Skip to main content
Science Advances logoLink to Science Advances
. 2022 Mar 18;8(11):eabj7698. doi: 10.1126/sciadv.abj7698

Enhancing ionic conductivity in solid electrolyte by relocating diffusion ions to under-coordination sites

Lei Zhu 1,, Youwei Wang 2,3,, Junchao Chen 1,4,†,*, Wenlei Li 5, Tiantian Wang 2,3, Jie Wu 1, Songyi Han 1, Yuanhua Xia 6, Yongmin Wu 1, Mengqiang Wu 5, Fangwei Wang 7, Yi Zheng 1, Luming Peng 4, Jianjun Liu 2,3,8,*, Liquan Chen 7, Weiping Tang 1,9,*
PMCID: PMC8932667  PMID: 35302845

Abstract

Solid electrolytes are highly important materials for improving safety, energy density, and reversibility of electrochemical energy storage batteries. However, it is a challenge to modulate the coordination structure of conducting ions, which limits the improvement of ionic conductivity and hampers further development of practical solid electrolytes. Here, we present a skeleton-retained cationic exchange approach to produce a high-performance solid electrolyte of Li3Zr2Si2PO12 stemming from the NASICON-type superionic conductor of Na3Zr2Si2PO12. The introduced lithium ions stabilized in under-coordination structures are facilitated to pass through relatively large conduction bottlenecks inherited from the Na3Zr2Si2PO12 precursor. The synthesized Li3Zr2Si2PO12 achieves a low activation energy of 0.21 eV and a high ionic conductivity of 3.59 mS cm−1 at room temperature. Li3Zr2Si2PO12 not only inherits the satisfactory air survivability from Na3Zr2Si2PO12 but also exhibits excellent cyclic stability and rate capability when applied to solid-state batteries. The present study opens an innovative avenue to regulate cationic occupancy and make new materials.


Ionic conductivity of solid electrolytes is enhanced with expanded bottleneck and reduced coordination for conductive ions.

INTRODUCTION

All-solid-state batteries (ASSBs) are attracting ever increasing attention for solving the intrinsic drawbacks of current Li-ion batteries, such as leakage and flammability of liquid electrolytes, and limited energy densities (1, 2). The success of ASSBs depends on solid electrolytes with satisfying Li-ionic conductivities. A high electrochemical stability at the interface of the solid electrolytes with electrode materials in a wide range of working potentials also plays an important role in designing ASSBs, for any electrochemical instability of solid electrolytes may result in interfacial decomposition products with a low ionic conductivity (3, 4). Recent studies on ASSBs have made considerable progress in high-conductivity inorganic solid electrolytes, which are generally classified into sulfide and oxide solid electrolytes (5). The most remarkable features of sulfide solid electrolytes such as Li10GeP2S12 (6) and Li6PS5X (X = Cl, Br, I) (7, 8) are their high ionic conductivity, comparable with organic liquid electrolytes at room temperatures, but low stability against Li metal to decompose into interfacial side product of Li2S (9). Oxide solid electrolytes such as Li7La3Zr2O12 (10) and LiPON (11) usually exhibit higher electrochemical stability at the interface in working potentials (12). Despite tremendous research advancements in oxide solid electrolytes, it still faces a large challenge to further improve their ionic conductivity to meet the requirement of practicable application (13).

In the past decades, a great deal of effort has been made for developing oxide solid electrolytes such as LISICON (14), NASICON (15, 16), perovskite (17, 18), garnet (10, 19), and LiPON (20) systems. The basic step in diffusion paths is Li-ion migration between two stable sites through the high-energy transition state. In these compounds, Li-ions usually occupy the lower-energy hexahedral sites and commonly surmounting high-energy and narrow bottleneck (transition state) into tetrahedral sites (Fig. 1A). Reducing the activation energy heights of transition states along the long-range diffusion path is of much importance in increasing ionic conductivity. For many years, optimization of these solid electrolytes has largely proceeded by substituting skeleton structural elements to enlarge the bottleneck and reduce the migration barrier (3, 21), which only achieve the ionic conductivities of the order of 10−6 to 10−3 S cm−1 at room temperature and the activation energy of 0.25 to 0.6 eV. Achieving a low migration barrier requires the bottleneck between two different polyhedrons to satisfy a very specific size. Statistical analysis indicates that those NASION-type lithium solid electrolytes with a bottleneck size of >2.05 Å exhibit a much lower migration barrier (Fig. 1B) (22). Intrinsically, regulating bottleneck size corresponds to change of polyhedron size, which may lead to Li+ occupancy changes from high-coordination to under-coordination sites for retaining structural stabilization (Fig. 1C). Therefore, the regulating coordination structure between Li+ ions and O2− anions provides an important strategy for breaking through this contradiction between occupancy stability and ionic conductivity.

Fig. 1. Relationships among coordination modulation, bottleneck, and migration energy barrier.

Fig. 1.

(A) Structural schematic of coordination engineering for the transition between LiO6 hexahedrons and LiO4 tetrahedrons. (B) Migration barrier of Li+ ions from LiO6 hexahedrons to LiO4 adjacent tetrahedrons. (C) Relations between bottleneck and the migration energy barrier of lithium-ion in NASICON-type lithium solid electrolytes (22).

In this work, a skeleton-retained cationic exchange approach based on Na3Zr2Si2PO12 (NZSP) is proposed to synthesize advanced solid electrolyte Li3Zr2Si2PO12 (LZSP), in which Li+ ions self-adaptively occupy under-coordinated sites and maintain relatively large bottleneck sizes by inheriting the pristine framework. The obtained LZSP is quite distinct from the low-ionic conductivity LiZr2(PO4)3 system, which is prepared by the common sol-gel method or solid-state method with a high calcination temperature of 900° to 1200°C (12, 23, 24). In combination with electrochemical impedance spectroscopy (EIS) and density functional theory (DFT) calculations, we demonstrate that the designed LZSP material with under-coordination Li-ions achieves a bulk ionic conductivity of up to 3.59 × 10−3 S cm−1 at room temperature and further performs excellent electrochemical, thermal, and air stability, which shows potential promising for practical application. The present work opens an innovative avenue to design new solid electrolytes by skeleton-retained cationic exchange approach.

RESULTS

Synthesis and local coordination of Li+ ions

The NASICON-type superionic conductor NZSP has both relatively high sodium ionic conductivity, ~10−4 S cm−1, and excellent structural stability with strong resistance against air and exceptional thermal stability (25, 26). Previous studies reveal that two-thirds of Na+ ions occupy the hexa-coordinated sites in the NZSP lattice structure, while the others occupy the octa-coordinated sites (25). In the three-dimensional (3D) channels enclosed by SiO44−/PO43− and ZrO68− polyhedrons, the hexa-coordinated Na+ ions contribute to the ionic conductivity by following the migration path of occupied octahedrons and unoccupied tetrahedrons. However, the octa-coordinated Na+ ions perform poor transport activity. Therefore, the structure-stable NZSP is used as a precursor to synthesize the corresponding solid electrolyte of LZSP through a skeleton-retained cationic exchange approach.

The sodium superionic conductor NZSP was synthesized by a sol-gel method (27). To accomplish full cationic exchange from Na+ to Li+, two basic principles about the selection and usage of solid electrolyte precursor, liquid carrier, and lithium compound [NZSP, 1-ethyl-3-methylimidazolium bis(trifluoromethylsulfonyl) imide (EMIIM), and bis(trifluoromethane) sulfonimide lithium (LiTFSI; Li+ cation coordinating with the N element of the TFSI anion), respectively] should be satisfied: (i) The selected liquid carrier (EMIIM in this work) is capable of dissolving the lithium compound (LiTFSI in this work) and simultaneously would not dissolve the solid electrolyte precursor; (ii) except for cation exchange, no other chemical reaction takes place. EMIIM and LiTFSI are selected as the solution environment and lithium source, respectively, to drive the cation exchange process. The operation temperature and concentration difference are intrinsically two key driving forces for promoting the cation substitution. Therefore, optimizing concentration difference and operation temperature is of great importance to balance in exchange effectiveness and damaging the skeleton of the solid electrolyte precursor due to inadequate and excessive additive. As shown in fig. S1, Na+ ions inside NZSP are thoroughly replaced by Li+ ions within LiTFSI dissolved EMIIM at 453 K (NZSP is not soluble in EMIIM). Unlike the commonly adopted molten salt ion-exchange method that requires high-temperature calcination (>673 K) (28), this approach can be conducted at a relatively low temperature (≤453 K), giving rise to the ability to restrain phase transition and keep the pristine framework. The usage of liquid carrier, ionic liquid—an essential Li+ ionic conductor, enables the fast removal of Na+ ions and continuous entry of Li+ ions, where liquids without ionic conductivity (e.g., water) may fail. In addition, the NZSP powders are more readily dispersed evenly in ionic liquids than solidified Li+ carrier, which further facilitates the Li+/Na+ ionic exchange.

The textures of “precursor” NZSP and “product” LZSP nanocrystals are determined by diffraction techniques and Rietveld refinement analysis (Fig. 2). The refined powder synchrotron x-ray diffraction (XRD) and neutron powder diffraction (NPD) patterns of NZSP and LZSP nanopowders (Fig. 2, A and B) crystallize with a monoclinic phase, which belongs to the crystal group of C12/c1 (space group #15). In the locally amplified synchrotron XRD patterns in Fig. 2A (see untruncated data in fig. S2), the peaks of (200) and (400) as well as the (020) and (021) facets all shift to higher angles from NZSP to LZSP, manifesting the lattice contraction of the a and b axis (a axis: 15.6 → 14.9 Å; b axis: 9.1 → 8.6 Å), respectively. Compared to those peaks, the signal of (111) facet just shows a less offset degree, corresponding to the slight elongation along the c axis (9.2 → 9.3 Å). NPD was then used to investigate the coordination evolution of conductive ions during the replacement of Na+ ions with Li+ ions. The NPD patterns of NZSP and LZSP (Fig. 2B) are deconvoluted by Rietveld refinements, from which the Wyckoff sites of each atom and the structural parameters consistent with the above results obtained from synchrotron XRD (tables S1 to S4). Particularly, the NPD technique is powerful in revealing the accurate atomic coordination of light elements (e.g., H and Li) (29, 30). Three intense diffraction peaks [(200), (021), and (130) facets] aroused by Na+ ions and Li+ ions are extracted (Fig. 2B and fig. S3). These characterizations show that the retained skeleton performs a little contraction to accommodate the small size of Li+ ions. Scanning electron microscopy (SEM) images (Fig. 2B, inset) exhibit the similar sizes of NZSP and LZSP particles (around 600 nm), indicating the maintained microscopic structure at the submicrometer level during cationic exchange. Energy-dispersive x-ray spectroscopy shows a homogeneous distribution of all constituent elements for NZSP and LZSP solids (fig. S4).

Fig. 2. Crystal structures of NZSP and LZSP.

Fig. 2.

(A) Refined powder synchrotron XRD and (B) NPD patterns of NZSP and LZSP powders with the Rietveld method. The SEM images of NZSP and LZSP are inserted in (B). All these patterns are recorded at room temperature. (C) Retained polyhedral skeletons containing the evolution path of Na+ ions to Li+ ions. (D) Evolution of Na+ ion to Li+ ion polyhedrons within unchanged skeleton.

The retained skeleton structure is further supported with 17O solid-state nuclear magnetic resonance (ssNMR), which is a high-sensitive technique to characterize the local structure of oxides at the atomic level (31, 32). As shown in fig. S5, their alike resonances observed in the 17O ssNMR spectra show that NZSP and LZSP crystals have a similar skeleton structure. In the skeleton structure, Wyckoff sites of Zr, Si/P, and O ions fully occupy the Wyckoff sites of 8f, 4e/8f, and 8f, respectively (figs. S6 and S7 and tables S1 and S2). In contrast, Na+ and Li+ occupancy sites in NZSP and LZSP exhibit an obvious difference, respectively. For NZSP, Na+ ions are distributed in three Wyckoff sites, fractional occupancy 4d (80%), full occupancy 4e (100%), and 8f (60%), respectively, corresponding to hexa-coordinated NaI, octa-coordinated NaII, and hexa-coordinated NaIII (fig. S2 and table S1) (22, 25). These high-coordinated polyhedrons with Na+ occupancy are connected with under-coordinated tetrahedrons, forming 3D migration channels such as -octahedron-tetrahedron-octahedron-, -dodecahedron-tetrahedron-dodecahedron-, and, -octahedron-tetrahedron-dodecahedron-. When cationic exchange is fully accomplished, Li+ ions either invade Na+-occupied polyhedrons (LiII → NaII) or create pioneering occupancy in tetrahedrons following the migration channels of -octahedron-tetrahedron (LiIV)-octahedron- and -dodecahedron-tetrahedron (LiIV)-dodecahedron-. These Li+ ions in polyhedrons are relaxed into equivalent under-coordination structures with penta-coordinated LiII and tetra-coordinated LiIII, LiIV, and LiV sites (fig. S7 and table S2). Our XRD and NPD analysis indicates that cationic exchange induces coordination transformation from high-coordinated occupancy (hexa- and octa-coordinated) of Na+ ions into under-coordinated occupancy (tetra- and penta-coordinated Li+) of Li+ ions. Except for 6% occupancy NaIIIO6 → LiIIIO4, most of the NaO6 octahedrons become vacant and Li+ ions are fully occupied in the tetrahedrons connected with two octahedrons. The Li+ replacement makes 100% occupancy NaIIO8 dodecahedrons into 43.8% occupancy LiIIO5 (V denotes vacancy site; fig. S7). As shown in Fig. 2 (C and D), the ionic substitution from Na+ to Li+ ions does not result in skeleton structure change. However, Na+ ions in NZSP occupy high-coordination structures (hexa-/octahedral sites), while exchanged Li+ ions in LZSP preferably occupy under-coordination structures such as tetrahedron and pentahedron. Their occupancy structure stability is further discussed in the DFT calculations.

Detailed coordination evolution and ion transport from Na+ ions to Li+ ions during the cationic exchange were studied through quantitative 23Na ssNMR and DFT calculations. As the process of Li+ ions substituting Na+ ions undergoes, the peaks in the 23Na ssNMR spectra move from high frequencies to low frequencies in fig. S8, which means that under-coordinated Na+ ions are more preferential to be substituted by Li+ ions than high-coordinated ones. The NMR results are in accord with the thermodynamic calculations in fig. S9, in which the formation energies of Li+-substituted under-coordinated Na+ ions (NaI sites and NaIII sites) are about 0.06 eV lower per Li+-substituted Na+ than those of Li+-substituted high-coordinated Na+ ions (NaII sites). In addition, except for NZSP and LZSP, two intermediate compounds, LiNa2Zr2Si2PO12 and Li2NaZr2Si2PO12, were further identified by controlling exchange temperature and time, and then tested with EIS (fig. S10). The presence of intermediates indicates that the skeleton-retained cationic exchange approach can also be used to develop hybrid inorganic superionic conductors (see figs. S10 to S14 and more discussions in text S1).

Ionic conductivity of the LZSP solid electrolyte

Before electrochemical tests, the content of unswapped Na+ ions and residual ionic liquid was determined to evaluate their potential influences. As shown in fig. S15 and table S5, there are no detectable Na, N, and F relevant impurity; their potential influence on electrochemical performances is therefore not considered further. The ionic conductivity of LZSP solid electrolyte was subsequently tested with EIS. In Fig. 3A, the Nyquist plots record from 10° to 60°C comprised one semicircle at the high- to medium-frequency region and a tail at the low-frequency region, which are attributed to the impedance of grain boundary (Rgb) and the Li+ transfer resistance between electrolyte and Au blocking electrodes. The semicircle corresponding to the bulk resistance (Rbulk) is missing, which is limited by the frequency measuring range. The values of Rbulk and Rgb, however, can be fitted with an equivalent circuit of the form Rbulk, Rgb, Qgb, and Qelectrode, where R is the resistance and Q is the constant phase element (10, 33). LZSP pellets were used to assemble Au/LZSP/Au symmetric cells. Generally, to densify powdered ceramic solid electrolytes, the most commonly used approach is to press the powders into pellets and then sinter the pellets under high temperatures (typically 1200°C or higher) (19, 34). However, thermal analysis indicates that LZSP powders would undergo phase transition after 742°C (fig. S16). Although its phase stability temperature of 742°C is lower than that (1000°C) of usual NASICON-type solid electrolytes (12, 35), this temperature is much higher than the reported cathode dissociation temperature in the previous literature (36). Therefore, a hot-pressing sintering process at 700°C for 1 hour under a fixed pressure of 70 MPa was carried out to prepare LZSP pellets (the synchrotron XRD and 17O ssNMR spectra in fig. S17 prove that the structure of LZSP annealed under 700°C keeps unchanged). The prepared LZSP pellet (fig. S18) is characterized to have a thickness of 2.4 mm and a relative density of about 83% (the porosity is 17%). At 20°C, it exhibits the bulk and total Li+ conductivities of 3.59 × 10−3 (fig. S19) and 8.75 × 10−6 S cm−1, respectively. When temperature is increased to 60°C, the bulk and total Li+ conductivities of LZSP are enhanced to 9.0 × 10−3 and 1.64 × 10−5 S cm−1, respectively. Two approaches could be applied to make denser pellet and improve total ionic conductivity (3739). On the one hand, we prepared composite pellets for solid-state batteries, where SN and LiTFSI were added to fill the pores between LZSP particles. With a higher relative density of 95%, the total ionic conductivity increases to 8.71 × 10−4 S cm−1 at 20°C, as shown in fig. S20. On the other hand, the denser LZSP pellet could be prepared through advanced vacuum hot-pressing technique. The outstanding conductivity benefits from the expanded bottleneck sizes and the improved Li+ transportation. Moreover, the ionic conductivity increases in line with elevated temperature, which accords with the Arrhenius curves (40). A reliable activation energy for the mobility of Li+ ions in LZSP is accordingly normalized to be 0.21 eV (Fig. 3B). The comparative study indicates that LZSP has the top value of bulk ionic conductivity among all reported NASICON-type solid electrolytes and is superior to other undoped oxide solid electrolytes (which are marked with horizontal lines in Fig. 3C).

Fig. 3. Comparison of LZSP with other undoped oxide solid electrolytes on ionic conductivities.

Fig. 3.

(A) Nyquist plots of LZSP pellets from 10° to 60°C. (B) Arrhenius plot of the ionic conductivity of LZSP. (C) Summary of bulk ionic conductivities of oxide solid electrolytes at room temperature; the materials are extracted from previous studies whose details are displayed in table S6.

Diffusion mechanism and kinetics

To reveal the influence mechanism of under-coordinated Li+ configuration (tetra-/penta-coordinated Li+ ions) on ionic conductivity of LZSP, DFT calculations combined with ab initio molecular dynamics (AIMD) simulations were carried out to characterize Li+ occupancy, structural stability, and ionic conductivity of LZSP. The structural evolution of LZSP was investigated in fig. S14. Compared with the LZSP configuration of Li+ ions substituting Na+ ions in NZSP without structural relaxation, the LZSP configuration within structural relaxation has a further reduced volume with the contraction from 1.14 to 10.2%. Meanwhile, some of the Li+ ions, which take the initial sites of hexa-coordinated NaI sites and NaIII sites, move to the sites of tetra-coordinated LiIV sites and the other Li+ ions move slightly off the sites of octa-coordinated NaII sites to form penta-coordinated LiII sites with Wyckoff site transformation from 4e to 8f (see fig. S21 and table S7). Compared with hexa-coordinated Li+ ion in traditional phosphatic solid electrolytes, the coordination of penta-/tetra-coordinated Li+ ions in produced LZSP is decreased. The coordination reduction of Li+ ions, accompanied by lattice contraction, leads to a 0.19 eV/formula lower free energy. Furthermore, different LZSP configurations, by tuning LiII occupancies from 50 to 31.25% with a constant difference of 6.25%, are built to screen the thermodynamically stable structure. The configuration with 43.8% occupancy LiII sites is found to have the lowest free energy, of which the ab place is compressed and the c axis is expanded compared with other LZSP configurations (fig. S14). The kinetic stability of screened LZSP configuration was investigated by mean square displacement (MSD) from AIMD simulations in canonical ensemble (NVT). As the temperature increases from 300 to 1000 K, the MSDs of Si/P and Zr atoms are near invariable and the maximal MSD of O atoms is ~0.2 Å2 (fig. S22). Furthermore, the phonon spectrum of screened LZSP configuration was calculated. As shown in figs. S23 and S24, the imaginary frequencies are largely ascribed to Li atoms and slightly ascribed to O atoms, which have electrostatic interactions with Li atoms. As a result, the frame of screened LZSP configuration with 43.8% occupancy LiII sites is stable and is used to investigate the Li+ ionic conductivity of LZSP.

To reveal the predominance of the under-coordinated structure in high-conductivity solid electrolytes, we calculated Li+ and Na+ migration barriers within LZSP and NZSP in the dilute limit of single Li+/Na+ and without other cations presented in corresponding compounds, respectively. The migration paths with corresponding relative energies of NZSP and LZSP are presented in Fig. 4 (A and B), respectively. An obvious discrepancy is different coordination structures of low-energy intermediates. The low-energy intermediates in NZSP have an octa-/hexa-coordinated structure character, whereas those in LZSP are changed into penta-/tetra-coordinated sites. In terms of calculated potential energy profiles, NZSP presents an isolated NaIV-NaI-NaIV-NaIII migration channel in which Na+ ions are difficult to surmount the migration barrier of 0.53 eV into NaII sites. The nonactivity of 100% occupancy NaII is further characterized by AIMD simulations at 600 K in Fig. 5. In a notable contrast, the LiII sites with 43.8% occupancy are fully activated by reducing the migration barrier. The Li+ migration barrier (0.29 eV) from the penta-coordinated LiII site to the tetra-coordinated LiIV site is 0.15 eV lower than the corresponding Na+ migration barrier (0.44 eV) from the octa-coordinated NaII site to the tetra-coordinated NaIV site. Intrinsically, the active effect in LZSP plays an important role in not only making the run-through channel network structure but also increasing mobile Li-ion, both of which can effectively enhance ionic conductivity.

Fig. 4. Single-ion migration mechanisms of Na+ ion and Li+ ion from CI-NEB.

Fig. 4.

(A) Energy barriers for single-ion migration of Na+ ion following the NaII-NaIV-NaI-NaIV-NaIII trajectory. (B) Energy barriers for single-ion migration of Li+ ion following the LiII-LiIV-LiI-LiIV-LiIII trajectory. The blue and purple balls stand for Na+ ions and Li+ ions, respectively. The olive, light green, pink, and sky-blue polyhedrons identify the NaIIO8 dodecahedron, LiIIO5 pentahedron, NaIVO4/LiIVO4 tetrahedrons, and NaIO6/LiIO6/ NaIIIO6/LiIIIO6 octahedrons, respectively.

Fig. 5. Diffusion kinetics of NZSP and LZSP from AIMD simulations.

Fig. 5.

(A) Diffusion pathways in monoclinic NZSP at 600 K. (B) Diffusion pathways in monoclinic LZSP at 600 K. (C) Arrhenius plot of Na+ diffusivity D as a function of temperature T in NZSP. (D) Arrhenius plot of Li+ diffusivity D as a function of temperature T in LZSP.

The lower migration barrier should be attributed to a suitable bottleneck size for migration ions. Previous studies show that the transition state structure of ionic conduction commonly corresponds to Li+ passing through the bottleneck between two polyhedrons (27). According to the Shannon effective ionic radius table, Na+ and Li+ ions have obviously different critical sizes of the bottleneck, 2.37 Å for Na-O and 1.94 Å for Li-O (41). Those migration paths, of which the bottlenecks are smaller than this critical size, would block the migration of Na+/Li+ ions. Since cationic exchange leads to little effect on the bottleneck size, 2.16 Å for the migration path of NaII-NaIV and 2.13 Å for the migration path of LiII-LiIV, Li+ ions are capable of passing through the bottleneck but Na+ ions fail. The comparison shows that NaII migration is inactive in NZSP but exchanged LiII migration is active in LZSP.

In comparison, the transition state along migration paths LiIV-LiI-LiIV and NaIV-NaI-NaIV exhibits clear different structures, although they have a similar barrier height (both are 0.17 eV). The former corresponds to the bottleneck structure between tetrahedron and octahedron, whereas the latter corresponds to the hexa-coordinated LiI structure. Their bottleneck sizes are very similar, 2.37 Å for the migration path of NaIV-NaI and 2.38 Å for the migration path of LiIV-LiI. The transition state structure change shows that such a large-size bottleneck does not inhibit Li+ migration. This firstly reported migration mechanism, enlarging the bottleneck size to reduce the migration barrier, is of much importance to design the structure of solid electrolytes. Recently, Zhang et al. (27) reported the high-energy site of NaV in NZSP by increasing the Na+-ion concentration. In this work, we also found low-occupancy LiV sites in LZSP, originating from unoccupied LiII sites. These high-energy sites of LiV increase the coulombic repulsions of Li+ ions following the trajectory path of -octahedron-tetrahedron (LiIV)-octahedron- and, depending on the concentration of high-energy sites, enhance the correlated migration of Li+ ions.

Direct comparison for two DFT-calculated potential energy profiles (Fig. 5) demonstrates that cationic exchange has made a notably lower migration barrier in under-coordinated LiII-LiIV than in high-coordinated NaII-NaIV. To verify the activation effect, AIMD simulations at a temperature of 600 K were performed to reveal the migration path change induced by cationic exchange. The Na+ and Li+ diffusion pathways of NZSP and LZSP from AIMD simulations under the same simulation conditions are presented in Fig. 5 (A and B), respectively. It is obvious that octa-coordinated NaII sites with 100% occupancy exhibit negligible ionic migration into the NaI-NaIII pathway. But hexa-coordinated NaI and NaIII with fractional occupancy present continuous and intertwined diffusion pathways, indicating the 2D migration mechanism. In contrast, LZSP presents the 3D migration mechanism since LiII sites in the corresponding position of NaII become active by forming the LiII-LiIV migration pathway. The activation effect is attributed to the size of the bottleneck located in the under-coordinated LiII-LiIV segment, which is 0.19 Å larger than the sum of Li+ and O2− radii.

According to extensive AIMD simulations at different temperatures, diffusion lengths of Na+ and Li+ ions at different sites in NZSP and LZSP are correlated with simulated temperatures. The fitted activation energies for different sites are also presented in Fig. 5 (C and D). It is obvious that the activation energies from AIMD in Fig. 5 are much lower than the migration barrier of the single-ion direct-hopping mechanism in Fig. 4 because the correlated migration is the preferred conduction mechanism in LZSP and NZSP, which was studied by Zhang et al. (27). The activation energy of NaII (0.22 eV) is nearly two times of those of NaI and NaIII (0.12 eV). However, the activation energy of LiII is comparable with those of LiIII, LiIV, and LiV. Octa-coordinated NaII sites with full occupancy only maintain equivalent vibration and do not exhibit the migration path. In contrast, the LiII sites in the similar position are turned into active ion migration along the LiII-LiIV path, which should be attributed to fractional occupancy and lower migration barrier of LiII. The activation effect plays an important role in not only turning the 2D into 3D migration path but also increasing the percentage of mobile ion rising from 67% in NZSP to 100% in LZSP. As a result of increased mobile ion and decreased migration barrier, the ionic conductivity of LZSP is theoretically predicted as high as 1.46 × 10−2 S cm−1.

In addition to thermal stability and ionic conductivity, other properties such as chemical stability against air and the tolerance of voltage are also investigated under electrochemical conditions. At room temperature, the LZSP powders are aged for 0, 100, and 200 hours within air to test the environmental stability. The 6Li ssNMR, powder synchrotron XRD, and C 1s x-ray photoelectron spectra (XPS) of these samples were recorded afterward and shown in fig. S25. No Li2CO3 relevant signals (a typical production as a result of air erosion) (42) were observed whenever the aged period was manifesting the excellent stability against air (see more discussion in text S2). Similar to the robust NZSP with the advantage of commendable air survivability (the exchange of protons with the surfaces of NZSP is tough, which prevents air-exposed NZSP powders from deteriorating) (43, 44), the derived LZSP electrolyte inherits its excellent stability against moisture and CO2 and requires no more extra seal measures if used in practical storage and transportation. To further reveal electrochemical stability, first-principles thermodynamic calculations were used to identify the thermal phase equilibria at different potentials for LZSP. According to our calculated reaction enthalpies, the LZSP is identified to have an electrochemical stability window of ~4 V (fig. S26), which is the top value among all the oxide solid electrolytes. This wide electrochemical window may be attributed to a high endothermicity (4.58 eV) of charge reaction formation of SiO2 + P2O5 + ZrO2 + O2 and a low exothermicity (0.48 eV) of discharge reaction formation of Li7Si3 + Li3P + Zr3O + Li2O.

Electrochemical test of solid-state batteries

To better eliminate the grain boundary impedance among LZSP nanoparticles, a thermal-assisted cold sintering approach was used to prepare dense LZSP composite pellets (45, 46). Furthermore, to negate the high interfacial resistance between the electrolyte and the electrode, soft polyethylene oxide–ethylene carbonate (PEO-EC) buffer layers were coated on the two surfaces of the LZSP composite pellets in the solid-state cells through an in situ solidification method (see more discussion in text S3 and figs. S27 to S30). A Li/LZSP/Li symmetric cell was fabricated to demonstrate the Li-ion transport capability of the LZSP solid electrolyte. As presented in Fig. 6A, lithium plating/stripping in the symmetrical Li/LZSP/Li cell remains stable for 2000 hours at room temperature with a current density of 0.1 mA cm−2, indicating an excellent stability against lithium metal. In Fig. 6B, measured by an Au/LZSP/Li cell, the electrochemical window of the solid electrolyte is stable up to 6 V versus Li/Li+, which is enough for the requirement of lithium batteries (all the subsistent batteries can withstand a charging voltage of up to 5 V) (47).

Fig. 6. The electrochemical performance of battery devices based on the LZSP solid electrolyte.

Fig. 6.

(A) Cycling performance of a Li/LZSP/Li symmetric cell tested at a current density of 0.1 mA cm−2. (B) Current-voltage curve of an Au/LZSP/Li cell. (C) Initial galvanostatic charge and discharge profile of a LiNi0.5Co0.2Mn0.3O2(NCM523)/LZSP/Li cell. (D) Cycling performance of a NCM523/LZSP/Li cell. (E) Rate performance of a NCM523/LZSP/Li cell.

High ionic conductivity, good stability against lithium metal and air, and wide stability window make LZSP one of the best candidate electrolytes [e.g., Li7La3Zr2O12 (LLZO) (10), Li10GeP2S12 (6, 48), and Li3M(M = Y, Sc, In)Cl6 (4953)] for developing commercial solid-state lithium batteries (table S8). A full cell containing the LZSP electrolyte pellet, the buffer layers, the NCM523 cathode, and the anode of Li metal was then designed to make a demonstration. Figure 6C shows the initial charge and discharge curves of the NCM523/LZSP/Li cell at 0.1 C (0.1 C = 15 mA g−1) at room temperature. The discharge specific capacity and coulombic efficiency of the first cycle are 162.0 mAh g−1 and 85.8%, respectively. After 100 cycles, there is a minor capacity fading for the cell and it delivered a specific capacity of 146.0 mAh g−1, equal to 90.1% of the initial discharge capacity (Fig. 6D), which surpasses the relevant performance of NCM523-based (cathode) and Li metal–based (anode) batteries using liquid electrolytes (fig. S31). Besides, rate performance was also tested at the current rate of 0.2, 0.5, 1.0, and 2.0 C, respectively, and high specific discharging capacities of 157.5, 141.8, 116.2, and 96.9 mAh g−1 were maintained (Fig. 6E). Compared to previously reported NCM523/Li cells based on other solid electrolytes (table S9), without any liquid wetting, the rate capacity, cycling performance, and discharge-specific capacity of this NCM523/LZSP/Li solid-state batteries are in the highest tier.

The synthesis approach of skeleton-retained cationic exchange could be commonly used to produce more solid electrolyte, even extensive inorganic materials, by controlling suitable reaction conditions. Several reactions such as Na3La(PO4)2 (NLP) → Li3La(PO4)2 (LLP) and Li1.3Al0.3Ti1.7(PO4)3 (LATP) → Na1.3Al0.3Ti1.7(PO4)3 (NATP) were attempted and exhibit effective cationic exchange. As shown in fig. S32, after the cationic exchange, the products of LLP and NATP respectively kept the skeleton structures of NLP and LATP completely. Together with detailed LZSP characterizations, it is expected that the approach could be used to synthesize new inorganic compounds with cationic composition and structure change. Its success largely depends on the rational selection of the solid electrolyte precursor, liquid carrier, and lithium compound.

DISCUSSION

In conclusion, skeleton-retained cationic exchange approach is used to produce a high-performance solid electrolyte with under-coordinated conduction Li+ ions and expanded bottleneck sizes. Using the NASICON-type solid electrolyte NZSP as a precursor, an unreported solid electrolyte, LZSP, with under-coordinated Li+ ions is harvested. The imported lithium cations acclimatize themselves to the relatively large skeleton by the reduction of Li coordination, resulting in low migration barriers for lithium cations to hop along 3D continuous migration channels. As a result, the bulk ionic conductivity of LZSP is up to 3.59 × 10−3 S cm−1 at room temperature, which is the top value among the current oxide solid electrolytes. In addition, the LZSP material not only shows satisfactory environmental survivability but also performs wide electrochemical window and long stability against Li metal in the use of LZSP-based solid-state batteries, indicating it as an advanced solid electrolyte with commercial promise. The present study indicates that the synthesis approach of skeleton-retained cationic exchange not only is effective in regulating the coordination structure of diffusion ions for improving ionic conductivity but also could be used to make new inorganic materials.

MATERIALS AND METHODS

Preparation of NZSP

The synthesis of NZSP powders is referenced from Zhang et al. (27). Three milliliters of tetraethyl orthosilicate, 10 ml of water, and 5 ml of ethanol were mixed and then stirred at room temperature for 10 min. After adding in 50 ml of water and 17 g of citric acid, the pH value of this mixed solution was controlled at 1.5 with HNO3 dropwise. Stirring at 333 K for 1 hour, a solution consisting of 5.132 g of ZrO(NO3)2·xH2O, 2.037 g of NaNO3, and 1.055 g of (NH4)2HPO4 was added while stirring at 353 K. The resulting solution was then transferred into an oil bath pan to evaporate residual water at 353 K for 24 hours. The obtained yellow dried gel was ball-milled at 600 rpm and then calcined at 1073 K in air. After grinding through an agate mortar, white NZSP powders were harvested.

Process of skeleton-retained cation exchange

In a typical skeleton-retained cation exchange procedure, 22.996 g of LiTFSI was dissolved in 40 ml of EMIIM in an Ar-filled glove box. The mixture was stirred at 333 K for 30 min and then cooled down to room temperature. Afterward, 2.829 g of NZSP was mixed with the above solution. The resulting white suspension was then sealed in a 100-ml reactor and heated in an oven at 453 K for 24 hours with a swing speed of 50 rpm. Subsequently, the white sediment was centrifuged, washed with ethanol, and dried at 353 K for 6 hours. All these above steps should be performed three times to yield pure LZSP powders (without residual sodium cation). After calcination at 773 K for 5 hours to remove the possible residual EMIIM, the target LZSP material is harvested. The exchanged ionic liquid can be reused through renewable process of zeolite adsorption, which is favorable to reduce cost of scale-up production.

Process of 17O isotopic labeling

In the 17O isotopic labeling process, 200 mg of NZSP or LZSP was sealed in a glass tube and then annealed at 673 K under 1 × 10−4 torr for 3 hours to get rid of surface-adsorbed impurities. Cooling down to room temperature, 90% 17O-enriched O2 gas (50 mbar, Isotec Inc.) was introduced and the glass tube was sealed again to heat at 673 K for 24 hours. Last, the samples were evacuated at 1 × 10−4 torr for 30 min to remove residual 17O-enriched O2 gas.

Characterization

Synchrotron XRD measurements were performed at the Beamline BL14B1 of Shanghai Synchrotron Radiation Facility (SSRF) in China. A medium-energy (18 keV, k = 0.6887 Å) x-ray beam (200 μm by 200 μm) is used. The NPD experiments were carried out at room temperature using high-resolution neutron powder diffractometer at China Mianyang Research Reactor. The incident beam with a wavelength λ = 1.8846 Å was vertically focused by a Ge (511) monochromator at a fixed take-off angle of 120°. Diffraction patterns were collected from 10° to 150° with a step increment of 0.06°. 23Na, 17O, and 6Li magic angle spinning NMR (MAS NMR) measurements were recorded on a 9.4-T Bruker Avance III 400 spectrometer at Larmor frequencies of 105.8, 54.2, and 58.8 MHz, respectively. Dry powders to be measured were packed into 4.0-mm ZrO2 rotors in a N2-filled glove box. 23Na, 17O, and 6Li NMR shifts were referenced to H2O at 0.0 parts per million (ppm), NaCl at 7.21 ppm, and LiCl at 0 ppm, respectively. Thermal gravimetric analysis data were obtained with a NETZSCH STA 449C instrument. XPS were recorded on Thermo ESCALAB 250 X with Al Kα (1486.6 eV) as core excitation. Binding energies expressed in XPS were corrected by referencing C 1s as 284.8 eV.

Computational methods

First-principles calculations were used to calculate the structural and electronic properties as implemented in the Vienna Ab initio Simulation Package (54). The plane-wave projector-augmented wave method (55, 56) with an energy cutoff of 400 eV and Monkhorst-pack k-meshes (57) were used to investigate the structural and transport properties by using generalized gradient approximation with Perdew-Burke-Ernzerh exchange-correlation functional (58). All atomic coordinates and lattice parameters are fully relaxed until the force convergence of 0.01 eV Å−1 and the energy convergence of 10−4 eV atom−1 are achieved. AIMD calculations were carried out in the canonical ensemble (NVT) using a Nosé thermostat at 300, 600, 800, 1000, and 1200 K with a time step of 3 fs for 30 ps to calculate MSDs and at 600 K for an appropriate time to investigate the diffusion pathways of conductive ions (59). AIMD and climbing image nudged elastic band (CI-NEB) (60) calculations revealed the transport properties of Na+ and Li+ ions via a correlated migration mechanism.

Fabrication of LZSP pellets

A hot-pressing sintering process was applied to obtain pure LZSP pellets. The LZSP powders were placed into the mold for dry pressing and sintered at 700°C under a fixed pressure of 70 MPa for 1 hour. After natural cooling to room temperature, the LZSP electrolyte pellet with a diameter of 10 mm was obtained. To better eliminate the grain boundary impedance among LZSP nanoparticles in the pure LZSP pellets, a thermal-assisted cold sintering approach was used to prepare dense LZSP composite pellets (45, 46). First, succinonitrile (SN) and LiTFSI with a mass ratio of 85:15 were mixed uniformly. Second, the LZSP powders were added and ground for 30 min. Last, the composite powders [90 weight % (wt %) LZSP, 10 wt % SN, and LiTFSI] were pressed at 130°C for 1 hour, and 810-MPa uniaxial pressure was applied. After natural cooling to room temperature, the LZSP composite pellet with a diameter of 10 mm was obtained. All operations were carried out in a drying room with a dew point at −40°C.

Pure LZSP pellets prepared by the hot-pressing sintering process were used to measure the ionic conductivities. LZSP composite pellets prepared by the thermal-assisted cold sintering approach were used to measure other electrochemical properties.

Electrochemical tests

The ionic conductivities of the electrolyte pellets were calculated from Nyquist plots, which were performed on an electrochemical workstation (Bio-Logic, VSP-300) with an AC amplitude of 50 mV from 7 MHz to 0.1 Hz by sputtering Au blocking electrodes on both sides of the pellets. The bulk and total ionic conductivities were obtained by fitting results using the ZView software. The ionic conductivities were calculated through σ = L/(R ∙ S), where σ, L, R, and S denote the ionic conductivity, the thickness of the pellet, the resistance, and the effective area of the electrode, respectively. For the Li/LZSP/Li symmetric cells, to improve the interface contact between the electrolyte and electrode, a PEO-EC buffer layer precursor solution was prepared by dispersing PEO (Mw = 600,000, Sigma-Aldrich) and EC (99.0%, Adamas) in dimethyl carbonate and stirred for 30 min at 60°C. The PEO-EC buffer layer precursor solution was casted onto on one side of the LZSP composite pellet and naturally volatilized for 12 hours to remove the dimethyl carbonate solvent, and two lithium foils were pasted on the buffer layers. The Li/LZSP/Li symmetric cells were investigated using a LAND battery tester (CT2001A, Wuhan LAND Electronics) under a current density of 0.1 mA cm−2.

The Au/LZSP/Li coin cells were assembled with a lithium foil pasted on the buffer layer as the counter electrode and a gold film sprayed on the other side of the pellet as the working electrode. The electrochemical window measurement was scanned in a range of −0.5 to 6.0 V at 1 mV s−1 at room temperature using a CHI604E electrochemical workstation.

The charge-discharge properties were evaluated by NCM523/LZSP/Li cells with soft PEO-EC buffer layers adhered to both sides of the LZSP composite pellet. The NCM523 cathode was composed of 77 wt % NCM523, 10 wt % Super P, 10 wt % polyvinylidene difluoride, and 3 wt % LiTFSI. The NCM523/LZSP/Li cell was tested between 2.8 and 4.3 V at current densities of 15, 30, 75, 150, and 300 mA g−1.

Acknowledgments

We would like to thank Y. Xia and W. Feng at Fudan University, T. Liu at Tongji University, X. Zhang at Shanghai Institute of Applied Physics, and X. Xia at Nanjing University of Science and Technology for their invaluable discussions and help in this work.

Funding: This work was financially supported by the National Key Research and Development Program of China (no. 2018YFB0905400), National Natural Science Foundation of China (nos. 21973107, 21972066, 22133005, and 22103093), Shanghai Academic Research Leader Program (20XD1404100), Shanghai Scientific Research Program of China (no. 19511107700), Shanghai Rising-Star Program (no. 18QB1402600), Shanghai Sailing Program (18YF1427300), and Science and Technology Commission of Shanghai Municipality (18511109400 and 18520723000).

Author contributions: L.Z., Y. Wang, J.C., J.L., and W.T. conceived and developed the idea of this project. L.Z., J.C., and W.L. carried out the synthesis of NZSP and LZSP materials. L.Z. and J.C. collected the synchrotron XRD data and analyzed them with Rietveld refinement. Y.X., F.W., and J.C. collected the NPD data and analyzed them with Rietveld method. Y. Wang, T.W., and J.L. conducted the DFT calculations. J.C. and L.P. performed 17O isotope enrichment and collected and analyzed all the NMR spectra. L.Z., J.W., S.H., W.L., M.W., Y. Wu, Y.Z., and W.L. performed the electrochemical tests. L.Z., Y. Wang, J.C., L.P., J.L., L.C., and W.T. wrote the manuscript, and all authors discussed the experiments and final manuscript.

Competing interests: The authors declare that they have no competing interests.

Data and materials availability: All data needed to evaluate the conclusions in the paper are present in the paper and/or the Supplementary Materials.

Supplementary Materials

This PDF file includes:

Figs. S1 to S32

Tables S1 to S9

Texts S1 to S3

References

REFERENCES AND NOTES

  • 1.Manthiram A., Yu X., Wang S., Lithium battery chemistries enabled by solid-state electrolytes. Nat. Rev. Mater. 2, 16103 (2017). [Google Scholar]
  • 2.Gao Z., Sun H., Fu L., Ye F., Zhang Y., Luo W., Huang Y., Promises, challenges, and recent progress of inorganic solid-state electrolytes for all-solid-state lithium batteries. Adv. Mater. 30, 1705702 (2018). [DOI] [PubMed] [Google Scholar]
  • 3.Chen R., Li Q., Yu X., Chen L., Li H., Approaching practically accessible solid-state batteries: Stability issues related to solid electrolytes and interfaces. Chem. Rev. 120, 6820–6877 (2019). [DOI] [PubMed] [Google Scholar]
  • 4.Schwietert T. K., Arszelewska V. A., Wang C., Yu C., Vasileiadis A., de Klerk N. J. J., Hageman J., Hupfer T., Kerkamm I., Xu Y., van der Maas E., Kelder E. M., Ganapathy S., Wagemaker M., Clarifying the relationship between redox activity and electrochemical stability in solid electrolytes. Nat. Mater. 19, 428–435 (2020). [DOI] [PubMed] [Google Scholar]
  • 5.Han F., Zhu Y., He X., Mo Y., Wang C., Electrochemical stability of Li10GeP2S12 and Li7La3Zr2O12 solid electrolytes. Adv. Energy Mater. 6, 1501590 (2016). [Google Scholar]
  • 6.Kamaya N., Homma K., Yamakawa Y., Hirayama M., Kanno R., Yonemura M., Kamiyama T., Kato Y., Hama S., Kawamoto K., Mitsui A., A lithium superionic conductor. Nat. Mater. 10, 682–686 (2011). [DOI] [PubMed] [Google Scholar]
  • 7.Zhao Y., Daemen L. L., Superionic conductivity in lithium-rich anti-perovskites. J. Am. Chem. Soc. 134, 15042–15047 (2012). [DOI] [PubMed] [Google Scholar]
  • 8.Lü X., Howard J. W., Chen A., Zhu J., Li S., Wu G., Dowden P., Xu H., Zhao Y., Jia Q., Antiperovskite Li3OCl superionic conductor films for solid-state Li-ion batteries. Adv. Sci. 3, 1500359 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Wenzel S., Randau S., Leichtweiß T., Weber D. A., Sann J., Zeier W. G., Janek J., Direct observation of the interfacial instability of the fast ionic conductor Li10GeP2S12 at the lithium metal anode. Chem. Mater. 28, 2400–2407 (2016). [Google Scholar]
  • 10.Murugan R., Thangadurai V., Weppner W., Fast lithium ion conduction in garnet-type Li7La3Zr2O12. Angew. Chem. Int. Ed. 46, 7778–7781 (2007). [DOI] [PubMed] [Google Scholar]
  • 11.Suzuki N., Inaba T., Shiga T., Electrochemical properties of LiPON films made from a mixed powder target of Li3PO4 and Li2O. Thin Solid Films 520, 1821–1825 (2012). [Google Scholar]
  • 12.Li Y., Zhou W., Chen X., Lü X., Cui Z., Xin S., Xue L., Jia Q., Goodenough J. B., Mastering the interface for advanced all-solid-state lithium rechargeable batteries. Proc. Natl. Acad. Sci. U.S.A. 113, 13313–13317 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Kim K. J., Balaish M., Wadaguchi M., Kong L. P., Rupp J., Solid-state Li–metal batteries: Challenges and horizons of oxide and sulfide solid electrolytes and their interfaces. Adv. Energy Mater. 11, 2002689 (2020). [Google Scholar]
  • 14.Deng Y., Eames C., Fleutot B., David R., Chotard J., Suard E., Masquelier C., Islam M. S., Enhancing the lithium ion conductivity in lithium superionic conductor (LISICON) solid electrolytes through a mixed polyanion effect. ACS Appl. Mater. Interfaces 9, 7050–7058 (2017). [DOI] [PubMed] [Google Scholar]
  • 15.Hou M., Liang F., Chen K., Dai Y., Xue D., Challenges and perspectives of NASICON-type solid electrolytes for all-solid-state lithium batteries. Nanotechnology 31, 132003 (2020). [DOI] [PubMed] [Google Scholar]
  • 16.Jian Z., Hu Y.-S., Ji X., Chen W., NASICON-structured materials for energy storage. Adv. Mater. 29, 1601925 (2017). [DOI] [PubMed] [Google Scholar]
  • 17.Ladenstein L., Simic S., Kothleitner G., Rettenwander D., Wilkening H. M. R., Anomalies in bulk ion transport in the solid solutions of Li7La3M2O12 (M = Hf, Sn) and Li5La3Ta2O12. J. Phys. Chem. C 124, 16796–16805 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Mei A., Jiang Q.-H., Lin Y.-H., Nan C.-W., Lithium lanthanum titanium oxide solid-state electrolyte by spark plasma sintering. J. Alloys Compd. 486, 16796–16805 (2009). [Google Scholar]
  • 19.Wang C., Fu K., Kammampata S. P., McOwen D. W., Samson A. J., Zhang L., Hitz G. T., Nolan A. M., Wachsman E. D., Mo Y., Thangadurai V., Hu L., Garnet-type solid-state electrolytes: Materials, interfaces, and batteries. Chem. Rev. 120, 4257–4300 (2020). [DOI] [PubMed] [Google Scholar]
  • 20.Lacivita V., Westover A. S., Kercher A., Phillip N. D., Yang G., Veith G., Ceder G., Dudney N. J., Resolving the amorphous structure of lithium phosphorus oxynitride (Lipon). J. Am. Chem. Soc. 140, 11029–11038 (2018). [DOI] [PubMed] [Google Scholar]
  • 21.Wang Y., Richards W. D., Ong S. P., Miara L. J., Kim J. C., Mo Y., Ceder G., Design principles for solid-state lithium superionic conductors. Nat. Mater. 14, 1026–1031 (2015). [DOI] [PubMed] [Google Scholar]
  • 22.Bachman J. C., Muy S., Grimaud A., Chang H.-H., Pour N., Lux S. F., Paschos O., Maglia F., Lupart S., Lamp P., Giordano L., Shao-Horn Y., Inorganic solid-state electrolytes for lithium batteries: Mechanisms and properties governing ion conduction. Chem. Rev. 116, 140–162 (2016). [DOI] [PubMed] [Google Scholar]
  • 23.Gao J., Wu J., Han S., Zhang J., Zhu L., Wu Y., Zhao J., Tang W., A novel solid electrolyte formed by NASICON-type Li3Zr2Si2PO12 and poly(vinylidene fluoride) for solid state batteries. Funct. Mater. Lett. 14, 2140001 (2021). [Google Scholar]
  • 24.Belam W., Sol-gel chemistry synthesis and DTA-TGA, XRPD, SIC and 7Li, 31P, 29Si MAS-NMR studies on the Li-NASICON Li3Zr2-ySi2-4yP1+4yO12 (0⩽y⩽0.5) system. J. Alloys Compd. 551, 267–273 (2013). [Google Scholar]
  • 25.Shannon R. D., Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. 32, 751–767 (1976). [Google Scholar]
  • 26.Yang K., Chen L., Ma J., He Y.-B., Kang F., Progress and perspective of Li1+xAlxTi2-x(PO4)3 ceramic electrolyte in lithium batteries. InfoMat. 3, 1195–1217 (2021). [Google Scholar]
  • 27.Zhang Z., Zou Z., Kaup K., Xiao R., Shi S., Avdeev M., Hu Y.-S., Wang D., He B., Li H., Huang X., Nazar L. F., Chen L., Correlated migration invokes higher Na+-ion conductivity in NaSICON-type solid electrolytes. Adv. Energy Mater. 9, 1902373 (2019). [Google Scholar]
  • 28.Vera V. B., Igor L. S., Ion exchange conversion of solid electrolyte, potassium sodiostannate, into isomorphous metastable sodium stannate. Mendeleev Commun. 26, 246–247 (2016). [Google Scholar]
  • 29.Malavasi L., Fisher C. A. J., Islam M. S., Oxide-ion and proton conducting electrolyte materials for clean energy applications: Structural and mechanistic features. Chem. Soc. Rev. 39, 4370–4387 (2010). [DOI] [PubMed] [Google Scholar]
  • 30.Zhang Z., Shao Y., Lotsch B., Hu Y.-S., Li H., Janek J., Nazar L. F., Nan C.-W., Maier J., Armand M., Chen L., New horizons for inorganic solid state ion conductors. Energ. Environ. Sci. 11, 1945–1976 (2018). [Google Scholar]
  • 31.Chen J., Wu X.-P., Hope M. A., Qian K., Halat D. M., Liu T., Li Y., Shen L., Ke X., Wen Y., Du J.-H., Magusin P. C. M. M., Paul S., Ding W., Gong X.-Q., Grey C. P., Peng L., Polar surface structure of oxide nanocrystals revealed with solid-state NMR spectroscopy. Nat. Commun. 10, 5420 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Chen J., Hope M. A., Lin Z., Wang M., Liu T., Halat D. M., Wen Y., Chen T., Ke X., Magusin P. C. M. M., Ding W., Xia X., Wu X.-P., Gong X.-Q., Grey C. P., Peng L., Interactions of oxide surfaces with water revealed with solid-state NMR spectroscopy. J. Am. Chem. Soc. 142, 11173–11182 (2020). [DOI] [PubMed] [Google Scholar]
  • 33.Van Den Broek J., Afyon S., Rupp J. L. M., Interface-engineered all-solid-state Li-ion batteries based on garnet-type fast Li+ conductors. Adv. Energy Mater. 6, 1600736 (2016). [Google Scholar]
  • 34.Han X., Gong Y., Fu K. K., He X., Hitz G. T., Dai J., Pearse A., Liu B., Wang H., Rubloff G., Mo Y., Thangadurai V., Wachsman E. D., Hu L., Negating interfacial impedance in garnet-based solid-state Li metal batteries. Nat. Mater. 16, 572–579 (2017). [DOI] [PubMed] [Google Scholar]
  • 35.DeWees R., Wang H., Synthesis and properties of NaSICON-type LATP and LAGP solid electrolytes. ChemSusChem 12, 3713–3725 (2019). [DOI] [PubMed] [Google Scholar]
  • 36.MacNeil D. D., Lu Z., Chen Z., Dahn J. R., A comparison of the electrode/electrolyte reaction at elevated temperatures for various Li-ion battery cathodes. J. Power Sources 108, 8–14 (2002). [Google Scholar]
  • 37.Xiao W., Wang J., Fan L., Zhang J., Li X., Recent advances in Li1+xAlxTi2-x(PO4)3 solid-state electrolyte for safe lithium batteries. Energy Storage Mater. 19, 379–400 (2019). [Google Scholar]
  • 38.Zhao P., Wen Y., Cheng J., Cao G., Jin Z., Ming H., Xu Y., Zhu X., A novel method for preparation of high dense tetragonal Li7La3Zr2O12. J. Power Sources 344, 56–61 (2017). [Google Scholar]
  • 39.Zhao P., Xiang Y., Wen Y., Li M., Zhu X., Zhao S., Jin Z., Ming H., Cao G., Garnet-like Li7-xLa3Zr2-xNbxO12 (x = 0–0.7) solid state electrolytes enhanced by self-consolidation strategy. J. Eur. Ceram. Soc. 38, 5454–5462 (2018). [Google Scholar]
  • 40.Pfenninger R., Struzik M., Garbayo I., Stilp E., Rupp J. L. M., A low ride on processing temperature for fast lithium conduction in garnet solid-state battery films. Nat. Energy 4, 475–483 (2019). [Google Scholar]
  • 41.Hong H. Y.-P., Crystal structures and crystal chemistry in the system Na1+xZr2SixP3−xO12. Mater. Res. Bull. 11, 173–182 (1976). [Google Scholar]
  • 42.Sharafi A., Kazyak E., Davis A. L., Yu S., Thompson T., Siegel D. J., Dasgupta N. P., Sakamoto J., Surface chemistry mechanism of ultra-low interfacial resistance in the solid-state electrolyte Li7La3Zr2O12. Chem. Mater. 29, 7961–7968 (2017). [Google Scholar]
  • 43.Mauvy F., Siebert E., Fabry P., Reactivity of NASICON with water and interpretation of the detection limit of a NASICON based Na+ ion selective electrode. Talanta 48, 293–303 (1999). [DOI] [PubMed] [Google Scholar]
  • 44.Guin M., Indris S., Kaus M., Ehrenberg H., Tietz F., Guillon O., Stability of NASICON materials against water and CO2 uptake. Solid State Ion. 302, 102–106 (2017). [Google Scholar]
  • 45.Lee W., Lyon C. K., Seo J.-H., Hallman R. L., Leng Y., Wang C.-Y., Hickner M. A., Randall C. A., Gomez E. D., Ceramic-salt composite electrolytes from cold sintering. Adv. Funct. Mater. 29, 1807872 (2019). [Google Scholar]
  • 46.Liu Y. L., Liu J., Sun Q., Wang D., Adair K. R., Liang J., Zhang C., Zhang L., Lu S., Huang H., Song X., Sun X., Insight into the microstructure and ionic conductivity of cold sintered NASICON solid electrolyte for solid-state batteries. ACS Appl. Mater. Interfaces 11, 27890–27896 (2019). [DOI] [PubMed] [Google Scholar]
  • 47.Andre D., Kim S.-J., Lamp P., Lux S. F., Maglia F., Paschosa O., Stiaszny B., Future generations of cathode materials: An automotive industry perspective. J. Mater. Chem. A 3, 6709–6732 (2015). [Google Scholar]
  • 48.Chen S., Xie D., Liu G., Mwizerwa J. P., Zhang Q., Zhao Y., Xu X., Yao X., Sulfide solid electrolytes for all-solid-state lithium batteries: Structure, conductivity, stability and application. Energy Storage Mater. 14, 58–74 (2018). [Google Scholar]
  • 49.Asano T., Sakai A., Ouchi S., Sakaida M., Miyazaki A., Hasegawa S., Solid halide electrolytes with high lithium-ion conductivity for application in 4 V class bulk-type all-solid-state batteries. Adv. Mater. 30, 1803075 (2018). [DOI] [PubMed] [Google Scholar]
  • 50.Li X., Liang J., Chen N., Luo J., Adair K. R., Wang C., Banis M. N., Sham T.-K., Zhang L., Zhao S., Lu S., Huang H., Li R., Sun X., Water-mediated synthesis of a superionic halide solid electrolyte. Angew. Chem. Int. Ed. 131, 16579–16584 (2019). [DOI] [PubMed] [Google Scholar]
  • 51.Liang J., Li X., Wang S., Adair K. R., Li W., Zhao Y., Wang C., Hu Y., Zhang L., Zhao S., Lu S., Huang H., Li R., Mo Y., Sun X., Site-occupation-tuned superionic LixScCl3+x halide solid electrolytes for all-solid-state batteries. J. Am. Chem. Soc. 142, 7012–7022 (2020). [DOI] [PubMed] [Google Scholar]
  • 52.Zahiri B., Patra A., Kiggins C., Yong A. X. B., Ertekin E., Cook J. B., Braun P. V., Revealing the role of the cathode–electrolyte interface on solid-state batteries. Nat. Mater. 20, 1392–1400 (2021). [DOI] [PubMed] [Google Scholar]
  • 53.Li X., Liang J., Yang X., Adair K. R., Wang C., Zhao F., Sun X., Progress and perspectives on halide lithium conductors for all-solid-state lithium batteries. Energ. Environ. Sci. 13, 1429–1461 (2020). [Google Scholar]
  • 54.Kresse G., Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 54, 11169–11186 (1996). [DOI] [PubMed] [Google Scholar]
  • 55.Blochl P. E., Projector augmented-wave method. Phys. Rev. B 50, 17953–17979 (1994). [DOI] [PubMed] [Google Scholar]
  • 56.Kresse G., Joubert D., From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 59, 1758–1775 (1999). [Google Scholar]
  • 57.Monkhorst H. J., Pack J. D., Special points for brillouin-zone integrations. Phys. Rev. B 13, 5188–5192 (1976). [Google Scholar]
  • 58.Perdew J. P., Burke K., Ernzerhof M., Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865–3868 (1996). [DOI] [PubMed] [Google Scholar]
  • 59.Nosé S., A unified formulation of the constant temperature molecular dynamics methods. J. Chem. Phys. 81, 511–519 (1984). [Google Scholar]
  • 60.Udagawa T., Suzuki K., Tachikawa M., Multicomponent molecular orbital-climbing image-nudged elastic band method to analyze chemical reactions including nuclear quantum effect. ChemPhysChem 16, 3156–3160 (2015). [DOI] [PubMed] [Google Scholar]
  • 61.M. H. Levitt, Chapter 7: NMR of other commonly studied nuclei, in Spin Dynamics: Basics of Nuclear Magnetic Resonance (John Wiley & Sons, 2002). [Google Scholar]
  • 62.Chen J., Zhu L., Jia D., Jiang X., Wu Y., Hao Q., Xia X., Ouyang Y., Peng L., Tang W., Liu T., LiNi0.8Co0.15Al0.05O2 cathodes exhibiting improved capacity retention and thermal stability due to a lithium iron phosphate coating. Electrochim. Acta 312, 179–187 (2019). [Google Scholar]
  • 63.Du M., Liao K., Lua Q., Shao Z., Recent advances in the interface engineering of solid-state Li-ion batteries with artificial buffer layers: Challenges, materials, construction, and characterization. Energ. Environ. Sci. 12, 1780–1804 (2019). [Google Scholar]
  • 64.Zhang X., Liu T., Zhang S., Huang X., Xu B., Lin Y., Xu B., Li L., Nan C.-W., Shen Y., Synergistic coupling between Li6.75La3Zr1.75Ta0.25O12 and poly(vinylidene fluoride) induces high ionic conductivity, mechanical strength, and thermal stability of solid composite electrolytes. J. Am. Chem. Soc. 139, 13779–13785 (2017). [DOI] [PubMed] [Google Scholar]
  • 65.Nie K., Hong Y., Qiu J., Li Q., Yu X., Li H., Chen L., Interfaces between cathode and electrolyte in solid state lithium batteries: Challenges and perspectives. Front. Chem. 6, 616 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Ma X., Xu Y., Zhang B., Xue X., Wang C., He S., Lin J., Yang L., Garnet Si–Li7La3Zr2O12 electrolyte with a durable, low resistance interface layer for all-solid-state lithium metal batteries. J. Power Sources 453, 227881 (2020). [Google Scholar]
  • 67.Pitawala H. M. J. C., Dissanayake M. A. K. L., Seneviratne V. A., Combined effect of Al2O3 nano-fillers and EC plasticizer on ionic conductivity enhancement in the solid polymer electrolyte (PEO)9LiTf. Solid State Ion. 178, 885–888 (2007). [Google Scholar]
  • 68.Nicotera I., Ranieri G. A., Terenzi M., Chadwick A. V., Webster M. I., A study of stability of plasticized PEO electrolytes. Solid State Ion. 146, 143–150 (2002). [Google Scholar]
  • 69.Ratner M. A., Shriver D. F., Ion transport in solvent-Free polymers. Chem. Rev. 88, 109–124 (1988). [Google Scholar]
  • 70.Liu Y., Li C., Li B., Song H., Cheng Z., Chen M., He P., Zhou H., Germanium thin film protected lithium aluminum germanium phosphate for solid-state Li batteries. Adv. Energy Mater. 8, 1702374 (2018). [Google Scholar]
  • 71.Aono H., Sugimoto E., Sadaoka Y., Imanaka N., Adachi G., Ionic Conductivity of solid electrolytes based on lithium titanium phosphate. J. Electrochem. Soc. 137, 1023–1027 (1990). [Google Scholar]
  • 72.Nairn K., Forsyth M., Every H., Greville M., MacFarlane D. R., Polymer-ceramic ion-conducting composites. Solid State Ion. 86-88, 589–593 (1996). [Google Scholar]
  • 73.Birke P., Salam F., Doring S., Weppner W., A first approach to a monolithic all solid state inorganic lithium battery. Solid State Ion. 118, 149–157 (1999). [Google Scholar]
  • 74.Yu J., Duan S., Huang B., Jin H., Xie S., Li J., Spatially resolved electrochemical strain of solid-state electrolytes via high resolution sequential excitation and its implication on grain boundary impedance. Small Methods 4, 2000308 (2020). [Google Scholar]
  • 75.Li S.-C., Cai J.-Y., Lin Z.-X., Phase relationships and electrical conductivity of Li1+xGe2-xAlxP3O12 and Li1+xGe2-xGrxP3O12 systems. Solid State Ion. 28–30, 1265–1270 (1988). [Google Scholar]
  • 76.Aono H., Sugimoto E., Sadaoka Y., Imanaka N., Adachi G., Electrical properties and sinterability for lithium germanium phosphate Li1+xMxGe2-x(PO4)3, M=Al, Cr, Ga, Fe, Sc, and in systems. Bull. Chem. Soc. Jpn. 65, 2200–2204 (1992). [Google Scholar]
  • 77.Fu J., Fast Li ion conducting glass-ceramics in the system Li2O-Al2O3-GeO2-P2O5. Solid State Ion. 104, 191–194 (1997). [Google Scholar]
  • 78.Xu X., Wen Z., Wu X., Yang X., Gu Z., Lithium ion-conducting glass-ceramics of Li1.5Al0.5Ge1.5(PO4)3-xLi2O (x = 0.0–0.20) with good electrical and electrochemical properties. J. Am. Ceram. Soc. 90, 2802–2806 (2007). [Google Scholar]
  • 79.Nikodimos Y., Tsai M.-C., Abrha L. H., Weldeyohannis H. H., Chiu S.-F., Bezabh H. K., Shitaw K. N., Fenta F. W., Wu S.-H., Su W.-N., Yang C.-C., Hwang B. J., Al-Sc dual-doped LiGe2(PO4)3—A NASICON-type solid electrolyte with improved ionic conductivity. J. Mater. Chem. A 8, 11302–11313 (2020). [Google Scholar]
  • 80.Chang C.-M., Lee Y. I., Hong S.-H., Park H.-M., Spark plasma sintering of LiTi2(PO4)3-based solid electrolytes. J. Am. Ceram. Soc. 88, 1803–1807 (2005). [Google Scholar]
  • 81.Luo Z., Qin C., Xu W., Liang H., Lei W., Shen X., Lu A., Crystal structure refinement, microstructure and ionic conductivity of ATi2(PO4)3 (A=Li, Na, K) solid electrolytes. Ceram. Int. 46, 15613–15620 (2020). [Google Scholar]
  • 82.Yamamoto H., Tabuchi M., Takeuchi T., Kageyama H., Nakamura O., Ionic conductivity enhancement in LiGe2(PO4)3 solid electrolyte. J. Power Sources 68, 397–401 (1997). [Google Scholar]
  • 83.Feng J., Xia H., Lai M. O., Lu L., NASICON-structured LiGe2(PO4)3 with improved cyclability for high-performance lithium batteries. J. Phys. Chem. C 113, 20514–20520 (2009). [Google Scholar]
  • 84.Hu Y.-W., Raistrick I. D., Huggins R. A., Ionic conductivity of lithium phosphate-doped lithium orthosilicate. Mater. Res. Bull. 11, 1227–1230 (1976). [Google Scholar]
  • 85.Khorassani A., Izquierdo G., West A. R., The solid electrolyte system Li3PO4-Li4SiO4. Mater. Res. Bull. 16, 1561–1567 (1981). [Google Scholar]
  • 86.Wang D., Zhong G., Lia Y., Gong Z., McDonald M. J., Mi J.-X., Fu R., Shi Z., Yang Y., Enhanced ionic conductivity of Li3.5Si0.5P0.5O4 with addition of lithium borate. Solid State Ion. 283, 109–114 (2015). [Google Scholar]
  • 87.Sudo R., Nakata Y., Ishiguro K., Matsui M., Hirano A., Takeda Y., Yamamoto O., Imanishi N., Interface behavior between garnet-type lithium-conducting solid electrolyte and lithium metal. Solid State Ion. 262, 151–154 (2014). [Google Scholar]
  • 88.Ramakumar S., Janani N., Murugan R., Influence of lithium concentration on the structure and Li+ transport properties of cubic phase lithium garnets. Dalton Trans. 44, 539–552 (2015). [DOI] [PubMed] [Google Scholar]
  • 89.Zhang Y., Luo D., Luo W., Du S., Deng Y., Deng J., High-purity and high-density cubic phase of Li7La3Zr2O12 solid electrolytes by controlling surface/volume ratio and sintering pressure. Electrochim. Acta 359, 136965 (2020). [Google Scholar]
  • 90.Thangadurai V., Kaack H., Weppner W. J. F., Novel fast lithium ion conduction in garnet-type Li5La3M2O12 (M = Nb, Ta). J. Am. Ceram. Soc. 86, 437–440 (2003). [Google Scholar]
  • 91.Nemori H., Matsuda Y., Matsui M., Yamamoto O., Takeda Y., Imanishi N., Relationship between lithium content and ionic conductivity in the Li5+2xLa3Nb2-xScxO12 system. Solid State Ion. 266, 9–12 (2014). [Google Scholar]
  • 92.Wang W. G., Wang X. P., Gao Y. X., Fang Q. F., Lithium-ionic diffusion and electrical conduction in the Li7La3Ta2O13 compounds. Solid State Ion. 180, 1252–1256 (2009). [Google Scholar]
  • 93.Wang W. G., Wang X. P., Gao Y. X., Hao G. L., Ma W. Q., Fang Q. F., Internal friction study on the lithium ion diffusion of Li5La3M2O12 (M = Ta, Nb) ionic conductors. Solid State Ion. 13, 1760–1764 (2011). [Google Scholar]
  • 94.Wang Y., Lai W., High ionic conductivity lithium garnet oxides of Li7-xLa3Zr2-xTaxO12 compositions. Electrochem. Solid State 15, A68–A71 (2012). [Google Scholar]
  • 95.Kotobuki M., Kanamura K., Fabrication of all-solid-state battery using Li5La3Ta2O12 ceramic electrolyte. Ceram. Int. 39, 6481–6487 (2013). [Google Scholar]
  • 96.Inaguma Y., Chen L., Itoh M., Nakamura T., Uchida T., Ikuta H., Wakihara M., High ionic conductivity in lithium lanthanum titanate. Solid State Commun. 86, 689–693 (1993). [Google Scholar]
  • 97.Tang H., Xu J., Enhanced electrochemical performance of LiFePO4 coated with Li0.34La0.51TiO2.94 by rheological phase reaction method. Mater. Sci. Eng. B 178, 1503–1508 (2013). [Google Scholar]
  • 98.Le H. T. T., Kalubarme R. S., Ngo D. T., Jang S.-Y., Jung K.-N., Shin K.-H., Park C.-J., Citrate gel synthesis of aluminum-doped lithium lanthanum titanate solid electrolyte for application in organic-type lithium-oxygen batteries. J. Power Sources 274, 1188–1199 (2015). [Google Scholar]
  • 99.Le H. T. T., Ngo D. T., Kim Y.-J., Park C.-N., Park C.-J., A perovskite-structured aluminium-substituted lithium lanthanum titanate as a potential artificial solid-electrolyte interface for aqueous rechargeable lithium-metal-based batteries. Electrochim. Acta 248, 232–242 (2017). [Google Scholar]
  • 100.Itoh M., Inaguma Y., Jung W.-H., Chen L., Nakamura T., High lithium ion conductivity in the perovskite-type compounds Ln1/2Li1/2TiO3 (Ln = La, Pr, Nd, Sm). Solid State Ion. 70-71, 203–207 (1994). [Google Scholar]
  • 101.Stramare S., Thangadurai V., Weppner W., Lithium lanthanum titanates: A review. Chem. Mater. 15, 3974–3990 (2003). [Google Scholar]
  • 102.Zhou Q., Xu B., Chien P.-H., Li Y., Huang B., Wu N., Xu H., Grundish N. S., Hu Y.-Y., Goodenough J. B., NASICON Li1.2Mg0.1Zr1.9(PO4)3 solid electrolyte for an all-solid-state Li-metal battery. Small Methods 4, 2000764 (2020). [Google Scholar]
  • 103.Kosova N. V., Devyatkina E. T., Stepanov A. P., Buzlukov A. L., Lithium conductivity and lithium diffusion in NASICON-type Li1+xTi2–xAlx(PO4)3 (x = 0;0.3) prepared by mechanical activation. Ionics 14, 303–311 (2008). [Google Scholar]
  • 104.Zhu Y. Z., He X. F., Mo Y. F., Origin of outstanding stability in the lithium solid electrolyte materials: Insights from thermodynamic analyses based on first-principles calculations. ACS Appl. Mater. Interfaces 7, 23685–23693 (2015). [DOI] [PubMed] [Google Scholar]
  • 105.Wu X. M., Li X. H., Zhang Y. H., Xu M. F., He Z. Q., Synthesis of Li1.3Al0.3Ti1.7(PO4)3 by sol–gel technique. Mater. Lett. 58, 1227–1230 (2004). [Google Scholar]
  • 106.Kotobuki M., Koishi M., Preparation of Li1.5Al0.5Ti1.5(PO4)3 solid electrolyte via a sol-gel route using various Al sources. Ceram. Int. 39, 4645–4649 (2013). [Google Scholar]
  • 107.Breuer S., Prutsch D., Ma Q., Epp V., Preishuber-Pflugl F., Tietz F., Wilkening M., Separating bulk from grain boundary Li ion conductivity in the sol-gel prepared solid electrolyte Li1.5Al0.5Ti1.5(PO4)3. J. Mater. Chem. A 3, 21343–21350 (2015). [Google Scholar]
  • 108.Feng J. K., Lu L., Lai M. O., Lithium storage capability of lithium ion conductor Li1.5Al0.5Ge1.5(PO4)3. J. Alloys Compd. 501, 255–258 (2010). [Google Scholar]
  • 109.Illbeigi M., Fazlali A., Kazazi M., Mohammadi A. H., Effect of simultaneous addition of aluminum and chromium on the lithium ionic conductivity of LiGe2(PO4)3 NASICON-type glass-ceramics. Solid State Ion. 289, 180–187 (2016). [Google Scholar]
  • 110.Thompson T., Yu S., Williams L., Schmidt R. D., Garcia-Mendez R., Wolfenstine J., Allen J. L., Kioupakis E., Siegel D. J., Sakamoto J., Electrochemical window of the Li-ion solid electrolyte Li7La3Zr2O12. ACS Energy Lett. 2, 462–468 (2017). [Google Scholar]
  • 111.Ohta S., Kobayashi T., Asaoka T., High lithium ionic conductivity in the garnet-type oxide Li7−xLa3(Zr2−x, Nbx)O12 (x = 0–2). J. Power Sources 196, 3342–3345 (2011). [Google Scholar]
  • 112.Wei J., Ogawa D., Fukumura T., Hirose Y., Hasegawa T., Epitaxial strain-controlled ionic conductivity in Li-ion solid electrolyte Li0.33La0.56TiO3 thin films. Cryst. Growth Des. 15, 2187–2191 (2015). [Google Scholar]
  • 113.Chen C. H., Amine K., Ionic conductivity, lithium insertion and extraction of lanthanum lithium titanate. Solid State Ion. 144, 51–57 (2001). [Google Scholar]
  • 114.Yamaki J., Ohtsuka H., Shodai T., Rechargeable lithium thin film cells with inorganic electrolytes. Solid State Ion. 86-88, 1279–1284 (1996). [Google Scholar]
  • 115.Bruce P. G., West A. R., Ionic conductivity of LISICON solid solutions, Li2+2xZn1-xGeO4. J. Solid State Chem. 44, 354–365 (1982). [Google Scholar]
  • 116.Gao J., Zhao Y. S., Shi S. Q., Li H., Lithium-ion transport in inorganic solid state electrolyte. Chin. Phys. B 25, 018211 (2016). [Google Scholar]
  • 117.Li W., Wu G., Araújo C. M., Scheicher R. H., Blomqvist A., Ahuja R., Xiong Z., Feng Y., Chen P., Li+ ion conductivity and diffusion mechanism in α-Li3N and β-Li3N. Energ. Environ. Sci. 3, 1524–1530 (2010). [Google Scholar]
  • 118.Rabenau A., Lithium nitride and related materials case study of the use of modern solid state research techniques. Solid State Ion. 6, 277–293 (1982). [Google Scholar]
  • 119.Rea J. R., Foster D. L., Mallory P. R., Inc C., High ionic conductivity in densified polycrystalline lithium nitride. Mater. Res. Bull. 14, 841–846 (1979). [Google Scholar]
  • 120.Xu R., Wu Z., Zhang S., Wang X., Xia Y., Xia X., Huang X., Tu J., Construction of all-solid-state batteries based on a sulfur-graphene composite and Li9.54Si1.74P1.44S11.7Cl0.3 solid electrolyte. Chem. Eur. J. 23, 13950–13956 (2017). [DOI] [PubMed] [Google Scholar]
  • 121.Seino Y., Ota T., Takada K., Hayashi A., Tatsumisago M., A sulphide lithium super ion conductor is superior to liquid ion conductors for use in rechargeable batteries. Energ. Environ. Sci. 7, 627–631 (2014). [Google Scholar]
  • 122.Boulineau S., Courty M., Tarascon J. M., Viallet V., Mechanochemical synthesis of li-argyrodite Li6PS5X (X=Cl, Br, I) as sulfur-based solid electrolytes for all solid state batteries application. Solid State Ion. 221, 1–5 (2012). [Google Scholar]
  • 123.Rangasamy E., Liu Z., Gobet M., Pilar K., Sahu G., Zhou W., Wu H., Greenbaum S., Liang C., An iodide-based Li7P2S8I superionic conductor. J. Am. Chem. Soc. 137, 1384–1387 (2015). [DOI] [PubMed] [Google Scholar]
  • 124.Zhou D., Shanmukaraj D., Tkacheva A., Armand M., Wang G. X., Polymer electrolytes for lithium-based batteries: Advances and Prospects. Chem 5, 2326–2352 (2019). [Google Scholar]
  • 125.Zhang J., Zhao J., Yue L., Wang Q., Chai J., Liu Z., Zhou X., Li H., Guo Y., Cui G., Chen L., Safety-reinforced poly(propylene carbonate)-based all-solid-state polymer electrolyte for ambient-temperature solid polymer lithium batteries. Adv. Energy Mater. 5, 1501082 (2015). [Google Scholar]
  • 126.Chen K., Shen Y., Jiang J., Zhang Y., Lin Y., Nan C.-W., High capacity and rate performance of LiNi0.5Co0.2Mn0.3O2 composite cathode for bulk-type all-solid-state lithium battery. J. Mater. Chem. A 2, 13332–13337 (2014). [Google Scholar]
  • 127.Shao Y., Wang H., Gong Z., Wang D., Zheng B., Zhu J., Lu Y., Hu Y.-S., Guo X., Li H., Huang X., Yang Y., Nan C.-W., Chen L., Drawing a soft interface: An effective interfacial modification strategy for garnet-type solid-state Li batteries. ACS Energy Lett. 3, 1212–1218 (2018). [Google Scholar]
  • 128.Liu B., Zhang L., Xu S., McOwen D. W., Gong Y., Yang C., Pastel G. R., Xie H., Fu K., Dai J., Chen C., Wachsman E. D., Hu L., 3D lithium metal anodes hosted in asymmetric garnet frameworks toward high energy density batteries. Energy Storage Mater. 14, 376–382 (2018). [Google Scholar]
  • 129.Liu T., Zhang Y., Zhang X., Wang L., Zhao S.-X., Lin Y.-H., Shen Y., Luo J., Li L., Nan C.-W., Enhanced electrochemical performance of bulk type oxide ceramic lithium batteries enabled by interface modification. J. Mater. Chem. A 6, 4649–4657 (2018). [Google Scholar]
  • 130.Chen L., Li W. X., Fan L. Z., Nan C.-W., Zhang Q., Intercalated electrolyte with high transference number for dendrite-free solid-state lithium batteries. Adv. Funct. Mater. 29, 1901047 (2019). [Google Scholar]
  • 131.Jiang T., He P., Wang G., Shen Y., Nan C.-W., Fan L.-Z., Solvent-free synthesis of thin, flexible, nonflammable garnet-based composite solid electrolyte for all-solid-state lithium batteries. Adv. Energy Mater. 10, 1903376 (2020). [Google Scholar]
  • 132.Zhang B., Chen L., Hu J., Liu Y., Liu Y., Feng Q., Zhu G., Fan L.-Z., Solid-state lithium metal batteries enabled with high loading composite cathode materials and ceramic-based composite electrolytes. J. Power Sources 442, 227230 (2019). [Google Scholar]
  • 133.Huang S., Yang H., Hu J., Liu Y., Wang K., Peng H., Zhang H., Fan L.-Z., Early lithium plating behavior in confined nanospace of 3D lithiophilic carbon matrix for stable solid-state lithium metal batteries. Small 15, 1904216 (2019). [DOI] [PubMed] [Google Scholar]
  • 134.Ruan Y., Lu Y., Huang X., Su J., Sun C., Jin J., Wen Z., Acid induced conversion towards a robust and lithiophilic interface for Li-Li7La3Zr2O12 solid-state batteries. J. Mater. Chem. A 7, 14565–14574 (2019). [Google Scholar]
  • 135.Wang G. X., He P. G., Fan L. Z., Asymmetric polymer electrolyte constructed by metal-organic framework for solid-state, dendrite-free lithium metal battery. Adv. Funct. Mater. 31, 2007198 (2021). [Google Scholar]
  • 136.Zhuang Z., Yang L., Ju B., Lei G., Zhou Q., Liao H., Yin A., Deng Z., Tang Y., Qin S., Tu F., Ameliorating interfacial issues of LiNi0.5Co0.2Mn0.3O2/poly (propylene carbonate) by introducing graphene interlayer for all-solid-state lithium batteries. ChemistrySelect 5, 2291–2299 (2020). [Google Scholar]
  • 137.Zhao L., Fu J., Du Z., Jia X., Qu Y., Yu F., Du J., Chen Y., High-strength and flexible cellulose/PEG based gel polymer electrolyte with high performance for lithium ion batteries. J. Membr. Sci. 593, 117428 (2020). [Google Scholar]
  • 138.Zha W. P., Li W. W., Ruan Y. D., Wang J. C., Wen Z. Y., In situ fabricated ceramic/polymer hybrid electrolyte with vertically aligned structure for solid-state lithium batteries. Energ. Storage Mater. 36, 171–178 (2021). [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Figs. S1 to S32

Tables S1 to S9

Texts S1 to S3

References


Articles from Science Advances are provided here courtesy of American Association for the Advancement of Science

RESOURCES