1. Introduction
Crucial role of surface as well as interface phenomena taking place in the ABO
3 (in the following called ABO for simplicity) perovskites are very serious issues in today’s condensed matter physics [
1,
2,
3,
4,
5,
6,
7,
8,
9,
10,
11,
12,
13,
14,
15,
16,
17,
18,
19,
20,
21,
22,
23,
24,
25]. All BaSnO
3 (BSO), BaTiO
3 (BTO), SrTiO
3 (STO), PbTiO
3 (PTO), CaTiO
3 (CTO), BaZrO
3 (BZO), SrZrO
3 (SZO), and CaZrO
3 (CZO) perovskites belong to the group of ABO perovskite oxides [
26,
27,
28,
29]. In our case, A = Ba, Sr, Pb, or Ca, whereas B = Sn, Ti, or Zr. ABO perovskites have plenty of commercially essential functions. Applications include actuators, capacitors, charge storage apparatus, and countless others [
30,
31]. For example, BSO may be used as the protonic conductor [
32], with application potential for fuel cells [
32]. BTO is examined as one of the inexpensive preparation price substitutes for wide band gap semiconductors with application prospects in numerous optoelectric devices [
33]. Doped STO is an excellent anode material for solid oxide fuel cells [
34] as well as a perfect choice for photocatalytic applications [
35]. PTO ferroelectric perovskite is an interesting material for numerous high-temperature as well as high- frequency piezoelectric applications [
36]. CTO perovskite oxide has received recognition in the last years due to strong catalytic activity [
37]. BZO perovskite is attractive as a thermal barrier coating material, used in gas turbine engines, which work at elevated temperatures [
38]. SZO perovskite has been widely used as a catalyst [
39], luminescent material [
40], and proton conductor [
41]. CZO perovskite has numerous technologically important applications, including capacitors and resonators, and also as humidity sensors [
42]. For that reason, it is evident that in the last 25 years, ABO perovskite (001) surfaces were massively explored experimentally as well as theoretically [
43,
44,
45,
46,
47,
48,
49,
50,
51,
52,
53,
54,
55,
56,
57,
58,
59,
60,
61,
62,
63,
64,
65,
66,
67,
68,
69,
70,
71,
72,
73,
74,
75,
76,
77,
78,
79,
80,
81,
82].
In contrast to the ABO perovskite neutral (001) surfaces, their charged, polar, and thereby very complex (111) surfaces are considerably less studied. For example, to the best of our knowledge, BSO (111) surfaces have never been investigated before, neither experimentally nor theoretically. Accordingly, in this paper, we performed the first in the world ab initio calculations, dealing with polar and charged BSO (111) surfaces. BTO perovskite (111) surfaces were analyzed experimentally by Hagendorf et al. [
83,
84] by means of STM, XPS, and LEED methods. Recently, Chun et al. [
8] examined the (111) surface termination of a BTO single crystal employing the combined density functional theory (DFT) as well as X-ray photoelectron spectroscopy (XPS) methods. They computed the (111) surface stability of the BTO stoichiometric Ti and BaO
3-terminations applying the DFT + U formalism [
8]. Finally, Eglitis [
85] performed ab initio hybrid DFT computations for BTO (111) surfaces and demonstrated that the surface energy of the BaO
3-terminated (111) surface is considerably larger than for the Ti-terminated (111) surface. Pojani et al. [
86] performed semi-empirical Hartree–Fock (HF) calculations and discussed the polarity effects on the STO (111) surfaces. Biswas et al. [
87] detected the upper atomic layer of STO (001), (011), and (111) surfaces using the time of flight mass spectroscopy. These authors found [
87] that all (001), (011), and (111) surface orientations exhibit the Ti-rich surface [
87]. Sekiguchi et al. [
88] investigated the changes in STO (111) polar surface structures by AFM, AES, and XPS experimental techniques as a function of the atmosphere [
88]. Tanaka et al. [
89] observed the clean (111) surfaces of reduced STO crystals by means of STM and RHEED experiments [
89]. They observed two different STO (111) surface structures [
89]. Namely, the first possible STO (111) surface structure has the SrO
3-x outermost layer, whereas the second possible STO (111) surface structure has the Ti outermost layer [
89]. Lastly, Eglitis [
90] performed ab initio computations of SrO
3- and Ti-terminated STO (111) polar surface atomic relaxations, surface rumplings, and energies, as well as charge redistributions and Γ-Γ band gaps [
90]. Using the GGA exchange-correlation functional, Pang et al. [
91] calculated the structural and electronic properties as well as the stabilities of four different terminations of cubic PTO (111) surfaces. For instance, they computed the directly cleaved Ti- and PbO
3-terminated as well as constructed TiO- and PbO
2-terminated PTO (111) polar surfaces [
91]. Eglitis [
85] performed ab initio B3LYP calculations for PbO
3 and Ti-terminated PTO (111) surfaces and found that for both terminations, the PTO (111) surface energies are considerably larger than the PTO (001) surface energies [
3]. Liu et al. [
92] constructed the stoichiometric as well as nonstoichiometric terminations for the CTO (111) surfaces. The LDA computations for the CTO (111) surface and cleavage energies, as well as surface electronic and atomic structure, and surface grand potential, were performed [
92]. Eglitis [
90,
93] performed ab initio computations for the polar CTO (111) surfaces using the B3LYP hybrid exchange-correlation functional. The atomic and electronic structure as well as surface energies and Ti-O chemical bond populations of the Ti and CaO
3-terminated polar CTO (111) surfaces were computed [
90,
93]. The only existing ab initio computations dealing with BaO
3- and Zr-terminated polar BaZrO
3 (111) surfaces were performed by Eglitis [
94,
95]. Namely, these authors performed ab initio B3LYP computations for BZO polar (111) surface atomic relaxations and electronic structure as well as energetics for two possible BaO
3 and Zr polar (111) surface terminations [
94,
95]. The only available ab initio computations dealing with polar SrZrO
3 (111) surfaces were performed by Eglitis et al. [
85,
96]. Specifically, Eglitis et al. [
85,
96] performed ab initio hybrid B3LYP computations for surface relaxations and electronic structure as well as surface energies of Zr- and SrO
3-terminated SrZrO
3 perovskite polar (111) surfaces. Lastly, the only recent existing and preliminary ab initio B3LYP computations dealing with CZO perovskite polar (111) CaO
3- and Zr-terminated surface atomic relaxations and surface energies were performed by Eglitis and Jia [
11].
According to the XRD measurement results obtained by Janifer et al. [
97], barium stannate (BSO) is a single-phase cubic perovskite [
97,
98]. Moreover, BSO has a wide optical band gap equal to 3.1 eV [
97,
98], and the cubic lattice parameter
a is identical to 4.119 Å [
97,
98]. BTO perovskite exhibits three phase transitions [
99,
100]. At high temperatures, BTO perovskite has a cubic structure with a symmetrical group
) [
99]. The BTO perovskite structure changes from this cubic
) to a tetragonal structure (
P4mm) at 403 K temperature [
99]. Then, at 278 K temperature, the BTO perovskite structure changes to orthorhombic (
Amm2) [
99]. Lastly, the BTO perovskite structure changes to rhombohedral (
R3m) at a temperature of 183 K [
99,
100]. According to experiments performed by Wemple [
101], the BTO room temperature band gaps are 3.38 eV and 3.27 eV, accordingly, for light polarized parallel as well as perpendicular to the ferroelectric
c axis [
101]. STO perovskite has only one structural phase transition at a temperature of 110 K [
102]. This phase transition reduces the STO perovskite symmetry from high symmetry cubic to tetragonal [
102]. The experimentally detected STO perovskite direct (Γ-Γ) band gap energy in the room temperature cubic phase is equal to 3.75 eV [
103]. PTO perovskite displays the single-phase transition at 763 K temperature from the high-temperature cubic phase to the tetragonal ferroelectric ground state [
104]. The PTO perovskite Γ-Γ band gap, measured at room temperature in its tetragonal ferroelectric ground state, is equal to 3.4 eV [
105]. Ali et al. [
106], using the careful Rietveld analysis of the neutron as well as X-ray powder diffraction data, discovered that the CTO perovskite displays two structural phase transitions [
106]. Namely, from a high-temperature cubic CTO structure
) at 1634 ± 13 K temperature to a tetragonal structure (
I4/mcm) [
106]. Finally, at 1498 ± 25 K temperature, CTO perovskite displays a phase transition to the ground state orthorhombic structure (
Pbnm) [
106]. The CTO perovskite Γ-Γ band gap, according to the experimental results obtained by Ueda et al. in the orthorhombic phase [
107], is approximately 3.5 eV [
107]. According to Knight [
108], the BZO perovskite is cubic at all measured temperatures within the temperature range from 4.2 K to 450 K [
108]. Namely, BZO always has the cubic perovskite structure with the symmetry group
) [
108]. Therefore, the experimental BZO Γ-Γ band gap is equal to 5.3 eV [
105]. SZO perovskite goes through three structural phase transitions [
109]. Initially, below 970 K temperature, SZO perovskite is orthorhombic with the symmetry group
Pnma [
109]. In the temperature range between 970 K and 1020 K, the SZO perovskite belongs to the other orthorhombic symmetry group
Cmcm [
109]. At 1020 K temperature, SZO perovskite transfers into the tetragonal structure with the symmetry group
I4/mcm [
109]. Finally, at a temperature above 1360 K, the SZO perovskite turns into a cubic structure with the symmetry group
) [
109]. According to the optical conductivity measurements performed by Lee et al. [
110], the SZO Γ-Γ gap within the orthorhombic phase, which is stable at room temperature, is equal to 5.6 eV [
110]. CZO perovskite, at low temperatures, has an orthorhombic structure (
Pbnm), which is stable up to temperatures of 2173 K ± 100 [
111,
112]. At higher temperatures, CZO perovskite has a cubic structure with a symmetrical group
) [
111,
112]. According to Rosa et al. [
113], the experimental CZO band gap of the orthorhombic structure at room temperature is equal to 5.7 eV [
113].
The intention of our review paper was to execute essential additional computations dealing mostly with BSO neutral (001) and especially polar (111) surfaces that have never been studied before. After completing necessary supplementary first principles computations for ABO perovskite (001) as well as (111) surfaces, we carefully analyzed our results and identified the systematic tendencies frequent for all eight BSO, BTO, STO, PTO, CTO, BZO, SZO, and CZO perovskite neutral (001) [
43,
44,
45,
46,
47,
48,
49,
50,
51,
52,
53,
54,
55,
56,
57,
58,
59,
60,
61,
62,
63,
64,
65,
66,
67,
68,
69,
70,
71,
72,
73,
74,
75,
76,
77,
78,
79,
80,
81,
82,
114,
115,
116,
117,
118,
119,
120] and polar (111) [
83,
84,
85,
86,
87,
88,
89,
90,
91,
92,
93,
94,
95,
96] surfaces. This paper research primarily investigates the surface properties. Concretely, the atomic relaxation, surface energy, chemical bond covalency, as well as electronic band gaps of ABO perovskites (e.g., BSO, BTO, STO, etc.), are established at both neutral (001) and polar (111) surfaces. The study aims to enhance the understanding of how surface characteristics influence the properties of these materials. We want to emphasize that the topic dealing with the theoretical investigation of ABO perovskite surfaces is both original and highly relevant to the field of condensed matter physics and materials science. While ABO perovskites have been extensively studied, the specific examination of polar (111) surfaces in the context of BSO, particularly with B3LYP computations, addresses a notable gap in the literature. Previous research primarily focused on (001) surfaces [
1,
2,
3,
4,
5,
7,
8,
9,
10,
11,
22,
43,
45,
48,
50,
56,
57,
58,
59,
60,
61], making this investigation significant for understanding surface phenomena that impact applications in electronics and catalysis. This study adds considerable depth to the existing literature by providing the first computational analysis of polar (111) surfaces for BSO, and it includes a comprehensive study among various perovskites. The detailed examination of atomic relaxations, surface energies, and bond covalency enhances understanding of how surface properties differ for these ABO perovskite materials compared to bulk behavior, offering insights that can guide future experimental and theoretical investigations.
2. Simulation Approach and Surface Models
In order to carry out the first principles of DFT-B3PW or DFT-B3LYP computations, we employed the CRYSTAL computer program [
121]. The CRYSTAL code makes use of Gaussian-type functions (GTFs) localized on all perovskite atoms. They are the basis for the expansion of the crystalline orbitals [
122]. The superiority of the CRYSTAL ab initio code [
121,
122] is its capability to compute confined 2D (001) and (111) ABO perovskite slabs without forced periodicity along the
z-axis [
121]. In order to make use of the linear combination of atomic orbitals (LCAO-GTF) method [
121,
122], it is advantageous to utilize the optimized basis sets (BS) [
121]. The optimized BS for BTO, STO, and PTO perovskites was developed in Ref. [
123]. The BS for Sn was taken from the CRYSTAL computer code [
121]. The relevant BS, used for CTO, BZO, SZO, and CZO perovskites and their surfaces, are described in Refs. [
2,
124,
125,
126]. All our ABO perovskite bulk as well as (001) and (111) surface computations were performed by means of the B3PW [
127,
128] or B3LYP [
129] hybrid exchange-correlation functionals. Both B3PW and B3LYP functionals give comparable results for ABO perovskites [
2,
3,
4,
9,
56]. However, it is important to note that the hybrid exchange-correlation functionals, like B3PW or B3LYP, allow us to achieve an excellent agreement with the experiment for the Γ-Γ band gaps of the related complex oxide materials, whereas the GGA-PBE and LDA exchange-correlation functionals, as a rule, underestimate the relevant Γ-Γ band gaps [
2,
3,
4,
9,
56]. On the other hand, it is well known that the HF approach considerably overestimates the Γ-Γ band gaps [
2,
3,
4,
9,
56]. We performed the reciprocal space integration by checking out the Brillouin zone for the 5 atom ABO perovskite cubic unit cell, applying the 8 × 8 × 8 times increased Pack Monkhorst mesh [
130] for the ABO perovskite bulk [
2,
3,
4,
11,
124,
125,
126,
131] as well as the 8 × 8 × 1 times increased mesh for their (001) and (111) surfaces [
2,
3,
4,
11,
85,
90,
124,
125,
126,
131].
The ABO perovskite (001) surfaces were described using 2D slabs (
Figure 1 and
Figure 2). These slabs subsist from nine planes located perpendicular to the [001] crystal direction.
Namely, to compute ABO perovskite (001) surfaces, we considered slabs consisting of nine alternating BO
2 and AO layers (
Figure 1 and
Figure 2). The mirror symmetries of the slabs were retained regarding their central layers (
Figure 1 and
Figure 2). Our computed 23-atom-containing slab with BO
2-terminated surfaces as well as the 22-atom slab with AO-terminated surfaces are depicted in
Figure 1 and
Figure 2, respectively. Both these slabs are non-stoichiometric. They have unit-cell formulas A
4B
5O
14 and A
5B
4O
13, respectively (
Figure 1 and
Figure 2). The succession of the ABO-perovskite (001) surface layers as well as the definitions of the interplane separations Δd
12, Δd
23, and the surface rumpling s are pictured in
Figure 1.
In contrast to the ABO perovskite neutral (001) surfaces, the polar (111) surfaces subsist of charged AO
3 and B planes, as depicted in
Figure 3. The ABO perovskite polar (111) surfaces have been characterized with 2D slabs, subsisting of nine planes perpendicular to the [
111] direction in the crystal. In order to compute the ABO perovskite polar (111) surfaces, we employed slabs subsisting of nine alternating B and AO
3 layers (
Figure 3 and
Figure 4). The first slab is terminated by B planes and subsists of a supercell consisting of 21 atoms (B-AO
3-B-AO
3-B-AO
3-B-AO
3-B) (
Figure 4a). Another slab is terminated by AO
3 planes (
Figure 4b). The corresponding supercell contains 24 atoms (AO
3-B-AO
3-B-AO
3-B-AO
3-B-AO
3) (
Figure 4b). Both these slabs are non-stoichiometric (
Figure 3 and
Figure 4). They have unit cell formulas A
4B
5O
12 and A
5B
4O
15, respectively (
Figure 4). As we know from earlier ab initio computations, devoted to the polar ABO perovskite (111) surfaces [
11,
86,
92,
132,
133,
134], the powerful electron reallocation takes place for these terminations in order to eliminate the polarity [
11,
86,
92,
132,
133,
134]. Consequently, the AO
3- as well as B-terminated ABO-perovskite (111) surfaces maintain their insulating character, and such computations can thereby be realized [
11,
86,
92,
132,
133,
134].
The first step for the ABO perovskite (001) and (111) surface energy computations is to calculate the relevant cleavage energies. Our computed cleavage energies are equally shared amid the created surfaces [
2,
3,
4,
11,
85,
94,
135]. Namely, the ABO perovskite (001) and (111) surfaces arise as a result of simultaneous (001) as well as (111) cleaved perovskite crystal [
2,
3,
4,
11,
85,
94,
135]. In our executed ABO perovskite (001) surface cleavage energy computations, the nine-layer AO- and BO
2-terminated slabs together embody 45 atoms corresponding to nine ABO perovskite unit cells:
where E
slabunr(AO) and E
slabunr(BO
2) are the unrelaxed AO- and BO
2-terminated ABO perovskite nine layer (001) slab total energies. E
bulk denotes the total energy of the ABO perovskite bulk unit cell containing five atoms. The factor equal to ¼ implies that four surfaces were generated due to the ABO perovskite (001) cleavage. In the second step, the AO- and BO
2-terminated nine-layer (001) slab relaxation energies, taking into account the relaxation of slabs from both sides, were computed, as follows:
where λ is AO or BO
2. E
slabrel(λ) is relaxed from both sides with regard to the AO- or BO
2-terminated (001) slab total energy. E
slabunr(λ) is the total energy for unrelaxed AO or BO
2-terminated ABO perovskite (001) slab. Lastly, the AO- or BO
2-terminated ABO perovskite (001) surface energy should be calculated using the following equation:
With this, we computed the ABO perovskite (111) surface as well as cleavage energies. Again, B- and AO
3-terminated ABO perovskite (111) surfaces are complementary. For that reason, the cleavage energy is the same for both AO
3 and B-terminated ABO perovskite (111) surfaces. Thereby, the cleavage energy for the complementary surface E
cl(AO
3 + B) may be computed from the total energies of nine-layer unrelaxed AO
3- and B-terminated ABO perovskite (111) slabs, as follows:
where E
slabunr(B) is our computed total energy of unrelaxed 21 atoms containing B-terminated ABO perovskite nine-layer (111) slab. E
slabunr(AO
3) is our computed total energy for 24 atoms containing a nine-layer unrelaxed AO
3-terminated ABO perovskite (111) slab. The relaxation energies for AO
3- and B-terminated ABO perovskite (111) surfaces can be obtained via the following equation:
where β = AO
3 or B describes the ABO perovskite (111) surface termination. In the end, the ABO perovskite AO
3- or B-terminated (111) surface energy is equal to the sum of the cleavage (4) and relaxation (5) energies, as follows:
4. Conclusions
As we can see from
Figure 8, the hybrid exchange-correlation functionals B3LYP and B3PW allow us to achieve a fair agreement with the experiments for ABO perovskite bulk Γ-Γ band gaps [
2,
3,
4,
9,
56]. At the same time, the HF method considerably overestimates the bulk Γ-Γ band gaps, whereas the PWGGA method underestimates them [
2,
3,
4,
9,
56]. This was the key reason why we choose the B3LYP and B3PW hybrid exchange-correlation functionals for our ABO perovskite bulk and surface computations [
2,
3,
4,
9,
56].
We completed B3LYP and B3PW computations for AO and BO
2-terminated (001) as well as AO
3- and B-terminated (111) surfaces of BSO, BTO, STO, PTO, CTO, BZO, SZO, and CZO perovskites. We observed that most of the upper-layer atoms for AO- and BO
2-terminated ABO perovskite (001) surfaces relax inward. The two exceptions from this systematic trend are the upward relaxation of oxygen atoms on the TiO
2-terminated PTO (001) surface by (0.31% of
a0) (
Table 4) as well as on the SrO-terminated STO (001) surface by (0.84% of
a0) (
Table 5). In contrast, all second-layer metal atoms relax upward. Also, practically all second-layer oxygen atoms relax upward. Only the oxygen atoms on SnO
2- and BaO-terminated BSO as well as SrO-terminated SZO (001) surfaces relax inward by a very small relaxation magnitude of (0.04, 0.07, and 0.05% of
a0, respectively) (
Table 4 and
Table 5). Lastly, almost all third-layer atoms, again, relax inward.
This tendency is less pronounced for atomic relaxation of first-, second-, and third-layer atoms for AO
3- and B-terminated ABO perovskite (111) surfaces (
Table 10 and
Table 11). Namely, 20 from 24, or 83.33%, of ABO perovskite (111) surface upper layer atoms relax inward. Nevertheless, only 14 from 24, or 58.33%, of ABO perovskite (111) surface second layer atoms relax upward. Finally, again, only 58.33% of the third layer (111) surface atoms relax inward. For almost all ABO perovskites, their (001) surface rumplings
s are considerably larger for AO-terminated compared to BO
2-terminated surfaces. For example, B3PW- or B3LYP-computed surface rumplings
s for AO-terminated STO, BSO, and CTO (001) surfaces (5.66, 1.36, and 7.89) are considerably larger than the respective surface rumplings
s for BO
2-terminated STO, BSO, and CTO (001) surfaces (2.12, 0.70, and 1.61, respectively) (
Table 6).
On the contrary, the ABO perovskite (001) surface energies, for both AO- and BO
2-terminations, are essentially equivalent. The largest computed ABO perovskite (001) surface energy is for the BO
2-terminated BSO (001) surface (1.39 eV), whereas the smallest is for the TiO
2-terminated PTO (001) surface (0.74 eV) (
Table 9). The ABO perovskite polar (111) surface energies always are substantially larger than their neutral (001) surface energies. Namely, the largest ABO perovskite polar (111) surface energy is for the CaO
3-terminated CZO (111) surface (9.62 eV), whereas the smallest is for the Ti-terminated CTO (111) surface (4.18 eV) (
Table 13 and
Figure 15). In most cases, the surface energies of AO
3-terminated ABO perovskite polar (111) surfaces are considerably larger than their B-terminated surface energies. Specifically, for BTO, STO, PTO, CTO, BZO, SZO, and CZO perovskites, their AO
3-terminated (111) surface energy is considerably larger than their B-terminated (111) surface energy (
Table 13 and
Figure 15). The only omission from this systematic trend is BSO perovskite, where the Sn-terminated (111) surface energy (5.20 eV) is 0.07 eV larger than the BaO
3-terminated (111) surface energy (5.13 eV).
Our computations illustrate a noticeable boost in the B-O chemical bond covalency near the ABO perovskite BO
2-terminated (001) surface related to the bulk. Particularly, for BSO, BTO, STO, PTO, CTO, BZO, SZO, and CZO perovskites, their bulk B-O chemical bond covalency (+0.284
e, +0.098
e, +0.088
e, +0.098
e, +0.084
e, +0.108
e, +0.092
e, +0.086
e) increases near the BO
2-terminated ABO perovskite (001) surface (+0.298
e, +0.126
e, +0.118
e, +0.114
e, +0.114
e, +0.132
e, +0.114
e, +0.102
e). The absolutely largest B-O chemical bond covalency is near the SnO
2-terminated BSO perovskite (001) surface (+0.298
e), whereas the smallest B-O chemical bond covalency is between Ti-O atoms (+0.084
e) in the CTO perovskite bulk (
Table 2 and
Table 7). Our computed ABO perovskite bulk Γ-Γ band gaps are almost always reduced near the AO- and BO
2-terminated neutral (001) surfaces as well as in most cases also near the AO
3- and B-terminated polar (111) surfaces.
We want to stress that our B3LYP and B3PW computations of ABO perovskite neutral (001) as well as polar (111) surface characteristics are very helpful in order to interpret processes, where surfaces exhibit a key role [
155,
156,
157]. For instance, in the chemistry and physics of ABO perovskite surface reactions, (001) and (111) surface as well as interface aspect, and adsorption. Specifically, a wide variety of useful technological applications of ABO perovskites counting electrooptical and piezoelectrical devices as well as fuel cells and microelectrodes have inspired their theoretical research [
155,
156,
157].