2022 Volume 90 Issue 6 Pages 067005
Technology for ion conduction in Li2S nanostructure has attracted considerable interest for achieving the intrinsic performance of Li-S batteries and for obtaining highly conductive materials. Herein, a quantitative relationship between ion transport and relaxation behavior in the Li2S–CaS nanocomposite is revealed based on complex conductivity and electric modulus formalism. Li2S–CaS nanocomposites prepared by high-energy ball milling show higher conductivity and activation energy for conduction than pure Li2S. The activation energy for relaxation is lower than the activation energy for conduction in 80Li2S·20CaS (mol%). The long-range ion transport involves high activation energy compared with the hopping of carriers at localized states in 80Li2S·20CaS. Additionally, the dual doping of Ca and halogen reduces the activation energy for conduction and improves the conductivity of Li2S. The findings show important insight for understanding ion transport in nanocomposites containing a Li2S nanostructure.

Lithium sulfide is attractive as a cathode material for Li–S batteries owing to its interesting properties, e.g., high electrochemical capacity (1,167 mAh g−1).1,2 A Li-S battery based on a Li2S cathode allows for coupling with a high-capacity anode such as Si and Sn; this is an appealing alternative for conventional Li ion batteries.1,3 However, the low electronic and ionic conductivity of Li2S limits the complete utilization of its intrinsic capacity. Achievements of high-power density in Li-S batteries require understanding the transport properties of Li2S. Recently, multivalent cation doping to Li2S improved its ionic conductivity and extracted a high battery performance as a cathode in all-solid-state Li-S batteries.4–6 Both Li2S–Al2S3 and Li2S–MgS solid solution demonstrated a higher conductivity (by two orders of magnitude of 10−6 S cm−1 at room temperature) than the milled Li2S because of the nanoionics. In addition, AlI3-doped Li2S cathode materials offered enhanced reversible capacity.7 Based on solid-state chemistry, the multiple unique nature of nanostructured materials containing Li2S has inspired several devices that use rechargeable batteries.
Lithium sulfide plays an important role as a starting material for obtaining highly conductive materials for solid-state batteries.8 Studies on sulfide ionic conductors were originally conducted in Li2S-based glass conductors. In the 1980s, Li2S-based glass, e.g., Li2S–P2S5–LiI,9 delivered an ionic conductivity of the order of 10−3 S cm−1 at room temperature. Extensive studies have discovered crystalline ionic conductors of the order 10−2 S cm−1 at room temperature, such as Li10GeP2S12,10 Li7P3S11,11 and argyrodite.12,13 These material classes have evolved through cation substitution on P-site, as well as anion substitution. Recently, cation substitution on the site for a charge carrier identified superionic conductors involving Li5.4Al0.2PS5Br,14 Li5.35Ca0.1PS4.5Cl1.55,13 and Na2.730Ca0.135PS4.15 Interestingly, both ionic conductivity and activation energy of the Na2.730Ca0.135PS4 increased from a pristine Na3PS4 solid electrolyte. Ca doping in Li ion and sodium ion conductors is of considerable interest for expressing fast ionic conduction.13,16,17
Herein, we report the effect of adding Ca ions and dual ions on ion-transport properties in a Li2S structure. Note that (100 − x) Li2S·xCaS (mol%) nanocomposites, prepared by high-energy ball milling, exhibit crystallite sizes ranging from 7 to 31 nm. The Li2S–CaS nanocomposites demonstrate higher conductivity and activation energy for conduction than the milled pure Li2S. To understand the fundamental relationship between activation energy for conduction and relaxation in (100 − x)Li2S·xCaS, we examined the complex conductivity and electric modulus formalism. The activation energy for electric relaxation is lower than the activation energy for conduction in 80Li2S·20CaS. The result indicates that long-range ion transport, including the contribution of a grain boundary, involves higher energy barriers than the hopping of carriers at localized states. The dual doping of Ca and halogen into Li2S reduces the activation energy of conduction and improved the conductivity. These important insights can provide an understanding of ion transport properties in functional materials for batteries involving Li2S.
Lithium sulfide (99.9 %, Mitsuwa), CaS (99.99 %, Kojundo Laboratory), CaF2 (99.99 %, Sigma-Aldrich). CaCl2 (99.99 %, Sigma-Aldrich), CaBr2 (99.98 %, Sigma-Aldrich), or CaI2 (99.999 %, Sigma-Aldrich) were mixed at appropriate molar ratios. Then, the powders obtained were loaded in 45 ml of zirconia pots with φ 10 mm zirconia balls and subjected to high-energy ball milling with planetary ball milling (Fritsch Pulverisete 7 Premium line) at a rotation speed of 510 rpm for 10 h (1 h milling; 15 min pause).
2.2 Material characterizationPowder X-ray diffraction (XRD) measurements were performed under 2θ = 10°–80° in a step interval of 0.02° at a scan rate of 0.2° min−1 using a Rigaku Ultima IV diffractometer. An X-ray beam was generated using CuKα radiation (40 kV, 30 mA). For XRD measurements, we used an XRD holder with a beryllium window (Rigaku). Subsequently, structural refinements were analyzed by the Rietveld method using the RIETAN-FP computer program.18 The crystallite size, D, and the microstrain, ε, for phases in the prepared nanocomposites were calculated from the Williamson–Hall method presented in Eq. 1:
| \begin{equation} \beta \cos\theta = C\epsilon \sin\theta + \frac{K\lambda}{D}, \end{equation} | (1) |
The overall conductivities of prepared samples were determined by alternating current impedance spectroscopy (SI 1260, Solatron) in a frequency range of 1 MHz to 10 Hz under a dry Ar flow. To fabricate cells for electrochemical impedance spectroscopy (EIS) measurements, each sample (∼80 mg) was filled in a holder made of polyether ether ketone (PEEK) with two stainless steel rods as blocking electrodes. We applied a uniaxial pressure of 256 MPa to pellets having diameters of ∼10.0 mm at room temperature. The complex conductivity, σ$^{*}$, of samples was determined from the complex impedance, Z$^{*}$, obtained using EIS data and is reported as follows:
| \begin{equation} \sigma^{*} = \sigma' + j\sigma'' = \frac{l}{AZ^{*}} = \frac{l}{AZ'} + j\frac{l}{AZ''}, \end{equation} | (2) |
The complex electric modulus, M$^{*}$, was determined from Eqs. 3 and 4, which are related to complex conductivity and permittivity, respectively, as follows:
| \begin{equation} \varepsilon^{*} = \varepsilon'' - j\varepsilon'' = \frac{\sigma''}{\omega \varepsilon_{0}} - j\frac{\sigma'}{\omega\varepsilon_{0}} \end{equation} | (3) |
| \begin{align} M^{*} &= \frac{1}{\varepsilon^{*}} = \frac{1}{(\varepsilon' - j\varepsilon'')} \\ &= M' + jM'' = \frac{\varepsilon'}{(\varepsilon')^{2} + (\varepsilon'')^{2}} + j\frac{\varepsilon''}{(\varepsilon')^{2} + (\varepsilon'')^{2}}, \end{align} | (4) |
Figure 1a shows the XRD patterns of (100 − x)Li2S·xCaS with Si powder as a standard. The XRD patterns for all samples show diffraction peaks attributed to a crystalline phase of Li2S with the space group Fm-3m. These patterns show no change in peak position and no side reaction after adding CaS. The intensity of the peak at 27° decreases with an increasing CaS content, whereas the intensity of the most intense peak at 31° in the XRD pattern of the CaS phase increases. These experimental results indicate that Li2S–CaS prepared by high-energy ball milling comprises a mixture of Li2S and CaS phases without the substitution of Ca in Li ions. To examine lattice distortion and mean crystallite size of mechanically treated Li2S–CaS, we performed Rietveld refinement based on XRD data. Figure 1b shows the Rietveld refinement pattern for 80Li2S·20CaS. A multi-phase crystal model, including Li2S and CaS phases, offers R factors with good fitness, which are sufficiently low to support the extracted structural model. Rwp, RB, and RF obtained for the Li2S phase are 10.1 %, 2.4 %, and 2.0 %, respectively. Figure 1c shows the mean crystallite sizes and the microstrain as a function of composition x in (100 − x)Li2S·xCaS. Both the mean crystallite size and microstrain were estimated using the Williamson–Hall method (Figs. S1a and S1b). The Williamson–Hall plot for the sample with x = 50 deviates from a linear relationship because of the anisotropic line broadening. The Li2S–CaS prepared by high-energy ball milling forms nanocrystalline composites with crystallite sizes ranging from ∼7–31 nm. The crystallite sizes demonstrate a decreasing trend with an increasing CaS content. The microstrain of the Li2S phase in the composites increases with the addition of CaS. Figure 2 shows FE-SEM images of (100 − x)Li2S·xCaS particles (0 ≤ x ≤ 40). The particle sizes of all samples are identical, which are estimated to be ∼50 µm. The particle morphology is rougher with increasing CaS content. On the other hand, no calcium atoms were detected on the particle surface of the Li2S-CaS doping from EDS mapping (Fig. S2). This experimental result indicates that the Li2S-CaS forms nanocomposite particles covered with the Li2S phase.

(a) The XRD patterns of (100 − x)Li2S·xCaS with added Si powder. The bottom profile shows CaS as a reference. (b) The Rietveld refinement pattern for 80Li2S·20CaS. The observed diffraction intensities and calculated patterns are denoted by red plus signs and a solid green line, respectively. The blue trace at the bottom represents differences between calculated and experimental patterns. The short green bars below the observed and calculated profiles indicate the positions of the allowed Bragg reflections of Si, CaS, and Li2S in order from the bottom. (c) The crystallite sizes and the microstrain of the Li2S and CaS phases in (100 − x)Li2S·xCaS. Circles and squares correspond to the Li2S and CaS phases, respectively.

FE-SEM images of (100 − x)Li2S·xCaS particles (0 ≤ x ≤ 40).
Figure 3a shows the real part σ′ of the complex conductivity of 80Li2S·20CaS as a function of frequency. At lower frequencies, the impedance spectra of obtained nanocrystalline samples demonstrate an obvious direct current (DC) plateau, which indicates long-range ion transport.19 The electronic conductivity of Li2S is lower than the ionic conductivity by at least two orders of magnitude in the temperature range from 30 to 450 °C; thus, the observed DC plateau is attributed to ion transport.20 The frequency-dependent dispersive region is observed at a higher frequency. Frequently, this observation has been interpreted as a fingerprint of the forward and backward jumps of carriers involving a short length scale.21,22 This suggests that the carrier repeats the polarization and its relaxation at a stable site for the carrier. A higher temperature allows for long-range ion transport in samples over a broader frequency range. In the high-frequency region, the frequency dependence of conductivity obeys Jonscher’s power law,19 i.e., σ ∝ ωn where ω is the angular frequency and exponent n is the degree of interaction between mobile ions as well as the lattice around them. Most ionic conducting compounds demonstrated no deviation from a value between 0.6 and 1.0 of exponent n.19 However, the exponent n demonstrates a value of >1 in 80Li2S·20CaS at lower temperatures than 90 °C. Furthermore, dispersive behavior in pure Li2S yields an exponent n of 1.2 at a temperature from 130 to 170 °C (Fig. S3). The higher exponent n may have originated from the migration of carriers from the occupied site to other sites via quantum mechanical tunnelling between asymmetric double-well potentials.23,24 This finding is consistent with the previous computational results that Li ions in the Li2S crystal exist in isolated tetrahedral sites in the face-centered cubic sulfur framework.25 In contrast, the 80Li2S·20CaS demonstrates an exponent n of 0.63 at 110 °C. The power law exponent n is related to the effective dimension of ion diffusion pathways. The exponent of the 80Li2S·20CaS indicates the presence of 3-dimensional Li-ion pathways in the structure.26 All samples follow the Arrhenius equation shown in Eq. 5:
| \begin{equation} \sigma_{\text{DC}}T = \sigma_{0}\exp\left(-\frac{E_{\text{a},\text{DC}}}{k_{\text{B}}T}\right), \end{equation} | (5) |
| \begin{equation} \sigma_{0} = \frac{\gamma Nq^{2}a_{0}^{2}}{6k_{\text{B}}}H_{\text{R}}f_{0}, \end{equation} | (6) |

(a) Conductivity isotherms of 80Li2S·20CaS at different temperatures ranging from 50 °C to 150 °C. The solid lines represent fitting as per Jonscher’s power law, i.e., σ ∝ ωn at the dispersive regime. (b) Arrhenius plots of the DC conductivities for (100 − x)Li2S·xCaS (0 ≤ x ≤ 50). (c) Activation energies and pre-exponential factors for (100 − x)Li2S·xCaS (0 ≤ x ≤ 50).
The investigation of dielectric peaks has long been an effective technique for examining microscopic motion in amorphous solids.22 Recently, an electric modulus analysis revealed ion dynamics in crystalline solid electrolytes.31–33 To establish quantitative relationships between the activation energy for conduction and relaxation in (100 − x)Li2S·xCaS, we investigated based on complex electric modulus formalism. Figure 4a shows the imaginary part of electric modulus, M′′, at different temperatures versus frequency in 90Li2S·10CaS. Figure S4 shows the electric modulus dependence frequency for samples with other compositions. The relaxation peaks appear in all samples, and their positions shift toward higher frequencies with increasing temperature. The relaxation peak shows a transition from the overall averaged long-range to the short-range carrier migration, which corresponds to the relaxation frequency, fmax. The mean relaxation time, τ, is determined from the τ = 1/2πfmax relationship and follows the Arrhenius equation in Eq. 7 as follows:
| \begin{equation} \tau = \tau_{0}\exp\left(-\frac{E_{\text{a},\tau}}{k_{\text{B}}T}\right), \end{equation} | (7) |

(a) The imaginary part of the modulus, M′′, versus the frequency of 90Li2S·10CaS at temperatures ranging from 70 °C to 150 °C. (b) The τM−1 rate versus the inverse temperature of (100 − x)Li2S·xCaS. Dashed lines represent the linear fit. (c) For comparison, the activation energy for electric relaxation and conduction as a function of composition x in (100 − x)Li2S·xCaS.
Dual doping is an effective strategy for improving ionic conductivity. In addition to CaS doping, we examined the effect of adding CaX2 (X = F, Cl, Br, and I) on structures and ionic conductivities of Li2S. Figure 5 shows the XRD patterns of (100 − x)Li2S·xCaX2 containing Si powder as a standard sample. For all samples, the XRD patterns of (100 − x)Li2S·xCaX2 (X = F, Cl, Br, and I) are observed as the peaks attributed to Bragg reflection of Li2S crystalline phase with the space group Fm-3m. The peak intensity decreases and peak line-width broadens with an increasing CaX2 (X = F, Cl, Br, and I) content. This shows the disordered structure arising from the substitution of the anion. Note that lithium halide is formed with an x value of 20 in (100 − x)Li2S·xCaX2 (X = Cl and Br), as well as an x value of 15 in (100 − x)Li2S·xCaI2. No change in the peak position following the addition of CaX2 (X = F and Cl) is observed; however, the peak shifts toward a lower angle after adding CaX2 (X = Br and I) (Fig. S5). These peak shifts indicate that the halogen had been substituted for S2− because the variations of the lattice plane spacing correspond to the size of ionic radii of the added halogen. Unsubstituted excess halogen reacts with Li ions, thus forming a lithium halide in the samples. Despite the large difference between S2− and F− (1.84 versus 1.33 Å), the peak position in Li2S–CaF2 is independent of composition. Based on chemical defects, Lorger et al. reported that in Li2S doped with LiF, fluoride ions created comparatively large interstitial sites in the antifluorite structure of Li2S.20

The XRD patterns of (a) the (100 − x)Li2S·xCaF2, (b) (100 − x)Li2S·xCaCl2, (c) (100 − x)Li2S·xCaBr2, and (d) (100 − x)Li2S·xCaI2 with added Si powder as a standard sample.
Figure 6 plots the temperature dependence of conductivities for (100 − x)Li2S·xCaX2 (X = F, Cl, Br, and I). The ionic conductivities of all samples containing additives significantly improve compared with that of milled Li2S. Among each material system, 90Li2S·10CaI2, 80Li2S·20CaBr2, 85Li2S·15CaCl2, and 95Li2S·5CaF2 demonstrate the highest ionic conductivity of 3.6 × 10−5, 1.3 × 10−5, 1.5 × 10−5, and 1.8 × 10−7 S cm−1 at 50 °C, respectively. This improvement is attributed to the creation of Li+ vacancies following the substitution of halogen and the nanoionics. All of the samples follow the Arrhenius equation in Eq. 5 and demonstrate lower activation energy than pure Li2S (Fig. 7). The intrinsic high electron-attracting ability of F− and Cl− could lead to a higher activation energy for ion transport because conduction at the applied temperature is determined by the electrostatic association that causes ionic charge carriers to be partially trapped at dopant sites.20 However, the activation energy involves no correlation with the ionic radius of the introduced halogen in this study. The considerable difference in activation energy among material systems could not be explained only by the slightly different trapping energy based on the dopant halogen species. A high electronegativity for F and Cl ions may cause a disordered local framework structure in Li2S crystals or a strong amorphization, in addition to the expansion of conduction pathways.34 A systematic investigation reported that 90Li2S·10CaI2 demonstrates the highest ionic conductivity at 50 °C among the prepared samples owing to the presence of Li ion vacancies and expansion of Li ion pathway when introducing I− with a larger ionic radius. Moreover, the ionic conductivity of 90Li2S·10CaI2 is higher than 80Li2S·20LiI with identical I− doping levels, which may be described by the enhanced ion transport in the interface between the I-doped Li2S and CaS phases (Fig. S6). The activation energy of 90Li2S·10CaI2 reflects 0.53 eV; this is higher than that of 80Li2S·20LiI (0.48 eV). Based on the electric modulus analysis for Li2S·CaS composites, the increase in ionic conductivity and activation energy is anticipated to have originated from increased energy barriers near the interface of the CaS nanocrystalline.

Arrhenius plots of the DC conductivities for (a) the (100 − x)Li2S·xCaF2, (b) (100 − x)Li2S·xCaCl2, (c) (100 − x)Li2S·xCaBr2, and (d) (100 − x)Li2S·xCaI2.

Activation energies for conduction in (100 − x) Li2S·xCaX2 (X = F, Cl, Br, and I).
Our study reports the influence of CaS and CaX2 (X = F, Cl, Br, and I) doping on the ion transport properties of Li2S. The Li2S-CaS composites involve crystallite sizes ranging from 7 to 31 nm, and demonstrated higher conductivity and activation energy for ion transport than the milled pure Li2S. We demonstrate the quantitative relationship between the activation energy for conduction and relaxation in (100 − x)Li2S·xCaS, based on complex conductivity and electric modulus formalism. In 80Li2S·20CaS, Ea,τ values are lower than Ea,DC values. This suggests that long-range ion transport in the sample involves higher activation energy compared with the hopping of carriers at localized states, which should be influenced by the increase in energy barriers near the interface between the Li2S and CaS nanocrystallines. The dual doping of Ca and halogen allows for decreasing the activation energy for conduction and increasing the conductivity of Li2S. A systematic investigation reports that 90Li2S·10CaI2 demonstrates the highest ionic conductivity of 3.6 × 10−5 S cm−1 at 50 °C among the dual doped samples; this conductivity is higher than 80Li2S·20LiI with identical I− doping levels. Moreover, the activation energy of 90Li2S·10CaI2 (0.53 eV) is higher than that of 80Li2S·20LiI (0.48 eV). These findings support the existence of increased energy barriers near the interface of the CaS nanocrystalline. These fundamental insights can help gain an understanding of ion transport properties in functional materials for batteries involving Li2S.
Part of this study was supported by the Advanced Low Carbon Technology Specially Promoted Research for Innovative Next Generation Batteries (JST-ALCA-SPRING: JPMJAL 1301) program of the Japan Science and Technology Agency. The authors would like to thank Enago (www.enago.jp) for the English language review.
The data that support the findings of this study are openly available under the terms of the designated Creative Commons License in J-STAGE Data at https://doi.org/10.50892/data.electrochemistry.19793662.
Hirotada Gamo: Conceptualization (Equal), Data curation (Lead), Formal analysis (Lead), Investigation (Lead), Methodology (Lead), Visualization (Lead), Writing – original draft (Lead)
Nguyen Huu Huy Phuc: Conceptualization (Equal), Supervision (Equal), Writing – review & editing (Supporting)
Mika Ikari: Investigation (Supporting), Writing – review & editing (Supporting)
Kazuhiro Hikima: Project administration (Equal), Writing – review & editing (Lead)
Hiroyuki Muto: Funding acquisition (Supporting)
Atsunori Matsuda: Funding acquisition (Lead), Project administration (Equal), Resources (Lead), Supervision (Equal), Writing – review & editing (Supporting)
The authors declare no competing financial interest.
Advanced Low Carbon Technology Research and Development Program: JPMJAL 1301
N. H. H. Phuc: Present address: Faculty of Materials Technology, Ho Chi Minh City University of Technology (HCMUT), 268 Ly Thuong Kiet Str., Dist. 10, Ho Chi Minh City, Vietnam
N. H. H. Phuc: Present address: Vietnam National University Ho Chi Minh City, Linh Trung Ward, Thu Duc Dist., Ho Chi Minh City, Vietnam
H. Gamo: ECSJ Student Member
K. Hikima and A. Matsuda: ECSJ Active Members
The Williamson–Hall plots for Li2S and CaS phase in (100 − x)Li2S·xCaS (Fig. S1), SEM-EDS images of in (100 − x)Li2S·xCaS (Fig. S2), densities of the pellet (Table S1), conductivity isotherms of Li2S (Fig. S3), the imaginary part of modulus M′′ versus the frequency of Li2S, 80Li2S·20CaS, 70Li2S·30CaS, and 50Li2S·50CaS (Fig. S4), a magnified view of the XRD patterns for (100 − x)Li2S·xCaF2, (100 − x)Li2S·xCaCl2, (100 − x)Li2S·xCaBr2, and (100 − x)Li2S·xCaI2 (Fig. S5), an Arrhenius plot of the DC conductivities for 80Li2S·20LiI (Fig. S6).