Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Synergic effect of CaI2 and LiI on ionic conductivity of solution-based synthesized Li7P3S11 solid electrolyte

Tran Anh Tuabc, Tran Viet Toanab, Luu Tuan Anhab, Le Van Thangabc and Nguyen Huu Huy Phuc*abc
aFaculty of Materials Technology, Ho Chi Minh City University of Technology (HCMUT), 268 Ly Thuong Kiet Str., Dist. 10, Ho Chi Minh City, Vietnam. E-mail: nhhphuc@hcmut.edu.vn
bVietnam National University Ho Chi Minh City, Linh Trung Ward, Thu Duc Dist., Ho Chi Minh City, Vietnam
cVNU-HCM Key Laboratory for Material Technologies, Ho Chi Minh City University of Technology (HCMUT), 268 Ly Thuong Kiet Str., Dist. 10, Ho Chi Minh City, Vietnam

Received 17th January 2024 , Accepted 8th February 2024

First published on 14th February 2024


Abstract

Li7P3S11 doped with CaX2 (X = Cl, Br, I) and LiI solid electrolytes were successfully prepared by liquid-phase synthesis using acetonitrile as the reaction medium. Their structure was investigated using XRD, Raman spectroscopy and SEM-EDS. The data obtained from complex impedance spectroscopy was analyzed to study the ionic conductivity and relaxation dynamics in the prepared samples. The XRD results suggested that a part of CaX2 and LiI incorporated into the structure of Li7P3S11, while the remaining part existed at the grain boundary of the Li7P3S11 particle. The Raman peak positions of PS43− and P2S74− ions in samples 90Li7P3S11–5CaI2 and 90Li7P3S11–5CaI2–5LiI had shifted as compared to the Li7P3S11 sample, showing that CaI2 addition affected the vibration of PS43− and P2S74− ions. EDS results indicated that CaI2 and LiI were well dispersed in the prepared powder sample. The ionic conductivity at 25 °C of sample 90Li7P3S11–5CaI2–5LiI reached a very high value of 3.1 mS cm−1 due to the improvement of Li-ion movement at the grain boundary and structural improvement upon CaI2 and LiI doping. This study encouraged the application of Li7P3S11 in all-solid-state Li-ion batteries.


1. Introduction

Sulfide-based solid electrolytes (SE) have high potential for application in all-solid-state batteries because of their relatively high ionic conductivity and suitable mechanical properties compared with oxide-based and polymer-based SEs.1,2 Li7P3S11, Li9.54Si1.74P1.44S11.7Cl0.3, Li5.35Ca0.1PS4.5Cl1.55 and Li10.1P2.95Sb0.05S12I exhibited excellent ionic conductivities of about 17, 25, 10.2 and 5.9 mS cm−1 at 25 °C, respectively.3–6 Among them, Li7P3S11 has been thoroughly studied since it was invented.7 Li7P3S11 is usually prepared by solid-state reaction and achieves a highest ionic conductivity of about 17 mS cm−1 at 25 °C.8 Researchers have studied the effects of many different substances such as Li3PO4, Li3BO3, Li2ZrO3 and P2O5 on the ionic conductivity of Li7P3S11.9–12 Murakami et al. studied 6/7Li and 31P solid-state NMR to investigate the origin of high ionic conductivity of Li7P3S11 and found that the significant motion fluctuation of the P2S7 tetrahedral unit above 310 K facilitated the Li-ion movement, resulting in high ionic conductivity.13 Seino et al. found that Li7P3S11 glass ceramic with the crystallinity ranging from 50% to 80% reached its highest ionic conductivity.14

The liquid phase synthesis of Li7P3S11 has been recently introduced and takes the advantage of electrode composite preparation.15,16 Dimethoxy ethane was the first solvent ever used to prepared Li7P3S11.17 Ethyl acetate was recently employed to synthesize Li7P3S11 which exhibited high ionic conductivity of 1.05 mS cm−1 at 25 °C.18 Among of the solvents that has been employed in Li7P3S11, acetonitrile (ACN) is the most common one because the prepared SEs exhibited relatively high ionic conductivity at 25 °C, ranging from 0.8 to 1.2 mS cm−1.15,19,20 The solvent-based synthesis of Li7P3S11 is complicated and could be described as dissolution-evaporation process. It was found that Li2S reacts with P2S5 in ACN to form soluble Li4P2S7 and Li3PS4 precipitate at a molar ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]1. The residue obtained after solvent removal are Li4P2S7·ACN and Li3PS4·ACN. Li7P3S11 was formed simultaneously with ACN removal from these phases during heat treatment at high temperature. There are many studies on how to increase the ionic conductivity of Li7P3S11 produced by solid-phase reaction methods, such as using doping substances or creating defects in the crystal lattice.21–23 However, there are still lack of the information about the method to improve the ionic conductivity of Li7P3S11 prepared using liquid-phase synthesis. It was reported that CaI2 and CaS addition enhanced the conductivity of Li7P3S11.16,24 Li7P3S11 prepared using a mixture of Li2S, P2S5, and excess elemental sulfur in a mixed solvent of acetonitrile, tetrahydrofuran, and ethanol also exhibited high ionic conductivity of about 1.2 mS cm−1 at 25 °C.25

In this study, the ionic conductivity of Li7P3S11 prepared using ACN was enhanced by CaX2 (X = Cl, Br, I) and LiI doping. The 95Li7P3S11–5CaI2 solid electrolyte exhibited the ionic conductivity of about 1.0 mS cm−1 at 25 °C. The 90Li7P3S11–5CaI2–5LiI solid electrolyte showed high ionic conductivity of about 3.1 mS cm−1 at 25 °C, which was comparable to those prepared by solid-state reaction. The data obtained from AC impedance spectroscopy was interpreted in terms of conductivity isotherms, dielectric constant and dielectric loss. It was found that CaI2 and LiI doping enhanced the Li-ion movement at grain boundary and P2S74− ion motion, thus improved ionic conductivity.

2. Experimental

Chemicals

Li2S (99.9%, Macklin), P2S5 (99%, Macklin), CaCl2 (99.99%, Macklin), CaBr2 (99.99%, Aladin), CaI2 (99.999%, Macklin), LiI (99.999%, Macklin) and super dehydrated acetonitrile (Aldrich) were used as-received without any further treatment.

Liquid-phase synthesis of Li7P3S11

1.5 g of Li2S and P2S5 (7[thin space (1/6-em)]:[thin space (1/6-em)]3 in molar ratio) was weighted and put into a three-necked flask together with 40 ml of ACN. The mixture was stirred at 300 rpm and 50 °C for 24 h, the solvent was then evaporated at 80 °C under low pressure. The residue was carefully grounded using agate mortar prior to be heat treated at 270 °C for 2 h in Ar atmosphere to obtain Li7P3S11 solid electrolyte (hereafter denote as LPS).

Liquid-phase synthesis of 95Li7P3S11–5CaX2 (X = Cl, Br, I)

1.5 g of Li2S and P2S5 (7[thin space (1/6-em)]:[thin space (1/6-em)]3 in molar ratio) and appropriate amount of CaX2 to form 95Li7P3S11–5CaX2 (molar ratio) was put into a three-necked flask together with 40 ml of ACN. The mixture was stirred at 300 rpm and 50 °C for 24 h, the solvent was then evaporated at 80 °C under low pressure. The residue was carefully grounded using agate mortar prior to be heat treated at 270 °C for 2 h in Ar atmosphere to obtain 95Li7P3S11–5CaX2 solid electrolytes (hereafter denote as 5CaX2).

Liquid-phase synthesis of 90Li7P3S11–5CaI2–5LiI

1.5 g of Li2S and P2S5 (7[thin space (1/6-em)]:[thin space (1/6-em)]3 in molar ratio), appropriate amount of CaI2 and LiI to form 90Li7P3S11–5CaI2–5LiI (molar ratio) was put into a three-necked flask together with 40 ml of ACN. The mixture was stirred at 300 rpm and 50 °C for 24 h, the solvent was then evaporated at 80 °C under low pressure. The residue was carefully grounded using agate mortar prior to be heat treated at 270 °C for 2 h in Ar atmosphere to obtain 90Li7P3S11–5CaI2–5LiI solid electrolyte (hereafter denote as 5CaI2–5LiI).

Structural characterization

The structure of the prepared samples was characterized with X-ray diffraction (XRD; X8, Bruker), Raman spectroscopy (Horiba LabRam HR spectrometer, 532 nm) and SEM (S4800, Hitachi) and EDS (ULTIM MAX, Oxford Instrument).

The samples were prepared in an Ar-filled glove box. The prepared sample was loaded into an air-tide sample holder for characterization.

AC electrochemical impedance spectroscopy

The electrical conductivity measurement was performed on a pellet prepared by uniaxially cold pressing approximately 100 mg of the powder at a pressure of 510 MPa as reported previously.26 The alternating-current impedance spectroscopy measurement was carried out using a potentiostat (PGSTAT302N, Autolab, Herisau, Switzerland) from 10 MHz to 10 Hz. Since 5CaI2–5LiI exhibited high ionic conductivity at room temperature, a pellet with thickness of about 4.2 mm was prepared to get accurate data.

3. Results and discussion

Fig. 1a showed the XRD patterns of LPS, 5CaX2 and 5CaI2–5LiI solid electrolytes. The patterns of all samples could be assigned to Li7P3S11 crystal phase.7,27 No peaks of CaX2 and LiI was detected. The position of the most intense peak in the doped samples was almost similar to that of LPS indicating that only a small amount of CaX2 incorporated to the crystal structure of Li7P3S11. It was also reported that the full doping amount of CaI2 to Li7P3S11 was about 3 mol%.16 Ujiie et al. reported that LiX (X = F, Cl, Br, I) dissolved into LPS crystal structure at less than 10% molar ratio.27 Hikima et al. found that CaS and CaI2 mainly remained at the grain boundary of Li7P3S11.24 Thus, the patterns in Fig. 1a suggested that a part of CaX2 and LiI incorporated into the structure of LPS, while the remained part existed at the grain boundary of the LPS particle. Fig. 1b showed the Raman spectra of LPS, 5CaI2 and 5CaI2–5LiI. Fig. 1c and d showed the deconvolution spectra of LPS and 5CaI2, respectively. The spectrum of LPS had two peaks at 421 and 411 cm−1, which corresponded to the vibration of PS43− and P2S74−.28 The spectrum of 5CaI2 also revealed two peaks of PS43− and P2S74− located at 421 and 403 cm−1. The spectrum of 5CaI2–5LiI showed two peaks located at 420 and 400 cm−1, which could be assigned to the local structure unit of PS43− and P2S74− in Li7P3S11, respectively. The peak positions of PS43− and P2S74− ions in samples 5CaI2 and 5CaI2–5LiI had shifted as compared to the LPS sample showing that CaI2 addition affected the vibration of PS43− and P2S74− ions. The results from Raman spectra demonstrated that a portion of CaI2 incorporated into the crystal structure of LPS. SEM-EDS results for 5CaI2–5LiI is shown in Fig. 1e. The SEs are in the form of particles with a size of several tens of micrometers. EDS results indicated that CaI2 and LiI were well dispersed in the prepared powder sample. It can be reasonably concluded that Ca2+ and I were doped in the Li7P3S11 structure based on the experimental results: the disappearance of the peak corresponding to CaI2 and LiI in the XRD results, the peak shift in the Raman spectrum, Ca and I are also dispersed in the particles as shown in the EDS results.
image file: d4ra00442f-f1.tif
Fig. 1 Structural characterization of the prepared samples. (a) XRD patterns of Li7P3S11 (LPS), 95Li7P3S11–5CaX2 (5CaX2) and 90Li7P3S11–5CaI2–5LiI (5CaI2–5LiI) solid electrolytes; (b) Raman spectra of LPS, 5CaI2 and 5CaI2–5LiI; (c and d) deconvolution of the main Raman peaks of LPS and 5CaI2, respectively; (e) SEM-EDS results of 5CaI2–5LiI.

The temperature dependence of the ionic conductivity of LPS, 5CaX2 and 5CaI2–5LiI solid electrolytes were illustrated in Fig. 2a. The electrochemical impedance spectra of 5CaI2–5LiI, which were employed to calculate ionic conductivity, were illustrated in Fig. 2b. Those of LPS and 5CaI2 were shown in Fig. S1 (ESI). The ionic conductivity at 25 °C of LPS, 5CaCl2, 5CaCl2, 5CaCl2 and 5CaI2–5LiI were 0.15, 0.19, 0.25, 1.0 and 3.1 mS, respectively. At 70 °C, the ionic conductivity of all the samples increased and reached the value of 1.05, 2.0, 2.2, 10.1 and 13 mS for LPS, 5CaCl2, 5CaCl2, 5CaCl2 and 5CaI2–5LiI, respectively. It could be seen that log10(σ) satisfied an almost linear dependence on inversed temperature and therefore followed Arrhenius equation σ = σ0[thin space (1/6-em)]exp(−Ea,DC/(kBT)). The activation energy Ea,DC was then calculated and shown in Table 1. The activation energy Ea,DC of LPS, 5CaCl2, 5CaCl2, 5CaCl2 and 5CaI2–5LiI were 36, 37, 36, 30 and 25 kJ mol−1, respectively. The results showed that CaI2 and LiI addition greatly improved the ionic conductivity and activation energy of LPS.


image file: d4ra00442f-f2.tif
Fig. 2 (a) Temperature dependence of Li7P3S11, 5CaX2 (X = Cl, Br, I) and 5CaI2–5LiI solid electrolytes; (b) electrochemical impedance spectra of the prepared 5CaI2–5LiI solid electrolyte.
Table 1 DC activation energy Ea,DC, activation energy Ea,m of ion migration at grain boundary, and characteristic time τ0,m of ion migration at grain boundary
  LPS 5CaCl2 5CaBr2 5CaI2 5CaI2–5LiI
Ea,DC/kJ mol−1 36 37 36 30 25
Ea,m/kJ mol−1 37 38 37 31 19
τ0,m/s 5.0 × 10−10 3.4 × 10−10 4.4 × 10−10 6.3 × 10−9 4.1 × 10−7


The study of dielectric properties provides information on electrical energy decay in materials with alternating electric fields. The real part of the dielectric constant, ε′, reflects the amount of energy stored in the form of polarization when an electric field is applied.29 The real part of the permittivity, ε′, was calculated using the following equation:

image file: d4ra00442f-t1.tif
where C0 = ε0(A/d) is the free space capacitance of the cell, ε0 is the permittivity of the free space (8.854 × 10−12 F m−1), and A and d are the surface area and thickness, respectively, of the sample pellet.

Fig. 3 showed the frequency dependence of the real part of dielectric constant, ε′, of (a) LPS, (b) 5CaI2, (c) 5CaI2–5LiI measured at different temperature and (d) frequency dependence of ε′ of LPS, 5CaCl2, 5CaBr2, 5CaI2, 5CaI2–5LiI measured at room temperature. The observed upturn in all plots at lower frequencies could be attributed to the electrode–electrolyte interface polarization. This is because the accumulation of charged ions near the electrode leads to the formation of space charge layer, which in turn block the electric field and enhance electrical polarization. As frequency increased, the dielectric constant decreased. This phenomenon is a typical property of ionic conducting materials.30 An increase in ε′ with an increase in temperature was observed at low and intermediate frequency region in all the samples suggesting that charge carrier movement at grain boundary was thermally activated (Fig. 3 and S2). The plots of LPS, 5CaCl2 and 5CaBr2 at room temperature continuously decreased in intermediate and high frequency region; however, the plots of 5CaI2 and 5CaI2–5LiI at room temperature exhibited a maxima at 105–106 Hz (Fig. 3d). The maxima were observed at 105–106 Hz in all the plots at 50 °C or above (Fig. 3 and S1). Those observation suggested that CaI2 and LiI addition enhanced dielectric properties of Li7P3S11. This observation was in agreement with Raman spectra (Fig. 1b) and ionic conductivity at room temperature of the prepared samples (Fig. 2).


image file: d4ra00442f-f3.tif
Fig. 3 Frequency dependence of the real part of dielectric constant, ε′, of (a) LPS, (b) 5CaI2, (c) 5CaI2–5LiI measured at different temperature and (d) frequency dependence of ε′ of LPS, 5CaCl2, 5CaBr2, 5CaI2, 5CaI2–5LiI measured at room temperature.

Fig. 4a–c showed the frequency dependence of the loss factor, tan[thin space (1/6-em)]δ, of LPS, 5CaI2 and 5CaI2–5LiI, respectively. The plot at room temperature of LPS showed two peaks at low and high frequency, which could be attributed to ion migration at grain boundary and bulk (Fig. 4a). As temperature increased, two peaks appeared at low frequency region. The one at lower frequency was assigned to electrode polarization. The peak at higher frequency shifted toward high frequency as temperature increased and could be attributed to grain boundary migration. The migration energy Ea,m of the Li+ ion moving at grain boundary could be obtained from the temperature dependence of the peak position at low frequency in tan[thin space (1/6-em)]δ. Temperature dependence of the inverse relaxation time (τm−1 = Fmax) was plotted, and the migration characteristic time τ0,m was then obtained by numerically fitting the data using the Arrhenius equation: τm = τ0,m[thin space (1/6-em)]exp(−Ea,m/(kBT)). The obtained Ea,m and τ0,m were illustrated in Table 1. The migration activation energy at grain boundary Ea,m of the prepared samples were similar to DC activation energy Ea,DC, suggesting that charge carrier moving at grain boundary was the critical process in those samples. The characteristic migration time of LPS, 5CaCl2 and 5CaBr2 were 5.0 × 10−10, 3.4 × 10−10 and 4.4 × 10−10 s, respectively. The characteristic migration time of 5CaI2 and 5CaI2–5LiI were 6.3 × 10−9 and 4.1 × 10−7 s, respectively. Therefore, the charge carrier movement at grain boundary was enhanced by CaI2 and LiI addition; it changed from short range diffusion in LPS to long range diffusion in 5CaI2 and 5CaI2–5LiI. This phenomenon was also consistent with the article by Tu et al., which reported the influence of LiI doping on the impedance at the grain boundary of Li2S–AlI3 solid solution.31


image file: d4ra00442f-f4.tif
Fig. 4 Frequency dependence of the loss factor, tan[thin space (1/6-em)]δ, of (a) LPS, (b) 5CaI2 and (c) 5CaI2–5LiI.

4. Conclusion

Li7P3S11 solid electrolytes doped with CaX2 (X = Cl, Br, I) and LiI were successfully prepared by liquid-phase synthesis using ACN as reaction medium. Results from XRD, Raman spectroscopy and SEM-EDS proved that CaX2 and LiI incorporated into the crystal structure of Li7P3S11. Results from AC electrochemical impedance spectroscopy showed that CaI2 and LiI addition enhanced dielectric properties of Li7P3S11. In addition, the charge carrier movement at grain boundary changed from short range diffusion in LPS to long range diffusion in 5CaI2 and 5CaI2–5LiI. As a result, very high ionic conductivity of 3.1 mS cm−1 at 25 °C was obtained.

Conflicts of interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Acknowledgements

We acknowledge Ho Chi Minh City University of Technology (HCMUT), VNU-HCM for supporting this study.

References

  1. S. Chen, D. Xie, G. Liu, J. P. Mwizerwa, Q. Zhang, Y. Zhao, X. Xu and X. Yao, Sulfide solid electrolytes for all-solid-state lithium batteries: structure, conductivity, stability and application, Energy Storage Mater., 2018, 14, 58–74 CrossRef.
  2. J. Lau, R. H. DeBlock, D. M. Butts, D. S. Ashby, C. S. Choi and B. S. Dunn, Sulfide Solid Electrolytes for Lithium Battery Applications, Adv. Energy Mater., 2018, 8, 201800933 CrossRef.
  3. Y. Seino, T. Ota, K. Takada, A. Hayashi and M. Tatsumisago, A sulphide lithium super ion conductor is superior to liquid ion conductors for use in rechargeable batteries, Energy Environ. Sci., 2014, 7, 627–631 RSC.
  4. Y. Kato, S. Hori, T. Saito, K. Suzuki, M. Hirayama, A. Mitsui, M. Yonemura, H. Iba and R. Kanno, High-power all-solid-state batteries using sulfide superionic conductors, Nat. Energy, 2016, 1, 16030 CrossRef CAS.
  5. P. Adeli, J. D. Bazak, A. Huq, G. R. Goward and L. F. Nazar, Influence of Aliovalent Cation Substitution and Mechanical Compression on Li-Ion Conductivity and Diffusivity in Argyrodite Solid Electrolytes, Chem. Mater., 2020, 33, 146–157 CrossRef.
  6. Y. Subramanian, R. Rajagopal and K.-S. Ryu, Synthesis of Sb-doped Li10P3S12I solid electrolyte and their electrochemical performance in solid-state batteries, J. Energy Storage, 2024, 78, 109943 CrossRef.
  7. F. Mizuno, A. Hayashi, K. Tadanaga and M. Tatsumisago, New, Highly Ion-Conductive Crystals Precipitated from Li2S–P2S5 Glasses, Adv. Mater., 2005, 17, 918–921 CrossRef CAS.
  8. M. K. Tufail, N. Ahmad, L. Yang, L. Zhou, M. A. Naseer, R. Chen and W. Yang, A panoramic view of Li7P3S11 solid electrolytes synthesis, structural aspects and practical challenges for all-solid-state lithium batteries, Chin. J. Chem. Eng., 2021, 39, 16–36 CrossRef CAS.
  9. B. Huang, X. Yao, Z. Huang, Y. Guan, Y. Jin and X. Xu, Li3PO4 -doped Li7P3S11 glass-ceramic electrolytes with enhanced lithium ion conductivities and application in all-solid-state batteries, J. Power Sources, 2015, 284, 206–211 CrossRef CAS.
  10. M. Eom, S. Choi, S. Son, L. Choi, C. Park and D. Shin, Enhancement of lithium ion conductivity by doping Li3BO3 in Li2S-P2S5 glass-ceramics electrolytes for all-solid-state batteries, J. Power Sources, 2016, 331, 26–31 CrossRef CAS.
  11. P. Lu, F. Ding, Z. Xu, J. Liu, X. Liu and Q. Xu, Study on (100-x)(70Li2S-30P2S5 )-xLi2ZrO3 glass-ceramic electrolyte for all-solid-state lithium-ion batteries, J. Power Sources, 2017, 356, 163–171 CrossRef CAS.
  12. Y. Guo, H. Guan, W. Peng, X. Li, Y. Ma, D. Song, H. Zhang, C. Li and L. Zhang, Enhancing the electrochemical performances of Li7P3S11 electrolyte through P2O5 substitution for all-solid-state lithium battery, Solid State Ionics, 2020, 358, 115506 CrossRef CAS.
  13. M. Murakami, K. Shimoda, S. Shiotani, A. Mitsui, K. Ohara, Y. Onodera, H. Arai, Y. Uchimoto and Z. Ogumi, Dynamical Origin of Ionic Conductivity for Li7P3S11 Metastable Crystal As Studied by 6/7Li and 31P Solid-State NMR, J. Phys. Chem. C, 2015, 119, 24248–24254 CrossRef CAS.
  14. Y. Seino, M. Nakagawa, M. Senga, H. Higuchi, K. Takada and T. Sasaki, Analysis of the structure and degree of crystallisation of 70Li2S–30P2S5 glass ceramic, J. Mater. Chem. A, 2015, 3, 2756–2761 RSC.
  15. M. Calpa, N. C. Rosero-Navarro, A. Miura and K. Tadanaga, Electrochemical performance of bulk-type all-solid-state batteries using small-sized Li7P3S11 solid electrolyte prepared by liquid phase as the ionic conductor in the composite cathode, Electrochim. Acta, 2019, 296, 473–480 CrossRef CAS.
  16. N. H. H. Phuc, H. Gamo, K. Hikima, H. Muto and A. Matsuda, Preparation of CaI2-Doped Li7P3S11 by Liquid-Phase Synthesis and Its Application in an All-Solid-State Battery with a Graphite Anode, Energy Fuels, 2022, 36, 4577–4584 CrossRef CAS.
  17. S. Ito, M. Nakakita, Y. Aihara, T. Uehara and N. Machida, A synthesis of crystalline Li7P3S11 solid electrolyte from 1,2-dimethoxyethane solvent, J. Power Sources, 2014, 271, 342–345 CrossRef CAS.
  18. J. Zhou, Y. Chen, Z. Yu, M. Bowden, Q. R. S. Miller, P. Chen, H. T. Schaef, K. T. Mueller, D. Lu, J. Xiao, J. Liu, W. Wang and X. Zhang, Wet-chemical synthesis of Li7P3S11 with tailored particle size for solid state electrolytes, Chem. Eng. J., 2022, 429, 132334 CrossRef CAS.
  19. R. C. Xu, X. H. Xia, Z. J. Yao, X. L. Wang, C. D. Gu and J. P. Tu, Preparation of Li7P3S11 glass-ceramic electrolyte by dissolution-evaporation method for all-solid-state lithium ion batteries, Electrochim. Acta, 2016, 219, 235–240 CrossRef CAS.
  20. M. Calpa, N. C. Rosero-Navarro, A. Miura and K. Tadanaga, Preparation of sulfide solid electrolytes in the Li2S–P2S5 system by a liquid phase process, Inorg. Chem. Front., 2018, 5, 501–508 RSC.
  21. L. Zhou, M. K. Tufail, N. Ahmad, T. Song, R. Chen and W. Yang, Strong Interfacial Adhesion between the Li2S Cathode and a Functional Li7P2.9Ce0.2S10.9Cl0.3 Solid-State Electrolyte Endowed Long-Term Cycle Stability to All-Solid-State Lithium-Sulfur Batteries, ACS Appl. Mater. Interfaces, 2021, 13, 28270–28280 CrossRef CAS PubMed.
  22. A. Hayashi, K. Minami, S. Ujiie and M. Tatsumisago, Preparation and ionic conductivity of Li7P3S11−z glass-ceramic electrolytes, J. Non-Cryst. Solids, 2010, 356, 2670–2673 CrossRef CAS.
  23. K. Minami, A. Hayashi, S. Ujiie and M. Tatsumisago, Electrical and electrochemical properties of glass–ceramic electrolytes in the systems Li2S–P2S5–P2S3 and Li2S–P2S5–P2O5, Solid State Ionics, 2011, 192, 122–125 CrossRef CAS.
  24. K. Hikima, I. Kusaba, H. Gamo, N. H. H. Phuc, H. Muto and A. Matsuda, High Ionic Conductivity with Improved Lithium Stability of CaS- and CaI2-Doped Li7P3S11 Solid Electrolytes Synthesized by Liquid-Phase Synthesis, ACS Omega, 2022, 7, 16561–16567 CrossRef CAS PubMed.
  25. H. Gamo, J. Nishida, A. Nagai, K. Hikima and A. Matsuda, Solution Processing via Dynamic Sulfide Radical Anions for Sulfide Solid Electrolytes, Adv. Energy Sustainability Res., 2022, 3, 2200019 CrossRef CAS.
  26. M. Atsunori, G. Hirotada and N. H. H. Phuc, Preparation of cubic Na3PS4 by liquid-phase shaking in methyl acetate medium, Heliyon, 2019, 5, e02760 CrossRef PubMed.
  27. S. Ujiie, T. Inagaki, A. Hayashi and M. Tatsumisago, Conductivity of 70Li2S·30P2S5 glasses and glass–ceramics added with lithium halides, Solid State Ionics, 2014, 263, 57–61 CrossRef CAS.
  28. B. Zhao, J. Wu, Z. Wang, W. Ma, Y. Shi, Y. Jiang, J. Jiang, X. Liu, Y. Xu and J. Zhang, Incorporation of lithium halogen in Li7P3S11 glass-ceramic and the interface improvement mechanism, Electrochim. Acta, 2021, 390, 138849 CrossRef CAS.
  29. D. M. A. Basset, S. Mulmi, M. S. El-Bana, S. S. Fouad and V. Thangadurai, Synthesis and characterization of novel Li-stuffed garnet-like Li5+2xLa3Ta2-xGdxO12 (0 </= x </= 0.55): structure-property relationships, Dalton Trans., 2017, 46, 933–946 RSC.
  30. A. Molak, M. Paluch, S. Pawlus, J. Klimontko, Z. Ujma and I. Gruszka, Electric modulus approach to the analysis of electric relaxation in highly conducting (Na0.75Bi0.25)(Mn0.25Nb0.75)O3ceramics, J. Phys. D: Appl. Phys., 2005, 38, 1450–1460 CrossRef CAS.
  31. T. A. Tu, N. H. H. Phuc, L. T. Q. Anh and T. V. Toan, Preparation of Li2S-AlI3-LiI Composite Solid Electrolyte and Its Application in All-Solid-State Li-S Battery, Batteries, 2023, 9, 290–300 CrossRef.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ra00442f

This journal is © The Royal Society of Chemistry 2024