We report a new ion conductor—Na11Sn2SbSe12—as a possible candidate for a solid catholyte in composite cathodes for all-solid state Na-batteries, which exhibits a room temperature ionic conductivity of 0.15 ± 0.03 mS cm−1 and an activation energy of 0.39 ± 0.02 eV. The sulfide solid solutions of Na11Sn2SbSe12, namely, Na11Sn2SbS12−xSex (x = 1 and 6), were also investigated through a combination of Rietveld refinement against powder x-ray diffraction data and electrochemical impedance spectroscopy to reveal the complex structure–property relationships governing ion transport in this class of materials. Meanwhile, broadening of the Na-ion diffusion pathways in Na11Sn2SbSe12 is expected to facilitate Na-ion transport compared to the sulfide-rich member of the solid solution, the opposite holds: increasing the Se fraction in Na11Sn2SbS12−xSex leads to a little change in the activation energy but a reduction in the ionic conductivity. We ascribe this to the lowering of the prefactor, σ0, in the Arrhenius relationship with increasing lattice “softening” as a function of higher Se content.

Li-ion batteries are currently the exclusive energy storage choice for many applications, such as electromobility. However, the cost of lithium is anticipated to increase over the next decade as the market share of EVs increases and other energy storage-dependent applications such as mini-grids, drones, and robots become more market-intensive. This is exacerbated by the fact that much of the untapped lithium reserves are located in remote or politically sensitive areas.1–3 In addition, demand for cobalt, still a key component in Li-ion battery cathodes, and low producer inventories can create a bottleneck in the move toward improved energy storage systems.2,4 Promising alternatives are Na-ion batteries (NIB). Sodium technology is considered to be a suitable choice in terms of battery cost, safety, and raw materials abundance.1 Sodium is expected to be the next targeted element after lithium, based on its relatively low atomic weight, an electrochemical potential close to that of Li, and sustainability.1,3,5 A multitude of “cobalt-free” positive electrode materials have been developed for NIB applications.6–8 These cathodes include sulfur. Room-temperature Na–S batteries have prompted extensive research interest due to their high-charge storage capacity and abundance of both sodium and sulfur,9,10 but the low electronic conductivity of sulfur (5 × 10−28 mS cm−1) leads to very low utilization of the active material in the electrode as well as sluggish electrochemical reaction processes.9 Elemental selenium, compared to sulfur, is regarded as an alternative promising cathode material for sodium (and lithium)-ion batteries11,12 due to its considerably higher electrical conductivity (1 mS cm−1 for Se vs 0.5 × 10−24 mS cm−1 for S)11 and a comparable volumetric capacity (3253 Ah l−1 based on a density of 4.82 g cm−3 vs 3467 Ah l−1 based on a density of 2.07 g cm−3).13,14 However, the organic liquid electrolytes used in Na-ion batteries have safety concerns due to their flammable nature and have cost factors that limit the realization of NIBs.15–17 Solid-state NIBs with the integration of solid-state electrolytes (SEs) that form the basis for all-solid state batteries (ASSBs) are an alternative because they can potentially address these factors.5,18,19 Na-ion SSEs, an essential component for realizing ASSBs, have been the subject of several studies that have led to the development of a wide range of new materials. Extensive development has been achieved in chalcogenide sodium-ion conductors;20 for example, in 2015, the superionic conductor, cubic-Na3PSe4 (σ = 1.16 mS cm−1), was synthesized by Zhang et al.,18 followed by computational and experimental investigations of this material by Bo et al.21 The crystal structure of the antimony analog, cubic-Na3SbSe4, was first identified in 1989,22 although its ionic conductivity properties (σ = 3.7 mS cm−1; Ea = 0.19 eV) were not measured until 2018. In both cubic-Na3PSe4 and cubic-Na3SbSe4 electrolytes, the enhancement of ion conductivity was observed compared to their sulfide-based counterparts that exhibit values of 0.2 and 2.8 mS cm−1, respectively.23,24 The opposite is true for the “NSPS” sulfide, Na11Sn2PS12, whose ion conductivity was reported in the range of 3.7–2.5 mS cm−1, respectively.25,26 The synthesis of its antimony analog, Na11Sn2SbS12, was reported at about the same time,27 and the phospho-selenium analog, Na11Sn2PSe12, was announced by two research groups who reported lower but similar ionic conductivities of 1.74 and 2.15 mS cm−1 at 25 °C.28 The same selenide—prepared by a mechanochemical method—yielded an even slightly lower conductivity of 1 mS cm−1 after annealing, but, nonetheless, its very good performance as an electrolyte was demonstrated in Na–Se ASSBs.29 The electrochemical performance of the selenium-poor antimonide Na11Sn2SbS11.5Se0.5 in ASSBs has also been reported; powder x-ray diffraction (PXRD) was utilized to identify compositions on the Se poor side (x < 0.8 in Na11Sn2SbS12−xSex),30 but the structure was not resolved by either powder or single-crystal methods, and structure–property relationships were not elucidated. Despite the increasing effort in developing Se-based solid electrolytes, the number of Na–Se ion conductors reported today is still limited, and compared to sulfides and oxides, selenides present a rather large field of discovery.

Here, we report that Se-substitution in Na11Sn2SbS12 leads to the solid solution series Na11Sn2SbS12−xSex (x = 1, 6, and 12). The end member (x = 12) is a new Na-ion conductor (σi = 0.15 ± 0.03 mS cm−1) with an activation energy of 0.39 ± 0.02 eV. Differential scanning calorimetry (DSC) measurements show that the material is stable to at least 500 °C. Our work encompasses structural characterization using both single-crystal and powder x-ray diffraction (PXRD), bond valence site energy (BVSE)31,32 calculations to obtain the topology of ion-migration pathways in the framework, and electrochemical impedance spectroscopy. A combination of single-crystal x-ray diffraction and BVSE studies show that Na-ion conduction in Na11Sn2SbSe12 relies on three-dimensional (3D) diffusion pathways in the framework involving the vast majority of the Na sites. As shown by single-crystal x-ray diffraction studies of Na11Sn2SbS12−xSex (x = 1, 6, and 12), the Se substitution enlarges the unit cell and increases Se2− occupancy on the three 32g Wyckoff positions, indicating the complete solubility of selenium in the framework. A decrease in ionic conductivity and an increase in activation energy with increasing Se content are observed in the chalcogenide solutions, Na11Sn2SbS12−xSex (x = 0 → 12). This contradicts expectations of higher conductivity based on a larger lattice volume, wider Na-ion channels, and a more polarizable anion framework but can be explained by consideration of the role of the prefactor in the Arrhenius relationship.

Na11Sn2SbS12−xSex (x = 1, 6, and 12) were synthesized by combining Na2S (99%, Sigma-Aldrich), Na2Se (98%, Alfa Aesar), Na2Se with Sb2S3 (99%, Sigma-Aldrich), S powder (99.9%, Sigma-Aldrich), Se powder (99%, Sigma-Aldrich), Sn powder (99%, Sigma-Aldrich), and Sb powder (99%, Sigma-Aldrich) in the targeted ratios. Na2Se was prepared using elemental sodium and selenium with naphthalene as an electron transfer agent in tetrahydrofuran (THF) according to the procedure reported by Thompson and Boudjouk.33 The precursors were combined according to the respective molar ratios of the targeted compositions and ball milled at 400 rpm for 33 h to yield the amorphous mixtures that were loaded into glassy carbon crucibles and vacuum sealed in quartz tubes. The mixtures were heated to 750 °C at a heating rate of 94 °C/h, held at this temperature for 5 h, and then slow-cooled over 79 h from 750 to 545 °C, followed by ice quenching. The resulting agglomerate of crystals was gently separated to pick suitable candidates for single-crystal x-ray diffraction studies.

Block-shaped, single crystals of Na11Sn2SbS12−xSex (x = 1, 6, and 12) with dimensions of 0.070 × 0.060 × 0.020; 0.160 × 0.110 × 0.040; and 0.130 × 0.090 × 0.030 mm3, respectively, were obtained for crystal structure determination. X-ray diffraction data were collected at 280 K using a Bruker Kappa diffractometer equipped with a Smart Apex II CCD and graphite-monochromated Mo-Kα radiation. The single crystals were protected by a Paratone-N oil and liquid nitrogen flow using an Oxford Cryostream controller 700 at 280 K. Data were collected by scanning ω and ϕ with an increment of 0.3° in two groups of frames at different angles for complete data using the APEX II suite strategy. The exposure time was set to 30 s/frame. After unit cell indexation, the data were corrected for Lorentz and polarization effects, and multi-scan absorption corrections were applied using SADABS within the APEX II software suite. The structures were solved using the direct method, and all atoms were refined anisotropically by the least-squares fitting method incorporated in the SHELXTL package. The crystallographic structural details and bond distances are summarized in Tables I and S2–S10.

TABLE I.

Atomic coordinates, occupation factor, and isotropic displacement parameters of Na11Sn2SbSe12 obtained from single-crystal x-ray diffraction at 280 K.

AtomWyckoff sitexyzSOFU(eq) (Å2)
Sn 16e 0.207 53(2) 1/2 1/4 0.019 00(6) 
Sb 8a 1/4 3/8 0.018 27(7) 
Se1 32g 0.107 99(2) 0.396 42(2) 0.199 25(2) 0.024 96(7) 
Se2 32g 0.310 43(2) 0.403 68(2) 0.302 85(2) 0.021 82(7) 
Se3 32g 0.090 94(2) 0.359 43(2) 0.424 74(2) 0.030 67(8) 
Na1 32g 0.256 40(12) 0.271 57(11) 0.377 89(5) 0.926(6) 0.044 8(6) 
Na2 16d 1/4 0.247 18(9) 0.869(8) 0.060 5(11) 
Na3 16e 0.043 27(13) 1/2 1/4 0.962(7) 0.041 8(7) 
Na4 16c 1/4 1/4 1/4 0.894(8) 0.055 6(10) 
Na5 16f 0.227 92(10) 0.522 08(10) 3/8 0.969(8) 0.051 5(9) 
Na6 8b 1/4 1/8 0.058(10) 0.020(11) 
AtomWyckoff sitexyzSOFU(eq) (Å2)
Sn 16e 0.207 53(2) 1/2 1/4 0.019 00(6) 
Sb 8a 1/4 3/8 0.018 27(7) 
Se1 32g 0.107 99(2) 0.396 42(2) 0.199 25(2) 0.024 96(7) 
Se2 32g 0.310 43(2) 0.403 68(2) 0.302 85(2) 0.021 82(7) 
Se3 32g 0.090 94(2) 0.359 43(2) 0.424 74(2) 0.030 67(8) 
Na1 32g 0.256 40(12) 0.271 57(11) 0.377 89(5) 0.926(6) 0.044 8(6) 
Na2 16d 1/4 0.247 18(9) 0.869(8) 0.060 5(11) 
Na3 16e 0.043 27(13) 1/2 1/4 0.962(7) 0.041 8(7) 
Na4 16c 1/4 1/4 1/4 0.894(8) 0.055 6(10) 
Na5 16f 0.227 92(10) 0.522 08(10) 3/8 0.969(8) 0.051 5(9) 
Na6 8b 1/4 1/8 0.058(10) 0.020(11) 

X-ray diffraction measurements were performed on a PANalytical Empyrean instrument outfitted with a PIXcel bidimensional detector using Cu-Kα radiation. The microcrystalline powders were loaded into 0.3 mm capillaries in an argon-filled glovebox. Rietveld quality patterns were recorded in Debye–Scherrer geometry using a parabolic x-ray mirror in the incident beam. Rietveld refinements for Na11Sn2SbS12−xSex (x = 1, 6, and 12) were performed using the models initially obtained from single-crystal XRD analysis. Rietveld refinements of each crystal structure were performed, where the scale factor, Chebyshev background, peak shape, and lattice parameters were simultaneously refined using the software package TOPAS 6 (Bruker-AXS). Refinement constraints that were used were as follows: the atomic coordinates and atomic displacement parameters were fixed to be the same as those obtained from the single crystal studies.

AC impedance spectroscopy was initially performed on Na11Sn2SbS12−xSex solid solutions (x = 1, 6, and 12) using a VMP3 potentiostat/galvanostat (Bio-Logic Science Instruments) in the frequency range from 1 MHz to 0.1 Hz under 2–3 tons of constant pressure. The powders were cold pressed at 2 tons with stainless steel electrodes in a 10 mm diameter die to form pellets. The ionic conductivity was obtained using the equation σ = t/RA, where R is the total resistance of the solid electrolyte, t is the sample thickness (0.8 mm average), and A is the area of the solid electrolyte. The Nyquist plots of Na11Sn2SbS12−xSex (x = 1, 6, and 12) at room temperature were composed of one semicircle followed by a linear diffusion tail in the low-frequency region. The bulk and grain boundary contributions could not be deconvoluted, and therefore, the conductivities correspond to the total conductivity. The semicircle was fitted with a parallel circuit composed of a resistor (R) in parallel with a constant phase element (CPE) for the ionic transport, in series with a CPE that represents the blocking electrodes.34,35 The temperature dependent ion conductivity measurements of Na11Sn2SbS12−xSex (x = 1, 6, and 12) were performed using the same custom-made SwagelokTM cell. The powders were pelletized at 2 metric tons of pressure, placed between two indium foils, and pressed at a lower pressure (∼1 metric ton) to avoid indium being forced into the pellet and causing a short circuit. Impedance was measured at frequencies from 10 MHz to 0.3 Hz, every 5 °C, over a temperature range from 30 to 65 °C.

To separate the bulk conductivity of the grains from the grain boundaries in Na11Sn2SbSe12, EIS was also conducted from 35 MHz to 1 Hz at temperatures from −100 to 100 °C every 10 °C. Data were recorded with a Novocontrol Concept 80 spectrometer that was connected to a Quatro cryo system and equipped with a zero-gradient synchrotron (ZGS) active sample cell (Novocontrol). Au thin films were sputter deposited with a thickness of 200 nm on both sides of the samples. The bulk conductivity of Na11Sn2SbS12 was determined by fitting the data (Nyquist plots) over the range from −100 to −40 °C (Fig. S5). Each of the semicircles was fitted with a parallel circuit composed of a resistor (R) in parallel with a constant phase element (CPE) for the ionic transport, in series with a CPE that represents the blocking electrodes.

DC polarization measurements Na11Sn2SbSe12 and c-NaSbSe2 were performed by placing pellets between two stainless steel rods of 10 mm diameter and pressing them at 2 metric tons. Applied voltages of 0.25, 0.50, and 0.75 V for 30 min each were used. The experimental density of Na11Sn2SbSe12 was 98.8% of the theoretical value. This value was calculated from the sample geometry and mass and compared to the theoretical density of 3.49 g/cm3.

BVSE calculations were performed utilizing the SoftBV program with the bond valence parameter set developed by Chen and Adams.36 The structural models obtained by single-crystal diffraction were used as an input. Na ion diffusion pathways were identified with the regions of low bond valence site energy for a dense grid of points with a resolution of 0.1 Å covering the crystal structure using the transferable Morse-type SoftBV forcefield. BVSE maps with an isosurface level of 0.4 eV over the global minimum are shown.

Between 5 and 10 mg of Na11Sn2SbSe12 was placed in an aluminum pan and sealed under an argon atmosphere. Differential scanning calorimetry curves were obtained using a 10 °C/min heating and cooling rate up to 500 °C.

The precursors for Na11Sn2SbS12−xSex (x = 1, 6, and 12) were mixed by mechanochemical ball-milling (see Sec. II). The PXRD patterns of the as-milled precursors for Na11Sn2SbS12−xSex are shown in Fig. S1 (supplementary material). Halo patterns with no crystalline reflections that could be assigned to the starting materials were observed. After the mechanochemical process, a heat treatment was then conducted on the amorphous precursor mixtures (see Sec. II) to crystallize the desired materials. Figure 1(a) presents the room temperature PXRD patterns of the heat-treated Na11Sn2SbS12−xSex (x = 1, 6, and 12) materials, with the pattern of our previously reported Na11Sn2SbS12 (x = 0)27 shown for comparison. As illustrated in Fig. 1(b), owing to the larger ionic radius of Se vs S (1.98 vs 1.84 Å), as the Se content increases, the Bragg reflections, i.e., (408), show a shift toward higher d-spacing relative to Na11Sn2SbS12, indicating the successful synthesis of a chalcogenide solid solution. Different impurity phases are observed throughout the Na11Sn2SbS12−xSex solid solution (Table S1); secondary-phase formation of Na3SbS4 and NaSbSe2 are identified for x = 6 and x = 12, respectively, while a small unidentified impurity was formed for x = 1 (also observed in x = 0). The addition of small amounts of excess selenium (1 wt. %) was implemented in Na11Sn2SbSe12 to compensate for any sodium loss during synthesis (750 °C) and reduce the secondary phase formation of NaSbSe2 (Table S1), although the PXRD pattern still reveals its formation as a minority phase. Differential scanning calorimetry (DSC) was used to investigate the thermal stability of the Na11Sn2SbSe12 end member. The DSC curve of Na11Sn2SbSe12 (Fig. S2) shows no other significant contributions but only very minor signals at 176 and 318 °C, indicating that Na11Sn2SbSe12 is stable up to 500 °C.

FIG. 1.

(a) Room-temperature powder x-ray diffraction patterns of Na11Sn2SbS12−xSex (x = 0, 1, 6, and 12) with Na3SbS4, NaSbSe2, and an unidentified impurity as labeled; (b) Expanded view on the right showing the peak shift in the (408) reflection toward the lower 2θ angle.

FIG. 1.

(a) Room-temperature powder x-ray diffraction patterns of Na11Sn2SbS12−xSex (x = 0, 1, 6, and 12) with Na3SbS4, NaSbSe2, and an unidentified impurity as labeled; (b) Expanded view on the right showing the peak shift in the (408) reflection toward the lower 2θ angle.

Close modal

The opportunity to establish a detailed structural understanding of the solid solution series relies on the successful growth of single crystals of appropriate size and quality for single-crystal XRD studies. These were obtained for all three compositions (see Sec. II). Based on the single-crystal data (Tables I and S2–S10), we find the crystal structures of Na11Sn2SbS12−xSex (x = 1, 6, and 12) are isostructural to Na11Sn2SbS12 and were solved in the same tetragonal space group I41/acd (no. 142) Z = 8. A comparison of the unit cell parameters of Na11Sn2SbS12−xSex materials is presented in Fig. 2. The linear expansion of the unit cell volume from 5236 Å3 for x = 0 to 5933 Å3 for x = 12, with increasing Se content (+13% volume increase) stems from a significant increase in the lattice parameters (+4.3% in a and +4.1% in c), in accordance with the results obtained from powder XRD.

FIG. 2.

Unit cell parameters of Na11Sn2SbS12−xSex with x = 0, 1, 6, and 12 obtained from single x-ray diffraction data at 280 K: (a) volume; (b) lattice parameters a and c.

FIG. 2.

Unit cell parameters of Na11Sn2SbS12−xSex with x = 0, 1, 6, and 12 obtained from single x-ray diffraction data at 280 K: (a) volume; (b) lattice parameters a and c.

Close modal

Figure 3(a) shows a representative powder x-ray diffraction pattern and Rietveld refinement against the x-ray data for Na11Sn2SbSe12. The refinements for Na11Sn2SbS12−xSex (x = 1, 6, and 12) materials confirmed the formation of the targeted phases together with different impurity phases (Fig. S3). For x = 6, 4.75 wt. % of Na3SbS4 is present, while for x = 12, 12.5 wt. % of NaSbSe2 is observed. An estimate of the amorphous content in x = 12 (∼17%) was obtained by fitting the total area of the pattern (accounting for both crystalline and amorphous contributions) using TopasTM [Eq. (S1)], although this represents an overestimate since the glass capillary contributes to the background at ∼20° and 50° in 2θ. Figure 3(b) displays the crystal structure of Na11Sn2SbSe12 along the [100] and [001] directions. The structure of Na11Sn2SbS12−xSex encompasses two Wyckoff positions, 8a and 16e, for the polyhedral anions [Sb(Q)4]3− and [Sn(Q)4]4− (Q = S/Se), respectively. As expected, the volume of these polyhedra increased with the degree of selenium substitution (Fig. S4). Na-ions are distributed over six distinct crystallographic sites, five of which are octahedrally coordinated by chalcogenide ions and connected through face-sharing, while the “interstitial” Na(6) (8b Wyckoff position) site is loosely bonded to eight anions forming a quasi-cubic environment as previously described for Na11Sn2SbS12.27 

FIG. 3.

Rietveld refinements of XRD data of (a) Na11Sn2SbSe12. NaSbSe2 is present as a minor secondary phase (12.5 wt. %). The black circles correspond to the data points, the red line denotes the calculated pattern, and the difference map is shown in blue. Olive vertical ticks correspond to Bragg reflections. (b) View of the structure of Na11Sn2SbSe12 along the [100] and [001] directions. Thermal ellipsoids are drawn at 50% probability. Color code: Partially vacant sites Na(2) and Na(4): light-rose; almost fully occupied sites Na(1), Na(3), and Na(5): red; low occupied site Na(6): pink; Sn: gray; Sb: purple; Se: orange; SnSe4 and SbSe4 polyhedral are shown in gray and purple, respectively.

FIG. 3.

Rietveld refinements of XRD data of (a) Na11Sn2SbSe12. NaSbSe2 is present as a minor secondary phase (12.5 wt. %). The black circles correspond to the data points, the red line denotes the calculated pattern, and the difference map is shown in blue. Olive vertical ticks correspond to Bragg reflections. (b) View of the structure of Na11Sn2SbSe12 along the [100] and [001] directions. Thermal ellipsoids are drawn at 50% probability. Color code: Partially vacant sites Na(2) and Na(4): light-rose; almost fully occupied sites Na(1), Na(3), and Na(5): red; low occupied site Na(6): pink; Sn: gray; Sb: purple; Se: orange; SnSe4 and SbSe4 polyhedral are shown in gray and purple, respectively.

Close modal

Representative single-crystal data for Na11Sn2SbSe12 is shown in Table I. The Na content was refined to 11.15(6), slightly higher than the targeted stoichiometry but the same as reported for the x = 0 composition.27 The effect of any additional minor sodium content on the conductivity is expected to be negligible, as shown by first principle studies.37 Two of the five sodium sites are almost, but not quite, fully occupied [Na(3), site occupancy factor: SOF = 0.96 and Na(5), SOF = 0.97], and three sites are partially vacant [Na(1), SOF = 0.93; Na(2), SOF = 0.87; and Na(4), SOF = 0.89]. The sixth site is almost unoccupied [Na(6), SOF = 0.058]. In Na11Sn2SbS12−xSex (x = 1, 6, and 12), as expected and confirmed by bond length data (Tables S5, S8, and S10), the Na–Q bond length increases as the Se content increases (from 2.70 Å in x = 1 to 2.92 Å in x = 12), while all the Na+–Na+ interatomic distances are in the range of 3.5–3.67 Å in x = 12, resulting in equidistant Na-ion transport through the framework. Additional structural features, such as the increase of Se2− occupancy on the three 32g Wyckoff positions as the Se content increases in Na11Sn2SbS12−xSex, were observed (Tables I, S3, and S6 and Fig. S4). The overall change in the Se occupancy on all three lattice sites suggests no site preference for Se.

Based on the single-crystal diffraction data of Na11Sn2SbS12−xSex materials, all five sodium sites [Na(1)–Na(5)] have an occupancy higher than 80%, while the probability that a sodium ion occupies the Na(6) site is rather low (SOF <14%) (Tables I, S3, and S6). The Na(6) site is a higher-energy site that is mostly vacant. Our previous work38 and that of Oh et al.37 unequivocally rule out Na-ion diffusion through the Na(6) site as an important pathway for Na-ion conduction. The diffusion channels created by face-sharing Na octahedra in the Na11Sn2SbSe12 framework form a strategic alternating “full\vacancy\full” arrangement created by the partially vacant sodium sites at the cross-over points [32g (Na(1)) ∼SOF = 93%; 16d (Na(2)) ∼SOF = 87%; 16c (Na(4)) ∼SOF = 89%] alternating with the almost fully occupied sodium sites [16e (Na(3))] and 16f (Na(5)) that spans in all three crystallographic directions (Table I). This alternating arrangement is essentially the same as in the sulfide Na11Sn2SbS12 (and its P analog).25,27 There is a slightly higher occupancy of the cross-over sites compared to Na11Sn2SbS12 [Na(1) ∼SOF = 89% and Na(2) ∼SOF = 84%] that may affect transport, but to a very minor degree.

The long-range Na-ion transport properties of Na11Sn2SbS12−xSex materials were examined by using the conventional complex impedance formalism (Table S11). A comparison between the ionic conductivities and activation energies obtained from EIS data for Na11Sn2SbS12−xSex (x = 1, 6, and 12) is shown in Figs. 4(a) and 4(b). The error in the obtained conductivities was determined from the span in the measurements of various samples of the same composition, and the values of the impedance analyses are presented in Table S12. The room temperature ionic conductivity decreases from 0.56 mS cm−1 for Na11Sn2SbS1227 to 0.34(2) mS cm−1 for Na11Sn2SbS11Se and to 0.15(3) mS cm−1 for Na11Sn2SbSe12, while the activation energy increases only slightly within error from 0.3427 to 0.37(2) and 0.39(2) eV, respectively [Figs. 5(a), 5(b), and S5]. Meanwhile, the total ionic conductivity of Na11Sn2SbSe12 may be slightly impacted by the minority phase NaSbSe2 (∼12 wt. %, Table S1), the overall trend in lower conductivity with increasing Se content suggests this is mainly governed by the bulk properties of the materials. This conclusion is especially true from x = 0 → 6, where the impurity fraction is only in the region of 4%–5%. The higher occupancy of the cross-over points, as mentioned above, may impact fast ion conduction by lowering the frequency of cooperative Na-ion hopping both in the ab plane and along the c axis.

FIG. 4.

(a) Room temperature Nyquist impedance plot of Na11Sn2SbS12−xSex (x = 1, 6, and 12) with the respective equivalent circuit shown. The fitted data are presented in Table S11; (b) Arrhenius plot of the total conductivity of Na11Sn2SbS12−xSex (x = 1, 6, and 12) from 30 to 65 °C.

FIG. 4.

(a) Room temperature Nyquist impedance plot of Na11Sn2SbS12−xSex (x = 1, 6, and 12) with the respective equivalent circuit shown. The fitted data are presented in Table S11; (b) Arrhenius plot of the total conductivity of Na11Sn2SbS12−xSex (x = 1, 6, and 12) from 30 to 65 °C.

Close modal
FIG. 5.

Data for Na11Sn2SbS12−xSex (x = 0, 1, 6, and 12). (a) Ionic conductivity as a function of Se content; (b) activation energy, Ea. The data for Na11Sn2SbS12 were taken from Ref. 27. The dotted lines are shown for guidance to the eye.

FIG. 5.

Data for Na11Sn2SbS12−xSex (x = 0, 1, 6, and 12). (a) Ionic conductivity as a function of Se content; (b) activation energy, Ea. The data for Na11Sn2SbS12 were taken from Ref. 27. The dotted lines are shown for guidance to the eye.

Close modal

As expected, Se substitution gives rise to an increase in the unit cell lattice parameters and hence the unit cell volume [5236 Å3 for S vs 5933 Å3 for Se, Fig. 2(a)], which should favor Na-ion diffusion. However, the decrease in ionic conductivity and increase in activation energy as the Se content increases contradict expectations based on the larger lattice volume and more polarizable anion framework of Na11Sn2SbSe12. The Se−2 substitution in Na11Sn2SbS12−xSex makes the lattice softer, i.e., creating longer/weaker bonds (Fig. S6) affecting all cation–anion interactions, but leads to an unexpected trend in increasing activation energy with a more polarizable anion framework. We note that lower ion conductivity for Se vs S analog is not without precedent; for example, as mentioned in the introduction, Na11Sn2PS12 exhibits an ion conductivity between 3.7 and 2.5 mS cm−1,25,26 whereas the phospho-selenium analog, Na11Sn2PSe12, displays lower ionic conductivities between 1.7 and 2.15 mS cm−1.4,28

A report on argyrodites Li6PS5X (X = Cl, Br, I)39 reveals the contradictory contributions to the conductivity from the Arrhenius prefactor and Ea, namely, that as the lattice polarizability increases, Ea often decreases due to softening of the lattice, but the prefactor σ0 also decreases owing to a decrease in the jump frequency. That leads to an overall decrease in conductivity with the increasing softness of the anion lattice in the argyrodites. For the Na11Sn2SbS12−xSex solid solution, a relatively small change in Ea is observed on the Se substitution (from 0.34 eV for x = 0 to 0.39 eV for x = 12), suggesting that the drop in the prefactor (see Fig. 4)—as dictated by the higher polarizability of Seδ− compared with Sδ−—likely governs the ionic conductivities of Na11Sn2SbS12−xSex. This may also explain the lower conductivity of Na11Sn2PSe12 vs Na11Sn2PS12 (see above). Further studies will be conducted to deconvolute the influence of lattice vibrations (softening of the lattice) on ionic transport, but these are beyond the current scope of the work.

The BVSE approach was used to explore the potential ion migration pathways in Na11Sn2SbSe12. Figures 6(a)6(c) display the bond valence model of Na-ion migration pathways in Na11Sn2SbSe12 along the a,b-plane and c-axis, respectively, with the structure in the polyhedral motif shown for comparison [Figs. 6(b)6(d)]. The BVSE maps, depicted as isosurfaces of constant bond valence site energy, suggest 3D interconnected network pathways that are displayed as yellow isosurfaces. These involve all the Na sites that form the major conduction pathways in the framework, for example, Na(4)-Na(2)-Na(4) and Na(2)-Na(3)-Na(2) chains along the a,b-plane [Fig. 6(b)] and Na(4)-Na(1)-Na(3)-Na(1)-Na(4) chains along the c-axis [Fig. 6(d)].

FIG. 6.

(a) Bond valence map model of Na-ion migration pathways in Na11Sn2SbSe12 visualized as isosurfaces of constant bond valence site energy in the ab plane; (b) View of the diffusion channels interconnecting Na(4)–Na(2)–Na(4) and Na(2)–Na(3)–Na(2) along the ab plane; (c) Bond valence map model of Na-ion migration pathways visualized as isosurfaces along the c axis; (d) View of the diffusion channels interconnecting Na(4)–Na(1)–Na(3)–Na(4) sites along the c axis. Color code: Na(2), Na(3), and Na(4): light rose; Na(1): red.

FIG. 6.

(a) Bond valence map model of Na-ion migration pathways in Na11Sn2SbSe12 visualized as isosurfaces of constant bond valence site energy in the ab plane; (b) View of the diffusion channels interconnecting Na(4)–Na(2)–Na(4) and Na(2)–Na(3)–Na(2) along the ab plane; (c) Bond valence map model of Na-ion migration pathways visualized as isosurfaces along the c axis; (d) View of the diffusion channels interconnecting Na(4)–Na(1)–Na(3)–Na(4) sites along the c axis. Color code: Na(2), Na(3), and Na(4): light rose; Na(1): red.

Close modal

The room temperature electronic conductivity of Na11Sn2SbSe12 was measured by the DC polarization of the SS/Na11Sn2SbSe12/SS symmetric cell. Figure S7(a) shows the DC polarization curves at three different voltages: 0.25, 0.50, and 0.75 V. The small potential applied to Na11Sn2SbSe12 between the two SS blocking electrodes leads to a decrease in the current until a steady-state value is reached. From a linear fit of voltage vs stabilized current [Fig. S7(b)], the electronic conductivity for Na11Sn2SbSe12 was estimated to be 4.9 ± 0.6 × 10−4 mS cm−1, which is three orders of magnitude lower than the ionic conductivity (1.5 × 10−1 mS cm−1). The secondary-phase c-NaSbSe2 should not contribute to the overall electronic conductivity as σe is similar (2 × 10−3 mS cm−1; Fig. S8), and its low mass fraction will not result in discrete percolation pathways. Meanwhile the propensity of Na11Sn2SbSe12 to reduction at a sodium electrode would obviate the application of Na11Sn2SbSe12 as a separator, the other properties of this 3D Na-ion conductor (fast Na-ion transport and a thermally stable structure) make it a potential candidate as a SE in composite cathode materials for ASSB applications.

In this work, Se2− substitution in Na11Sn2SbS12−xSex x = 1, 6, and 12 led to the synthesis of the new Na-ion conductor, Na11Sn2SbSe12, with a room temperature ionic conductivity of 0.15 mS cm−1 and an activation energy of 0.39 eV. Based on the single-crystal diffraction studies performed on Na11Sn2SbS12−xSex (x = 1, 6, and 12), we find that with increasing Se2− substitution, a uniform increase in lattice parameters and occupancy of the three 32g Wyckoff positions by Se2− was observed. Bond valence energy site calculation results show that the Na11Sn2SbSe12 framework is comprised of a 3D interconnected network likely involving all Na sites forming the major conduction pathways, as previously reported for the sulfide analog. Meanwhile the increase of the Na channel volume with Se2− substitution should be more beneficial for Na-ion transport, increasing the Se fraction in Na11Sn2SbS12−xSex leads to an increase in the activation energy (0.34 eV for x = 0 to 0.39 eV for x = 12) and a reduction in the ionic conductivity (0.56 mS cm−1 for x = 0 to 0.15 mS cm−1 for x = 12). The higher occupancy of the cross-over point [Na(1) with a SOF of 93%] may impact the ion conduction by lowering the prospect of cooperative 3D Na-ion diffusion, but this is not expected to be a major contribution. We note that such “anomalous” behavior has been previously reported, as the conductivity of Na11Sn2PS12 (3.7–2.5 mS cm−1) is higher than that of its Se analog, Na11Sn2PSe12 (1.7–2.15 mS cm−1). We attribute this to the fact that a decrease in the prefactor, σ0, owing to a softer anion lattice, may govern the decrease in conductivity when Ea remains fairly constant. Further studies need to be conducted to deconvolute the influence of lattice vibrations (softening of the lattice) on ionic transport. Nonetheless, these detailed structural solutions of the solid solution series—Na11Sn2SbS12−xSex—provide a better understanding of the transport properties of this structure type.

See the supplementary material for additional information on the crystallography and electrochemistry data.

L.F.N. acknowledges the NSERC for financial support of this work through their Discovery and Canada Research Chair programs and partial support from the Joint Center for Energy Storage Research, an Energy Innovation Hub funded by the U.S. Department of Energy (DOE), Office of Science, Basic Energy Sciences. E.P.R. thanks Lawrence Livermore National Laboratory for their support via the career development program. D.R. acknowledges financial support from the Austrian Federal Ministry for Digital and Economic Affairs, the National Foundation for Research, Technology and Development, and the Christian Doppler Research Association (Christian Doppler Laboratory for Solid-State Batteries).

The authors have no conflicts to disclose.

Erika P. Ramos: Conceptualization (equal); Formal analysis (lead); Investigation (lead); Writing – original draft (lead); Writing – review & editing (equal). Abdeljalil Assoud: Investigation (equal); Methodology (equal). Laidong Zhou: Investigation (equal); Validation (equal). Abhinandan Shyamsunder: Investigation (equal). Daniel Rettenwander: Investigation (equal). Linda F. Nazar: Conceptualization (equal); Funding acquisition (lead); Supervision (lead); Writing – review & editing (lead).

The data that support the findings of this study are available within the article and its supplementary material.

1.
V.
Palomares
,
P.
Serras
,
I.
Villaluenga
,
K. B.
Hueso
,
J.
Carretero-González
, and
T.
Rojo
,
Energy Environ. Sci.
5
,
5884
(
2012
).
2.
M.
Li
,
J.
Lu
,
Z.
Chen
, and
K.
Amine
,
Adv. Mater.
30
,
1800561
(
2018
).
3.
J.-J.
Kim
,
K.
Yoon
,
I.
Park
, and
K.
Kang
,
Small Methods
1
,
1700219
(
2017
).
4.
M.
Duchardt
,
S.
Neuberger
,
U.
Ruschewitz
,
T.
Krauskopf
,
W. G.
Zeier
,
J.
Schmedt auf der Günne
,
S.
Adams
,
B.
Roling
, and
S.
Dehnen
,
Chem. Mater.
30
,
4134
(
2018
).
5.
Z.
Yu
,
L.
Shang
,
J.
Hwan Seo
,
D.
Wang
,
X.
Luo
,
Q.
Huang
,
S.
Chen
,
J.
Lu
,
X.
Li
,
Z.
Kui Liu
, and
D.
Wang
,
Adv. Mater.
29
,
1605561
(
2017
).
6.
Juliette
Billaud
,
J. Raphaele
Clément
,
A. Robert
Armstrong
,
Jesus
Canales-Vázquez
,
Patrick
Rozier
,
Clare  P.
Grey
, and
Peter G.
Bruce
,
J. Am. Chem. Soc.
136
,
17243
(
2014
).
7.
N.
Yabuuchi
,
M.
Kajiyama
,
J.
Iwatate
,
H.
Nishikawa
,
S.
Hitomi
,
R.
Okuyama
,
R.
Usui
,
Y.
Yamada
, and
S.
Komaba
,
Nat. Mater.
11
,
512
(
2012
).
8.
J. R.
Kim
and
G. G.
Amatucci
,
J. Electrochem. Soc.
163
,
A696
(
2016
).
9.
H.
Wang
,
Y.
Jiang
, and
A.
Manthiram
,
Adv. Energy Mater.
8
,
1701953
(
2018
).
10.
S.
Xin
,
Y.-X.
Yin
,
Y.-G.
Guo
, and
L.-J.
Wan
,
Adv. Mater.
26
,
1261
(
2014
).
11.
J.
Ding
,
H.
Zhou
,
H.
Zhang
,
T.
Stephenson
,
Z.
Li
,
D.
Karpuzov
, and
D.
Mitlin
,
Energy Environ. Sci.
10
,
153
(
2017
).
12.
X.
Li
,
J.
Liang
,
J.
Luo
,
C.
Wang
,
X.
Li
,
Q.
Sun
,
R.
Li
,
L.
Zhang
,
R.
Yang
,
S.
Lu
,
H.
Huang
, and
X.
Sun
,
Adv. Mater.
31
,
1808100
(
2019
).
13.
C.
Luo
,
Y.
Xu
,
Y.
Zhu
,
Y.
Liu
,
S.
Zheng
,
Y.
Liu
,
A.
Langrock
, and
C.
Wang
,
ACS Nano
7
,
8003
(
2013
).
14.
L.
Wang
,
X.
Zhang
,
L.
Deng
,
J.
Tang
,
H.
Deng
,
W.
Hu
, and
Z.
Liu
,
ACS Appl. Mater. Interfaces
11
,
4995
(
2019
).
15.
C.
Sun
,
J.
Liu
,
Y.
Gong
,
D. P.
Wilkinson
, and
J.
Zhang
,
Nano Energy
33
,
363
(
2017
).
16.
S.
Xin
,
Y.
You
,
S.
Wang
,
H.-C.
Gao
,
Y.-X.
Yin
, and
Y.-G.
Guo
,
ACS Energy Lett.
2
,
1385
(
2017
).
17.
C.
Zhou
,
S.
Bag
, and
V.
Thangadurai
,
ACS Energy Lett.
3
,
2181
(
2018
).
18.
L.
Zhang
,
K.
Yang
,
J.
Mi
,
L.
Lu
,
L.
Zhao
,
L.
Wang
,
Y.
Li
, and
H.
Zeng
,
Adv. Energy Mater.
5
,
1501294
(
2015
).
19.
L.
Zhang
,
D.
Zhang
,
K.
Yang
,
X.
Yan
,
L.
Wang
,
J.
Mi
,
B.
Xu
, and
Y.
Li
,
Adv. Sci.
3
,
1600089
(
2016
).
20.
J. W.
Heo
,
A.
Banerjee
,
K. H.
Park
,
Y. S.
Jung
, and
S.-T.
Hong
,
Adv. Energy Mater.
8
,
1702716
(
2018
).
21.
S.-H.
Bo
,
Y.
Wang
,
J. C.
Kim
,
W. D.
Richards
, and
G.
Ceder
,
Chem. Mater.
28
,
252
(
2016
).
22.
Brigitte
Eisenmann
and
Roland
Zagler
,
Z. Naturforsch.
44b
,
249
256
(
1989
).
23.
A.
Hayashi
,
K.
Noi
,
A.
Sakuda
, and
M.
Tatsumisago
,
Nat. Commun.
3
,
856
(
2012
).
24.
D.
Zhang
,
X.
Cao
,
D.
Xu
,
N.
Wang
,
C.
Yu
,
W.
Hu
,
X.
Yan
,
J.
Mi
,
B.
Wen
,
L.
Wang
, and
L.
Zhang
,
Electrochim. Acta
259
,
100
(
2018
).
25.
Z.
Zhang
,
E.
Ramos
,
F.
Lalère
,
A.
Assoud
,
K.
Kaup
,
P.
Hartman
, and
L. F.
Nazar
,
Energy Environ. Sci.
11
,
87
(
2017
).
26.
M.
Duchardt
,
U.
Ruschewitz
,
S.
Adams
,
S.
Dehnen
, and
B.
Roling
,
Angew. Chem., Int. Ed.
57
,
1351
(
2018
).
27.
E. P.
Ramos
,
Z.
Zhang
,
A.
Assoud
,
K.
Kaup
,
F.
Lalère
, and
L. F.
Nazar
,
Chem. Mater.
30
,
7413
(
2018
).
28.
Z.
Yu
,
S.-L.
Shang
,
D.
Wang
,
Y. C.
Li
,
H. P.
Yennawar
,
G.
Li
,
H.-T.
Huang
,
Y.
Gao
,
T. E.
Mallouk
,
Z.-K.
Liu
, and
D.
Wang
,
Energy Storage Mater.
17
,
70
(
2019
).
29.
R. P.
Rao
,
X.
Zhang
,
K. C.
Phuah
, and
S.
Adams
,
J. Mater. Chem. A
7
,
20790
(
2019
).
30.
H.
Wan
,
L.
Cai
,
W.
Weng
,
J. P.
Mwizerwa
,
J.
Yang
, and
X.
Yao
,
J. Power Sources
449
,
227515
(
2020
).
31.
S.
Adams
and
R. P.
Rao
, in
Bond Valences
, edited by
I. D.
Brown
and
K. R.
Poeppelmeier
(
Springer
,
Berlin, Heidelberg
,
2014
), pp.
129
159
.
32.
S.
Adams
and
R. P.
Rao
,
Phys. Status Solidi A
208
,
1746
(
2011
).
33.
D. P.
Thompson
and
P.
Boudjouk
,
J. Org. Chem.
53
,
2109
(
1988
).
34.
P.
Vadhva
,
J.
Hu
,
M. J.
Johnson
,
R.
Stocker
,
M.
Braglia
,
D. J. L.
Brett
, and
A. J. E.
Rettie
,
ChemElectroChem
8
,
1930
(
2021
).
35.
V.
Vivier
and
M. E.
Orazem
,
Chem. Rev.
122
,
11131
(
2022
).
36.
37.
K.
Oh
,
D.
Chang
,
I.
Park
,
K.
Yoon
, and
K.
Kang
,
Chem. Mater.
31
,
6066
(
2019
).
38.
Z.
Zhang
,
P.-N.
Roy
,
H.
Li
,
M.
Avdeev
, and
L. F.
Nazar
,
J. Am. Chem. Soc.
141
,
19360
(
2019
).
39.
M. A.
Kraft
,
S. P.
Culver
,
M.
Calderon
,
F.
Böcher
,
T.
Krauskopf
,
A.
Senyshyn
,
C.
Dietrich
,
A.
Zevalkink
,
J.
Janek
, and
W. G.
Zeier
,
J. Am. Chem. Soc.
139
,
10909
(
2017
).

Supplementary Material