in uence of inversion on mg mobility and … of inversion on mg mobility and electrochemistry in...

17
Influence of inversion on Mg mobility and electrochemistry in spinels Gopalakrishnan Sai Gautam, 1, 2, * Pieremanuele Canepa, 2, Alexander Urban, 2 Shou-Hang Bo, 2 and Gerbrand Ceder 3, 2, 1 Department of Materials Science and Engineering, Massachusetts Institute of Technology, Cambridge, MA 02139, USA 2 Materials Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, CA 94720, USA 3 Department of Materials Sciences and Engineering, University of California Berkeley, CA 94720, USA Magnesium oxide and sulfide spinels have recently attracted interest as cathode and electrolyte materials for energy-dense Mg batteries, but their observed electrochemical performance depends strongly on synthesis conditions. Using first principles calculations and percolation theory, we explore the extent to which spinel inversion influences Mg 2+ ionic mobility in MgMn2O4 as a pro- totypical cathode, and MgIn2S4 as a potential solid electrolyte. We find that spinel inversion and the resulting changes of the local cation ordering give rise to both increased and decreased Mg 2+ migration barriers, along specific migration pathways, in the oxide as well as the sulfide. To quantify the impact of spinel inversion on macroscopic Mg 2+ transport, we determine the percolation thresholds in both MgMn2O4 and MgIn2S4. Furthermore, we analyze the impact of inversion on the electrochemical properties of the MgMn2O4 cathode via changes in the phase behavior, average Mg insertion voltages and extractable capacities, at varying degrees of inversion. Our results confirm that inversion is a major performance limiting factor of Mg spinels and that synthesis techniques or compositions that stabilize the well-ordered spinel structure are crucial for the success of Mg spinels in multivalent batteries. I. INTRODUCTION Multivalent (MV) batteries, such as those based on Mg 2+ , 1,2 can potentially achieve high volumetric energy density via facile non-dendritic stripping/deposition on an energy-dense metal anode. 3–5 However, the develop- ment of viable MV technology is hindered by poor Mg diffusivity in oxide cathodes as well as poor Coulombic efficiencies in liquid electrolytes. 2,5–7 One pathway to improve Mg migration in solids is to utilize host structures where Mg occupies an unfavor- able coordination environment. 8–10 Spinels with compo- sition AM 2 X 4 (A = Mg, M = metal cations, X = O or S) are appealing structures in this regard because of their tetrahedrally-coordinated Mg sites, rather than the preferred octahedral coordination of Mg. Theoreti- cal calculations indeed predict reasonable Mg 2+ migra- tion barriers (550 - 750 meV) in both oxide and sulfide spinels. 11,12 Note that oxide spinels have long been used as cathodes and anodes in commercial Li-ion batteries. 13–19 Spinel-Mn 2 O 4 is a particularly promising, energy- dense, MV cathode, as it is one of the few oxides 20–25 to have shown electrochemically reversible Mg 2+ intercalation. 26,27 However, the cyclable Mg con- tent, i.e., the observed capacity, seems to depend strongly on the synthesis conditions. 26–28 Several studies on the MgMn 2 O 4 structure 29–33 have indicated that the spinel is prone to inversion, i.e., Mg/Mn antisite disorder (see Section II), where the degree of inversion can range from 20% 30 to 60%. 29 It has further been argued that the propensity of Mn 3+ to disproportionate into Mn 2+ and Mn 4+ promotes spinel inversion and phase transformations. 16,34 Since inversion directly affects the local cation arrangement, it may significantly impact the Mg 2+ ionic mobility. 35,36 For the rational design of im- proved Mg battery cathodes it is, therefore, crucial to understand how inversion in oxide spinels affects Mg 2+ migration. Inversion is not a phenomenon unique to oxides, and other chalcogenide spinels such as sulfides, which are also important cathode materials in MV technology, 12 are also known to exhibit inversion. 37,38 A recent combined the- oretical and experimental study has identified ternary sulfide and selenide spinels as promising Mg-ion conduc- tors with potential applications as solid electrolytes in MV batteries. 38 Solid electrolytes combine the advantage of improved safety with a high Mg transference num- ber. Three promising compounds were reported, namely, MgSc 2 Se 4 , MgSc 2 S 4 , and MgIn 2 S 4 . 38 MgIn 2 S 4 spinel had previously been reported, 39,40 and the available literature as well as our own synthesis attempts (Figure S1 in Sup- porting Information, SI *# ) indicate that the compound is prone to inversion, where the degree of inversion can be as high as 85% (Table S2 in SI). In the present work, motivated by the importance of the spinel structure for MV battery technology, we explore the influence of spinel inversion on Mg mobil- ity in ternary oxides and sulfides, using MgMn 2 O 4 and MgIn 2 S 4 as the prototype for each class of spinels. We consider all possible local cation environments that arise due to inversion and compute the activation barriers for Mg migration in each scenario using first-principles calcu- *# Electronic Supporting Information available free of charge online at http://dx.doi.org/10.1021/acs.chemmater.7b02820

Upload: doanhuong

Post on 04-May-2018

218 views

Category:

Documents


1 download

TRANSCRIPT

Influence of inversion on Mg mobility and electrochemistry in spinels

Gopalakrishnan Sai Gautam,1, 2, ∗ Pieremanuele Canepa,2, †

Alexander Urban,2 Shou-Hang Bo,2 and Gerbrand Ceder3, 2, ‡

1 Department of Materials Science and Engineering,Massachusetts Institute of Technology, Cambridge, MA 02139, USA

2 Materials Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, CA 94720, USA†

3 Department of Materials Sciences and Engineering,University of California Berkeley, CA 94720, USA

Magnesium oxide and sulfide spinels have recently attracted interest as cathode and electrolytematerials for energy-dense Mg batteries, but their observed electrochemical performance dependsstrongly on synthesis conditions. Using first principles calculations and percolation theory, weexplore the extent to which spinel inversion influences Mg2+ ionic mobility in MgMn2O4 as a pro-totypical cathode, and MgIn2S4 as a potential solid electrolyte. We find that spinel inversion andthe resulting changes of the local cation ordering give rise to both increased and decreased Mg2+

migration barriers, along specific migration pathways, in the oxide as well as the sulfide.To quantify the impact of spinel inversion on macroscopic Mg2+ transport, we determine the

percolation thresholds in both MgMn2O4 and MgIn2S4. Furthermore, we analyze the impact ofinversion on the electrochemical properties of the MgMn2O4 cathode via changes in the phasebehavior, average Mg insertion voltages and extractable capacities, at varying degrees of inversion.Our results confirm that inversion is a major performance limiting factor of Mg spinels and thatsynthesis techniques or compositions that stabilize the well-ordered spinel structure are crucial forthe success of Mg spinels in multivalent batteries.

I. INTRODUCTION

Multivalent (MV) batteries, such as those based onMg2+,1,2 can potentially achieve high volumetric energydensity via facile non-dendritic stripping/deposition onan energy-dense metal anode.3–5 However, the develop-ment of viable MV technology is hindered by poor Mgdiffusivity in oxide cathodes as well as poor Coulombicefficiencies in liquid electrolytes.2,5–7

One pathway to improve Mg migration in solids is toutilize host structures where Mg occupies an unfavor-able coordination environment.8–10 Spinels with compo-sition AM2X4 (A = Mg, M = metal cations, X = Oor S) are appealing structures in this regard becauseof their tetrahedrally-coordinated Mg sites, rather thanthe preferred octahedral coordination of Mg. Theoreti-cal calculations indeed predict reasonable Mg2+ migra-tion barriers (∼ 550 − 750 meV) in both oxide andsulfide spinels.11,12 Note that oxide spinels have longbeen used as cathodes and anodes in commercial Li-ionbatteries.13–19

Spinel-Mn2O4 is a particularly promising, energy-dense, MV cathode, as it is one of the fewoxides20–25 to have shown electrochemically reversibleMg2+ intercalation.26,27 However, the cyclable Mg con-tent, i.e., the observed capacity, seems to depend stronglyon the synthesis conditions.26–28 Several studies on theMgMn2O4 structure29–33 have indicated that the spinelis prone to inversion, i.e., Mg/Mn antisite disorder(see Section II), where the degree of inversion canrange from 20%30 to 60%.29 It has further been arguedthat the propensity of Mn3+ to disproportionate intoMn2+ and Mn4+ promotes spinel inversion and phasetransformations.16,34 Since inversion directly affects the

local cation arrangement, it may significantly impact theMg2+ ionic mobility.35,36 For the rational design of im-proved Mg battery cathodes it is, therefore, crucial tounderstand how inversion in oxide spinels affects Mg2+

migration.

Inversion is not a phenomenon unique to oxides, andother chalcogenide spinels such as sulfides, which are alsoimportant cathode materials in MV technology,12 are alsoknown to exhibit inversion.37,38 A recent combined the-oretical and experimental study has identified ternarysulfide and selenide spinels as promising Mg-ion conduc-tors with potential applications as solid electrolytes inMV batteries.38 Solid electrolytes combine the advantageof improved safety with a high Mg transference num-ber. Three promising compounds were reported, namely,MgSc2Se4, MgSc2S4, and MgIn2S4.38 MgIn2S4 spinel hadpreviously been reported,39,40 and the available literatureas well as our own synthesis attempts (Figure S1 in Sup-porting Information, SI∗#) indicate that the compoundis prone to inversion, where the degree of inversion canbe as high as ∼ 85% (Table S2 in SI).

In the present work, motivated by the importanceof the spinel structure for MV battery technology, weexplore the influence of spinel inversion on Mg mobil-ity in ternary oxides and sulfides, using MgMn2O4 andMgIn2S4 as the prototype for each class of spinels. Weconsider all possible local cation environments that arisedue to inversion and compute the activation barriers forMg migration in each scenario using first-principles calcu-

∗# Electronic Supporting Information available free of charge onlineat http://dx.doi.org/10.1021/acs.chemmater.7b02820

2

lations. The high requirement for the ionic conductivityin solid electrolytes typically demands migration barriersto be < 500 meV, as observed in solid Li-conductors,41

while cathodes can operate under lower ionic mobilities(barriers ∼ 750 meV, see Section IV A)2 as the requiredlength is less than for a conductor. Hence, we limit acces-sible Mg2+ migration paths to those with a barrier lessthan 500 meV and 750 meV for operation as a solid elec-trolyte and cathode, respectively. We will use MgMn2O4

as the prototype cathode for which we restrict barriersto 750 meV and MgIn2S4 as an example of an electrolyte(barriers < 500 meV).

Our results indicate that inversion, in both solid elec-trolytes and cathodes, can simultaneously cause a de-crease in activation barriers across certain migration tra-jectories while increasing the barriers across others, lead-ing to a complex interplay of opening and closing of spe-cific Mg migration pathways. To quantify the impactof these variations in the microscopic activation barri-ers on macroscopic Mg diffusion, we estimate the criticalMg concentrations (percolation thresholds) required tofacilitate Mg2+ diffusion through the structure at differ-ent degrees of inversion. Note that Mg extraction fromthe cathode material creates Mg-vacancies that can affectthe percolation properties. For example, vacancies cancause migration pathways that are inactive in the fullydischarged composition to become accessible. Hence, fora cathode, we examine the variation of the percolationthreshold with vacancy content in the spinel lattice. Inelectrolytes, the Mg concentration does not significantlyvary and we do not consider the effect of Mg-vacanciesin MgIn2S4. Our estimates indicate that stoichiomet-ric MgMn2O4 and MgIn2S4 spinels remain percolatingup to ∼ 55–59% and 44% inversion, respectively. Fi-nally, we discuss the impact of spinel inversion on Mg-electrochemistry in the Mn2O4 cathode by evaluating the0 K phase diagram, average voltages and the accessibleMg capacity at various degrees of inversion.

While previous studies have analyzed the impact of in-version on structural, thermal, electronic, and magneticproperties,30,42–45 the effect on Mg mobility in spinelshas not yet been investigated. Understanding the influ-ence of inversion on ion mobility will provide guidelines totune the synthesis and electrochemical conditions of bothcathodes and solid electrolytes, not only in MV systemsbut also in existing Li-ion architectures.46 Finally, our re-sults emphasize the importance of the topology of cationsites in setting the migration behavior within a generalanion framework.47

II. STRUCTURE

A spinel configuration is a specific ordering of cationsites (A and M in AM2X4) in a face-centered cubic (FCC)packing of anion sites (X), as shown in Figure 1. In a“normal” spinel, half of the octahedral (oct) sites, i.e.,16d, are occupied by M atoms (Mn/In, blue octahedra

in Figure 1), while 1/8 of the tetrahedral (tet) sites (8a)are occupied by A (Mg, orange tetrahedra) cations.

Polyhedra in the spinel structure share faces, edgesand corners, as summarized in Table I. For example, the8a sites that are occupied by A are face-sharing withvacant (Vac) 16c oct sites (dashed red square in Fig-ure 1a), edge-sharing with vacant 48f tet (dashed redtriangle) and corner-sharing with vacant tet (48f, 8b)and M-containing 16d oct sites.48 Face-sharing polyhe-dra have the lowest cation–cation distance, leading to thehighest level of electrostatic repulsion, followed by edge-sharing and subsequently corner-sharing polyhedra.49 In-deed, the 16c, 48f and 8b sites are vacant in spinel lattices(8b not shown in Figure 1) since they face-share with oc-cupied 8a or 16d sites.

Inversion in a spinel structure refers to the collec-tion of anti-site defects in the 8a (A) and 16d (M) sub-lattices, as shown in Figure 1b. The degree of inversion,i, is defined as the fraction of 8a sites occupied by Mcations, with a value of 0 (or 0%) and 1 (100%) indi-cating a normal and a fully inverted spinel, respectively.Thus, cations A and M are exchanged in inverted spinels(green arrows in Figure 1b), leading to a stoichiometry ofA1−iMi[Ai/2M1−(i/2)]2X4, compared to AM2X4 in nor-mal spinels.

A. Possible Mg-hops

Figure 2 and Table II summarize the possible localcation arrangements in a spinel structure that can origi-nate from inversion. The orange, blue, and green polyhe-dra in Figure 2 correspond to Mg, M, and mixed (Mg/M)occupation, respectively, with the arrows in each panelindicating the Mg migration trajectory. The dashed rect-angles and triangles signify vacancies. Grey polyhedracorrespond to 8a sites that are either cation occupied orvacant. While Figure 2a indicates the migration trajec-tory in a normal spinel, panels b, c, d, and e depict thepossible Mg-hops that can occur in an inverted spinel.The sub-panels in Figure 2b correspond to slices alongperpendicular directions, i.e., the 8a sites in the left sub-panel of Figure 2b are perpendicular to the plane of thepaper in the right sub-panel.

In a normal spinel, the rate for Mg diffusion is de-termined by the hop between adjacent 8a tet sites face-sharing with a 16c octahedron, as shown in Figure 2a.Hence, the migration topology is tet− oct− tet, and re-ferred to as “Hop 1” in our work. The intermediate 16csite in Hop 1 shares edges with six 16d oct sites (“ring”sites) that are occupied by M cations (2 out of 6 ring sitesare shown in Figure 2a). It was recently proposed8,11,12

that the migration barrier in normal spinels, both ox-ides and sulfides, is predominantly set by the size of theshared triangular face (not shown in Figure 2a) betweenthe 8a tet and 16c oct sites.

Along the tet−oct− tet migration pathway in invertedspinels (referred to as “Hop 2”) the 16d ring sites can be

3

Mg

MM

Mg

FIG. 1: Schematic of a normal (a) and an inverted (b) spinel MgM2X4 (M = Mn, In and X = O, S). The blue and orangepolyhedra correspond to the M (16d, oct) and Mg (8a, tet). The dashed rectangle indicates the vacant 16c, oct site and thedashed triangle the vacant 48f tet site. In (b), green arrows display the exchange of Mg and M sites, leading to inversion inthe spinel.

TABLE I: Notations used in the AM2X4 structure of Figure 1. Vac indicates vacancy. No. sites is normalized against theconventional (cubic) cell of a normal spinel with 32 anions.

Site Coordination Ion in normal spinel Sharing neighbors No. sites

Face Edge Corner

8a tet A (Mg2+) 16c 48f 48f, 16d, 8b 8

16d oct M (Mn3+,4+/In3+) 8b, 48f 16c, 16d 8a, 48f 16

16c oct Vac 8a, 48f 16d, 16c 8b, 48f 16

48f tet Vac 16d, 16c 8a, 8b, 48f 8a, 8b, 16c, 16d 48

8b tet Vac 16d 48f 48f, 16c, 8a 8

occupied by both M and Mg cations, as indicated by thesix green polyhedra in the right sub-panel of Figure 2b.To evaluate Mg2+ migration along Hop 2, we consideredmultiple configurations from 1 ring site occupied by Mgto all 6 ring sites being occupied by Mg. Since eachring site occupancy (e.g., 2/6 or 3/6 Mg) correspondsto a large number of possible cation decorations on thering sites, we used the decoration that had the lowestelectrostatic energy, as obtained by minimizing the Ewaldenergy of the unit cell50 using classical charges in thespinel framework. The specific cation arrangements usedto evaluate the Mg migration barriers along Hop 2 aredisplayed in Figure S13.

As inversion leads to Mg2+ occupancy of 16d sites,Mg-hopping across 16d sites must also be considered. A16d − 16d hop can occur through two possible tetrahe-dral intermediate sites, the 8b and 48f . The 8b sitestypically share all their triangular faces with occupied16d sites and are therefore not open to Mg2+ migrationdue to high electrostatic repulsion, as shown by previous

studies.8,36,47 However, the 48f sites share 2 triangularfaces with vacant 16c sites, enabling them to act as viableintermediate sites for Mg2+ hopping. As such, we onlyconsider the 16d − 16d hop via the 48f as intermediatesite, leading to a 16d−48f−16d topology (Figures 2c, d,and e). The 48f shares one of its edges with an 8a tet site(Table I), where the “edge-8a” can be occupied by Mg2+

(“Hop 3”, Figure 2c), M3+/4+ (“Hop 4”, Figure 2d) or avacancy (“Hop 5”, Figure 2c). Additionally, across Hops3, 4, and 5, we consider two scenarios where the 8a sitesthat share a corner with the 48f (“corner-8a”, grey poly-hedra in Figure 2) are either occupied by cations or leftvacant.

B. Percolation theory

While activation barriers for the various cation ar-rangements in Figure 2 determine the active Mg2+ migra-tion hops (or channels) on the atomic scale, the macro-

4

16d

M

16d

M

8a

Mg

8a

Vac

16c

oct tet oct

tet oct tet a) Hop 1 b) Hop 2

c) Hop 3 d) Hop 4 e) Hop 5

16d

Mg/M

16d

Mg/M

8a

Mg

8a

Vac

16c16c

16d

Mg/M

16d

Mg/M

16d

Mg/M

16d

Mg/M

16d

Mg/M

16d

Mg/M

16d

Mg

16d

Vac

48f

8a

Mg

(edge)

8a

8a

16d

Mg

16d

Vac

48f

8a

M

(edge)

8a

8a

16d

Mg

16d

Vac

48f

8a

Vac

(edge)

8a

8a

FIG. 2: Local cation environments and various Mg hops considered in an inverted spinel structure. In all migration scenarios aMg atom migrates from an occupied site (indicated by solid black circles) to an adjacent vacant site (dashed black rectangles),along the trajectory indicated by the arrows. Hops 1 (a) and 2 (b) occur with a tet→ oct→ tet topology, while hops 3 (c), 4(d), and 5 (e) occur along an oct → tet → oct pathway. Blue and orange polyhedra correspond to Mg and M (M = Mn, In),while green polyhedra indicate mixed M/Mg occupancy. In the case of Hops 3, 4, and 5 the 8a sites corner-sharing with theintermediate 48f site are shown as grey polyhedra. The notation “edge” in panels (c), (d) and (e) corresponds to the 8a sitethat edge-shares with the 48f . Vac indicates vacancy.

scopic diffusion of Mg2+, which is essential for (dis)chargeof cathodes or ionic conduction in solid electrolytes, de-pends on the existence of a percolating network of activemigration channels. As the 8a−16c−8a channels form apercolating network throughout the spinel structure, sto-ichiometric normal spinels with Mg in 8a enable macro-scopic diffusion of Mg2+ as long as the 8a− 16c− 8a hopis open, i.e., the migration barrier for Hop 1 is below athreshold value. However, inversion leads to mixing ofcation occupancies in both the 8a and 16d sites, poten-

tially causing some 8a− 16c− 8a channels to close (dueto higher Mg2+ migration barriers along Hop 2) whileopening new channels typically closed in a normal spinel(e.g., Hops 3, 4, or 5). Hence, in addition to identify-ing facile microscopic hops, it is important to considerwhether a percolating network of low-barrier migrationchannels exists. Analogous studies have been done onLi+ percolation in rocksalt lattices.36

In percolation theory, the site percolation problem51–54

identifies the critical concentration, x = xcrit, at which

5

TABLE II: Summary of all hops considered for evaluating Mg2+ mobility in inverted spinels, where M = Mn, In and Vac = Va-cancy. The neighbor column indicates the site that edge-shares with the intermediate site in the corresponding hop. The lastcolumn signifies the (maximum) number of configurations, along each migration trajectory, for which migration barriers havebeen calculated in this work. For example along Hop 3, the corner-8a sites being cation-occupied and vacant are the twoconfigurations considered.

Hop Topology Intermediate site neighbor(s) # configurations

1 8a− 16c− 8a (tet− oct− tet) 16d (oct, M) 1

2 8a− 16c− 8a (tet− oct− tet) 16d (oct, Mg/M) 6

3 16d− 48f − 16d (oct− tet− oct) 8a (tet, Mg) 2

4 16d− 48f − 16d (oct− tet− oct) 8a (tet, M) 2

5 16d− 48f − 16d (oct− tet− oct) 8a (tet, Vac) 2

an infinite network of contiguous connected sites existsin an infinite lattice of randomly occupied sites. In termsof ionic diffusion, xcrit sets the “percolation threshold”,above which percolating channels exist in a given struc-ture and macroscopic ion diffusion is feasible. Whilepercolation thresholds are accessible analytically for 2Dlattices,53 Monte-Carlo (MC) simulations need to be usedto estimate xcrit in 3D structures.36,55,56

The existence of a percolating diffusion network in astructure at a certain x (> xcrit) does not imply thatall ions in the structure can be (reversibly) extracted.Mg sites that are not part of a percolating network willform isolated clusters throughout the structure so thatthe amount of extractable ions is lower than the totalconcentration, i.e., xext < x. The quantity xext canbe assumed to correspond to the capacity of a cathodematerial. Numerically, xext is also estimated from MCsimulations.36

In summary, the two central quantities obtained frompercolation MC simulations are the Mg concentration be-yond which macroscopic diffusion is feasible (xcrit) andthe fraction of extractable Mg ions in a percolating struc-ture (xext). In order to study Mg diffusion in spinels, wemodified the nearest neighbor model (normally consid-ered in site percolation estimations) to include occupan-cies up to the 3rd nearest neighbor (i.e., corner-sharingsites in Table I). Two Mg sites in a given spinel ar-rangement are considered connected only if the migrationchannel linking them is open (i.e., the migration barrieris below an upper-limit). Thus, a percolating networkof Mg sites is formed solely via open migration channels.Whether a channel is considered open will depend on themigration barrier for Mg hopping through it.

III. METHODS

The computational approaches to predict propertiesrelevant to cathode materials have recently been reviewedby Urban et al.57 Also, the ability of Density Func-tional Theory (DFT)58,59 methods to predict materialswith novel properties has been amply demonstrated.60

As a result, all calculations in this work are done withDFT as implemented in the the Vienna Ab Initio Sim-

ulation Package,61,62 and employing the Projector Aug-mented Wave theory.63 An energy cut-off of 520 eV isused for describing the wave functions, which are sampledon a well-converged k -point (4×4×4) mesh. The elec-tronic exchange-correlation is described by the semi-localPerdew-Burke-Ernzerhof (PBE)64 functional of the Gen-eralized Gradient Approximation (GGA). Calculationson MgxMn2O4 are always initialized with an ideal cubicstructure while allowing for potential tetragonal distor-tions during the geometry relaxation as the spinel canbe either cubic (xMg ∼ 0) or tetragonal (xMg ∼ 1) basedon the concentration of Jahn-Teller active Mn3+ ions.The computed c/a ratio for the tetragonal-MgMn2O4

structure is in excellent agreement with experimentalreports29,65 (see Section S12). For voltage and 0 Kphase diagram calculations of MgxMn2O4, the PBESolexchange-correlation functional66 is used to improve thedescription of the energetics,67 while a Hubbard U cor-rection of 3.9 eV is added to remove spurious self-interaction of the Mn d -electrons.68–70

The activation barrier calculations are performed withthe Nudged Elastic Band (NEB) method.71,72 The barri-ers are calculated in a conventional spinel cell (32 anions),which ensures a minimum distance of ∼ 8 A between theelastic bands and reduces fictitious interactions with pe-riodic images. We verified that migration barriers do notchange appreciably (< 3% deviation) when equivalentcalculations are performed in larger supercells (see Fig-ure S2). Seven images are introduced between the initialand final end points to capture the saddle point and themigration trajectory. All NEB results are based on thePBE functional, without Hubbard U.10,11 The migrationbarriers in spinel-MgIn2S4 are calculated with compen-sating electrons added as a background charge to ensurecharge-neutrality of the structure at non-stoichiometricMg concentrations.

As migration barriers are calculated in the conven-tional spinel cell, the degree of inversion (i) that canbe modeled is constrained by the migration trajectoryunder consideration, in both the oxide and the sulfide.For example, along Hop 2 (Figure 2b), a 3/6 Mg ringsite occupancy leads to 3 Mn/In atoms in the 8a sites,and consequently results in i ∼ 3/8 = 0.375. Similarly,the barrier calculations along the 16d− 48f − 16d topol-

6

ogy (Hops 3, 4, and 5), which require a minimum of 2Mg atoms in the 16d sites (or 2 Mn/In sites in the 8a),correspond to i ∼ 0.25.

Monte-Carlo simulations are used to estimate the Mgpercolation thresholds (xcrit) and the fraction of ex-tractable Mg ions (xext). A 6×6×6 supercell of the prim-itive spinel structure is used, which corresponds to 1728anion atoms (Figure S6 plots convergence behavior withsupercell size). In MC simulations, a network of Mgsites is considered percolating when it spans the peri-odic boundaries of the simulation cell in one or moredirections.73 Inversion in the spinel is introduced duringMC sweeps by labelling a number of random 8a and 16dsites, corresponding to the degree of inversion, as part ofthe “Mg sub-lattice”. For example, the Mg sub-lattice ina normal spinel consists of all 8a sites. However, in aninverted spinel (with the degree of inversion i) the Mgsub-lattice will be composed of (1-i)% of all 8a sites and(i/2)% of all 16d sites.

To evaluate the M composition at which percolationoccurs, a MC sweep is performed with the followingsteps:73 (i) the supercell is initialized with M atomsin both M and Mg “sub-lattices”, corresponding to aM3X4 (X = O, S) stoichiometry, (ii) M atoms on theMg sub-lattice are randomly changed to Mg, (iii) afterall Mg sub-lattice sites are changed (i.e., a stoichiometryof MgM2X4 is attained), M atoms on the M sub-latticeare randomly flipped to Mg. During an MC sweep, oncea Mg atom replacement results in the formation of a per-colating network, the current Mg concentration (xMg) istaken as an estimate of the percolation threshold (xcrit),while for x > xcrit, the fraction of sites within the per-colating network, xext, is stored. The values of xcrit

and xext are averaged over 2000 MC sweeps to guaran-tee well-converged estimates. The effect of vacancies onMg percolation in the Mn-spinel is captured by initializ-ing the Mg sub-lattice with varying vacancy concentra-tions, at a given degree of inversion, corresponding to aVaczMn3−zO4 stoichiometry (z ≤ 1). Whenever vacan-cies are initialized in a supercell, only the Mn atoms arechanged to Mg during a MC sweep.

IV. RESULTS

A. MgMn2O4

Figure 3 plots the ranges of Mg2+ migration barriersin MgxMn2O4 (y-axis) for all hops of Figure 2 and Ta-ble II, while the raw data is included in Figure S3 of theSI. The migration barriers are calculated with respect tothe absolute energies of the end points, nominally iden-tical for a given Mg2+ hop. However, there are a fewcases where the end point energies are different, sincethe local symmetry of the cation decoration is brokendifferently across the end points (e.g., 3/6 hop in Fig-ure S3b). In such cases, the barrier is reported withrespect to the end point with the lowest energy. The

dotted black line in Figure 3 is the upper-limit of theMg migration barrier, as required for reasonable batteryperformance,2 and is used to determine the percolationthresholds (see Section IV C). For a MgxMn2O4 cathodeparticle of size ∼ 100 nm being (dis)charged at a C/3 rateat 60◦C, the migration barrier upper-limit is ∼ 750 meV(the upper-limit decreases to ∼ 660 meV at 298 K).2

Since full-cell Mg batteries so far have displayed superiorperformance at ∼ 60◦C than at 25◦C,1,74 the value of∼ 750 meV has been used as the cut-off to differentiate“open” and “closed” Mg2+ migration channels. In termsof notations, the fractions used in Hop 2 (e.g., 1/6, 2/6,etc., yellow rectangle in Figure 3) correspond to the frac-tion of 16d ring sites (Figure 2b) that are occupied byMg2+. The terms “8a empty” and “8a full” along Hops3, 4, and 5 in Figure 3 indicate that the corner-8a sites(Figures 2c, d, and e) are vacant and occupied by cations,respectively. xMg in Figure 3 is the Mg concentration inthe cell used for the barrier estimation, corresponding tothe “dilute Mg” (xMg ∼ 0, solid red lines) and “dilutevacancy” (xMg ∼ 1, dashed blue lines) limits.

Mg migration barriers along Hop 1 (tet−oct−tet, nor-mal spinel) at the dilute Mg and dilute vacancy limits are∼ 717 meV and ∼ 475 meV, respectively (red rectangle inFigure 3), in good agreement with previous studies.8,11,75

Note that the dilute Mg (vacancy) limit for Hop 1 corre-sponds to the regime when no 8a sites, other than thoserequired to model the hop, are occupied by Mg (vacan-cies). Since the migration barriers at both Mg concentra-tion limits are below ∼ 750 meV, Hop 1 is always openfor Mg migration. Barriers along Hop 2 (yellow rectanglein Figure 3) decrease initially with Mg occupation of the16d ring sites (∼ 393 meV at 2/6 vs. 536 meV at 1/6) be-fore increasing beyond 750 meV at 5/6 and 6/6 Mg. Thenon-monotonic variation of the migration barriers alongHop 2 is due to the gradual destabilization of the 16csite. The increasing instability of the 16c also changesthe migration energy profile (Figure S3b) from “valley”-like8 at 1/6 Mg to “plateau”-like at 5/6 Mg. Figure S14shows the Mg migration barriers along Hop 2 when thering sites are occupied by vacancies instead of Mg2+.

In the case of the oct − tet − oct Hops 3 and 4 (greenand cyan rectangles in Figure 3), which respectively havetet Mg and Mn edge-sharing with the intermediate 48fsite, the barriers vary drastically based on Mg contentand occupancy of the corner-8a sites. For example, at(i) xMg ∼ 0 and vacant corner-8a, the barrier along Hop3 (∼ 592 meV) is well below the upper-bound of 750 meV,while the barrier is comparable along Hop 4 (∼ 743 meV).At (ii) xMg ∼ 0 and cation-occupied corner-8a, the barri-ers along Hops 3 and 4 increase significantly (∼ 1388 meVand ∼ 1418 meV) and surpass the upper-limit set foropen channels. Eventually, at (iii) xMg ∼ 1 (cation-occupied corner-8a), the barriers decrease to ∼ 845 meVand ∼ 784 meV along Hops 3 and 4, respectively. Notethat the barriers along Hops 3 and 4 in Figure 3 are cal-culated at a degree of inversion, i ∼ 0.25. At a higherdegree of inversion (i ∼ 1) and xMg ∼ 1 (cation-occupied

7

1 2 3 4 5Hop

200

400

600

800

1000

1200

1400M

igra

tion

ba

rrie

r (m

eV

)x

Mg~ 0

xMg

~ 1

2/6

3/6

1/6

4/6

5/6

6/6

8a empty

8a full

8a empty

8a full

8a empty

8a full

475

717

1055

393

592

13881418

743703

319

FIG. 3: Ranges of Mg2+ migration barriers along the hops considered in spinel-MgxMn2O4. The dotted black line indicatesthe upper-limit of migration barriers (∼ 750 meV) used to distinguish open and closed migration channels in percolationsimulations. Solid red and dashed blue lines correspond to dilute Mg (xMg ∼ 0) and dilute vacancy (xMg ∼ 1) limits. Fractionsalong Hop 2 indicate the occupancy of Mg2+ in the 16d ring sites, while the legend “8a full (empty)” corresponds to cation-occupied (vacant) corner-8a sites along Hops 3 – 5. The barriers along Hop 1 are calculated at i ∼ 0, while Hops 3 – 5 havebeen done at i ∼ 0.25. Along Hop 2, i varies with Mg occupancy of the ring sites, ranging from i ∼ 0.125 at 1/6 Mg to i ∼ 0.75at 6/6 Mg. The raw data from Nudged Elastic Band calculations are displayed in Figure S3 of the SI.

corner-8a), the barrier is ∼ 1039 meV along Hop 4 (Fig-ure S5). Hence, from the data of Figure 3, Hop 3 is con-sidered closed for Mg migration whenever the corner-8asites are cation-occupied, while Hop 4 is always consid-ered a closed channel.

Mg migration barriers decrease significantly if theedge-8a is vacant (i.e., along Hop 5). For example, themigration barriers along Hop 5 (purple rectangle in Fig-ure 3) are well below that of Hops 3 and 4 across the sce-narios of (i) low Mg, vacant corner-8a (319 meV for Hop 5vs. 592 and 743 meV for Hops 3 and 4, respectively), (ii)low Mg, cation-occupied corner-8a (703 meV vs. 1388 and1418 meV), and (iii) high Mg, cation-occupied corner-8a(570 meV vs. 845 and 784 meV). Hence, Hop 5 is alwaysopen for Mg migration, since the barriers are below the

upper limit of 750 meV.

In summary, the tet − oct − tet pathway (Hops 1 and2) remains open for Mg migration in MgMn2O4 until ahigh degree of Mg occupation on the 16d ring sites (i.e.,≥ 5/6 Mg) is present, which corresponds to high degreesof inversion (i > 0.625). The oct − tet − oct pathway isopen only when the edge-8a is vacant (Hop 5) or whenthe corner-8a are vacant with Mg in the edge-8a (Hop 3).

B. MgIn2S4

Figure 4 plots the Mg2+ migration barriers in MgIn2S4

for the hops of Figure 2 (the raw data are shown in Fig-ure S4). Since we consider MgIn2S4 as an ionic con-

8

ductor, off-stoichiometric Mg concentrations are not ofinterest. Hence, all hops in Figure 4 are evaluated atthe dilute vacancy limit (xMg ∼ 1, dashed blue lines inFigure 4). The fractions used (1/6, 2/6, etc.) in Fig-ure 4 are the number of 16d ring sites occupied by Mg2+

in Hop 2. Along Hops 3 – 5, we use cation-occupiedcorner-8a sites (i.e., “8a full” in Figures S4c, d, and e).The upper-limit of the Mg migration barrier for classify-ing open and closed migration channels (as indicated bythe dotted black line in Figure 4) is set to ∼ 500 meV,based on migration barriers of ∼ 400 – 500 meV observedin fast Li-ion conductors, such as Garnets and Si-basedthio-LISICONs.41

In the case of Hop 1, the barrier is ∼ 447 meV, well be-low the upper limit of ∼ 500 meV. Mg migration barriersalong Hop 2 (yellow rectangle in Figure 4) follow trendssimilar to that of MgMn2O4 (Figure 3). For example,at low Mg occupation of the ring sites (1/6 or 2/6 Mg),the barrier is below the limits for percolating diffusion,before increasing beyond 500 meV at higher Mg contentin the ring sites (> 3/6 Mg). Also, the shape of the mi-gration energy curve changes from a “valley” at 1/6 Mg(solid black line in Figure S4b) to a “plateau” beyond2/6 Mg (solid red line in Figure S4b), indicating that the16c site becomes progressively unstable with increasingMg occupation of the ring 16d.

Along the 16d−48f−16d pathways (Hops 3, 4 and 5),the migration barriers are always higher than 500 meV,irrespective of the occupancy of the edge-8a. Indeed, themagnitude of the barriers are ∼ 683 meV, ∼ 531 meV,and ∼ 504 meV for Mg-occupied, In-occupied and vacantedge-8a, respectively, indicating that the oct − tet − octpathway will not be open for Mg2+ migration.

C. Percolation thresholds

Based on the data of Figures 3 and 4, and the up-per limits of Mg migration barriers set for MgMn2O4

(750 meV) and MgIn2S4 (500 meV), we compiled a listof conditions that enable the opening of the possible hopsin Table III. For example, Hop 1 (8a− 8a) is open for allvalues of xMg and i for both MgxMn2O4 and MgIn2S4.Both the oxide and the sulfide spinel exhibit high bar-riers (> 1 eV) for a 16d − 8a hop (Figure S8), whichwould limit Mg transfer between an octahedral 16d siteand an adjacent tetrahedral 8a site. Thus, in our perco-lation simulations, the 8a − 8a (Hops 1 and 2) and the16d− 16d (Hops 3, 4, and 5) channels remain decoupled,and a percolating network consists solely of either 8a−8aor 16d− 16d channels.

Figures 5a and b plot the percolation threshold (xcrit,black lines), at various degrees of inversion (i) inMn3−xO4 and In3−xS4. The dashed yellow lines indicatethe stoichiometric spinel, i.e., M:X = 2:4. The blue (red)shaded region corresponds to Mg concentration rangeswhich do (do not) exhibit percolation. The x-axis inFigure 5 begins at a M3X4 (i.e., 50% M-excess or 100%

TABLE III: Summary of rules used during percolation sim-ulations with the conditions for an open channel. The upperlimit of migration barriers used to distinguish between openand closed channels is 750 meV and 500 meV for MgMn2O4

and MgIn2S4, respectively.

Hop Topology Open under condition

MgMn2O4 – 750 meV

1 8a− 16c− 8a Always open

2 8a− 16c− 8a Max. 4/6 ring sites with Mg

3 16d− 48f − 16d Corner 8a vacant

4 16d− 48f − 16d Always closed

5 16d− 48f − 16d Always open

MgIn2S4 – 500 meV

1 8a− 16c− 8a Always open

2 8a− 16c− 8a Max. 2/6 ring sites with Mg

3 16d− 48f − 16d Always closed

4 16d− 48f − 16d Always closed

5 16d− 48f − 16d Always closed

Mg-deficient) configuration and spans concentrations upto Mg1.5M1.5X4 (i.e., 25% M-deficient, 50% Mg-excess).Generally, percolation thresholds in the M-excess domain(i.e., xcrit < 1) are desirable as this implies that the sto-ichiometric spinel will possess percolating networks andwill facilitate macroscopic Mg transport.

In the case of cathodes (Mn2O4), Mg deintercalationfrom the framework creates vacancies, which can facili-tate the formation of Mg percolating networks by open-ing certain migration channels (e.g., Hop 5 in MgMn2O4,Figure 3). Therefore, we explored the variation of thepercolation threshold with vacancy concentration (“z” inFigure 5a) in the Mn-spinel. For the sake of simplic-ity, x in Figure 5 refers to the sum of Mg and vacancyconcentrations. For example, x = 1 and z = 0.5 (greencircle on the dashed black line) in Figure 5a indicatesa composition of Mg0.5Vac0.5Mn2O4, while x = 0.6 andz = 0 (green square) corresponds to the M-excess spinel-Mg0.6Vac0Mn2.4O4.

The percolation threshold in the absence of vacan-cies (z = 0) is indicated by the solid black line in Fig-ure 5a. When vacancies are introduced, the thresholddecreases, as indicated by the xcrit at z = 0.4 (dottedblack line) or 0.5 (dashed) consistently exhibiting lowervalues than xcrit at z = 0 in Figure 5a. For example,at x = 0.8 and i = 0.5 (indicated by the green star inFigure 5a), the spinel does not form a percolating net-work when there are no vacancies (z = 0, Mg0.8Mn2.2O4),since xcrit ∼ 0.88 > 0.8. However, the structure canpercolate Mg when vacancies are introduced (z = 0.5,Mg0.3Vac0.5Mn2.2O4), as xcrit reduces to ∼ 0.52 < 0.8.In a case such as this, the initial cathode structure maynot be percolating, but introducing vacancies in the ini-tial part of the charge can create a percolating zone onthe cathode particle surface through which further Mg-removal can occur. However, upon discharge the perco-

9

683

531

504

1 2 3 4 5Hop

300

400

500

600

700

800M

igra

tion

ba

rrie

r (m

eV

)x

Mg~ 1

4472/6

1/6

4/6

5/6

3/6

6/6

458468

517

548

575

651

FIG. 4: Mg2+ migration barriers along each possible hop in spinel-MgIn2S4. The dotted black line indicates the upper-limitof migration barriers (∼ 500 meV) used to distinguish open and closed migration channels in percolation simulations. Dashedblue lines indicate the dilute vacancy (xMg ∼ 1) limit. Fractions along Hop 2 indicate the occupancy of Mg2+ in the 16d ringsites, while the corner-8a sites are cation-occupied across Hops 3 – 5. The barrier along Hop 1 is calculated at i ∼ 0, whileHops 3 – 5 have been done at i ∼ 0.25. Along Hop 2, i varies with Mg occupancy of the ring sites, ranging from i ∼ 0.125 at1/6 Mg to i ∼ 0.75 at 6/6 Mg. The raw data from Nudged Elastic Band calculations are displayed in Figure S4.

lating structure could easily become non-percolating ifpolarization increases the surface Mg concentration toorapidly.

At any degree of inversion, the magnitude of xcrit

varies non-monotonically and reduces only up to a va-cancy content, z = 0.4 or 0.5 (see Figure S7a). Indeed,at x = 0.8 and i = 0.5 (green star), an increase in z be-yond 0.5 (such as z = 0.6, Mg0.2Vac0.6Mn2.2O4), causesthe xcrit to increase to ∼ 0.6, but the spinel continuesto percolate. Thus, the shaded grey region in Figure 5a,which is bound by the z = 0.4, 0.5 and 0 lines representsthe extent of variation of xcrit with vacancy content inthe cathode. Notably, the lowest value of xcrit is obtainedat z = 0.4 for 0 ≤ i ≤ 0.35 and 0.595 ≤ i < 0.77, and atz = 0.5 for 0.35 ≤ i ≤ 0.595, respectively.

The stoichiometric {Mg/Vac}Mn2O4 spinel at i = 0(dashed yellow line in Figure 5a), permits macroscopicMg diffusion, since the percolation threshold (xcrit ∼0.44 for z = 0 – 0.4) is in the Mn-excess domain (i.e.,xcrit < 1). When vacancies are absent in the sto-ichiometric spinel (z = 0), which corresponds to thedischarged MgMn2O4 composition, the structure perco-lates Mg up to i ∼ 0.55. Upon charging, the presenceof vacancies (z = 0.5) enables Mg percolation withinMg0.5Vac0.5Mn2O4 up to i ∼ 0.59. At higher degreesof inversion (0.59 < i < 0.77), the oxide spinel requiresMn-deficient concentrations (i.e., x > 1) to facilitate Mgpercolation, as illustrated by xcrit ∼ 1.05− 1.13 (z = 0.5– 0) at i = 0.6. At i > 0.77, the oxide does not forma percolating Mg network at any level of Mn-deficiency

10

a)2O4 - 750 meV

b) MgIn2S4 - 500 meV

MgMn

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4x in (Mgx−zVacz)Mn3−xO4

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

Deg

ree

ofsp

inel

inve

rsio

n(i

) z=0z=0.4z=0.5

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4x in MgxIn3−xS4

0.0

0.1

0.2

0.3

0.4

0.5

Deg

ree

ofsp

inel

inve

rsio

n(i

) z=0

FIG. 5: The critical concentration for Mg percolation (xcrit) in the (Mgx−zVacz)Mn3−xO4 (a) and MgxIn3−xS4 (b) spinels areplotted as black lines at different degrees of spinel inversion i. The stoichiometric spinel concentration (M:X = 2:4) is indicatedby the dashed yellow lines. Note that the zero on the x-axis corresponds to a stoichiometry of M3X4 (M = Mn/In and X =O/S). z indicates the vacancy concentration in the structure. The shaded red (blue) region in both panels indicates the Mgconcentration range where macroscopic Mg diffusion is not possible (possible). The shaded grey region in panel (a) refers tothe range of variation of the percolation threshold with vacancy content in the oxide cathode. The green circle, square and starin panel (a) correspond to sample scenarios discussed in the text.

(for z ≤ 1) in the lattice.

In stoichiometric ionic conductors, such as MgIn2S4,the vacancy concentration is low and therefore vacanciesare not expected to play a major role in macroscopicMg transport. Specifically in MgIn2S4, vacancies do not

open additional migration channels, as indicated by theclosed Hop 5 in Figure 4. Indeed, the percolation thresh-old in the In-spinel does not change up to a vacancycontent, z = 0.2 in the structure (see Figure S7b). Atz = 0, the xcrit in In3−xS4 (solid black line in Figure 5b)

11

increases continuously with increase in inversion, withxcrit ∼ 0.435, and 0.74 at i = 0, and 0.4, respectively.Thus, at low i, stoichiometric MgIn2S4 should exhibitsignificant ionic conductivity. However, at higher degreesof inversion (i > 0.44), the sulfide spinel does not formpercolating networks at any Mg-concentration, owing tothe absence of open 16d − 16d channels in combinationwith the 8a − 8a channels being closed beyond 2/6 Mgring site occupancy (Table III).

In general, mobility requirements in an ionic conduc-tor are more stringent than in a cathode, consistent withthe stricter cut-off of 500 meV we applied to the mi-gration barriers in MgIn2S4.41,76 Indeed, a sulfide spinelMg-cathode (such as MgxTi2S4

74) exhibiting similar ac-tivation barriers with inversion as MgIn2S4 will not suf-fer from any percolation bottlenecks, since the barriersacross all cation arrangements are well below the milder750 meV cut-off set for cathodes (Figure 4).

D. Impact of inversion on cathode electrochemistry

Under ideal conditions, the structure of an ionic con-ductor (such as MgIn2S4) should not undergo significantchanges during operation. Thus, the extent of inver-sion should, in principle, be measured using characteri-zation experiments post-synthesis (the calculated forma-tion energies of various inverted configurations in spinel-MgIn2S4 are plotted in Figure S11). However in a cath-ode material such as MgxMn2O4, which can generatemobile Mn2+ ions (Figure S9) through disproportion-ation of Mn3+, the degree of inversion (i) can changeduring electrochemical cycling.16,34 Consequently, struc-tural changes in a cathode during cycling should mani-fest themselves as changes in the voltage profile and ob-served capacity, which can be benchmarked with theo-retical predictions.2,21 To evaluate the effect of inversionon the voltage profile of MgxMn2O4, we calculated thephase diagram and energy of the intercalation system at0 K as a function of Mg content under various degrees ofinversion.16,17,21,77

To evaluate the ground state hull of the MgxMn2O4

system, we enumerated over 400 Mg-vacancy configu-rations, at different Mg concentrations (xMg = 0, 0.25,0.5, 0.75 and 1) and different degrees of inversion (i=0, 0.25, 0.5, 0.75 and 1). Figure 6a displays structureswith formation energies (y-axis) below 200 meV/Mn2O4

at different Mg concentrations (x-axis), and the forma-tion energies of all the Mg-vacancy configurations consid-ered are plotted in Figure S10 of the SI. Notably, forma-tion energies in Figure 6a have been referenced to thenon-inverted (i = 0), empty Mn2O4 and magnesiated(MgMn2O4) spinel configurations. For each configura-tion, the degree of inversion is indicated by the corre-sponding symbol used, ranging from i = 0 (black circles)to i = 1 (red stars).

Overall, the MgxMn2O4 system is phase separatingat 0 K across non-inverted (i = 0) MgMn2O4 and

Mn2O4 domains, since the ground state hull of the system(dashed black line in Figure 6a) only exhibits two config-urations (i.e., MgMn2O4 and Mn2O4). Some solubilityat low Mg content may be possible given the low positivemixing energy at xMg = 0.25 for the non-inverted spinel(Eformation ∼ 14 meV/Mn2O4). At higher Mg content,the formation energies are very high for the non-invertedspinel (Figure S10), making a solid solution behavior veryunlikely. Inversion becomes likely to occur at intermedi-ate Mg compositions, as the low positive formation ener-gies are on the scale of the configurational entropy. Forexample, Eformation ∼ 11 meV/Mn2O4 at i = 0.25 andxMg = 0.5 (green square at x = 0.5 in Figure 6a). Hence,inversion at intermediate states of magnesiation is likely.While Mg by definition has to be mobile in Mn2O4 tooperate as a cathode, Mn mobility, which is required forspinel inversion to occur, depends strongly on its valencestate.16,34 Typically, Mn3+ can be mobile through a tem-porary disproportionation mechanism, generating mobileMn2+ (Figure S9).16,34

Figure 6b plots the average voltages as a function ofxMg at different i by taking the lowest Eformation con-figuration at each i and xMg.77 The average voltage forMg insertion in the non-inverted (i = 0) configurationis ∼ 2.84 V (dashed black line in Figure 6b), in agree-ment with previous theoretical estimates.11,78 Inversiondoes increase the average insertion voltage (averaged overxMg = 0 to 1) marginally compared to the normal spinel,with specific values of ∼ 2.92, 2.99, 2.97 and 2.99 V ati = 0.25, 0.5, 0.75 and 1, respectively. Notably, the phasebehavior of the MgxMn2O4 system under inversion willbe different compared to the normal spinel due to theformation of metastable inverted states at intermediateMg compositions.

The extractable Mg content (xext, see Section II B), ob-tained as a function of inversion from our Monte-Carlosimulations, indicates the extractable capacity of a cath-ode particle, and is shown in Figure 6c for stoichiomet-ric MgMn2O4. The y-axis indicates the % of the cath-ode’s theoretical capacity (∼ 270 mAh/g for MgMn2O4),that can be cycled reversibly. At low degrees of in-version, the extractable capacity in the stoichiometricspinel decreases roughly linearly with the degree of in-version, reaching ∼ 41% (∼ 110 mAh/g) at i = 0.4.The extractable Mg content decreases more rapidly fromi = 0.4 to i = 0.5, before stabilizing around ∼ 15%(∼ 40 mAh/g) between i = 0.5 and 0.6. Eventually, noneof the Mg becomes extractable beyond i = 0.61, reflect-ing the trends in the percolation thresholds (xcrit ∼ 0.59at stoichiometric MgMn2O4, Figure 5a) at high degreesof inversion. Note that, the overall amount of cyclableMg from a cathode particle is influenced both by theextractable Mg (shown in Figure 6c) and by the phasebehavior as a function of xMg. For example, if the Mgremoval occurs via a two-phase reaction (as is the casefor the non-inverted spinel), then the presence of a non-percolating layer on the surface may prevent extractionof Mg from the bulk, even if percolation conditions are

12

0 0.25 0.5 0.75 1x in Mg

xMn

2O

4

i = 0i = 0.25

i = 0.5i = 0.75

i = 1a)

0

20

40

60

80

100

120

140

160

180

200

Fo

rma

tion

en

erg

y (

me

V/M

n2O

4)

0 0.25 0.5 0.75 1x in Mg

xMn

2O

4

2.7

2.8

2.9

3

3.1

3.2

3.3

3.4

3.5

Ave

rag

e V

olta

ge

(V

) i = 0i = 0.25i = 0.5i = 0.75i = 1

b)

0 0.2 0.4 0.6 0.8 1Degree of inversion (i)

0

20

40

60

80

100

Extr

acta

ble

ca

pa

city

(%

)

c)

FIG. 6: (a) Ground state hull (or 0 K phase diagram) of the MgxMn2O4 system, with the zero of the formation energyreferenced to the non-inverted (i=0) magnesiated (MgMn2O4) and empty (Mn2O4) spinel configurations. (b) Average voltagecurves under i in MgxMn2O4, obtained using the lowest formation energy structures at each i across Mg concentrations. (c)The percentage of the theoretical capacity that can be reversibly extracted is plotted as a function of inversion in stoichiometricMgMn2O4.

still favorable in the bulk material.

V. DISCUSSION

In this work, we have used DFT-based NEB calcu-lations to assess the changes in the activation barrier

for Mg2+ migration arising from inversion in both ox-ide (MgMn2O4) and sulfide (MgIn2S4) structures. Fromour results (Figures 3 and 4), we can conclude that inver-sion has a significant impact on both oxides and sulfides,by opening and closing specific migration trajectories.In order to extrapolate the impact of the various Mg2+

migration barriers on macroscopic Mg diffusion, we esti-

13

mated the percolation thresholds under different degreesof spinel inversion. Furthermore, we analyzed the impactof spinel inversion on cathode properties of MgxMn2O4

by evaluating the average voltages and practical capaci-ties at different degrees of inversion.

A. Factors influencing barriers in MgMn2O4

Trends from activation barriers of Figure 3 suggest thatMg migration along the 8a− 16c− 8a pathways (Hops 1and 2) can improve significantly with Mg occupation ofthe 16d ring sites (up to 4/6 Mg), at low degrees of in-version. Additionally, the 16d− 48f − 16d channels openfor Mg migration whenever the edge-8a is vacant. How-ever, high degrees of inversion detrimentally affect Mg2+

motion, due to the closing of both 16d−16d (corner- andedge-8a become occupied by the metal cation) and 8a−8achannels (high migration barriers at high Mg in the ringsites). Although we have specifically considered the caseof spinel-MgxMn2O4, similar trends can be expected forother oxide spinels, given the similarity in Mg migrationbarriers along Hop 1 with different 3d-metals.11

Previous studies have used electrostatic considerationsto partially explain trends in Li+ activation barriers ina Mn2O4 spinel.46 Indeed, the reduction in Mg migra-tion barriers along Hops 1, 3, 4, and 5 (Figure 3) withincreasing Mg concentration can be attributed to lowerelectrostatic repulsions at the corresponding intermedi-ate sites caused by the reduction of Mn4+ to Mn3+. Forexample, the barrier reduces from 717 to 475 meV alongHop 1 and 1388 to 845 meV along Hop 3, as xMg in-creases from ∼ 0 to ∼ 1. However, Mg2+ activation bar-riers generally depend on steric and bonding constraintsin addition to electrostatics, which are often difficult todeconvolute over a range of NEB calculations. For exam-ple, the Mg2+ activation barriers across Hop 2 (yellow barin Figure 3) at low Mg occupation in the ring sites (1/6,2/6) are lower than Hop 1 (red bar, Figure 3), whichmay be attributed to reduced electrostatic repulsion onthe intermediate 16c (due to Mg2+ replacing higher va-lent Mn in the ring sites). However, barriers along Hop2 increase beyond Hop 1 and eventually beyond the limitof ∼ 750 meV at higher Mg in the ring sites (5/6, 6/6),despite lower electrostatic repulsion. Thus, the high Mgcontent in the ring sites decreases the stability of the in-termediate 16c. One possible reason for the instabilityof the 16c site could arise from charge-deficient oxygenatoms being shared with adjacent, Mg2+-occupied (in-stead of Mn3+/4+) 16d sites. Indeed, the instability ofthe 16c (e.g., in the case of 6/6 Mg in Hop 2) is quantifiedby longer (DFT-based) ∼ 2.3 A Mg–O bonds, comparedto ∼ 2.08 A in 16d with Mg (along the same hop) and∼ 2.13 A in rocksalt MgO.78

For the 16d − 48f − 16d hops in Figure 3 (Hops 3–5), electrostatic effects are more dominant than for thetet−oct−tet hops (Hops 1, 2), primarily due to the inter-mediate 48f edge-sharing with an 8a. Indeed, the cation

centers in edge-sharing tetrahedra are closer (∼ 2.15 Aexperimentally between 48f and 8a in an ideal LiMn2O4-spinel79) than in edge-sharing octahedra (∼ 2.88 A be-tween 16c and 16d). Consequently, the Mg barriers areconsistently lower with a vacant edge-8a (Hop 5, Fig-ure 3) compared to Mg/Mn-filled edge-8a (Hops 3, 4 inFigure 3). Also, Mg2+ activation barriers (at xMg ∼ 0)increase significantly when the corner-8a sites are cation-occupied rather than vacant (Figure 3). A closer lookat the cation-cation distances across corner-sharing 48fand 8a (∼ 2.88 A in ideal LiMn2O4) reveals that thecorner-sharing tetrahedra within a spinel framework mayexperience electrostatic repulsion as high as edge-sharedoctahedra (i.e., 16c and 16d). Thus, the combinationof cation-cation repulsion arising from both edge- andcorner-8a sites results in the high barriers along Hops 3and 4.

B. Barriers in sulfides vs. oxides

Activation barriers calculated in MgIn2S4 (Figure 4)exhibit similar trends to MgMn2O4 (Figure 3), resultingfrom analogous trends in electrostatics, steric and bond-ing environments. However, the absolute changes in bar-riers in the sulfide are remarkably lower than the oxide.For example, the absolute difference between the lowestand the highest Mg migration barriers of MgMn2O4 (atxMg ∼ 1) across Hops 1 through 5 is ∼ 662 meV (1055 –393 meV), while this is a much lower ∼ 236 meV (683 –447 meV) for MgIn2S4. Similarly, the barriers along the16d − 48f − 16d trajectory are far less sensitive to theedge-8a occupancy in the sulfide (504–673 meV) than inthe oxide (570–845 meV at xMg ∼ 1). Surprisingly, themigration barrier with an edge-8a occupied by Mg2+ ishigher (∼ 683 meV) than when the edge-8a is occupiedby In3+ (∼ 531 meV), suggesting that the In–S bondingenvironment screens the higher In3+ charge better thanthe Mg–S bonds screen Mg2+.

Lower activation barriers for Mg in sulfides have beenreported before,11,12,38 which have been assigned to ro-bust electrostatic screening, high polarizability, higherdegree of covalency and large volume per anion of S2−

compared to O2−.2,76 For example, a MgxTi2S474 cath-

ode will not suffer from any percolation bottlenecks, if thebarriers across all cation arrangements are similar to thecalculated values in MgIn2S4 (i.e., < 750 meV, Figure 4).But a more stringent upper-bound of ∼ 500 meV on thebarrier in a solid-state conductor41,76 indicates that in-version can significantly affect a sulfide ionic conductorby closing all 16d − 16d channels and several 8a − 8achannels with high Mg in the 16d ring (Figure 4). Sinceionic mobility is expected to improve with larger anionsand higher covalency (such as Se2− compared to S2−

and O2−), inversion is expected to affect Mg-mobilityto a lesser extent in Mg-containing Se-spinels, such asMgSc2Se4, compared to oxides and sulfides.

14

C. Percolation under inversion

Estimations of percolation thresholds (xcrit) in theMgxMn3−xO4 system (Figure 5a) indicate that spinelinversion should not detrimentally affect macroscopicMg2+ diffusion across the structure up to a fairly highdegree of inversion, i ∼ 0.55− 0.59. However, Mg-excessconcentrations are required to ensure percolating net-works form at i = 0.6 − 0.7, while the spinel completelyceases to percolate Mg beyond i = 0.77 (Figure 5a).Given the preponderance of conversion reactions underMg-excess concentrations in the oxide spinel, specificallythe decomposition of MgxMn3−xO4 (x > 1) into MgOand MnO,2 it is of paramount importance that the chem-ically synthesizable, stoichiometric {Mg/Vac}Mn2O4 re-mains percolating. Efforts should be made to reduce orprecisely control the amount of inversion (i.e., i < 0.6),by carefully tuning synthesis temperature and coolingrate29,30 during MgMn2O4 synthesis.

Higher Mg conductivity, as is required for a solid stateelectrolyte, demands a lower cut-off for the migrationbarrier along a pathway. In the case of MgIn2S4, wherewe used a 500 meV cut-off, MC simulations indicate thatthe stoichiometric spinel should remain percolating up toi ∼ 0.44. However, high degrees of inversion (i ∼ 0.85)can be observed during MgIn2S4 synthesis (Figure S1).As a result, strategies to limit inversion (i.e., i < 0.44) insulfide spinel ionic conductors, such as chemical dopingand careful calibration of synthesis conditions, need tobe sought.

D. Voltages and capacities

Inversion can also significantly impact electrochemicalproperties, such as phase behavior, average voltages andextractable capacities in an oxide-spinel cathode (Fig-ure 6). For example, the average voltage for Mg inter-calation, across xMg = 0 − 1 in the Mn2O4-spinel, ishigher in an inverted spinel compared to a normal spinel(Figure 6b). Mg intercalation experiments in spinel-Mn2O4 have reported a marginally higher average volt-age (∼ 2.9 V)26 than predicted for the normal spinel(∼ 2.84 V), with extraction voltages as high as ∼ 3.5 Vduring the charging cycle, which might be an indicationof the spinel inverting during electrochemistry. Also, thecalculated 0 K phase diagram of the Mg-Mn2O4 system(Figure 6a) suggests that the tendency to invert is thehighest at an intermediate Mg concentration, as indi-cated by low Eformation (< 50 meV/Mn2O4) configu-rations with i = 0.25 at xMg = 0.5. Hence, the degreeof inversion in the Mn-spinel can indeed change dynam-ically during electrochemical Mg cycling, especially dueto the presence of mobile Mn2+ ions (Figure S9). As re-ported by Ling et al.80, the mobility of Mn2+ within thespinel can also depend on the local arrangement of Mg2+

ions. Thus, from the data in Figure 6a, we expect the de-gree of inversion to vary largely between 0 and 0.25 dur-

ing Mg-cycling. Also, previous Mg-cycling experimentsin spinel-Mn2O4 have reported solvent co-intercalationbased phase transformations,28,81 which can be aided bythe presence of mobile Mn2+ ions.

Additionally, the first Mg-site that will be (de-)intercalated in the spinel will depend on the degree of in-version on the surface of the cathode particle. For exam-ple, if the degree of inversion is ∼ 0 on the surface, thenMg2+ ions present in the 8a sites will be de-intercalatedfirst from magnesiated-MgMn2O4. Similarly, in a par-tially inverted surface of a discharged cathode, the Mg2+

ions in 16d sites that are connected via Hop 3 channelswill be extracted as well as those in 8a sites connectedvia Hop 1 and open Hop 2 channels. In the case of apartially inverted surface in a charged-Mn2O4 cathode,the Mg2+ ions are more likely to first insert into 16dchannels connected via Hop 5, since Hop 5 exhibits lowerMg migration barriers compared to Hop 1 (Figure 2) atxMg ∼ 0.

Since the percolation threshold in the oxide cathodecan change with the vacancy concentration during Mg(de)intercalation (Figure 5a), a dynamic change in thedegree of inversion during Mg-cycling can cause polar-ization within the cathode particle. For example, if ichanges from 0.55 to 0.59 while charging the MgMn2O4

cathode, during the following discharge the spinel is per-colating only up to Mg0.5Vac0.5Mn2O4 (z = 0.5 in Fig-ure 5a) at i = 0.59. For further discharge into the struc-ture, i.e., from Mg0.5Vac0.5Mn2O4 to MgMn2O4, a re-duction in i to 0.55 is necessary, which can lead to hys-teresis in the voltages during the charge and dischargecycles. Importantly, the extractable Mg content in stoi-chiometric MgMn2O4 decreases continuously with inver-sion, reaching values of ∼ 63% (171 mAh/g) and ∼ 17%(46 mAh/g) at i = 0.25 and 0.5 (Figure 6c), respectively.Thus, strategies to minimize changes in i, during Mg2+

cycling, such as cation-doping of Mn to prevent Mn2+

generation, should be employed to ensure reversible Mg(de)intercalation.

VI. CONCLUSION

Spinels are promising materials in the development ofmultivalent battery electrodes and solid electrolytes butare prone to antisite disorder in the form of spinel inver-sion. With the example of two prototypical oxide andsulfide spinels, MgMn2O4 (cathode) and MgIn2S4 (solidelectrolyte), we demonstrated that inversion can signifi-cantly impact both Mg-ion mobility and electrochemicalproperties. Using first-principles calculations, we ana-lyzed the migration barrier for Mg2+ hopping in differentlocal cation arrangements, and found that inversion canboth open and close select migration pathways on theatomic scale. To quantify the influence of local barrierchanges on the macroscopic transport of Mg2+ ions, wedetermined the minimal M-deficiency x in MgxM3−xX4

required for percolation. Using a cut-off of 750 meV and

15

500 meV for cathodes and solid electrolytes, respectively,we found that the stoichiometric MgMn2O4 and MgIn2S4

compositions are Mg percolating up to ∼ 55–59% and44% inversion. Since the degree of inversion in the spinelsconsidered in this work may vary between 20% and 85%depending on the method of preparation,29,30,38 a care-ful calibration of the synthesis conditions is essential toensure sufficient Mg transport and to reduce the resul-tant impedance. In addition, spinel inversion can affectthe electrochemical properties of cathode materials bychanging the phase behavior, average voltage, and ex-tractable capacities. Specifically, we find that the degreeof inversion can change dynamically during electrochem-ical Mg cycling, as indicated by the 0 K phase diagramof the MgxMn2O4 system and the activation barriers forMn2+ hopping. Notably, even low degrees of inversion(i < 0.4) can detrimentally reduce the extractable capac-ity in stoichiometric MgMn2O4, with an estimated 15%decrease in capacity with every 10% increase in inversion.Thus, spinel inversion can hinder the electrochemical per-formance of both cathodes and solid electrolytes in MVsystems and synthesis efforts must always be made tostabilize the normal spinel structure.

Given that the Mg2+ migration barriers over a range ofoxide11 and sulfide spinels12 show similar trends, we ex-

pect similar behavior upon inversion in other spinel ma-terials. Finally, the framework developed in this work,particularly the data reported on percolation thresh-olds and extractable Mg, is readily transferable to otherspinels that have potential applications in Li-ion, Na-ion,Ca/Zn-multivalent and other battery fields.

Acknowledgments

The current work is fully supported by the Joint Centerfor Energy Storage Research (JCESR), an Energy Inno-vation Hub funded by the U.S. Department of Energy,Office of Science and Basic Energy Sciences. This studywas supported by Subcontract 3F-31144. The authorsthank the National Energy Research Scientific Comput-ing Center (NERSC) for providing computing resources.Use of the Advanced Photon Source at Argonne NationalLaboratory was supported by the U.S. Department of En-ergy, Office of Science, Office of Basic Energy Sciences,under Contract No. DE-AC02-06CH11357. The authorsdeclare no competing financial interests. GSG is thank-ful to Daniel C. Hannah at Lawrence Berkeley NationalLaboratory for a thorough reading of the manuscript.

∗ Electronic address: [email protected]† G. S. Gautam and P. Canepa contributed equally to the

work‡ Electronic address: [email protected], [email protected] D. Aurbach, Z. Lu, A. Schechter, Y. Gofer, H. Gizbar,

R. Turgeman, Y. Cohen, M. Moshkovich, and E. Levi,Nature 407, 724 (2000), ISSN 0028-0836, URL http:

//dx.doi.org/10.1038/35037553.2 P. Canepa, G. S. Gautam, D. C. Hannah, R. Malik, M. Liu,

K. G. Gallagher, K. A. Persson, and G. Ceder, Chem. Rev.(2017).

3 H. D. Yoo, I. Shterenberg, Y. Gofer, G. Gershinsky,N. Pour, and D. Aurbach, Energy Environ. Sci. 6, 2265(2013), ISSN 1754-5706, URL http://dx.doi.org/10.

1039/C3EE40871J.4 J. O. Besenhard and M. Winter, ChemPhysChem

3, 155 (2002), ISSN 1439-7641, URL http:

//dx.doi.org/10.1002/1439-7641(20020215)3:2<155::

AID-CPHC155>3.0.CO;2-S.5 J. Muldoon, C. B. Bucur, and T. D. Gregory, Chem. Rev.

114, 11683 (2014), URL http://dx.doi.org/10.1021/

cr500049y.6 P. Canepa, S. Jayaraman, L. Cheng, N. Rajput, W. D.

Richards, G. Sai Gautam, L. A. Curtiss, K. Persson, andG. Ceder, Energy Environ. Sci. 8, 3718 (2015), ISSN 1754-5706, URL http://dx.doi.org/10.1039/C5EE02340H.

7 P. Canepa, G. S. Gautam, R. Malik, S. Jayaraman,Z. Rong, K. R. Zavadil, K. Persson, and G. Ceder, Chem.Mater. 27, 3317 (2015), URL http://dx.doi.org/10.

1021/acs.chemmater.5b00389.8 Z. Rong, R. Malik, P. Canepa, G. Sai Gautam, M. Liu,

A. Jain, K. Persson, and G. Ceder, Chem. Mater. 27,

6016 (2015), ISSN 1520-5002, URL http://dx.doi.org/

10.1021/acs.chemmater.5b02342.9 I. Brown, Acta Crystallogr. Sect. B 44, 545 (1988), URLhttp://dx.doi.org/10.1107/S0108768188007712.

10 G. Sai Gautam, P. Canepa, R. Malik, M. Liu,K. Persson, and G. Ceder, Chem. Commun. 51, 13619(2015), ISSN 1364-548X, URL http://dx.doi.org/10.

1039/C5CC04947D.11 M. Liu, Z. Rong, R. Malik, P. Canepa, A. Jain,

G. Ceder, and K. A. Persson, Energy Environ. Sci. 8,964 (2015), ISSN 1754-5706, URL http://dx.doi.org/

10.1039/C4EE03389B.12 M. Liu, A. Jain, Z. Rong, X. Qu, P. Canepa, R. Malik,

G. Ceder, and K. Persson, Energy Environ. Sci. 9, 3201(2016), URL http://dx.doi.org/10.1039/C6EE01731B.

13 M. S. Whittingham, Chem. Rev. 104, 4271 (2004),ISSN 1520-6890, URL http://dx.doi.org/10.1021/

cr020731c.14 M. Thackeray, P. Johnson, L. de Picciotto, P. Bruce,

and J. Goodenough, Mater. Res. Bull. 19, 179 (1984),ISSN 0025-5408, URL http://dx.doi.org/10.1016/

0025-5408(84)90088-6.15 M. Wagemaker, A. Van Der Ven, D. Morgan, G. Ceder,

F. Mulder, and G. Kearley, Chemical Physics 317, 130(2005).

16 J. Reed, G. Ceder, and A. V. D. Ven, Electrochem. Solid-State Lett. 4, A78 (2001), URL http://dx.doi.org/10.

1149/1.1368896.17 A. Van der Ven, C. Marianetti, D. Morgan, and G. Ceder,

Solid State Ionics 135, 21 (2000).18 E. Ferg, R. Gummow, A. De Kock, and M. Thackeray, J.

Electrochem. Soc. 141, L147 (1994).

16

19 D. Andre, S.-J. Kim, P. Lamp, S. F. Lux, F. Maglia,O. Paschos, and B. Stiaszny, J. Mater. Chem. A 3, 6709(2015), ISSN 2050-7496, URL http://dx.doi.org/10.

1039/C5TA00361J.20 G. Gershinsky, H. D. Yoo, Y. Gofer, and D. Aurbach,

Langmuir 29, 10964 (2013), ISSN 1520-5827, URL http:

//dx.doi.org/10.1021/la402391f.21 G. Sai Gautam, P. Canepa, A. Abdellahi, A. Urban,

R. Malik, and G. Ceder, Chem. Mater. 27, 3733 (2015),ISSN 1520-5002, URL http://dx.doi.org/10.1021/acs.

chemmater.5b00957.22 G. Sai Gautam, P. Canepa, W. D. Richards, R. Malik,

and G. Ceder, Nano Lett. 16, 2426 (2016), URL http:

//dx.doi.org/10.1021/acs.nanolett.5b05273.23 N. Sa, T. L. Kinnibrugh, H. Wang, G. S. Gautam,

K. W. Chapman, J. T. Vaughey, B. Key, T. T. Fis-ter, J. W. Freeland, D. L. Proffit, et al., Chem. Mater.28, 2962 (2016), URL http://dx.doi.org/10.1021/acs.

chemmater.6b00026.24 S. Tepavcevic, Y. Liu, D. Zhou, B. Lai, J. Maser, X. Zuo,

H. Chan, P. Kral, C. S. Johnson, V. Stamenkovic, et al.,ACS Nano 9, 8194 (2015), ISSN 1936-086X, URL http:

//dx.doi.org/10.1021/acsnano.5b02450.25 P. Novak, V. Shklover, and R. Nesper, Z. Phys. Chem.

185, 51 (1994), URL http://dx.doi.org/10.1524/zpch.

1994.185.Part_1.051.26 C. Kim, P. J. Phillips, B. Key, T. Yi, D. Nordlund, Y.-

S. Yu, R. D. Bayliss, S.-D. Han, M. He, Z. Zhang, et al.,Adv. Mater. 27, 3377 (2015), ISSN 0935-9648, URL http:

//dx.doi.org/10.1002/adma.201500083.27 Z. Feng, X. Chen, L. Qiao, A. L. Lipson, T. T. Fister,

L. Zeng, C. Kim, T. Yi, N. Sa, D. L. Proffit, et al., ACSAppl. Mater. Interfaces 7, 28438 (2015), URL http://dx.

doi.org/10.1021/acsami.5b09346.28 S. Kim, K. W. Nam, S. Lee, W. Cho, J.-S. Kim, B. G.

Kim, Y. Oshima, J.-S. Kim, S.-G. Doo, H. Chang, et al.,Angew. Chem. Int. Ed. 54, 15094 (2015), ISSN 1433-7851,URL http://dx.doi.org/10.1002/anie.201505487.

29 K. Irani, A. Sinha, and A. Biswas, J. Phys. Chem.Solids 23, 711 (1962), URL http://dx.doi.org/10.1016/

0022-3697(62)90530-9.30 L. Malavasi, P. Ghigna, G. Chiodelli, G. Maggi,

and G. Flor, J. Solid State Chem. 166, 171 (2002),ISSN 0022-4596, URL http://dx.doi.org/10.1006/

jssc.2002.9577.31 M. Rosenberg and P. Nicolau, Phys. Stat. Sol. 6, 101

(1964).32 R. Manaila and P. Pausescu, Phys. Stat. Sol. 9, 385 (1965).33 N. Radhakrishnan and A. Biswas, Phys. Stat. Sol. (a) 37,

719 (1976).34 J. Reed and G. Ceder, Chem Rev 104, 4513 (2004).35 J. Lee, A. Urban, X. Li, D. Su, G. Hautier, and G. Ceder,

Science 343, 519 (2014).36 A. Urban, J. Lee, and G. Ceder, Adv. Energy Mater. 4,

1400478 (2014).37 Y. Seminovski, P. Palacios, P. Wahnon, and R. Grau-

Crespo, Appl. Phys. Lett. 100, 102112 (2012).38 P. Canepa, S.-H. Bo, G. S. Gautam, B. Key, W. D.

Richards, Y. Wang, J. Li, and G. Ceder, under review(2017).

39 H. Hahn and W. Klingler, Z. Anorg. Allg. Chem. 263, 177(1950).

40 M. Wakaki, O. Shintani, T. Ogawa, and T. Arai, Jpn. J.Appl. Phys. 19, 255 (1980).

41 J. C. Bachman, S. Muy, A. Grimaud, H.-H. Chang,N. Pour, S. F. Lux, O. Paschos, F. Maglia, S. Lupart,P. Lamp, et al., Chem. Rev. 116, 140 (2016), URL http:

//dx.doi.org/10.1021/acs.chemrev.5b00563.42 A. Walsh, S.-H. Wei, Y. Yan, M. Al-Jassim, J. A. Turner,

M. Woodhouse, and B. Parkinson, Phys. Rev. B 76, 165119(2007).

43 D. Das and S. Ghosh, J. Phys.: Condens. Matter 29,055805 (2016).

44 L. Schwarz, Z. Galazka, T. M. Gesing, and D. Klimm,Cryst. Res. Technol. 50, 961 (2015).

45 D. Santos-Carballal, A. Roldan, R. Grau-Crespo, andN. H. de Leeuw, Phys. Rev. B 91, 195106 (2015).

46 B. Xu and S. Meng, J. Power Sources 195, 4971 (2010).47 G. S. Gautam, X. Sun, V. Duffort, L. F. Nazar, and

G. Ceder, J. Mater. Chem. A 4, 17643 (2016), URLhttp://dx.doi.org/10.1039/C6TA07804D.

48 K. E. Sickafus, J. M. Wills, and N. W. Grimes, J. Am.Ceram. Soc. 82, 3279 (1999).

49 L. Pauling, The nature of the chemical bond and the struc-ture of molecules and crystals: an introduction to mod-ern structural chemistry, vol. 18 (Cornell university press,1960).

50 P. P. Ewald, Ann. Phys. 369, 253 (1921).51 D. Stauffer and A. Aharony, Introduction to percolation

theory (CRC press, 1994).52 A. Hunt and R. Ewing, Percolation theory for flow in

porous media (Lecture Notes in Physics 674) (Soil Sci SocAmerica, 2006).

53 M. B. Isichenko, Rev Mod Phys 64, 961 (1992).54 J. W. Essam, Rep. Prog. Phys. 43, 833 (1980).55 S. C. Van der Marck, Int J Mod Phys C 9, 529 (1998).56 C. D. Lorenz and R. M. Ziff, Phys Rev E 57, 230 (1998).57 A. Urban, D.-H. Seo, and G. Ceder, npj Comput.

Mater. 2, 16002 (2016), URL http://dx.doi.org/10.

1038/npjcompumats.2016.2.58 P. Hohenberg and W. Kohn, Phys Rev 136, B864 (1964).59 W. Kohn and L. J. Sham, Phys Rev 140, A1133 (1965).60 A. Jain, Y. Shin, and K. A. Persson, Nature Reviews Ma-

terials 1, 15004 (2016).61 G. Kresse and J. Hafner, Phys. Rev. B 47, 558 (1993).62 G. Kresse and J. Furthmuller, Phys Rev B 54, 11169

(1996).63 G. Kresse and D. Joubert, Phys. Rev. B 59, 1758 (1999).64 J. P. Perdew, K. Burke, and M. Ernzerhof, Phys Rev Lett

77, 3865 (1996).65 N. Sanjana, A. Biswas, and A. Sinha, J. Sci. Ind. Res. B

19, 415 (1960).66 J. P. Perdew, A. Ruzsinszky, G. I. Csonka, O. A. Vydrov,

G. E. Scuseria, L. A. Constantin, X. Zhou, and K. Burke,Phys. Rev. Lett. 100, 136406 (2008).

67 D. A. Kitchaev, H. Peng, Y. Liu, J. Sun, J. P. Perdew, andG. Ceder, Phys. Rev. B 93, 045132 (2016), URL http:

//dx.doi.org/10.1103/PhysRevB.93.045132.68 V. I. Anisimov, J. Zaanen, and O. K. Andersen, Phys. Rev.

B 44, 943 (1991).69 F. Zhou, M. Cococcioni, C. A. Marianetti, D. Morgan,

and G. Ceder, Physical Review B 70, 235121 (2004), URLhttp://dx.doi.org/10.1103/PhysRevB.70.235121.

70 A. Jain, G. Hautier, S. P. Ong, C. J. Moore, C. C. Fischer,K. A. Persson, and G. Ceder, Phys. Rev. B 84, 045115(2011).

71 G. Henkelman and H. Jonsson, J. Chem. Phys. 113, 9978(2000).

17

72 D. Sheppard, R. Terrell, and G. Henkelman, J. Chem.Phys. 128, 134106 (2008).

73 M. Newman and R. Ziff, Phys. Rev. Lett. 85, 4104 (2000).74 X. Sun, V. Duffort, P. Bonnick, Z. Rong, M. Liu,

K. Persson, G. Ceder, and L. F. Nazar, Energy Environ.Sci. 9, 2273 (2016), URL http://dx.doi.org/10.1039/

C6EE00724D.75 C. Ling, R. Zhang, and F. Mizuno, ACS Appl. Mater. In-

terfaces 8, 4508 (2016).76 Y. Wang, W. D. Richards, S. P. Ong, L. J. Miara, J. C.

Kim, Y. Mo, and G. Ceder, Nat. Mater. 14, 1026 (2015),URL http://dx.doi.org/10.1038/nmat4369.

77 M. K. Aydinol, A. F. Kohan, G. Ceder, K. Cho, andJ. Joannopoulos, Phys. Rev. B 56, 1354 (1997), URLhttp://dx.doi.org/10.1103/PhysRevB.56.1354.

78 A. Jain, S. P. Ong, G. Hautier, W. Chen, W. D.Richards, S. Dacek, S. Cholia, D. Gunter, D. Skinner,G. Ceder, et al., APL Materials 1, 011002 (2013), ISSN2166532X, URL http://link.aip.org/link/AMPADS/v1/

i1/p011002/s1&Agg=doi.79 S. Bagcı, H. Tutuncu, S. Duman, E. Bulut, M. Ozacar,

and G. Srivastava, J. Phys. Chem. Solids 75, 463 (2014).80 C. Ling and F. Mizuno, Chem. Mater. 25, 3062

(2013), ISSN 1520-5002, URL http://dx.doi.org/10.

1021/cm401250c.81 X. Sun, V. Duffort, B. L. Mehdi, N. D. Browning, and

L. F. Nazar, Chem. Mater. 28, 534 (2016), ISSN 1520-5002, URL http://dx.doi.org/10.1021/acs.chemmater.

5b03983.