Next Article in Journal
The Study of C3H6 Impact on Selective Catalytic Reduction by Ammonia (NH3-SCR) Performance over Cu-SAPO-34 Catalysts
Previous Article in Journal
Crystal Imperfections of Industrial Vanadium Phosphorous Oxide Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Water Splitting on Multifaceted SrTiO3 Nanocrystals: Computational Study

1
Institute of Solid State Physics, University of Latvia, Kengaraga 8, LV1063 Riga, Latvia
2
Institute of Physics, University of Tartu, W.Ostwaldi 1, 50411 Tartu, Estonia
*
Authors to whom correspondence should be addressed.
Catalysts 2021, 11(11), 1326; https://doi.org/10.3390/catal11111326
Submission received: 12 October 2021 / Revised: 28 October 2021 / Accepted: 30 October 2021 / Published: 31 October 2021
(This article belongs to the Topic Electromaterials for Environment & Energy)

Abstract

:
Recent experimental findings suggest that strontium titanate SrTiO3 (STO) photocatalytic activity for water splitting could be improved by creating multifaceted nanoparticles. To understand the underlying mechanisms and energetics, the model for faceted nanoparticles was created. The multifaceted nanoparticles’ surface is considered by us as a combination of flat and “stepped” facets. Ab initio calculations of the adsorption of water and oxygen evolution reaction (OER) intermediates were performed. Our findings suggest that the “slope” part of the step showed a natural similarity to the flat surface, whereas the “ridge” part exhibited significantly different adsorption configurations. On the “slope” region, both molecular and dissociative adsorption modes were possible, whereas on the “ridge”, only dissociative adsorption was observed. Water adsorption energies on the “ridge” ( 1.50 eV) were significantly higher than on the “slope” ( 0.76 eV molecular; 0.83 eV dissociative) or flat surface ( 0.79 eV molecular; 1.09 eV dissociative).

1. Introduction

Strontium titanate SrTiO3 (STO) is a well-known material for water splitting [1,2,3,4,5,6,7,8,9,10,11,12]. The process of water adsorption and dissociation was studied in detail [13,14,15]. The effects of doping are investigated in [16]. Recent developments in nanocrystal synthesis offered materials with enhanced charge separation achieved by heterojunction [17,18], mesocrystallinity [19], or the exposed anisotropic facets [20,21]. Nanoparticles synthesized by Takata et al. [20] were made from STO doped by aluminium and photodeposited cocatalysts Rh/Cr2O3 and CoOOH, and demonstrated a quantum efficiency of up to 96% in the range of 350 to 360 nm. Synthesized six and eighteen-facet STO nanocrystals, as described in [21], demonstrated high catalytic activity in water splitting. When doped by Pt and Co3O4 on particular facets, these nanoparticles exhibited even higher performance. Such an improvement is attributed to the unique properties of anisotropic facets of the particles.
One of the key properties of a high-performance water splitting material is a low charge recombination rate. Adsorption of water and oxygen evolution reaction (OER) intermediates on stepped surfaces is expected to be qualitatively different than that on flat surfaces. Featuring surfaces of different orientations, the 18-facet nanoparticle provides a natural platform for efficient charge separation. The six-facet nanoparticle is essentially a cube with the {0 0 1} faces. Its edges, however, can be considered as a different reaction area from the {0 0 1} flat parts. Currently, to the best of our knowledge, the structure of the surface of these nanoparticles is described only at the nanoscale. To reveal the properties of different reaction areas of multi-faceted nanoparticles, an atomistic model has to be designed and tested.
In 18-facet STO nanoparticle, the {0 0 1} facets are combined with the facets parallel to the {1 1 0} crystallographic plane. Although for the real material, the surfaces {0 0 1} and {1 0 0} are not equivalent [22], for the present study, the distortion of the perovskite structure is irrelevant and the orientation of the surfaces are given relative to the cubic phase. As it was shown in [23], the ideal polar {1 1 0} surface is unstable. Its stabilization can be achieved by forming steps of the more stable {0 0 1} orientation [24].
In the present study, we propose an atomistic model of the {0 0 1} stepped surface, which is relevant to both six as well as eighteen-facet STO nanoparticles. On this surface, we simulate OER, as suggested by Nørskov [25]. The four-step reaction includes adsorption of H2O, HO*, O*, and HOO* species. The results will be used to perform thermodynamic simulations of the OER to obtain over-potential η OER values.
An extensive investigation of OER on the flat STO surface was performed by Cui et al. [14]. For the flat {0 0 1}, the STO surface of the over-potential value of 0.66 V was obtained.

2. Methods and Computational Details

We performed density functional theory (DFT) calculations using the Vienna Ab initio Simulation Package (VASP) [26,27,28,29]. Computational details are listed in Table 1. Relaxed rhombohedral SrTiO3 phase ( R 3 ¯ c ) with the optimized lattice constant a 0 of 3.92 Å was used. We compared flat surface and stepped surface, as shown in Figure 1 and Figure 2, respectively. Adsorbate was placed on both terminations of the slab to neutralize the electric dipole moment.
Adsorption energy E ads , x for the configuration x was calculated by Equation (1), where E x is the total energy of the configuration with the adsorbate, E surf is the total energy of the corresponding surface without the adsorbate, E H 2 O is the total energy of water, E H 2 is the total energy of molecular hydrogen, and coefficients c 1 and c 2 are determined so that the total number of particles on the left hand side of the equation is zero. Factor 1 2 is a result of the adsorbate being placed on both terminations of the slab. Lower adsorption energy corresponds to the stronger binding.
All figures were created in the VESTA visualization system [30].
E ads , x = 1 2 ( E x ( E surf + c 1 E H 2 O + c 2 E H 2 ) )

3. Results and Discussion

3.1. Water Adsorption

It is important to understand how water adsorption on stepped surfaces distinguishes from that on flat surfaces. On flat surfaces, the most preferable water adsorption site is atop titanium. The stepped model of faceted surfaces features three regions: the ridge, slope, and gully, marked on Figure 2.
On flat surfaces, two adsorption modes are possible: molecular (Figure 3a; E ads = 0.79   eV ) and dissociative (Figure 3b; E ads = 1.09   eV ), with dissociative being more energetically favorable, which is in agreement with Reference [13]. In [13], it was demonstrated that there is no significant transition barrier ( 0.09 eV) between the two adsorption modes. On the stepped surface, the situation is more complex and each adsorption region should be discussed separately.
On the Slope region, there are several possible adsorption configurations. The most energetically favorable one is dissociative adsorption along Slope (Figure 4d; E ads = 0.83   eV ), followed closely by molecular adsorption (Figure 4b,c with E ads of 0.76   eV and 0.71   eV , respectively). Molecular adsorption energy on the slope was loosely dependent on the orientation of water molecules. Dissociation towards the gully or ridge was less favorable (Figure 4e,f with E ads of 0.62   eV and 0.41   eV , respectively). Although, similarly to flat surfaces, one of the possible dissociative configurations on the slope was more favorable than molecular ones, the difference between energies was not as large, thus it cannot be unequivocally concluded, wherein the adsorption mode dominates.
On the Ridge region, only one configuration was observed: water dissociation accompanied by spontaneous oxygen vacancy formation (Figure 4a). It also had the lowest adsorption energy (strongest adsorption) of 1.50 eV among all the tested configurations. The investigation of the question regarding whether the ridge breaks down irreversibly or whether the vacancy is healed during subsequent water adsorption is out of scope of this paper.
On the Gully region, only one dissociative configuration was found, but only hydrogen was adsorbed, while oxygen was on the slope. Moreover, this configuration had a relatively weak binding (Figure 4g, E ads = 0.49   eV ), hence we did not investigate it further.

3.2. Oxygen Evolution Reaction (OER) Intermediates

To perform thermodynamic simulations to estimate STO photo-catalytic activity, it is necessary to compute adsorption energies for oxygen evolution reaction (OER) intermediates: HO*, O*, and HOO*, where the star * denotes the active adsorption site. All energies are compiled in Table 2.
Results for the flat surface are shown in Figure 5. HO* had only one possible configuration, as shown in Figure 5a. There were two possible configurations for O*: where oxygen from the adsorbate bonded to surface oxygen, denoted as O*(Osurf) (Figure 5b), and where oxygen was atop titanium, denoted as O* (Figure 5c). The O*(Osurf) adsorption energy was much lower ( 2.87 eV versus 3.52 eV), thus it was more energetically favorable. The HOO* adsorbate also had two configurations: one where hydrogen bonded to surface oxygen (Figure 5d), denoted as H(Osurf)OO*, and the other where hydrogen bonded to the adsorbate’s oxygen (Figure 5e), denoted as HOO*. In [14], only the second configuration was mentioned, although its adsorption energy was significantly higher than that of the H(Osurf)OO* configuration: 4.24 eV versus 3.64 eV.
The adsorption of intermediates on the slope (Figure 6) was similar to that of the flat surface: only one HO* configuration (Figure 6a), O*(Osurf) (Figure 6b), and O* (Figure 6c), and two HOO* configurations, namely H(Osurf)OO* (Figure 6d) and HOO* (Figure 6e) were observed. O*(Osurf) was more energetically favorable than O* ( 2.93 eV versus 4.11 eV) and H(Osurf)OO* was more favorable than HOO* ( 3.48 eV versus 4.52 eV), analogous to the flat surface. HO* on the slope had higher adsorption energy than HO* on the flat surface ( 1.35 eV versus 0.77 eV).
Results for the intermediates on the ridge region are shown in Figure 7. For each intermediate, only one adsorption configuration was observed. HO* (Figure 7a) and HOO* (Figure 7c), similarly to the water adsorption, were accompanied by spontaneous oxygen vacancy formation, while O* bonded between surface oxygen and titanium (Figure 7b).

4. Conclusions

We have performed a detailed investigation of water adsorption and oxygen evolution reaction (OER) intermediate adsorption on strontium titanate SrTiO3 flat and stepped surfaces. In contrast to the flat surface, the stepped surface, a significant part of which comprises the ridge region, demonstrated high adsorption energies as well as pronounced structural transformations caused by the adsorbate. Our findings suggest that:
  • The ridge region permits dissociative water adsorption only, accompanied by spontaneous formation of oxygen vacancy;
  • Results for the flat surface are in agreement with other computational studies [13];
  • On the slope region, both molecular and dissociative adsorption modes are possible;
  • Adsorption of both water and its intermediates on the slope region is similar to that on flat surfaces;
  • Except for atomic hydrogen, no adsorption was observed on the gully region; and
  • There are different adsorption configurations of OER intermediates possible on flat surfaces and slope regions.

Author Contributions

Conceptualization, Y.A.M. and E.A.K.; data curation, M.S. and V.K.; formal analysis, M.S.; funding acquisition, D.B. and E.A.K.; investigation, M.S., Y.A.M., G.Z. and V.K.; methodology, Y.A.M., D.B. and E.A.K.; project administration, D.B. and E.A.K.; resources, Y.A.M. and E.A.K.; software, M.S.; supervision, Y.A.M., D.B. and E.A.K.; validation, M.S., Y.A.M., G.Z. and V.K.; visualization, M.S.; writing—original draft, M.S.; writing—review and editing, Y.A.M., D.B. and E.A.K. All authors have read and agreed to the published version of the manuscript.

Funding

The financial support of M-ERA.NET2 Sun2Chem project is greatly acknowledged by E.K. Authors thank Marjeta Maček Kržmanc and Chi-Sheng Wu, for the fruitful discussions. The financial support of FLAG-ERA JTC project To2Dox is acknowledged by Y.A.M. This paper is based upon the work from COST Action 18234, supported by COST (European Cooperation in Science and Technology). The support is greatly acknowledged by Y.A.M. and V.K. The grant No. 1.1.1.2/VIAA/l/16/147 (1.1.1.2/16/I/001) under the activity of Post-doctoral research aid is greatly acknowledged by M.S. and D.B.

Acknowledgments

The Institute of Solid State Physics, University of Latvia (Latvia) as the Centre of Excellence has received funding from the European Union’s Horizon 2020 Frame-work Programme H2020-WIDESPREAD-01-2016-2017-Teaming Phase2 under grant agreement No. 739508, project CAMART2. The computer resources were provided by the Stuttgart Supercomputing Center (project DEFTD 12939) and Latvian Super Cluster (LASC).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kato, H.; Kudo, A. Visible-Light-Response and Photocatalytic Activities of TiO2 and SrTiO3 Photocatalysts Codoped with Antimony and Chromium. J. Phys. Chem. B 2002, 106, 5029–5034. [Google Scholar] [CrossRef]
  2. Miyauchi, M.; Nakajima, A.; Watanabe, T.; Hashimoto, K. Photocatalysis and Photoinduced Hydrophilicity of Various Metal Oxide Thin Films. Chem. Mater. 2002, 14, 2812–2816. [Google Scholar] [CrossRef]
  3. Konta, R.; Ishii, T.; Kato, H.; Kudo, A. Photocatalytic Activities of Noble Metal Ion Doped SrTiO3 under Visible Light Irradiation. J. Phys. Chem. B 2004, 108, 8992–8995. [Google Scholar] [CrossRef]
  4. Yu, S.C.; Huang, C.W.; Liao, C.H.; Wu, J.C.; Chang, S.T.; Chen, K.H. A novel membrane reactor for separating hydrogen and oxygen in photocatalytic water splitting. J. Membr. Sci. 2011, 382, 291–299. [Google Scholar] [CrossRef]
  5. Kato, H.; Kobayashi, M.; Hara, M.; Kakihana, M. Fabrication of SrTiO3 exposing characteristic facets using molten salt flux and improvement of photocatalytic activity for water splitting. Catal. Sci. Technol. 2013, 3, 1733. [Google Scholar] [CrossRef]
  6. Wang, B.; Shen, S.; Guo, L. SrTiO3 single crystals enclosed with high-indexed 023 facets and 001 facets for photocatalytic hydrogen and oxygen evolution. Appl. Catal. B Environ. 2015, 166–167, 320–326. [Google Scholar] [CrossRef]
  7. Ham, Y.; Hisatomi, T.; Goto, Y.; Moriya, Y.; Sakata, Y.; Yamakata, A.; Kubota, J.; Domen, K. Flux-mediated doping of SrTiO3 photocatalysts for efficient overall water splitting. J. Mater. Chem. A 2016, 4, 3027–3033. [Google Scholar] [CrossRef] [Green Version]
  8. Foo, G.S.; Hood, Z.D.; Wu, Z. Shape Effect Undermined by Surface Reconstruction: Ethanol Dehydrogenation over Shape-Controlled SrTiO3 Nanocrystals. ACS Catal. 2018, 8, 555–565. [Google Scholar] [CrossRef]
  9. Li, D.; Yu, J.C.C.; Nguyen, V.H.; Wu, J.C.; Wang, X. A dual-function photocatalytic system for simultaneous separating hydrogen from water splitting and photocatalytic degradation of phenol in a twin-reactor. Appl. Catal. B Environ. 2018, 239, 268–279. [Google Scholar] [CrossRef]
  10. Kampouri, S.; Stylianou, K.C. Dual-Functional Photocatalysis for Simultaneous Hydrogen Production and Oxidation of Organic Substances. ACS Catal. 2019, 9, 4247–4270. [Google Scholar] [CrossRef]
  11. Kanazawa, T.; Nozawa, S.; Lu, D.; Maeda, K. Structure and Photocatalytic Activity of PdCrOx Cocatalyst on SrTiO3 for Overall Water Splitting. Catalysts 2019, 9, 59. [Google Scholar] [CrossRef] [Green Version]
  12. Saleem, Z.; Pervaiz, E.; Yousaf, M.U.; Niazi, M.B.K. Two-dimensional materials and composites as potential water splitting photocatalysts: A review. Catalysts 2020, 10, 464. [Google Scholar] [CrossRef]
  13. Guhl, H.; Miller, W.; Reuter, K. Water adsorption and dissociation on SrTiO3 (001) revisited: A density functional theory study. Phys. Rev. B Condens. Matter Mater. Phys. 2010, 81, 155455. [Google Scholar] [CrossRef] [Green Version]
  14. Cui, M.; Liu, T.; Li, Q.; Yang, J.; Jia, Y. Oxygen Evolution Reaction (OER) on Clean and Oxygen Deficient Low-Index SrTiO3 Surfaces: A Theoretical Systematic Study. ACS Sustain. Chem. Eng. 2019, 7, 15346–15353. [Google Scholar] [CrossRef]
  15. Holmström, E.; Spijker, P.; Foster, A.S. The interface of SrTiO3 and H2O from density functional theory molecular dynamics. Proc. R. Soc. A Math. Phys. Eng. Sci. 2016, 472. [Google Scholar] [CrossRef] [Green Version]
  16. Chen, H.C.; Huang, C.W.; Wu, J.C.S.; Lin, S.T. Theoretical Investigation of the Metal-Doped SrTiO3 Photocatalysts for Water Splitting. J. Phys. Chem. C 2012, 116, 7897–7903. [Google Scholar] [CrossRef]
  17. Afroz, K.; Moniruddin, M.; Bakranov, N.; Kudaibergenov, S.; Nuraje, N. A heterojunction strategy to improve the visible light sensitive water splitting performance of photocatalytic materials. J. Mater. Chem. A 2018, 6, 21696–21718. [Google Scholar] [CrossRef]
  18. Maček Kržmanc, M.; Daneu, N.; Čontala, A.; Santra, S.; Kamal, K.M.; Likozar, B.; Spreitzer, M. SrTiO3/Bi4Ti3O12Nanohetero structural Platelets Synthesized by Topotactic Epitaxy as Effective Noble-Metal-Free Photocatalysts for pH-Neutral Hydrogen Evolution. ACS Appl. Mater. Interfaces 2021, 13, 370–381. [Google Scholar] [CrossRef] [PubMed]
  19. Li, X.; Yu, J.; Jaroniec, M. Hierarchical photocatalysts. Chem. Soc. Rev. 2016, 45, 2603–2636. [Google Scholar] [CrossRef]
  20. Takata, T.; Jiang, J.; Sakata, Y.; Nakabayashi, M.; Shibata, N.; Nandal, V.; Seki, K.; Hisatomi, T.; Domen, K. Photocatalytic water splitting with a quantum efficiency of almost unity. Nature 2020, 581, 411–414. [Google Scholar] [CrossRef]
  21. Mu, L.; Zhao, Y.; Li, A.; Wang, S.; Wang, Z.; Yang, J.; Wang, Y.; Liu, T.; Chen, R.; Zhu, J.; et al. Enhancing charge separation on high symmetry SrTiO3 exposed with anisotropic facets for photocatalytic water splitting. Energy Environ. Sci. 2016, 9, 2463–2469. [Google Scholar] [CrossRef]
  22. Tomio, T.; Miki, H.; Tabata, H.; Kawai, T.; Kawai, S. Control of electrical conductivity in laser deposited SrTiO3 thin films with Nb doping. J. Appl. Phys. 1994, 76, 5886–5890. [Google Scholar] [CrossRef]
  23. Noguera, C. Polar oxide surfaces. J. Phys. Condens. Matter 2000, 12, R367–R410. [Google Scholar] [CrossRef]
  24. Bachelet, R.; Valle, F.; Infante, I.C.; Sánchez, F.; Fontcuberta, J. Step formation, faceting, and bunching in atomically flat SrTiO3 (110) surfaces. Appl. Phys. Lett. 2007, 91, 251904. [Google Scholar] [CrossRef]
  25. Man, I.C.; Su, H.Y.; Calle-Vallejo, F.; Hansen, H.A.; Martínez, J.I.; Inoglu, N.G.; Kitchin, J.; Jaramillo, T.F.; Nørskov, J.K.; Rossmeisl, J. Universality in Oxygen Evolution Electrocatalysis on Oxide Surfaces. ChemCatChem 2011, 3, 1159–1165. [Google Scholar] [CrossRef]
  26. Kresse, G.; Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B 1993, 47, 558–561. [Google Scholar] [CrossRef]
  27. Kresse, G.; Hafner, J. Ab initio molecular-dynamics simulation of the liquid-metal–amorphous-semiconductor transition in germanium. Phys. Rev. B 1994, 49, 14251–14269. [Google Scholar] [CrossRef] [PubMed]
  28. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  29. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  30. Momma, K.; Izumi, F. VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data. J. Appl. Crystallogr. 2011, 44, 1272–1276. [Google Scholar] [CrossRef]
  31. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Vanderbilt, D. Soft self-consistent pseudopotentials in a generalized eigenvalue formalism. Phys. Rev. B 1990, 41, 7892–7895. [Google Scholar] [CrossRef]
  33. Kresse, G.; Hafner, J. Norm-conserving and ultrasoft pseudopotentials for first-row and transition elements. J. Phys. Condens. Matter 1994, 6, 8245–8257. [Google Scholar] [CrossRef]
  34. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef] [Green Version]
  35. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  36. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
Figure 1. Flat surface cell. (a) Front view; (b) Isometric view.
Figure 1. Flat surface cell. (a) Front view; (b) Isometric view.
Catalysts 11 01326 g001
Figure 2. Stepped surface cell. The one-ridge adsorption area; two-slope adsorption area; and three-gully adsorption area. (a) Front view; (b) Isometric view.
Figure 2. Stepped surface cell. The one-ridge adsorption area; two-slope adsorption area; and three-gully adsorption area. (a) Front view; (b) Isometric view.
Catalysts 11 01326 g002
Figure 3. Water adsorption configuration on the stepped surface. (a) Molecular adsorption on the flat surface: E ads = 0.79 eV . (b) Dissociative adsorption on the flat surface: E ads = 1.09 eV .
Figure 3. Water adsorption configuration on the stepped surface. (a) Molecular adsorption on the flat surface: E ads = 0.79 eV . (b) Dissociative adsorption on the flat surface: E ads = 1.09 eV .
Catalysts 11 01326 g003
Figure 4. Water adsorption configurations on the stepped surface. (a) Adsorption on the ridge region. E ads = 1.50 eV . (b) Molecular adsorption on the slope region with water oriented towards the gully. E ads = 0.76 eV . (c) Molecular adsorption on the slope region with water oriented towards the ridge. E ads = 0.71 eV . (d) Dissociative adsorption on the slope region with hydrogen migrated along the slope. E ads = 0.83 eV . (e) Dissociative adsorption on the slope region with hydrogen migrated towards the gully. E ads = 0.62 eV . (f) Dissociative adsorption on the slope region with hydrogen migrated towards the ridge. E ads = 0.41 eV . (g) Dissociative adsorption on the gully region. E ads = 0.49 eV .
Figure 4. Water adsorption configurations on the stepped surface. (a) Adsorption on the ridge region. E ads = 1.50 eV . (b) Molecular adsorption on the slope region with water oriented towards the gully. E ads = 0.76 eV . (c) Molecular adsorption on the slope region with water oriented towards the ridge. E ads = 0.71 eV . (d) Dissociative adsorption on the slope region with hydrogen migrated along the slope. E ads = 0.83 eV . (e) Dissociative adsorption on the slope region with hydrogen migrated towards the gully. E ads = 0.62 eV . (f) Dissociative adsorption on the slope region with hydrogen migrated towards the ridge. E ads = 0.41 eV . (g) Dissociative adsorption on the gully region. E ads = 0.49 eV .
Catalysts 11 01326 g004
Figure 5. OER intermediates on the flat surface. (a) HO* adsorbate on the flat surface: E ads = 0.77 eV . (b) O*(Osurf) adsorbate on the flat surface: E ads = 2.87 eV . (c) O* adsorbate on the flat surface: E ads = 3.52 eV . (d) H(Osurf)OO* adsorbate on the flat surface: E ads = 3.64 eV . (e) HOO* adsorbate on the flat surface, standard configuration: E ads = 4.24 eV .
Figure 5. OER intermediates on the flat surface. (a) HO* adsorbate on the flat surface: E ads = 0.77 eV . (b) O*(Osurf) adsorbate on the flat surface: E ads = 2.87 eV . (c) O* adsorbate on the flat surface: E ads = 3.52 eV . (d) H(Osurf)OO* adsorbate on the flat surface: E ads = 3.64 eV . (e) HOO* adsorbate on the flat surface, standard configuration: E ads = 4.24 eV .
Catalysts 11 01326 g005
Figure 6. OER intermediates on the slope of the stepped surface. (a) HO* adsorbate on the slope region: E ads = 1.35 eV . (b) O*(Osurf) adsorbate on the slope region: E ads = 2.93 eV . (c) O* adsorbate on the slope region: E ads = 4.11 eV . (d) H(Osurf)OO* adsorbate on the slope region: E ads = 3.48 eV . (e) HOO* adsorbate on the slope region: E ads = 4.52 eV .
Figure 6. OER intermediates on the slope of the stepped surface. (a) HO* adsorbate on the slope region: E ads = 1.35 eV . (b) O*(Osurf) adsorbate on the slope region: E ads = 2.93 eV . (c) O* adsorbate on the slope region: E ads = 4.11 eV . (d) H(Osurf)OO* adsorbate on the slope region: E ads = 3.48 eV . (e) HOO* adsorbate on the slope region: E ads = 4.52 eV .
Catalysts 11 01326 g006
Figure 7. OER intermediates on the ridge of the stepped surface. (a) HO* adsorbate on the ridge region: E ads = 1.24 eV . (b) O* adsorbate on the ridge region: E ads = 2.40 eV . (c) HOO* adsorbate on the ridge region: E ads = 3.13 eV .
Figure 7. OER intermediates on the ridge of the stepped surface. (a) HO* adsorbate on the ridge region: E ads = 1.24 eV . (b) O* adsorbate on the ridge region: E ads = 2.40 eV . (c) HOO* adsorbate on the ridge region: E ads = 3.13 eV .
Catalysts 11 01326 g007
Table 1. Computational details.
Table 1. Computational details.
SoftwareVASP 6 [27,28,29]
Exchange-correlation functionalGGA-PBE [31]
PseudopotentialsUltra Soft [32,33] potentials using the
Projector Augmented Wave (PAW) method [34,35]
SmearingGaussian smearing
Ti-valence configuration 3 p 6 3 d 2 4 s 2 , valence 10, energy cutoff 222 eV,
generated 07.09.2000
Sr-valence configuration 4 s 2 4 p 6 5 s 2 , valence 10, energy cutoff 229 eV,
generated 07.09.2000
O-valence configuration 2 s 2 2 p 4 , valence 6, energy cutoff 400 eV,
generated 08.04.2002
H-valence configuration 1 s 1 , valence 1, energy cutoff 250 eV,
generated 15.06.2001
Spin polarizationNon-spin polarized calculation
Plane wave basis set cut-off520 eV
Flat surface geometry (Figure 1) 2 a 0 × 4 a 0 surface cell, seven layers-thick, 20 Å vacuum
gap, 144 atoms
Stepped surface geometry (Figure 2) 2 ( 2 a 0 × 2 a 0 ) , 2 2 a 0 thickness, 10 Å vacuum gap,
104 atoms
Flat surface k-point mesh 4 × 2 × 2 Monkhorst-Pack [36]
Stepped surface k-point mesh 4 × 4 × 2 Monkhorst-Pack [36]
Table 2. Adsorption energies of oxygen evolution reaction (OER) intermediates on different types of surfaces.
Table 2. Adsorption energies of oxygen evolution reaction (OER) intermediates on different types of surfaces.
Surface TypeHO*, eVO*, eVHOO*, eV
Flat surface0.77 (Figure 5a)2.87 (Figure 5b)/ 3.52 (Figure 5c)3.64 (Figure 5d)/ 4.24 (Figure 5e)
Slope1.35 (Figure 6a)2.93 (Figure 6b)/ 4.11 (Figure 6c)3.48 (Figure 6d)/ 4.52 (Figure 6e)
Ridge1.24 (Figure 7a)2.40 (Figure 7b)3.13 (Figure 7c)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sokolov, M.; Mastrikov, Y.A.; Zvejnieks, G.; Bocharov, D.; Kotomin, E.A.; Krasnenko, V. Water Splitting on Multifaceted SrTiO3 Nanocrystals: Computational Study. Catalysts 2021, 11, 1326. https://doi.org/10.3390/catal11111326

AMA Style

Sokolov M, Mastrikov YA, Zvejnieks G, Bocharov D, Kotomin EA, Krasnenko V. Water Splitting on Multifaceted SrTiO3 Nanocrystals: Computational Study. Catalysts. 2021; 11(11):1326. https://doi.org/10.3390/catal11111326

Chicago/Turabian Style

Sokolov, Maksim, Yuri A. Mastrikov, Guntars Zvejnieks, Dmitry Bocharov, Eugene A. Kotomin, and Veera Krasnenko. 2021. "Water Splitting on Multifaceted SrTiO3 Nanocrystals: Computational Study" Catalysts 11, no. 11: 1326. https://doi.org/10.3390/catal11111326

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