The Superion Conductor Na11.1Sn2.1P0.9Se12 - ACS Publications

potential shortages, Na+ ion batteries attract renewed interest.7,10,11. Nowadays, both ... Page 1 of 7 ..... Conductor. J. Power Sources 2016, 329, 5...
32 downloads 0 Views 845KB Size
Subscriber access provided by Kaohsiung Medical University

Article

The Superion Conductor Na11.1Sn2.1P0.9Se12: Lowering the Activation Barrier of Na+ Conduction in Quaternary 1-4-5-6 Electrolytes Marc Duchardt, Sven Neuberger, Uwe Ruschewitz, Thorben Krauskopf, Wolfgang G. Zeier, Jörn Schmedt auf der Günne, Stefan Adams, Bernhard Roling, and Stefanie Dehnen Chem. Mater., Just Accepted Manuscript • DOI: 10.1021/acs.chemmater.8b01656 • Publication Date (Web): 24 May 2018 Downloaded from http://pubs.acs.org on May 24, 2018

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

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

Page 1 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

The Superion Conductor Na11.1Sn2.1P0.9Se12: Lowering the Activation Barrier of Na+ Conduction in Quaternary 1-4-5-6 Electrolytes Marc Duchardt,† Sven Neuberger,# Uwe Ruschewitz,‡ Thorben Krauskopf,§ Wolfgang G. Zeier,§,∥ Jörn Schmedt auf der Günne,# Stefan Adams,∆ Bernhard Roling,†* Stefanie Dehnen†* † Fachbereich Chemie and Wissenschaftliches Zentrum für Materialwissenschaften (WZMW), Philipps-Universität Marburg, Hans-Meerwein-Straße 4, D-35043 Marburg, Germany. ‡ Department of Chemistry, University of Cologne, Greinstraße 6, D-50939 Köln, Germany. § Institute of Physical Chemistry, Justus-Liebig-University Giessen, Heinrich-Buff-Ring 17, D-35392 Giessen, Germany. ∥ Center for Materials Research (LaMa), Justus-Liebig-University Giessen, Heinrich-Buff-Ring 16, D-35392 Giessen, Germany. # Inorganic Materials Chemistry, Department of Chemistry/Biology, University of Siegen, 57068 Siegen, Germany. ∆ Department of Materials Science and Engineering, National University of Singapore, Singapore, Singapore 117576. ABSTRACT: We report on the first quaternary selenide-based Na+ superionic solid electrolyte, Na11.1Sn2.1P0.9Se12 (further denoted as NaSnPSe), which shows virtually the same room temperature Na+ ion conductivity (3.0 mS/cm) as the current record holder for sulfidebased systems, Na11Sn2PS12 (denoted as NaSnPS), but with a considerably lower activation energy of 0.30 eV. Both electrolytes belong to the currently highly topical class of solids comprising group I, IV, V, and VI atoms, which we summarize as 1-4-5-6 electrolytes. Herein, they are compared to each other with regard to their structural characteristics and the resulting ion transport properties. The lower activation energy of Na+ ion transport in NaSnPSe is well in line with the results of speed of sound measurements, Raman spectroscopy, bondvalance site energy calculations and molecular dynamics simulations, all of which point to a lower lattice rigidity and to weaker Nachalcogen interactions as compared to NaSnPS.

In order to proceed to a sustainable economy, modern society has to turn its back on fossil fuels.1 But the most important alternatives, wind and solar energy, can only be produced on a fluctuating basis, giving rise to the need for safe, reliable and affordable ways to store electrical energy.2,3 High-energy Li+ ion batteries can surely contribute to the mitigation of this problem. But already today they are widely used for a multitude of different devices like for example smartphones, laptops or electric cars. Therefore, justifiable doubts arise whether the future lithium supply will meet the continuously increasing demand, and at what price.2–5 Another pressing issue for Li+ ion batteries resides in the already beginning shortage in the supply of cobalt which represents a crucial component in Li+ ion cathode materials.6 For Na+ ion batteries on the other hand, a multitude of cobalt-free cathode materials exist.7–9 Even though a further replacement of cobalt inside Li+ ion cathode materials would surely reduce the demand for this raw material, the demand for Li would remain unaffected thus raising the battery price considerably.6 In the context of these potential shortages, Na+ ion batteries attract renewed interest.7,10,11 Nowadays, both commercial Li+ ion batteries and the still experimental room temperature Na+ ion batteries draw on organic carbonates as electrolytes.12–14 But, as frequently reported, the usage of these electrolytes poses safety risks due to potential leakages and their flammability.1 Solid electrolytes (SEs), on the other hand, represent a suitable alternative, especially regarding their high ionic conductivities. Among the SEs, chalcogenides have attracted much attention due to their greater softness and hence easier processibility compared to oxides. Fast-ion conducting sulfide-based Li+ ion conductors have been known for many years now.15 The ternary system Li–P–S has

been intensively studied by many groups.16–21 In 2011 Kamaya et al. succeeded in the pioneering discovery of Li10GeP2S12, which adopts the newly-found LGPS-type structure. In a comprehensive theoretical study in 2013, Ceder and coworkers investigated various compounds of the general formula type Li10TtP2E12 (Tt = Si, Ge, Sn, E = O, S, Se), that were supposed to adopt an LGPS-type structure.22 From this study, important trends for Li+ superionic conductors were derived that suggested increasing conductivity in the order Sn  Ge  Si and O  S  Se. Later experimental studies succeeded in the realization of several of the proposed substitutions, e.g. the replacement of expensive Ge by Si or Sn.23– 27 The resulting compounds turned out to be isostructural to the mother compound LGPS, but while incorporation of Sn indeed leads to a lower conductivity due to a tightening along the diffusion channels in z-direction,25 the expected increase for the Si homologue was disproved experimentally.24 However, except for a minor substitution,28 the suggested complete replacement of sulfur with selenium was never accomplished until now (cf. Figure S1). Hence, none of the respective quaternary selenides Li10+xTt1+xP2-xSe12 (Tt = Si, Ge, Sn) were reported, although the ternary Li4SnSe4, with an ionic conductivity of 2·10-5 S/cm, is known. Quaternary sodium-ion compounds of the type Na10+xSn1+xP2-xS12 have been reported only recently.29–31 The foundation was led by theoretical studies by the groups of Waghmare32 and Ceder,33 followed by the observation of a very high Na+ ion conductivity for Na11Sn2PS12 (3.7 mS/cm) in spite of a relatively large activation barrier (0.39 eV).30 Inspired by further theoretical studies22 and by the known fact that softening of the lattice may diminish the activation barrier,34 which would allow to envisage lower-

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

temperature applications of according electrolytes, besides their general suitability for usage in the rather new field of sodiumselenium batteries,35–37 we currently expand the studies towards the related selenides. To explore whether sodium, with its greater ionic radius as compared to lithium, might stabilize a potential Na11Sn2PSe12, we tried to transfer the newly found Sn:P ratio of 2:1 in this quaternary sulfide (which generally deviates from the ratio found in lithium compounds, Tt:P = 1:2) to the respective selenides.29,30,38 As a result, we present in this study the new superionic sodiumion conductor Na11.1Sn2.1P0.9Se12, which represents the first quaternary selenium-based Na+ superionic conductor reported to date. We note that Na11.1Sn2.1P0.9Se12 together with Na11Sn2PS1229,30 and the very recently reported Na4-xSn1-xSbxS438 (Figure 1) are the first examples of a new structural class of chalcogenide-based sodium ion conductors, for which we suggest the abbreviations NaSnPSe and NaSnPS, respectively, in analogy to LGPS (Li10GeP2S12) and LSnPS (Li10SnP2S12) used for the structurally related lithium ion conductors.23,26,27,39 Both these Li+ and Na+ ion conductors together form the overarching class of 1-4-5-6 electrolytes, named after the main group affiliation of the constituting elements.

Figure 1. Overview of the hitherto known combinations of tetrel (Tt), pentel (Pn) and chalcogen (E) atoms in quaternary 1-4-5-6 electrolytes, forming the Li+ compounds with the generalized formula Li10TtPn2E12 (left) and the Na+ compounds with the generalized formula Na11Tt2PnE12 (right). The newly found NaSnPSe is denoted as a purple sphere. The title compound NaSnPSe was synthesized in a high temperature reaction. Stoichiometric amounts of Na2Se, Sn, red phosphorous and selenium powder were mixed, ground in an agate mortar and vacuum-sealed in a quartz ampoule. A general temperature profile comprised a heating rate of 40 °C/h up to 600 °C, dwelling for 12 h and subsequent cooling down at a rate of 20 °C/h. In order to obtain single crystals of NaSnPSe (and also for Na11Sn2PS12, see Supporting Information), a slower cooling rate of 2 °C/h was applied. As typical for chalcogenide-based alkali metal superionic conductors, both phases are hygroscopic and (especially in the case of E = Se) also sensitive to (humid) air, thus need to be handled under dry and non-oxidative conditions. According to both single crystal and powder X-ray diffraction experiments, NaSnPSe is isostructural to the previously reported sulfide NaSnPS, and thus crystallizes in the space group I41/acd (Figure 2). The lattice parameters a and c increase from 13.5309(14) Å to 14.1562(10) Å (+4.4%) and from 27.125(4) Å to 28.322(2) Å (+4.6%) at 100 K (Table S2), respectively. The associated unit cell volume increases from 4966.2 Å to 5675.7 Å (+14.3%), which matches very well the increase reported for the substitution of selenium for sulfur in the structurally related ternary compound Na3PS4 both in its cubic (+14.3%) and its tetragonal modification (+13.9%).40 A detailed comparison of the sulfide versus selenide structure is provided in the Supporting Information (Table S3). However, although NaSnPSe is isostructural to NaSnPS, thus comprising tetrahedral [PSe4]3– and [SnSe4]4– anions (with one crystallographically independent site for each type of anion) and

six crystallographically independent Na+ sites, we also identified notable differences in their structural details. First we found different site occupational factors for the Na1–Na6 sites in comparison with the respective sites in the sulfide (Table S3). Second, the atomic position of the P atom (Wyckoff site 8a) revealed to be partially occupied by ca. 10% of a second Sn atom (thus increasing the amount of sodium atoms from 11 to 11.1 per formula unit). Third, we found a rotational disorder of the tetrahedral anion at this position in a 9:1 ratio. The (P/Sn)–Se bond lengths refine to the same (average) value in both orientations, indicating a rotational disorder of both [PSe4]3– and [SnSe4]4– tetrahedra. Despite extensive attempts and good single crystal data, it was not possible to obtain a reasonable superstructure model, so that we describe the structure in the mentioned disordered model. Whether the disorder of the tetrahedra is dynamic or static is not clear until now, but it clearly points towards a higher degree of positional freedom for the anionic positions as compared to the sulfide.

Figure 2. Cut-off from the crystal structure of NaSnPSe, viewed along [001] (a) and [100] (b). Sn atoms and [SnSe4]4– tetrahedra are drawn in grey, P atoms and [PSe4]3– tetrahedra in orange, Se atoms in dark red, and Na+ cations in different shades of blue in order to differentiate between the six crystallographically distinct sites. For a comprehensive visualization of the [PS4]3– tetrahedral, both possible orientations are shown. Their superimposition leads to the appearance as a cube. c: Rietveld refinement of a diffraction pattern collected at beamline BL9 of the DELTA synchrotron radiation facility, Dortmund/Germany (λ = 0.49594 Å; T = 295 K). Experimental data points (black crosses), calculated profile (red solid line) and difference curve (blue curve below) are shown. Purple vertical bars mark the positions of Bragg reflections. The refinement residuals are given in the inset. The ionic conductivity of annealed pellets of polycrystalline Na11Sn2PSe12 was examined by means of impedance spectroscopy over broad temperature and frequency ranges, enabling us to differentiate between grain, grain boundary, and electrode contributions to the overall impedance. In Figure 3a, the Nyquist plot of a NaSnPSe pellet measured at –100 °C is depicted. The spectrum shows exemplarily the three detected processes. The highfrequency semicircle is characterized by a capacitance of 23 pF/cm2 and is thus assigned to ion transport in the crystalline grains. A smaller intermediate-frequency semicircle is observed, which is characterized by a capacitance of 10 nF/cm2 indicating that this semicircle reflects a grain boundary impedance. The spike observed in the low frequency regime is ascribed to the electrode polarization. From a fit to the equivalent circuit, the grain resistance Rg and the grain boundary resistance Rgb were

ACS Paragon Plus Environment

Page 2 of 7

Page 3 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials obtained, and the grain conductivity σg   ∙  ∙ g and the total conductivity σtotal   ∙  ∙  g  gb  were calculated. In the respective Nyquist plot at 20 °C (Figure S2), the former two processes are not resolved any more due to the high ionic conductivity of the compound. Figure 3b displays an Arrhenius plot of the grain conductivities σg of both NaSnPSe and NaSnPS.30 The grain conductivity of NaSnPSe could be obtained in the temperature range from –120 to 0 °C and it exhibits an activation energy EA of 0.30 eV. This value is considerably lower than the one measured recently for Na11Sn2PS12.30 The extrapolated room-temperature grain conductivity of NaSnPSe, on the other hand, amounts to σg  3.0 mS/cm (σtot = 1.7 mS/cm) and is thus slightly lower than the conductivities observed for the best sulfide-based superionic Na+ conductor NaSnPS with σg = 3.7 mS/cm (σtot = 3.0 mS/cm) and the best selenide-based superionic conductor Na3SbSe4.30,41 However, this value is higher than for other high-performance chalcogenide-based electrolytes, like Na3SbS442,43 and Na3PSe4.44 The high Na+ ion mobility is confirmed by the fast spin lattice relaxation of 23Na, which yielded a relaxation time of T1 = 4.7(3) · 10−4 s. The static 23Na NMR powder spectrum shows a single peak only (Figure 3c). Even with an excitation band width in the MHz regime, no satellite transitions could be detected, which indicates that all Na+ ions are involved in fast ionic motion on the NMR time scale.

Figure 3. a: Nyquist plot of NaSnPSe after annealing, together with the equivalent circuit utilized for fitting the spectra. The different contributions to the total impedance, obtained from the fits, are highlighted in different colors. b: Arrhenius plot of the grain and total conductivity of NaSnPSe obtained from the mean values of measurements on three annealed NaSnPSe pellets. For comparison, the grain conductivity of NaSnPS is also shown30. c: Static 23Na NMR spectrum of NaSnPSe (90° pulse length of 0.5 µs). In contrast to the theoretical prediction for the related LGPS-type and Na7P3E11-type electrolytes (E = O, S, Se),45 the replacement of sulfur by selenium does apparently not lead to a conductivity increase. Even though the new NaSnPS-type electrolytes differ in both their stoichiometry and structure from their lithium analogues, one would intuitively assume to raise the ionic conductivity by expanding the lattice parameters and the polarizability of the lattice, like it was for instance observed for the replacement of oxygen by sulfur in Li10GeP2S12-xOx46 and for the replacement of sulfur by selenium in Na3PS4-xSex.40

In order to investigate whether for the new NaSnPS and NaSnPSe families of Na+ ion conductors, the same correlations for the lattice dynamics and the ionic transport can be observed as for Na3PS4−xSex and Li6PS5X (X = Cl, Br, I)34,40 we implemented ultrasonic speed of sound measurements. Substituting sulfur by the more easily polarizable selenium leads to a lower speed of sound and thus a lower Debye frequency νD, which points to a lower lattice rigidity. The positive correlation between the Debye frequency and the activation energy of the conduction process, as shown in Figure S2, is in good accordance with the trends recently reported by Kraft et al. and Krauskopf et al.34,40 The assumptions regarding the enhanced structural freedom and the results found for the activation energy of the conductivity are supported by Raman spectra (Figure 4). A blue shift of the Raman peaks associated with the symmetric stretches of the Sn–S and P– S bonds (A1,Sn–S and A1,P–S) by 5 cm–1 is observed for NaSnPS when compared to the ternary sulfides NaSnS and NaPS. This blue shift implies stronger Sn–S and P–S bonds in the quaternary compound, which should go along with weaker Na···S interactions. Qualitatively, the same blue shift is also observed for the A1,Sn–Se and A1,P–Se mode in NaSnPSe, when compared to Na4SnSe4 and Na3PSe4. Here, the A1,P–Se mode is even blue shifted by 8 cm–1. Thus, one might suspect that the strengthening of the P–Se bond in the quaternary solid and the concomitant weakening of the Na···Se interaction results in a more notable widening of the whole structure, as compared to the quaternary sulfide. Hence, we conclude the following: (a) the Na···chalcogen interaction is weaker in the selenide, which parallels the lower activation barrier observed in this Na+ ion conductor, (b) this observation is accompanied by new features like partial disorder of Sn versus P sites and the rotational disorder of the [PSe4]3– unit.

Figure 4. a, b: Raman spectra of NaSnPS in comparison with those of Na3PS4 and Na4SnS4. c, d: Raman spectra of NaSnPSe in comparison with those of Na3PSe4 and Na4SnSe4 (b and d are zooms of a and c, respectively). Additional bond valence site energy (BVSE) calculations47,48 for Na+ in various local structure models were conducted in order to investigate the influence of the observed structural disorder on the ion transport model (Figure 5 and Figure S7). Besides the partially occupied Na sites, the BVSE calculations shown in Figure 5a consistently reveal one type of low energy interstitial site, the position of which depends on the orientation of the (P/Sn)(Se3/Se3A) tetrahedra. When comparing pathways in NaSnPSe with those in NaSnPS, both phases exhibit a dense 3D

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

pathway network that may be thought of comprising two interpenetrating 3D pathway networks with similar migration barriers involving either Na1, Na4, Na5 and i1 or Na2, Na3, Na6 and i1. As discussed above, the larger unit cell volume and higher polarizability (cf. Figure S2) of the selenide will contribute to a reduction of the migration barriers in relaxed local structure models from 0.35 eV to 0.25 eV (c.f. Figure S8), but may not be the only factor. The seemingly counterintuitive combination of lower migration barrier with slightly lower room temperature conductivity may be caused by different relative contributions of the two sub-networks to the overall conductivity: Figure S8 shows that Na6 forms an extremely shallow minimum in NaSnPS, while the site energy of Na6 in NaSnPSe is considerably lowered and the minimum adopts a more regular shape. Along with the drastically reduced (= vanished) number of vacancies on the Na6 site, this reduced local mobility of Na6 limits the relative contribution of the (slightly higher energy) Na2–Na6–Na3–i1 pathway branches. Moreover, Sn on the Sn2/P site will impede transport via adjacent i1 site, while the local mobility of ions on the i1 sites around P will be higher in NaSnPSe. Overall, this reduces the density of the effective pathway network in local structure models of NaSnPSe, thus counterbalancing the lower migration barriers.

electrolytes based on the softBV forcefield49 and such relatively small supercells, the experimental conductivity is somewhat underestimated, while the observed activation energy of 0.28 eV for the temperature range up to 350 K harmonizes to the experimental findings (cf. Figure 3b). In accordance with the experiments, we find similar room-temperature conductivities, despite the clearly lower activation energies for NaSnPSe compared to NaSnPS. The Na+ trajectories for the supercell model harmonize with the predictions from the BVSE models (cf. Figures S9 and S10). It may be noticed that, as suggested from the static pathway models, the mobility in the low-energy local pathway cluster Na24–Na6–Na32 is somewhat less pronounced in the selenide. Furthermore, it is important to mention that the observed differences between NaSnPSe and NaSnPS are not caused by the higher Sn content of NaSnPSe, since in the MD simulations, only a small difference in the Na+ ion conductivity between a so far hypothetical stoichiometric compound Na11Sn2PSe12 and the Sn-enriched compound Na11.1Sn2.1P0.9Se12 (NaSnPSe) is observed (Figure S11). In conclusion, the first example of a quaternary selenium-based alkali superionic conductor, Na11.1Sn2.1P0.9Se12 (NaSnPSe) has been synthesized. With this new compound, we extended the spectrum of a new promosing structural class of Na+ conductors, which may be summarized as 1-4-5-6 electrolytes, towards the so far unprecedented selenides. While the Na+ ion conductivity is nearly as high as for the related sulfide (NaSnPS), the activation energy of the conduction process is significantly lower as a consequence of to the softer crystal lattice of the selenide, which is advantageous for battery-based energy storage at temperatures considerably lower than 298 K. Further substitutions are likely to additionally foster the ionic conductivity in future works as it has been shown for the lithium counterparts.

ASSOCIATED CONTENT Supporting Information The experimental data, more details on the computational studies as well as results of the structure refinements for both compounds (Tables S4 – S7) are reported in the Supporting Information. The crystallographic information files (CIF) with full structural information are deposited as Supporting Information. The Supporting Information is available free of charge on the ACS Publications website (PDF).

AUTHOR INFORMATION Figure 5. a: BVSE models of migration barriers for various relaxed local structure models of NaSnPSe, containing P(Se3)4 (violet squares); Sn2(Se3A)4 (dark grey diamonds), P(Se3A)4 (light blue triangles) or Sn2(Se3)4 (light grey circles). The lowest energy 3-dimensional (3D) pathway (migration barrier of 0.25 eV) is formed in the P(Se3)4 model by Na4–i1–Na1–i1–Na5 networks (including the split positions from the SCXRD structure). A separate 3D pathway is formed by interconnecting local low energy Na2-Na6-Na3 paths via the interstitial site i1. In the case of the Sn(Se3A)4 environment, these paths percolate with a migration barrier of 0.4 eV. At an only slightly higher migration barrier (0.41-0.44 eV) both networks merge to a common dense pathway network. b: Arrhenius plot of results from NVT molecular dynamics (MD) simulations for Na11.1Sn2.1P0.9Se12 and Na11Sn2PS12. Lines are linear fits as a guide to the eye. The activation barriers mentioned are based on regressions over the linear regime up to 350 K. The predictions of the static BVSE models were further tested by empirical molecular dynamics simulations of Na11Sn2PS12 (NaSnPS) and Na11.1Sn2.1P0.9Se12 (NaSnPSe), see Figure 5b. As commonly found for NVT simulations of sulfide-based solid

Corresponding Author [email protected].

Author Contributions The manuscript was written through contributions of all authors.

ACKNOWLEDGMENT M.D. acknowledges a fellowship by the Marburg Research Academy (MARA). The authors thank Melanie Werker and Dr. Christian Sternemann for collecting synchrotron powder diffraction data and DELTA (Dortmund/Germany) for providing synchrotron radiation. B. R. would like to thank the German Federal Ministry for Economy and Technology for financial support in the framework of the SME Central Innovation Programme (ZIM Grant ZF4430001ZG7). St. A. acknowledges support from the NUS “Centre for Energy Research” grant as well as from National Research Foundation, Prime Minister’s Office, Singapore under the Competitive Research Programme (CRP Award NRF-CRP 10-2012-6).

ACS Paragon Plus Environment

Page 4 of 7

Page 5 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

REFERENCES (1) (2)

(3) (4)

(5) (6)

(7)

(8)

(9)

(10)

(11)

(12)

(13)

(14)

(15)

(16)

(17)

(18)

(19)

(20)

(21)

(22)

Armand, M.; Tarascon, J.-M. Building Better Batteries. Nature 2008, 451, 652–657. Ellis, B. L.; Nazar, L. F. Sodium and Sodium-Ion Energy Storage Batteries. Curr. Opin. Solid State Mater. Sci. 2012, 16, 168–177. Hwang, J.-Y.; Myung, S.-T.; Sun, Y.-K. Sodium-Ion Batteries: Present and Future. Chem. Soc. Rev. 2017, 46, 3529–3614. Kundu, D.; Talaie, E.; Duffort, V.; Nazar, L. F. The Emerging Chemistry of Sodium Ion Batteries for Electrochemical Energy Storage. Angew. Chemie - Int. Ed. 2015, 54, 3432–3448. Slater, M. D.; Kim, D.; Lee, E.; Johnson, C. S. Sodium-Ion Batteries. Adv. Funct. Mater. 2013, 23, 947–958. Vaalma, C.; Buchholz, D.; Weil, M.; Passerini, S. A Cost and Resource Analysis of Sodium-Ion Batteries. Nat. Rev. Mater. 2018, 3, 18013. Billaud, J.; Clément, R. J.; Armstrong, A. R.; Canales-Vázquez, J.; Rozier, P.; Grey, C. P.; Bruce, P. G. β-NaMnO2: A HighPerformance Cathode for Sodium-Ion Batteries. J. Am. Chem. Soc. 2014, 136, 17243–17248. Kim, J. R.; Amatucci, G. G. NaMn2-xNixO4Derived from Mesoporous LiMn2-xNixO4: High-Voltage Spinel Cathode Materials for Na-Ion Batteries. J. Electrochem. Soc. 2016, 163, A696–A705. Yabuuchi, N.; Kajiyama, M.; Iwatate, J.; Nishikawa, H.; Hitomi, S.; Okuyama, R.; Usui, R.; Yamada, Y.; Komaba, S. P2-Type Nax[Fe1/2Mn1/2]O2 Made from Earth-Abundant Elements for Rechargeable Na Batteries. Nat. Mater. 2012, 11, 512–517. Kim, H.; Kim, H.; Ding, Z.; Lee, M. H.; Lim, K.; Yoon, G.; Kang, K. Recent Progress in Electrode Materials for Sodium-Ion Batteries. Adv. Energy Mater. 2016, 6, 1600943. Li, Y.; Hu, Y.-S.; Li, H.; Chen, L.; Huang, X. A Superior LowCost Amorphous Carbon Anode Made from Pitch and Lignin for Sodium-Ion Batteries. J. Mater. Chem. A 2015, 4, 96–104. Ponrouch, A.; Monti, D.; Boschin, A.; Steen, B.; Johansson, P.; Palacín, M. R. Non-Aqueous Electrolytes for Sodium-Ion Batteries. J. Mater. Chem. A 2015, 3, 22–42. Bhide, A.; Hofmann, J.; Katharina Dürr, A.; Janek, J.; Adelhelm, P. Electrochemical Stability of Non-Aqueous Electrolytes for Sodium-Ion Batteries and Their Compatibility with Na0.7CoO2. Phys. Chem. Chem. Phys. 2014, 16, 1987– 1998. Ponrouch, A.; Marchante, E.; Courty, M.; Tarascon, J.-M.; Palacín, M. R. In Search of an Optimized Electrolyte for Na-Ion Batteries. Energy Environ. Sci. 2012, 5, 8572–8583. Quartarone, E.; Mustarelli, P. Electrolytes for Solid-State Lithium Rechargeable Batteries: Recent Advances and Perspectives. Chem. Soc. Rev. 2011, 40, 2525–2540. Murayama, M.; Sonoyama, N.; Yamada, A.; Kanno, R. Material Design of New Lithium Ionic Conductor, Thio-LISICON, in the Li2S–P2S5 System. Solid State Ionics 2004, 170 , 173–180. Liu, Z.; Fu, W.; Payzant, E. A.; Yu, X.; Wu, Z.; Dudney, N. J.; Kiggans, J.; Hong, K.; Rondinone, A. J.; Liang, C. Anomalous High Ionic Conductivity of Nanoporous β-Li3PS4. J Am Chem Soc. 2013, pp 975–978. Hayashi, A.; Muramatsu, H.; Ohtomo, T.; Hama, S.; Tatsumisago, M. Improvement of Chemical Stability of Li3PS4 Glass Electrolytes by Adding MxOy (M = Fe, Zn, and Bi) Nanoparticles. J. Mater. Chem. A 2013, 1, 6320–6326. Dietrich, C.; Sadowski, M.; Sicolo, S.; Weber, D. A.; Sedlmaier, S. J.; Weldert, K. S.; Indris, S.; Albe, K.; Janek, J.; Zeier, W. G. Local Structural Investigations, Defect Formation, and Ionic Conductivity of the Lithium Ionic Conductor Li4P2S7. Chem. Mater. 2016, 28, 8764–8773. Chu, I. H.; Nguyen, H.; Hy, S.; Lin, Y. C.; Wang, Z.; Xu, Z.; Deng, Z.; Meng, Y. S.; Ong, S. P. Insights into the Performance Limits of the Li7P3S11 Superionic Conductor: A Combined FirstPrinciples and Experimental Study. ACS Appl. Mater. Interfaces 2016, 8, 7843–7853. Wang, Y.; Lu, D.; Bowden, M.; El Khoury, P. Z.; Han, K. S.; Deng, Z. D.; Xiao, J.; Zhang, J.-G.; Liu, J. Formation Mechanism of Li7P3S11 Solid Electrolytes through Liquid Phase Synthesis. Chem. Mater. 2018, 30, 990–997. Ong, S. P.; Mo, Y.; Richards, W. D.; Miara, L.; Lee, H. S.;

(23)

(24)

(25)

(26)

(27)

(28)

(29)

(30)

(31)

(32)

(33)

(34)

(35)

(36)

(37)

(38)

(39)

(40)

(41)

Ceder, G. Phase Stability, Electrochemical Stability and Ionic Conductivity of the Li10±1MP2X12 (M = Ge, Si, Sn, Al or P, and X = O, S or Se) Family of Superionic Conductors. Energy Environ. Sci. 2013, 6, 148–156. Bron, P.; Dehnen, S.; Roling, B. Li10Si0.3Sn0.7P2S12 - A LowCost and Low-Grain-Boundary-Resistance Lithium Superionic Conductor. J. Power Sources 2016, 329, 530–535. Hori, S.; Suzuki, K.; Masaaki, H.; Kato, Y.; Saito, T.; Yonemura, M.; Kanno, R. Synthesis, Structure, and Ionic Conductivity of Solid Solution, Li10+δM1+δP2−δS12 (M = Si, Sn). Faraday Discuss. 2014, 176, 83–94. Krauskopf, T.; Culver, S. P.; Zeier, W. G. The Bottleneck of Diffusion and Inductive Effects in Li10Ge1-xSnxP2S12. Chem. Mater. 2018, 30, 1791–1798. Kuhn, A.; Gerbig, O.; Zhu, C.; Falkenberg, F.; Maier, J.; Lotsch, B. V. A New Ultrafast Superionic Li-Conductor: Ion Dynamics in Li11Si2PS12 and Comparison with Other Tetragonal LGPSType Electrolytes. Phys. Chem. Chem. Phys. 2014, 16, 14669– 14674. Bron, P.; Johansson, S.; Zick, K.; Schmedt auf der Günne, J.; Dehnen, S.; Roling, B. Li10SnP2S12: An Affordable Lithium Superionic Conductor. J. Am. Chem. Soc. 2013, 135, 15694– 15697. Yang, K.; Dong, J.; Zhang, L.; Li, Y.; Wang, L. Dual Doping: An Effective Method to Enhance the Electrochemical Properties of Li10GeP2S12-Based Solid Electrolytes. J. Am. Ceram. Soc. 2015, 98, 3831–3835. Zhang, Z.; Ramos, E.; Lalère, F.; Assoud, J.; Kaup, K.; Hartman, P.; Nazar, L. F. Na11Sn2PS12: A New Solid State Sodium Superionic Conductor. Energy Environ. Sci. 2018, 11, 87–93. Duchardt, M.; Ruschewitz, U.; Adams, S.; Dehnen, S.; Roling, B. Vacancy-Controlled Na+ Superion Conduction in Na11Sn2PS12. Angew. Chemie Int. Ed. 2018, 57, 1351–1355. Yu, Z.; Shang, S.-L.; Gao, Y.; Wang, D.; Li, X.; Liu, Z.-K.; Wang, D. A Quaternary Sodium Superionic Conductor Na10.8Sn1.9PS11.8. Nano Energy 2018, 47, 325–330. Kandagal, V. S.; Bharadwaj, M. D.; Waghmare, U. V. Theoretical Prediction of a Highly Conducting Solid Electrolyte for Sodium Batteries: Na10GeP2S12. J. Mater. Chem. A 2015, 3, 12992–12999. Richards, W. D.; Tsujimura, T.; Miara, L. J.; Wang, Y.; Kim, J. C.; Ong, S. P.; Uechi, I.; Suzuki, N.; Ceder, G. Design and Synthesis of the Superionic Conductor Na10SnP2S12. Nat. Commun. 2016, 7, 11009. Kraft, M. A.; Culver, S. P.; Calderon, M.; Bo, F.; Krauskopf, T.; Senyshyn, A.; Dietrich, C.; Zevalkink, A.; Janek, J.; Zeier, W. G. Influence of Lattice Polarizability on the Ionic Conductivity in the Lithium Superionic Argyrodites Li6PS5X (X = Cl , Br , I ). 2017, 139, 10909–10918. Luo, C.; Xu, Y.; Zhu, Y.; Liu, Y.; Zheng, S.; Liu, Y.; Langrock, A.; Wang, C. Selenium@Mesoporous Carbon Composite with Superior Lithium and Sodium Storage Capacity. ACS Nano 2013, 7, 8003–8010. Li, Q.; Liu, H.; Yao, Z.; Cheng, J.; Li, T.; Li, Y.; Wolverton, C.; Wu, J.; Dravid, V. P. Electrochemistry of Selenium with Sodium and Lithium: Kinetics and Reaction Mechanism. ACS Nano 2016, 10, 8788–8795. Ding, J.; Zhou, H.; Zhang, H.; Stephenson, T.; Li, Z.; Karpuzov, D.; Mitlin, D. Exceptional Energy and New Insight with a Sodium–selenium Battery Based on a Carbon Nanosheet Cathode and a Pseudographite Anode. Energy Environ. Sci. 2017, 10, 153–165. Heo, J. W.; Banerjee, A.; Park, K. H.; Jung, Y. S.; Hong, S.-T. New Na-Ion Solid Electrolytes Na4−xSn1−xSbxS4 (0.02 ≤ X ≤ 0.33) for All-Solid-State Na-Ion Batteries. Adv. Energy Mater. 2018, 4, 1702716. Kamaya, N.; Homma, K.; Yamakawa, Y.; Hirayama, M.; Kanno, R.; Yonemura, M.; Kamiyama, T.; Kato, Y.; Hama, S.; Kawamoto, K.; et al. A Lithium Superionic Conductor. Nat. Mater. 2011, 10, 682–686. Krauskopf, T.; Pompe, C.; Kraft, M. A.; Zeier, W. G. Influence of Lattice Dynamics on Na+ Transport in the Solid Electrolyte Na3PS4−xSex. Chem. Mater. 2017, 29, 8859 − 8869. Wang, N.; Yang, K.; Zhang, L.; Yan, X.; Wang, L.; Xu, B.

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(42)

(43)

(44)

(45)

Improvement in Ion Transport in Na3PSe4–Na3SbSe4 by Sb Substitution. J. Mater. Sci. 2017, 18–21. Banerjee, A.; Park, K. H.; Heo, J. W.; Nam, Y. J.; Moon, C. K.; Oh, S. M.; Hong, S. T.; Jung, Y. S. Na3SbS4: A Solution Processable Sodium Superionic Conductor for All-Solid-State Sodium-Ion Batteries. Angew. Chemie - Int. Ed. 2016, 55, 9634– 9638. Wang, H.; Chen, Y.; Hood, Z. D.; Sahu, G.; Pandian, A. S.; Keum, J. K.; An, K.; Liang, C. An Air-Stable Na3SbS4 Superionic Conductor Prepared by a Rapid and Economic Synthetic Procedure. Angew. Chemie - Int. Ed. 2016, 55, 8551– 8555. Zhang, L.; Yang, K.; Mi, J.; Lu, L.; Zhao, L.; Wang, L.; Li, Y.; Zeng, H. Na3PSe4: A Novel Chalcogenide Solid Electrolyte with High Ionic Conductivity. Adv. Energy Mater. 2015, 5, 3–6. Wang, Y.; Richards, W. D.; Bo, S.-H.; Miara, L. J.; Ceder, G.

(46)

(47)

(48)

(49)

Computational Prediction and Evaluation of Solid-State Sodium Superionic Conductors Na7P3X11 (X = O, S, Se). Chem. Mater. 2017, 29, 7475–7482. Sun, Y.; Suzuki, K.; Hara, K.; Hori, S.; Yano, T.; Hara, M.; Hirayama, M.; Kanno, R. Oxygen Substitution Effects in Li10GeP2S12 Solid Electrolyte. J. Power Sources 2016, 324, 798– 803. Adams, S.; Rao, S.; P., R. Understanding Ionic Conduction and Energy Storage Materials with Bond-Valence-Based Methods. In Bond Valences; 2014; pp 129–159. Adams, S.; Rao, R. P. High Power Lithium Ion Battery Materials by Computational Design. Phys. Status Solidi 2011, 208, 1746–1753. Chem, J. M.; Adams, S.; Rao, R. P. Structural Requirements for Fast Lithium Ion Migration in Li10GeP2S12. J. Mater. Chem. 2012, 22, 7687–7691.

ACS Paragon Plus Environment

Page 6 of 7

Page 7 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials TOC Graphic

ACS Paragon Plus Environment