High-Throughput Computational Screening of Perovskites for

Jun 20, 2016 - Antoine A. Emery, James E. Saal,. †. Scott Kirklin, Vinay I. Hegde, and Chris Wolverton*. Department of Materials Science and Enginee...
10 downloads 0 Views 3MB Size
Subscriber access provided by University of Birmingham

Article

High-Throughput Computational Screening of Perovskites for Thermochemical Water Splitting Applications Antoine A. Emery, James E. Saal, Scott Kirklin, Vinay I. Hegde, and Chris Wolverton Chem. Mater., Just Accepted Manuscript • DOI: 10.1021/acs.chemmater.6b01182 • Publication Date (Web): 20 Jun 2016 Downloaded from http://pubs.acs.org on June 25, 2016

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.

Chemistry of Materials 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 48

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

Chemistry of Materials

High-Throughput Computational Screening of Perovskites for Thermochemical Water Splitting Applications Antoine A Emery, James E Saal a, Scott Kirklin, Vinay I Hegde, and Chris Wolverton∗ Department of Materials Science and Engineering, Northwestern University, Evanston, Illinois 60208, USA E-mail: [email protected] a Current

address: QuesTek Innovations, LLC, Evanston, IL 60201

Abstract The use of hydrogen as fuel is a promising avenue to aid in the reduction of greenhouse effect gases released in the atmosphere. In this work, we present a highthroughput density functional theory (HT-DFT) study of 5,329 cubic and distorted perovskites ABO3 compounds to screen for thermodynamically favorable two-step thermochemical water splitting (TWS) materials. From a dataset of more than 11,000 calculations, we screened materials based on: (a) thermodynamic stability, and (b) oxygen vacancy formation energy that allow favorable TWS. From our screening strategy, we identify 139 materials as potential new candidates for TWS application. Several of these compounds, such as CeCoO3 and BiVO3 , have not been experimentally explored yet for TWS and present promising avenues for further research. We show that taking into consideration all phases present in the A-B-O ternary phase, as opposed to only calculating the formation energy of a compound, is crucial to assess correctly

1

ACS Paragon Plus Environment

Chemistry of Materials

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

Page 2 of 48

the stability of a compound as it reduces the number of potential candidates from 5,329 to 383. Finally, our large dataset of compounds containing stabilites, oxidation states and ionic sizes allowed us to revisit the structural maps for perovskites by showing stable and unstable compounds simultaneously.

Introduction Hydrogen used as fuel presents several advantages. It can be used in fuel cells for stationary as well as mobile applications and the product of its combustion is water, making it potentially a green alternative to fossil fuels 1–4 . Currently, hydrogen is mainly produced by steam reforming of natural gas 5,6 which decreases the environmental advantage of its carbon free combustion. Thus, exploring ways to produce hydrogen in a sustainable and carbon-neutral fashion, is important for its widespread adoption as a fuel. Among different carbon-free routes to produce hydrogen, such as photoelectrochemical (PEC) and electrolytic processes, solar driven thermochemical water splitting (TWS) cycles, i.e. the use of solar thermal energy to drive a set of chemical reactions, have the advantage of using the entire solar spectrum, thus leading to higher theoretical efficiencies. 7–13 Specifically, hydrogen can be produced by splitting the water molecule in a two-step thermochemical cycle as follows: δ MOx → MOx−δ + O2 2

(1)

MOx−δ + δH2O → MOx + δH2

(2)

where M is a metal, MOx its corresponding metal oxide, and δ is the degree of oxygen off-stoichiometry, i.e. the amount of oxygen loss when reducing the metal oxide which is a function of temperature, pressure and the metal oxide. Reaction 1 is the thermal reduction step (TR) occuring typically around 2000 K and Reaction 2 is the gas splitting

2

ACS Paragon Plus Environment

Page 3 of 48

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

Chemistry of Materials

step (GS) occurring at lower temperature (around 1000 K). 10 Using a similar concept, carbon monoxide (CO), a hydrocarbon fuel precursor, can be produced by splitting of the carbon dioxide (CO2 ) molecule. 9 In both cases, the metal oxide and the associated thermodynamics of Eq. (1) and (2) are critical aspects that determine the efficiency of the gas splitting. 14,15 For practical fuel production, metal oxides are required which increase the efficiency and rates at which the thermal reduction and gas splitting reactions occur. Ideally, the materials should be reduced at as low a temperature as possible to avoid energy loss due to heat transfer and thus increase the overall efficiency of the process. 10,11,16,17 However, as the temperature of reduction is also connected to that of the GS, it cannot be too low as to disallow gas splitting. As phase transformations often negatively impact the kinetics and the cyclability of both reactions, it is desirable to have a metal oxide that can accommodate large amounts of oxygen vacancies without changing structure. Other considerations include structural stability at high temperature, as well as cost and availability of the metal oxides. Thus, the choice of the metal oxide is key in terms of the conditions under which solar thermochemical cycles reactions will take place, as well as the capacity of water splitting. In the literature, 280 thermochemical water-splitting cycles were reviewed by Abanades et al. 16 Out of those, 30 were selected as technologically feasible based on temperature of operation, process complexity, cost and toxicity of materials among other criteria. 16 Cycles used successfully to split water include zinc oxide (ZnO/Zn) 17,18 , tin oxide (SnO2 /SnO) 19,20 , iron oxide 8 and ceria 21–26 based cycles. Ceria was proposed by Otsuka et al. 21 as a candidate for water splitting and was studied thoroughly as it appears to be one of the most promising materials for water splitting applications. 21–26 Indeed, it can accommodate a large quantity of oxygen vacancies without phase transformation. 21–26 As a consequence, fast kinetics of thermal reduction and gas splitting reactions were shown by Abanades et al. 22 and ceria was cycled 500 times exhibiting stable oxygen release and

3

ACS Paragon Plus Environment

Chemistry of Materials

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

hydrogen production after 100 cycles. 25 One major drawback of the ceria based cycle is the high temperature needed during the thermal reduction step. In the search for improved materials for TWS, ABO3 perovskites have been recently proposed for water splitting. 27–31 Many perovskites show a high degree of oxygen offstoichiometry. Furthermore, the alloying potential of both the A and B cation sites makes a very large compositional space for promising materials. Lanthanum based perovskites, such as Mn4+ and Sr2+ doped LaAlO3 , have been experimentally studied and showed promising results in terms of oxygen release during thermal reduction (up to eight times larger than ceria). 27 Additionally, La1− x Srx MnO3 exhibited higher solar-to-fuel efficiency than ceria under lower temperature conditions. 28 Finally, Nalbandian et al. tested La1− x Srx MO3 (M=Mn, Fe, x=0, 0.3, 0.7, 1) perovskites in a membrane reactor to produce hydrogen in an isothermal and continuous fashion. 29 The remarkable stability of the perovskite structure with respect to its constituents elements and its ability to tolerate a high degree of oxygen off-stoichiometry, suggests that there could be potentially superior perovskite compounds awaiting discovery. 32 As the number of possible compositions for ABO3 structures is too large to be completely explored experimentally, in this work, we use high-throughput density functional theory (HT-DFT) to search efficiently for novel perovskite (ABO3 ) metal oxides for thermochemical water splitting applications. In this work, 73 elements were chosen as possible A and B cations (see Figure 3), resulting in 5,329 (732 ) ABO3 perovskite stoichiometric compounds. Today’s computational resources enable high-throughput DFT where large numbers of compounds (typically ten or hundred thousands) are calculated to create databases containing energies of numerous materials. 33–38 Databases can then be searched for materials with desired properties, thus accelerating materials design, as shown successfully for various applications, 34,39–42 most notably a perovskites search for photoelectrochemical water splitting. 43,44 Here, we build on such approaches to screen for thermochemical water splitting perovskites.

4

ACS Paragon Plus Environment

Page 4 of 48

Page 5 of 48

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

Chemistry of Materials

We start with the central premise that tuning the thermodynamics of oxide reduction and compound stability is necessary to enable improved material performance of the thermochemical water splitting process. Two filters are used to screen our high-throughput DFT dataset for potential water splitting oxides: 1) Stability: A perovskite compound must be thermodynamically stable in order to be considered as a promising candidate. In this work, stability refers to 0 K and 0 bar DFT stability of ABO3 with respect to all possible combinations of phases present in the A-B-O ternary phase diagram. In other words, the ABO3 perovskite should be on the ground state convex hull, which is defined as the envelope connecting the lowest energy compounds at every composition in the phase space. 45 2) Oxygen vacancy formation energy: In order to split water, both thermal reduction and gas splitting reactions must be thermodynamically favorable, i.e. have a negative Gibbs free energies. Meredig and Wolverton 14,15 showed that, for typical TR and GS temperatures, these conditions impose limits on the enthalpy and entropy of reduction, creating a window where both reactions are thermodynamically favorable. In the present study, we use these limits on the enthalpy of reduction as a second filter. To avoid false negative classification of compounds and to account for the approximations inherent to the high-throughput character of this study, we use a slightly larger oxygen vacancy formation energy window; we consider compounds that have an oxygen vacancy formation energy between 2.5 and 5.0 eV/O atom to be suitable for TWS. After calculating 5,329 different compositions in the cubic, rhombohedral, tetragonal and orthorhombic distortions of perovskite, we found 383 stable compounds. Among those, 139 fell within the suitable oxygen vacancy formation energy range for water splitting. In addition, we identified that rare earth elements and 3d-transition metals are prominent on the A- and B-sites of stable perovskites, respectively. The high-throughput database of compounds and energies allowed us to reconsider structural maps for the perovskite phase across a very wide range of possible A- and B-site chemistries.

5

ACS Paragon Plus Environment

Chemistry of Materials

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

Methodology DFT All DFT calculations were performed using the Vienna ab initio simulation package (VASP). 46,47 Projector-augmented wave method (PAW) potentials 48 are used with the Perdew-Burke-Ernzerhof (PBE) 49 generalized gradient approximation to the exchangecorrelation functional. To improve the description of localized charge density of some 3d transition metals and actinides (V, Cr, Mn, Fe, Co, Ni, Cu, Th, U, Np, Pu), DFT+U is used. 50,51 Any calculations containing actinides or 3d elements (Sc-Cu) are spin-polarized with ferromagnetic alignment of spins. We note that this approach does not capture antiferromagnetism which is present in certain perovskites. 52 Based on a study by Stevanovic et al., who calculated ternary compounds with up to ten different relative spin orientations, this computational error was found to be of the order of 0.01-0.02 eV/atom and thus is considered negligible for the purpose of this paper. 53 Calculations are performed through the Open Quantum Materials Database (OQMD, www.oqmd.org) framework. 34,54 The OQMD contains the energies of over 440,000 compounds comprising ∼40,000 structures from the Inorganic Crystal Structure Database (ICSD) 55,56 as well as more than 400,000 theoretical prototype structures. Calculation settings are explained in more detail in Kirklin et al. 54

Assessment of T = 0 K Stability The first filter that will be used to assess the perovskite compounds is stability. To determine, for a given A-B-O system, which phases are stable it is necessary to find, as a function of composition, the set of phases that have an energy lower than any other structures or any other linear combinations of structures. These ground state phases are then linked with tie lines to form a convex hull. The stability of a perovskite is given in terms of the energy difference between the perovskite and the convex hull, also referred as the hull 6

ACS Paragon Plus Environment

Page 6 of 48

Page 7 of 48

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

Chemistry of Materials

(a)

(b)

(c)

(d)

¯ 221). Green atoms are the A-site cation, Figure 1: (a) Cubic perovskite structure (Pm3m, blue atoms are the B-site cations that are embedded in oxygen (red atoms) octahedron. (b) ¯ 167). (c) Tetragonal distortion (P4mm, 99). (d) OrthorhomRhombohedral distortion (R3c, bic distortion (Pnma, 62). The unit cell (black lines ) of the cubic, rhombohedral, tetragonal and orthorhombic structure contains 5, 10, 5 and 20 atoms respectively. Distortions are exaggerated to be visible. distance, is defined as: ABO3 3 ∆Hstab = ∆H ABO − ∆H f f

(3)

where ∆H f is the convex hull energy at the ABO3 composition (not including the perovskite under consideration). Stable perovskites that are already reported and already in the OQMD will have a stability energy of zero (Eq. 3). Stable perovskites that are discovered during the course of this work will have a negative stability energy (Eq. 3) and perovskites ABO3 3 with a positive value of ∆Hstab are unstable. Finally, ∆H ABO is the DFT formation f

enthalpy of the perovskite calculated as follows: 3 ∆H ABO = E( ABO3 ) − µ A − µ B − 3 ∗ µO f

(4)

where E( ABO3 ) is the DFT total energy of the ABO3 compound, µ A , µ B and µO are the chemical potentials of the A, B and oxygen species respectively. In most cases, µ A and µ B are DFT 0 K total energies of the crystalline elements. However, for elements exhibiting

7

ACS Paragon Plus Environment

Chemistry of Materials

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

a phase transition between 0 K and 300 K (Na, Ti, Sn, O, Hg) and elements with DFT+U correction (V, Cr, Mn, Fe, Co, Ni, Cu, Th, U, Np, Pu), chemical potentials are fitted to experimental values as detailed in Ref 54 . In particular, the oxygen chemical potential (µO ) is fitted to experimental oxide ∆H f values. Extensions of the grand canonical linear programming method (GCLP) 57 are used to construct convex hulls for every calculated system. 54,57 All 440,000 phases present in the OQMD are included in the stability calculation.

Structural Distortions of Perovskite Aside from the ideal cubic structure (Figure 1 (a)), perovskites can exhibit several distortions. 32 The most frequently encountered are rhombohedral, tetragonal and orthorhombic distortions 32 (Figure 1 (b), (c) and (d)). Without experimental data for a given composition, or without using costly crystal structure prediction tools, it is not possible, a priori, to know what is the most stable structure of any given compound. As calculating every distortion (Figure 1) for all the 5,329 stoichiometries would be computationally costly, we started by calculating the 5,329 different compounds in the undistorted cubic perovskite structure. To see the general effect of distortions on the convex hull stability of perovskites, we randomly selected one third (1,776) of the compositions and performed calculations of their distorted perovskite structures. Although the structural distortions often lower the energy of the ideal structure, we find no case (of the 1,776) where the distortions lower the energy by more than 0.5 eV/atom. As a result, we only need to calculate all distortions for all cubic perovskites with a stability (Eq. 3) lower than 0.5 eV/atom. With this approach, we calculated 2,162 stoichiometries that required calculation of their rhombohedral, tetragonal and orthorhombic distortions. Hence, the total number of calculations is approximately 12,000.

8

ACS Paragon Plus Environment

Page 8 of 48

Page 9 of 48

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

Chemistry of Materials

Oxygen Vacancy Formation Energy The second filter that is used to screen potential candidates for water splitting applications is the oxygen vacancy formation energy. For a general case, the oxygen vacancy formation energy, per vacancy, is calculated as follows: ∆EvO =

1 1 E( ABO3−δ ) + µO − E( ABO3 ) δ δ

(5)

where δ is the oxygen off-stoichiometry, E( ABO3 ) is the DFT total energy of the defect-free cubic perovskite cell and µO is the same chemical potential as used in Eq. 4. Finally, E( ABO3−δ ) is the DFT total energy of a cubic supercell containing an oxygen vacancy. In the dilute limit, ∆EvO in Eq. 5 should be converged as a function of supercell size, until there is negligible interaction between the vacancy and its periodic images. However, large supercells are too computationally costly for our high-throughput approach. Hence, we use small supercell A2 B2 O5 structures, which correspond to a 2x1x1 supercell size, and have a small unit cell (9 atoms) which enables the calculation of the oxygen vacancy formation energy for all perovskites in a high-throughput manner. Additionally, Curnan & Kitchin, who calculated oxygen vacancy formation energies for LaBO3 and SrBO3 (B = Sc-Cu) with different degrees of simplification, showed that trends in the oxygen vacancy formation energy are largely unaffected by the supercell size. 58 We confirmed this finding by performing calculations of several LaBO3 perovskite vacancy formation energies for various supercell sizes, as shown in Figure 2.

Oxidation States and Ionic Size To keep our high-throughput methodology exhaustive, we did not restrict any compounds based on their charge balance or oxidation state prior to performing the calculations. However, to compare our stability results with previous study as well as drawing structure maps, it is necessary to estimate the oxidation state of A and B atoms in the perovskite 9

ACS Paragon Plus Environment

Chemistry of Materials

  (2Y >H92 DWRP@

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

Page 10 of 48

    

DWRPVXSHUFHOO WKLVZRUN DWRPVXSHUFHOO DWRPVXSHUFHOO H[SHULPHQW /D&U2 /D0Q2 /D)H2

/D&R2

/D1L2

Figure 2: Comparison of oxygen vacancy formation energy calculated, in a highthroughput way, in the present work with oxygen vacancy formation energies calculated with larger supercell. Data are taken from Lee et al. 52 for the 40 atoms supercells and Deml et al. 59 for the 80 atoms supercells. Experimental data are taken from Kuo et al. 60 and Nowotny et al. 61 for LaMnO3 , Mizusaki et al. 62 for LaFeO3 and Mizusaki et al. 63 for LaCoO3 . compositions. For that purpose, we used a bond valence method (BVM) 64 to obtain the oxidation states of the cations for each of the 5,329 compositions that we studied. By fixing the oxidation state of oxygen to -2, we calculated the oxidation state of both cations using a BVM as implemented in pymatgen. 65 Nine elements used in our study (Tc, Os, Pt, Au, Pm, Ac, Pa, Np and Pu) do not have bond valence s. Thus, compounds containing these elements are labeled as having an unknown charge state. With the oxidation states and the coordination numbers (12, 6 and 2 for the A-, B-atom and oxygen, respectively) of all the atoms in the structure, we used Shannon radii, 66,67 which were tabulated by Seshadri and Basu, 68 to estimate the size of each elements.

10

ACS Paragon Plus Environment

Page 11 of 48

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

Chemistry of Materials

Results And Discussion Stability The results for perovskite stability of all compositions are summarized in Table 1. A vast 3 majority (92%) of the compounds have negative formation energy (∆H ABO as defined in f

Eq. 4). This result is expected as we are combining electropositive elements, in our case metals and semi-metals, to oxygen, the second highest electronegative element. A negative formation energy only indicates stability of a compound with respect to its constituent elements and is thus a necessary but not sufficient condition for stability. In this work, we ABO3 consider compounds with a stability (∆Hstab ) lower than 25 meV/atom (approximately

kT at room temperature) to be either stable or nearly so. Out of all the compounds that 3 have ∆H ABO < 0, only a small fraction (4%) are stable with respect to all the phases f

present in the A-B-O convex hull. We report 383 stable perovskite compounds which is considerably more than what was reported in a similar recent study. 69 We explain ¨ this discrepancy by the fact that, as opposed to Korbel et al., we included lanthanides as potential cations. Indeed, lanthanide-containing perovskites represent approximatively 50% of the compounds we predict to be stable. In Table 1, rhombohedral, tetragonal or orthorhombic compounds that are stable and relax to a higher symmetry group (i.e rhombohedral to cubic, tetragonal to cubic, tetragonal to rhombohedral, orthorhombic to tetragonal, orthorhombic to rhombohedral or orthorhombic to cubic) are labeled as stable in their highest symmetry group. The orthorhombic distortion accounts for 84% of all the stable perovskites, which is in accordance with the literature stating that space group 62 is the most common perovskite structure 70 . The difference between negative formation energy and stability shows the importance of having a large and complete database, including theoretical prototypes, available when computing the high-throughput stability of compounds by first-principles calculations. The database is of particular relevance when dealing with compounds that contain elements that have not been thoroughly 11

ACS Paragon Plus Environment

Chemistry of Materials

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

Page 12 of 48

explored experimentally as missing phases in those systems can lead to inaccurate stability calculations. Our high-throughput study allows us to obtain a complete picture of the stability of the perovskite compounds in terms of frequency of elements on the A- and B-sites. Figure 3 shows the number of stable occurrences for each elements on the A- or B-site. Out of the 73 elements we considered for this study, 44 appear in stable perovskites on the A-site and 57 appear on the B-site. Three elements (B, Mg, Zn) do not appear in any stable ABO3 perovskites, regardless of the other metal or whether these elements are placed on the A or B sites. Conversely, 30 elements are predicted to form perovskites in both the A- and the B-site. Numerous perovskites containing elements that are not extensively studied experimentally such as lanthanides and actinides are predicted to be stable and present opportunities for new materials discovery. Alkali metals, alkaline earths, rare earths and 3d-transition metals are heavily represented as elements forming stable perovskite structure which is consistent with other work in the literature 71,72 . To see similar behavior among elements from the same group, we plotted the stable perovskites in a network graph by clustering elements by groups (see Figure 4). In this plot, each connecting line represents a stable compounds (i.e. 383 tie-lines are drawn). Each disk represents an element that is present in at least one stable perovskite (i.e. 70 disks are displayed). Elements are colored and clustered by groups and their size is proportional to the amount of lines connected to them. The curvature of the lines is always clockwise which gives an information on which elements is on the A- and the B-site. For instance, the lines connecting the rare earth and the 3d transition metals are all convex, indicating that the rare earths are on the A-sites and the 3d transition metals are on the B-sites. We observe a high occurrence of alkali metals, alkaline earths and rare earths on the A-sites and 3d-transition metals and 4d-transition metals on the B-sites. This picture is in agreement similar studies found in literature 32 . We find a total of 383 stable perovskites, shown in Figure 5. We observe a cluster of

12

ACS Paragon Plus Environment

Page 13 of 48

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

Chemistry of Materials

A

Li Be

B

B

Na Mg

Al

K Ca Sc Ti Rb Sr

Y

V

Si

Cr Mn Fe Co Ni Cu Zn Ga Ge As

Zr Nb Mo Tc Ru Rh Pd Ag Cd In Sn Sb Te

Cs Ba Lu Hf Ta W Re Os Ir

Pt Au Hg Tl Pb Bi

La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Ac Th Pa U Np Pu 0

4

8

12 16 20 24 # of stable occurrences on A and B sites

28

32

Figure 3: ABO3 perovskite stability by frequency of A- and B-site cations. Periodic table containing all the elements that were considered for the A- and B-sites. Elements are color-coded based on the number of stable occurence on the A- or B-site.

13

ACS Paragon Plus Environment

Chemistry of Materials

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

Page 14 of 48

Figure 4: Network graph of stable perovskites in terms of cation frequency. Each disk represents an element that can be on the A- or B-site, their size is proportional to the number of lines connected to them and their color and positioning helps clustering elements from the same group. Each line corresponds to a stable compound, their color is a mixing of both disks that they connect. The curvature of the lines is always clockwise. TM stands for transition metals.

14

ACS Paragon Plus Environment

Page 15 of 48

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

Chemistry of Materials

compounds containing rare earth on the A-site and transition metals on the B-site, which is likely due to the multiple possible oxidation states for those elements, allowing them to be in the +1/+5, +2/+4 or +3/+3 A/B cation oxidation state configuration and form charge neutral compounds with numerous other elements. This result is in agreement with the observation that the B-site atom is almost always smaller the A-site atoms in the perovskite structure as it is embedded in a rigid octahedron of oxygen. 73–75 Additionally, we see very few compounds containing transition metals (rows labeled 3d, 4d and 5d in Figure 5) on the A-site, probably for a similar reason: transition metals are generally smaller than the other elements and thus preferentially tend to occupy the B-site. Conversely, most of the perovskites predicted to be stable in the undistorted cubic geometry contain bigger elements such as alkali and alkaline-earth metals (Li, Be, Na, Mg, K, Ca, Rb, Sr, Cs and Ba) on the A-site. It is worth noting that the high occurrence of distorted orthorhombic perovskites showcases their 0 K ground state. However, perovskites often exhibit a phase transformation between low and high temperature, which is likely to happen at temperatures required for TWS 32 . With so few rhombohedral and tetragonal perovskites, no obvious trends emerge in their constituting elements. Table 1: Breakdown of calculated stable perovskites by the type of structural distortion 3 favored. ∆H ABO corresponds to the formation energy with respect to the constituent f ABO3 elements of the perovskite and ∆Hstab is the stability of a compound with respect to all other phases present in the A-B-O phase diagram

Distortion

Number calculated Cubic 5329 Rhombohedral 2162 Tetragonal 2162 Orthorhombic 2162 Total 11,815

3 ∆H ABO f @ FXELFSHURYVNLWH UKRPERKHGUDOSHURYVNLWH  

















WHWUDJRQDOSHURYVNLWH RUWKRUKRPELFSHURYVNLWH

$OO&&RPSRXQGV &KDUJH%DO DQFHG RPSRXQGV 

U$ > @

U$ > @

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

Page 24 of 48

E

XQVWDEOH

D















      U% > @ U% > @

24

ACS Paragon Plus Environment

Page 25 of 48

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

Chemistry of Materials

Figure 6: Perovskite structure maps of radii of A- and B-cations (r A vs r B ). Colored symbols are stable perovskites and greyed out symbols are unstable. After calculating the oxidation state of both cations for each compounds using the bond valence method, 64 Shannon radii were used for each element. 66,67 The dashed line represents the x=y line. (a) Plot containing all compounds (5,329 compositions) and (b) Plot containing only charge balanced compounds (2215 compositions).

25

ACS Paragon Plus Environment

Chemistry of Materials

$OO&RPSRXQGV



D



 

 







 

  





FXELFSHURYVNLWH UKRPERKHGUDOSHURYNVLWH

  7ROHUDQFHIDFWRU



 



WHWUDJRQDOSHURYVNLWH RUWKRUKRPELFSHURYVNLWH

XQVWDEOH

&KDUJH%DODQFHG&RPSRXQGV



E



 

 

 







 

  

1XPEHURIVWDEOHFRPSRXQGV







  7ROHUDQFHIDFWRU

26



ACS Paragon Plus Environment



 

1XPEHURIVWDEOHFRPSRXQGV

1XPEHURIXQVWDEOHFRPSRXQGV



1XPEHURIXQVWDEOHFRPSRXQGV

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

Page 26 of 48

Page 27 of 48

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

Chemistry of Materials

Figure 7: Frequency of stable perovskites in function of their tolerance factor. Tolerance √ factor is definied as r A + rO = 2(r B + rO ). 82 Shannon radii were used for to compute the tolerance factors of each compound. 66,67 The y-axis is normalized by the number of stable perovskites (389) for the stable compounds and unstable compounds (4940) for the unstable compounds. (a) Plot containing all compounds (5,329 compositions) and (b) Plot containing only charge balanced compounds (2215 compositions).

27

ACS Paragon Plus Environment

Chemistry of Materials



(2Y >H92 DWRP@

    

D 









$%2 +VWDE >H9DWRP@

FXELFSHURYVNLWH UKRPERKHGUDOSHURYVNLWH

WHWUDJRQDOSHURYVNLWH RUWKRUKRPELFSHURYVNLWH









 

(2Y >H92 DWRP@

(2Y >H92 DWRP@

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

Page 28 of 48

  

 

&HULD

 D

   E











$%2  >H9DWRP@  +VWDE $%2

+VWDE



28

  >H9DWRP@

ACS Paragon Plus Environment







Page 29 of 48

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

Chemistry of Materials

Figure 8: Perovskite oxygen vacancy formation energy (∆EvO ) plotted against stability ABO3 (∆Hstab ). The green windows represents the oxygen vacancy formation energy target. (a) The stabilities of distortions with the lowest energy are plotted. For the 3167 (53292162) compounds where we did not calculate distortions, the cubic stability is plotted. (b) Blowup of the stable phases region. Experimental and calculated oxygen vacancy formation energies of ceria are taken from Chiang et al., 84 Yang et al. 85 and Murgida et al. 86

29

ACS Paragon Plus Environment

Chemistry of Materials

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

Page 30 of 48

Summary and Conclusions Going beyond binary oxides for thermochemical water splitting applications opens a large composition space that is unreasonably big to be entirely explored experimentally. In this work, we used high-throughtput density functional theory to screen ABO3 perovskites based on thermodynamic considerations. We did an exhaustive search of the all the possible ABO3 combinations, without filtering for charge neutrality prior to performing the calculations which lead to the discovery of stable perovskites that have hard-to-predict oxidation states. We used two filters, compounds stability and oxygen vacancy formation energy, to isolate potential candidates for water splitting. The stability filter showed the importance of considering all competing phases present in the ternary A-B-O phase diagram to assess the stability of a compounds accurately. We found the majority of the stable perovskites to be orthorhombic with rare earth elements on the A-site and 3d-transition metals on the B-site. Plotting stable and unstable compounds in structural maps and computing their tolerance factor lead to the conclusion that purely geometrical argument are not sufficient to describe completely the formability of perovskites. Finally, we identified 139 perovskites that are predicted to be thermodynamically favorable for water splitting applications, some of those not reported in the literature. The high-throughput methodology presented in this paper shows the benefit of using first-principles calculations to efficiently screen an exhaustively large number of compounds at once. It provides a baseline for further studies involving more detailed exploration of a restricted number of those compounds.

Acknowledgement The authors would like to thank the Institute for Sustainability and Energy at Northwestern (ISEN) for its support in the form of a booster grant. This work was also funded by the U.S. Department of Energy, Office of Science, Basic Energy Sciences under Grant DE30

ACS Paragon Plus Environment

Page 31 of 48

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

Chemistry of Materials

FG02-07ER46433. S.K. was supported by the Center for Electrochemical Energy Science (CEES), an Energy Frontier Research Center funded by the US Department of Energy, Office of Science and Office of Basic Sciences. V.I.H. was supported by the National Science Foundation through grant DMR-1309957. This research used computational resources provided by the Quest high performance computing facility at Northwestern University which is jointly supported by the Office of the Provost, the Office for Research, and Northwestern University Information Technology as well as the National Energy Research Scientific Computing Center, a DOE Office of Science User Facility supported by the Office of Science of the U.S. Department of Energy under Contract No. DE-AC02-05CH11231. The authors also acknowledge helpful discussions with Kenneth R. Poeppelmeier. Table 3: List of perovskites, ordered by ascending vacancy formation energy, for water ABO3 splitting application. All these compounds pass the stability (∆Hstab < 0.025 eV/atom) O and oxygen vacancy formation energy (2.5 < ∆Ev < 5 eV/O atom) screens. Compounds in bold are not reported in any of the review papers used for the literature survey. 71,76–78 Oxidation states were calculated using bond valence parameters 64 as implemented in pymatgen. 65 Formula

Stable Distortion

ABO3 ∆Hstab

∆EvO

[eV/atom]

[eV/O atom]

Ox. N. A

Ox. N.B

3

TbCoO3

orthorhombic

0.001

2.517

3

EuOsO3

orthorhombic

-0.130

2.517

unknown

BiMnO3

orthorhombic

0.019

2.550

3

3

BiFeO3

orthorhombic

0.012

2.593

3

3

NaNpO3

orthorhombic

-0.095

2.664

unknown

AuPaO3

rhombohedral

-0.081

2.683

unknown

EuSbO3

orthorhombic

-0.036

2.699

3

3

LuFeO3

orthorhombic

0.009

2.732

3

3

EuRuO3

orthorhombic

-0.113

2.753

2

4

YbPuO3

orthorhombic

0.024

2.789

unknown

31

ACS Paragon Plus Environment

Chemistry of Materials

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

Page 32 of 48

Table 3 – continued Formula

Stable Distortion

ABO3 ∆Hstab

∆EvO

[eV/atom]

[eV/O atom]

Ox. N. A

Ox. N.B

AcRuO3

orthorhombic

0.006

2.793

unknown

SmCoO3

orthorhombic

-0.018

2.937

3

YbTcO3

orthorhombic

0.003

2.953

unknown

LaCoO3

orthorhombic

-0.023

2.970

3

PmCoO3

orthorhombic

-0.022

2.996

unknown

NaOsO3

orthorhombic

-0.024

3.020

unknown

LiTcO3

orthorhombic

0.006

3.047

unknown

HoFeO3

orthorhombic

0.008

3.100

3

3

NdCoO3

orthorhombic

-0.001

3.130

3

3

SrRuO3

orthorhombic

-0.017

3.142

2

4

YFeO3

orthorhombic

-0.113

3.196

3

3

DyFeO3

orthorhombic

0.008

3.199

3

3

CeCoO3

orthorhombic

-0.044

3.216

3

3

TbFeO3

orthorhombic

0.008

3.270

3

3

EuReO3

orthorhombic

-0.085

3.289

2

4

PmInO3

orthorhombic

0.023

3.292

unknown

CuUO3

rhombohedral

0.011

3.294

1

KOsO3

cubic

0.015

3.300

unknown

SrOsO3

orthorhombic

-0.077

3.331

unknown

AcCoO3

rhombohedral

0.000

3.336

unknown

BiVO3

orthorhombic

0.000

3.346

3

3

PrInO3

orthorhombic

0.005

3.349

3

3

YbSnO3

orthorhombic

0.023

3.357

2

4

GdFeO3

orthorhombic

-0.022

3.390

3

3

32

ACS Paragon Plus Environment

3

3

5

Page 33 of 48

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

Chemistry of Materials

Table 3 – continued Formula

Stable Distortion

ABO3 ∆Hstab

∆EvO

[eV/atom]

[eV/O atom]

Ox. N. A

Ox. N.B

NdInO3

orthorhombic

0.013

3.400

3

3

LaInO3

orthorhombic

0.018

3.403

3

3

CaTcO3

orthorhombic

0.017

3.425

unknown

YMnO3

orthorhombic

0.011

3.474

3

3

TbMnO3

orthorhombic

0.002

3.526

3

3

CeInO3

orthorhombic

-0.004

3.536

3

3

CaVO3

orthorhombic

-0.020

3.569

2

4

EuPuO3

orthorhombic

-0.103

3.589

unknown

GdMnO3

orthorhombic

0.002

3.606

3

3

SmFeO3

orthorhombic

-0.029

3.630

3

3

NaTcO3

cubic

0.000

3.635

unknown

EuNpO3

orthorhombic

-0.001

3.640

unknown

PuZrO3

orthorhombic

0.021

3.640

unknown

BaOsO3

cubic

-0.113

3.677

unknown

EuSnO3

orthorhombic

-0.193

3.692

2

PmFeO3

orthorhombic

-0.072

3.706

unknown

SmMnO3

orthorhombic

0.001

3.736

2

4

AgUO3

rhombohedral

-0.009

3.754

1

5

PmMnO3

orthorhombic

-0.064

3.769

unknown

SrPuO3

orthorhombic

-0.121

3.789

unknown

SrVO3

orthorhombic

-0.001

3.792

2

4

TmVO3

orthorhombic

-0.011

3.822

2

4

YbVO3

orthorhombic

-0.105

3.829

2

4

NdFeO3

orthorhombic

-0.035

3.836

3

3

33

ACS Paragon Plus Environment

4

Chemistry of Materials

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

Page 34 of 48

Table 3 – continued Formula

Stable Distortion

ABO3 ∆Hstab

∆EvO

[eV/atom]

[eV/O atom]

Ox. N. A

Ox. N.B

EuTcO3

orthorhombic

-0.164

3.840

unknown

BiCrO3

orthorhombic

0.012

3.849

3

3

NdMnO3

orthorhombic

0.005

3.849

3

3

KTcO3

cubic

0.000

3.883

unknown

ErVO3

orthorhombic

0.005

3.887

3

3

PrMnO3

orthorhombic

0.002

3.889

3

3

LaMnO3

orthorhombic

0.000

3.927

3

3

LuCrO3

orthorhombic

-0.011

3.932

3

3

HoVO3

orthorhombic

0.006

3.932

2

4

SrSnO3

orthorhombic

-0.047

3.943

2

4

CeMnO3

orthorhombic

-0.085

3.955

3

3

PrFeO3

orthorhombic

-0.086

3.974

3

3

PuHfO3

orthorhombic

-0.042

3.987

unknown

YVO3

orthorhombic

-0.037

3.990

3

AcInO3

orthorhombic

-0.027

3.994

unknown

DyVO3

orthorhombic

0.005

3.995

2

4

YbGeO3

orthorhombic

0.001

4.029

2

4

TmCrO3

orthorhombic

-0.022

4.037

3

3

LaFeO3

orthorhombic

0.007

4.043

3

3

TbVO3

orthorhombic

0.004

4.045

2

4

YbWO3

orthorhombic

-0.075

4.046

2

4

UScO3

orthorhombic

0.013

4.052

3

3

LiUO3

rhombohedral

0.003

4.062

1

5

CsUO3

cubic

0.009

4.073

1

5

34

ACS Paragon Plus Environment

3

Page 35 of 48

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

Chemistry of Materials

Table 3 – continued Formula

Stable Distortion

ABO3 ∆Hstab

∆EvO

[eV/atom]

[eV/O atom]

Ox. N. A

Ox. N.B

3

CeFeO3

orthorhombic

-0.092

4.100

3

AcMnO3

orthorhombic

-0.116

4.100

unknown

GdVO3

orthorhombic

0.005

4.112

3

3

EuVO3

orthorhombic

-0.075

4.123

2

4

ErCrO3

orthorhombic

0.002

4.126

3

3

CuPaO3

rhombohedral

-0.184

4.134

unknown

AcFeO3

orthorhombic

-0.115

4.141

unknown

SrTcO3

cubic

0.002

4.154

unknown

BaSnO3

cubic

0.001

4.158

2

4

TlUO3

cubic

-0.105

4.189

1

5

HoCrO3

orthorhombic

-0.033

4.191

3

3

BaPuO3

rhombohedral

0.010

4.211

unknown

YbMoO3

orthorhombic

0.007

4.239

2

4

SmVO3

orthorhombic

-0.064

4.244

2

4

DyCrO3

orthorhombic

0.003

4.245

3

3

YCrO3

orthorhombic

-0.020

4.254

3

3

TmGaO3

orthorhombic

0.019

4.279

3

3

PmVO3

orthorhombic

-0.069

4.301

unknown

TbCrO3

orthorhombic

-0.047

4.319

3

3

NaReO3

cubic

0.007

4.345

1

5

NaUO3

orthorhombic

0.003

4.356

1

5

EuGeO3

orthorhombic

-0.216

4.370

2

4

GdCrO3

orthorhombic

-0.052

4.371

3

3

AgPaO3

rhombohedral

-0.295

4.373

unknown

35

ACS Paragon Plus Environment

Chemistry of Materials

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

Page 36 of 48

Table 3 – continued Formula

Stable Distortion

ABO3 ∆Hstab

∆EvO

[eV/atom]

[eV/O atom]

Ox. N. A

Ox. N.B

NdVO3

orthorhombic

-0.077

4.379

3

3

UAlO3

cubic

-0.004

4.389

3

3

BaNpO3

orthorhombic

-0.088

4.442

unknown

PrVO3

orthorhombic

-0.083

4.464

3

3

UVO3

orthorhombic

0.021

4.467

3

3

YGaO3

orthorhombic

-0.090

4.492

3

3

NpVO3

orthorhombic

-0.032

4.514

unknown

RbUO3

cubic

-0.050

4.519

1

5

SmCrO3

orthorhombic

-0.064

4.544

3

3

CeVO3

orthorhombic

-0.090

4.548

3

3

PuAlO3

orthorhombic

-0.071

4.577

unknown

GdGaO3

orthorhombic

0.007

4.590

3

PmCrO3

orthorhombic

-0.067

4.600

unknown

LaVO3

orthorhombic

-0.083

4.615

3

3

NdCrO3

orthorhombic

-0.075

4.684

3

3

NaMoO3

cubic

-0.004

4.704

1

5

DyTiO3

orthorhombic

0.022

4.721

2

4

SmGaO3

orthorhombic

0.019

4.741

3

3

PrCrO3

orthorhombic

-0.079

4.772

3

3

PmGaO3

orthorhombic

-0.023

4.795

unknown

TbTiO3

orthorhombic

0.015

4.803

2

4

CeCrO3

orthorhombic

-0.081

4.838

3

3

CeErO3

orthorhombic

0.023

4.848

3

3

LaCrO3

orthorhombic

-0.071

4.851

3

3

36

ACS Paragon Plus Environment

3

Page 37 of 48

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

Chemistry of Materials

Table 3 – continued Formula

Stable Distortion

ABO3 ∆Hstab

∆EvO

[eV/atom]

[eV/O atom]

Ox. N. A

Ox. N.B

NdGaO3

orthorhombic

-0.006

4.867

3

3

GdTiO3

orthorhombic

0.023

4.871

2

4

NpCrO3

orthorhombic

-0.021

4.872

unknown

EuMoO3

orthorhombic

-0.119

4.931

2

4

LaGaO3

orthorhombic

0.015

4.940

3

3

PrGaO3

orthorhombic

-0.056

4.942

3

3

CeTmO3

orthorhombic

0.015

4.959

3

3

NdLuO3

orthorhombic

0.022

4.969

3

3

AcVO3

orthorhombic

-0.106

4.974

unknown

References ¨ (1) Schlapbach, L.; Zuttel, A. Hydrogen-storage materials for mobile applications. Nature 2001, 414, 353–358. (2) Weber, A.; Ivers-Tiff´ee, E. Materials and concepts for solid oxide fuel cells (SOFCs) in stationary and mobile applications. J. Power Sources 2004, 127, 273–283. (3) Singhal, S. Solid oxide fuel cells for stationary, mobile, and military applications. Solid State Ionics 2002, 152-153, 405–410. (4) von Spakovsky, M.; Olsommer, B. Fuel cell systems and system modeling and analysis perspectives for fuel cell development. Energy Convers. Manag. 2002, 43, 1249–1257. (5) Koroneos, C.; Dompros, A.; Roumbas, G.; Moussiopoulos, N. Life cycle assessment of hydrogen fuel production processes. Int. J. Hydrogen Energy 2004, 29, 1443–1450.

37

ACS Paragon Plus Environment

Chemistry of Materials

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

Page 38 of 48

(6) Momirlan, M.; Veziroglu, T. The properties of hydrogen as fuel tomorrow in sustainable energy system for a cleaner planet. Int. J. Hydrogen Energy 2005, 30, 795–802. (7) De Beni, G.; Marchetti, C. A Chemical Process to Decompose Water Using Nuclear Heat. 163rd Natl Meet. Am. Chem. Soc., Div. Fuel Chem. 1972, 110–133. (8) Nakamura, T. Hydrogen production from water utilizing solar heat at high temperatures. Sol. Energy 1977, 19, 467–475. (9) Kodama, T. High-temperature solar chemistry for converting solar heat to chemical fuels. Prog. Energy Combust. Sci. 2003, 29, 567–597. ¨ (10) Roeb, M.; Neises, M.; Monnerie, N.; Call, F.; Simon, H.; Sattler, C.; Schmucker, M.; Pitz-Paal, R. Materials-Related Aspects of Thermochemical Water and Carbon Dioxide Splitting: A Review. Materials (Basel). 2012, 5, 2015–2054. (11) Smestad, G. P.; Steinfeld, A. Review: Photochemical and Thermochemical Production of Solar Fuels from H2 O and CO2 Using Metal Oxide Catalysts. Ind. Eng. Chem. Res. 2012, 51, 11828–11840. (12) Wang, Z.; Roberts, R.; Naterer, G.; Gabriel, K. Comparison of thermochemical, electrolytic, photoelectrolytic and photochemical solar-to-hydrogen production technologies. Int. J. Hydrogen Energy 2012, 37, 16287–16301. (13) Romero, M.; Steinfeld, A. Concentrating solar thermal power and thermochemical fuels. Energy Environ. Sci. 2012, 5, 9234–9245. (14) Meredig, B.; Wolverton, C. First-principles thermodynamic framework for the evaluation of thermochemical H2 O- or CO2 -splitting materials. Phys. Rev. B 2009, 80, 245119. (15) Meredig, B.; Wolverton, C. Erratum: First-principles thermodynamic framework for

38

ACS Paragon Plus Environment

Page 39 of 48

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

Chemistry of Materials

the evaluation of thermochemical H2 O- or CO2 -splitting materials. Phys. Rev. B 2011, 83, 239901. (16) Abanades, S.; Charvin, P.; Flamant, G.; Neveu, P. Screening of water-splitting thermochemical cycles potentially attractive for hydrogen production by concentrated solar energy. Energy 2006, 31, 2805–2822. (17) Steinfeld, A. Solar hydrogen production via a two-step water-splitting thermochemical cycle based on Zn/ZnO redox reactions. Int. J. Hydrogen Energy 2002, 27, 611–619. (18) Loutzenhiser, P. G.; Meier, A.; Steinfeld, A. Review of the Two-Step H2 O/CO2 Splitting Solar Thermochemical Cycle Based on Zn/ZnO Redox Reactions. Materials (Basel). 2010, 3, 4922–4938. (19) Abanades, S.; Charvin, P.; Lemont, F.; Flamant, G. Novel two-step SnO2 /SnO watersplitting cycle for solar thermochemical production of hydrogen. Int. J. Hydrogen Energy 2008, 33, 6021–6030. (20) Abanades, S. CO2 and H2 O reduction by solar thermochemical looping using SnO2 /SnO redox reactions: Thermogravimetric analysis. Int. J. Hydrogen Energy 2012, 37, 8223–8231. (21) Otsuka, K.; Hatano, M.; Morikawa, A. Decomposition of water by cerium oxide of δ-phase. Inorganica Chim. Acta 1985, 109, 193–197. (22) Abanades, S.; Flamant, G. Thermochemical hydrogen production from a two-step solar-driven water-splitting cycle based on cerium oxides. Sol. Energy 2006, 80, 1611– 1623. (23) Abanades, S.; Legal, A.; Cordier, A.; Peraudeau, G.; Flamant, G.; Julbe, A. Investigation of reactive cerium-based oxides for H2 production by thermochemical two-step watersplitting. J. Mater. Sci. 2010, 45, 4163–4173. 39

ACS Paragon Plus Environment

Chemistry of Materials

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

Page 40 of 48

(24) Furler, P.; Scheffe, J. R.; Steinfeld, A. Syngas production by simultaneous splitting of H2 O and CO2 via ceria redox reactions in a high-temperature solar reactor. Energy Environ. Sci. 2012, 5, 6098–6103. (25) Chueh, W. C.; Falter, C.; Abbott, M.; Scipio, D.; Furler, P.; Haile, S. M.; Steinfeld, A. High-Flux Solar-Driven Thermochemical Dissociation of CO2 and H2 O Using Nonstoichiometric Ceria. Science 2010, 330, 1797–1801. (26) Le Gal, A.; Abanades, S. Catalytic investigation of ceria-zirconia solid solutions for solar hydrogen production. Int. J. Hydrogen Energy 2011, 36, 4739–4748. (27) McDaniel, A. H.; Miller, E. C.; Arifin, D.; Ambrosini, A.; Coker, E. N.; O’Hayre, R.; Chueh, W. C.; Tong, J. Sr- and Mn-doped LaAlO3−δ for solar thermochemical H2 and CO production. Energy Environ. Sci. 2013, 6, 2424–2428. (28) Scheffe, J. R.; Weibel, D.; Steinfeld, A. LanthanumStrontiumManganese Perovskites as Redox Materials for Solar Thermochemical Splitting of H2 O and CO2 . Energy & Fuels 2013, 27, 4250–4257. (29) Nalbandian, L.; Evdou, A.; Zaspalis, V. La1− x Srx MO3 (M = Mn, Fe) perovskites as materials for thermochemical hydrogen production in conventional and membrane reactors. Int. J. Hydrogen Energy 2009, 34, 7162–7172. (30) McDaniel, A.; Ambrosini, A.; Coker, E.; Miller, J.; Chueh, W.; O’Hayre, R.; Tong, J. Nonstoichiometric Perovskite Oxides for Solar Thermochemical H2 and CO Production. Energy Procedia 2014, 49, 2009–2018. (31) Demont, A.; Abanades, S.; Beche, E. Investigation of Perovskite Structures as OxygenExchange Redox Materials for Hydrogen Production from Thermochemical Two-Step Water-Splitting Cycles. J. Phys. Chem. C 2014, 118, 12682–12692.

40

ACS Paragon Plus Environment

Page 41 of 48

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

Chemistry of Materials

(32) Ishihara, T. Perovskite Oxide for Solid Oxide Fuel Cells; Fuel Cells and Hydrogen Energy; Springer US: Boston, MA, 2009; pp 1–296. (33) Hafner, J.; Wolverton, C.; Ceder, G. Toward Computational Materials Design: The Impact of Density Functional Theory on Materials Research. MRS Bull. 2006, 31, 659–668. (34) Saal, J. E.; Kirklin, S.; Aykol, M.; Meredig, B.; Wolverton, C. Materials Design and Discovery with High-Throughput Density Functional Theory: The Open Quantum Materials Database (OQMD). JOM 2013, 65, 1501–1509. (35) Curtarolo, S.; Setyawan, W.; Wang, S.; Xue, J.; Yang, K.; Taylor, R. H.; Nelson, L. J.; Hart, G. L.; Sanvito, S.; Buongiorno-Nardelli, M.; Mingo, N.; Levy, O. AFLOWLIB.ORG: A distributed materials properties repository from highthroughput ab initio calculations. Comput. Mater. Sci. 2012, 58, 227–235. (36) Curtarolo, S.; Setyawan, W.; Hart, G. L.; Jahnatek, M.; Chepulskii, R. V.; Taylor, R. H.; Wang, S.; Xue, J.; Yang, K.; Levy, O.; Mehl, M. J.; Stokes, H. T.; Demchenko, D. O.; Morgan, D. AFLOW: An automatic framework for high-throughput materials discovery. Comput. Mater. Sci. 2012, 58, 218–226. (37) Jain, A.; Hautier, G.; Moore, C. J.; Ping Ong, S.; Fischer, C. C.; Mueller, T.; Persson, K. A.; Ceder, G. A high-throughput infrastructure for density functional theory calculations. Comput. Mater. Sci. 2011, 50, 2295–2310. (38) Landis, D. D.; Hummelshoj, J. S.; Nestorov, S.; Greeley, J.; Dulak, M.; Bligaard, T.; Norskov, J. K.; Jacobsen, K. W. The Computational Materials Repository. Comput. Sci. Eng. 2012, 14, 51–57. (39) Ceder, G.; Chiang, Y.-M.; Sadoway, D. R.; Aydinol, M. K.; Jang, Y.-I.; Huang, B. Identification of Cathode Materials for Lithium Batteries Guided by First-Principles Calculations. Nature 1998, 392, 694–696. 41

ACS Paragon Plus Environment

Chemistry of Materials

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

Page 42 of 48

(40) Curtarolo, S.; Morgan, D.; Persson, K.; Rodgers, J.; Ceder, G. Predicting Crystal Structures with Data Mining of Quantum Calculations. Phys. Rev. Lett. 2003, 91, 135503. (41) Greeley, J.; Jaramillo, T. F.; Bonde, J.; Chorkendorff, I.; Nørskov, J. K. Computational high-throughput screening of electrocatalytic materials for hydrogen evolution. Nat. Mater. 2006, 5, 909–913. (42) Armiento, R.; Kozinsky, B.; Fornari, M.; Ceder, G. Screening for high-performance piezoelectrics using high-throughput density functional theory. Phys. Rev. B 2011, 84, 014103. (43) Castelli, I. E.; Olsen, T.; Datta, S.; Landis, D. D.; Dahl, S.; Thygesen, K. S.; Jacobsen, K. W. Computational screening of perovskite metal oxides for optimal solar light capture. Energy Environ. Sci. 2012, 5, 5814–5819. (44) Castelli, I. E.; Landis, D. D.; Thygesen, K. S.; Dahl, S.; Chorkendorff, I.; Jaramillo, T. F.; Jacobsen, K. W. New cubic perovskites for one- and two-photon water splitting using the computational materials repository. Energy Environ. Sci. 2012, 5, 9034–9043. (45) Barber, C. B.; Dobkin, D. P.; Huhdanpaa, H. The quickhull algorithm for convex hulls. ACM Trans. Math. Softw. 1996, 22, 469–483. ¨ (46) Kresse, G.; Furthmuller, 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. ¨ (47) Kresse, G.; Furthmuller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. (48) Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmentedwave method. Phys. Rev. B 1999, 59, 1758–1775.

42

ACS Paragon Plus Environment

Page 43 of 48

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

Chemistry of Materials

(49) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. (50) Dudarev, S. L.; Botton, G. A.; Savrasov, S. Y.; Humphreys, C. J.; Sutton, A. P. Electronenergy-loss spectra and the structural stability of nickel oxide: An LSDA+U study. Phys. Rev. B 1998, 57, 1505–1509. (51) Wang, L.; Maxisch, T.; Ceder, G. Oxidation energies of transition metal oxides within the GGA+U framework. Phys. Rev. B 2006, 73, 195107. (52) Lee, Y.-L.; Kleis, J.; Rossmeisl, J.; Morgan, D. Ab initio energetics of LaBO3 (001) (B = Mn , Fe, Co, and Ni) for solid oxide fuel cell cathodes. Phys. Rev. B 2009, 80, 224101. (53) Stevanovi´c, V.; Lany, S.; Zhang, X.; Zunger, A. Correcting density functional theory for accurate predictions of compound enthalpies of formation: Fitted elemental-phase reference energies. Phys. Rev. B 2012, 85, 115104. ¨ (54) Kirklin, S.; Saal, J. E.; Meredig, B.; Thompson, A.; Doak, J. W.; Aykol, M.; Ruhl, S.; Wolverton, C. The Open Quantum Materials Database (OQMD): assessing the accuracy of DFT formation energies. npj Comput. Mater. 2015, 1, 15010. (55) Bergerhoff, G.; Hundt, R.; Sievers, R.; Brown, I. D. The inorganic crystal structure data base. J. Chem. Inf. Model. 1983, 23, 66–69. (56) Belsky, A.; Hellenbrandt, M.; Karen, V. L.; Luksch, P. New developments in the Inorganic Crystal Structure Database (ICSD): accessibility in support of materials research and design. Acta Crystallogr. Sect. B Struct. Sci. 2002, 58, 364–369. (57) Akbarzadeh, A. R.; Ozolins, V.; Wolverton, C. First-Principles Determination of Multicomponent Hydride Phase Diagrams: Application to the Li-Mg-N-H System. Adv. Mater. 2007, 19, 3233–3239.

43

ACS Paragon Plus Environment

Chemistry of Materials

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

Page 44 of 48

(58) Curnan, M. T.; Kitchin, J. R. Effects of Concentration, Crystal Structure, Magnetism, and Electronic Structure Method on First-Principles Oxygen Vacancy Formation Energy Trends in Perovskites. J. Phys. Chem. C 2014, 118, 28776–28790. (59) Deml, A. M.; Stevanovi´c, V.; Holder, A. M.; Sanders, M.; O’Hayre, R.; Musgrave, C. B. Tunable Oxygen Vacancy Formation Energetics in the Complex Perovskite Oxide Srx La1− x Mny Al1−y O3 . Chem. Mater. 2014, 26, 6595–6602. (60) Kuo, J.; Anderson, H.; Sparlin, D. Oxidation-reduction behavior of undoped and Sr-doped LaMnO3 nonstoichiometry and defect structure. J. Solid State Chem. 1989, 83, 52–60. (61) Nowotny, J.; Rekas, M. Defect Chemistry of (La,Sr)MnO3 . J. Am. Ceram. Soc. 2005, 81, 67–80. (62) Mizusaki, J.; Yoshihiro, M.; Yamauchi, S.; Fueki, K. Nonstoichiometry and defect structure of the perovskite-type oxides La1− x Srx FeO3−δ . J. Solid State Chem. 1985, 58, 257–266. (63) Mizusaki, J.; Mima, Y.; Yamauchi, S.; Fueki, K.; Tagawa, H. Nonstoichiometry of the perovskite-type oxides La1− x Srx CoO3−δ . J. Solid State Chem. 1989, 80, 102–111. (64) O’Keefe, M.; Brese, N. E. Atom sizes and bond lengths in molecules and crystals. J. Am. Chem. Soc. 1991, 113, 3226–3229. (65) Ong, S. P.; Richards, W. D.; Jain, A.; Hautier, G.; Kocher, M.; Cholia, S.; Gunter, D.; Chevrier, V. L.; Persson, K. A.; Ceder, G. Python Materials Genomics (pymatgen): A robust, open-source python library for materials analysis. Comput. Mater. Sci. 2013, 68, 314–319. (66) Shannon, R. D.; Prewitt, C. T. Effective ionic radii in oxides and fluorides. Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem. 1969, 25, 925–946. 44

ACS Paragon Plus Environment

Page 45 of 48

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

Chemistry of Materials

(67) Shannon, R. D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. Sect. A 1976, 32, 751–767. (68) Seshadri, R.; Basu, R. Periodic table of the elements. 2002. ¨ S. M.; Marques, M.; Botti, S. Stability and electronic properties of new inor(69) Korbel, ganic perovskites from high-throughput ab initio calculations. J. Mater. Chem. C 2016, 4, 3157–3167. (70) Zhou, J.-S.; Goodenough, J. B. Universal Octahedral-Site Distortion in Orthorhombic Perovskite Oxides. Phys. Rev. Lett. 2005, 94, 065501. (71) Roth, R. Classification of perovskite and other ABO3 -type compounds. J. Res. Natl. Bur. Stand. (1934). 1957, 58, 75–88. (72) Bhalla, A. S.; Guo, R.; Roy, R.; Ruyan, A. S. B.; Rustum, G. The perovskite structure - a review of its role in ceramic science and technology. Mater. Res. Innov. 2000, 4, 3–26. (73) Woodward, P. M. Octahedral Tilting in Perovskites. I. Geometrical Considerations. Acta Crystallogr. Sect. B Struct. Sci. 1997, 53, 32–43. (74) El-Mellouhi, F.; Brothers, E. N.; Lucero, M. J.; Bulik, I. W.; Scuseria, G. E. Structural phase transitions of the metal oxide perovskites SrTiO3 , LaAlO3 , and LaTiO3 studied with a screened hybrid functional. Phys. Rev. B 2013, 87, 035107. (75) Zhou, J.-S.; Alonso, J. A.; Pomjakushin, V.; Goodenough, J. B.; Ren, Y.; Yan, J.-Q.; Cheng, J.-G. Intrinsic structural distortion and superexchange interaction in the orthorhombic rare-earth perovskites RCrO3 . Phys. Rev. B 2010, 81, 214115. (76) Giaquinta, D. M.; zur Loye, H.-C. Structural Predictions in the ABO3 Phase Diagram. Chem. Mater. 1994, 6, 365–372. (77) Li, C.; Soh, K. C. K.; Wu, P. Formability of ABO3 perovskites. J. Alloys Compd. 2004, 372, 40–48. 45

ACS Paragon Plus Environment

Chemistry of Materials

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

Page 46 of 48

(78) Zhang, H.; Li, N.; Li, K.; Xue, D. Structural stability and formability of ABO3 -type perovskite compounds. Acta Crystallogr. Sect. B Struct. Sci. 2007, 63, 812–818. (79) Huan, T. D.; Amsler, M.; Marques, M. A. L.; Botti, S.; Willand, A.; Goedecker, S. Low-Energy Polymeric Phases of Alanates. Phys. Rev. Lett. 2013, 110, 135502. (80) Muller, O.; Roy, R. The Major Ternary Structural Families; Springer- Verlag: New York, Heidelberg, Berlin, 1974; pp 1–487. (81) Kumar, A.; Verma, A. S.; Bhardwaj, S. R. Prediction of Formability in Perovskite-Type Oxides. Open Appl. Phys. J. 2008, 1, 11–19. (82) Goldschmidt, V. M. Die Gesetze der Krystallochemie. Naturwissenschaften 1926, 14, 477–485. (83) Castelli, I. E.; Jacobsen, K. W. Designing rules and probabilistic weighting for fast materials discovery in the Perovskite structure. Model. Simul. Mater. Sci. Eng. 2014, 22, 055007. (84) Chiang, Y.-M.; Lavik, E.; Blom, D. Defect thermodynamics and electrical properties of nanocrystalline oxides: pure and doped CeO2 . Nanostructured Mater. 1997, 9, 633–642. (85) Yang, Z.; Luo, G.; Lu, Z.; Hermansson, K. Oxygen vacancy formation energy in Pd-doped ceria: A DFT+U study. J. Chem. Phys. 2007, 127, 074704. (86) Murgida, G. E.; Ferrari, V.; Ganduglia-Pirovano, M. V.; Llois, A. M. Ordering of oxygen vacancies and excess charge localization in bulk ceria: A DFT+U. Phys. Rev. B 2014, 90, 115120. (87) Feng, Z. a.; El Gabaly, F.; Ye, X.; Shen, Z.-X.; Chueh, W. C. Fast vacancy-mediated oxygen ion incorporation across the ceria-gas electrochemical interface. Nat. Commun. 2014, 5, 4374.

46

ACS Paragon Plus Environment

Page 47 of 48

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

Chemistry of Materials

(88) Hansen, H. A.; Wolverton, C. Kinetics and Thermodynamics of H2 O Dissociation on Reduced CeO2 (111). J. Phys. Chem. C 2014, 118, 27402–27414.

47

ACS Paragon Plus Environment

Chemistry of Materials

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 9: For Table of Contents Only. Tiff version included in the zip file the TOC folder.

48

ACS Paragon Plus Environment

Page 48 of 48