Water Oxidation Catalysis with Fe2O3 Constrained at the Nanoscale

Mar 1, 2017 - Water Oxidation Catalysis with Fe2O3 Constrained at the Nanoscale. Meir Haim Dahan and Maytal Caspary Toroker. Department of Materials S...
0 downloads 11 Views 2MB Size
Subscriber access provided by University of Newcastle, Australia

Article 2

3

Water Oxidation Catalysis with FeO Constrained at the Nano-Scale Meir Haim Dahan, and Maytal Caspary Toroker J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.6b12666 • Publication Date (Web): 01 Mar 2017 Downloaded from http://pubs.acs.org on March 2, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry C is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 20

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 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Water Oxidation Catalysis with Fe2O3 Constrained at the Nano-scale Meir Haim Dahan and Maytal Caspary Toroker* Department of Materials Science and Engineering, Technion - Israel Institute of Technology, Haifa 3200003, Israel Keywords: DFT, iron oxide, catalysis, OER, water splitting.

Abstract Photo-electrochemical cells containing iron (III) oxide (Fe2O3) have attracted extensive investigations due to their ability to convert solar energy into chemical energy by water splitting. Recently, fabrication of nano-scaled Fe2O3 have been adopted for photoelectrochemical cells in order to increase solar energy absorption and reduce slow diffusion length of charge carriers. In order to understand how nano-scaled confinement influences catalytic efficiency, we perform Density Functional Theory +U calculations of water oxidation on a thin slab of Fe2O3(0001). We consider possible hydrogen vacancies that may appear at high pH and voltage and find that promoting hydrogen vacancy formation improves catalytic efficiency. We also analyze the effect of geometrical strain on the slab that may result from deposition on a substrate. We conclude that nano-Fe2O3 should be grown on a substrate with a similar lattice constant in order to reduce strain and improve catalytic efficiency.

*Corresponding author: E-mail: [email protected] , Tel.: +972 4 8294298.

ACS Paragon Plus Environment

The Journal of Physical Chemistry

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 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1. Introduction Solar energy conversion with photo-electrochemical cells has attracted a significant amount of research and development.1-3 One of the most studied materials in the past couple of decades is iron (III) oxide (Fe2O3).4-7 The many advantages of Fe2O3 include having proper band edge positions and band gap for catalysis and solar energy absorption, respectively. The outstanding advantage which makes Fe2O3 superior to other suggested materials is long-lasting stability.8 Yet, a major limitation is low mobility and high electron-hole recombination rate.9-11 This limitation can be avoided by nano-structuring.12-15 Among the different nanostructures proposed for Fe2O3, nanoribbons are interesting architectures due to their low-dimensionality that may be useful for better solar energy absorption and conversion. There are reports showing that synthesizing thin nanoribbons of Fe2O3 is possible.16-19 As a result, theoretical investigations have also considered models for Fe2O3 nanoribbons.20 However, no study has evaluated how low-dimensionality would influence the water oxidation catalysis ability of Fe2O3. Lower dimensional materials are widely used for many catalytic applications.21 The main reason is the larger surface area that can more efficiently absorb solar energy. In addition, the larger surface contributes additional active sites. Other factors may play a role since the electronic structure also changes with the reduction of dimensionality. Some examples that have been investigated include graphene and transition metal dichalcogenide nanostructures.22-23 In this paper, we explore the contribution of nano-scaling to catalysis for Fe2O3. We constructed periodic slab models of thin nano-ribbons of Fe2O3(0001) and used Density Functional Theory + U (DFT+U) to model the water oxidation catalytic mechanism. Our results show that reducing the number of atomic layers does not change the overpotential required for water oxidation. However, hydrogen vacancies that may appear at smaller scales reduce the overpotential. In addition, we applied strain on the nano-ribbon in order to mimic the geometrical effect of a substrate and found that a large mismatch would not be favorable for catalysis.

2. Methods and Calculation Details Spin-polarized Density Functional Theory calculations were performed with the VASP program.24-25 The DFT+U formalism of Duradev et al. was used with the Perdew-Burke-Ernzerhof (PBE) functional and an on-site Coulomb repulsion26 with an effective U-J term of 4.3 eV for Fe, since we wish to compare to previous calculations with this methodology that was shown successful in predicting electronic structure27-28 and catalytic ability of Fe2O3.29-38 Projected-augmented wave (PAW) potentials replaced the core electrons of Fe 1s2s2p3s and O 1s.39-40

ACS Paragon Plus Environment

Page 2 of 20

Page 3 of 20

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 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

The bulk unit cell structure of Fe2O3 was cleaved at the (0001) facet since this is one of the stable facets at aqueous environment and since we wanted to compare to previous literature that studied this facet.29-30 However, in this work we chose a single Fe2O3 monolayer of Fe atoms, fully coordinated with oxygen atoms and hydroxyl groups. The chemical components of the sheet are: Fe2O6H6, that is a fully hydroxylated Fe2O3. Each one of the two Fe atom is octahedrally coordinated with six hydroxyl groups. Two of the hydroxyl groups are chemically bonded to both Fe atoms. Other parameters were the same as in ref. 29-30 for comparison: the vacuum length was taken to be 10 Å, the energy cutoff and k-point grid were chosen to be 700 eV and a Gammacentered k-point mesh of 3x3x1, respectively. The ion positions in the slabs were converged until the force components on all ions were less than 0.03 eV/Å. Final geometries obtained are provided in the Supporting Information (SI). Vacancy formation energy was calculated by the formula:41 1 ∆𝐸𝑓 (𝐻) = 𝐸ℎ𝑜𝑠𝑡 − 𝐸𝑑𝑒𝑓𝑒𝑐𝑡 − 𝐸(𝐻2 ) 2 where 𝐸ℎ𝑜𝑠𝑡 is the energy of the host surface cell (denote *OH, i.e. terminating hydroxyl group) with no H vacancy or reaction intermediate adsorbed, 𝐸𝑑𝑒𝑓𝑒𝑐𝑡 is the energy of the surface unit cell with a H vacancy, and 𝐸(𝐻2 ) is the energy of a hydrogen diatomic molecule. The surface energy is calculated from the formula: ∆𝐸𝑠𝑢𝑟𝑓 = [𝐸𝑠𝑢𝑟𝑓 − 𝐸𝑏𝑢𝑙𝑘 − 3 ∙ 𝐸(𝐻2 𝑂)]/2𝐴 where 𝐸𝑠𝑢𝑟𝑓 is the total energy of the surface, 𝐸𝑏𝑢𝑙𝑘 is the total energy of the Fe2O3 bulk, three is the number of H2O groups binding to the surface and forming hydroxyl groups on the surface, and A is the surface area. We considered slabs representing reaction intermediates of the following fivestep mechanism that has been studied in refs. 29-30, but modified to reflect alkaline operating conditions, * + H2O  *H2O

(1)

*H2O + h+ + OH-  *OH + H2O

(2)

*OH + h+ + OH-  *O + H2O

(3)

*O + h+ + OH-  *OOH

(4)

*OOH + h+ + OH-  * + O2 + H2O (5) where intermediate "*" is a slab with an oxygen vacancy termination available for adsorption and, for example, intermediate "*H2O" is a slab with an adsorbed H2O molecule. Overall, this reaction mechanism includes one adsorption reaction and four

ACS Paragon Plus Environment

The Journal of Physical Chemistry

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 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

deprotonations. In practice, we remove an H atom rather than a proton in each step since under electrochemical conditions the excess electron is also removed from the surface as a result of applying voltage. The free energy for each reaction was calculated by subtracting reactants and products and adding our calculated Zero Point Energy (ZPE) corrections and entropic contributions. In the supporting information we provide details of ZPE calculation for all slabs, showing negligible differences from either removing atomic layers, adding hydrogen vacancies, or applying strain; hence, we ultimately used ZPE correction from the thick slabs. The applied voltage is accounted for by adding a constant energy -eU of 1 eV to the free energies of reactions 1-5 with respect to the normal hydrogen electrode (NHE). The pH is incorporated by adding −𝐾𝐵 𝑇 ∙ 𝑙𝑛10 ∙ 𝑝𝐻 (𝐾𝐵 is Boltzmann constant and 𝑇 is the temperature of 298.15 K) to the free energy calculations. The overpotential is defined as voltage needed to add to the calculated electrochemical potential so that all reaction free energies are negative.29 The free energies were calculated in the presence of added hydrogen vacancies or strain, i.e. contraction or elongation of the surface lattice constants horizontal to the surface plane. Hydrogen vacancies were added to each reaction intermediate slab at various locations. As seen in Figure 1, we considered both H vacancy below and aside the active site. The unit cells of all slabs are obtained from the bulk unit cell and normally only the ion positions are optimized. In some cases mentioned explicitly, full lattice vectors is also accounted for in order to consider surface reconstruction that appears at the nano-scale. The contraction or elongation of the slab volume by X% was done by multiplying the surface lattice vectors by √𝑋.

Figure 1. Unit cell for the nano-scaled Fe2O3(0001) surface showing locations of hydrogen vacancies for the first reaction intermediate denoted as *vac. Red, brown, and white spheres represent oxygen, iron, and hydrogen atoms, respectively.

3. Results In this section we present calculated free energies of water oxidation for a thin "ribbon" of Fe2O3(0001) in the presence of hydrogen vacancies and strain (see Figure 2). Hydrogen vacancies are anticipated to have strong dominance for a thin material at operating conditions of high pH and voltage. Our calculations reveal that hydrogen vacancy formation energy for a thin nanoribbon is 0.56 eV under pH=14 and E=1 Volt. Furthermore, geometrical strain may be induced by a substrate. We explain that

ACS Paragon Plus Environment

Page 4 of 20

Page 5 of 20

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 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

hydrogen vacancies charge the surface and could improve catalytic performance, while geometrical strain alters the workfunction and has an unfavorable effect on catalysis.

Figure 2. Catalytic cycle of water oxidation with Fe2O3(0001) having one hydrogen vacancy (denoted hydrogen 1 in the text). The geometries are fully optimized. Red, brown, and white spheres represent oxygen, iron, and hydrogen atoms, respectively. Our central result regarding the addition of hydrogen vacancies is that they slightly decrease the free energy of reaction 3 (see Table 1), which determines the overpotential in the stoichiometric case. As a result, the overpotential may decrease. However, when the concentration of hydrogen vacancies is large or when the vacancies are close to the active site (hydrogen vacancy 2 is closer than hydrogen vacancy 1 to the active site), the decrease in the free energy of reaction 3 is accompanied by an increase in the free energy of reaction 4, which is now determining and increasing the overpotential. We note that the free energies of a thicker slab (with the reaction happening on both sides of the slab in order to correct for inter-unit cell dipole interactions) are similar to the free energies for the thin nano-Fe2O3; hence, reducing the number of layers has little effect on catalysis. Table 1. Free energies and overpotentials for Fe2O3(0001) with or without hydrogen vacancies at pH=14 and E=1 Volts. Units are in eV. Results for thick slabs are obtained from ref. 34. The reaction in all thin slabs occurs on one side of the slab, while the reaction on the thick slab occurs at two sides. Free energies without potential and pH corrections are given in the supporting information. Reaction

1 2 3

No No H vacancies, vacancies, vacancy thick slab thin slab 1

-1.78 -1.86 -0.01

-2.18 -1.99 0.05

-1.98 -0.73 -0.16

H vacancy 2

-1.60 -0.78 -0.38

ACS Paragon Plus Environment

H H vacancies vacancies 1 and 2 2 and 3

-1.56 -0.36 -0.40

-1.59 -0.36 -0.28

The Journal of Physical Chemistry

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 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

4 5 overpotential

-0.15 -0.92 0.71

0.03 -0.62 0.77

-0.16 -1.69 0.56

0.27 -2.23 0.99

Page 6 of 20

0.09 -2.48 0.81

-0.12 -2.37 0.60

The changes in the free energies as a result of hydrogen vacancies can be visualized by observing the cumulative free energies (Figure 3). The most dominant feature is the rise of the cumulative free energy in reaction 2. We observed a similar rise in reaction 2 in our previous study,30 when considering Fe vacancies in thick Fe2O3. Both Fe and H vacancies generate excess holes in the lattice that inhibit reaction 2, which involves creating additional holes. The hole generated by a H vacancy is located on an Fe atom according to the magnetization changes observed: the magnetization on an Fe atom reduced from 4.3 to 3.6 Bohr magneton as a result of an excess hole). The surface charging as a result of hydrogen vacancies can explain the trends in the free energies. Hydrogen vacancies create holes and destabilizes the surface. Specifically, intermediates *vac, *OH2, and *OOH are now less neutral and stable. Therefore, the free energy of reactions 2 and 4 increase, while reaction 5 decreases.

Figure 3. Cumulative free energies for water oxidation reaction for Fe2O3(0001) in the a) pure, b) H vacancy, and c) 2H vacancy case. The locations of the vacancies are denoted as "H vacancy 1" and "H vacancies 1 and 2" as shown in Figure 1, respectively. In addition to hydrogen vacancies, strain may have a large effect on nanostructure functionality. In order to explore the effect of strain on catalysis, we considered both geometrical compression and expansion of the lattice vectors (see catalytic cycle in Figure 4). We find that the best performance occurs when there is no geometrical strain. As seen in Table 2, the free energy of reaction 3 (that determines the overpotential at no strain) decreases when the lattice vectors are shorter. However, this favorable decrease comes at the expense of an increase in the free energy of reaction 4 and as a result, an increase in the overpotential. The increase in overpotential is also observed when the lattice constants increase after allowing full relaxation (0.87 Volt in Table 2).

ACS Paragon Plus Environment

Page 7 of 20

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 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 4. Catalytic cycle of water oxidation with Fe2O3(0001) having strain of 10% contraction. The geometries are fully optimized. Red, brown, and white spheres represent oxygen, iron, and hydrogen atoms, respectively. The atomic distance 𝛿 shown at the bottom of each intermediate (in Å) refers to the smaller distance that results from contraction between, for example, the oxygen atoms marked with the yellow crosses.

Table 2. Free energies and overpotentials for Fe2O3(0001) with or without strain at pH=14 and E=1 Volts. Values without operating conditions are given in SI. "Lattice relaxed" refers to fully relaxing the ions and lattice constants of the slab. The percentage is given for the expansion or contraction of the volume. Units are in eV. Free energies without potential and pH corrections are given in the supporting information. Reaction

No strain

1 2 3 4 5 overpotential

-1.78 -1.86 -0.01 -0.15 -0.92 0.71

Lattice 5% 10% 5% 10% 15% relaxed contract contract expand expand expand

-1.55 -1.82 0.15 -0.48 -1.02 0.87

-1.62 -1.97 0.01 0.15 -1.29 0.88

-1.68 -1.88 -0.04 0.27 -1.38 0.99

-2.15 -1.98 0.09 -0.10 -0.59 0.81

-1.65 -1.94 0.14 -0.35 -0.91 0.86

-1.59 -1.89 0.20 -0.50 -0.94 0.92

20% expand

-1.53 -1.82 0.27 -0.54 -1.09 0.99

The effect of strain on the overpotential is apparent in the cumulative free energy curves. As seen in Figure 5, there is a monotonous decrease in reaction step 3 and a rise in reaction step 4. The effect of strain on the reaction free energies can be explained from the changes in the work function. As seen in Figure 6, the valence band edge position decreases (work function increases) when the lattice constants of the slab are elongated. The larger distance between Fe-O bonds localized the atomic orbitals and as a consequence the energy required to extract an electron is larger. The larger work function inhibits *OH slab deprotonation and electron extraction that are occurring in reaction 3, and therefore the free energy of this reaction is larger (Table 2). This result implies that expansion stabilized intermediate *OH relative to intermediate *O, and therefore the free energy of reaction 3 increases while reaction 4 decreases.

ACS Paragon Plus Environment

The Journal of Physical Chemistry

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 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 5. Cumulative free energies for water oxidation reaction for Fe2O3(0001) in the a) pure, b) %10 contraction, and c) 10% expansion.

Figure 6. Valence band edge position for Fe2O3(0001) under applied strain for reaction intermediate *OH. The red single dot shows the valence band edge with a H vacancy and no strain. We note that a major advantage of Fe2O3 is the proper band gap and stability. We calculated the band gap on the Fe2O3 nanoribbon and found a value of 2.6 eV, which is higher than the bulk band gap of 2.1 eV,27 consistent with the quantum size effect. The surface energy is negative -0.31 J/m2 as a result of terminating the surface with hydroxyl groups and indicates the creation of the surface is favorable and therefore stable. Finally, we considered a combination of having hydrogen vacancy 1 and performing geometrical changes. The corresponding free energies for reactions 1-5 are -0.31, 1.11, 1.59, 1.81, and 0.23 eV when contracting 5% of the volume and having a

ACS Paragon Plus Environment

Page 8 of 20

Page 9 of 20

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 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

hydrogen vacancy. After full geometry optimization with a hydrogen vacancy, the free energies are -0.04, 1.62, 2.01, 1.47, and -0.63 eV. We observe that the free energy required for reaction 2 rises as a result of vacancy-generated holes charging the surface. At 5% contraction with a hydrogen vacancy, the free energy of reaction 3 decreases while that of reaction 4 rises. But elongation resulting from full geometry optimization increases the free energy of reaction 3 and the overpotential.

4. Conclusions The water oxidation reaction was modeled for a nano-ribbon of Fe2O3(0001). The model included hydrogen vacancies and geometrical strain that may be dominant at the nano-scale. We considered H vacancies at different positions, as well as contraction and expansion of the unit cell's lattice vectors. Our calculations reveal that H vacancies may reduce the overpotential required for water oxidation. The surface charging with vacancy-generated holes makes the first deprotonation step difficult, but the second deprotonation step requires less energy and the overpotential is lower. We further find that strain is unfavorable for catalyzing water oxidation. Elongating the lattice vectors localizes the atomic orbitals and increases the work function. With a larger work function, electron extraction during reaction 3 is prohibited and the overpotential increases. In contrast, contracting the lattice vectors decreases the work function and reaction 3 becomes more plausible, but in turn reaction 4 has a higher free energy that increases overpotential. Since any strain, contraction or expansion, increases the overpotential, we suggest to grow nano-Fe2O3 on a substrate with similar lattice constants in order to reduce mismatch and strain for maintaining good catalytic ability.

Acknowledgements This research was supported by the Nancy and Stephen Grand Technion Energy Program, the I-CORE Program of the Planning and Budgeting Committee, and The Israel Science Foundation (Grant No. 152/11). This work was supported by the post LinkSCEEM-2 project, funded by the European Commission under the 7th Framework Programme through Capacities Research Infrastructure, INFRA-2010-1.2.3 Virtual Research Communities, Combination of Collaborative Project and Coordination and Support Actions (CP-CSA) under grant agreement no RI-261600.

Supporting information available Further details on free energy calculations and unit cells used for the surfaces are provided in the supporting information.

ACS Paragon Plus Environment

The Journal of Physical Chemistry

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 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Table of Contents Image

5. References 1. Sivula, K.; van de Krol, R., Semiconducting Materials for Photoelectrochemical Energy Conversion. Nat, Rev. Mat. 2016, 1, 15010. 2. Zaera, F., The Surface Chemistry of Catalytic Reactions: Progress and Challenges. J. Phys. Chem. Lett. 2015, 6, 4115-411.6 3. Hisatomi, T.; Kubota, J.; Domen, K., Recent Advances in Semiconductors for Photocatalytic and Photoelectrochemical Water Splitting. Chem. Soc. Rev. 2014, 43, 75207535. 4. Zandi, O.; Hamann, T. W., The Potential Versus Current State of Water Splitting with Hematite. Phys. Chem. Chem. Phys. 2015, 17, 22485-22503. 5. Hayes, D.; Hadt, R. G.; Emery, J. D.; Cordones, A. A.; Martinson, A. B. F.; Shelby, M. L.; Fransted, K. A.; Dahlberg, P. D.; Hong, J.; Zhang, X., et al., Electronic and Nuclear Contributions to Time-Resolved Optical and X-Ray Absorption Spectra of Hematite and Insights into Photoelectrochemical Performance. Ener. Env. sci. 2016, 9, 3754-3769. 6. Tamirat, A. G.; Rick, J.; Dubale, A. A.; Su, W.-N.; Hwang, B.-J., Using Hematite for Photoelectrochemical Water Splitting: A Review of Current Progress and Challenges. Nano. Horizons 2016, 1, 243-267. 7. Soares, M. R. S.; Goncalves, R. H.; Nogueira, I. C.; Bettini, J.; Chiquito, A. J.; Leite, E. R., Understanding the Fundamental Electrical and Photoelectrochemical Behavior of a Hematite Photoanode. Phys. Chem. Chem. Phys. 2016, 18, 21780-21788. 8. Dias, P.; Vilanova, A.; Lopes, T.; Andrade, L.; Mendes, A., Extremely Stable Bare Hematite Photoanode for Solar Water Splitting. Nano Ener. 2016, 23, 70.799. Neufeld, O.; Caspary Toroker, M., Play the Heavy: An Effective Mass Study for ΑFe2O3 and Corundum Oxides. J. Chem. Phys. 2016, 144, 164704.

ACS Paragon Plus Environment

Page 10 of 20

Page 11 of 20

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 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

10. Barroso, M.; Pendlebury, S. R.; Cowan, A. J.; Durrant, J. R., Charge Carrier Trapping, Recombination and Transfer in Hematite ([Small Alpha]-Fe2O3) Water Splitting Photoanodes. Chem. Sci. 2013, 4, 2724-2734. 11. Iordanova, N.; Dupuis, M.; Rosso, K. M., Charge Transport in Metal Oxides: A Theoretical Study of Hematite Α-Fe2O3. J. Chem. Phys. 2005, 122, 144.305 12. Dotan, H.; Kfir, O.; Sharlin, E.; Blank, O.; Gross, M.; Dumchin, I.; Ankonina, G.; Rothschild, A., Resonant Light Trapping in Ultrathin Films for Water Splitting. Nat Mater 2013, 12, 158-164. 13. Boudoire, F.; Toth, R.; Heier, J.; Braun, A.; Constable, E. C., Hematite Nanostructuring Using Electrohydrodynamic Lithography. Appl. Surf. Sci. 2014, 305, 62-66. 14. Warren, S. C.; Voïtchovsky, K.; Dotan, H.; Leroy, C. M.; Cornuz, M.; Stellacci, F.; Hébert, C.; Rothschild, A.; Grätzel, M., Identifying Champion Nanostructures for Solar WaterSplitting. Nat Mater 2013, 12, 842-849. 15. Townsend, T. K.; Sabio, E. M.; Browning, N. D.; Osterloh, F. E., Photocatalytic Water Oxidation with Suspended Alpha-Fe2O3 Particles-Effects of Nanoscaling. Ener. Env. sci . .4270-4275 ,4 ,2011 16. Xu, R.; Yan, H.; He, W.; Su, Y.; Nie, J.-C.; He, L., Ultrathin Α-Fe2O3 Nanoribbons and Their Moiré Patterns. J. Phys. Chem. C 2012, 116, 6879-6883. 17. Liao, L.; Zheng, Z.; Yan, B.; Zhang, J. X.; Gong, H.; Li, J. C.; Liu, C.; Shen ,Z. X.; Yu, T., Morphology Controllable Synthesis of Α-Fe2O3 1d Nanostructures: Growth Mechanism and Nanodevice Based on Single Nanowire. J. Phys. Chem. C 2008, 112, 1078410788. 18. Sarkar, D.; Mandal, M.; Mandal, K., Design and Synthesis of High Performance Multifunctional Ultrathin Hematite Nanoribbons. ACS Appl. Mat. Int. 2013, 5, 11995-12004. 19. Chao, W.; Yiqian, W.; Xuehua, L.; Huaiwen, Y.; Jirong, S.; Lu, Y.; Guangwen, Z.; Federico, R., Structure Versus Properties in Α -Fe 2 O 3 Nanowires and Nanoblades. Nanotech. 2016, 27, 035702. 20. Wanaguru, P.; An, J.; Zhang, Q., DFT+U Study of Ultrathin Α-Fe2O3 Nanoribbons from (110) and (104) Surfaces. J. Appl. Phys. 2016, 119, 084302. 21. Zhou, M.; Lou, X. W. D.; Xie, Y., Two-Dimensional Nanosheets for Photoelectrochemical Water Splitting: Possibilities and Opportunities. Nano Today 2013, 8, 598-618. 22. Lei, X.; Yu, K.; Li, H.; Tang, Z.; Zhu, Z., First-Principle and Experiment Framework for Charge Distribution at the Interface of the Molybdenum Dichalcogenide Hybrid for Enhanced Electrochemical Hydrogen Generation. The Journal of Physical Chemistry C 2016, 120, 15096-15104. 23. Rodriguez, A. M.; Munoz-Garcia, A. B.; Crescenzi, O.; Vazquez, E.; Pavone, M., Stability of Melamine-Exfoliated Graphene in Aqueous Media: Quantum-Mechanical Insights at the Nanoscale. Phys. Chem. Chem. Phys. 2016, 18, 22203-22209. 24. Kresse, G.; Hafner, J., \Textit{Ab Initio} Molecular Dynamics for Liquid Metals. Phys. Rev. B 1993, 47, 558-561. 25. Kresse, G.; Furthmüller, J., Efficiency of Ab-Initio Total Energy Calculations for Metals and Semiconductors Using a Plane-Wave Basis Set. Comp. Mater. Sci. 1996, 6, 15-50. 26. Perdew, J. P.; Burke, K.; Ernzerhof, M., Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 19.3865-3868 ,77 ,96 27. Mosey, N. J.; Liao, P.; Carter, E. A., Rotationally Invariant Ab Initio Evaluation of Coulomb and Exchange Parameters for DFT+U Calculations. J. Chem. Phys. 2008, 129, 014103. 28. Nguyen, M.-T.; Seriani, N.; Gebauer, R., Defective Α-Fe2O3(0001): An Ab Initio Study. ChemPhysChem 2014, 15, 2930-2935. 29. Liao, P.; Keith, J. A.; Carter, E. A., Water Oxidation on Pure and Doped Hematite (0001) Surfaces: Prediction of Co and Ni as Effective Dopants for Electrocatalysis. J. Am. Chem. Soc.13296-13309 ,134 ,2012 .

ACS Paragon Plus Environment

The Journal of Physical Chemistry

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 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

30. Toroker, M. C., Theoretical Insights into the Mechanism of Water Oxidation on Nonstoichiometric and Titanium-Doped Fe2O3(0001). J. Phys. Chem. C 2014, 118, 2316223167. 31. Neufeld, O.; Toroker, M. C., Platinum-Doped Α-Fe2O3 for Enhanced Water Splitting Efficiency: A DFT+U Study. J. Phys. Chem. C 2015, 119, 5836-5847. 32. Yatom, N.; Neufeld, O.; Caspary Toroker, M., Toward Settling the Debate on the Role of Fe2O3 Surface States for Water Splitting. J. Phys. Chem. C 2015, 119.24789-24795 , 33. Yatom, N.; Toroker, M., Hazardous Doping for Photo-Electrochemical Conversion: The Case of Nb-Doped Fe2O3 from First Principles. Molecules 2015, 20, 19668. 34. Neufeld, O.; Yatom, N.; Caspary Toroker, M., A First-Principles Study on the Role of an Al2o3 Overlayer on Fe2O3 for Water Splitting. ACS Catal. 2015, 5, 7237-7243. 35. Zhang, X.; Klaver, P.; van Santen, R.; van de Sanden, M. C. M.; Bieberle-Hütter, A., Oxygen Evolution at Hematite Surfaces: The Impact of Structure and Oxygen Vacancies on Lowering the Overpotential. J. Phys. Chem. C 2016, 120, 18201-18208. 36. Zhang, X.; Cao, C.; Bieberle-Hutter, A., Orientation Sensitivity of Oxygen Evolution Reaction on Hematite. J. Phys. Chem. C 2016. 37. Hellman, A.; Pala, R. G. S., First-Principles Study of Photoinduced Water-Splitting on Fe2O3. J. Phys. Chem. C 2011, 115, 12901-12907. 38. Nguyen, M.-T.; Piccinin, S.; Seriani, N.; Gebauer, R., Photo-Oxidation of Water on Defective Hematite(0001). ACS Catal. 2015, 5, 715-721. 39. Blöchl, P. E ,.Projector Augmented-Wave Method. Phys. Rev. B 1994, 50, 1795317979. 40. Kresse, G.; Joubert, D., From Ultrasoft Pseudopotentials to the Projector AugmentedWave Method. Phys. Rev. B 1999, 59, 1758-1775. 41. Fidelsky, V.; Toroker, M. C., Enhanced Water Oxidation Catalysis of Nickel Oxyhydroxide through the Addition of Vacancies. J. Phys. Chem. C 2016, 120, 25405-25410.

ACS Paragon Plus Environment

Page 12 of 20

Page 13 of 20

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 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 1 82x34mm (300 x 300 DPI)

ACS Paragon Plus Environment

The Journal of Physical Chemistry

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 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2 177x130mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 14 of 20

Page 15 of 20

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 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 3 Figure 3 85x50mm (300 x 300 DPI)

ACS Paragon Plus Environment

The Journal of Physical Chemistry

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 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 4 177x99mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 16 of 20

Page 17 of 20

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 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 5 Figure 5 85x50mm (300 x 300 DPI)

ACS Paragon Plus Environment

The Journal of Physical Chemistry

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 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 6 85x50mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 18 of 20

Page 19 of 20

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 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Table of Contents Image Table of Contents Image 177x127mm (300 x 300 DPI)

ACS Paragon Plus Environment

The Journal of Physical Chemistry

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 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Cover graphics Cover graphics 177x129mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 20 of 20