Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Healable and conductive sulfur iodide for solid-state Li–S batteries

Abstract

Solid-state Li–S batteries (SSLSBs) are made of low-cost and abundant materials free of supply chain concerns. Owing to their high theoretical energy densities, they are highly desirable for electric vehicles1,2,3. However, the development of SSLSBs has been historically plagued by the insulating nature of sulfur4,5 and the poor interfacial contacts induced by its large volume change during cycling6,7, impeding charge transfer among different solid components. Here we report an S9.3I molecular crystal with I2 inserted in the crystalline sulfur structure, which shows a semiconductor-level electrical conductivity (approximately 5.9 × 10−7 S cm−1) at 25 °C; an 11-order-of-magnitude increase over sulfur itself. Iodine introduces new states into the band gap of sulfur and promotes the formation of reactive polysulfides during electrochemical cycling. Further, the material features a low melting point of around 65 °C, which enables repairing of damaged interfaces due to cycling by periodical remelting of the cathode material. As a result, an Li–S9.3I battery demonstrates 400 stable cycles with a specific capacity retention of 87%. The design of this conductive, low-melting-point sulfur iodide material represents a substantial advancement in the chemistry of sulfur materials, and opens the door to the practical realization of SSLSBs.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Structure and property characterizations of S9.3I.
Fig. 2: The electrochemical performance of S9.3I and elemental S cathodes in solid-state Li–S cells.
Fig. 3: The working mechanism of S9.3I cathode in Li–S9.3I cell at 25 °C.
Fig. 4: Repair of the S9.3I cathode/LPS solid electrolyte interface in Li–S9.3I cells by remelting.

Similar content being viewed by others

Data availability

The data that support the findings of this study are available from the corresponding author upon request.

References

  1. Bruce, P. G., Freunberger, S. A., Hardwick, L. J. & Tarascon, J. M. Li-O2 and Li-S batteries with high energy storage. Nat. Mater. 11, 19–29 (2011).

    Article  ADS  PubMed  Google Scholar 

  2. Randau, S. et al. Benchmarking the performance of all-solid-state lithium batteries. Nat. Energy 5, 259–270 (2020).

    Article  ADS  CAS  Google Scholar 

  3. Yang, Z., Huang, H. & Lin, F. Sustainable electric vehicle batteries for a sustainable world: perspectives on battery cathodes, environment, supply chain, manufacturing, life cycle, and policy. Adv. Energy Mater. 12, 2200383 (2022).

    Article  CAS  Google Scholar 

  4. Yue, J., Yan, M., Yin, Y. X. & Guo, Y. G. Progress of the interface design in all-solid-state Li-S batteries. Adv. Funct. Mater. 28, 1707533 (2018).

    Article  Google Scholar 

  5. Guo, W. et al. Artificial dual solid-electrolyte interfaces based on in situ organothiol transformation in lithium sulfur battery. Nat. Commun. 12, 3031 (2021).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  6. Fu, K. et al. Three-dimensional bilayer garnet solid electrolyte based high energy density lithium metal–sulfur batteries. Energy Environ. Sci. 10, 1568–1575 (2017).

    Article  CAS  Google Scholar 

  7. Lin, Z., Liu, Z. C., Dudney, N. J. & Liang, C. D. Lithium superionic sulfide cathode for all-solid lithium-sulfur batteries. ACS Nano 7, 2829–2833 (2013).

    Article  CAS  PubMed  Google Scholar 

  8. Ji, X., Lee, K. T. & Nazar, L. F. A highly ordered nanostructured carbon-sulphur cathode for lithium-sulphur batteries. Nat. Mater. 8, 500–506 (2009).

    Article  ADS  CAS  PubMed  Google Scholar 

  9. Pan, H. et al. Carbon-free and binder-free Li-Al alloy anode enabling an all-solid-state Li-S battery with high energy and stability. Sci. Adv. 8, eabn4372 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Yang, X., Luo, J. & Sun, X. Towards high-performance solid-state Li-S batteries: from fundamental understanding to engineering design. Chem. Soc. Rev. 49, 2140–2195 (2020).

    Article  CAS  PubMed  Google Scholar 

  11. Wu, J., Liu, S., Han, F., Yao, X. & Wang, C. Lithium/sulfide all-solid-state batteries using sulfide electrolytes. Adv. Mater. 33, 2000751 (2021).

    Article  CAS  Google Scholar 

  12. Nagao, M., Hayashi, A. & Tatsumisago, M. High-capacity Li2S–nanocarbon composite electrode for all-solid-state rechargeable lithium batteries. J. Mater. Chem. 22, 10015–10020 (2012).

    Article  CAS  Google Scholar 

  13. Chen, Z. et al. Bulk/interfacial synergetic approaches enable the stable anode for high energy density all solid-state lithium−sulfur batteries. ACS Energy Lett. 7, 2761–2770 (2022).

    Article  CAS  Google Scholar 

  14. Yao, X. et al. High performance all-solid-state lithium–sulfur batteries enabled by amorphous sulfur-coated reduced graphene oxide cathodes. Adv. Energy Mater. 7, 1602923 (2017).

    Article  Google Scholar 

  15. Saßnick, H. D. & Cocchi, C. Electronic structure of cesium-based photocathode materials from density functional theory: performance of PBE, SCAN, and HSE06 functionals. Electron. Struct. 3, 027001 (2021).

    Article  ADS  Google Scholar 

  16. Hautier, G. et al. Phosphates as lithium-ion battery cathodes: an evaluation based on high-throughput ab initio calculations. Chem. Mater. 23, 3495–3508 (2011).

    Article  CAS  Google Scholar 

  17. Liu, G., Niu, P., Yin, L. & Cheng, H.-M. α-Sulfur crystals as a visible-light-active photocatalyst. J. Am. Chem. Soc. 134, 9070–9073 (2012).

    Article  CAS  PubMed  Google Scholar 

  18. Abass, A. K. & Ahmad, N. H. Indirect band gap investigation of orthorhombic single crystals of sulfur. J. Phys. Chem. Solids 47, 143–145 (1986).

    Article  ADS  CAS  Google Scholar 

  19. Chen, X. et al. Dynamically preferred state with strong electronic fluctuations from electrochemical synthesis of sodium manganite. Matter 5, 735–750 (2022).

    Article  CAS  Google Scholar 

  20. Lee, Y. G. et al. High-energy long-cycling all-solid-state lithium metal batteries enabled by silver-carbon composite anodes. Nat. Energy 5, 299–308 (2020).

    Article  ADS  CAS  Google Scholar 

  21. Ye, L. & Li, X. A dynamic stability design strategy for lithium metal solid state batteries. Nature 593, 218–222 (2021).

    Article  ADS  CAS  PubMed  Google Scholar 

  22. Li, C. et al. A quasi-intercalation reaction for fast sulfur redox kinetics in solid-state lithium–sulfur batteries. Energy Environ. Sci. 15, 4289–4300 (2022).

    Article  CAS  Google Scholar 

  23. Zhang, H. et al. Designer anion enabling solid-state lithium-sulfur batteries. Joule 3, 1689–1702 (2019).

    Article  CAS  Google Scholar 

  24. Zhang, Y. et al. Se as eutectic accelerator in sulfurized polyacrylonitrile for high performance all-solid-state lithium-sulfur battery. Energy Storage Mater. 21, 287–296 (2019).

    Article  Google Scholar 

  25. Li, X. et al. High-performance Li-SeSx all-solid-state lithium batteries. Adv. Mater. 31, 1808100 (2019).

    Article  Google Scholar 

  26. Li, M. et al. Solid-state lithium–sulfur battery enabled by thio-LiSICON/polymer composite electrolyte and sulfurized polyacrylonitrile cathode. Adv. Funct. Mater. 30, 1910123 (2020).

    Article  CAS  Google Scholar 

  27. Wang, D. et al. Realizing high-capacity all-solid-state lithium-sulfur batteries using a low-density inorganic solid-state electrolyte. Nat. Commun. 14, 1895 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Liu, Y., Meng, X., Wang, Z. & Qiu, J. A Li2S-based all-solid-state battery with high energy and superior safety. Sci. Adv. 8, eabl8390 (2022).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  29. Patel, M. U. et al. X-ray absorption near-edge structure and nuclear magnetic resonance study of the lithium-sulfur battery and its components. ChemPhysChem 15, 894–904 (2014).

    Article  CAS  PubMed  Google Scholar 

  30. Liang, X. et al. A highly efficient polysulfide mediator for lithium-sulfur batteries. Nat. Commun. 6, 5682 (2015).

    Article  ADS  PubMed  Google Scholar 

  31. Nandasiri, M. I. et al. In-situ chemical imaging of solid-electrolyte interphase layer evolution in Li–S batteries. Chem. Mater. 29, 4728–4737 (2017).

    Article  CAS  Google Scholar 

  32. Yang, C. et al. Unique aqueous Li-ion/sulfur chemistry with high energy density and reversibility. Proc. Natl Acad. Sci. USA 114, 6197–6202 (2017).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  33. Li, X. et al. Highly stable halide-electrolyte-based all-solid-state Li-Se batteries. Adv. Mater. 34, 2200856 (2022).

    Article  CAS  Google Scholar 

  34. Lu, Y., Zhao, C. Z., Huang, J. Q. & Zhang, Q. The timescale identification decoupling complicated kinetic processes in lithium batteries. Joule 6, 1172–1198 (2022).

    Article  CAS  Google Scholar 

  35. Guo, Q., Lau, K. C. & Pandey, R. Thermodynamic and mechanical stability of crystalline phases of Li2S2. J. Phys. Chem. C 123, 4674–4681 (2019).

    Article  CAS  Google Scholar 

  36. Sofekun, G. O. et al. The rheology of liquid elemental sulfur across the λ-transition. J. Rheol. 62, 469–476 (2018).

    Article  ADS  CAS  Google Scholar 

  37. Wan, T., Saccoccio, H. M., Chen, C. & Ciucci, F. Influence of the discretization methods on the distribution of relaxation times deconvolution: implementing radial basis functions with DRT tools. Electrochim. Acta 184, 483–499 (2015).

    Article  CAS  Google Scholar 

  38. Stoll, S. & Schweiger, A. EasySpin, a comprehensive software package for spectral simulation and analysis in EPR. J. Magn. Reson. 178, 42–55 (2006).

    Article  ADS  CAS  PubMed  Google Scholar 

  39. Ravel, B. & Newville, M. ATHENA, ARTEMIS, HEPHAESTUS: data analysis for X-ray absorption spectroscopy using IFEFFIT. J. Synchrotron Radiat. 12, 537–541 (2005).

    Article  CAS  PubMed  Google Scholar 

  40. Hammersley, A. P., Svensson, S. O., Hanfland, M., Fitch, A. N. & Hausermann, D. Two-dimensional detector software: from real detector to idealised image or two-theta scan. High Press. Res. 14, 235–248 (1996).

    Article  ADS  Google Scholar 

  41. Qiu, X., Thompson, J. W. & Billinge, S. J. L. PDFgetX2: a GUI-driven program to obtain the pair distribution function from X-ray powder diffraction data. J. Appl. Crystallogr. 37, 678–678 (2004).

    Article  ADS  CAS  Google Scholar 

  42. Blochl, P. E. Projector augmented-wave method. Phys. Rev. B 50, 17953–17979 (1994).

    Article  ADS  CAS  Google Scholar 

  43. Kresse, J. F. G. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 54, 1169–1186 (1996).

    Article  Google Scholar 

  44. Heyd, J., Scuseria, G. E. & Ernzerhof, M. Hybrid functionals based on a screened Coulomb potential. J. Chem. Phys. 118, 8207–8215 (2003).

    Article  ADS  CAS  Google Scholar 

  45. Tang, W., Sanville, E. & Henkelman, G. A grid-based Bader analysis algorithm without lattice bias. J. Phys. Condens. Matter 21, 084204 (2009).

    Article  ADS  CAS  PubMed  Google Scholar 

  46. Ong, S. P. et al. Python Materials Genomics (pymatgen): a robust, open-source python library for materials analysis. Comput. Mater. Sci. 68, 314–319 (2013).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

This work was supported by the Advanced Research Projects Agency–Energy, US Department of Energy (DOE), under contract no. DE-AR0000781. M.L.H.C. and S.P.O. acknowledge the support from the Materials Project, funded by the US DOE, Office of Science, Office of Basic Energy Sciences, Materials Sciences and Engineering Division, under contract no. DEAC02-05-CH11231 (Materials Project Program No. KC23MP). Computing resources were provided by the National Energy Research Scientific Computing Center (NERSC) and the Advanced Cyberinfrastructure Coordination Ecosystem: Services & Support (ACCESS) under grant no. DMR-150014. S.T. and E.H. are supported by the Assistant Secretary for Energy Efficiency and Renewable Energy (EERE), Vehicle Technology Office of the US DOE through the Advanced Battery Materials Research (BMR) Program under contract no. DE-SC0012704. This research used 28-ID−2, 8-BM and 7-BM beamlines of the National Synchrotron Light Source II, US DOE Office of Science User Facilities, operated for the DOE Office of Science by Brookhaven National Laboratory under contract no. DE-SC0012704. C. Wu, Z.F. and Y.Y. are supported by the US DOE’s Office of EERE under the Vehicle Technologies Program under contract no. DE-EE0008864. This work made use of the shared facilities of the UC Santa Barbara MRSEC (grant no. DMR-720256), a member of the Materials Research Facilities Network (http://www.mrfn.org). This research used the Electron Microscopy facility of the Center for Functional Nanomaterials (CFN), which is a US DOE Office of Science User Facility, at Brookhaven National Laboratory, under contract no. DE-SC0012704. Canhui Wang and Chao Wang were supported by the Advanced Research Projects Agency–Energy, US DOE, under contract no. DE-AR0001191. FIB and SEM characterizations were performed at the San Diego Nanotechnology Infrastructure (SDNI) of UCSD supported by the NSF (grant no. ECCS1542148). Raman facilities were supported by the NSF through UCSD MRSEC, grant no. DMR-201192. The authors acknowledge the use of facilities and instrumentation at the UC Irvine (UCI) Materials Research Institute (IMRI), which was supported in part by the NSF through the UCI MRSEC (grant no. DMR-2011967). XPS facilities were funded in part by the NSF Major Research Instrumentation. We thank the Molecular Mass Spectrometry Facility at UC San Diego for performing the MALDI-TOFMS measurement, which is supported by the NSF under grant no. CHE−1338173. Part of this work was performed in the Cordx Yufeng Li Collaboratory.

Author information

Authors and Affiliations

Authors

Contributions

J.Z. and P.L. conceptualized the idea and designed all of the experiments. S.P.O. and M.L.H.C. performed the theoretical calculations and drafted the results. S.T. and E.H. conducted the XAS and X-ray PDF tests. S.W. performed the Raman, DSC, cryo-SEM and XPS measurements. C. Wu, Z.F. and Y.Y. collected the cross-sectional SEM images of the full cells at different states, including anode interfaces and cathode interfaces. H.N. collected EPR data, H.N. and R.J.C. analysed and wrote up the results. Canhui Wang and Chao Wang performed the cryo-TEM. Y.X. and E.E.F. performed the XRD tests. Q.R.S.M. did the in situ heating XRD test. H.L., S.Y., G.H., J. Holoubek, J. Hong and C.S. helped with the synthesis of materials, cell fabrications, electrochemical performance tests, data analysis, discussion and revision. C.J.B. helped with the revised manuscript preparation and discussion. J.Z., P.L. and S.P.O. drafted the manuscript with input and revision from all authors. P.L. and S.P.O. supervised the research. J.Z. and M.L.H.C. contributed equally to this work.

Corresponding authors

Correspondence to Shyue Ping Ong or Ping Liu.

Ethics declarations

Competing interests

P.L. and J.Z. report a US provisional patent application filed on February 13, 2023, Serial No. _63/484,659, based on this work.

Peer review

Peer review information

Nature thanks Pouya Partovi-Azar and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data figures and tables

Extended Data Fig. 1 Healable interfaces in SSLSBs with a low-melting-point sulfur iodide, and the synthesis and characterization of sulfur iodide materials.

(a) Schematic of a solid state battery with elemental sulfur as the active material. Poor solid/solid contact develops during cycling due to volume changes of the active material. (b) Schematic of a solid state battery with sulfur iodide as the active material. Ideal active material/electrolyte interface is achieved through periodical heating to melt the cathode, thus healing the interface. (c) Illustration of the procedures to prepare S9.3I. (d) The DSC curves of sulfur iodide with different S:I ratios in the temperature window of 20−150 °C. With S:I ratios decreasing from 1:0 to 9.3:1, the typical sulfur endothermic peak gradually disappears to transition to one phase melting behavior. Beyond 6:1, the exothermic peaks assigned to iodine are observed. (e) Spectra obtained during in situ heating XRD of S9.3I from 30 to 130 °C with an 5 °C interval. (f) Raman spectra of sulfur iodide with different ratios, which features an iodine peak located at 423.8 cm−1 when the S:I ratio is less than 9.3:1. The cryo-SEM image (g) and the corresponding EDX mapping images of I (h) and S (i) element distribution of S9.3I after melting. (j) XRD of S, I2, S35.7I, S15.9I, S9.3I, S6I. In the low angle range, the relative intensity of the broadened peak centered at ~1.5° gradually increases with the increase of iodine, but from 9.3:1 to 6:1, the iodine diffraction peaks also emerge. (k) Mass spectra of S and S9.3I. Compared to S, S9.3I shows two additional peaks centered at 127.0 and 253.8 m/z, which are attributed to I and I2 species respectively. No peaks can be indexed to species containing S-I bonds. (l) The PDF of sulfur iodide materials after melting with different S and I ratios, elemental sulfur and iodine. When the ratio of S to I is below 9.3:1, bulk I2 is present as indicated by the signals from the short and long range structures.

Extended Data Fig. 2 Computed Crystal and electronic structure of S-I compounds.

(a) Computed XRD patterns of elemental S and predicted S9.6I structure. (b) Computed PDF of S9.6I. We attribute any minor discrepancies to the fact that the DFT relaxations are carried out at 0 K, while the experimental data is obtained at 300 K. (c) Computed XRD patterns of S, S9.6I, S8I and S with S8 ring vacancy structure. (d) HSE projected density of states of S, S9.6I and S8I. Elemental projected density of states for S (e) and S9.6I (f) using SCAN functionals.

Extended Data Fig. 3 Electronic structure and conductivities of sulfur iodide materials.

(a) Variable temperature X-band EPR spectra collected on S9.3I and elemental sulfur. The radical concentration in S9.3I grows with increasing temperature, but not in S. (b) Temperature evolution of the EPR signal intensity (top panel), linewidth (bottom panel), and g-factor (middle panel) of S9.3I. The g-factor for S9.3I is approximately 1.99 and does not change appreciably with increasing temperature (bottom panel), suggesting that similar radical species are produced over the entire temperature range. (c) EPR spectra collected at 100 K on S9.3I before and after melting. The electronic conductivity of S35.7I (d), S15.9I (e), and S6I (f) measured by using a potentiostatic test at room temperature.

Extended Data Fig. 4 SEM and elemental mapping of S and S9.3I mixed with LPS after heating at 100 °C.

The SEM image of S/LPS mixture (a) and S9.3I/LPS mixture (b) after milling and heating at 100 °C. (c) The cross sectional cryo-FIB SEM images and corresponding element distribution images of S/LPS for S and P. The distribution of S and P is inhomogeneous. (d) Distribution of S, P, and I in S9.3I/LPS. The distribution in the entire area is homogeneous.

Extended Data Fig. 5 Electrochemical performance data of solid state cell components.

The cycling stability and capacity of a LPS/VGCF cathode at 100 °C (a) and 25 °C (c). The corresponding voltage profiles at 100 °C (b) and at 25 °C (d). Capacity contribution of LPS is very limited and the cycling stability of LPS is very poor either at 100 °C or 25 °C with low coulombic efficiencies. The long-term cycling stability of Li/SP/LPSCl/SP/Li symmetric cells at 100 °C (e) and 25 °C (f) at a current density of 0.3 mA cm−2.

Extended Data Fig. 6 The performance of sulfur iodide cathode at different conditions and their comparison with other SSLSBs.

The dQ/dV curves during delithiation of S9.3I (a) and S (b) cathode based on the charging curves at 100 °C from Figs. 2a and b. (c) The cycling stability of S, S35.7I, S15.9I, S9.3I and S6I cathode at 1.6 A g−1 and at 100 °C. (d) The corresponding cycling coulombic efficiency in (c). (e) Rate capability of S9.3I cathode at 100 °C at current densities of 0.32, 0.48, 0.80, 1.28, 1.60, 2.40, 3.20, 4.00, 4.80 and 5.60 A g−1. Rate performance comparison between Li-S9.3I solid state batteries with other SSLSBs: based on the weight of (f) only S and (g) S composite, red-colored symbols represent Li metal anode based SSLSBs while other cells used non-Li metal anodes. (h) The cycling performance of S9.3I cathode with a high mass loading of 4.2 mg cm−2 at 0.32 A g−1 and at 100 °C. (i) The corresponding voltage profiles at different cycles in (h). (j) Cycling stability comparison between Li-S9.3I solid state batteries with other reported SSLSBs. The cycling performance of the S9.3I cathode at varied temperatures and current densities: (k) 80 °C, 0.8 A g−1, (l) 60 °C, 0.8 A g−1, (m) 40 °C, 0.32 A g−1. It should be noted that a low current density was applied to the cells tested at all temperatures during the initial formation cycle.

Extended Data Fig. 7 Morphological study of the Li-SP-LPSCl interface during cycling in full cells.

The cross-sectional SEM of Li/SP/LPSCl interface at different states: pristine (a), fully discharged to 1.5 V (b) and fully charged back to 3.0 V (c). It should be noted that all of the anode interfaces were cut from the full cells with the configuration of S9.3I-LPS-VGCF/LPS/LPSCl/SP/Li with a high mass loading of S9.3I 4.2 mg cm−2 running at 100 °C and at a current density of 0.32 A g−1. At the pristine state (a), both the Li/SP interface and SP/LPSCl SSE interface are well connected. After discharge (b), the thickness of Li decreases from ~50 μm to ~30 μm, the 20 μm difference corresponding to an area capacity of ~4.0 mAh cm−2. After charging back to 3.0 V (c), the thickness of Li recovers back to ~50 μm. Importantly, both the Li/SP interface and SP/LPSCl SSE interface are well maintained without the growth of Li dendrites during the discharge and charge processes.

Extended Data Fig. 8 EIS study of Li-S9.3I and Li-S full cell at 25 °C during initial cycles.

(a) The initial discharge/charge curves of a Li-S9.3I cell, and the corresponding dQ/dV curves (b) derive from (a). (c) The EIS of the Li-S9.3I cell at different states of discharge. (d) The EIS of the Li-S9.3I cell at different states of charge. The DRT analysis of EIS at different states of discharge (e) related to (c), and states of charge (f) related to (d). (g) The initial discharge/charge curves of a Li-S cathode, and the corresponding dQ/dV curves (h) derive from (g). (i) The EIS of the Li-S cell at different states of discharge. (j) The EIS of the Li-S cell at different states of charge. The DRT analysis of EIS at different states of discharge (k) related to (i), and states of charge (l) related to (j). The impedance of S9.3I cathode is much lower than that of pure S cathode during cycling.

Extended Data Fig. 9 Characterizations of intermediates during cycling to diagnose the S9.3I cathode working mechanism.

(a) The corresponding ratios of S0, Sb, St, and S2- in the S9.3I cathode in Fig. 3b at different discharge/charge states. (b) Ex situ Raman spectra of S9.3I during the discharge/charge processes. The ex situ Raman spectra suggest the good reversibility of S9.3I cathode at RT. X-band EPR spectra obtained on (c) the pristine S9.3I cathode, and on (d) the SSE and carbon mixture (LPS-VGCF). (e) Ex situ EPR data obtained on cathode samples harvested from cells stopped at various stages of the initial discharge/charge processes. (f) Parameters obtained from fits of the EPR spectra in (e), including the ratios of the signal intensities from the S9.3I cathode material and from the VGCF additive, and the linewidth of the S9.3I EPR signal. The ratio of S9.3I radicals to carbon decreases during initial discharge and grows upon subsequent charge (f). This trend is consistent with the following model, assuming that polysulfide chains that form in the solid-state exist as dianions and cannot be detected in our EPR measurements, unlike EPR-active polysulfide radicals formed in solution through disproportionation of polysulfide dianions (as in conventional Li-S cells). (g) The S2p high resolution XPS of comparison samples, including LPS/VGCF mixture, S cathode discharge to 1.3 V (D-1.3 V) and chemically lithiated S9.3I prepared by reacting with Li powder with an equivalent capacity of ~800 mAh g−1. (h) Digital photos of elemental S cathode discharged to 1.3 V, S9.3I cathode discharged to 1.3 V, chemically lithiated S9.3I immersed in THF solution and standard Li2S4, Li2S6 THF solutions. (i) The UV-Vis spectra of corresponding samples in (h). The S9.3I cathode discharged to 1.3 V and chemically lithiated S9.3I suspensions exhibit the color of LiSx solutions immediately, but elemental S cathode discharged to 1.3 V doesn’t show any change.

Extended Data Fig. 10 Structure and property evolution of S9.3I cathode during cycling.

(a) XRD of chemically lithiated S9.3I. A chemically lithiated S9.3I particle with a few micrometers (b) and corresponding iodine element mapping (c) shows a uniform iodine distribution. EDX mapping images of S9.3I cathode at different states: (d) Pristine, (e) after 10 cycles and heated over 48 h at 100 °C, and the corresponding element analysis. There is only a very small S: I atomic ratio change after 10 cycles and then heated at 100 °C for over 48 h, indicating the thermal stability of S9.3I cathode. (f) EIS spectra and (g) corresponding DRT analysis of Li-S9.3I full cell at different cycles.

Extended Data Table 1 Comparison of experimental bulk S8 cell parameters with PBE, optB88 and SCAN computed values
Extended Data Table 2 Comparison of cycling stability and rate capacities of S9.3I and other previously reported high performance S composites cathodes in SSLSBs

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhou, J., Holekevi Chandrappa, M.L., Tan, S. et al. Healable and conductive sulfur iodide for solid-state Li–S batteries. Nature 627, 301–305 (2024). https://doi.org/10.1038/s41586-024-07101-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-024-07101-z

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing