Masato Torii,
Atsushi Sakuda*,
Kota Motohashi
and
Akitoshi Hayashi
Department of Applied Chemistry, Graduate School of Engineering, Osaka Metropolitan University, 1-1 Gakuen-cho, Naka-ku, Sakai, Osaka 599-8531, Japan. E-mail: saku@omu.ac.jp; Fax: +81-72-2549910; Tel: +81-72-2549331
First published on 6th August 2025
All-solid-state batteries have emerged as alternative rechargeable batteries offering high energy density and enhanced safety. However, suppressing their mechanical degradation is challenging. In particular, inorganic solid electrolytes must form mechanically stable solid–solid interfaces with electrode active materials, making the examination of their elastic properties essential for creating robust interfaces. Pugh's ratio (B/G) serves as a key parameter for estimating ductility, with a desirable value exceeding 1.75—a criterion originally proposed for polycrystalline metals. In this study, the elastic properties of Li-ion-conducting crystalline electrolytes were comprehensively evaluated via first-principles calculations. The calculated mechanical properties of their crystal structures were classified based on their anion elements. The elastic moduli of sulfide and halide crystals were relatively lower than those of oxide and nitride materials. The Pugh ratios of sulfide crystals were generally higher than 1.75, while those of oxide crystals clustered around 1.75 and nitride crystals typically fell below this threshold. Additionally, a nonlinear correlation between mean atomic volume and elastic constants was observed. Among the various electrolytes, Li2SO4 exhibited exceptional elastic properties: α-Li2SO4 demonstrated a significantly high B/G value of 4.28, indicating distinctive ductility.
Solid electrolytes are classified primarily based on their anion elements and crystal unit cell structures. Sulfide solid electrolytes are notable for their high moldability and ionic conductivity.5 They exhibit superior formability and can be densified solely by pressing under ambient temperature.6 Among sulfide solid electrolytes, Li-ion conductors such as Li10GeP2S12 (LGPS) type electrolytes demonstrate the highest ionic conductivity, exceeding 10−2 S cm−1.2,7,8 The argyrodite Li6PS5Cl crystal family is another typical sulfide solid electrolyte, with ionic conductivity reaching 10−2 S cm−1.9,10
Oxide solid electrolytes are also widely used due to their electrochemical stability. Garnet-type Li7La3Zr2O12 (LLZO) exhibits both high ionic conductivity (∼1 mS cm−1) and stability against lithium Li metal.11,12 Some nitride solid electrolytes also possess high ionic conductivity; for instance, Li3N electrolytes exceed 10−3 S cm−1.13 Recently, halide solid electrolytes have garnered attention because of their high ionic conductivity and oxidation resistance. Asano et al. first reported the novel Li3YCl6 chloride electrolyte, which exhibited ionic conductivity exceeding 10−3 S cm−1.14 Following this discovery, a variety of Li3MCl6-type chloride electrolytes have been investigated.15
From a mechanical perspective, solid electrolytes must possess adequate deformability and ductility to accommodate the expansion and contraction of electrode active materials, thereby minimizing stress accumulation at the interface. The elastic modulus is a fundamental property that quantifies the resistance of a material against deformation under microscopic elastic strain. During elastic deformation, stress and strain are proportional, with the elastic modulus serving as the proportionality constant. In general, a lower elastic modulus corresponds to easier deformation, making solid electrolytes with low elastic moduli preferable for such applications.
Various proportional constants are defined based on the plane and direction of stress and strain, and this information is typically organized in a four-dimensional elastic tensor. For ease of interpretation or computation, this tensor is often reduced to a two-dimensional form. The Voigt–Reuss–Hill approximation is commonly employed to predict elastic moduli from the tensor components.16 The primary elastic moduli used for evaluating crystalline materials include the polycrystalline Young's modulus (E), bulk modulus (B), and shear modulus (G). Young's modulus represents the proportionality constant between stress and strain in the uniaxial direction, while the polycrystalline Young's modulus (E) provides an averaged value for polycrystalline materials. The bulk modulus (B) quantifies the relationship between isotropic stress and volumetric strain, whereas the shear modulus (G) is related to angular displacement under equilibrium shear force. Additionally, Poisson's ratio (ν), defined as the ratio of transverse strain to axial strain in response to uniaxial stress, is often used for mechanical characterization.
Pugh's ratio, the shear-to-bulk modulus ratio (B/G), is a key parameter for evaluating the mechanical ductility of solid electrolytes.17,18 Originally proposed by Pugh in the context of polycrystalline metals, this criterion assesses a material's ability to undergo plastic deformation.17 Plastic deformation occurs when a material is strained beyond its elastic limit, and its likelihood depends on the material's elastic properties. Pugh suggests that polycrystalline metal materials, with a B/G ratio exceeding 1.75, are more likely to exhibit plastic deformation. However, this standard, derived for metals, has not been validated for ceramic materials, such as those used in inorganic solid electrolytes.
Deng et al. previously conducted a comprehensive investigation of the elastic moduli of typical solid electrolytes,19 viz. alkali superionic conductors, using first-principles calculations, uncovering trends influenced by anion species, structural frameworks, and alkali ions. Nevertheless, they did not include nitrides or the more recently explored chloride solid electrolytes.14,15 Furthermore, additional investigations on a wider range of electrolytes, beyond those covered in their study, are required.
In this study, we aimed to provide more comprehensive insights into the elastic properties of Li-ion conductors by examining a broader selection of electrolyte materials. The calculated elastic moduli are systematically categorized based on the anion types and crystal structures of the solid electrolytes. Using these findings, alongside an analysis of Pugh's ratio (B/G), the study identifies solid electrolytes with high deformability and ductility, which are essential for accommodating the expansion and contraction of electrodes.
The Monkhorst–Pack scheme28 was utilized to determine k-point distributions and the irreducible Brillouin zone. As an exception, a Γ-centered 1 × 1 × 1 k-point mesh was adopted for LLZO due to computational cost considerations. Lattice constants and ionic positions of ordinary crystal structures were fully relaxed, ensuring that the final forces on all relaxed atoms were less than 0.01 eV Å−1. On-site Coulomb term (U) values of 2.50, 3.50, 4.00, and 2.30 eV were applied for the Ti-3d, Zr-4d, Nb-4d, and La-5f orbitals, respectively, following prior studies.29–32 We incorporated the effect of van der Waals interactions by applying the DFT-D3 method with the Becke–Johnson damping function to all calculations.33 Spin polarization was not considered in the computations. The “mean atomic volume” (MAV) was defined as the ratio of the cell volume to the total number of atoms, with the cell volume determined from lattice constants obtained through structural optimization in this study.
Elastic moduli were determined from the computational results of elastic tensors calculated using the Voigt–Reuss–Hill approximation methods.16 The bulk modulus (B) and shear modulus (G) are obtained as the average values (the Hill's prediction values) of those calculated using the Voigt's (BV and GV) and the Reuss's (BR and GR) prediction methods (eqn (1)–(6)). In these equations, each component of the inverse elasticity tensor is represented by Sij. The polycrystalline Young's modulus (E) and Poisson's ratio (ν) were estimated using eqn (7) and (8), respectively.
9BV = (C11 + C22 + C33) + 2(C12 + C23 + C31) | (1) |
1/BR = (S11 + S22 + S33) + 2(S12 + S23 + S31) | (2) |
B = (BV + BR)/2 | (3) |
15GV = (C11 + C22 + C33) − (C12 + C23 + C31) + 3(C44 + C55 + C66) | (4) |
15/GR = 4(S11 + S22 + S33) − 4(S12 + S23 + S31) + 3(S44 + S55 + S66) | (5) |
G = (GV + GR)/2 | (6) |
E = 9BG/(3B + G) | (7) |
ν = (3B − 2G)/2(3B + G) | (8) |
These elastic properties were extracted from the elastic tensors using VASPKIT.34 Notably, these calculations were conducted by assuming a temperature of 0 K.
In some of the electrolytes investigated in this study, certain Li sites exhibit partial occupancies or disorder. While such site occupancies can influence the calculated elastic properties, the Li positions with the highest occupancies and/or highest symmetry were selected for the structural models used in the calculations. This approach was adopted to ensure a consistent and representative comparison across different materials.
For the argyrodite-type structures, Li6PS5Cl, Li6PS5Br, Li6PS5I, and Li6SbS5I were selected. Li6PS5Br and Li6PS5I were included because Br and I are commonly used as halogens alongside Cl.9 Li6SbS5I, an argyrodite material previously reported by our group, demonstrated an ionic conductivity of 2.1 × 10−6 S cm−1, which was higher than that of Li6PS5I.35
For LGPS-type structures, Li10GeP2S12, Li10SnP2S12 and Li10SiP2S12 were selected. Additionally, several thio-LISICON-type electrolytes, including LixXS4 solid electrolytes (X = B, Al, Si, Ga, Ge, Sn, and Sb),36–42 were included as sulfide solid electrolytes. These materials can serve as solid electrolytes themselves or as end members for other electrolytes, such as LGPS-type materials. Li4P2S6 was also included as a relatively stable solid electrolyte, maintaining its crystal structure up to 950 °C in vacuum and 280 °C in air.43 The low-temperature (LT) phase of Li7PS6 was introduced as another argyrodite electrolyte.44
Beyond garnet-type structures, various oxide electrolytes with high ionic conductivities45 were considered, including Na super-ionic conductor (NASICON)-type LiM2(PO4)3 electrolytes (M = Ti, Ge, Zr, and Hf),46 Li super-ionic conductor (LISICON)-type Li2ZnGeO4 electrolyte,47 lithium phosphorus oxynitride (LiPON) electrolytes such as Li2PNO2,48 perovskite electrolytes (Li1/8La5/8TiO3 and Li1/2La1/2TiO3),49 and anti-perovskite electrolytes (Li3OCl, Li3OBr, and Li3OCl0.5Br0.5).50 ortho-Oxyacid electrolytes such as Li3PO4 and Li2SO4, although generally exhibiting low ionic conductivities, were also considered. Our previous studies reported electrolytes in the Li3BO3–Li2SO451 and Li4GeO4–Li3VO452 systems using ortho-oxyacid electrolytes as end members.
Nitride electrolytes other than Li3N include LiSi2N3, Li7PN4, and LiPN2.53,54 Halide electrolytes such as LiAlCl4, Li2CdCl4, Li2MgCl4, and Li2ZnI455–57 were studied alongside Li3YCl6.14 Li3MCl6-based electrolytes for various M elements have also been explored, with Li3InCl6 often used together with Li3YCl6 owing to its high ionic conductivity.58 These solid electrolyte materials were selected based on a previously reported review on their ionic conductivities.59
To calculate the elastic properties of these materials, input structural files were primarily obtained from the Materials Project60 or the ICSD database.61 The exception was the input file for the Li6SbS5I crystal structure, which was separately created based on the Li6PS5Cl structure.
Type | Formula | Space group | Crystal system | E/GPa | B/GPa | G/GPa | ν | B/G | MAV/cm3 mol−1 | Ref. |
---|---|---|---|---|---|---|---|---|---|---|
Argyrodite | Li6PS5Cl | F![]() |
Cubic | 22.1 | 28.7 | 8.1 | 0.37 | 3.54 | No data | 19 |
Argyrodite | Li6PS5Br | F![]() |
Cubic | 25.3 | 29.0 | 9.3 | 0.35 | 3.11 | No data | 19 |
Argyrodite | Li6PS5I | F![]() |
Cubic | 30.0 | 29.9 | 11.3 | 0.33 | 2.65 | No data | 19 |
Argyrodite | Li6PS5Cl | F![]() |
Cubic | 27.44 | 34.73 | 10.03 | 0.37 | 3.46 | 11.74 | This study |
Argyrodite | Li6PS5Br | F![]() |
Cubic | 30.11 | 35.19 | 11.09 | 0.36 | 3.17 | 11.75 | This study |
Argyrodite | Li6PS5I | F![]() |
Cubic | 35.10 | 35.14 | 13.16 | 0.33 | 2.67 | 11.88 | This study |
Argyrodite | Li6SbS5I | F![]() |
Cubic | 38.27 | 32.88 | 14.65 | 0.31 | 2.24 | 12.87 | This study |
Thio-LISICON (LGPS-type) | Li10GeP2S12 | P42mc | Tetragonal | 37.2 | 30.4 | 14.4 | 0.30 | 2.12 | 11.59 | 61 |
Thio-LISICON (LGPS-type) | Li10GeP2S12 | P42mc | Tetragonal | 21.7 | 27.3 | 7.9 | 0.37 | 3.46 | No data | 19 |
Thio-LISICON (LGPS-type) | Li10SnP2S12 | P42mc | Tetragonal | 29.1 | 23.5 | 11.2 | 0.29 | 2.10 | No data | 19 |
Thio-LISICON (LGPS-type) | Li10SiP2S12 | P42mc | Tetragonal | 24.8 | 27.8 | 9.2 | 0.35 | 3.02 | No data | 19 |
Thio-LISICON (LGPS-type) | Li10GeP2S12 | P42mc | Tetragonal | 28.03 | 32.40 | 10.34 | 0.36 | 3.13 | 10.89 | This study |
Thio-LISICON (LGPS-type) | Li10SnP2S12 | P42mc | Tetragonal | 26.01 | 30.97 | 9.56 | 0.36 | 3.24 | 11.14 | This study |
Thio-LISICON (LGPS-type) | Li10SiP2S12 | P42mc | Tetragonal | 28.21 | 32.74 | 10.40 | 0.36 | 3.15 | 10.77 | This study |
Thio-LISICON | β-Li3PS4 | Pnma | Orthorhombic | 29.5 | 23.3 | 11.4 | 0.29 | 2.04 | No data | 19 |
Thio-LISICON | γ-Li3PS4 | Pmn21 | Orthorhombic | 33.4 | 32.9 | 12.6 | 0.33 | 2.61 | No data | 19 |
Thio-LISICON | β-Li3PS4 | Pnma | Orthorhombic | 33.49 | 29.28 | 12.79 | 0.30 | 2.29 | 11.34 | This study |
Thio-LISICON | γ-Li3PS4 | Pmn21 | Orthorhombic | 36.82 | 39.34 | 13.70 | 0.34 | 2.87 | 11.23 | This study |
Thio-LISICON | Li4SnS4 | Pnma | Orthorhombic | 42.90 | 38.45 | 16.32 | 0.31 | 2.36 | 11.30 | This study |
Thio-LISICON | Li4GeS4 | Pnma | Orthorhombic | 46.42 | 40.73 | 17.72 | 0.31 | 2.30 | 10.65 | This study |
Thio-LISICON | Li5AlS4 | P21/m | Monoclinic | 61.46 | 43.55 | 24.3 | 0.26 | 1.79 | 9.36 | This study |
Thio-LISICON | Li3BS3 | Pnma | Orthorhombic | 35.96 | 30.16 | 13.82 | 0.30 | 2.18 | 10.15 | This study |
Thio-LISICON | Li5GaS4 | P21/m | Monoclinic | 59.65 | 43.17 | 23.49 | 0.27 | 1.84 | 9.45 | This study |
Thio-LISICON | Li3SbS4 | Pmn21 | Orthorhombic | 33.86 | 35.48 | 12.62 | 0.34 | 2.81 | 12.57 | This study |
Other sulfides | Li7P3S11 | P![]() |
Triclinic | 21.9 | 23.9 | 8.1 | 0.35 | 2.95 | No data | 19 |
Other sulfides | Li7P3S11 | P![]() |
Triclinic | 28.37 | 29.52 | 10.59 | 0.34 | 2.79 | 11.39 | This study |
Other sulfides | Li7PS6 | Pna21 | Orthorhombic | 51.88 | 41.15 | 20.11 | 0.29 | 2.05 | 9.95 | This study |
Other sulfides | Li4P2S6 | P![]() |
Trigonal | 81.50 | 46.14 | 33.80 | 0.21 | 1.37 | 9.99 | This study |
Other sulfides | Li2SiS3 | Cmc21 | Orthorhombic | 47.91 | 44.53 | 18.14 | 0.32 | 2.45 | 11.28 | This study |
Other sulfides | Li2SnS3 | C2/c | Monoclinic | 92.23 | 51.41 | 38.40 | 0.20 | 1.34 | 10.47 | This study |
Type | Formula | Space group | Crystal system | E/GPa | B/GPa | G/GPa | ν | B/G | MAV/cm3 mol−1 | Ref. |
---|---|---|---|---|---|---|---|---|---|---|
ortho-Oxyacid | γ-Li3PO4 | Pnma | Orthorhombic | 103.4 | 72.5 | 40.9 | 0.26 | 1.77 | No data | 19 |
ortho-Oxyacid | γ-Li3PO4 | Pnma | Orthorhombic | 114.23 | 82.32 | 45.02 | 0.27 | 1.83 | 5.82 | This study |
ortho-Oxyacid | β-Li3PO4 | Pmn21 | Orthorhombic | 118.56 | 83.40 | 46.93 | 0.26 | 1.78 | 5.71 | This study |
ortho-Oxyacid | α-Li2SO4 | F![]() |
Cubic | 26.77 | 41.19 | 9.62 | 0.39 | 4.28 | 9.61 | This study |
ortho-Oxyacid | β-Li2SO4 | P21/c | Monoclinic | 50.36 | 40.92 | 19.44 | 0.29 | 2.10 | 6.92 | This study |
ortho-Oxyacid | Li3BO3 | P21/c | Monoclinic | 120.49 | 87.31 | 47.44 | 0.27 | 1.84 | 4.47 | This study |
ortho-Oxyacid | Li3VO4 | Pmn21 | Orthorhombic | 93.64 | 72.93 | 36.41 | 0.29 | 2.00 | 6.25 | This study |
ortho-Oxyacid | Li4GeO4 | Cmcm | Orthorhombic | 137.49 | 94.91 | 54.62 | 0.26 | 1.74 | 5.69 | This study |
ortho-Oxyacid | Li4SiO4 | P![]() |
Triclinic | 146.80 | 91.08 | 59.61 | 0.23 | 1.53 | 5.40 | This study |
ortho-Oxyacid | Li4SiO4 | P21/m | Monoclinic | 127.91 | 88.39 | 50.81 | 0.26 | 1.74 | 5.39 | This study |
ortho-Oxyacid | Li5AlO4 | Pbca | Orthorhombic | 143.72 | 93.98 | 57.71 | 0.24 | 1.63 | 5.45 | This study |
LISICON | Li2ZnGeO4 | Pc | Monoclinic | 115.16 | 91.82 | 44.60 | 0.30 | 2.06 | 6.55 | This study |
Oxynitride | Li2PNO2 | Cmc21 | Orthorhombic | 160.02 | 100.65 | 64.78 | 0.24 | 1.55 | 5.66 | This study |
Garnet | Li7La3Zr2O12 | I41/acd | Tetragonal | 175.1 | 127.4 | 68.9 | 0.27 | 1.85 | No data | 19 |
Garnet | Li5La3Nb2O12 | Ia![]() |
Cubic | 141.1 | 111.3 | 54.8 | 0.29 | 2.03 | No data | 19 |
Garnet | Li5La3Ta2O12 | Ia![]() |
Cubic | 144.2 | 112.0 | 56.1 | 0.29 | 2.00 | No data | 19 |
Garnet | Li6.25Al0.25La3Zr2O12 | Ia![]() |
Cubic | 162.6 | 112.4 | 64.6 | 0.26 | 1.74 | No data | 62 |
Garnet | Li6.5La3Zr1.5Ta0.5O12 | Ia![]() |
Cubic | 154.9 | 99.2 | 62.5 | 0.24 | 1.59 | No data | 62 |
Garnet | Li7La3Zr2O12 | I41/acd | Tetragonal | 165.37 | 120.54 | 65.04 | 0.27 | 1.85 | 6.85 | This study |
Garnet | Li5La3Nb2O12 | Ia![]() |
Cubic | 133.67 | 106.52 | 51.78 | 0.29 | 2.06 | 7.66 | This study |
Garnet | Li5La3Ta2O12 | Ia![]() |
Cubic | 141.93 | 112.40 | 55.03 | 0.29 | 2.04 | 7.54 | This study |
NASICON | LiTi2(PO4)3 | R![]() |
Trigonal | 143.7 | 95.0 | 57.6 | 0.25 | 1.65 | No data | 19 |
NASICON | LiTi2(PO4)3 | R![]() |
Trigonal | 146.16 | 100.62 | 58.10 | 0.26 | 1.73 | 7.45 | This study |
NASICON | LiGe2(PO4)3 | R![]() |
Trigonal | 183.79 | 121.42 | 73.65 | 0.25 | 1.65 | 6.90 | This study |
NASICON | LiZr2(PO4)3 | R![]() |
Trigonal | 105.33 | 78.26 | 41.28 | 0.28 | 1.90 | 8.41 | This study |
NASICON | LiHf2(PO4)3 | R![]() |
Trigonal | 122.81 | 85.83 | 48.68 | 0.26 | 1.76 | 8.14 | This study |
Perovskite | Li1/8La5/8TiO3 | Pmm2 | Orthorhombic | 233.9 | 179.0 | 91.2 | 0.28 | 1.96 | No data | 19 |
Perovskite | Li1/2La1/2TiO3 | P2/c | Monoclinic | 262.5 | 183.5 | 104.0 | 0.26 | 1.76 | No data | 19 |
Perovskite | Li1/8La5/8TiO3 | Pmm2 | Orthorhombic | 220.64 | 165.6 | 86.33 | 0.28 | 1.92 | 7.57 | This study |
Perovskite | Li1/2La1/2TiO3 | P2/c | Monoclinic | 270.36 | 170.75 | 109.4 | 0.24 | 1.56 | 7.18 | This study |
Anti-perovskite | Li3OCl | Pm![]() |
Cubic | 99.7 | 55.7 | 41.5 | 0.20 | 1.34 | No data | 19 |
Anti-perovskite | Li3OBr | Pm![]() |
Cubic | 92.8 | 52.3 | 38.5 | 0.20 | 1.36 | No data | 19 |
Anti-perovskite | Li3OCl | Pm![]() |
Cubic | 113.89 | 62.22 | 47.65 | 0.19 | 1.31 | 6.63 | This study |
Anti-perovskite | Li3OBr | Pm![]() |
Cubic | 108.80 | 59.62 | 45.49 | 0.20 | 1.31 | 7.14 | This study |
Anti-perovskite | Li3OCl0.5Br0.5 | Pm![]() |
Cubic | 110.85 | 60.78 | 46.34 | 0.20 | 1.31 | 6.9 | This study |
Type | Formula | Space group | Crystal system | E/GPa | B/GPa | G/GPa | ν | B/G | MAV/cm3 mol−1 | Ref. |
---|---|---|---|---|---|---|---|---|---|---|
Nitride | α-Li3N | P6/mmm | Hexagonal | 105.79 | 65.83 | 42.93 | 0.23 | 1.53 | 6.24 | This study |
Nitride | β-Li3N | P63/mmc | Hexagonal | 120.43 | 70.18 | 49.60 | 0.21 | 1.41 | 5.18 | This study |
Nitride | Li3BN2 | P42/mnm | Tetragonal | 140.51 | 83.64 | 57.59 | 0.22 | 1.45 | 5.36 | This study |
Nitride | Li7PN4 | P![]() |
Cubic | 221.44 | 105.52 | 96.26 | 0.15 | 1.10 | 4.83 | This study |
Nitride | LiSi2N3 | Cmc21 | Orthorhombic | 300.80 | 167.44 | 125.27 | 0.20 | 1.34 | 5.79 | This study |
Nitride | LiPN2 | I![]() |
Tetragonal | 254.93 | 149.84 | 104.78 | 0.22 | 1.43 | 5.51 | This study |
Chloride | Li3YCl6 | P![]() |
Trigonal | 41.9 | 28.9 | 16.6 | No data | 1.74 | No data | 15 |
Chloride | Li3InCl6 | P![]() |
Trigonal | 44.5 | 30.3 | 17.7 | No data | 1.71 | No data | 15 |
Chloride | Li3YCl6 | P![]() |
Trigonal | 41.05 | 28.41 | 16.30 | 0.26 | 1.74 | 12.54 | This study |
Chloride | Li3InCl6 | P![]() |
Trigonal | 43.13 | 29.80 | 17.13 | 0.26 | 1.74 | 12.04 | This study |
Chloride | LiAlCl4 | P21/c | Monoclinic | 20.37 | 13.45 | 8.16 | 0.25 | 1.65 | 14.05 | This study |
Chloride | Li2CdCl4 | Imma | Orthorhombic | 33.00 | 33.19 | 12.37 | 0.33 | 2.68 | 12.61 | This study |
Chloride | Li2MgCl4 | Imma | Orthorhombic | 37.53 | 34.57 | 14.23 | 0.32 | 2.43 | 11.73 | This study |
Iodide | Li2ZnI4 | Pnma | Orthorhombic | 26.46 | 16.42 | 10.74 | 0.23 | 1.53 | 17.99 | This study |
Sulfide electrolytes exhibited generally low elastic moduli and high B/G ratios, indicating their high deformability and ductility. Specifically, the majority of their elastic moduli—E, B, and G—were below 50, 40, and 20 GPa, respectively. Moreover, most of their Pugh's ratios (B/G) exceeded the critical value of 1.75, which is commonly used to differentiate between brittle and ductile materials. These findings underscore the inherent deformability and ductility of sulfide electrolytes. Among them, argyrodite-type and LGPS-type electrolytes displayed notably low elastic moduli and B/G ratios exceeding 2.5 (except for Li6SbS5I, with B/G = 2.24). Particularly, their shear moduli (G) were remarkably low (<15.0 GPa), signifying a high likelihood of plastic deformation within their crystal structures. Additionally, materials such as β-Li3PS4, Li3BS3, Li3SbS4, and Li7P3S11 demonstrated relatively low elastic moduli, with γ-Li3PS4, Li3SbS4, and Li7P3S11 exhibiting B/G values of approximately 2.8.
Conversely, exceptions such as Li4P2S6 and monoclinic Li2SnS3 exhibited relatively high elastic moduli. The polycrystalline Young's moduli E for these materials reached 81.50 and 92.23 GPa, respectively, whereas their shear moduli (G) were 33.80 and 38.40 GPa. These values were significantly higher than those of other sulfide materials and more than twice as high as those of argyrodite and LGPS-type materials. Despite their elevated shear modulus, their bulk moduli were only marginally larger than those of other sulfide materials (Li4P2S6: 46.14 GPa, Li2SnS3: 51.41 GPa). This combination of a relatively high shear modulus and moderate bulk modulus yielded a Pugh's ratio of <1.5. Similarly, Li5AlS4 and Li5GaS4 also exhibited relatively high elastic moduli and a low B/G ratio of approximately 1.75. Overall, Pugh's ratios among sulfide electrolytes tend to be highest for materials belonging to the argyrodite and LGPS-type families. In contrast, the other thio-LISICON materials exhibit moderately high values, yet generally lower than those of argyrodite and LGPS. In addition, the other sulfide systems typically show even lower Pugh's ratios.
To validate our computational results, we compared them with elastic moduli experimentally determined via ultrasonic pulse-echo measurements. We have previously measured the elastic moduli of four amorphous materials synthesized through mechanical milling: Li3PS4, Li7P3S11, Li2SiS3, and Li4GeS4. These results are summarized in Table S2. Although the experimentally observed trend in elastic moduli is generally consistent with the computational predictions, the absolute values obtained from experiments are lower than those calculated. This discrepancy is likely due to three primary contributing factors: (1) amorphous materials typically exhibit larger mean atomic volumes due to the presence of free volume; (2) the relative density (i.e., packing density) of the pelletized samples is below 100%; and (3) the lattice constants at 0 K tend to be smaller than those at room temperature, which can lead to an overestimation of the elastic moduli in the computational results.
The oxide electrolytes generally exhibited relatively high elastic moduli, with the Pugh's ratio B/G for most of them being less than 2.0. Specifically, their polycrystalline Young's moduli E generally ranged from 90 to 170 GPa, with the exception of Li2SO4 and the perovskite-type materials. Among these, Li2SO4, which contains both sulfur and oxygen, demonstrated unique elastic properties. Both α-Li2SO4 and β-Li2SO4 exhibited relatively low elastic moduli. Notably, α-Li2SO4 had an exceptionally high B/G value of 4.28, whereas the value of β-Li2SO4 was somewhat lower (B/G = 2.10). The α-phase of Li2SO4, which exhibits higher ionic conductivity than the β-phase, undergoes a phase transition from b to a at 577 °C, enhancing its ionic conductivity through the paddlewheel mechanism.64
While Li2SO4 alone is not widely recognized as a high-conductivity solid-state electrolyte at room temperature, its incorporation into Li3BO3-based glass and glass–ceramic electrolytes has been reported to enhance formability.65 This improved formability is essential for achieving high-density electrolyte pellets, leading to higher ionic conductivity. The improved formability is particularly attributed to the ductility of Li2SO4. Previous studies have highlighted the relevance of Pugh's ratio (B/G) as an indicator of ductility, with materials exhibiting a Pugh's ratio greater than 1.75 generally regarded as ductile. In this study, α-Li2SO4 exhibited relatively high Pugh's ratios, suggesting their intrinsic ductility. Such ductile behavior is expected to promote more efficient compaction during the pressing process, leading to higher pellet densities.
Additionally, α-Li2SO4 exhibited a larger MAV compared to β-Li2SO4 and other oxide electrolytes, which was comparable to those of sulfides. This characteristic is thought to be a key factor contributing to its lower E and G values. Interestingly, the bulk moduli of α-Li2SO4 and β-Li2SO4 were nearly identical, despite the significant difference in their MAVs.
Perovskite-type structures generally displayed relatively high elastic moduli, with E values exceeding 200 GPa, B values of approximately 170 GPa, and G values of approximately 100 GPa. Moreover, anti-perovskite structures exhibited significantly lower B/G ratios (<1.5). These materials had relatively low bulk moduli (of approximately 60 GPa), contributing to their reduced B/G ratios.
Nitride electrolyte structures also displayed high elastic constants and low B/G ratios. Specifically, their B/G values (<1.5) were generally lower than those of oxides (∼1.75). Among the nitrides, α-Li3N exhibited relatively low elastic moduli (E = 105.79 GPa, B = 65.83 GPa, G = 42.93 GPa); however, these values were still higher than those of other electrolyte materials, particularly sulfides and halides. The elastic moduli of Li7PN4 and LiSi2N3 were comparable to those of perovskite oxides. Notably, the shear modulus of LiSi2N3 was approximately 125 GPa, marking the highest value among all materials analyzed in this study.
In contrast, the elastic moduli of halides were generally low and comparable to those of sulfides. Specifically, LiAlCl4 exhibited significantly low elastic moduli (E = 20.37 GPa, B = 13.45 GPa, G = 8.16 GPa). However, their B/G ratios varied among the halide materials selected in this study. Notably, Li2CdCl4 had a higher Pugh's ratio (B/G = 2.68), suggesting that its ductility is comparable to that of several sulfide materials. In contrast, the B/G of Li3MCl6-type (M = Y, In) was approximately 1.75, indicating that their ductility is not superior to that of sulfide materials. The elastic moduli of these materials were calculated in a previous study,15 and their values were generally consistent with those reported in this study.
Fig. 1 shows the relationship between the calculated shear modulus (G) and bulk modulus (B) values from the computational results. Fig. 1(a) presents the results for sulfide and halide crystals, which generally exhibit relatively low elastic moduli. Most of their Pugh's ratios (B/G) are higher than 1.75 (above the brown dotted line), indicating that these electrolytes are likely to be ductile materials. The mechanical ductility of actual solid electrolytes (which are in the form of powder compacts) within a battery cannot be fully predicted solely based on the calculated Pugh's ratio. Fig. 1(b) shows the computation results for oxides and nitrides, which have higher elastic constants. All the oxides are generally distributed around B/G = 1.75 (within the range of 1.5–2.0), except for α-Li2SO4. By contrast, the nitrides are divided into two groups: those with B/G ratios of ∼1.75 and <1.75. Furthermore, the B/G ratio of α-Li2SO4, indicated by the red circle in the figure, is uniquely positioned compared to other materials.
Fig. 2 illustrates the relationship between MAV and (a) bulk modulus (B), (b) polycrystalline Young's modulus (E), and (c) shear modulus (G) in our computational results. The data indicate a negative nonlinear correlation between MAV and elastic constants across all electrolyte types. These correlations align with the relationship between MAV and Young's modulus observed in glass electrolytes reported by our group.18 Oxide and nitride electrolytes typically have low MAV and high elastic moduli, while sulfides and halides exhibit the opposite trend, with high MAV and low elastic moduli. Notably, Li2SO4 displays a unique distribution in Fig. 2. Despite having an MAV similar to that of oxides, β-Li2SO4 shows elastic moduli comparable to those of sulfides. Moreover, the plot for α-Li2SO4 in Fig. 2 aligns within the distribution of sulfides, despite containing oxygen, showing elastic properties similar to sulfides that do not contain oxygen.
![]() | ||
Fig. 2 Correlation between mean atomic volume and (a) bulk moduli B, (b) polycrystalline elastic moduli E, and (c) shear moduli G of various crystal structures of Li-ion conductors. |
Notably, oxide and nitride electrolytes display a wide range of elastic constants even with similar MAV values. This suggests that factors beyond MAV, such as crystal structure and constituent elements, also play a significant role in determining the elastic moduli of nitrides and oxides. In contrast, the variation in elastic constants is relatively small for the sulfide and halide electrolytes selected in this study.
Detailed information on the solid electrolyte materials used in this study and experimentally obtained elastic moduli of amorphous electrolytes prepared via mechanical milling are available in SI. See DOI: https://doi.org/10.1039/d5ma00733j
This journal is © The Royal Society of Chemistry 2025 |