Tatsuya
Joutsuka
Department of Materials Science and Engineering, Ehime University, 3 Bunkyo-cho, Matsuyama, Ehime 790-8577, Japan. E-mail: joutsuka.tatsuya.zk@ehime-u.ac.jp
First published on 25th March 2025
Strontium titanate (SrTiO3 or STO) is one of the promising photocatalysts for sustainable energy applications. Using the density functional theory (DFT) calculations, we herein study the structural and electronic factors contributing to its high photocatalytic activity and facet dependence. The constrained DFT method revealed that the hole polarons in bulk and surface STO are localized primarily on oxygen atoms. In contrast, electron polarons in bulk STO tend to delocalize over oxygen atoms unless stabilized by oxygen vacancies. The stability of hole polarons is higher at the surface O site of the (110) surface compared to the (001) surfaces. In addition, the oxygen vacancy is stable specifically at the TiO2-terminated (001) surface. These findings provide an atomic-level insight into the relationship between polaron stability and facet dependence of photocatalysis, paving the way for the design of more efficient STO-based photocatalysts.
In photocatalysts, the catalytic activity varies significantly across different crystal facets.1,7,8 Controlling these facets is therefore crucial for optimizing photocatalyst performance. In the experiment of STO,1 it was observed that a hole polaron that is employed in the oxygen-evolution reaction (OER) migrates to the (110) surface, whereas an electron polaron that is employed in hydrogen-evolution reaction (HER) migrates to the (001) surface. This highlights the significance of the stability of electron and hole polarons at the surface in understanding the photocatalytic activity of STO.
Density functional theory (DFT) calculations have been widely employed to clarify the physical properties of STO,9,10 such as the surface structures of the catalysts.11–16 STO has various facets, such as SrO-terminated (001) facet,12 with varying stabilities depending on the experimental conditions.17,18 Therefore, elucidating the facet dependence, which has been shown to be important for various photocatalysts,7,19,20 is crucial to further improve the photocatalytic performance. However, the intricate relationship between the surface structure and the reaction mechanism—particularly the stability of polarons—remains poorly understood.
This study aims to elucidate the stability of polarons and its facet dependence to better understand the reaction mechanism behind the high photocatalytic activity in STO. Our study begins with an analysis of the bulk structure in STO, focusing on the behaviour of hole and electron polarons as well as oxygen vacancies. To simulate the localization of these polarons and the dynamics of polaron transfer, we applied the constrained DFT (CDFT) method21–24 to charge-transfer reactions based on first-principles calculations.7,25,26 The transmission coefficient is also calculated to account for the dynamical effects on the polaron transfer. We then explore the structure of the (001) and (110) surfaces and the facet-dependence stability of polarons to better understand the reaction mechanism of STO photocatalysts. The adsorption of a water molecule and influence of OH group at the STO surface are also considered, given the potential impact on the charge transfer.
The remainder of this paper is structured as follows: Section 2 details the computational and parameters employed for the DFT calculations. In Section 3, we present and discuss the computational results, including the bulk and surface structures, polaron stability, transmission coefficient, and adsorption of a water molecule, with a focus on understanding the facet-dependent photocatalytic activity. Finally, Section 4 summarizes the key findings of the study.
We employ the CDFT method in this study to localize polarons. Some previous studies have used geometric coordinates as reaction coordinates to calculate energy profiles in polaron transfer by interpolating the coordinates between reactant and product states.33 However, this approach is limited to energetics and can hardly be applied to molecular dynamics. In contrast, the CDFT method can utilize energy-gap coordinates (defined in eqn (1) below),25,26 which enables sampling of intermediate states in molecular dynamics simulations. The thermal fluctuations in MD simulations can influence charge transfer processes, such as polaron transfer. For example, at the TiO2 surface, polarons can transfer among various surface sites in MD simulations at finite temperatures,34 indicating the thermal fluctuations of polaron stability. The thermal fluctuations are also crucial when discussing free energies in charge-transfer reactions. For instance, in the proton transfer of aqueous silicic acid, the calculated reaction free energy is 13.9 kcal mol−1—significantly lower than the enthalpy of reaction (29.3 kcal mol−1) in the absence of solvent-induced fluctuations.26 In addition, this capability enables the calculation of the transmission coefficient, as described below.
In the CDFT calculations, the atomic radii of the Becke density partitioning method22,35 were 1.85 Å for Sr, 1.36 Å for Ti, 0.63 Å for O, 0.32 Å for H from ref. 36. The convergence threshold of the charge constraint is 10−3 electrons. The snapshots were drawn with VESTA.37
The initial bulk geometry of cubic perovskite STO was based on the experimental cell length of 3.8996 Å.42 A supercell was constructed by expanding the unit cell threefold along a, b, and c axes. The resulting supercell consists of 27 units of SrTiO3 (135 atoms in total). Both the atomic positions and cell lengths were optimized in geometry optimizations. In the following calculations, the cell length of supercell was fixed to the optimized cell length.
Subsequently, an excess positive charge (+1) was added to the systems, and CDFT MD simulations started from the above-mentioned five snapshots and performed for 1 ps in the NVT ensemble as an equilibration run. During these simulations, the magnetization constraint of 0.48 (as detailed in Section 3.1.2) was applied to a single oxygen atom. Following equilibration, 5 ps diabatic MD simulation with the CDFT method was performed, resulting in the total simulation time of 25 ps (5 initial conditions × 5 ps each).
To calculate the transmission coefficient, we employed the energy-gap coordinate. The vertical energy-gap coordinate for the CDFT method with nuclear configuration R was by25,26
s = ER (R) − EP (R), | (1) |
We next computed the time-correlation function (TCF) of the time derivative of the energy-gap coordinate,
C(t) = 〈ṡ(0)ṡ(t)〉, | (2) |
To minimize the sampling noise arising from the limited simulation time of the MD simulations, we multiplied the Gaussian window function of exp(−t2/τ2) with τ = 0.5 ps to the TCF. The TCF was truncated to 2048 steps (equivalent to 4.096 ps). To assign the vibrational modes in the spectral density, we also performed the normal mode analysis using the optimized geometry. In this calculation, the increment to be used to construct the Hessian with finite difference method was set to 0.01 bohr.
The transmission coefficient was evaluated by50
![]() | (3) |
Because the (001) surfaces are not stoichiometric, the non-standard formula was required to calculate the surface energy. To calculate the surface energy of the (001) surfaces, the unrelaxed cleavage energy surface was calculated as follows:12
![]() | (4) |
![]() | (5) |
Finally, the surface energy was calculated by summing the unrelaxed cleavage energy and relaxation energy
Esurf(Λ) = E(unr)surf(Λ) + Erel(Λ). | (6) |
In contrast, the surface energy at the (110) surface was computed by a standard formula
![]() | (7) |
Eads = E(substrate + adsorbate) − E(substrate) − E(adsorbate), | (8) |
Here, E(substrate + adsorbate), E(substrate), and E(adsorbate) are the energies of the substrate and adsorbate, substrate and adsorbate, respectively. In this work, the substrate was the STO slab, and the adsorbate was a water molecule.
The experimentally measured indirect band gap energy is 3.25 eV, while the direct band gap energy is 3.75 eV.51 The agreement with the experimental band gap is also the best for the hybrid PBE functional (Table 1). These results indicate that the inclusion of HF exchange improves the agreement with the experimental physical properties. The smaller computed band gaps without HF exchange have been observed in the previous calculations using the local density approximation, where the indirect and the direct band gap energies of STO determined from the band structure calculations are 1.89 and 2.22 eV.51
Because the hybrid PBE functional more accurately reproduces the experimental physical properties, it was used for all subsequent calculations. In addition, the cell length of supercell was fixed to the optimized cell length in the following calculations.
Fig. 2b shows the projected density of states (PDOS) of bulk STO using the hybrid PBE functional. Most of the electronic states in the valence band originate from orbitals of oxygen, while most of the electronic states in the conduction band are derived from orbitals of titanium. No significant contribution from orbitals of strontium and titanium is observed near the Fermi energy.
Fig. 3a also shows the (yellow) isosurface of spin density of the localized hole. While primarily localized on the specified oxygen atom, the hole extends also to four neighbouring oxygen atoms, each exhibiting the smaller Hirshfeld spin density of 0.10. This indicates that, in STO, a hole tends to be slightly more delocalized compared to in anatase TiO2,7 in which the Hirshfeld spin density of a localized hole on an O atom typically exceeds 0.8. It is noted that if the hybrid PBE functional is changed to the PBE functional, a hole delocalizes as shown in Fig. S2a (ESI†), indicating the necessity of HF exchange for polaron localization. In addition, using the revPBE functional with 10.5% HF exchange localized a hole as shown in Fig. S2b (ESI†) and yields the magnetization of 0.51 on an O atom by Hirshfeld analysis. This shows that the different generalized gradient approximation functionals with the same fraction of HF exchange can give similar magnetizations for hole polarons, showing the robustness of a stable hole polaron in bulk SrTiO3 within the DFT functionals employed in this work.
The distance between the donor and acceptor oxygen atoms slightly decreases from 2.79 Å in the neutral state to 2.73 Å when the polaron is localized on the donor oxygen atom to stabilize the positive charge of a polaron. On the other hand, both Sr- and Ti–OD bond lengths respectively increase on average from 2.79 and 1.98 Å (neutral state) to 2.83 and 2.06 Å (polaron-localized state), because the positive charge of a polaron repels Sr and Ti cations. (Detailed bond lengths are provided in Fig. S1a, ESI.†) These structural adjustments promote the stabilization of the localized positive charge on the oxygen atom. Similar structural changes have also been observed in TiO2.8
The PDOS in the hole-polaron-stabilized state is shown in Fig. 4. A mid-gap state associated with a hole on the OD atom appears within the band gap, positioned slightly above the valence band maximum (VBM) at 0.69 eV. The contributions from the other atoms are close to those in Fig. 2b, indicating that the electronic structure of atoms, aside from the OD atom, is largely unaffected by localization of a polaron.
![]() | ||
Fig. 4 PDOS of bulk STO in the presence of a hole polaron as shown in Fig. 3a. The positive and negative DOS correspond to the alpha and beta spin states. The DOS projected on the donor O atom is magnified by ten times and marked in red filled area. |
Fig. 5a shows the diabatic potential energy surface (PES) calculated via the CDFT method for the hole polaron transfer from a donor oxygen atom (OD) to an acceptor oxygen atom (OA). Fig. 5b and c respectively show the optimized bulk STO and the localized polaron at the transition state (TS) and product state. To calculate the diabatic potential energy surface, we adopted the following mapping potential according to the linear mixing scheme25,26
Eα = (1 − α)ER + αEP. | (9) |
By varying the perturbation parameter α from 0 (reactant) to 1 (product) in constant increments of Δα, we can generate the diabatic states. The midpoint value, α = 0.5, corresponds to the TS of the reaction. The orbital in the adiabatic state was obtained by calculating the Kohn–Sham orbital using the optimized geometry through the CDFT method.
The calculated electronic coupling of polaron transfer in the bulk phase is 0.20 eV by the CDFT method. This value slightly exceeds the barrier height of 0.15 eV in the diabatic PES of Fig. 5a. In contrast, the barrier height in the adiabatic PES, also shown in Fig. 5a, is 0.05 eV. The barrier height is comparable to the thermal energy of kBT = 0.03 eV at 298.15 K. This low activation energy is close to the experimental activation energy of 0.06–0.10 eV in N ion implanted SrTiO352 and 0.10 eV in degraded crystal of SrTiO3.53 In the calculation of the adiabatic PES, we adopted the geometry optimized with the mapping potential and obtained the Kohn–Sham orbital after removing the constraint on the magnetization. We note that quantitative comparison with experimental results is challenging for the activation energy due to differences between experiments and calculations. For instance, SrTiO3 employed in experiments contains dopants and defects that can influence polaron transfer, while our calculations employed pristine SrTiO3. In addition, using the linear-interpolation method,33 the activation energy increases to 0.11 eV, which is higher than that of 0.05 eV in the adiabatic PES using the CDFT method. The lower activation energy by the CDFT method arises from the geometry relaxation at the TS.
The spin density at TS, shown in Fig. 5b, reveals that the hole is delocalized over the two oxygen atoms of the donor (D) and acceptor (A). The distance between the two O atoms of the donor and acceptor slightly increases from 2.73 Å (polaron on the donor O atom) to 2.77 Å (TS) due to repulsion between the polarons. Meanwhile, both Sr–OD and Ti–OD bond lengths change from 2.83 and 2.06 Å (reactant state) to 2.83 and 2.02 Å (TS) on average, respectively. These structural changes arise from the repulsion between the positively charged polaron and the Sr and Ti cations. (Individual bond lengths are provided in Fig. S1a, ESI.†)
![]() | ||
Fig. 6 Fourier transform of the time correlation function (eqn (2)) of the time derivative of the energy-gap coordinate for the hole polaron transfer in bulk STO using the hybrid PBE functional. The ordinate is plotted as a logarithmic scale. |
Normal mode analysis was conducted to assign the vibrational modes (see Fig. S3 for the illustrations, ESI†) by selecting the normal modes, the frequencies of which are the closest to the above-mentioned three peaks in Fig. 6. The lowest-frequency mode is assigned to the vibrations of the neighbouring Sr and Ti atoms, with a normal mode frequency of 205 cm−1. The middle- and high-frequency modes are primarily associated with the O motions in the two neighbouring TiO6 clusters and one TiO6 cluster, with the normal mode frequencies of 420 and 697 cm−1, respectively. These findings suggest that the collective vibrational modes of oxygen, which modulates the neighbouring OO distance and O–Ti–O angles, are strongly coupled to the polaron transfer dynamics.
We then optimized the geometry of bulk STO with one oxygen vacancy and one excess electron, because an O vacancy can induce an electron localization as observed in TiO2.55Fig. 7b shows the spin density of localized electrons. In contrast to the hole localized on O atoms, the electron polaron in part localized to the O vacancy site yet the other spin densities uniformly delocalize over the surrounding Ti atoms next to the oxygen vacancy. The magnetization of Ti atoms next to the oxygen vacancy ranges from −0.269 (down spin) to 0.241 (up spin). Fig. 7d shows the calculated DOS. The electron trapped state lies around 2.7 eV. Since an electron polaron does not localize in the same manner as a hole polaron and the oxygen vacancy seems to cause the localization, the stability of oxygen defects is investigated in the following surface calculations. Fig. 7e and f show the TiO5 clusters next to O vacancy in the neutral state and the electron-polaron-localized state, respectively. Upon electron localization, the Ti–O bond lengths adjacent to O vacancy increase by 0.01–0.09 Å due to charge repulsion between the electron polaron and oxygen anions, while the Hirshfeld charge of Ti atoms next to O vacancy decreases by 0.08–0.12. This electron localization near O vacancy is facilitated by the high local structural flexibility of the lattice around O vacancy, as well as a more adaptable charge environment, which are more difficult in pristine SrTiO3 without O vacancy as mentioned above.
The TiO2-terminated (001) surface and the O-terminated (110) surface also exhibit the similar changes of bond lengths. At the TiO2-terminated (001) surface in Fig. 8b, d[Sr1–O1] = 2.65 Å is also smaller than that in bulk (2.79 Å). In contrast, d[Ti1–O2] is 1.83 Å, which is smaller than that in the bulk value of 1.98 Å.
In contrast to the (001) surfaces, the (110) surface has asymmetry due to the presence of a surface-top O atom. The (110) surface has two types of surface O atoms (Fig. 8c): O1, which protrudes from the surface SrTiO layer, and O2, which is in the surface SrTiO layer. The Sr1–O1 bond length of 2.56 Å is much smaller than that in the bulk value of 2.79 Å. The (110) surface exhibits the lowest coordination number of three (two by Sr atoms and one by a Ti atom) for the surface O atoms among the three surfaces examined in this work. This is in contrast to the SrO-terminated (001) surface, where the coordination number is five (four by Sr atoms and one by a Ti atom), and the TiO2-terminated (001) surface, where it is four (two by Sr atoms and two by Ti atoms).
The calculated surface energies of the optimized SrO- and TiO2-(001) and (110) surfaces using the hybrid PBE functional are 1.22, 1.23, and 1.25 eV per surface unit cell. Our computational results are consistent with trends observed in the previous study using the B3PW functional,12 where the order of the surface energy is reproduced as 1.15 eV (SrO-terminated (001) surface) <1.23 eV (TiO2-terminated (001) surface) <1.54 eV (O-terminated (110) surface) though the differences among the facets are larger than our results, probably due to the different computational conditions. The small difference in surface energies may not be experimentally resolvable. However, both (001) and (110) surfaces have been observed in experiments,1 indicating that these surfaces can coexist in STO. In real photocatalytic processes, which occur in aqueous environments, the stability of different facets may vary. While investigating this aspect is important, it lies beyond the scope of this study. Consequently, as all facets are plausible, this study considers all three surfaces in the subsequent discussions.
Surface | Molecular | Dissociative |
---|---|---|
SrO-(001) | −0.89 | −1.44 |
TiO2-(001) | −0.77 | −0.98 |
(110) | −0.70 | −1.71 |
![]() | ||
Fig. 9 Molecular and dissociative adsorptions of a water molecule adsorbed on the (a) SrO- and (b) TiO2-terminated STO (001) and (c) (110) surfaces using the hybrid PBE functional. The enlarged views for the region enclosed by dashed lines are also shown, including the bond lengths in Å. Both molecular and dissociative adsorptions are shown. In panel (a), the definition of the depth coordinate for Fig. 10–12 is also displayed. Colour code: green: Sr, red: O, light blue: Ti, and white: H. |
At the SrO-terminated (001) surface, a molecular adsorption is more stable than on the other surfaces, with an Sr–O bond length is 2.72 Å. The H atoms in the adsorbed water molecule point toward the surface normal direction. For dissociative adsorption, a water molecule splits to an H atom, which bonds to the surface O atom, and an OH group, which bridges the neighbouring surface Sr atoms with the Sr–O bond lengths of 2.55 and 2.62 Å. The computed adsorption energy for molecular adsorption closely matches the previous calculations ranging from −0.67 to −1.1 eV, while that for dissociative adsorption is smaller than the previous calculations ranging from −0.78 to −1.1 eV by various DFT functionals.13
For molecular adsorption at the TiO2-terminated (001) surface, the O atom of a water molecule adsorbs to the surface Ti atom. Unlike at the SrO-terminated (001) surface, the H atoms in an adsorbed water molecule points toward the surface parallel direction. For dissociative adsorption, a water molecule splits to an OH group on the surface Ti atom and H atom on the neighbouring surface O atom. The computed adsorption energies in our DFT calculations are similar to the previous DFT calculations, ranging from −0.64 to −1.0 eV for molecular adsorption and ranging from −0.61 to −0.87 eV for dissociative adsorption using various DFT functionals.13
At the (110) surface, molecular adsorption can also be found on Sr atom. The adsorption energy for molecular adsorption is much larger than the dissociative case, and a water molecule strongly prefers to split on the surface. The Sr–O bond is not completely along the surface-normal direction due to the presence of the surface-top O atom. For dissociative adsorption, the dissociated proton bonds to the surface O atom, while the OH group bridges two Sr and Ti atoms. The stronger adsorption for dissociative adsorption was also reported previously.14
![]() | ||
Fig. 10 (a) Potential energy surface of hole polaron transfer and (b)–(d) polaron holes localized at the surface O atoms at the SrO- and TiO2-terminated STO (001) and (110) surfaces using the hybrid PBE functional. The energy at the bulk site that is the most distant from the surface was set to zero. The definition of the depth coordinate is displayed in Fig. 9a. The yellow lobes were displayed with the isosurface levels of (b) 0.0030, (c) 0.0033, and (d) 0.0031. |
As shown in Fig. 10b, when a hole is at the outermost O atom of the SrO-terminated STO (001) surface, the Ti–O bond length slightly decreases by 8% as the Ti atom moves toward the surface, while Sr–O bond length remains almost the same. For the TiO2-terminated STO (001) surface as shown in Fig. 10c, Ti–O bond length increases by 5%, and the protrusion of O atom along the surface normal direction is pronounced. When a hole is localized on the surface O atom at the STO (110) surface, Ti- and Sr–O bond lengths respectively increase from 1.74 and 2.56 Å to 1.84 (6% increase) and 2.73 Å (7% increase), as shown in Fig. 10d. Such structural changes in terms of bond lengths upon localization of a hole were also observed in bulk discussed in Section 3.1.2. The most stable polaron at the STO (110) surface may be related with the flexibility of surface O atom. In fact, the surface O atom of the STO (110) surface has the lowest coordination number of three as discussed in Section 3.2.1, and therefore is thought to be the most flexible among the three surfaces considered in this study.
To investigate the impact of hydroxylation, we computed the potential energy profile as a function of polaron position at the STO surfaces on which one water molecule split into OH and H (see the Section 3.2.2 for the adsorption structures). The energy profile in Fig. 11 remains largely unchanged from that at the pristine surfaces, except at the surface OH site. At the TiO2-terminated (001) surface, however, the energy at the surface oxygen site decreases compared to the pristine surface. However, the polaron transfer from the surface O site to the OH group is energetically unfavourable with an increase of 0.13 eV. At the SrO-terminated (001) surface, a polaron on the OH group is the most stable among the OH groups at the three surfaces studied in this work. Nevertheless, the polaron transfer should be highly unlikely due to the instability of polaron at the surface O site. In the case of the (110) surface, the energy rises to 0.14 eV at the OH site. This suggests that the hole transfer proceeds through the surface oxygen site rather than an OH group during oxidation reactions. This contrasts with the case of anatase TiO2,7 where the presence of OH group stabilizes a hole polaron.
![]() | ||
Fig. 11 Same as Fig. 10a, yet OH group is attached the surfaces from water adsorption using the hybrid PBE functional. |
We further computed the potential energy surface for various positions of oxygen vacancies at the STO surfaces, as shown in Fig. 12. At all surfaces, the energy drops at the surface O sites. This would be because the structural changes at surface are easier than in bulk. The TiO2-terminated (001) surface exhibits the more pronounced energy drop than the other surfaces. This indicates that O vacancy can be formed more readily at the TiO2-terminated (001) surface. A higher concentration of O vacancies at the TiO2-terminated (001) surface implies more trapping sites for electron polaron, consistent with the experimental findings that an electron polaron transfers to the (001) surface to induce the HER reactions.1
Our investigation into the bulk STO structure revealed that the hybrid PBE functional closely more accurately reproduces experimental physical properties of STO, particularly in terms of cell length and band gap energy, making it a reliable choice for subsequent calculations. The simulation of hole polaron transfer demonstrated that the process in bulk STO has a low activation barrier, with a calculated barrier height close to the thermal energy at room temperature. In bulk STO, a hole polaron localizes on O atoms, while an electron polaron can localize with the existence of an oxygen vacancy. The dynamical effects of polaron transfer were also examined through by computing the transmission coefficient, and the hole polaron transfer occurs through an adiabatic charge transfer mechanism.
When exploring the surface properties, we considered both SrO and TiO2-terminated (001) surfaces and the O-terminated (110) surface, which are known for their stability. The calculations of surface energies and adsorption energies provided a quantitative understanding of the surface stability and the interaction with a water molecule and the STO surface. The dissociative adsorption of a water molecule is energetically favoured, and the computed adsorption energies are consistent with the previous DFT studies. At the STO surfaces, a hole is localized on the surface O atom of the (110) surface most stably, and an OH group the surface does not enhance the stability. In contrast, the oxygen vacancy is most stable on the TiO2-terminated (001) surface. The insights gained from this study can inform future efforts in developing advanced STO photocatalysts for sustainable energy applications.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4cp04725g |
This journal is © the Owner Societies 2025 |