our B3LYP and B3PW calculations were compared with the relevant experimental data.
CrystalMethodOptical Bulk Band Gap at Γ–Γ Point
Ab Initio CalculationsExperimental Results
WO3B3LYP4.953.74 [70]
SrTiO3B3PW3.963.75 [43]
BaTiO3B3PW3.553.2 [71]
PbTiO3B3PW4.323.4 [72]
CaTiO3B3PW4.183.5 [73]
Table 5. WO2-terminated WO3 as well as TiO2-terminated SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) surface atom relaxation for upper-three surface layers (in percent of the bulk lattice constant).
Table 5. WO2-terminated WO3 as well as TiO2-terminated SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) surface atom relaxation for upper-three surface layers (in percent of the bulk lattice constant).
Surfaces, (001)WO3SrTiO3BaTiO3PbTiO3CaTiO3
LayerIonWO2-Term.TiO2-Term.TiO2-Term.TiO2-Term.TiO2-Term.
MethodB3LYPB3PWB3PWB3PWB3PW
1B−2.07−2.25−3.08−2.81−1.71
O+0.42−0.13−0.35+0.31−0.10
2AAbsent+3.55+2.51+5.32+2.75
O+0.11+0.57+0.38+1.28+1.05
3B−0.01----
O0.00----
Table 6. Our B3LYP- or B3PW-calculated surface rumplings s and relative displacements Δdij between the three near-surface planes for the WO2-terminated WO3 and the TiO2-terminated BaTiO3, PbTiO3, CaTiO3 and SrTiO3 (001) surfaces as a percent of the bulk material lattice constant. The available experimental data are listed for comparison purposes.
Table 6. Our B3LYP- or B3PW-calculated surface rumplings s and relative displacements Δdij between the three near-surface planes for the WO2-terminated WO3 and the TiO2-terminated BaTiO3, PbTiO3, CaTiO3 and SrTiO3 (001) surfaces as a percent of the bulk material lattice constant. The available experimental data are listed for comparison purposes.
MaterialMethodWO2- or TiO2-Terminated (001) Surface
sΔd12Δd23
WO3B3LYP+2.49--
BaTiO3B3PW+2.73−5.59+2.51
PbTiO3B3PW+3.12−8.13+5.32
CaTiO3B3PW+1.61−4.46+2.75
SrTiO3B3PW+2.12−5.80+3.55
RHEED exp. [74]+2.6+1.8+1.3
LEED exp. [75]+2.1 ± 2+1 ± 1−1 ± 1
Table 7. Our B3LYP- or B3PW-calculated atomic shift magnitudes D (in Å) as well as the effective atomic charges Q (in e) and nearest atomic chemical bond populations P (in e) for the WO2- and TiO2-terminated WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) surfaces.
Table 7. Our B3LYP- or B3PW-calculated atomic shift magnitudes D (in Å) as well as the effective atomic charges Q (in e) and nearest atomic chemical bond populations P (in e) for the WO2- and TiO2-terminated WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) surfaces.
WO2 and TiO2-Term. (001) SurfacesWO3SrTiO3BaTiO3PbTiO3CaTiO3
LayerProperty IonWO2-Ter.TiO2-Ter.TiO2-Ter.TiO2-Ter.TiO2-Ter.
1DB−0.078−0.088−0.123−0.111−0.066
Q +2.783+2.291+2.307+2.279+2.278
P +0.108+0.118+0.126+0.114+0.114
DO+0.016−0.005−0.014+0.012−0.004
Q −1.146−1.296−1.280−1.184−1.267
P −0.014−0.014−0.038+0.044+0.016
2DA -+0.139+0.101+0.209+0.106
Q -+1.850+1.767+1.275+1.754
P -−0.008−0.030+0.008+0.006
DO+0.004+0.022+0.015+0.050+0.041
Q −0.925−1.365−1.343−1.167−1.324
P +0.064+0.080+0.090+0.080+0.086
3DB−0.0004----
Q +3.001+2.348+2.365+2.335+2.326
P +0.144+0.096+0.104+0.108+0.090
DO 0.000----
Q −1.037−1.384−1.371−1.207−1.354
P −0.032−0.010−0.034+0.018+0.008
Table 8. Our B3LYP or B3PW calculated W–O or Ti–O chemical bond populations for WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 bulk and for WO2- or TiO2-terminated (001) surfaces (in e).
Table 8. Our B3LYP or B3PW calculated W–O or Ti–O chemical bond populations for WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 bulk and for WO2- or TiO2-terminated (001) surfaces (in e).
Material MethodW-O or Ti-O Chemical Bond Populations
BulkWO2, TiO2-Term. (001)
WO3B3LYP+0.142+0.108 (W(I)—O(I))
+0.142+0.278 (W(I)—O(II))
SrTiO3B3PW+0.088+0.118
BaTiO3B3PW+0.098+0.126
PbTiO3B3PW+0.098+0.114
CaTiO3B3PW+0.084+0.114
Table 9. B3LYP- or B3PW-calculated Γ–Γ band gaps for WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 bulk as well as their WO2 or TiO2-terminated (001) surfaces.
Table 9. B3LYP- or B3PW-calculated Γ–Γ band gaps for WO3, SrTiO3, BaTiO3, PbTiO3 and CaTiO3 bulk as well as their WO2 or TiO2-terminated (001) surfaces.
MaterialMethodCalculated Band Gap at Γ–Γ Point
BulkWO2, TiO2-Term. (001)
WO3B3LYP4.951.16
SrTiO3B3PW3.963.95
BaTiO3B3PW3.552.96
PbTiO3B3PW4.323.18
CaTiO3B3PW4.183.30
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Eglitis, R.I.; Purans, J.; Jia, R. Comparative Hybrid Hartree-Fock-DFT Calculations of WO2-Terminated Cubic WO3 as Well as SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) Surfaces. Crystals 2021, 11, 455. https://doi.org/10.3390/cryst11040455

AMA Style

Eglitis RI, Purans J, Jia R. Comparative Hybrid Hartree-Fock-DFT Calculations of WO2-Terminated Cubic WO3 as Well as SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) Surfaces. Crystals. 2021; 11(4):455. https://doi.org/10.3390/cryst11040455

Chicago/Turabian Style

Eglitis, R. I., Juris Purans, and Ran Jia. 2021. "Comparative Hybrid Hartree-Fock-DFT Calculations of WO2-Terminated Cubic WO3 as Well as SrTiO3, BaTiO3, PbTiO3 and CaTiO3 (001) Surfaces" Crystals 11, no. 4: 455. https://doi.org/10.3390/cryst11040455

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

Article Metrics

Back to TopTop