Ab Initio DFT+U Analysis of Oxygen Vacancy Formation and ...€¦ · Ab Initio DFT+U Analysis of...

9
Ab Initio DFT+U Analysis of Oxygen Vacancy Formation and Migration in La 1x Sr x FeO 3δ (x = 0, 0.25, 0.50) Andrew M. Ritzmann, Ana B. Muñ oz-García, Michele Pavone, John A. Keith, § and Emily A. Carter* ,§,Department of Chemical and Biological Engineering, Princeton University, Princeton, New Jersey 08544-5263, United States Department of Chemical Sciences, University of Naples Federico II, Naples 80126, Italy § Department of Mechanical and Aerospace Engineering, Program in Applied and Computational Mathematics, and Andlinger Center for Energy and the Environment, Princeton University, Princeton, New Jersey 08544-5263, United States * S Supporting Information ABSTRACT: Incorporating mixed oxygen-ion-electron conducting (MIEC) cathode materials is a promising strategy to make intermediate-temperature solid oxide fuel cells (IT-SOFCs) viable; however, a lack of fundamental understanding of oxygen transport in these materials limits their development. Density functional theory plus U (DFT+U) calculations are used to investigate how the Sr concentration aects the processes that govern oxygen ion transport in La 1x Sr x FeO 3δ (LSF, x = 0, 0.25, and 0.50). Specically, we show that oxygen vacancies compensate holes introduced by Sr and that this compensation facilitates oxygen vacancy formation in LSF. We also nd that oxygen migration in LaFeO 3 is accompanied by electron transfer in the opposite direction. Our results explicitly identify and clarify the role of electron-decient substitutions in promoting oxygen diusion in LSF. This atomic level insight is important for enabling rational design of iron-based SOFC cathode materials. KEYWORDS: solid-oxide fuel cells, La 1x Sr x FeO 3 , oxygen vacancies, oxygen migration INTRODUCTION Building a sustainable energy future requires clean and ecient technologies based on renewable fuels. 1 Fuel cells promise to produce electrical power cleanly and eciently via electro- chemistry; however, high overpotentials associated with the oxygen reduction reaction (ORR) rate at the cathode currently limit their widespread deployment. Mitigating these over- potentials requires either expensive ORR catalysts (e.g., Pt) or high operating temperatures. Solid oxide fuel cells (SOFCs, which consist of solid state anode, cathode, electrolyte, and interconnect materials) use high operating temperatures (T op > 800 °C) 2 to overcome the ORR overpotential, thus permitting power generation from a variety of fuel streams (e.g., H 2 , CH 4 , and syngas). 35 While high operating temperatures eschew the need for precious metal catalysts, they detrimentally cause premature aging of the SOFC. 6 Decreasing the ORR overpotential in SOFC devices should enable higher perform- ance, lower operating temperatures, and longer lifetimes. Recent work has focused on developing intermediate-temper- ature SOFCs (IT-SOFCs) that operate between 600 and 800 °C. 7 One class of IT-SOFC uses mixed oxygen-ion-electron conducting (MIEC) cathode materials that increase the active region of the cathode, 8 thereby reducing the ORR over- potential. Although all SOFC cathodes must conduct electrons, MIEC cathodes improve upon traditional cathodes (e.g., La 1x Sr x MnO 3 ) by allowing bulk oxygen transport. The success of MIEC cathodes shows that bulk oxygen transport in the cathode determines the viability of IT-SOFCs. While some successful MIEC materials have been proposed recently, 9,10 La 1x Sr x Co 1y Fe y O 3δ (LSCF), an ABO 3 perovskite material with substitutions in both the A- and B-sites, remains the reference MIEC for SOFC cathodes. LSCF typically contains comparable amounts of La and Sr and signicantly more Fe than Co (e.g., La 0.6 Sr 0.4 Co 0.2 Fe 0.8 O 3δ ). 11 At the limit of no cobalt, one obtains the iron-only parent material, La 1x Sr x FeO 3δ (LSF). Fundamental insight into the electronic structure and oxygen transport of LSF will in turn elucidate the behavior of LSCF by isolating the role of A-site substitutions. Because IT-SOFC applications require facile oxygen ion transport, higher oxygen diusion coecients are desirable. In LSF, oxygen diusion occurs via a vacancy-mediated hopping mechanism. 12,13 The resulting self-diusion coecient (D o) has the form = D CD o v v (1) where C v is the concentration of oxygen vacancies and D v is the vacancy diusion coecient. D v takes an Arrhenius form where the activation enthalpy (ΔH mig ) is the energy barrier for an oxygen ion hopping into an unoccupied neighboring site. In LSF, the oxygen vacancies do not interact. 14 The free energy of oxygen vacancy formation (ΔG f,vac ) governs the vacancy concentration. Therefore, the sum of ΔH f,vac and ΔH mig denes Received: April 1, 2013 Revised: June 26, 2013 Published: June 27, 2013 Article pubs.acs.org/cm © 2013 American Chemical Society 3011 dx.doi.org/10.1021/cm401052w | Chem. Mater. 2013, 25, 30113019

Transcript of Ab Initio DFT+U Analysis of Oxygen Vacancy Formation and ...€¦ · Ab Initio DFT+U Analysis of...

Page 1: Ab Initio DFT+U Analysis of Oxygen Vacancy Formation and ...€¦ · Ab Initio DFT+U Analysis of Oxygen Vacancy Formation and Migration in La 1‑xSr xFeO 3‑δ (x = 0, 0.25, 0.50)

Ab Initio DFT+U Analysis of Oxygen Vacancy Formation andMigration in La1‑xSrxFeO3‑δ (x = 0, 0.25, 0.50)Andrew M. Ritzmann,† Ana B. Munoz-García,‡ Michele Pavone,‡ John A. Keith,§ and Emily A. Carter*,§,∥

†Department of Chemical and Biological Engineering, Princeton University, Princeton, New Jersey 08544-5263, United States‡Department of Chemical Sciences, University of Naples Federico II, Naples 80126, Italy§Department of Mechanical and Aerospace Engineering, ∥Program in Applied and Computational Mathematics, and AndlingerCenter for Energy and the Environment, Princeton University, Princeton, New Jersey 08544-5263, United States

*S Supporting Information

ABSTRACT: Incorporating mixed oxygen-ion-electron conducting (MIEC) cathodematerials is a promising strategy to make intermediate-temperature solid oxide fuel cells(IT-SOFCs) viable; however, a lack of fundamental understanding of oxygen transport inthese materials limits their development. Density functional theory plus U (DFT+U)calculations are used to investigate how the Sr concentration affects the processes thatgovern oxygen ion transport in La1‑xSrxFeO3‑δ (LSF, x = 0, 0.25, and 0.50). Specifically, weshow that oxygen vacancies compensate holes introduced by Sr and that this compensationfacilitates oxygen vacancy formation in LSF. We also find that oxygen migration in LaFeO3 isaccompanied by electron transfer in the opposite direction. Our results explicitly identify andclarify the role of electron-deficient substitutions in promoting oxygen diffusion in LSF. Thisatomic level insight is important for enabling rational design of iron-based SOFC cathodematerials.

KEYWORDS: solid-oxide fuel cells, La1‑xSrxFeO3, oxygen vacancies, oxygen migration

■ INTRODUCTION

Building a sustainable energy future requires clean and efficienttechnologies based on renewable fuels.1 Fuel cells promise toproduce electrical power cleanly and efficiently via electro-chemistry; however, high overpotentials associated with theoxygen reduction reaction (ORR) rate at the cathode currentlylimit their widespread deployment. Mitigating these over-potentials requires either expensive ORR catalysts (e.g., Pt) orhigh operating temperatures. Solid oxide fuel cells (SOFCs,which consist of solid state anode, cathode, electrolyte, andinterconnect materials) use high operating temperatures (Top >800 °C)2 to overcome the ORR overpotential, thus permittingpower generation from a variety of fuel streams (e.g., H2, CH4,and syngas).3−5 While high operating temperatures eschew theneed for precious metal catalysts, they detrimentally causepremature aging of the SOFC.6 Decreasing the ORRoverpotential in SOFC devices should enable higher perform-ance, lower operating temperatures, and longer lifetimes.Recent work has focused on developing intermediate-temper-ature SOFCs (IT-SOFCs) that operate between 600 and 800°C.7 One class of IT-SOFC uses mixed oxygen-ion-electronconducting (MIEC) cathode materials that increase the activeregion of the cathode,8 thereby reducing the ORR over-potential. Although all SOFC cathodes must conduct electrons,MIEC cathodes improve upon traditional cathodes (e.g.,La1‑xSrxMnO3) by allowing bulk oxygen transport. The successof MIEC cathodes shows that bulk oxygen transport in thecathode determines the viability of IT-SOFCs.

While some successful MIEC materials have been proposedrecently,9,10 La1‑xSrxCo1‑yFeyO3‑δ (LSCF), an ABO3 perovskitematerial with substitutions in both the A- and B-sites, remainsthe reference MIEC for SOFC cathodes. LSCF typicallycontains comparable amounts of La and Sr and significantlymore Fe than Co (e.g., La0.6Sr0.4Co0.2Fe0.8O3‑δ).

11 At the limitof no cobalt, one obtains the iron-only parent material,La1‑xSrxFeO3‑δ (LSF). Fundamental insight into the electronicstructure and oxygen transport of LSF will in turn elucidate thebehavior of LSCF by isolating the role of A-site substitutions.Because IT-SOFC applications require facile oxygen ion

transport, higher oxygen diffusion coefficients are desirable. InLSF, oxygen diffusion occurs via a vacancy-mediated hoppingmechanism.12,13 The resulting self-diffusion coefficient (Do) hasthe form

=D C Do v v (1)

where Cv is the concentration of oxygen vacancies and Dv is thevacancy diffusion coefficient. Dv takes an Arrhenius form wherethe activation enthalpy (ΔHmig) is the energy barrier for anoxygen ion hopping into an unoccupied neighboring site. InLSF, the oxygen vacancies do not interact.14 The free energy ofoxygen vacancy formation (ΔGf,vac) governs the vacancyconcentration. Therefore, the sum of ΔHf,vac and ΔHmig defines

Received: April 1, 2013Revised: June 26, 2013Published: June 27, 2013

Article

pubs.acs.org/cm

© 2013 American Chemical Society 3011 dx.doi.org/10.1021/cm401052w | Chem. Mater. 2013, 25, 3011−3019

Page 2: Ab Initio DFT+U Analysis of Oxygen Vacancy Formation and ...€¦ · Ab Initio DFT+U Analysis of Oxygen Vacancy Formation and Migration in La 1‑xSr xFeO 3‑δ (x = 0, 0.25, 0.50)

the apparent activation energy for the oxygen diffusioncoefficient where A is the preexponential factor which containsthe entropy of oxygen vacancy formation.

= − Δ +ΔD Ae H H k TO

( )/f,vac mig B (2)

Oxygen diffusion measurements on LSF (for x = 0, 0.1, and0.25) indicate that x does not significantly affect ΔHmig.

13

However, thermogravimetric analysis shows that increasing xsignificantly raises the oxygen vacancy concentration.15

Adding Sr clearly improves oxygen transport in LSF, but aconclusive explanation of the fundamental physics driving thisimprovement remains absent. In addition to vacancy-mediatedoxygen transport, the defect chemistry and electrical con-ductivity of LSF are intimately related. The host material,LaFeO3, is a G-type antiferromagnetic insulator,16 althoughlanthanum vacancies (VLa

/// in Kroger-Vink notation17) lead tolow p-type conductivity at ambient oxygen partial pressure.18

Furthermore, while similar ABO3 materials La1‑xSrxMnO3 andLa1‑xSrxCoO3 become metallic when x > 0.2, LSF remains anantiferromagnetic semiconductor for x < 0.9.19,20 Substitutionaldefects (SrLa

/), lanthanum vacancies (VLa///), oxygen vacancies

(VO••), and electrons/holes play important roles in the defect

chemistry of LSF.14 Large ionic radii suppress interstitial defectswithin the perovskite structure.Computational studies have already provided some micro-

scopic insight into defect chemistry and ionic transport in LSF.Jones and Islam21 studied defect chemistry and ion migration inLSF using classical potentials. However, conclusions drawnfrom simulations using classical, empirically-based potentialsrun the risk of depending strongly on the particularparameterization of the potentials rather than representingthe true physics. By contrast, first-principles quantummechanics calculations based on density functional theory(DFT) can offer unbiased insights by being free from system-dependent parameters. Indeed, several electronic structuredescriptions of LaFeO3 have been provided using DFT.22−26

Unfortunately, these predictions may suffer from self-interaction errors that cause DFT to fail for strongly correlatedmaterials (e.g., first row late transition metal oxides).27 Tomitigate these errors, the DFT+U method28 has been appliedto understanding the electronic structure29−31 and defectchemistry25,26,32,33 of LaFeO3. However, these previous studiesdid not consider either the Sr substitution or oxygen migrationin LaFeO3. While the perspective of Kuklja et al. discounts theDFT+U method for complex oxide materials (e.g., LSCF),25 itsvalidity for LSF remains an open question.Here we present an ab initio DFT+U34 analysis of the

electronic structure and oxygen vacancy formation process inLSF (x = 0, 0.25, and 0.50) and oxygen migration in LaFeO3.Our study clearly shows how electron-deficient A-sitesubstitutions enhance oxygen migration in LSF. This insightis highly relevant to rationally designing iron-based SOFCcathode materials. Since the migration barrier does not changesignificantly with Sr content,13 we use LaFeO3 as a model forinvestigating oxygen migration in LSF. Our study considers allpossible arrangements of La/Sr for a given composition withina 40-atom supercell and considers larger (160-atom) cellsdesigned to reproduce the random distribution of La/Sr in LSF.We show that DFT+U successfully models these materials andthat Sr substitution has a large effect on the oxygen vacancyformation enthalpy, which in turn has a dramatic effect on ionconductivity.

■ METHODS AND COMPUTATIONAL DETAILSWe employed three structural models to study LSF. First, anorthorhombic unit cell of LaFeO3 (Figure 1a) corresponding to the

experimental structure was used to validate the DFT+U model.Second, a pseudocubic 2 × 2 × 2 supercell of the simple perovskiteunit cell containing 40 atoms was used to calculate the free energy ofoxygen vacancy formation and the oxygen migration enthalpy.Previous studies have used the 40-atom cell to study oxygen vacanciesin LaMO3 (M = Cr, Mn, Fe, Co, Ni) perovskite materials.32,33 Weconstrained the lattice vectors of the second cell to remain cubic, butinternal ionic relaxations were allowed. This model successfullycaptures the known cooperative tilting of the FeO6 octahedra (Figure1b). To address the role of Sr substitution, we considered all uniquearrangements of La and Sr within the La sublattice (Figure 1c). Westudy the G-type antiferromagnetic state of LSF which is consistentwith neutron diffraction experiments.35 For validation, we used a 160-atom supercell of LSF to consider the effects of lower oxygen vacancyconcentrations on our predictions. We generated the 160-atomsupercell using the special quasirandom structures method (SQS) todetermine the La/Sr arrangement which best resembles the randomalloy.36 The cell was generated using the mcsqs code in version 3.0 ofthe Alloy Theoretic Automatic Toolkit (ATAT)37−39 by firstenumerating all possible supercells of the LaFeO3 orthorhombic unitcell (ensuring G-type antiferromagnetism in the resulting cells) for aspecified number of atoms. To minimize vacancy−vacancy inter-actions, we screened for supercells which maximized the distance froman oxygen atom to its periodic image. Finally, for the best candidatecells, we used the mcsqs code to find the La/Sr arrangement whichbest reproduced random mixing on the A-site sublattice for x = 0.25and x = 0.50. Both compositions yielded supercells (Figure 1d) withthe same shape and matched the random alloy pair-correlationfunctions to a cutoff of 8.8 Å. The vacancy−vacancy distances were∼7.9 Å in the 40-atom pseudocubic cell and ∼11.25 Å in the SQS cells.Calculations for other x values (e.g., x = 0.125) would requiresignificantly larger supercells and were not undertaken in this study.

Figure 1. a) The LaFeO3 orthorhombic unit cell used for methodvalidation, b) the pseudocubic cell used for modeling LSF, c) allunique La/Sr arrangements for x = 0, 0.25, and 0.50, and d) 160 atomsupercells for x = 0, x = 0.25, and x = 0.50. Structures visualized withVESTA.40

Chemistry of Materials Article

dx.doi.org/10.1021/cm401052w | Chem. Mater. 2013, 25, 3011−30193012

Page 3: Ab Initio DFT+U Analysis of Oxygen Vacancy Formation and ...€¦ · Ab Initio DFT+U Analysis of Oxygen Vacancy Formation and Migration in La 1‑xSr xFeO 3‑δ (x = 0, 0.25, 0.50)

The energy cost for oxygen vacancy formation determines thenonstoichiometry of LSF. We compute this energy by comparing theenergy of the supercell with an oxygen vacancy (a neutral oxygen atomis removed) to the same cell without an oxygen vacancy (referenced tohalf the energy of a gas phase O2 molecule in its triplet ground state).The vacancy concentrations are 1/24 (4.2%, δ = 0.125) and 1/96(1.0%, δ = 0.03125) in the pseudocubic (40-atom) and SQS (160-atom) supercells, respectively. These vacancy concentrations exceedthe experimental range when x = 0 (0.008−0.01%, T = 900−1200 °C)but are reasonable for x = 0.25 and x = 0.50 (1% and 4%,respectively).15 The simplest approximation of the vacancy formationenthalpy uses only electronic energies and neglects thermal andvibrational effects:

Δ = − +E E E E12f,vac defective perfect O2 (3)

Approximating ΔHf,vac at absolute zero includes the zero point energy(ZPE) contributions

ν

ν ν

Δ = Δ +

− +

=

− −

=

−⎜ ⎟⎛⎝

⎞⎠

H E h

h h

(0)12

12

12

12

i

N

i

j

N

j

f,vac f,vac1

3( 1) 3

,defective

1

3 3

perfect O2(4)

where h is Planck’s constant, the ν’s are vibrational frequenciescalculated within the harmonic oscillator (H.O.) approximation, and Nis the number of atoms in the perfect supercell. Approximating ΔHf,vacat finite temperature involves integrating the heat capacities (Cp) fromabsolute zero to the temperature of interest. Cp values for the solids(the sums in eq 5) are calculated using the H.O. approximation forvibrational contributions, while Cp for O2 also includes effects from therigid-rotor and ideal gas approximations for vibrational, rotational, andtranslational contributions.41 Because the PV contributions to enthalpyfor solids should be negligible, we assume that the internal energy ofeach solid is equivalent to its enthalpy, while the enthalpy of gas phaseO2 includes the PV term via its constant pressure heat capacity.

ν

ν

Δ = Δ +−

−−

+ ′ ′

ν

ν

=

− −

=

)

H T Hh

hC T dT

( ) (0)e 1

e 112

( ( )

i

Ni

h k T

j

Nj

h k T

T

f,vac f,vac1

3( 1) 3,defective

/

1

3 3,perfect

/ 0p,O

i

j

,defective B

,perfect B 2

(5)

Calculating the free energy of oxygen vacancy formation (ΔGf,vac)requires the entropy (S) of the solids and O2 as a function oftemperature.

Δ = Δ − ΔG T H T T S T( ) ( ) ( )f,vac f,vac f,vac (6)

Δ = − +S T S T S T S T( ) ( ) ( )12

( )f,vac vib,defective vib,perfect O2 (7)

Gryaznov et al. showed that including vibrational effects of thesolids can significantly impact oxygen vacancy formation free energiesin LSCF.42 We express the vibrational enthalpy and entropy as thesum of contributions from harmonic oscillators with the Γ-pointphonon frequencies of the perfect and defective crystals.43 Wecomputed the vibrational frequencies for the entire 40-atom supercelland found that all frequencies were real except for the three acousticmodes which were essentially zero. Sampling the Brillouin zone (BZ)beyond the Γ-point resulted in unphysical imaginary frequencies insome regions of the BZ. We attribute this issue to approximating theHessian matrix using finite differences of forces and to the constraintthat the lattice vectors remain cubic. To avoid issues arising from theseunphysical imaginary frequencies, we only use the Γ-point phononfrequencies in our calculations. For each composition, we compute thevibrational effects using frequencies calculated in the pseudocubic cellwith the isotropic La/Sr arrangements only (rightmost distributions ofSr shown in Figure 1c).

Electronic energies were obtained from spin-polarized Kohn−ShamDFT44,45 and DFT+U28 calculations with periodic boundaryconditions using the Vienna Ab Initio Simulation Package(VASP)46−48 version 5.2.2. The nuclei and core electrons wererepresented within the frozen-core projector-augmented wave49

approach using standard potentials for Fe (4s23d6), La(5s25p66s25d1), Sr (4s24p65s2), and ‘regular’ O (2s22p4) from theVASP library, where the outer core/valence electrons shown inparentheses are self-consistently optimized.50 Electron exchange andcorrelation was evaluated within the generalized gradient approx-imation (GGA) using the functional of Perdew, Burke, and Ernzerhof(PBE),51 although we also performed validation calculations (videinfra) with the local density approximation (LDA) and the moreexpensive hybrid functional of Heyd, Scuseria, and Ernzerhof (HSE).52

Orthorhombic LaFeO3 calculations required a 4 × 4 × 3 Monkhorst-Pack53 k-point mesh for convergence, while pseudocubic and SQS LSFcalculations required 4 × 4 × 4 and 3 × 3 × 3 Monkhorst-Pack k-pointmeshes, respectively. The planewave basis set was converged at akinetic energy cutoff of 750 eV. Integration over the first BZ usedGaussian smearing (σ = 0.05 eV) during structural relaxations. Thetetrahedron method with Blochl corrections54 for BZ integration wasused after geometry optimization to obtain projected density of states(PDOS), refined energies, and the final charge density for Baderanalysis.55,56 Total energies were converged to within 5 meV/formulaunit with these numerical settings. The Ueff parameter for Fe

3+ (4.3eV) in the DFT+U calculations was derived from unrestrictedHartree−Fock calculations on electrostatically embedded Fe2O3

Table 1. Experimental and Calculateda Lattice Constants (a, b, c), Axial and Equatorial Fe−O Bond Lengths (rz and rxy,Respectively), Axial and Equatorial O−Fe−O Bond Angles (αz and αxy, Respectively), Bulk Modulus (B0), Eigenvalue Gap (Eg),and Fe Magnetic Moment (μFe) for Orthorhombic LaFeO3

property expt LDA LDA+U GGA GGA+U HSEe

a/Å 5.553b 5.352 5.462 5.556 5.572 ----b/Å 5.563b 5.273 5.472 5.653 5.627 ----c/Å 7.862b 7.509 7.711 7.885 7.901 ----rz/Å 2.009b 1.905 1.971 2.025 2.028 ----rxy/Å 2.007b 1.903 1.976 2.046 2.035 ----

2.002b 1.901 1.961 2.024 2.024αz/° 156.32b 160.21 155.99 153.54 153.80 ----αxy/° 157.22b 161.89 157.01 153.72 154.69 ----B0/GPa ----- 206 188 155 157 ----Eg/eV 2.1c metal 2.31 0.75 2.53 3.63μFe/μB 4.6 ± 0.2d 0.78−0.89 4.1 3.7 4.2 4.1

aAll DFT+U calculations were performed with an ab-initio-obtained Ueff = 4.3 eV. bReference 65 (room temperature). cReference 16 (optical gap).dReference 35. eDFT-HSE single point calculation performed at the GGA+U geometry.

Chemistry of Materials Article

dx.doi.org/10.1021/cm401052w | Chem. Mater. 2013, 25, 3011−30193013

Page 4: Ab Initio DFT+U Analysis of Oxygen Vacancy Formation and ...€¦ · Ab Initio DFT+U Analysis of Oxygen Vacancy Formation and Migration in La 1‑xSr xFeO 3‑δ (x = 0, 0.25, 0.50)

clusters in earlier work.34 The transferability of this Ueff value wasrecently validated by applying it in the study of the double-perovskiteMIEC material Sr2FeMoO6.

10,57,58 The energies of defective cells werecalculated at the lattice constants of the perfect cell. We used thelattice vectors obtained at absolute zero because thermal expansionresults in only minor changes in the lattice vectors (∼1%) from lowtemperature to SOFC temperatures.59 The nudged elastic band (NEB)method60 was used to determine the minimum energy pathway foroxygen migration. The transition state for oxygen migration in LaFeO3was found using the climbing-image NEB (CI-NEB) method61,62 asimplemented in the VASP transition state theory tools from theUniversity of Texas.63,64

■ RESULTS AND DISCUSSION

Bulk Properties of LSF. We first determine the mostappropriate electronic structure method for treating LaFeO3using the orthorhombic unit cell (Figure 1a). Table 1 comparesstructural, electronic, and magnetic properties for antiferro-magnetic LaFeO3 calculated with various DFT, DFT+U, andhybrid DFT functionals. Notably, pure DFT-LDA incorrectlypredicts a metallic structure for LaFeO3, while pure DFT-GGApredicts a gap for LaFeO3. However, pure DFT-GGAsignificantly underestimates the gap and Fe magnetic moments.These errors likely arise from electron self-interaction inherentin pure density functionals that can be ameliorated with DFT+U; indeed, LDA+U and GGA+U give more accurate gaps andmagnetic moments. For completeness, we tested the moreexpensive HSE hybrid functional with a calculation at the GGA+U geometry. The magnetic moment from the HSE calculationagreed with the DFT+U calculation; however, HSE gave aneigenvalue gap far in excess of the experimental optical gap.From this, we conclude that DFT+U theory provides asufficient description of LaFeO3. Compared to LDA+U, GGA+U results give better lattice vectors and Fe−O bond lengths,all of which deviate from experiment by no more than 1.6%. Onthis basis, we conclude that GGA+U calculations provide thebest model for LaFeO3, and we use this level of theory for therest of our calculations. Previously, Shein et al.31 reported abulk modulus of ∼198 GPa for LaFeO3 calculated with LDA+Uwith Ueff = 5.4 eV chosen to reproduce the experimental bandgap and magnetic moment in LaFeO3.

30 Our LDA+U value of188 GPa, calculated with the ab initio Ueff = 4.3 eV, agrees wellwith their calculations, while our GGA+U value was 157 GPa.We expect the actual bulk modulus lies between these values,but, to the best of our knowledge, experimental measurementsare not available to confirm this expectation.The structural behavior of LSF under increased Sr content

may appear counterintuitive. Sr2+ has a larger ionic radius (1.44Å) than La3+ (1.36 Å),66 but when Sr2+ substitutes for La3+, theunit cell volume actually decreases (Figure 2a). Shorter Fe−Obonds accompany this decrease (Figure 2b), while Fe−O−Febond angles increase toward 180° (Figure 2c). This trendreproduces the experimental trends found in LSF,67 but theunderlying cause only becomes clear upon examination of theelectronic structure discussed next.As alluded to earlier, Sr substitution introduces holes into the

lattice of LSF to maintain charge neutrality in the crystal. Wedetermine the nature of the hole states that form upon Sr-substitution in LSF by analyzing its electronic structure, asrepresented by its PDOS (Figure 3). In LaFeO3, a clear gap isvisible above the Fermi level (Figure 3a). The valence bandmaximum consists of strongly mixed Fe 3d and O 2p states inthe majority channel and mostly O 2p states in the minoritychannel, consistent with nearly-closed-shell oxygen anions.

Figure 2. Structural parameters for the pseudocubic cell from GGA+Ucalculations: a) cell volume (Å3), b) Fe−O bond length (Å), and c)Fe−O−Fe bond angle (°) as a function of strontium content.

Figure 3. GGA+U PDOS for Fe 3d (black) and O 2p (red) states inLSF. a) x = 0, b) x = 0.25, and c) x = 0.50 in the pseudocubic (40-atom) cell. Positive values represent majority spin states while negativevalues represent minority spin states.

Chemistry of Materials Article

dx.doi.org/10.1021/cm401052w | Chem. Mater. 2013, 25, 3011−30193014

Page 5: Ab Initio DFT+U Analysis of Oxygen Vacancy Formation and ...€¦ · Ab Initio DFT+U Analysis of Oxygen Vacancy Formation and Migration in La 1‑xSr xFeO 3‑δ (x = 0, 0.25, 0.50)

Minority Fe 3d states dominate the conduction band minimum.Thus, we find that LaFeO3 is a ligand-to-metal charge transferinsulator where optical excitations involve transitions fromoccupied minority O 2p states to unoccupied minority Fe 3dstates. Our observation agrees with optical reflectance16 andsoft X-ray absorption68 experiments, lending credence to theuse of GGA+U to understand LSF’s electronic structure. For x= 0.25 (Figure 3b) and x = 0.50 (Figure 3c), the hole statesintroduced by Sr substitution appear just above the Fermi level.While the valence band now crosses the Fermi level, the gapremains clear. Thus, the GGA+U method correctly predicts p-type semiconductivity for these compositions. Furthermore, thePDOS indicate the hole states have mixed Fe 3d and O 2pcharacter and hence are at least somewhat delocalized (videinf ra), in contrast to the assumption of hopping type, localizedcarriers that was employed in analyzing the conductivity andSeebeck coefficients of LSF.14

Besides affecting the DOS, Sr substitution also affects themagnetic moment and charge on each atom (Table 2).Increasing Sr-content decreases magnetic moments of acorresponding number of Fe ions; holes therefore appear toaffect the Fe sublattice. However, with increasing Sr content,the Fe and La Bader charges55 remain essentially unchanged.The Bader charge on O becomes less negative by only 0.04electrons per O atom for each 0.25 increase in Sr mole fraction.Surprisingly, this small difference, when distributed over theentire oxygen sublattice, accounts for nearly all of the electronslost due to holes arising from Sr substitution, indicating holedelocalization on the oxygen sublattice. Moreover, the fact thatthe charge on Fe does not change but the Fe magnetic

moments are reduced indicates increased Fe−O bondcovalency upon Sr substitution rather than hole localizationon Fe as had been suggested previously.14,69,70 Our resultsagree with the 2005 study of Patrakeev, et al., which found noactivation energy for carrier mobility in LSF (x ≤ 0.5)indicating band-like semiconducting behavior.71

Oxygen Vacancy Formation in LSF. Now we describefundamental characteristics of oxygen vacancy formation inLSF. Each La/Sr arrangement shown in Figure 1c has severalpotential oxygen vacancy sites. Two Fe ions and four La/Srions coordinate each oxygen atom in LSF. We calculatedvacancy formation energies for removing an oxygen ion in eachLa/Sr environment present for each unit cell shown in Figure1c, leading to one, seven, and fifteen unique oxygen vacancycalculations for x = 0, x = 0.25, and x = 0.50, respectively.Vacancy formation energies (Table 3) are calculated using eqs3−7. The vibrational effects significantly impact the free energyof oxygen vacancy formation, but they do not alter the generaltrends. Therefore, we use the oxygen vacancy formationenthalpy at 700 °C (eq 4) for qualitative analysis and directcomparison to the experimentally-derived enthalpy. From x = 0to x = 0.25, our results show a steep decline in oxygen vacancyformation energy, which agrees with experimental measure-ments (T = 600−1200 °C) in this range (Figure 4).15 Since ourGGA+U results slightly overestimate the equilibrium latticeconstants (compared to experiment), our calculation alreadyaccounts for the small (∼1%) elongation of the lattice vectorsdue to thermal expansion.59 From x = 0.25 to x = 0.5, theoxygen vacancy formation enthalpy decreases in the pseudo-cubic cell, while the remaining SQS cells and experiment show

Table 2. Magnetic Moments (μB) and Bader Charges (e) for LSF with x = 0, x = 0.25, and x = 0.50 with and without an OxygenVacancy (VO

••) in the Pseudocubic (40-Atom) Cella,b,c

x = 0 x = 0.25d x = 0.50d

without VO•• with VO

•• without VO•• with VO

•• without VO•• with VO

••

μFe 4.16 4.17 ± 0.01 4.07 ± 0.12 4.16 4.11 ± 0.03 (×4)e 4.16 ± 0.013.73 ± 0.09 (×4)e

μFe* ---- 3.59 ---- 4.08 ---- 3.46qFe 1.70 1.69 ± 0.01 1.70 1.70 1.70 1.68 ± 0.03qFe* ---- 1.25 ---- 1.60 ---- 1.62qLa 2.08 2.06 ± 0.02 2.09 2.07 ± 0.01 2.10 2.08qSr ---- ---- 1.59 1.58 ± 0.01 1.59 1.58qO −1.26 −1.27 ± 0.01 −1.22 ± 0.01 −1.26 ± 0.01 −1.18 ± 0.03 −1.22 ± 0.04

aUncertainties give the range of calculated values (for different defect configurations) relative to the calculated mean. bFe* indicates the iron ionsadjacent to the oxygen vacancy. cVO

•• (in Kroger-Vink notation) denotes an oxygen vacancy where a neutral oxygen atom has been removed.dValues are reported for the isotropic La/Sr distributions when x = 0.25 and x = 0.50 (rightmost distributions in Figure 1c). e‘×4’ indicates four Feions (out of eight) within the supercell have this magnetic moment

Table 3. Average GGA+U Oxygen Vacancy Formation Energies (in eV) for LSF with x = 0, 0.25, and 0.50 in the Pseudocubic(40-Atom) and SQS (160-Atom) Supercellsa

x ΔEf,vac ΔHf,vac(0) ΔHf,vac(700 °C) ΔGf,vac(700 °C)

Pseudocubic Cell (40 Atoms)a

0 4.05 4.01 3.99 3.040.25 0.71 ± 0.17 0.82 ± 0.17 0.69 ± 0.17 0.14 ± 0.170.50 0.02 ± 0.28 0.09 ± 0.28 −0.01 ± 0.28 −0.67 ± 0.28

SQS Cell (160 Atoms)b

0.0 4.06 4.02 4.00 3.060.25 0.14 0.25 0.12 −0.440.50 0.31 0.38 0.28 −0.39

aUncertainties give the range of calculated values (for different defect configurations) relative to the calculated mean. bVibrational contributions areassumed to be the same as those found in the pseudocubic cell.

Chemistry of Materials Article

dx.doi.org/10.1021/cm401052w | Chem. Mater. 2013, 25, 3011−30193015

Page 6: Ab Initio DFT+U Analysis of Oxygen Vacancy Formation and ...€¦ · Ab Initio DFT+U Analysis of Oxygen Vacancy Formation and Migration in La 1‑xSr xFeO 3‑δ (x = 0, 0.25, 0.50)

relatively constant values in this range. Below, we discuss whythe pseudocubic and SQS models differ.Adding Sr initially causes a large drop in ΔHf,vac leading to

improved bulk oxygen transport. Bader charges and magneticmoments in LSF with and without an oxygen vacancy (Table2) help explain this decrease. When x = 0, creating an oxygenvacancy leads to reduced magnetic moments and Bader chargeson the two Fe ions adjacent to the vacancy. These results showthat the two electrons from the oxygen vacancy reduce theadjacent Fe ions (from Fe3+ to Fe2+) by localizing in their dmanifold. Reducing high spin Fe3+ to Fe2+ requires that theadded electron occupy a minority spin d orbital. As no otherminority spin d electrons are present on the Fe ion, no gain ind-d exchange energy exists to offset the additional Coulombrepulsion associated with adding another electron to the dmanifold, giving rise to the high vacancy formation energy.33

The presence of Sr significantly alters oxygen vacancyformation. Little electron density localizes on the Fe sitesadjacent to the vacancy in x = 0.25; instead, the average chargeon the oxygen ions becomes more negative, indicating thatcreating the vacancy provides electrons which disperse acrossthe oxygen sublattice. As discussed earlier, the holes introducedby Sr substitution withdraw charge from the oxygen sublattice.Creating a vacancy generates electrons, and these electronscompensate the holes arising from Sr substitution. The Fe ionremains in the favorable +3 oxidation state that has a half-filledd-shell, thereby maximizing the d-d exchange stabilization withno additional Coulomb repulsion as suffered in LaFeO3 vacancyformation. The electronegative oxygen ions regain the chargethey lost when Sr substituted for La. Similar to the x = 0.25LSF, the oxygen vacancies in x = 0.50 LSF provide electronswhich compensate some of the holes residing on the oxygen

sublattice. The process of creating an oxygen vacancy isessentially the same in x = 0.25 and x = 0.50, involvingfavorable charge compensation and no added Coulombrepulsion on Fe ions, which leads to very low oxygen vacancyformation energies. Our results for La0.5Sr0.5FeO3 differsignificantly from the DFT-GGA results of Mastrikov et al.26

who found that vacancy formation leads to charge localizationon the Fe ions adjacent to the vacancy (as we found forLaFeO3). This difference leads to a large disagreement betweenthe computed oxygen vacancy formation enthalpies (0.28 eV inthis work vs 3.39−3.60 eV from ref 26). Our results show fargreater agreement with experiment than the DFT-GGA results.Our improved agreement with experiment likely arises fromusing computational methods that more accurately representthe La1‑xSrxFeO3 family of materials (specif ically, using GGA+Umethod, treating the antiferromagnetic order appropriately, andusing the ‘regular’ instead of ‘sof t’ PAW potential for oxygen).Recall that our pseudocubic model predicts a decrease in

ΔHf,vac from x = 0.25 to x = 0.50 (Figure 4) which differs fromexperiment. In the previous paragraph, we showed that chargecompensation occurs in both cases and facilitates oxygenvacancy formation. However, the ratio of vacancies to holes(VO

··/SrLa/) differs in each case. Specifically, when x = 0.25,

VO··/SrLa

/ = 0.5 in the pseudocubic model, which is the limit forcharge compensation (two holes compensating one vacancy).When x = 0.50, additional uncompensated holes are present inthe pseudocubic supercell so VO

··/SrLa/ is less than 0.5.

Increasing the size of our supercell model lowers the oxygenvacancy formation energy for x = 0.25 while leaving x = 0.0 andx = 0.50 basically untouched because the x = 0.25 calculationno longer occurs at the charge compensation limit. Thus, theΔHf,vac values for x = 0.25 and x = 0.50 are similar. The narrowscatter in our 40-atom oxygen vacancy formation enthalpies(obtained for various local environments and for different La/Sr arrangements) indicates that the multiple minimum problemsometimes encountered for DFT+U72 is not a likely source oferror in our calculations. This does not explain why our ΔHf,vacvalues at 700 °C in the 160-atom SQS cell are 1.1, 0.88, and0.72 eV below the experimental values, for x = 0.00, x = 0.25,and x = 0.50, respectively. We attribute this difference primarilyto the known overbinding of O2 by DFT-GGA, which wouldshift our values upward by at least 0.42 eV (half of thedifference between the bond dissociation energy, D0, calculatedwith DFT-GGA and determined experimentally), in betteragreement with experiment. While we have used ΔHf,vac in ourpreceding discussion, a plot of ΔGf,vac as a function oftemperature for each composition is included for reference inthe Supporting Information.

Oxygen Migration in LSF. Oxygen migration will lead tolong-range diffusion once a significant concentration of oxygenvacancies exists in LSF. Since experimental evidence indicatesthat the energy barrier does not depend strongly on Srcontent,13 we use x = 0 (LaFeO3) to model oxygen migrationin LSF. Admittedly, the independence of the migration barrierwith respect to Sr content is at best a loose relationship sincesignificantly higher migration barriers were measured forLa0.6Sr0.4FeO3 (1.58−2.21 eV depending on PO2

)73 and(La0.6Sr0.4)0.99FeO3 (1.34 eV).74 LaFeO3 remains suitable fordelving into the migration process; however, it represents onlyan approximation to the migration barrier for other Srcompositions. We confirm the curved migration pathway(Figure 5a) previously predicted by Jones and Islam using

Figure 4. GGA+U vacancy formation enthalpy at 700 °C in LSF as afunction of Sr content. Experimental results from ref 15, DFT-GGAresults from ref 26, average 40-atom results (gray line), all 40-atomresults (gray circles), and 160-atom results (red line and squares).Even better agreement between our computed oxygen vacancyformation enthalpies and experimental values would be obtainedwith a correction for the known overbinding of the O2 molecule byDFT-GGA (vide inf ra). The small scatter in our 40-atom resultsindicates that the local chemical environment has a minimal effect onthe oxygen vacancy formation enthalpy. [a]Experimental results fromref 15 are the reported reaction enthalpies determined over the rangeof 600 °C-1200 °C for 2Fe3+ + O2‑ → 2Fe2+ + VO

·· + 1/2O2 (x = 0)and 2Fe4+ + O2‑ → 2Fe3+ + VO

·· + 1/2O2 (x = 0.25, 0.50) in LSF.Because they are determined from a fit using the van’t Hoff equation,the experimental values are temperature-independent. Experimentaluncertainties represent 2σ. [b]DFT-GGA results from Mastrikov et al.(ref 26) obtained using the 40-atom pseudocubic supercell withferromagnetic ordering.

Chemistry of Materials Article

dx.doi.org/10.1021/cm401052w | Chem. Mater. 2013, 25, 3011−30193016

Page 7: Ab Initio DFT+U Analysis of Oxygen Vacancy Formation and ...€¦ · Ab Initio DFT+U Analysis of Oxygen Vacancy Formation and Migration in La 1‑xSr xFeO 3‑δ (x = 0, 0.25, 0.50)

classical potentials.21 Our calculated oxygen migration barrier(ΔHmig with x = 0) is 0.79 eV (Figure 5b), agreeing nearlyperfectly with the measured migration barrier of 0.77 eVobtained from oxygen tracer diffusion measurements13 whilefalling below the value of 1.10 eV obtained from conductivityrelaxation experiments.75 Our study attempts to model the self-diffusion coefficient (DO) for LaFeO3 which most closelyresembles the tracer diffusion coefficient (DO*) and is furtherfrom the chemical diffusion (Dchem) coefficient obtained fromconductivity relaxation experiments. Therefore, our calculatedmigration barrier is in good agreement with experiment whencompared against the experimental results of greatest similarity.Our results agree well with Mastrikov et al.’s reported migrationbarrier of 0.75 eV;26 however, their DFT-GGA analysis utilizedferromagnetic LaFeO3 and the less-accurate ‘soft’ oxygen PAWpotential available in VASP. The similarity in the computedbarriers, in spite of several methodological differences indicatesthat steric/electrostatic effects play a dominant role indetermining the migration barrier. While ref 26 indicatesnegligible charge transfer from the migrating oxygen ion to theFe ion at the center of the migration pathway, we find adifferent charge transfer accompanies oxygen diffusion inLaFeO3. As mentioned earlier, oxygen migration involveshopping from an occupied site into an adjacent vacant site.These sites are each coordinated to two Fe ions, and one ofthese Fe ions is adjacent to both the initial and final sites. Themagnetic moment of the Fe ion proximal to the migration (FeA,which is coordinated to the vacancy in both the initial and finalstates) remains low (relative to the bulk μFe) throughout themigration process. The two other Fe ions are coordinated tothe vacant site in the initial state (FeB) or the final state (FeC).Before migration, FeA and FeB have lower magnetic momentsarising from the adjacent oxygen vacancy, while FeC shows abulk-like magnetic moment. After migration, FeA and FeC havelower magnetic moments, while FeB now has a bulklikemagnetic moment. During migration, there is an abrupt switchin the magnetic moments of FeB and FeC corresponding to anelectron transferring from FeB to FeC. An ionic picture explainsthe origin of this electron transfer. The vacancy lies betweenFeA and FeB in the initial state and FeA and FeC in the finalstate. In each state, the ions adjacent to the vacancy are in the+2 oxidation state. Thus, FeA remains Fe2+ throughout themigration process. FeB starts as Fe2+ and becomes Fe3+, whileFeC undergoes the reverse transformation. Since the totalnumber of electrons in the cell remains unchanged, an electronmust transfer from FeB to FeC during the migration process.

The potential energy surface along the migration pathway isslightly asymmetric because the electron transfer happens afterthe transition state. This means that the transition state moreclosely resembles the ξ = 0.25 image, while the ξ = 0.75 imagehas a different magnetic structure. Since the ξ = 0.75 image islinked to the transition state by a spring, it is pulled to a higherenergy on account of the abrupt magnetic transition thatoccurs. While this explanation holds when x = 0, the physics ofoxygen vacancy formation changes in the presence of Sr. Asdiscussed previously, the electrons arising from creating theoxygen vacancy delocalize on the O sublattice when x > 0.Therefore, this coupled electron-ion motion in oppositedirections should disappear with the addition of Sr. Breakingthis coupling should have a minimal effect on the oxygenmigration barrier, which is controlled by the movement of ions,not electrons.

■ CONCLUSIONS

We have presented a detailed first-principles quantummechanics analysis showing how Sr substitution influencesoxygen transport in LSF that provides a deeper understandingof oxygen transport in LSF cathodes. Specifically, we clarify thefundamental role of A-site dopants in promoting oxygenvacancy formation and therefore oxide ion transport in IT-SOFCs with MIEC cathodes. After showing the DFT+Umethod’s capability of successfully modeling LSF, wedemonstrated that the holes introduced by Sr substitutionhave mainly O 2p character. A key finding is that oxygenvacancy formation in LaFeO3 involves reducing the Fe ionsadjacent to the vacant site. This reduction is unfavorable andtherefore ΔHf,vac is large in LaFeO3. With Sr present, theelectrons left behind when a vacancy forms compensate for theholes introduced by Sr. The resulting ΔHf,vac is significantlylower than that in LaFeO3, resulting in more oxygen vacanciesand easier oxygen transport. We showed that oxygen diffusionin LaFeO3 is also intrinsically coupled to an electron transferbut the electronic structure of LSF suggests that this couplingwill disappear for ion transport in LSF. GGA+U calculationsindicate significant increases in oxygen diffusivity will occurwith increasing Sr concentration because of the correspondingincrease in oxygen vacancies.This refined illustration provided by the DFT+U method

shows how the interplay between electron-deficient substitu-tions, iron ions, and the oxygen sublattice can improve oxygentransport in LSF. As found in Sr2FeMoO6 based materi-als,10,57,58 the electrons from an oxygen vacancy in LSF (x > 0)

Figure 5. (a) Minimum energy pathway for oxygen migration in LaFeO3, (b) energy, and (c) Fe magnetic moments along the minimum energypathway at the GGA+U level of theory.

Chemistry of Materials Article

dx.doi.org/10.1021/cm401052w | Chem. Mater. 2013, 25, 3011−30193017

Page 8: Ab Initio DFT+U Analysis of Oxygen Vacancy Formation and ...€¦ · Ab Initio DFT+U Analysis of Oxygen Vacancy Formation and Migration in La 1‑xSr xFeO 3‑δ (x = 0, 0.25, 0.50)

distribute over the oxygen sublattice which produces veryfavorable oxygen vacancy formation enthalpies. This insight haswide potential application to iron-based SOFC cathodematerials, but this should still be confirmed by experiment.Further theoretical studies will address how substitutions in thetransition metal site (B-site doping) can further improveoxygen transport in LSF-based SOFC cathodes.

■ ASSOCIATED CONTENT*S Supporting InformationO2 thermodynamics from DFT-GGA results. Free energy ofoxygen vacancy formation (ΔGf,vac) as a function of temper-ature for x = 0, 0.25, 0.50. Optimized crystal structures for thepseudocubic and SQS supercells for x = 0, 0.25, 0.50. Thismaterial is available free of charge via the Internet at http://pubs.acs.org.

■ AUTHOR INFORMATIONCorresponding Author*E-mail: [email protected] ContributionsAll authors have given approval to the final version of themanuscript.NotesThe authors declare no competing financial interest.

■ ACKNOWLEDGMENTSThe simulations presented in this article were performed (inpart) on computational resources supported by the PrincetonInstitute for Computational Science and Engineering (PICS-ciE) and the Office of Information Technology's HighPerformance Computing Center at Princeton University. Partof this research used resources of the National Energy ResearchScientific Computing Center, which is supported by the Officeof Science of the U.S. Department of Energy under ContractNo. DE-AC02-05CH11231. We gratefully acknowledge thecomputing resources provided on “Fusion,” a 320-nodecomputing cluster operated by the Laboratory ComputingResource Center at Argonne National Laboratory. Researchwas carried out in in part at the Center for FunctionalNanomaterials, Brookhaven National Laboratory, which issupported by the U.S. Department of Energy, Office of BasicEnergy Sciences, under Contract No. DE-AC02-98CH10886.HeteroFoaM, an Energy Frontier Research Center funded bythe U.S. Department of Energy, Office of Science, Office ofBasic Energy Sciences, under award DE-SC0001061, supportedthis work.

■ ABBREVIATIONSSOFC, solid oxide fuel cell; MIEC, mixed ion-electronconductor; LSCF, La1‑xSrxCo1‑yFeyO3; LSF, La1‑xSrxFeO3

■ REFERENCES(1) Fanchi, J. R. Energy: Technology and Directions for the Future;Elsevier Academic Press: Burlington, MA, 2004.(2) Adler, S. B. Chem. Rev. 2004, 104, 4791−4844.(3) Gur, T. M.; Homel, M.; Virkar, A. V. J. Power Sources 2010, 195,1085−1090.(4) Homel, M.; Gur, T. M.; Koh, J. H.; Virkar, A. V. J. Power Sources2010, 195, 6367−6372.(5) Murray, E. P.; Tsai, T.; Barnett, S. A. Nature 1999, 400, 649−651.(6) Steele, B. C. H.; Heinzel, A. Nature 2001, 414, 345−352.

(7) Arregui, A.; Rodriguez-Martinez, L. M.; Modena, S.; Bertoldi, M.;van Herle, J.; Sglavo, V. M. Electrochim. Acta 2011, 58, 312−321.(8) Lu, Z.; Hardy, J.; Templeton, J.; Stevenson, J. J. Power Sources2012, 198, 90−94.(9) Mastrikov, Y. A.; Kuklja, M. M.; Kotomin, E. A.; Maier, J. EnergyEnviron. Sci. 2010, 3, 1544−1550.(10) Munoz-García, A. B.; Bugaris, D. E.; Pavone, M.; Hodges, J. P.;Huq, A.; Chen, F.; zur Loye, H.-C.; Carter, E. A. J. Am. Chem. Soc.2012, 134, 6826−6833.(11) Marinha, D.; Dessemond, L.; Djurado, E. J. Power Sources 2012,197, 80−87.(12) Ishigaki, T.; Yamauchi, S.; Mizusaki, J.; Fueki, K.; Naito, H.;Adachi, T. J. Solid State Chem. 1984, 55, 50−53.(13) Ishigaki, T.; Yamauchi, S.; Kishio, K.; Mizusaki, J.; Fueki, K. J.Solid State Chem. 1988, 73, 179−187.(14) Mizusaki, J.; Sasamoto, T.; Cannon, W. R.; Bowen, H. K. J. Am.Ceram. Soc. 1983, 66, 247−252.(15) Mizusaki, J.; Yoshihiro, M.; Yamauchi, S.; Fueki, K. J. Solid StateChem. 1985, 58, 257−266.(16) Arima, T.; Tokura, Y.; Torrance, J. B. Phys. Rev. B 1993, 48,17006−17009.(17) Kroger, F. A.; Vink, H. J. In Solid State Physics; Seitz, F.,Turnbull, D., Eds.; Academic Press: 1956; Vol. 3, pp 307−435.(18) Mizusaki, J.; Sasamoto, T.; Cannon, W. R.; Bowen, H. K. J. Am.Ceram. Soc. 1982, 65, 363−368.(19) Chainani, A.; Mathew, M.; Sarma, D. D. Phys. Rev. B 1993, 48,14818−14825.(20) Matsuno, J.; Mizokawa, T.; Fujimori, A.; Mamiya, K.; Takeda,Y.; Kawasaki, S.; Takano, M. Phys. Rev. B 1999, 60, 4605−4608.(21) Jones, A.; Islam, M. S. J. Phys. Chem. C 2008, 112, 4455−4462.(22) Sarma, D. D.; Shanthi, N.; Barman, S. R.; Hamada, N.; Sawada,H.; Terakura, K. Phys. Rev. Lett. 1995, 75, 1126−1129.(23) Mahadevan, P.; Shanthi, N.; Sarma, D. D. J. Phys.: Condens.Matter 1997, 9, 3129−3138.(24) Iglesias, M.; Rodríguez, A.; Blaha, P.; Pardo, V.; Baldomir, D.;Pereiro, M.; Botana, J.; Arias, J. E.; Schwarz, K. J. Magn. Magn. Mater.2005, 290−291, 396−399.(25) Kuklja, M. M.; Kotomin, E. A.; Merkle, R.; Mastrikov, Y. A.;Maier, J. Phys. Chem. Chem. Phys. 2013, 15, 5443−5471.(26) Mastrikov, Y. A.; Merkle, R.; Kotomin, E. A.; Kuklja, M. M.;Maier, J. Phys. Chem. Chem. Phys. 2013, 15, 911−918.(27) Terakura, K.; Oguchi, T.; Williams, A. R.; Kubler, J. Phys. Rev. B1984, 30, 4734−4747.(28) Anisimov, V. I.; Zaanen, J.; Andersen, O. K. Phys. Rev. B 1991,44, 943−954.(29) Solovyev, I.; Hamada, N.; Terakura, K. Phys. Rev. B 1996, 53,7158−7170.(30) Yang, Z.; Huang, Z.; Ye, L.; Xie, X. Phys. Rev. B 1999, 60,15674−15682.(31) Shein, I.; Shein, K.; Kozhevnikov, V.; Ivanovskii, A. Phys. SolidState 2005, 47, 2082−2088.(32) Lee, Y.-L.; Kleis, J.; Rossmeisl, J.; Morgan, D. Phys. Rev. B 2009,80, 224101.(33) Pavone, M.; Ritzmann, A. M.; Carter, E. A. Energy Environ. Sci.2011, 4, 4933−4937.(34) Mosey, N. J.; Liao, P.; Carter, E. A. J. Chem. Phys. 2008, 129,014103.(35) Koehler, W. C.; Wollan, E. O. J. Phys. Chem. Solids 1957, 2,100−106.(36) Zunger, A.; Wei, S.-H.; Ferreira, L. G.; Bernard, J. E. Phys. Rev.Lett. 1990, 65, 353−356.(37) Van de Walle, A.; Asta, M.; Ceder, G. Calphad 2002, 26, 539−553.(38) Van de Walle, A.; Ceder, G. JPE 2002, 23, 348−359.(39) Van de Walle, A. Calphad 2009, 33, 266−278.(40) Momma, K.; Izumi, F. J. Appl. Crystallogr. 2008, 41, 653−658.(41) McQuarrie, D. A.; Simon, J. D. Physical Chemistry: A MolecularApproach, 1st ed.; University Science Books: Sausalito, CA, 1997.

Chemistry of Materials Article

dx.doi.org/10.1021/cm401052w | Chem. Mater. 2013, 25, 3011−30193018

Page 9: Ab Initio DFT+U Analysis of Oxygen Vacancy Formation and ...€¦ · Ab Initio DFT+U Analysis of Oxygen Vacancy Formation and Migration in La 1‑xSr xFeO 3‑δ (x = 0, 0.25, 0.50)

(42) Gryaznov, D.; Finnis, M. W.; Evarestov, R.; Maier, J. 2012,arXiv:1206.6242.(43) Denbigh, K. G. The Principles of Chemical Equilibrium: WithApplications in Chemistry and Chemical Engineering, 4th ed.; CambridgeUniversity Press: 1981.(44) Hohenberg, P.; Kohn, W. Phys. Rev. 1964, 136, B864−B871.(45) Kohn, W.; Sham, L. J. Phys. Rev. 1965, 140, A1133−A1138.(46) Kresse, G.; Hafner, J. Phys. Rev. B 1993, 48, 13115−13118.(47) Kresse, G.; Furthmuller, J. Phys. Rev. B 1996, 54, 11169−11186.(48) Kresse, G.; Furthmuller, J. Comput. Mater. Sci. 1996, 6, 15−50.(49) Blochl, P. E. Phys. Rev. B 1994, 50, 17953−17979.(50) Kresse, G.; Joubert, D. Phys. Rev. B 1999, 59, 1758−1775.(51) Perdew, J. P.; Burke, K.; Ernzerhof, M. Phys. Rev. Lett. 1996, 77,3865−3868.(52) Heyd, J.; Scuseria, G. E.; Ernzerhof, M. J. Chem. Phys. 2003, 118,8207−8215.(53) Monkhorst, H. J.; Pack, J. D. Phys. Rev. B 1976, 13, 5188−5192.(54) Blochl, P. E.; Jepsen, O.; Andersen, O. K. Phys. Rev. B 1994, 49,16223−16233.(55) Bader, R. F. W. Atoms in Molecules: A Quantum Theory; OxfordUniversity Press: New York, USA, 1994.(56) Tang, W.; Sanville, E.; Henkelman, G. J. Phys.: Condens. Matter2009, 21, 084204.(57) Munoz-García, A. B.; Pavone, M.; Carter, E. A. Chem. Mater.2011, 23, 4525−4536.(58) Munoz-García, A. B.; Pavone, M.; Ritzmann, A. M.; Carter, E. A.Phys. Chem. Chem. Phys. 2013, 15, 6250−6259.(59) Selbach, S. M.; Tolchard, J. R.; Fossdal, A.; Grande, T. J. SolidState Chem. 2012, 196, 249−254.(60) Jonsson, H.; Mills, G.; Jacobsen, K. W. In Classical AndQuantum Dynamics In Condensed Phase Simulations; Berne, B. J.,Ciccotti, G., Coker, D. F., Eds.; World Scientific: Hackensack, NJ,1998; pp 385−404.(61) Henkelman, G.; Uberuaga, B. P.; Jonsson, H. J. Chem. Phys.2000, 113, 9901−9904.(62) Henkelman, G.; Jonsson, H. J. Chem. Phys. 2000, 113, 9978−9985.(63) Sheppard, D.; Terrell, R.; Henkelman, G. J. Chem. Phys. 2008,128, 134106−134106−10.(64) Vasp TST Tools. http://theory.cm.utexas.edu/vtsttools/(accessed Nov 25, 2012).(65) Marezio, M.; Dernier, P. D. Mater. Res. Bull. 1971, 6, 23−29.(66) Shannon, R. D. Acta Crystallogr., Sect. A 1976, 32, 751−767.(67) Striker, T.; Ruud, J.; Gao, Y.; Heward, W.; Steinbruchel, C. SolidState Ionics 2007, 178, 1326−1336.(68) Wadati, H.; Kobayashi, D.; Kumigashira, H.; Okazaki, K.;Mizokawa, T.; Fujimori, A.; Horiba, K.; Oshima, M.; Hamada, N.;Lippmaa, M.; Kawasaki, M.; Koinuma, H. Phys. Rev. B 2005, 71,035108.(69) Tai, L.-W.; Nasrallah, M. M.; Anderson, H. U.; Sparlin, D. M.;Sehlin, S. R. Solid State Ionics 1995, 76, 273−283.(70) Patrakeev, M. V.; Bahteeva, J. A.; Mitberg, E. B.; Leonidov, I. A.;Kozhevnikov, V. L.; Poeppelmeier, K. R. J. Solid State Chem. 2003, 172,219−231.(71) Patrakeev, M. V.; Leonidov, I. A.; Kozhevnikov, V. L.;Poeppelmeier, K. R. J. Solid State Chem. 2005, 178, 921−927.(72) Gryaznov, D.; Heifets, E.; Kotomin, E. Phys. Chem. Chem. Phys.2012, 14, 4482−4490.(73) Orikasa, Y.; Nakao, T.; Oishi, M.; Ina, T.; Mineshige, A.;Amezawa, K.; Arai, H.; Ogumi, Z.; Uchimoto, Y. J. Mater. Chem. 2011,21, 14013−14019.(74) Søgaard, M.; Vang Hendriksen, P.; Mogensen, M. J. Solid StateChem. 2007, 180, 1489−1503.(75) Wærnhus, I.; Grande, T.; Wiik, K. Top. Catal. 2011, 54, 1009−1015.

Chemistry of Materials Article

dx.doi.org/10.1021/cm401052w | Chem. Mater. 2013, 25, 3011−30193019