targeting superionic conductivity by turning on anion

19
Article Targeting Superionic Conductivity by Turning on Anion Rotation at Room Temperature in Fast Ion Conductors A combination of the maximum entropy method and AIMD simulations demonstrates that polyanion [PS 4 ] 3 rotation is facile in the fast ion conductors b-Li 3 PS 4 and its Si-substituted analog, Li 3.25 Si 0.25 P 0.75 S 4 , but absent in the non- conductive phase, g-Li 3 PS 4 . The increased entropy upon the substitution of Si for P stabilizes the high-temperature rotor phase (b-Li 3 PS 4 ) at room temperature. Joint- time correlation analysis and AIMD simulations show that [PS 4 ]/[SiS 4 ] anion rotational dynamics are coupled to and greatly enhance cation diffusion by widening the bottleneck for Li + -ion transport. Zhizhen Zhang, Hui Li, Kavish Kaup, Laidong Zhou, Pierre-Nicholas Roy, Linda F. Nazar [email protected] HIGHLIGHTS Polyanion rotation occurs in conductive b-Li 3 PS 4 , Li 3.25 Si 0.25 P 0.75 S 4 ; not in g-Li 3 PS 4 Facile anion rotational dynamics revealed by combination of MEM and AIMD studies Si doping stabilizes the b-Li 3 PS 4 rotor phase to room temperature by entropic effects Joint-time correlation analysis / anion dynamics coupled to (enhanced) cation mobility Zhang et al., Matter 2, 1667–1684 June 3, 2020 ª 2020 Elsevier Inc. https://doi.org/10.1016/j.matt.2020.04.027 ll

Upload: others

Post on 20-Jul-2022

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Targeting Superionic Conductivity by Turning on Anion

ll

Article

Targeting Superionic Conductivity by Turningon Anion Rotation at Room Temperature inFast Ion Conductors

Zhizhen Zhang, Hui Li, Kavish

Kaup, Laidong Zhou,

Pierre-Nicholas Roy, Linda F.

Nazar

[email protected]

HIGHLIGHTS

Polyanion rotation occurs in

conductive b-Li3PS4,

Li3.25Si0.25P0.75S4; not in g-Li3PS4

Facile anion rotational dynamics

revealed by combination of MEM

and AIMD studies

Si doping stabilizes the b-Li3PS4rotor phase to room temperature

by entropic effects

Joint-time correlation analysis /

anion dynamics coupled to

(enhanced) cation mobility

A combination of the maximum entropy method and AIMD simulations

demonstrates that polyanion [PS4]3� rotation is facile in the fast ion conductors

b-Li3PS4 and its Si-substituted analog, Li3.25Si0.25P0.75S4, but absent in the non-

conductive phase, g-Li3PS4. The increased entropy upon the substitution of Si for P

stabilizes the high-temperature rotor phase (b-Li3PS4) at room temperature. Joint-

time correlation analysis and AIMD simulations show that [PS4]/[SiS4] anion

rotational dynamics are coupled to and greatly enhance cation diffusion by

widening the bottleneck for Li+-ion transport.

Zhang et al., Matter 2, 1667–1684

June 3, 2020 ª 2020 Elsevier Inc.

https://doi.org/10.1016/j.matt.2020.04.027

Page 2: Targeting Superionic Conductivity by Turning on Anion

ll

Article

Targeting Superionic Conductivityby Turning on Anion Rotation at RoomTemperature in Fast Ion Conductors

Zhizhen Zhang,1,2 Hui Li,1,3 Kavish Kaup,1,2 Laidong Zhou,1,2 Pierre-Nicholas Roy,1

and Linda F. Nazar1,2,4,*

Progress and Potential

Solid-state Li-ion batteries, with

their integration of solid

electrolytes, have emerged as

alternatives to traditional liquid

electrolyte Li-ion batteries. They

have the potential to provide

higher energy density, longer

cycle life, and better safety for

large-scale electrochemical

energy-storage applications such

as electric vehicles. The

fundamental understanding of

how to achieve the necessary

superionic conduction in solid

electrolytes is of significant

importance to battery

development. While previous

studies mainly focused on factors

related to how the static

crystalline structure affects Li-

cation mobility, potential impact

from anion dynamics has been

largely neglected. Here, we

combine experimental and

computational studies to reveal

that anion dynamics can be

manipulated by the entropy-

driven stabilization of a high-

temperature rotor phase to room

temperature and show that

polyanion rotational motion in

lithium thiophosphates is coupled

to, and enhances, Li-ion diffusion.

SUMMARY

Tremendous interest in solid-state devices has prompted the devel-opment of descriptors that can be utilized to develop better mate-rials. Previous work mostly focused on the static structure, while dis-covery of the interplay between cation mobility and anion dynamicscan also provide a powerful descriptor. Here, we demonstrate that[PS4] rotation is remarkably facile in fast ion conductors b-Li3PS4 andits Si-substituted analog, Li3.25Si0.25P0.75S4, whereas it is absent inthe non-conductive phase, g-Li3PS4. Phonon calculations identifythe increased entropy upon substitution of Si for P as the origin ofthe stabilization of the high-temperature phase (b-Li3PS4) at roomtemperature. Joint-time correlation analysis and power spectrareveal that [PS4]/[SiS4] anion rotational dynamics are coupled toand greatly enhance cation diffusion via the paddle-wheel effectby widening the bottleneck for Li+-ion transport. These findingsshed light on the critical role of anion dynamics and can serve as ageneral guideline for the design of new conductors.

INTRODUCTION

Driven by the ever-increasing demand for advanced energy storage, technological ad-

vances in rechargeablebatteries are eagerly sought. Solid-state batteries are considered

among the most promising next-generation technologies as they offer the opportunity

for excellent safety, long life cycle, and high-energy density with the use of high-voltage

cathode materials, a fast-ion conducting solid electrolyte, and a Li metal anode.1–3 The

prospects of developing solid-state batteries based on fast ion conductors have moti-

vated tremendous endeavors from many researchers in this field. Practical fast ion

conducting candidates demand a high ion conductivity (R10�3 S cm�1 at working tem-

peratures).4 Among the choices of solid electrolytes, lithium thiophosphates exhibit

excellent ion conductive properties combined with good ductility. Prominent examples

include Li10MP2S12 (M =Ge, Sn),5,6 Li7P3S11,7,8 and Li6PS5X argyrodites (X = Cl, Br, I).9,10

Considerable enhancement in ion conductivity has been achieved through fundamental

principles, including volume effect,4 creating anion11 and cation disorder,12–14 invoking

cooperativemigration via cationconcentrationmanipulation,15,16 tuning theanion lattice

polarizability,9 and framework softness.9,17 While such studies have focused on the rele-

vance of the static structure to cation mobility, exploiting the anion dynamics

(reorientation/rotation) of the complex polyanions is of equal importance, which has

been relatively less explored.

It was proposed long ago that in structures containing polyanions with covalent

bonds (e.g., NO2�,18 PO4

3�,18,19 SO42�20), the rotational/reorientational motion

Matter 2, 1667–1684, June 3, 2020 ª 2020 Elsevier Inc. 1667

Page 3: Targeting Superionic Conductivity by Turning on Anion

llArticle

of these polyanions may improve cationic conductivity and decrease the associated

activation energy via the ‘‘paddle-wheel’’ mechanism.18 These phenomena were

observed to occur upon the order-disorder transformation into the superionic

high-temperature (HT) phases (rotor phases), while the materials transform revers-

ibly to the poorly conducting form upon a decrease in temperature. Most recently,

dynamic motion has been demonstrated in borohydride anions (BH4�,21 B10H10

2�,22

B12H122�,23,24 CB9H10

�,25 CB11H12�,26 and B11H14

�27) owing to the combined utili-

zation of advanced techniques such as quasi-elastic neutron scattering and ab initio

simulations. However, such anion reorientation mostly occurs at elevated tempera-

ture (with the exception of NaCB9H10,25 NaB11H14

�,27 0.7Li(CB9H10)-0.3Li

(CB11H12),28 iodide substituted LiBH4

29, and 75Li2S–25P2S5 glass30 and thus has

not received broad attention to date. Extending this concept to other, more general,

systems and stabilizing the rotor phase at or below room temperature—concomitant

with high conductivity—is thus highly anticipated.

The tendency of a complex anion to exhibit rotational disorder in an ionic crystal de-

pends on the crystal structure: energetically equivalent orientations for the polyan-

ion and a high polarizability of the static framework are necessary. Very recently, we

and other groups30–33 simultaneously reported on four [PS4] polyanion frameworks

that show a different tendency toward anion rotational disorder. Direct evidence

of anion rotational dynamics and strong theoretical support showing how the anion

dynamics enhance cation diffusion were provided in our work.25

Here, we report that anion rotation can be induced by entropy-driven stabilization of

the polyanion rotor phase down to room temperature by manipulating the anion lat-

tice. We have selected crystalline Li3PS4 as a model to demonstrate this concept,

although the strategy discussed here is broadly applicable to other materials.

Partially substituting P5+ in Li3PS4 with (Si4+ + Li+) stabilizes a superionic rotor phase

at room temperature, Li3.25[Si0.25P0.75]S4, which has an expanded volume similar to

that of the HT rotor phase, superionic b-Li3PS4. Neutron powder diffraction (NPD)

studies enable determination of the rotational disorder of the [PnS4] (Pn = P; Si) poly-

anion from an experimental perspective via the maximum entropy method (MEM),

while ab initio molecular dynamics (AIMD) simulations provide a comprehensive

probe of the anion trajectories, free energy landscape, and the coupled motion be-

tween cations and anions. Analysis of the bottlenecks to cation transport reveals that

[PS4]/[SiS4] anion rotation facilitates the Li+-ion diffusion by transiently opening up

the window for Li+-ion transport and thereby decreasing the Li+-ion migration bar-

rier. Turning on polyanion rotational dynamics is directly correlated with enhanced

cation diffusivity in the lattice, as suggested by our joint-time correlation function

analysis. Our findings elucidate that manipulating the reorientational/rotational

mobility of the polyanions influences the cation diffusivity and serves as a vital com-

plementary descriptor for the design of highly conductive fast ion conductors.

1Department of Chemistry, and the WaterlooInstitute of Nanotechnology, University ofWaterloo, Waterloo, ON N2L 3G1, Canada

2Joint Center for Energy Storage Research,Argonne National Laboratory, Argonne, IL 60439,USA

3Institute of Theoretical Chemistry, College ofChemistry, Jilin University, 2519 Jiefang Road,Changchun 130023, China

4Lead Contact

*Correspondence: [email protected]

https://doi.org/10.1016/j.matt.2020.04.027

RESULTS AND DISCUSSION

Cooperative Cation Diffusion Mechanism from AIMD Simulations

Li3PS4 exists as the thermodynamically stable low-temperature (LT) g phase at room

temperature (space group: Pmn21; a = 7.70729(11) A, b = 6.53521(10) A, c =

6.1365(7) A, V = 307.47(7) A3, Z = 2),34 which exhibits very low ion conductivity of

3 3 10�7 S$cm�1.35 It converts to the highly conductive HT phase b-Li3PS4 at

195�C (space group: Pnma; a = 12.8190(5) A, b = 8.2195(4) A, c = 6.1236(2) A,

V = 645.23(5) A3, Z = 4),34 which exhibits a conductivity of �10�3 S$cm�1 at

200�C.35 The extrapolated conductivity of b-Li3PS4 to room temperature is reported

1668 Matter 2, 1667–1684, June 3, 2020

Page 4: Targeting Superionic Conductivity by Turning on Anion

Figure 1. Polyanion Arrangement, Cation Diffusion Pathways, DiffusionMechanism, and Activation Energy in g-Li3PS4, b-Li3PS4, and Li3.25[Si0.25P0.75]

S4 from AIMD Simulations

(A) Arrangement of the PS4 polyanion in g-Li3PS4 in a 2a2b2c super cell. The arrangement of the [PS4] tetrahedron apices follows a T+ arrangement along

the c axis, where T+ (T�) denotes that the apex of the [PS4] tetrahedron is up (down); Li cation sublattice is also in an ordered manner.

(B) Arrangement of the PS4 polyanion complex in b-Li3PS4 in a 1.5a2b2c supercell. The [PS4]3� tetrahedra form an alternating T+ and T� arrangement.

Pink (with cross) and yellow spheres represent Li and S atoms, respectively. Green tetrahedra represent [PS4]3� anions.

(C and D) The Li+-ion diffusion pathway in b-Li3PS4 in a 1a2b2c supercell from AIMD at 1,050 K: (C) along the b axis; (D) in ac plane. b-Li3PS4 is a two-

dimensional (2D) conductor. Li+ ions transport mainly along the [101] direction in the ac plane, and negligible pathways are observed along the b axis.

(E and F) The Li+-ion diffusion pathway in Li3.25[Si0.25P0.75]S4 in a 1a2b2c supercell from AIMD at 1,050 K: (E) along the b axis; (F) in the ac plane.

Li3.25[Si0.25P0.75]S4 exhibits three-dimensional (3D) pathways and Li+-ion diffusivity exhibits a quasi-anisotropic nature.

(G) The distinctive part of the van Hove function obtained from AIMD simulations for Li3.25[Si0.25P0.75]S4 at 1,050 K;Gd(r, t) reflects the probability density

to find a different Li cation at a distance r and at time t from a place where at time t = 0 there was a Li cation.

(H) Calculated diffusivities from AIMD simulations at elevated temperatures (600 K, 750 K, 900 K, 1,050 K, and 1,200 K) in b-Li3PS4 and Si-doped

Li3.25[Si0.25P0.75]S4. Activation energies of 0.35 eV and 0.30 eV were extracted by fitting the diffusivities for b-Li3PS4 and Li3.25[Si0.25P0.75]S4, respectively.

llArticle

to be 8.9 3 10�7 S$cm�1.35,36 Nevertheless, b-Li3PS4 can be stabilized at room tem-

perature in a nanoporous form with ion conductivity of 1.63 10�4 S$cm�1, via a syn-

thetic route that utilizes a wet-chemical method (a tetrahydrofuran [THF] complex).36

The significant enhancement of conductivity by �2 orders of magnitude compared

with bulk b-Li3PS4 was claimed to arise from surface conduction induced by the

porous structure.36 Significant amounts of hydrogen in the material are also indi-

cated by a high background in NPD patterns, possibly from residual THF.37

The polyanions in the two modifications are arranged differently. The [PS4] tetra-

hedra in g-Li3PS4 feature a T+ arrangement along the c axis, where T+ (T�) denotesthat the apex of the [PS4] tetrahedron is up (down) (Figure 1A); while in b-Li3PS4, the

[PS4] tetrahedral arrangement is disordered and follows a zigzag arrangement (Fig-

ure 1B).34 The distinction in the [PS4] polyanion arrangement results in quite different

Li sites and occupancies in these two polymorphs: in the g phase, Li fully occupies

the tetrahedral sites (Wyckoff notation: 4b and 2a), and the apex of the [LiS4] tetra-

hedron follows the same ordering as that of [PS4] tetrahedron; in the b phase, the

zigzag arrangement of [PS4] polyanions allows Li to partition in both octahedral

(8d, fully occupied) and tetrahedral sites (4b and 4c, partially occupied).34 Our

Matter 2, 1667–1684, June 3, 2020 1669

Page 5: Targeting Superionic Conductivity by Turning on Anion

llArticle

previous study revealed that the arrangement of [PS4] polyanions in the b form can

be stabilized at room temperature by partial substitution of Si at the P site, redistrib-

uting Li atoms across five sites via splitting of both the 8d and 4c sites and resulting in

a room temperature conductivity of 1.23 10�3 S$cm�1 for the optimal composition,

Li3.25[Si0.25P0.75]S4 (details are the subject of a separate publication).38 This repre-

sents an increase of three orders of magnitude higher conductivity compared with

(the extrapolated value for) bulk b-Li3PS4 and one order of magnitude higher than

nanoporous b-Li3PS4.

To study the intrinsic reason for the enhanced conductivity caused by the incorpora-

tion of Si, we performed AIMD simulations on both b-Li3PS4 and Si-substituted

Li3.25[Si0.25P0.75]S4 at five elevated temperatures (600, 750, 900, 1,050, and 1,200

K) with a 1a2b2c supercell. This selected composition shows the best ion conductiv-

ity among the wide range of compositions we visited by AC impedance measure-

ments. The Li+-ion probability densities extracted from the trajectory at 1,050 K

for b-Li3PS4 and b-Li3.25[Si0.25P0.75]S4 are shown in Figures 1C and 1D and Figures

1E and 1F, respectively. These data confirm that Li+-ion transport in b-Li3PS4 is aniso-

tropic,37 in good agreement with a previous MEM study. The major pathways are

along the [101] direction in the ac plane (Figure 1D), while no connected channels

are available along the [010] direction (Figure 1C). In contrast, in the Si-substituted

material, Li+ ions diffuse quasi-isotropically in the three-dimensional (3D) framework:

namely, major pathways are observed along both the [101] and [�101] directions in

the ac plane (Figure 1E) and along the [010] direction (Figure 1F). Macroscopically,

the conduction mechanism of Li+ ions can be probed by computing the van Hove

correlation function (see Computational Details). The distinct-part Gd(r, t) is a real-

space radial distribution function that characterizes the spatial and temporal distri-

butions of pairs of lithium ions in the lattice. The high probability of one Li site being

occupied by another Li cation over a very short time duration (1–2 ps) (Figures 1G

and S1) confirms that Li cations migrate through correlated migration in

Li3.25[Si0.25P0.75]S4. Diffusion coefficients were obtained from the slope of the time

dependence of the cation mean-square displacements using the Einstein relation

(for details, see Computational Details). Li3.25[Si0.25P0.75]S4 exhibits higher diffusion

coefficients than b-Li3PS4 at all equivalent temperatures except 1,200 K, where the

values are almost identical (1.46 3 10�8 m2 s�1 for Li3.25[Si0.25P0.75]S4 versus

1.48 3 10�8 m2 s�1 for b-Li3PS4). A linear fit to the Arrhenius plot of diffusion coef-

ficients derived from AIMD yields an activation energy barrier of 0.30 eV for Li+-

ion diffusion in Li3.25[Si0.25P0.75]S4 (Figure 1H), which is in very good accord with

our experimental value of 0.31 eV from AC impedance studies.38 The predicted acti-

vation energy for Li+-ion diffusion in b-Li3PS4 is higher—0.35 eV—and is consistent

with the value of 0.36 eV for nanoporous b-Li3PS4, while Ea is reported to be 0.47

eV for bulk Li3PS4.35

Facile [PnS4] Polyanion Reorientational Disorder in b-Li3PS4 and

Li3.25[Si0.25P0.75]S4 via the Maximum Entropy Method

We first track the evolution of the lattice dynamics of the three structures, g-Li3PS4,

b-Li3PS4, and Li3.25[Si0.25P0.75]S4, from room temperature to elevated temperature

(350�C) using time-of-flight neutron diffraction studies (Figures 2A–2K). The

maximum entropy method (MEM) was employed to analyze the nuclear density dis-

tribution from the NPD data. The local structures of the polyanions in these three

compositions are displayed in Figures 2A, 2D, and 2G as a reference. Figures 2B

and 2C show the two-dimensional (2D) contour slice of the (001) plane at

z = 0.405 at data-collection temperatures of 30�C and 200�C for g-Li3PS4, respec-

tively. These two maps explicitly show that nuclear density of the [PS4] anion groups

1670 Matter 2, 1667–1684, June 3, 2020

Page 6: Targeting Superionic Conductivity by Turning on Anion

Figure 2. MEM Plots of g-Li3PS4, b-Li3PS4, and Li3.25[Si0.25P0.75]S4 from Variable Temperature Neutron Powder Diffraction

(A) Crystal structure of g-Li3PS4 (Pmn21) through the (001) plane, z = 0.30–0.50; [PS4] tetrahedra are denoted in green.

(B and C) Views of MEM plots of g-Li3PS4: (B) at 30�C through the (001) plane, z = 0.405; (C) at 200�C through the (001) plane, z = 0.405.

(D) Crystal structure of b-Li3PS4 (Pnma) through the (001) plane, z = 0.15–0.35.

(E and F) Views of MEM plots of b-Li3PS4 at 350�C: (E) through the (001) plane, z = 0.275; (F) through the (100) plane, x = 0.325.

(G) Crystal structure of Li3.25[Si0.25P0.75]S4 (Pnma) through (001) plane, z = 0.15–0.35. [SiS4]/[PS4] tetrahedra are denoted in green.

(H–K) Views of MEM plots of Li3.25[Si0.25P0.75]S4 through the (001) plane, z = 0.275 at (H) 30�C and (I) 300�C; through the (100) plane, x = 0.35 at (J) 30�Cand (K) 300�C.

llArticle

in g-Li3PS4 is highly localized (the red areas) at both room temperature and elevated

temperature (200�C), suggesting that no anion disorder and, hence, no anion reor-

ientation occurs in LT g-Li3PS4. Since the phase transformation between g-Li3PS4 and

b-Li3PS4 starts around 200�C, we carried out MEM analysis at 350�C, where the trans-formation to the b-Li3PS4 phase is complete. Unlike the polyanions in g-Li3PS4, the

[PS4] anions in b-Li3PS4 show a very dispersed nuclear density distribution, indicative

of unequivocal anion rotational disorder at 350�C in the ab plane (Figure 2E), and

minor rotational disorder in the bc plane (Figure 2F). Intriguingly, the [PS4]/[SiS4] an-

ions in Li3.25[Si0.25P0.75]S4 show prominent reorientational disorder at room temper-

ature (Figures 2H and 2J). In addition to the striking reorientational disorder in the ab

plane (Figures 2H and 2I), the [PS4]/[SiS4] moieties show more notable rotational dis-

order in the bc plane (Figures 2J and 2K). Upon heating to 300�C, the reorientation

of [PS4]/[SiS4] anions is even more pronounced, as evident from the more dispersed

and smeared nuclear density distribution (Figures 2I and 2K). These observations

imply that Si substitution shifts the highly reorientational nature of the HT b poly-

morph (rotor phase) to room temperature.

Entropy-Driven Stabilization of the Rotor Phase below or at Room

Temperature

To understand the driving force for the stabilization of the HT rotor phase down to

room temperature by Si substitution, we computed the contribution of phonons

to the entropy of the three phases (LT g-Li3PS4, HT b-Li3PS4, and Li3.25[Si0.25P0.75]

S4). As shown in Figure 3A, the entropy values of these three structures follow the

Matter 2, 1667–1684, June 3, 2020 1671

Page 7: Targeting Superionic Conductivity by Turning on Anion

Figure 3. Thermodynamic Properties of g-Li3PS4, b-Li3PS4, and Li3.25[Si0.25P0.75]S4 from Phonon Calculations

(A) Entropy.

(B) Helmholtz free energies at temperature range between 0 and 600 K.

(C) Phonon density of states. More pronounced phonon densities for b-Li3PS4 and Li3.25[Si0.25P0.75]S4 are evidenced in the low-frequency region, as

indicated by the pink triangular region.

llArticle

trend of g-Li3PS4 < b-Li3PS4 < Li3.25[Si0.25P0.75]S4 in the temperature range studied

(0–600 K). The calculated Helmholtz free energies (F = U� TS, where U is the internal

energy, T is the temperature in Kelvin, and S is the entropy), otherwise show

the trend of g-Li3PS4 > Li3.25[Si0.25P0.75]S4 > b-Li3PS4 between 0 and 330 K and

g-Li3PS4 > b-Li3PS4 > Li3.25[Si0.25P0.75]S4 between 330 and 600 K (Figure 3B).These findings suggest that the higher entropy of the rotor phases b-Li3PS4 and

Li3.25[Si0.25P0.75]S4 contributes more to the lower free energy of these two systems

than the LT phase g-Li3PS4. This demonstrates that Si doping stabilizes the rotor

phase at or below room temperature by entropy-driven effects. The LT phase

g-Li3PS4 shows a different phonon density of states (PDOS) (Figure 3C) compared

with the rotor phases (HT phase b-Li3PS4 and Li3.25[Si0.25P0.75]S4). Particularly,

more pronounced PDOS peaks appear at low frequency (below 3 THz) in the HT

phase b-Li3PS4 and Li3.25[Si0.25P0.75]S4 than in the LT phase g-Li3PS4. These modes

are a characteristic feature of the disorder in systems39 and are attributed to the

notable reorientational/rotational dynamics of the disordered polyanions in

b-Li3PS4 and Li3.25[Si0.25P0.75]S4. A similar phenomenon was observed in the PDOS

of HT versus LT LiBH4 in a previous study.40

Helmholtz Free Energy Potential for Anion Rotation

To gain a more comprehensive understanding of the anion reorientational/rota-

tional disorder in Li3.25[Si0.25P0.75]S4, we performed a detailed analysis of the

AIMD data. The preferred reorientation of the [PS4]/[SiS4] polyanions is reflected

by the spatial distribution of the P/Si, and sulfur (S) atoms that are extracted from

the trajectories of these species. Figure 4A shows the 3D probability distribution

of [PS4]/[SiS4] anion groups within the reference frame of the Li3.25[Si0.25P0.75]S4 crys-

tal lattice. It is evident that the [PS4]/[SiS4] anions exhibit significant reorientation/

rotational dynamics: several more maxima can be observed aside from the original

four static positions of the four S2� ligands. The S ligands of both the [PS4] and [SiS4]

anion groups exhibit dispersed spatial probability density, although the former is

slightly more uniform. The 2D distribution of the S ligands of the [PS4]/[SiS4] motifs

are obtained by mapping the S ligands in the spherical coordinates defined in Fig-

ure 4B. As shown in Figures 4C and 4D, more than four maxima are observed for the

four S ligands bonded to P/Si (denoted by a, b, c, d, e, f.), suggesting that facile

anion reorientation/rotational dynamics occur at this temperature within the

1672 Matter 2, 1667–1684, June 3, 2020

Page 8: Targeting Superionic Conductivity by Turning on Anion

Figure 4. 3D and 2D Density Distribution and Helmholtz Free Energy Surface of S Ligands in

Li3.25[Si0.25P0.75]S4(A) 3D probability distribution of [PnS4] (Pn = Si/P) in Li3.25[Si0.25P0.75]S4 from AIMD simulations at

1,200 K.

(B) Definition of the angles q and f in the reference frame of the crystal lattice, with q defined as the

angle between the P/Si–S bond and z axis and the angle f corresponding to the angle between the

x axis and the projection of the P/Si–S vector in the xy plane.

(C and D) 2D projected probability distribution of S ligand bonded to (C) P4 and (D) Si3.

(E and F) Helmholtz free energy surface of S ligands as a function of angle q and f in (E) P4 and (F) Si3.

The free energy A was computed as A(q, f) = –kBTln[r(q, f)], where kB is the Boltzmann constant, T is

temperature, and r(q, f) is the probability density distribution of the S ligands of the [PnS4] anions

from AIMD simulations. The S atoms bonded to both P and Si in Li3.25[Si0.25P0.75]S4 show a relatively

flat free energy surface, enabling them to easily rotate to their nearby state.

llArticle

timescales of AIMD. This is in full accordance with the distinctly observed reorienta-

tional disorder of [PS4]/[SiS4] anions in our MEM plots. Consistently, both [PS4] and

[SiS4] anions feature dispersive and disordered reorientation/rotation (Figures 4C

and 4D), implying soft reorientation/rotational dynamics. To quantitatively explore

these dynamics, we calculated the Helmholtz free energy of the four S ligands

bonded to P/Si. The Helmholtz free energy surface reflects the average properties

Matter 2, 1667–1684, June 3, 2020 1673

Page 9: Targeting Superionic Conductivity by Turning on Anion

llArticle

that relate to the thermodynamic states as a function of the coordinates. The rotation

routes and barriers can be determined from the coordinates and the heights of the

local minima and transition states. The Helmholtz free energy surfaces for [PS4] and

[SiS4] demonstrated in Figures 4E and 4F explicitly reveal that the S ligands exhibit a

very shallow and flat free energy landscape. The very low free energy barriers of re-

orientational/rotational dynamics (0.03–0.16 eV for [PS4] and 0.03–0.18 eV for [SiS4])

allow facile rotation of these S ligands to their adjacent minima. The [SiS4]4� polyan-

ions show a slightly higher free energy barrier (0.02 eV) for rotation than the [PS4]3�

polyanions, due to the more ionic nature of the Si–S bond compared with the P–S

bond and the higher charge on the complex. Nonetheless, the fairly low barriers

allow the facile rotation of both to the adjacent minima. These barriers are almost

identical to the free energy barrier for the [PS4] polyanion rotation in b-Li3PS4(0.04–0.17 eV, Figure S2). The anion rotational dynamics at lower temperatures in

Li3.25[Si0.25P0.75]S4 were also scrutinized. As shown in Figure S3, dramatic anion re-

orientation is evidenced at 1,050 K from the projected density of S atoms. The fairly

flat energy landscape is revealed by the free energy barrier heights of the anions

(0.07–0.25 eV for [PS4] and 0.08–0.25 eV for [SiS4]).

A more detailed analysis of the trajectory of the S ligands of the [PS4]/[SiS4] anions

was also performed (Figures 5, S4, and S5). As demonstrated in Figures 5A and

5B, while the P–S and Si–S bond lengths are fluctuating due to vibration/rotation,

the bonds are maintained upon [PnS4] polyanion rotation, indicating that the struc-

ture is stable even at elevated temperature. The angle q (the angle between the P–S/

Si–S vector and the z axis, as defined in Figure 4B) is tracked as an indicator of the

anion rotational dynamics. Remarkable changes in q with the simulation time in

Li3.25[Si0.25P0.75]S4 are observed. This reveals the facile and continuous rotation of

the S ligands of [PS4]/[SiS4] during the simulations. The rotation of S from one config-

uration to another occurs over a very short time duration (0.2–5 ps), and a full rotation

was finished in �5–50 ps. Dynamic behavior of [PS4]/[SiS4] anions at other tempera-

tures (900 and 1,050 K) and [PS4] anions in b-Li3PS4 at 1,200 K are shown in Figures

S4–S6. Clearly, [PS4]/[SiS4] polyanion rotation persists down to the simulation tem-

perature of 900 K within the limited timescale of our simulations, suggesting that

rotation in Li3.25[Si0.25P0.75]S4 is readily activated. Compared with Li3.25[Si0.25P0.75]

S4, almost identical facile polyanion rotation is observed in b-Li3PS4 at the equivalent

temperature (Figure S6).

Impacts of Anion Rotation: Widening the Bottleneck for Cation Diffusion

The phase transition of Li3PS4 from the room temperature g phase to the b rotor

phase is accompanied by a volume expansion of 6.7% per formula unit (307.47(7)

A3, Z = 2 for g-Li3PS4 versus 645.23(5) A3, Z = 4 for b-Li3PS4), as generally observed

upon the transformation to the HT rotor phase in other systems.18,41 This first-order

transition is concomitant with the release of the ordered arrangement of [PS4] poly-

anion at larger cell volumes (Figures 1A and 1B), which further creates Li+-ion disor-

der.34 The partial replacement of P5+ by (Si4++ Li+) stabilizes the rotor phase at room

temperature by entropic effects, as suggested by our phonon calculations, and has

the following influences relevant to the Li-ion conductivity: (1) it enhances themobile

carrier concentration; (2) it increases the free transport volume for the cations; and (3)

it induces anion reorientational/rotational disorder (Figures 2 and 3C), which further

enhances the cation diffusivity via a ‘‘paddle-wheel mechanism.’’ The former two fac-

tors have positive contributions to cation conductivity according to the classical

equation, s = nqu, where n is the mobile carrier concentration and u is the mobility.

Themobility u is probably favored by the enlarged free volume and flattened energy

landscape due to possibly more highly correlated motions induced by the increased

1674 Matter 2, 1667–1684, June 3, 2020

Page 10: Targeting Superionic Conductivity by Turning on Anion

Figure 5. Rotational Dynamics of [PnS4] (Pn = P; Si) Polyanions in Li3.25[Si0.25P0.75]S4(A and B) The bond lengths of (A) P4 and (B) Si3 and its four S ligands in Li3.25[Si0.25P0.75]S4 during the AIMD simulations at 1,200 K.

(C–F) The angle q (see reference frame in Figure 4B) of the four S atoms bonded to (C) P4, (D) P12, (E) Si3, and (F) Si4 as a function of the simulation time.

llArticle

Li concentration, but meanwhile is facilitated by the anion rotation invoked upon Si

substitution (see below). These factors cannot be teased apart. We note that the

room temperature conductivity of Li3.25[Si0.25P0.75]S4 is dramatically increased by a

factor of 103 compared with g-Li3PS4 and is comparable with the conductivity of

the rotor phase (b-Li3PS4) at HT (10�3 S cm�1 at �200�C), where the anion rotational

mobility is activated. We therefore propose that the high conductivity of the Si-

substituted Li3PS4 (Li3.25[Si0.25P0.75]S4) cannot be explained only by the increased

mobile carrier concentration and/or the split Li sites reported in our earlier work.38

Instead, the substitution of P5+ for (Li+ + Si4+) in Li3PS4 stabilizes the rotor phase

to room temperature by entropic effects, and consequently Li+-ion diffusion is

strongly facilitated by the paddle-wheel mechanism. This is demonstrated below.

Matter 2, 1667–1684, June 3, 2020 1675

Page 11: Targeting Superionic Conductivity by Turning on Anion

Figure 6. The Effects of [PS4] Anion Rotation: Transiently Widening the Bottleneck for Cation Transport

(A) Trajectories of the S ligands of [PnS4] (Pn = Si/P) polyanions in Li3.25[Si0.25P0.75]S4 within 1 ps from AIMD simulation. Small yellow spheres represent

the initial position of S atoms; orange, blue, green, and purple spheres indicate S trajectories.

(B and C) Trajectories of Li in the vicinity of one [PnS4] (Pn = Si/P) within 1 ps at 1,050 K: (B) when [PnS4] is undergoing rotation; (C) when [PnS4] rotation

pauses.

(D and E) Coordinates of one Li in the vicinity of the [PnS4] (Pn = Si/P): (D) during its rotation; (E) when the anion rotation pauses.

(F) Schematic plot of the effects of the polyanion rotation on the trigonal bottleneck that Li+ ions hop through; the blue solid circles represent S atoms

shared by the [PnS4] (Pn = Si/P) tetrahedron and [LiS4] or [LiS6] polyhedron. Red and blue arrows depict the overall pathways for Li-ion migration derived

from AIMD simulation, and the green arrows (right diagram) show the local Li+-ion path through the triangular face-sharing windows.

(G) Area of the trigonal bottleneck for Li+-ion migration when [PnS4] (Pn = Si/P) is undergoing rotation and when the rotation pauses.

llArticle

To establish a direct link between anion rotation and Li+-cation diffusion, we carried out

detailed analysis of the motions of the two species by scrutinizing the AIMD snapshots.

We randomly selected a duration of 1 ps to study the impact of anion rotation on the

Li+-ion diffusion. A snapshot with the embedded trajectories of S within 1 ps was ex-

tracted from the AIMD simulations and is presented in Figure 6A. The S ligands of the

[PnS4] (Pn = P/Si) groups in the supercell undergo facile rotational motion. A comparison

study was performed here to illustrate the effects of the anion rotation (Figures 6B–6E).

Figure 6B shows the cation trajectories at the vicinity of one [PnS4] (Pn = P/Si) moiety

when it undergoes rotation. During this 1 ps duration, distinct Li+-ion diffusion events

occur in a cooperative manner with the rotation of the anions, as indicated by the red

arrows. One of these diffusive motions was tracked by monitoring its coordinates. As

shown in Figure 6C, this Li+-ion migrates �4.26 A along the a axis and 4.08 A along

1676 Matter 2, 1667–1684, June 3, 2020

Page 12: Targeting Superionic Conductivity by Turning on Anion

llArticle

the c axis. This suggests a successful hop in the ac plane, noting that the Li-Li distance is

�3.56–3.82 A in this plane. A parallel analysis was conducted in another periodwhen the

rotation of this [PnS4] anion pauses. The trajectories of the Li ions explicitly show that

most of the Li+ ions linger around the equilibrium sites when [PnS4] rotation pauses,

although a few diffusional events are observed (Figure 6D). This is further confirmed

by the coordinates of the Li atom in Figure 6E.

To uncover the intrinsic factors underlying the contribution from the anion rotation,

we examined the trigonal bottleneck that Li+ ions need to hop through in order to

diffuse (Figure 6F, middle panel). A much wider bottleneck size is observed upon

the rotation of the S atom (Figure 6G), because it transiently alters the dimensions

of the triangular windows that the Li+ ions must squeeze through, similar to the

process of [PS4] polyanion rotation in Na11Sn2PS12.31 This is due to the fact that

the [LiS4] and [LiS6] cages share sulfur ligands with the [PS4]/[SiS4] polyanions, as

illustrated in Figure 6F (middle panel). The widening of this bottleneck as the

polyanion rotates decreases the energy barrier for successful Li+-ion jumps and

hence facilitates the Li+-ion diffusion. Such an effect is the so-called paddle-wheel

mechanism. This signifies the fluctuation of the anion landscape, which in turn can

cause frustration of the local and overall potential energy landscape experienced

by cations prior to and during successful diffusive processes. We propose that

the anion fluctuations provide the driving force for cations to redistribute along

the lower energy barrier path and, therefore, decrease the energy barrier for cation

diffusion.

Direct Proof of Coupled Anion Rotation and Cation Diffusion

To gain more generic evidence of the cation-anion interplay, we generated the power

spectrumof both cations and anions via Fourier transform of the velocity autocorrelation

function from AIMD simulations (Figure 7A). The frequency of the anion rotation is

important, as it needs to occur at the same timescale of the cation translational diffusion

to ensure that the two motions are coupled. Specific vibrational modes related to anion

frequency regimes can be identified: mid-frequency internal anion fluxional and

torsional modes (15 < f < 30 THz) and low-frequency vibrational, librational, and rota-

tional modes (f< 15 THz). The high degree of dispersion in the S-derived spectrum likely

derives from the propensity of S ligands bonded to Si/P to adopt a variety of positions,

whereby each position would exhibit a different dynamic frequency. This explains the

origin of the observed [PS4]/[SiS4] rotation in Li3.25[Si0.25P0.75]S4. Modes associated

with the [PS4]/[SiS4] anion group were derived by Fourier transform of the average

angular velocity of individual P/Si and S atoms (ui) (Figure 7A). A strong overlap at low

frequency is evident between the cation diffusional and anion rotational modes upon

comparing the Li-derived modes, the librational/rotational modes of the S atoms, and

the [PS4]/[SiS4] angular spectrum. This implies probable momentum transfer between

the anion rotationalmotion and cation vibrations prior to diffusive events.Wealso exam-

ined the power spectrum in b-Li3PS4 where a strong overlap between the Li- and

S-derived modes was also observed (Figure S7), again suggesting coupled motion be-

tween cation diffusion and anion rotation.

The impact of this anion rotation can be further evidenced from the Helmholtz free

energy barrier (A) for a Li+-ion hop from the first shell of the [PS4]/[SiS4] group to the

second shell. A very low A value is observed; namely, for Li+ ions to hop from the first

shell of the [PS4] group to the second shell, they need to overcome a free energy bar-

rier of 0.14 eV and a barrier of 0.17 eV from the first shell of [SiS4] to the second shell

(Figure 7B and Table 1). The free energy barrier for Li+ ions to migrate in b-Li3PS4 is

predicted to be 0.15 eV (Figure S7). Note that these values are fairly close to the free

Matter 2, 1667–1684, June 3, 2020 1677

Page 13: Targeting Superionic Conductivity by Turning on Anion

Figure 7. Coupled Anion Rotation and Cation Diffusion

(A) Power spectrum calculated via the Fourier transform of the velocity autocorrelation function for

Li+ ions (red curve) and S2� ions (yellow curve), and Fourier transform of the averaged angular

velocity autocorrelation function for [PS4] and [SiS4] anions in Li3.25[Si0.25P0.75]S4. The purple and

blue curves correspond to the quantity derived from the average angular velocity of individual P/Si

and S atoms (ui) in [PS4] and [SiS4] polyanions, respectively.

(B) Free energy barrier profile, A, for Li+-ion diffusion from the first coordinate shell to the second

shell of the [PS4]/[SiS4] anion group in Li3.25[Si0.25P0.75]S4; r is defined as the distance of Li to Si/P.

(C) 2D probability density distribution r2Dq; r of the P-S-Li angles (q) and the distance (r) between

S ligands and Li.

(D) 2D probability density distribution r2Dq;r of the Si-S-Li angles (q) and the distance (r) between S

ligands and Li.

(E and F) Residence time correlation functions ChhðtÞD for Li+ cations nearest to the polyhedral

anions: (E) [PS4] and (F) [SiS4]. The CPlPlðtÞD polyanions orientational correlation functions are also

shown, along with ChPlPlðtÞhðtÞD orientational correlation function restricted to Li+ cations nearest to

the polyhedral anion.

llArticle

energy barriers for anion rotation in the two rotor systems (0.03–0.16 eV for [PS4] an-

ions and 0.03–0.18 eV for [SiS4] anions in Li3.25[Si0.25P0.75]S4; 0.04–0.17 eV for [PS4]

groups in b-Li3PS4). This suggests that cation diffusional and anion rotational dy-

namics sustain a significant mutual effect.

1678 Matter 2, 1667–1684, June 3, 2020

Page 14: Targeting Superionic Conductivity by Turning on Anion

Table 1. Helmholtz Free Energy Barrier (A) for Li+-Ion Diffusion and [PS4]3�/[SiS4]

4� Polyanion

Rotation, as well as Activation Energy (Ea) for Li+-Ion Diffusion in Li3.25Si0.25P0.75S4 from AIMD

Simulations

Composition Li3.25[Si0.25P0.75]S4 b-Li3PS4

A for Li-cation diffusion (eV) 0.14 eV in the shell of [PS4]3� anions;

0.17 eV in the shell of [SiS4]4� anions

0.15 eV in the shellof [PS4]

3� anions

Ea (experiment, eV) 0.31 eV (Tang et al.27) 0.36 eV, 0.47 eV(Hanghofer et al.33)

Ea (predicted from AIMD, eV) 0.30 eV 0.35 eV

A for [SiS4]/[PS4] anionrotation (eV)

0.03–0.16 eV for [PS4]3� anions;

0.03–0.18 eV for [SiS4]4� anions

0.04-0.17 eV for[PS4]

3� anions

llArticle

To demonstrate the direct interaction of the mobile cations and the rotational

mobility of the anions, we quantitatively computed the correlation between the

two. Figures 7C and 7D show the 2D probability distribution (r2Dq;r ) of the P/Si-S-Li

angle and the distance between S ligands and Li atoms in the first shell. The (r2Dq;r )

of both [PS4] and [SiS4] groups are highly delocalized, signifying a strong correlation

between the polyanion and the Li+ cations. On the other hand, if there were negli-

gible correlation between Li+-ion diffusion and [PS4]/[SiS4] rotation, the r2Dq;r would

be localized in a discrete spot. These intriguing observations reveal significant

coupling between the angular and radial coordinates of the Li+ ion relative to the

[PS4]/[SiS4] group.

To verify this mechanism from a more quantitative perspective, we defined a resi-

dence time correlation function to assess the propensity of the Li+-ion to escape

from its local minima. This function decays very quickly for both [PS4] and [SiS4],

with a characteristic time of 9.5 ps and 12.8 ps, respectively, suggesting that Li+

ions stay slightly longer in the first shell of the [SiS4] anion compared with that of

the [PS4] anion. We also present the CPlPlðtÞD correlation function to depict the decay

of the orientation between the Pn–Li and Pn–S bond (Pn = P, Si) vectors (more details

regarding Pl are presented in Computational Details). The decay of CP1P1ðtÞ Dis fasterthan that of CP2P2ðtÞD as expected. Both CP1P1ðtÞD and CP2P2ðtÞD correlation functions

for [PS4] anions (2.4 ps for l = 1 and 1.1 ps for l = 2) show faster decay than those of

[SiS4] (2.6 ps for l = 1 and 1.5 ps for l = 2), since rotational motion of the former occurs

with faster decay. To further probe the contribution of anion reorientation to the

cation diffusional motion, we define a joint-time correlation function

ðChPlhðtÞPlðtÞDÞ, which measures how effectively anion reorientation facilitates the

cation diffusion. This joint correlation function quickly decays for [PS4] polyanions

with values of 1.7 ps for l= 1 and 0.6 ps for l= 2. For [SiS4] polyanions, the decay times

are slightly longer, namely 2.3 ps for l = 1 and 0.8 ps for l = 2. These findings verify

that the rotation of the anion groups creates disorder in the anion framework, which

in turn leads to a fluctuating potential and frustrated energy landscape perceived by

the cations. As a result, the activation energy for cation migration is lowered.

To further confirm whether the increased volume upon phase transition turns on the

anion rotation, we adjusted the free volume for cation diffusion/anion rotation by

varying the volume of Li3.25Si0.25P0.75S4. As shown in Figures S8 and S9, reducing

the volume by 10% hinders the [PnS4] anion rotation (this reduced volume is lower

than that of the LT phase g-Li3PS4) and leads to much higher rotational barrier

heights (0.07–0.26 eV for [PS4] anions and 0.06–0.26 eV for [SiS4] anions). The

reduced free volume, together with the more undulating anion free energy land-

scape, results in a lower cation diffusivity (7.603 10�9 m2 s�1 at 1,200 K). These find-

ings suggest that anion rotation is reduced upon volume contraction but not

Matter 2, 1667–1684, June 3, 2020 1679

Page 15: Targeting Superionic Conductivity by Turning on Anion

llArticle

extinguished. The origin of the activation of anion rotation upon the g-b phase tran-

sition or Si substitution in Li3PS4 remains an open question. Rotational entropy may

play a factor, but this is a future investigation not within the scope of this study.

CONCLUSIONS

Our findings indicate how anion dynamics influence cation mobility. Facile rotation of

the [PS4]/[SiS4] polyanions in Li3.25[Si0.25P0.75]S4 is confirmed by a combination

of neutron diffraction based on MEM analysis and AIMD simulations. Direct proof of

an interaction between rotational dynamics of the anions and translational diffusion of

the cations via a paddle-wheelmechanismwas illustrated byAIMD. Turning on the anion

rotational ability by leveraging the anion framework is of significance because such anion

rotation causes disorder in the anion framework and lowers the energy barrier landscape

for cation migration. Our work also highlights the importance of generating a ‘‘rotor

phase’’ at room temperature via anion tuning. The introduction of Si into the anions sta-

bilizes the HT rotor phase at room temperature by entropic effects and ‘‘turns on’’ the

anion rotation. Our findings suggest that there is significant potential to be exploited

in superionic conductors with isolated polyanion groups. Enhanced cation conductivity

can be achieved by leveraging the anions to invoke the paddle-wheel mechanism

(anion-cation interaction). This should serve as a guideline in the future design of fast

ion conductors. The concept can be extended to other systems, particularly where a

static anion framework would otherwise provide poor conductivity.We note that sulfides

are one of themost likely conductors when anion rotation can persist down to room tem-

perature, due to the larger polarizability of sulfides compared with oxides.

EXPERIMENTAL PROCEDURES

Resources Availability

Lead Contact

Linda F. Nazar (email: [email protected])

Materials Availability

This study did not generate new unique materials.

Data and Code Availability

All experimental data, computational data, and code are available upon reasonable

request to the Lead Contact author.

Preparation of Li3PS4 and Li3.25Si0.25P0.75S4Li3PS4 and Li3.25Si0.25P0.75S4 were synthesized as described in our previous paper.38

Neutron Powder Diffraction Measurements

Time-of-flight NPD measurements were performed on POWGEN at the Spallation

Neutron Source at the Oak Ridge National Laboratory. The samples were loaded

into a vanadium can sealed with a copper gasket and aluminum lid before measure-

ment. The NPD data were collected at room temperature, 200�C, 300�C, and 350�C.Data analysis by maximum entropy methods were performed using RIETAN-FP42

and Dysnomia codes.43

Computational Details

Geometry optimization was carried out by density functional theory calculations us-

ing the Vienna ab initio simulation package (VASP)44 with the projector augmented-

wave method.45 The generalized gradient approximation function parameterized by

Perdew-Burke-Ernzerhof (PBE)46 was used to describe the exchange correlation po-

tential. The cutoff energy in our calculations is set to 450 eV. The geometry

1680 Matter 2, 1667–1684, June 3, 2020

Page 16: Targeting Superionic Conductivity by Turning on Anion

llArticle

optimization was performed on the 1a32b32c super cell. The total energy and the

force on each atom were converged to within 10�5 eV and 0.01 eV A�1, respectively.

A G-centered k-point mesh of 3 3 2 3 2 was used for the Brillouin integrations.

The effects of phonons on the thermodynamic properties were calculated using

the VASP implementation of density functional perturbation (DFPT) and the

PHONONPY software package.47 For the electronic structure calculations, the

convergence criteria for the total energy and force on each atom were set to

10�10 eV and 0.001 eV A�1, respectively. For phonon calculations, supercells

(2a32b32c) and a fine k-point grid (2 3 2 3 1) for the three compositions,

g-Li3PS4, b-Li3PS4, and Li3.25[Si0.25P0.75]S4, were used to obtain accurate values of

forces.

AIMD simulations were carried out at a constant pressure within the canonical (NVT)

ensemble using a Nose thermostat48 to maintain constant temperature. The volume

of the unit cell was fixed to the value from the fully relaxed structure. All simulations

were thermalized to the specific temperature (600 K, 750 K, 900 K, 1,050 K, 1,200 K,

1,350 K), equilibrated for 5 ps, then equilibrated at the desired temperature for 450

ps. To keep the computational time reasonable for the relatively large unit cell, we

only performed integration in reciprocal space at the G-point.

Li+-ion probability densities were calculated from the atomic trajectories. The diffu-

sion coefficient is defined as

D = limt/N

�1

2dtj r!iðtÞ � r!ið0Þj2

�; (Equation 1)

where d is the dimension of the lattice on which the diffusion take place and t is the

elapsed time. The average-mean-square displacement,

Cj r!ðtÞ � r!ð0Þj2D = 1

N

XN

i = 1Cj r!iðt + t0Þ � r!iðt0Þj2D; (Equation 2)

is the averaged displacement of Li atoms over time t, r!iðtÞ is the displacement of the

ith Li ion at time t, and t0 is the initial time.D is obtained by linear fitting to the depen-

dence of average-mean-square displacement over 2dt.

The van Hove correlation function was calculated from the AIMD simulations.49 The

distinctive part Gdðr ; tÞ is computed as follows:

Gdðr ; tÞ = 1

4pr2rNdCXNd

isj

d�r � ��riðt0Þ � rjðt + t0Þ

���Dt0 ; (Equation 3)

where d is the Dirac delta function, r!iðtÞ represents the position of the ith particle at

time t. and Nd and r are the number of diffusing alkali ions in the unit cell and radial

distance, respectively. The average number density serves as the normalization fac-

tor in Gd that Gd/1 when r [ 1. For a given r and t, Gdðr ; tÞ describes the radial

distribution of N � 1 particles after a period of t with respect to the initial reference

particle.

We also computed the normalized 2D probability density distributionr2Dq;f as a func-

tion of q and f for the S ligands of [SiS4]/[PS4] anions from the AIMD simulations.

Here, q represents the angle between a selected Si/P–S bond and the z axis, and

f corresponds to the angle between the x axis and the projection of the same Si/

P–S vector in the xy plane. Based on this, the free energy surface of the S ligands

was computed via

Matter 2, 1667–1684, June 3, 2020 1681

Page 17: Targeting Superionic Conductivity by Turning on Anion

llArticle

Aðq; fÞ = � kBT ln r2Dq;f; (Equation 4)

where kB is the Boltzmann constant and T is temperature.

The one-dimensional (1D) Li probability distribution r1Dr as a function of the distance

(r) was also calculated from AIMD simulations. The corresponding 1D free energy

curve is calculated by

AðrÞ = � kBT ln r1Dr : (Equation 5)

The power spectrumwas calculated via the Fourier transform of the velocity autocor-

relation function (ðCPNi1 bvi ðtÞ ,bvi ðt + DtÞDÞ, in which i = Li or S in Li3.25[Si0.25P0.75]S4.Ni

is the total number for each type of atom in the supercell. In our work,NLi = 52,NSi =

4, NP = 12, and NS = 64.

The angular power spectrum for the entire [SiS4]/[PS4] anion group was obtained via

the Fourier transform of the averaged angular velocity autocorrelation function

ðCP cuk ðtÞ ,cuk ðt + DtÞDÞ, in which, cuk is the angular velocity of individual Sk ligand

of PS4 anion groups, which is calculated via

buk =br k 3 bvk

r2k: (Equation 6)

Here, br k and cvk are the position vector and velocity for each S atom relative to the

central of mass of [SiS4]/[PS4] anion group.

The timescale for the hopping of the Li+ cations to the nearest polyhedral anion was

calculated by analyzing the decay of the residence time correlation function. This

correlation function is defined as Ch h(t)D h Ch([R_cut-R(t = 0)] h[R_cut-R(t)]D, where

h is the Heaviside function. This function equals unity when R < R_cut and is zero

otherwise (R_cut = 6.1 A). Another correlation function, PlPlðtÞ, is defined to depict

that the decay of the orientation between the P/Si–Li and P/Si–S bond vectors, where

Pl is a Legendre Polynomial of degree l = 1, 2. The argument of Pl is the cosine of the

angle between the Pn–Li and Pn–S bond vectors. We refer to the l = 1 case as first

order and to the l = 2 case as second order. We have also combined the residence

time with the orientational dynamics in the ChPlh(t)Pl(t)D correlation. This joint-time

correlation function measures the contribution of anion orientational dynamics to

the Li+-cations.

SUPPLEMENTAL INFORMATION

Supplemental Information can be found online at https://doi.org/10.1016/j.matt.

2020.04.027.

ACKNOWLEDGMENTS

This work was supported by the Joint Center for Energy Storage Research, an Energy

Innovation Hub funded by the US Department of Energy (DOE), Office of Science,

Basic Energy Sciences. Neutron diffraction measurements at the POWGEN instru-

ment at Oak Ridge National Laboratory, Spallation Neutron Source, was sponsored

by the Scientific User Facilities Division, Office of Basic Energy Sciences, US DOE.

We would like to thank Dr. Ashfia Huq for the NPD data collection. L.F.N and

P.-N.R. also thank NSERC for platform support through the Discovery Grant and

Canada Research Chair programs. H.L. acknowledges the support of the National

Natural Science Foundation of China (grant nos. 21773081 and 21533003) and the

Talents Cultivation Program (Jilin University, China). L.F.N. gratefully acknowledges

Compute Canada for CPU time allocations.

1682 Matter 2, 1667–1684, June 3, 2020

Page 18: Targeting Superionic Conductivity by Turning on Anion

llArticle

AUTHOR CONTRIBUTIONS

Conceptualization, Z.Z. and L.F.N.; Methodology, Z.Z, H.L., and P.-N.R.; Formal

analysis, Z.Z. and H.L.; Investigation, Z.Z. and K.K.; Resources, L.F.N., Z.Z., and

L.Z.; Data Curation, H.L., Z.Z., and P.-N.R,; Writing – Original Draft, Z.Z.; Writing –

Review & Editing, Z.Z., L.F.N., and P.-N.R.; Visualization, Z.Z.; Supervision, Z.Z.

and L.F.N.; Project Administration, L.F.N.; Funding Acquisition, L.F.N.

DECLARATION OF INTERESTS

The authors declare no competing interests.

Received: November 28, 2019

Revised: March 24, 2020

Accepted: April 28, 2020

Published: June 3, 2020

REFERENCES

1. Bachman, J.C., Muy, S., Grimaud, A., Chang,H.-H., Pour, N., Lux, S.F., Paschos, O., Maglia,F., Lupart, S., and Lamp, P. (2015). Inorganicsolid-state electrolytes for lithium batteries:mechanisms and properties governing ionconduction. Chem. Rev. 116, 140–162.

2. Janek, J., and Zeier, W.G. (2016). A solid futurefor battery development. Nat. Energy 1, 16141.

3. Goodenough, J.B., and Park, K.-S. (2013). TheLi-ion rechargeable battery: a perspective.J. Am. Chem. Soc. 135, 1167–1176.

4. Zhang, Z., Shao, Y., Lotsch, B., Hu, Y.-S., Li, H.,Janek, J., Nazar, L.F., Nan, C.-W., Maier, J., andArmand, M. (2018). New horizons for inorganicsolid state ion conductors. Energy Environ. Sci.11, 1945–1976.

5. Kamaya, N., Homma, K., Yamakawa, Y.,Hirayama, M., Kanno, R., Yonemura, M.,Kamiyama, T., Kato, Y., Hama, S., andKawamoto, K. (2011). A lithium superionicconductor. Nat. Mater. 10, 682–686.

6. Bron, P., Johansson, S., Zick, K., Schmedt aufder Gunne, J., Dehnen, S., and Roling, B.(2013). Li10SnP2S12: an affordable lithiumsuperionic conductor. J. Am. Chem. Soc. 135,15694–15697.

7. Seino, Y., Ota, T., Takada, K., Hayashi, A., andTatsumisago, M. (2014). A sulphide lithiumsuper ion conductor is superior to liquid ionconductors for use in rechargeable batteries.Energy Environ. Sci. 7, 627–631.

8. Busche, M.R., Weber, D.A., Schneider, Y.,Dietrich, C., Wenzel, S., Leichtweiss, T.,Schroder, D., Zhang, W., Weigand, H., andWalter, D. (2016). In situ monitoring of fast Li-ion conductor Li7P3S11 crystallization inside ahot-press setup. Chem. Mater. 28, 6152–6165.

9. Kraft, M.A., Culver, S.P., Calderon, M., Bocher,F., Krauskopf, T., Senyshyn, A., Dietrich, C.,Zevalkink, A., Janek, J., and Zeier, W.G. (2017).Influence of lattice polarizability on the ionicconductivity in the lithium superionicargyrodites Li6PS5X (X= Cl, Br, I). J. Am. Chem.Soc. 139, 10909–10918.

10. Deiseroth, H.J., Kong, S.T., Eckert, H.,Vannahme, J., Reiner, C., Zaiss, T., and

Schlosser, M. (2008). Li6PS5X: a class ofcrystalline Li-rich solids with an unusually highLi+ mobility. Angew. Chem. Int. Ed. 47,755–758.

11. Wang, Y., Richards, W.D., Ong, S.P., Miara, L.J.,Kim, J.C., Mo, Y., and Ceder, G. (2015). Designprinciples for solid-state lithium superionicconductors. Nat. Mater. 14, 1026–1031.

12. Zhang, Z., Ramos, E., Lalere, F., Assoud, A.,Kaup, K., Hartman, P., and Nazar, L.F. (2018).Na11Sn2 PS12: a new solid state sodiumsuperionic conductor. Energy Environ. Sci. 11,87–93.

13. Ramos, E.P., Zhang, Z., Assoud, A., Kaup, K.,Lalere, F., and Nazar, L.F. (2018). Correlatingion mobility and single crystal structure insodium-ion chalcogenide-based solid statefast ion conductors: Na11Sn2PnS12 (Pn= Sb, P).Chem. Mater. 30, 7413–7417.

14. Schlem, R., Muy, S., Prinz, N., Banik, A.,Shao-Horn, Y., Zobel, M., and Zeier, W.G.(2019). Mechanochemical synthesis: a tool totune cation site disorder and ionic transportproperties of Li3MCl6 (M = Y, Er) superionicconductors. Adv. Energy Mater. https://doi.org/10.1002/aenm.201903719.

15. He, X., Zhu, Y., and Mo, Y. (2017). Origin of fastion diffusion in super-ionic conductors. Nat.Commun. 8, 15893.

16. Zhang, Z., Zou, Z., Kaup, K., Xiao, R., Shi, S.,Avdeev, M., Hu, Y.S., Wang, D., He, B., and Li,H. (2019). Correlated migration invokes higherNa+-ion conductivity in NaSICON-type solidelectrolytes. Adv. Energy Mater. 9, 1902373.

17. Krauskopf, T., Pompe, C., Kraft, M.A., andZeier, W.G. (2017). Influence of lattice dynamicson Na+ transport in the solid electrolyteNa3PS4-xSex. Chem. Mater. 29, 8859–8869.

18. Jansen, M. (1991). Volume effect orpaddle-wheel mechanism-fast alkali-metalionic conduction in solids with rotationallydisordered complex anions. Angew. Chem. Int.Ed. 30, 1547–1558.

19. Wilmer, D., Funke, K., Witschas, M., Banhatti,R.D., Jansen, M., Korus, G., Fitter, J., andLechner, R. (1999). Anion reorientation in an ion

conducting plastic crystal–coherentquasielastic neutron scattering from sodiumortho-phosphate. Phys. B Condens. Matter266, 60–68.

20. Karlsson, L., and McGreevy, R. (1995).Mechanisms of ionic conduction in Li2SO4 andLiNaSO4: paddle wheel or percolation? SolidState Ionics 76, 301–308.

21. Ikeshoji, T., Tsuchida, E., Morishita, T., Ikeda,K., Matsuo, M., Kawazoe, Y., and Orimo, S.-I.(2011). Fast-ionic conductivity of Li+ in LiBH4.Phys. Rev. B 83, 144301.

22. Udovic, T.J., Matsuo, M., Tang, W.S., Wu, H.,Stavila, V., Soloninin, A.V., Skoryunov, R.V.,Babanova, O.A., Skripov, A.V., and Rush, J.J.(2014). Exceptional superionic conductivity indisordered sodiumdecahydro-closo-decaborate. Adv. Mater. 26,7622–7626.

23. Verdal, N., Udovic, T.J., Stavila, V., Tang, W.S.,Rush, J.J., and Skripov, A.V. (2014). Anionreorientations in the superionic conductingphase of Na2B12H12. J. Phys. Chem. C 118,17483–17489.

24. Kweon, K.E., Varley, J.B., Shea, P., Adelstein,N., Mehta, P., Heo, T.W., Udovic, T.J., Stavila,V., and Wood, B.C. (2017). Structural, chemical,and dynamical frustration: origins of superionicconductivity in closo-borate solid electrolytes.Chem. Mater. 29, 9142–9153.

25. Tang, W.S., Matsuo, M., Wu, H., Stavila, V.,Zhou, W., Talin, A.A., Soloninin, A.V.,Skoryunov, R.V., Babanova, O.A., and Skripov,A.V. (2016). Liquid-like ionic conduction in solidlithium and sodiummonocarba-closo-decaborates near or at roomtemperature. Adv. Energy Mater. 6, 1502237.

26. Dimitrievska, M., Shea, P., Kweon, K.E., Bercx,M., Varley, J.B., Tang, W.S., Skripov, A.V.,Stavila, V., Udovic, T.J., and Wood, B.C. (2018).Carbon incorporation and anion dynamics assynergistic drivers for ultrafast diffusion insuperionic LiCB11H12 and NaCB11H12. Adv.Energy Mater. 8, 1703422.

27. Tang, W.S., Dimitrievska, M., Stavila, V., Zhou,W., Wu, H., Talin, A.A., and Udovic, T.J. (2017).Order-disorder transitions and superionic

Matter 2, 1667–1684, June 3, 2020 1683

Page 19: Targeting Superionic Conductivity by Turning on Anion

llArticle

conductivity in the sodium nido-undeca (carba)borates. Chem. Mater. 29, 10496–10509.

28. Kim, S., Oguchi, H., Toyama, N., Sato, T.,Takagi, S., Otomo, T., Arunkumar, D., Kuwata,N., Kawamura, J., and Orimo, S.-I. (2019). Acomplex hydride lithium superionic conductorfor high-energy-density all-solid-state lithiummetal batteries. Nat. Commun. 10, 1081.

29. Martelli, P., Remhof, A., Borgschulte, A.,Ackermann, R., Strassle, T., Embs, J.P., Ernst,M., Matsuo, M., Orimo, S.-I., and Zuttel, A.(2011). Rotational motion in LiBH4/LiI solidsolutions. J. Phys. Chem. A 115, 5329–5334.

30. Smith, J.G., and Siegel, D.J. (2020). Low-temperature paddlewheel effect in glassy solidelectrolytes. Nat Commun. 11, 1483.

31. Zhang, Z., Roy, P., Li, H., Avdeev, M., andNazar,L. (2019). Coupled cation-anion dynamicsehances cation mobility in room temperaturesuperionic solid-state electrolytes. J. Am.Chem. Soc. 141, 19360–19372.

32. Famprikis, T., Dawson, J.A., Fauth, F., Clemens,O., Suard, E., Fleutot, B., Courty, M., Chotard,J.-N., Islam, M.S., and Masquelier, C. (2019). Anew, superionic, plastic polymorph of the Na+

ion conductor Na3PS4. ACS Mater. Lett. 1,641–646.

33. Hanghofer, I., Gadermaier, B., and Wilkening,H.M.R. (2019). Fast rotational dynamics inargyrodite-type Li6PS5X (X: Cl, Br, I) as seen by31P nuclear magnetic relaxation–on cation-anion coupled transport in thiophosphates.Chem. Mater. 31, 4591–4597.

34. Homma, K., Yonemura, M., Kobayashi, T.,Nagao, M., Hirayama, M., and Kanno, R. (2011).Crystal structure and phase transitions of the

1684 Matter 2, 1667–1684, June 3, 2020

lithium ionic conductor Li3PS4. Solid StateIonics 182, 53–58.

35. Tachez, M., Malugani, J.-P., Mercier, R., andRobert, G. (1984). Ionic conductivity of andphase transition in lithium thiophosphateLi3PS4. Solid State Ionics 14, 181–185.

36. Liu, Z., Fu, W., Payzant, E.A., Yu, X., Wu, Z.,Dudney, N.J., Kiggans, J., Hong, K.,Rondinone, A.J., and Liang, C. (2013).Anomalous high ionic conductivity ofnanoporous b-Li3PS4. J. Am. Chem. Soc. 135,975–978.

37. Stoffler, H., Zinkevich, T., Yavuz, M., Senyshyn,A., Kulisch, J., Hartmann, P., Adermann, T.,Randau, S., Richter, F.H., and Janek, J. (2018).Li+-ion dynamics in b-Li3PS-4 observed byNMR:local hopping and long-range transport.J. Phys. Chem. C 122, 15954–15965.

38. Zhou, L., Assoud, A., Shyamsunder, A., Huq, A.,Zhang, Q., Hartmann, P., Kulisch, J., and Nazar,L.F. (2019). An entropically stabilized fast-ionconductor: Li3.25[Si0.25P0.75]S4. Chem. Mater.31, 7801–7811.

39. Chumakov, A., Sergueev, I., Van Burck, U.,Schirmacher, W., Asthalter, T., Ruffer, R.,Leupold, O., and Petry, W. (2004). Collectivenature of the boson peak and universaltransboson dynamics of glasses. Phys. Rev.Lett. 92, 245508.

40. Buchter, F., Łodziana, Z., Mauron, P., Remhof,A., Friedrichs, O., Borgschulte, A., Zuttel, A.,Sheptyakov, D., Strassle, T., and Ramirez-Cuesta, A. (2008). Dynamical properties andtemperature induced molecular disordering ofLiBH4 and LiBD4. Phys. Rev. B 78, 094302.

41. Sadikin, Y., Skoryunov, R.V., Babanova, O.A.,Soloninin, A.V., Lodziana, Z., Brighi, M.,Skripov, A.V., and Cerny, R. (2017). Aniondisorder in K3BH4B12H12 and its effect on cationmobility. J. Phys. Chem. C 121, 5503–5514.

42. .Izumi, F., and .Momma, K., eds. Three-dimensional visualization in powder diffraction.In Solid State Phenomena, vol. 130, D. Stro _zand M. Karolus, eds. (Trans Tech Publications),pp. 15–20.

43. Momma, K., Ikeda, T., Belik, A.A., and Izumi, F.(2013). Dysnomia, a computer program formaximum-entropy method (MEM) analysis andits performance in the MEM-based patternfitting. Powder Diffr. 28, 184–193.

44. N Nose, S. (1984). A unified formulation of theconstant temperature molecular dynamicsmethods. J. Chem. Phys. 81, 511–519.

45. Franceschetti, A., Wei, S.-H., and Zunger, A.(1994). Absolute deformation potentials of Al,Si, and NaCl. Phys. Rev. B 50, 17797.

46. Perdew, J.P., Burke, K., and Ernzerhof, M.(1996). Quantum theory group tulaneuniversity. Phys. Rev. Lett. 77, 3865–3868.

47. Togo, A., and Tanaka, I. (2015). First principlesphonon calculations in materials science. Scr.Mater. 108, 1–5.

48. Hoover, W.G. (1985). Canonical dynamics:equilibrium phase-space distributions. Phys.Rev. A 31, 1695.

49. Van Hove, L. (1954). Correlations in space andtime and Born approximation scattering insystems of interacting particles. Phys. Rev. 95,249.